0% found this document useful (0 votes)
22 views640 pages

Large Wood National Manual Final

The Large Wood National Manual provides guidance on the assessment, planning, design, and maintenance of large wood in fluvial ecosystems to restore ecological processes and functions. It emphasizes the importance of using large wood as a natural engineering solution for habitat restoration while ensuring scientific rigor in project design. The document serves as a resource for federal and state agencies, practitioners, and planners involved in river restoration efforts across the United States.

Uploaded by

Család Réder
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
22 views640 pages

Large Wood National Manual Final

The Large Wood National Manual provides guidance on the assessment, planning, design, and maintenance of large wood in fluvial ecosystems to restore ecological processes and functions. It emphasizes the importance of using large wood as a natural engineering solution for habitat restoration while ensuring scientific rigor in project design. The document serves as a resource for federal and state agencies, practitioners, and planners involved in river restoration efforts across the United States.

Uploaded by

Család Réder
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 640

LARGE WOOD NATIONAL MANUAL

ASSESSMENT, PLANNING, DESIGN, AND MAINTENANCE


OF LARGE WOOD IN FLUVIAL ECOSYSTEMS:
RESTORING PROCESS, FUNCTION, AND STRUCTURE

July 10, 2015


This page intentionally left blank.
LARGE WOOD NATIONAL MANUAL
ASSESSMENT, PLANNING, DESIGN, AND MAINTENANCE OF LARGE
WOOD IN FLUVIAL ECOSYSTEMS:
RESTORING PROCESS, FUNCTION, AND STRUCTURE

PREPARED FOR:
U.S. Bureau of Reclamation and U.S Army Corps of Engineers
Pacific Northwest Regional Office Engineer Research and Development
1150 North Curtis Road, Suite 100 Center Environmental Laboratory
Boise, ID 83706-1234 3909 Halls Ferry Road
Vicksburg, MS 39180

PREPARED BY:
ICF International
710 Second Avenue, Suite 550
Seattle, WA 98104
and
Natural Systems Design
1900 N Northlake Way #211
Seattle, WA 98105
and
Doug Shields Engineering

850 Insight Park Avenue


University, MS 38677

SUGGESTED CITATION:

Bureau of Reclamation and U.S. Army Corps of Engineers. 2015. National Large Wood
Manual: Assessment, Planning, Design, and Maintenance of Large Wood in Fluvial
Ecosystems: Restoring Process, Function, and Structure. 628 pages + Appendix. Available:
www.usbr.gov/pn/.

Large Wood National Manual July 2015


i
This page intentionally left blank.
Preface
The historical legacies of anthropogenic removal of large wood from rivers, reduction of wood inputs
through land use alterations and channel hardening, the loss of large riparian trees that once formed
stable snags, and alterations in wood transport ranging from levees to dams all have contributed to the
degradation of riverine ecosystems and decline of native species. Wood was not just debris that rivers
carried to the sea; it also altered channel morphology, fluvial processes, the storage of sediment and
organic matter, and even the evolution of landscapes. The role of wood in creating aquatic and riparian
habitat has led many regulatory agencies and fisheries advocates to recommend the reintroduction of
large wood. It makes sense that the methods and manner in which wood gets reintroduced would differ
based on hydro-geomorphic conditions and project goals. But the wide range of wood projects and their
performance cannot be explained just by differences in site conditions and goals.

This manual is intended to help establish more consistent methods to assess, design, and manage wood
projects to restore streams and rivers throughout the United States. Various federal and state agencies
are increasingly advocating that more wood be used as a softer, more cost-effective, and ecologically
beneficial engineering approach in restoration and mitigation projects to meet environmental mandates
and endangered species requirements, while maintaining traditional agency missions. The term softer
should only imply that wood is a natural part of a stream and, therefore, better fits within the context of
restoring natural conditions. But there should not be anything soft about the analysis and design of
wood projects: they should be conducted with the same scientific and engineering rigor as any river
project. The failure of wood placements in restoration is entirely the fault of the design, not the material.
By understanding the geomorphology, hydraulics, and geotechnical aspects of a project and with good
engineering, stable wood structures can be designed for various situations and longevities. In many
situations it may be desirable to place wood that can move, but designers should understand the fate
and function of such programs. In the end, it is stable wood that most directly benefits restoration, and
the underlying goal of wood projects should be to restore the function of wood until riparian forests are
able to supply the large trees that can sustain those functions. In highly constrained systems where that
may not be possible, engineering solutions can still be pursued to restore the function of wood well into
the future.

The roles of the U.S. Bureau of Reclamation (Reclamation) and U.S. Army Corps of Engineers (USACE) in
protecting native and listed species while meeting water delivery and managing flood risk, navigational,
and ecosystem restoration mandates have become increasingly diverse and in demand over the past few
decades. Reclamation and the USACE have missions that span the United States. Staff tasked with
developing designs for projects, providing technical support, or executing regulatory review must
ensure that these projects meet habitat improvement goals—and, in some cases, population
improvement metrics—with minimal risks and maximum benefits at reasonable cost. As public
stewards, Reclamation and the USACE are also tasked with ensuring due diligence with design of these
projects to prevent unanticipated harm to private landowners, infrastructure, or recreationalists on the
river. Noting Reclamation’s and the USACE’s shared missions, mandates, and broad geographic focus, an
interagency team of leaders and senior scientists recently recommended a cooperative effort to better
understand existing practice; develop collaborative assessment, design, and construction guidelines; and
improve standards for wood-based restoration engineering. In the past, the majority of large wood
design and implementation has occurred by practitioners. With increased involvement from the federal
sector in these types of projects, it is prudent to have a common set of guidelines for the use of wood in
restoration efforts that are used by agency staff and serve as a foundation for future planning, design,

Large Wood National Manual July 2015


i
Bureau of Reclamation and
U.S. Army Corps of Engineers Preface

implementation, and regulatory review. This document is intended to serve as the initial step in the
process of agency acceptance of developing standardized practices for the maintenance of existing, and
placement of new, wood structures in fluvial ecosystems by providing technical guidance. If appropriate,
formal agency acceptance of these guidelines will be determined at a later date.

This document is also meant to serve as a practical resource for planners and to help practitioners in the
restoration industry to understand more fully the roles of wood and how it should be reintroduced and
managed in fluvial ecosystems using both active (placement) and passive (recruitment and transport)
methods. In summary, this effort’s goal was to develop a comprehensive publication for the planning,
design, placement, maintenance, and assessment of large wood in rivers and streams, with an
overarching emphasis on restoring ecosystem forms, processes, and functions, given the current states
of science and practice. In fields as fast-changing as restoration ecology, design, and practice, the authors
here recognize—and hope—that this material will be improved with additional knowledge and
experience. That recognition, however, does not address the current need for technical guidance. We
believe the basic elements of this publication will hold true long into the future—particularly the
underlying premise that wood is a critical component of fluvial systems that will only become more
appreciated with additional research. As such, Reclamation and the USACE hope that this document
provides needed technical assistance to restoration practitioners as well as acts as a catalyst to drive
further innovations and improved benefits for aquatic ecosystem restoration.

D. J. Bandrowski, U.S. Bureau of Reclamation*

Jock Conyngham, U.S. Army Corps of Engineers

* Currently – Yurok Tribe

Large Wood National Manual July 2015


ii
DISCLAIMER
This document provides information to states, territories, authorized tribes, local governments,
watershed organizations, and the public regarding technical tools and sources of material for the
planning, design, placement, and maintenance of large wood in rivers. The document may refer to
statutory and regulatory provisions that contain legally binding requirements. The document does not
substitute for those provisions or regulations, nor is it a regulation itself. Thus, it does not impose
binding requirements on federal agencies, states, territories, authorized tribes, local governments,
watershed organizations, or the public and might not apply to a particular situation based upon the
circumstances. Federal agencies, state, territory, local government, and authorized tribe decision makers
retain the discretion to adopt approaches on a case-by-case basis that differ from this guidance. The use
of nonmandatory words like should, could, would, may, might, recommend, encourage, expect, and can in
this document means solely that something is suggested or recommended; it does not mean that the
suggestion or recommendation is legally required, that it imposes binding requirements, or that
following the suggestion or recommendation necessarily creates an expectation of any federal agency
approval.

This document is not intended to replace any existing planning guidelines previously adopted by federal
agencies, such as the Natural Resource Conservation Service’s Stream Restoration Design Handbook
(NEH 654); U.S. Army Corps of Engineers’ (USACE’s) Engineer Regulations (ERs) 1105-2-100, 1165-2-501,
and 1165-2-100; and U.S. Environmental Protection Agency’s Handbook for Developing Watershed Plans
to Restore and Protect our Water (2008) and A Quick Guide to Developing Watershed Plans to Restore and
Protect Our Waters (2013). Rather, it addresses how the use of large wood can be considered in concert
with these restoration planning processes.

Interested parties are free to raise questions and objections about the appropriateness of applying the
guidance provided herein to a specific situation, and Reclamation and USACE will consider whether the
recommendations in this guidance are appropriate in that situation. Reclamation and USACE may
change or add to this document in the future.

Large Wood National Manual July 2015


iii
Bureau of Reclamation and
U.S. Army Corps of Engineers Disclaimer

This page intentionally left blank.

Large Wood National Manual July 2015


iv
ACKNOWLEDGMENTS
Federal Agency Support
The idea for preparing this manual came from an interagency meeting of
researchers and leaders from the Bureau of Reclamation (Reclamation) and
the U.S. Army Corps of Engineers (USACE) Environmental Laboratory,
Engineer Research and Development Center, held in Sacramento, California
in May 2011. Participants at that meeting identified research and guidance
on the roles and utilization of large wood in riverine systems as a first-
priority national restoration need. Staff from Reclamation’s Sedimentation
and River Hydraulics Group and other offices and the USACE Environmental
Laboratory then organized a workshop in Seattle in February 2012, where
40 experts from federal, state, and local agencies; tribal governments;
academia; and the private sector gathered to refine research and application
needs. The lack of current design and application guidance was identified by that group as a critical
problem. That group helped with this publication’s initial scoping and served in various other roles.

David (DJ) Bandrowski and Jock Conyngham served as agency leads and
proponents for Reclamation and USACE, respectively, for the development of
this document. Their responsibilities including translating the
recommendations from earlier meetings into action; refining the manual’s
scope; securing funding from multiple sources; developing and executing a
support contract; and identifying appropriate authors, co-authors, and peer
reviewers. Finally, they each served as Chapter Authors and Lead Technical
Editors, and steered the manual through internal agency channels.
Reclamation’s primary financial support came from the Pacific Northwest
Region’s Columbia-Snake Recovery Office (CSRO) in Boise, Idaho, in addition to
the Research and Development Office, Science and Technology Program at the Denver Technical Service
Center. USACE support for the publication came primarily from the Engineering With Nature (EWN)
Program, in addition to the Ecosystem Management and Restoration Research Program (EMRRP), both
located at the Environmental Laboratory in Vicksburg, Mississippi.

Primary Contractors
ICF Jones and Stokes Inc., an affiliate of ICF International (hereafter ICF International)
prepared this document under contract with the U.S. Bureau of Reclamation. As such,
ICF International was the prime contractor, contributed to development of the
materials, managed the publication process, and was ultimately responsible for
producing the publication. To this end, 16 professionals on ICF International staff
contributed to this effort by serving as the managing editor, chapter authors, formatting and production
editors, production coordinator, and editorial support staff. In particular, Leo Lentsch, as Managing
Editor and Chapter Author, brought the skills that were often needed to maintain progress on this effort
while also contributing the perspective of a seasoned fish conservation biologist and natural resource
manager with over 35 years of experience. Additionally, Jasmin Mejia and Ken Cherry deserve special
recognition for their persistence in handling the final details of production.

Two primary firms assisted ICF International with preparation of this publication.

Large Wood National Manual July 2015


v
Bureau of Reclamation and
U.S. Army Corps of Engineers Acknowledgments

Natural Systems Design provided four scientists for this effort. They served as
a Lead Technical Editor, Chapter Authors, and editorial support staff. As a
Lead Technical Editor and Chapter Author, Tim Abbe brought a unique
perspective and wealth of information on the use of large wood in aquatic
restoration projects. His input was instrumental to preparing this document.

Doug Shields Engineering provided two individuals for this effort. They
served as a Lead Technical Editor, Chapter Author, and editorial support staff.
As a primary contributor to this document, Doug Shields’ influence was
unmeasurable. He brought insight, knowledge, and a “get it done” approach
that made completion of this effort possible.

Numerous additional individuals contributed text, provided source information, served as expert
reviewers, or provided invaluable constructive comments on the various drafts. The experience that
they bring, as captured in this manual, will promote the science and art of stream restoration design.

Other Contractors
Individuals or their associated firms/organizations that were paid for their contributions to this
document include: Tracy Drury (Anchor QEA, LLC), Rocco Fiori (Fiori Geoscience), Martin Fox (Fox
Environmental), James MacBroom (Milone & MacBroom, Inc.), Rebecca Manners (University of
Montana), Jordan Rosenfeld (British Columbia Ministry of Environment), Roy Schiff (Milone and
MacBroom, Inc.), Andrew Simon (Cardno ENTRIX), Emily Stanley (University of Wisconsin), Dana
Warren (Oregon State University), and Ellen Wohl (Colorado State University).

Technical Reviewers and Other Contributors


Perhaps, most notably, numerous individuals contributed their time: Paul Bakke (U.S. Fish and Wildlife
Service), Bob Banard (Washington Department of Fish and Wildlife), Janine Castro (U.S. Fish and
Wildlife Service), Brian Cluer (National Oceanic and Atmospheric Administration), Michelle Cramer
(Washington Department of Fish and Wildlife), Todd Crowl (Florida International University), David
Gaeuman (U.S. Bureau of Reclamation), Christopher Gippel (Fluvial Systems), Bob Gubernick (U.S. Forest
Service, Region 9), Casey Kramer ( Washington State Department of Transportation), Greg Koonce
(Interfluve, Inc.), Peter Lagasse (Ayers Associates), Jim Park (Washington State Department of
Transportation), Patricia Olson (Washington Department of Ecology), Roger Peters (U.S. Fish and
Wildlife Service), Rob Schanz (Washington State Department of Transportation), C. Anna Toline (U.S.
National Park Service), Doug Thompson (Connecticut College), and Rod Wittler (U.S. Bureau of
Reclamation).

Large Wood National Manual July 2015


vi
CONTRIBUTORS
Managing Editor
Leo D. Lentsch – ICF International

Lead Technical Editors (alphabetical)


Tim Abbe – Natural Systems Design (NSD)

David Bandrowski – U.S. Bureau of Reclamation

Jock Conyngham – U.S. Army Corps of Engineers

Doug Shields – Shields Engineering, LLC

Lead Formatting and Document Production Editors (alphabetical)


Kenneth Cherry – Editor, ICF International

Elizabeth Irvin – Editor, ICF International

Jasmin Mejia – Production Coordinator, ICF International

Chapter Authors (alphabetical)


Tim Abbe (Preamble, Chapters 1, 2, 4, 6, and 7) – Natural Systems Design

David Bandrowski (Preamble, Chapter 8) – U.S. Bureau of Reclamation

Brendan Belby (Chapters 4 and 7) – ICF International

Judsen Bruzgul (Chapter 5) – ICF International

Jock Conyngham (Preamble, Chapter 5) – U.S. Army Corps of Engineers

Chris Earle (Chapter 1) – ICF International

Gregg Ellis (Chapter 8) – ICF International

Leif Embertson (Chapters 6 and 7) – Natural Systems Design

Rocco Fiori (Chapter 8) – Fiori Geosciences

Martin Fox (Chapter 1) – Fox Environmental

Rocky Hrachovec (Chapters 6 and 8) – Natural Systems Design

Carl Jensen (Chapter 6) – ICF International

Leo Lentsch (Preamble, Chapters 1, 2, 3, 8, and 9) – ICF International

James MacBroom (Chapter 5) – Milone & MacBroom, Inc.

Katy Maher (Chapter 5) – ICF International

Large Wood National Manual July 2015


vii
Bureau of Reclamation and
U.S. Army Corps of Engineers Contributors

Rebecca Manners (Chapter 5) – University of Montana

Willis McConnaha (Chapters 3 and 9) – ICF International

Jordan Rosenfeld (Chapter 3) – British Columbia Ministry of Environment

Roy Schiff (Chapter 5) – Milone and MacBroom, Inc.

Doug Shields (Chapters 4, 6, and 8) – Shields Engineering, LLC

Tom Stewart (Chapter 3) – ICF International

C. Anna Toline (Chapter 9) – National Park Service

Dana Warren (Chapter 3) – Oregon State University

Ellen Wohl (Chapter 5) – Colorado State University

Editing and Production Support


Saadia Byram – Editor, ICF International

Kristen Lundstrom – Editor, ICF International

Laura Shields – Technical Assistant, Shields Engineering, LLC

Contracting
Joseph Pratt – Bureau of Reclamation

Mark Matthies – ICF International

Technical Reviewers
In addition to the Editors and Chapter Authors, numerous individuals took the time to review several
versions of the document and offer their useful suggestions and advice:

Paul Bakke – U.S. Fish and Wildlife Service

Bob Banard - Washington Department of Fish and Wildlife

Janine Castro – U.S. Fish and Wildlife Service

Brian Cluer – National Oceanic and Atmospheric Administration

Michelle Cramer - Washington Department of Fish and Wildlife

Todd Crowl – Florida International University: Southeast Environmental Research Center

Tracy Drury – Anchor QEA, LLC

Rocco Fiori – Fiori Geoscience

David Gaeuman – U.S. Bureau of Reclamation

Christopher Gippel - Fluvial Systems

Large Wood National Manual July 2015


viii
Bureau of Reclamation and
U.S. Army Corps of Engineers Contributors

Bob Gubernick – U.S. Forest Service, Region 9

Casey Kramer – Washington State Department of Transportation

Greg Koonce – Interfluve, Inc.

Peter Lagasse - Ayers Associates

Jim Park - Washington State Department of Transportation

Patricia Olson - Washington Department of Ecology

Roger Peters – U.S. Fish and Wildlife Service

Rob Schanz - Washington State Department of Transportation

Andrew Simon – Cardno ENTRIX

Emily Stanley – University of Wisconsin

Doug Thompson – Connecticut College

Rod Wittler – U.S. Bureau of Reclamation

Large Wood National Manual July 2015


ix
Bureau of Reclamation and
U.S. Army Corps of Engineers Contributors

This page intentionally left blank.

Large Wood National Manual July 2015


x
USER’S GUIDE
As mentioned in the preface, this national publication provides a basic understanding of the role of
wood in fluvial aquatic and riparian ecosystems and how it should be maintained, reintroduced, and/or
managed. It highlights the best available science, creative engineering, and policies associated with
restoring wood in rivers and streams (hereafter streams – see Glossary) as well as underscores the
significance of wood in fluvial ecosystems. It is also a source of practical information on how to assess
the need for wood, use wood in restoration projects, and manage wood that naturally enters streams. To
this end, this national publication provides resource managers and restoration practitioners with
comprehensive guidelines for the planning, design, placement, and maintenance of large wood in
streams with an emphasis on restoring ecosystem process and function. The document is organized into
10 chapters.

Chapter 1. Large Wood Introduction


This chapter provides an overview of the importance of wood in fluvial ecosystems as well as a
historical perspective on the use of wood in stream channels. As such, it provides a broad overview of
the use of wood in restoration projects. Main subjects include:

 Ecological Restoration – introduces the concept of ecological restoration and the key ecosystem
divisions across the United States.
 Large Wood – describes the importance of riparian forests and wood recruitment in fluvial
ecosystems.
 Ecological Functions of Wood – provides an overview of important biological and physical
functions of wood in fluvial ecosystems.
 History of Wood Management and Restoration in Streams and Rivers – provides a summary
and overview of the use of large wood in aquatic ecosystem restoration projects.

Chapter 2. Large Wood and the Fluvial Ecosystem Restoration Process


This chapter provides a general overview of the ecological restoration-planning and decision-making
process and how it applies to the overall planning and implementation of projects that use large wood to
restore process and function to fluvial aquatic ecosystems. It describes 12 important components to
consider when developing successful restoration projects. Inherent to the restoration process is the
recognition that suitable solutions may include a wide range of design elements, from simple changes in
resource management practices to major structural alterations, the selection of which depends on the
nature of each individual project. To this end, an integrated approach to the planning and
decision-making process provides the foundation for selecting and using appropriate tools and
procedures for placing wood in streams. Main subjects include:

 Ecological Restoration Process – describes ecological restoration and 12 important considerations


in the restoration process.

 Restoration Decision Making – at each step in the ecological restoration-planning process critical
decisions need to be made that will influence the outcome of a project. This section discusses
considerations for: (1) Planning Team Composition, (2) Scaling the Process, and (3) Integrating
Economics into the Restoration Process.

Large Wood National Manual July 2015


xi
Bureau of Reclamation and
U.S. Army Corps of Engineers User’s Guide

 Making Informed Restoration Decisions: A Structured Process – describes the application of


using a structured process as well as decision support tools for making informed restoration
decisions.

Chapter 3. Ecological and Biological Considerations


Restoration of large wood is often undertaken to achieve biological goals. Hence, the inherent
assumption of restoration of large wood is that habitat features in streams associated with wood are
positively related to the survival, persistence, and abundance of desired aquatic species and
communities and ecological functions. This chapter discusses the ecological and biological
considerations associated with large wood in streams. It focuses on stream ecology and the role of wood
as a biological habitat, specifically examining the role of wood in salmonid ecosystems and deriving
general principles that are applicable to other systems and species. Main subjects include:

 Ecological Functions of Large Wood – highlights the fact that large wood is a key structural
element in forested stream ecosystems worldwide. Large wood serves as a food resource for
microbes, fungi, and macroinvertebrates, As such, this section discusses the role of large wood in
habitat formation, aquatic food webs, and biogeochemical processes.

 Hyporheic Zone – this zone extends streams below the surface flow to include the “sponge” of
saturated substrate. This section describes the ecological functions associated with wood and the
hyporheic zone.

 Regional Differences in Large Wood Ecology – the biological and physical roles of large wood in
streams apply to a wide range of geographies and stream types. This section describes the
differences between geographic regions within the United States.

 Considering Assessing the Need for Wood Placement – describes important considerations in
determining the need to supplement wood in aquatic ecosystems, including: Fish Population
Dynamics and Instream Wood, Linking Habitat to Fish Population Dynamics, Fish Assemblages and
Large Wood, Wood as Habitat for Aquatic Invertebrates and Terrestrial Species, and Assessing the
Effectiveness of Wood Restoration.

 Scale and the River Continuum Concept – discusses the importance of scale as well as the river
continuum process as it relates to the placement of wood in channels.

 Key Findings and Uncertainties – summarizes and highlights key findings and uncertainties.

Chapter 4. Geomorphology and Hydrology Considerations


This chapter explores how trees and wood influence geomorphology and hydrology through such
activities as trapping sediment and organic matter, reducing rates of bank erosion, limiting long-term
rates of incision that influence valley formation, and providing habitat resilience to extreme elements.
The chapter provides an understanding of how wood can naturally influence a fluvial aquatic system
and how it can be used to restore it. The chapter also outlines areas of uncertainties and where further
research is necessary. In particular, information and descriptions regarding wood loading and longevity
in streams is lacking for many regions of the country including the Southwest, the Sierra Nevada, the
Great Plains, the lower Midwest, the South, the Mid-Atlantic, and the Alaskan Interior. A final key points
section provides a concise outline of the chapter, summarizing the geomorphic effects of wood in
streams, and the factors that influence the morphology and dynamics of a stream. Main subjects include:

Large Wood National Manual July 2015


xii
Bureau of Reclamation and
U.S. Army Corps of Engineers User’s Guide

 Geomorphology – discusses the process and factors influencing the formation and change of fluvial
geomorphology, including the flow of water through a channel network; the movement of sediment
and woody debris; the factors controlling channel form, the stability of steam beds and banks, and
the rate and magnitude to which channels move; and how large wood and logjams influence flow
conditions to alter the channels and floodplains.

 Hydrology – discusses the timing, rate, and mechanism of water movement through watersheds
and their role in the geomorphic processes for large woody material design. The section describes
how hydrological processes, namely streamflow hydrographs and flood wave dynamics, are affected
by riparian vegetation and large wood and explores the implications of this in terms of ecological
benefit and altered levels of flood protection.

 Key Findings and Uncertainties – summarizes and highlights key findings and uncertainties.

Chapter 5. Watershed-Scale and Long-Term Considerations


This chapter addresses issues of large wood supply and recruitment as well as the long-term viability of
large wood restoration projects. It discusses effects of climate change, effects of stochastic flooding and
storms on pulsed colluvial and alluvial recruitment, planning and infrastructure design for large wood
mobilization during peak flows, and the use of large wood in flood response. The chapter also outlines
areas of uncertainty and where further research is necessary, such as the transportation dynamics of
pulsed wood inputs from stochastic events, the effects of climate change on future peak flow hydrology,
and vegetative stress induced by base flow alteration resulting from climate change. A final key points
section provides a concise outline of the chapter, summarizing the capacity of a watershed system to
produce, supply, recruit, and transport large wood elements as well as the benefits in terms of stability
and habitat values resulting from large wood-based projects. Main subjects include:

 Corridor and Basin Management Concepts – explains the reasons for and the effects of the
truncation of wood supply to U.S. rivers through forest clearing and development.

 Flood Dynamics and Response – explores how forest dynamics, hillslope dynamics, river-network
dynamics, diota, and channel dynamics interact to govern the mechanisms, rates, and quantities of
wood recruitment. This section also describes the mechanisms for retention of large wood loads and
the role of floods in large wood management.

 Large Wood and River Crossing Interaction – discusses the influence of large wood on channel
equilibrium, stability, and instream habitat and how removal of large wood can negatively affect
long-term channel bed and bank stability.

 Large Wood’s Impact on Bridges and Culverts – discusses the role of wood accumulation near
bridges and culverts, leading to scour-inducing turbulence and contributing to bridge failure.

 Watershed-Scale Risk to Structures – discusses the sources of large wood pieces in rivers and
forested riverbanks and the risk of large wood blockage on bridges and culverts,

 Structure Vulnerability and Design Recommendations – discusses the vulnerability of bridges


and culverts to large wood jams and debris in relation to watershed and channel characteristics and
structure geometry.

 Floods, Recovery, and Large Wood – discusses post-flood evaluation of large wood loads within
the context of watershed wood budget of source, transport, and retention. The section discusses

Large Wood National Manual July 2015


xiii
Bureau of Reclamation and
U.S. Army Corps of Engineers User’s Guide

wood retention/removal alternatives in terms of the potential of long-term channel destabilization


and loss of habitat.

 Climate Change – discusses the pathways by which climate change may alter stream ecosystem
structure and function. The section examines how climate change will potentially impact ecological
processes related to large wood.

 Key Findings and Uncertainties – summarizes and highlights key findings and uncertainties.

Chapter 6. Engineering Considerations


This chapter provides an introduction to the engineering design of large wood placements in streams
and rivers. Large wood restoration projects require an interdisciplinary design capable of responding to
biological, physical, and social factors potentially affecting the security and long-term viability of the
structure. It explores hydraulic analysis, landscape architecture, types of structures utilized, and other
elements specific to the engineering challenges of large wood restoration projects. The chapter also
discusses areas where more information is needed and further research is necessary, for example, the
need for a basis to estimate the time required for natural regrowth of forests to sustain instream wood
levels. Specific information needs regarding wood piling size, species, and condition are also noted. A
final key points section provides a concise outline of the chapter, highlighting the role of wood in
assisting the recovery of degraded systems, the role of decay and erosion in large wood projects, and the
use and necessity of hydrologic and geomorphic modeling. Main subjects include:

 Design Life of Placed Wood – discusses the decay rates of large wood projects and the goal of
replacing natural wood sources and associated processes that will naturally replenish instream
large wood and floodplain.

 Level of Design Effort – discusses the appropriate level of effort and analysis for the design of large
wood structure projects.

 Design Decisions and Data Requirements – provides a series of data gathering and analysis
exercises to guide design decisions. Key design decisions relate to hydrology, reach layout,
materials, structure dimensions and details, hydraulics, sediment, vegetation, anchoring,
construction, and economics.

 Special Considerations or Urban Streams – provides some key parameters to consider in the
design of large wood structures in the urban environment. Considerations include extreme
modifications to water, sediment, and wood loading and the potential impacts on public
infrastructure and safety.

 Integrating Landscape Architecture – discusses large wood structure in the context of multiuse
landscape. The section discusses the interaction between the built and natural environmental in
terms of large wood structures.

 Key Findings and Uncertainties – summarizes and highlights key findings and uncertainties.

Chapter 7. Risk Considerations


This chapter provides an overview of how to assess risk when integrating wood into stream and river
restoration projects. It describes risks associated with the use of wood in stream and river restoration
projects such as loss or washout of wood placement, unintended geomorphic changes in river corridor

Large Wood National Manual July 2015


xiv
Bureau of Reclamation and
U.S. Army Corps of Engineers User’s Guide

conditions, rise in water elevation, and alteration of sediment transport. The chapter also outlines areas
of uncertainty where further research is necessary, including the need for region-specific information on
the impacts of wood removal (including channel incision resulting from in-stream wood removal) and
data on existing wood loading specific to location, size, and mobility of large wood pieces. The need for
guidelines pertaining to culvert design, wood-management following storms and floods, and legal
liability of wood placement are also noted. Main subjects include:

 Defining and Assessing Risk – discusses quantitative and qualitative approaches to assessing risk,
key elements of a risk assessment, and professional liability.

 Bridges and Culverts – discusses the role of describing downstream crossings in risk assessments.

 Key Findings and Uncertainties – summarizes and highlights key findings and uncertainties.

Chapter 8. Regulatory Compliance, Public Involvement, and Implementation


This chapter addresses the federal, state, and local regulations that control or may influence placement,
operation, and long-term operation and maintenance of large wood. It describes the regulatory
background, offers potential scenarios under which the regulations may apply, and provides potential
best management practices designers and installers should consider.

Public involvement through outreach during a large wood project may occur for several reasons. In
general, outreach will be associated with public noticing required by regulations, public outreach to
solicit design input and to build project support, and outreach to inform river users about the presence
of large wood to help ensure their long-term safety.

The chapter then turns to incorporating large wood structures into a larger aquatic or riparian
restoration project. The discussion includes grading in the project reach to accommodate a large wood
installation, implications for revegetation and irrigation system placement, erosion control, interpretive
and educational opportunities, and landscape aesthetics. Graphic standards for use in construction
document preparation are also described. The chapter also discusses areas of uncertainty and where
further research is necessary, such as guidance for the use of drones and webcams in monitoring
implementation, the development of approaches to inducing (rather than constructing) large wood
accumulations, and research into enhanced techniques for rapid revegetation of riparian zones and
floodplains. A final key points section provides a concise outline of the chapter, summarizing the issues
of contractual arrangements for procuring implementation services, maintaining a daily log as part of
implementation project management, and safety considerations. Main subjects include:

 Regulatory Compliance and Public Considerations – describes the types of federal, state, and
local regulations that control or may influence the initial placement and long-term operation and
maintenance of large wood.

 Public Involvement and Input – discusses the methods of public outreach during a large wood
project, including public noticing required by regulations, public outreach to solicit design input and
to build public support, and outreach to inform river users about the presence of large wood to help
ensure long-term safety.

 Examples of Regulatory Compliance Approaches – provides two example of regulatory


compliance approaches for large wood projects.

 Construction – examines legal issues and disputes arising from accidental injuries, cost overruns,
project failure, and construction-related risks.

Large Wood National Manual July 2015


xv
Bureau of Reclamation and
U.S. Army Corps of Engineers User’s Guide

 Safety – discusses potential safety issues, best management practices, personal protective
equipment, log handling, and other potential hazards associated with logging, construction, and
amphibious operations.

 Managing Environmental Impacts – discusses actions that may be used to minimize impacts on
water quality and ecological resources during construction of large wood projects.

 Maintenance and Adjustments – discusses maintenance, adjustment, and adaptive management


techniques that may be necessary to maintain large wood project functioning as intended.

 Key Findings and Uncertainties – summarizes and highlights key findings and uncertainties.

Chapter 9. Assessing Ecological Performance


This chapter discusses evaluation approaches to the ecological performance of large wood restoration
projects that address the uncertainty and associated risks that remain an inevitable part these projects.
The chapter identifies carefully designed evaluations and/or experiments, performance indicators, and
research designs, which would assist managers in making informed decisions regarding large wood
restoration projects. Main subjects include:

 Incorporating Best Science Practices – discusses the use of best science practices in restoration
projects, including using conceptual models and following scientific principles and guidelines.

 Measurable Outcomes and Performance Indicators – discusses the selection of appropriate


measurable outcomes and performance indicators pertaining to water quality, periphyton, aquatic
macroinvertebrates, and fish and aquatic vertebrate assemblage.

 Monitoring – discusses varieties of monitoring activities from ecosystem restoration projects,


including compliance monitoring, effectiveness monitoring, and long-term status and trend
monitoring.

 Research and Experimentation – discusses the role of research and experimentation in natural
resources management actions.

 Making Decisions and Choices – explores the role of adaptive management in restoration projects.

Chapter 10. Large Wood Bibliography


This chapter provides a bibliography of relevant scientific publications on the use of wood in stream and
river channels.

Large Wood National Manual July 2015


xvi
TABLE OF CONTENTS
List of Tables......................................................................................................................................... xxiv
List of Figures ....................................................................................................................................... xxvi
Glossary .............................................................................................................................................. xxxiii
List of Acronyms ....................................................................................................................................lxiii
List of Symbols .......................................................................................................................................lxv

Chapter 1 Large Wood Introduction ................................................................................................ 1-1


1.1 Need for and Purpose of this Manual ........................................................................................... 1-1
1.2 Ecological Restoration................................................................................................................... 1-2
1.3 Large Wood ................................................................................................................................... 1-5
1.3.1 Importance of Riparian Forests ....................................................................................... 1-6
1.3.2 Wood Loading in Natural Settings ................................................................................... 1-7
1.3.3 Historical Instream Wood Conditions ............................................................................ 1-10
1.3.4 Wood Recruitment Processes ........................................................................................ 1-11
1.3.5 Wood Management ....................................................................................................... 1-24
1.3.6 Wood Performance Standards ...................................................................................... 1-29
1.3.7 Wood Distribution within Channel Networks ................................................................ 1-30
1.3.8 Wood Longevity ............................................................................................................. 1-32
1.4 Ecological Functions of Wood ..................................................................................................... 1-34
1.4.1 Biological Functions ....................................................................................................... 1-35
1.4.2 Physical Functions .......................................................................................................... 1-38
1.5 History of the Use of Wood for Restoration in Streams ............................................................. 1-41
1.6 References .................................................................................................................................. 1-45
Chapter 2 Large Wood and the Fluvial Ecosystem Restoration Process ........................................... 2-1
Chapter 2 Large Wood and the Fluvial Ecosystem RESTORATION Process ......................................... 2-1
2.1 Introduction .................................................................................................................................. 2-1
2.2 Ecological Restoration Process ..................................................................................................... 2-2
2.2.1 Define the Problem and Develop Goals ........................................................................... 2-4
2.2.2 Assess Site Conditions ...................................................................................................... 2-7
2.2.3 Identify Opportunities and Constraints ........................................................................... 2-9
2.2.4 Define Risks and Uncertainties ...................................................................................... 2-14
2.2.5 Develop Design Considerations ..................................................................................... 2-16
2.2.6 Conduct Site Surveys ..................................................................................................... 2-18
2.2.7 Prepare and Evaluate Alternative Restoration Concepts .............................................. 2-21

Large Wood National Manual July 2015


xvii
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

2.2.8 Prepare Monitoring and Adaptive Management Plan................................................... 2-21


2.2.9 Prepare Detailed Design Plans ....................................................................................... 2-22
2.2.10 Complete Environmental and Regulatory Requirements .............................................. 2-23
2.2.11 Implement the Project ................................................................................................... 2-25
2.2.12 Monitor and Implement Adaptive Management Measures.......................................... 2-26
2.3 Restoration Decision Making ...................................................................................................... 2-26
2.3.1 Planning Team Composition .......................................................................................... 2-27
2.3.2 Scaling the Process......................................................................................................... 2-27
2.3.3 Integrating Socioeconomics into the Restoration Process ............................................ 2-28
2.3.4 Using Structured Decision Making ................................................................................. 2-34
2.4 References .................................................................................................................................. 2-40
Chapter 3 Ecological and Biological Considerations ......................................................................... 3-1
3.1 Introduction .................................................................................................................................. 3-1
3.2 Ecological Functions of Large Wood ............................................................................................. 3-2
3.2.1 Habitat Formation............................................................................................................ 3-3
3.2.2 Aquatic Food Webs .......................................................................................................... 3-6
3.2.3 Biogeochemical Functions ............................................................................................... 3-8
3.3 Hyporheic Zone ............................................................................................................................. 3-9
3.4 Regional Differences in Large Wood Ecology ............................................................................. 3-11
3.4.1 Western United States ................................................................................................... 3-11
3.4.2 Northeastern United States ........................................................................................... 3-12
3.4.3 Midwestern and Southeastern United States ............................................................... 3-12
3.4.4 Mountain West and Southwestern United States ......................................................... 3-13
3.5 Considering the Need for Wood Placement .............................................................................. 3-13
3.5.1 Fish Population Dynamics and Instream Wood ............................................................. 3-15
3.5.2 Linking Habitat to Fish Population Dynamics ................................................................ 3-16
3.5.3 Fish Assemblages and Large Wood ................................................................................ 3-21
3.5.4 Wood as Habitat for Aquatic Invertebrates and Terrestrial Species ............................. 3-22
3.5.5 Assessing the Effectiveness of Wood Restoration ......................................................... 3-22
3.6 Scale and the River Continuum Concept .................................................................................... 3-23
3.7 Uncertainties and Research Needs ............................................................................................. 3-27
3.8 Key Points.................................................................................................................................... 3-27
3.9 References .................................................................................................................................. 3-27
Chapter 4 Geomorphology and Hydrology Considerations .............................................................. 4-1
4.1 Introduction .................................................................................................................................. 4-1
4.2 Geomorphology ............................................................................................................................ 4-1
4.2.1 Wood Structures .............................................................................................................. 4-5

Large Wood National Manual July 2015


xviii
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

4.2.2 Big Trees ........................................................................................................................... 4-9


4.2.3 Hydraulic Influence of Wood ......................................................................................... 4-10
4.2.4 Channel Morphology ..................................................................................................... 4-22
4.2.5 Wood and Channel Incision ........................................................................................... 4-34
4.2.6 Wood and Bank Erosion................................................................................................. 4-46
4.2.7 Sediment Storage........................................................................................................... 4-53
4.2.8 Water Quality................................................................................................................. 4-53
4.3 Hydrology .................................................................................................................................... 4-54
4.3.1 Effects of Riparian Vegetation and Wood on Hydrology ............................................... 4-55
4.4 Uncertainties and Research Needs ............................................................................................. 4-63
4.5 Key Points.................................................................................................................................... 4-64
4.6 References .................................................................................................................................. 4-67
Chapter 5 Watershed-Scale and Long-Term Considerations ............................................................ 5-1
5.1 Introduction and Purpose ............................................................................................................. 5-1
5.2 Corridor and Basin Management Concepts .................................................................................. 5-1
5.3 Flood Dynamics and Response ..................................................................................................... 5-1
5.3.1 Pulsed Stochastic Inputs as a Large Wood Recruitment Mechanism .............................. 5-1
5.3.2 Large-Scale and Long-Term Considerations................................................................... 5-12
5.4 Large Wood and River Crossing Interaction ............................................................................... 5-16
5.5 Large Wood’s Impact on Bridges and Culverts ........................................................................... 5-17
5.5.1 National Overview ......................................................................................................... 5-17
5.6 Watershed-Scale Risk to Structures ............................................................................................ 5-19
5.6.1 System-scale and Local Large Wood Sources ................................................................ 5-19
5.6.2 Wood Transport to Bridges............................................................................................ 5-19
5.6.3 Critical Wood Size .......................................................................................................... 5-20
5.6.4 Bed Forms ...................................................................................................................... 5-20
5.6.5 Floodplain Wood ............................................................................................................ 5-20
5.6.6 Spoil Piles ....................................................................................................................... 5-21
5.7 Structure Vulnerability and Design Recommendations .............................................................. 5-22
5.7.1 Vulnerability ................................................................................................................... 5-22
5.7.2 Increasing Structure Resiliency ...................................................................................... 5-22
5.7.3 Improved Bridge and Culvert Design to Pass Large Wood ............................................ 5-22
5.7.4 Additional Bridge and Culvert Design Considerations ................................................... 5-30
5.8 Floods, Recovery, and Large Wood............................................................................................. 5-30
5.8.1 Large Wood Assessment ................................................................................................ 5-31
5.8.2 Large Wood Alternatives Analysis ................................................................................. 5-31
5.8.3 Large Wood Flood Recovery Design .............................................................................. 5-33

Large Wood National Manual July 2015


xix
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

5.9 Climate Change ........................................................................................................................... 5-34


5.9.1 Climate-Driven Processes Related to Large Wood ........................................................ 5-34
5.9.2 Recent and Future Climate Change ............................................................................... 5-35
5.9.3 Potential Climate Change Impacts on the Riverine Environment and Built
Infrastructure ................................................................................................................. 5-39
5.9.4 Large Wood Contribution to Reducing Climate Vulnerabilities in Riverine
Ecosystems and Built Infrastructure .............................................................................. 5-42
5.10 Conclusion ................................................................................................................................... 5-43
5.11 Uncertainties and Research Needs ............................................................................................. 5-44
5.12 Key Points.................................................................................................................................... 5-44
5.13 References .................................................................................................................................. 5-45
Chapter 6 Engineering Considerations ............................................................................................ 6-1
6.1 Overview ....................................................................................................................................... 6-1
6.2 Introduction .................................................................................................................................. 6-1
6.3 Area of Applicability ...................................................................................................................... 6-7
6.4 Design Life of Placed Wood .......................................................................................................... 6-9
6.5 Level of Design Effort .................................................................................................................. 6-13
6.6 Design Decisions and Data Requirements .................................................................................. 6-13
6.6.1 Hydrology ....................................................................................................................... 6-15
6.6.2 Reach Layout .................................................................................................................. 6-19
6.6.3 Select Types of Structures.............................................................................................. 6-27
6.6.4 Determine Dimensions .................................................................................................. 6-30
6.6.5 Select Wood Materials ................................................................................................... 6-32
6.6.6 Hydraulic Analysis .......................................................................................................... 6-32
6.6.7 Scour Analysis ................................................................................................................ 6-33
6.6.8 Bank Erosion .................................................................................................................. 6-34
6.6.9 Force and Moment Analysis .......................................................................................... 6-34
6.6.10 Planting Vegetation ....................................................................................................... 6-42
6.6.11 Constructability Assessment .......................................................................................... 6-43
6.7 Special Considerations for Urban Streams ................................................................................. 6-44
6.7.12 Design Discharge ............................................................................................................ 6-44
6.7.13 Floodplain Regulation .................................................................................................... 6-46
6.7.14 Existing Utilities .............................................................................................................. 6-46
6.7.15 Sediment and Debris ...................................................................................................... 6-47
6.7.16 Existing and Historic Structures ..................................................................................... 6-48
6.8 Integrating Landscape Architecture............................................................................................ 6-48
6.8.1 Landscape Integration ................................................................................................... 6-48

Large Wood National Manual July 2015


xx
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

6.8.2 Public Use Considerations ............................................................................................. 6-49


6.8.3 Graphic Standards .......................................................................................................... 6-49
6.9 Uncertainties and Research Needs ............................................................................................. 6-51
6.10 Key Points.................................................................................................................................... 6-52
6.11 References .................................................................................................................................. 6-54
Chapter 7 Risk Considerations ........................................................................................................ 7-1
7.1 Purpose ......................................................................................................................................... 7-1
7.2 Introduction .................................................................................................................................. 7-1
7.3 Defining and Assessing Risk .......................................................................................................... 7-5
7.3.1 Quantitative and Qualitative Risk Assessment ................................................................ 7-7
7.3.2 Elements of Risk Assessment ........................................................................................... 7-9
7.3.3 Professional Liability ...................................................................................................... 7-13
7.3.4 Defining Risk on Your Project ........................................................................................ 7-14
7.3.5 Reach Factors ................................................................................................................. 7-16
7.3.6 Large Wood Structure Factors ....................................................................................... 7-20
7.4 Bridges and Culverts ................................................................................................................... 7-23
7.5 Uncertainties and Research Needs ............................................................................................. 7-24
7.6 Key Points.................................................................................................................................... 7-24
7.7 References .................................................................................................................................. 7-25
Chapter 8 Regulatory Compliance, Public Involvement, and Implementation ................................. 8-1
8.1 Introduction .................................................................................................................................. 8-1
8.2 Regulatory Compliance and Public Involvement .......................................................................... 8-4
8.2.1 Federal Regulations ......................................................................................................... 8-4
8.2.2 State and Local Regulations ............................................................................................. 8-5
8.3 Public Involvement and Input ....................................................................................................... 8-8
8.4 Regulatory Compliance Approaches ............................................................................................. 8-9
8.4.1 Scenario 1: Project Site with an Endangered Species ...................................................... 8-9
8.4.2 Scenario 2: Erosion Control Project ................................................................................. 8-9
8.5 Construction.................................................................................................................................. 8-9
8.5.1 Construction Oversight .................................................................................................... 8-9
8.5.2 Water Management ...................................................................................................... 8-14
8.5.3 Excavation ...................................................................................................................... 8-17
8.5.4 Wood Placement ........................................................................................................... 8-18
8.5.5 Securing Wood ............................................................................................................... 8-21
8.5.6 Finish Work .................................................................................................................... 8-25
8.5.7 Typical Construction Equipment .................................................................................... 8-26
8.6 Safety .......................................................................................................................................... 8-36

Large Wood National Manual July 2015


xxi
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

8.6.1 Potential Safety Issues ................................................................................................... 8-36


8.6.2 Potential Best Management Practices ........................................................................... 8-37
8.6.3 Personal Protective Equipment ..................................................................................... 8-39
8.6.4 Log Handling .................................................................................................................. 8-39
8.6.5 Excavation and Earth Moving ........................................................................................ 8-39
8.6.6 Helicopters ..................................................................................................................... 8-41
8.6.7 Chainsaw Operation....................................................................................................... 8-41
8.7 Managing Environmental Impacts .............................................................................................. 8-42
8.7.1 Water Quality................................................................................................................. 8-42
8.7.2 Fish Exclusion ................................................................................................................. 8-42
8.7.3 Cultural Resources ......................................................................................................... 8-43
8.7.4 Noise .............................................................................................................................. 8-44
8.8 Maintenance and Adjustments................................................................................................... 8-45
8.8.1 Three Types of Maintenance ......................................................................................... 8-45
8.8.2 Maintenance Activities .................................................................................................. 8-46
8.8.3 Adjustments Based on Monitoring and Adaptive Management ................................... 8-46
8.9 Acknowledgments....................................................................................................................... 8-47
8.10 Uncertainties and Research Needs ............................................................................................. 8-47
8.11 Key Points.................................................................................................................................... 8-47
8.12 References .................................................................................................................................. 8-48
Chapter 9 Assessing Ecological Performance ................................................................................... 9-1
9.1 Introduction .................................................................................................................................. 9-1
9.2 Incorporating Best Science Practices ............................................................................................ 9-2
9.2.1 Using Best Available Knowledge ...................................................................................... 9-2
9.2.2 Using Conceptual Models ................................................................................................ 9-3
9.2.3 Following Scientific Principles and Guidelines ................................................................. 9-4
9.2.4 Existing Protocols and Indices.......................................................................................... 9-4
9.3 Measurable Outcomes and Performance Indicators .................................................................... 9-5
9.3.1 Water Quality................................................................................................................... 9-6
9.3.2 Periphyton ....................................................................................................................... 9-7
9.3.3 Aquatic Macroinvertebrates ............................................................................................ 9-7
9.3.4 Fish and Aquatic Vertebrate Assemblage ........................................................................ 9-8
9.4 Monitoring .................................................................................................................................... 9-8
9.4.1 Compliance Monitoring ................................................................................................... 9-8
9.4.2 Effectiveness Monitoring ................................................................................................. 9-9
9.4.3 Long-Term Status and Trend Monitoring ...................................................................... 9-10
9.4.4 Collect, Analyze, Synthesize, and Evaluate Data ........................................................... 9-10

Large Wood National Manual July 2015


xxii
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

9.5 Research and Experimentation ................................................................................................... 9-10


9.5.1 Research......................................................................................................................... 9-11
9.5.2 Before-and-After Studies ............................................................................................... 9-12
9.5.3 Pilot Projects .................................................................................................................. 9-12
9.6 Making Decisions and Choices .................................................................................................... 9-13
9.7 References .................................................................................................................................. 9-17
Chapter 10 Large Wood Bibliography.............................................................................................. 10-1

Appendix A Sample Implementation Contracts


A-1 Types of Federal Contracts Useful for Large Wood Projects ....................................................... A-1
A-2 Sample Documents for Hybrid Contracts .................................................................................... A-3
A-3 Sample Contract Language for Separate Harvest and Hauling Contract ..................................... A-9
A-4 Example—Safety and Health Provisions for Large Wood Placement Contracts ....................... A-10

Large Wood National Manual July 2015


xxiii
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

LIST OF TABLES
1-1 Distributions of Wood ................................................................................................................. 1-28

1-2 Minimum Wood Piece Volume Required to Qualify as a Key Piece (by Bankfull Width
Class) ....................................................................................................................................... 1-29

2-1 Steps in Structured Decision Making .......................................................................................... 2-36

4-1 Channel Reach Classification ...................................................................................................... 4-22

5-1 Culvert Failure Data .................................................................................................................... 5-18

5-2 Debris Countermeasures for Culverts and Bridges ..................................................................... 5-23

5-3 Large Wood Removal Recommendations................................................................................... 5-32

6-1 Limitations on Applicability of Large Wood Structures ................................................................ 6-8

6-2 Comparison of Desirability of Various Tree Species for Stream Structures ............................... 6-11

6-3 Levels of Design Effort for Instream Large Wood Structures ..................................................... 6-14

6-4 Key Engineering Issues for Instream Large Wood Structure Placement .................................... 6-15

6-5 Recommendations for Placement of Large Wood in Streams for Aquatic Habitat
Benefits ....................................................................................................................................... 6-20

6-6 Criteria for Spacing Intermittent Large Wood Structures along the Outside of
Meander Bends ........................................................................................................................... 6-22

6-7 Classification of Large Wood Instream Structures Based on Architecture ................................. 6-28

7-1 Important Project Characteristics Defining Existing Conditions and Geomorphic


Setting ....................................................................................................................................... 7-11

7-2 Important Elements for Consideration in Risk Assessment ........................................................ 7-12

7-3 Relative Risk of Instream Wood to Recreational River Users ..................................................... 7-22

8-1 Large Wood Regulatory Compliance Decision Analysis ................................................................ 8-6

8-2 Size Categories for Large Wood .................................................................................................. 8-12

8-3 Configurations for Instream Large Wood Placement ................................................................. 8-19

8-4 Comparison of Methods for Securing Instream Large Wood ..................................................... 8-21

8-5 Ballast Materials for Instream Large Wood Structures .............................................................. 8-24

8-6 Examples of Heavy Equipment Used in Large Wood Installation Including Machine
and Lift Weights as Appropriate ................................................................................................. 8-29

Large Wood National Manual July 2015


xxiv
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

8-7 Comparison of Pile-Driving Methods .......................................................................................... 8-33

8-8 Personal Protective Equipment and Attire for Large Wood Project Implementation ................ 8-40

Large Wood National Manual July 2015


xxv
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

LIST OF FIGURES
1-1 Map of Ecosystem Divisions, Regions, and Providences Across North America .......................... 1-3

1-2 Schematic of a Fluvial Aquatic Ecosystem .................................................................................... 1-4

1-3 Wood Loading Tends to Increase With Channel Size When Normalized to Bankfull
Width ......................................................................................................................................... 1-8

1-4 Example of High Wood Loading in a Large Channel (Nooksack River Delta, Northwest
Washington) .................................................................................................................................. 1-9

1-5 North American Forests .............................................................................................................. 1-16

1-6 The Median Instream Large Wood Volume (A) and Number of Pieces (B) According
to Adjacent Riparian Stand Age Class, at the Time of 1999–2000 Surveys ................................ 1-21

1-7 Wood Loading In Streams Throughout the United States and Other Regions Typically
Range from 1 to 2,000 Megagrams per Hectare ........................................................................ 1-27

1-8 The Percent Distribution of Large Wood to Group Size Class According to Five
Bankfull Width Classes ................................................................................................................ 1-31

1-9 Comparison of the Mean Percent Large Wood Volume by Four Lateral Zone
Distributions ................................................................................................................................ 1-31

1-10 (A) Example of Decay Curves for Three Common Pacific Northwest Tree Species; (B)
Example of Ancient Logjam More than 120 Years Old Exposed in the Right Bank of
South Fork Nooksack River, Washington .................................................................................... 1-33

1-11 Naturally Occurring Snag Embedded in Channel Thalweg, Androscoggin River near
Bethel, Maine .............................................................................................................................. 1-34

1-12 Lush Riparian Areas Even Occur in Arid Regions Where They Deliver Wood to
Streams, North Central Oregon .................................................................................................. 1-35

1-13 Large Trees Can Play a Major Role in the Morphology of Rivers, Such as this 2.4-
Meter Douglas Fir Across Carbon River, Washington ................................................................. 1-38

1-14 Logjam Deflecting the Hoh River in Northwest Washington ...................................................... 1-39

1-15 Relationship Between Large Logs (>30 centimeters) and Debris Dams in Adirondack
Streams with Bankfull Widths of 2 to 16 Meters, Northern New York ...................................... 1-41

1-16 Removal of Wood Leads to Channel Incision, Converting Alluvial Pool-Riffle Channels
to Bedrock and Damaging Habitat and Infrastructure, Such as this Bridge Failure in
the Mashel River, Western Washington ..................................................................................... 1-41

1-17 Stable Wood Bifurcates Flow Leading to Anabranching Channels when Undisturbed,
and Creates a Complex and Productive Habitat ......................................................................... 1-44

Large Wood National Manual July 2015


xxvi
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

1-18 A Buried Log More than 500 Years Old Forming Grade Control, Coal Creek, 2004,
Ozette River Tributary, Washington ........................................................................................... 1-44

2-1 Phases and Considerations Associated with Ecological Restoration Projects Using
Large Wood ................................................................................................................................... 2-3

2-2 Gravel Patch on Incising Bedrock Channel, Rickreall Creek, Oregon .......................................... 2-11

2-3 Woodward Creek Pipeline Crossing Wood Placement, Washington.......................................... 2-11

2-4 Eroding River Bank, Nisqually River, Washington ....................................................................... 2-12

2-5 Public Meeting ............................................................................................................................ 2-12

2-6 A Warning Sign on Wood Placement, South Fork Nooksack River, Washington ....................... 2-13

2-7 Excavation and Dewatering During Construction of an Engineered Logjam in Elwha


River, Washington ....................................................................................................................... 2-13

2-8 Recession of Honeycomb Glacier in North Cascades of Washington is an Example of


how Warming Climate Affects Hydrology................................................................................... 2-14

2-9 Bridge Improvements Done to Improve Wood Conveyance as Part of a Stream


Restoration Project ..................................................................................................................... 2-15

2-10 The Structured Decision Making Process ................................................................................... 2-35

3-1 Features of a Beverton-Holt Production Function ...................................................................... 3-16

4-1 Although Precipitation Increases Surficial Runoff, Erosion Rates Diminish (as
measured by sediment yield) due to the Influence of Vegetation ............................................... 4-6

4-2 Illustration of Several Basic Fundamentals of Fluvial Geomorphology, including


Spatial and Temporal Change over Time, the Importance of Sediment Budgets, and
the Role of Wood .......................................................................................................................... 4-7

4-3 (A) General Distribution of Natural Wood Accumulation Types Within a Watershed;
(B) Application of Four of Those Types to Engineered Logjam Structures ................................... 4-8

4-4 Wood is Typically the Largest Bed Material Entering Streams and Tends to Get
Larger in Lower Elevations of a Watershed (Larger Channels), the Inverse of Rock
Particles ....................................................................................................................................... 4-11

4-5 Big Trees Were Historically Common Along Streams Throughout the United States ................ 4-12

4-6 (A) Snags and Logjams, Were Common Throughout Much of the Missouri and other
Midwestern Rivers, as Depicted in this Illustration by George Catlin in 1832; (B)
Undated Photo, Circa Early 1900s, of a River on the Olympic Peninsula of
Washington Loaded with Sitka Spruce (Picea sitchensis) Snags ................................................. 4-13

Large Wood National Manual July 2015


xxvii
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

4-7 (A) Historic Changes to the Upper Willamette Transforming the Natural
Anabranching Morphology into a Single-Thread Channel; (B) Lower Taiya River, a
Wood-Rich Anabranching River in Southeastern Alaska ............................................................ 4-14

4-8 Comparison of an Alluvial River with Wood (Hoh River, Washington) to One Where
Wood Has Been Removed (Cowlitz River, Washington) ............................................................. 4-15

4-9 Flow Around a Stable Snag ......................................................................................................... 4-17

4-10 Process by Which a Snag Becomes Imbedded in a Channel Bed ................................................ 4-18

4-11 Natural Log Steps Influencing Water Elevations and Distribution of Shear Stress in
Fisher Creek in the North Cascades, Washington ....................................................................... 4-21

4-12 Examples of Alluvial (Gravel-Bed) Stream Channels With Low Wood Loading (A) and
High Wood Loading (B) ............................................................................................................... 4-23

4-13 (A) Correlation Between Percent of Large Wood Pools (with residual depth > 0.5
meter [1.6 feet]) Formed by Wood as a Function of Riparian Forest Stand Age; (B)
Frequency of Textural Patches as a Function of Wood Pieces per Reach for Streams
Draining the West Slope of Olympic Mountains in Northwestern Washington......................... 4-24

4-14 (A) Threshold of Effective Wood Loading Based on Pool Frequency as a Function of
Wood Loading per Square Meter of Channel Bed; (B) Size of Functional Wood in
Queets River Basin ...................................................................................................................... 4-27

4-15 Conceptual Illustration of How Wood Introduces Physical Complexity to a Simplified


Channel ....................................................................................................................................... 4-28

4-16 (A) Role of Natural Logjams in Reducing the Radius of Curvature of Channel
Meanders in the Queets River, Washington; (B) Based on Assumptions for Channel
Sizing Relative to Drainage Area, the Super Elevation Associated with Smaller Radii
of Curvature Results in an Increase in Water Elevations of 0.35–1.0 meters (1.1–3.3
feet), Demonstrating Another Way Logjams Increase Floodplain Connectivity and
Drive Side Channel Formation .................................................................................................... 4-29

4-17 Wood Forces Channel Complexity Such as Anabranching (a); the Removal of Wood
Can Transform These Multi-Thread Systems Into a Wide Single-Thread Channel (b);
Observations of the Upper Cowlitz River in Washington Show the Loss of Vegetated
Island Coincident With Increasing Channel Width (c) ................................................................ 4-30

4-18 Geomorphic Changes in Lower Elwha River, Washington, Associated with ELJ
Placement ................................................................................................................................... 4-31

4-19 Predicting Channel Planform Morphology Based on Formative Discharge (Q*),


Median Grain Size (D50), and Channel Slope ............................................................................. 4-32

Large Wood National Manual July 2015


xxviii
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

4-20 Illustration From White River in Western Washington Showing the Difference in
Cumulative Bank Length (2x channel length) for Unconfined Anabranching Reach
With Numerous Logjams Versus Confined Reaches ................................................................... 4-32

4-21 (A) Hydrograph Showing the Influence of a Large Channel Spanning Logjam in the
Deschutes River, South of Olympia, Washington; (B) Hysteresis Curve Showing How
the Logjam Has the Most Significant Effect on Head (Dz) During Rising Limb of
Hydrograph ................................................................................................................................. 4-33

4-22 (A) Dimensionless Plot of How Wood Obstructing 80% of the Ozette River,
Washington, Increases Water Elevations Using a 1D Hydraulic Model; (B) Channel
Spanning Logjam on Upper Yakima River, West of Easton, Kittitas County,
Washington ................................................................................................................................. 4-34

4-23 Wood in Steep (S=0.18) Headwater Channel of Olympic Peninsula, Washington ..................... 4-34

4-24 Log Strength Can Be Critical in Headwater Channels Where They Are Subjected to
Severe Forces Imposed by Debris Flows ..................................................................................... 4-37

4-25 Wood Stores Sediment thus Reducing Sediment Transport Capacity by Obstructing
Flow and Increasing Roughness, Thereby Increasing Sediment Storage Within a
Channel ....................................................................................................................................... 4-38

4-26 (A) Wood in Taylor Creek (Seattle) Is Trapping Sediment and Dissipating Flood
Energy; (B) Coal Creek in Nearby Bellevue also Experienced Increased Peak Flows
due to Urbanization but Was also Historically Cleared and Lacks Mature Riparian
Conditions and Is Undergoing Incision ....................................................................................... 4-39

4-27 Historic Channel Incision in the South Fork Nooksack River, Washington ................................. 4-40

4-28 Eroding Bank Along the Hoh River, Washington, Showing a Snag Pointing in Flow
Direction of a Relic Channel With its Invert Perched Over 2.4 Meters (8 Feet) Above
the Current River Bed ................................................................................................................. 4-41

4-29 A Single 2.5-Meter (8.2-Foot) Diameter Old Growth Douglas Fir (Pseudotsuga
menziesii) Impounding the Carbon River in Mt. Rainier National Park, Washington ................. 4-41

4-30 Conceptual Channel Evolution Model of Stream Experiencing Incision due to


Channelization ............................................................................................................................ 4-42

4-31 Channel Incision Poses a Serious Threat to Infrastructure Such as Pipelines, Bridge
Abutments and Piers, Water Intakes, and Road Grades ............................................................ 4-43

4-32 Geomorphologists Offer Direct Design Input on the Role of Wood and Bed Material
on Channel Morphology that Is Essential in Stream Restoration and Providing
Sustainable Solutions for Protecting Infrastructure ................................................................... 4-44

4-33 Natural Logjam Influence on Channel Aggradation and Terrace Construction in 4th
Order Alta Creek (A) and 6th Order Mainstem Queets River (B) ............................................... 4-45

Large Wood National Manual July 2015


xxix
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

4-34 (A) Forest Areas With Larger Trees Erode More Slowly Than Areas With Smaller
Trees Along the Hoh and Queets Rivers; (B) Breaking Data Into Two Categories
Greater and Less Than 53 Centimeters (21 Inches), There Is a Statistically Significant
Difference, With Areas With Larger Trees Eroding at less than Half the Rate of
Smaller Trees .............................................................................................................................. 4-47

4-35 (A) Erosion Into Mature Forests Along the Hoh River Recruits Large Snags That Form
Stable Obstructions (Key Pieces) in the Channel That Slow Erosion Rates; (B) Areas of
Industrial Forest or Agriculture (Trees Less Than 21 Inches) Erode at Over Twice the
Rate 4-48

4-36 Illustration of How Large Wood Influenced Channel Process and Morphology on the
South Fork Hoh River, Washington, from 1993 to 2013 ............................................................. 4-49

4-37 Clearing Mature Riparian Forests Eliminates Functional Wood Recruitment and
Alters Processes and Channel Form ............................................................................................ 4-50

4-38 Outside Olympic National Park Almost All Old Growth Forest Within the River Valley
and Adjacent Hillslopes Has Been Cut ........................................................................................ 4-51

4-39 Flow Velocity Fields Around Two Bends of the Lower Wabash River, Illinois ............................ 4-51

4-40 Illustration of How Rougher Banks Reduce River Velocities Near the Bank ............................... 4-52

4-41 Conceptual Diagram of the Effect of Riparian Vegetation on Discharge at the Scale of
a Plant, a Cross-Section, a Reach, and a Catchment ................................................................... 4-56

4-42 Sample of Simulated Waves Computed for Different Channel Shapes ...................................... 4-58

5-1 Influences on Wood Recruitment to River Corridors ................................................................... 5-2

5-2 Examples of Protruding Boulders Helping to Trap Wood Along Streams..................................... 5-7

5-3 Wood Deposited along the Top of Bank at the Outside of a Meander Bend on the
Dall River in Central Alaska ........................................................................................................... 5-8

5-4 Conceptual Illustration of Downstream Trends in Total Wood Load and Logjams
along a River Network................................................................................................................... 5-9

5-5 Impact that Reoperation of Dams, to Include More Natural Elements of the
Hydrograph, Can Have on a Riparian Ecosystem ........................................................................ 5-14

5-6 Conceptual Illustration of Wood-Related Feedback ................................................................... 5-16

5-7 Geomorphic Engineering Structure Sizing Method .................................................................... 5-28

5-8 Schematic of Culvert Performance at Varying Stages and Alignments ...................................... 5-29

5-9 Pathways by Which Climate Change May Alter Stream Ecosystem Structure and
Function ...................................................................................................................................... 5-35

5-10 Projected Temperature Change by 2071–2099 .......................................................................... 5-37

Large Wood National Manual July 2015


xxx
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

5-11 Streamflow Projections for River Basins in the Western United States ..................................... 5-38

5-12 Changes in Timing of Streamflow from Snowmelt ..................................................................... 5-39

5-13 Occurrence Probability of Trout Species as a Function of Air Temperature and


Winter High Flow Frequency ...................................................................................................... 5-40

5-14 Wood Inhibiting the Flow of Water Through a Culvert under Highway 4 Following
the Las Conchas, New Mexico Fire (2011) .................................................................................. 5-42

6-1 Impact of Spatial Scale and Relative Risk on Engineering Aspects of a Large Wood
Project ......................................................................................................................................... 6-2

6-2 Examples of Wood Placements Used to Stabilize River Banks ..................................................... 6-3

6-3 Climate Index for Wood Decay Hazard ....................................................................................... 6-10

6-4 Graphical Output of Mean Daily Flow Monthly Exceedance Analysis and Project-
Specific Salmonid Life Stages ...................................................................................................... 6-19

6-5 Large Wood Bed-Control Structures ........................................................................................... 6-21

6-6 Continuous and Intermittent, Spur-Type Large Wood Structures.............................................. 6-22

6-7 ELJ Spacing to Protect Road and Enhance Habitat Along the Cispus River, Gifford
Pinchot National Forest, Washington ......................................................................................... 6-23

6-8 Example of the Use of Two Sets of ELJs in Bank Protection Along Hoh River ............................ 6-24

6-9 Recommended Extent of Riprap Revetment for 110 o Bend ..................................................... 6-24

6-10 Valley Scale Restoration Approach to Limiting Bank Erosion Along Valley Margins .................. 6-25

6-11 Upper Quinault River Valley Floodplain and Side Channel Restoration ..................................... 6-26

6-12 Definition Sketch for Large Wood Geometric Variables ............................................................. 6-30

6-13 Typical Free Body Diagram for a Large Wood Structure ............................................................ 6-35

6-14 Entanglement of Logs in Riparian Stumps and Boles for Passive Restraint, Hylebos
Creek, Milton, Washington ......................................................................................................... 6-38

6-15 Definition Sketch for Derivation of Geotechnical Forces on a Horizontally Embedded


Log ....................................................................................................................................... 6-41

6-16 Construction of Temporary Ramp for Access to Channel for Large Wood Structure
Construction in Little Topashaw Creek, Mississippi .................................................................... 6-43

6-17 Examples of Large Wood Projects in Urban Settings of the Pacific Northwest .......................... 6-45

6-18 Large Wood Structure Graphic Standards .................................................................................. 6-50

7-1 Natural Logjam on Long Tom Creek near Venata, Oregon ........................................................... 7-2

Large Wood National Manual July 2015


xxxi
Bureau of Reclamation and
U.S. Army Corps of Engineers Contents

7-2 Scour Undermining Downstream Corner of an ELJ on Upper Quinault, Washington .................. 7-4

7-3 RiverRAT Screening Matrix ........................................................................................................... 7-8

7-4 Relative Quantity of Wood Within a Reach, the Subset with High Geomorphic and
Habitat Benefits, and the Subset that Causes Public Safety Concerns ....................................... 7-16

7-5 Risk Assessment Chart ................................................................................................................ 7-17

7-6 Natural Wood Accumulation in Idaho ........................................................................................ 7-18

7-7 Egress and Portage...................................................................................................................... 7-21

8-1 Use of Locally Sourced Large Wood ............................................................................................ 8-11

8-2 Equipment Useful for Handling Large Wood .............................................................................. 8-13

8-3 Diversion of a Small Stream Around a Construction Zone Through Plastic Pipe ........................ 8-14

8-4 Water Management Techniques for Large Wood Projects ........................................................ 8-15

8-5 Examples of Temporary Bridges Constructed for Large Wood Projects..................................... 8-17

8-6 Minimal Excavation for Placing First Layer of Large Wood ........................................................ 8-17

8-7 Pile of Slash Available for Use in Large Wood Project ................................................................ 8-20

8-8 Planting Willow Cuttings in Recent Sediment Deposits Adjacent to Placed Large
Wood Using Water Jetting .......................................................................................................... 8-26

8-9 Manual Labor Team Stockpiling Large Wood Prior to Stream Installation................................. 8-27

8-10 Belgian Draft Horses Moving Large Wood for Instream Placement ........................................... 8-27

8-11 Cable Yarding Large Wood for Transport to Channel ................................................................. 8-27

8-12 Sequence for Constructing Large Wood Structure with Vertically Driven Piles Used to
Secure the Structure ................................................................................................................... 8-34

8-13 Use of Helicopter to Transport Large Wood to Remote Project Site ......................................... 8-35

8-14 Log Skidder Mired in an Isolated Deposit of Highly Plastic Clay in a Stream Bed....................... 8-41

8-15 Construction Laborers Work to Secure Fabrics Around Large Wood Toe Placements
on the Outside of a Meander Bend in a Shallow Channel .......................................................... 8-41

9-1 Key Components of an Adaptive Management Framework ...................................................... 9-14

Large Wood National Manual July 2015


xxxii
GLOSSARY
Adaptive capacity An asset or resource’s ability to adjust and cope with existing climate
variability or future climate impacts.
Adaptive management An approach to management that addresses changing site and
project conditions, as well as taking into account new knowledge; a
management approach that incorporates monitoring of project
outcomes and uses the monitoring results to make revisions and
refinements to ongoing management and operations actions.
Adfluvial fish Species that hatch in rivers or streams, migrate to lakes as juveniles
to grow, and return to rivers or streams to spawn.
Aggradation Long-term sediment deposition that occurs on the bed of a channel;
the opposite is degradation or bed erosion.
Alignment Planform of a channel.
Allowable shear stress A threshold channel design technique whereby channel dimensions
design method are selected so that the average applied grain bed shear stress is less
than the allowable shear stress for the boundary material.
Allowable velocity The greatest mean velocity that will not cause a channel boundary to
erode.
Allowable velocity design A threshold channel design technique whereby channel dimensions
method are selected so that the applied velocity during design conditions is
less than the limiting velocity of the channel boundary.
Alluvial channel Streams and channels that have bed and banks formed of material
transported by the stream under present flow conditions. There is
an exchange of material between the inflowing sediment load and
the bed and banks of an alluvial channel.
Alluvial channel design A design approach whereby a channel configuration is selected so
that it is in balance with the inflowing sediment and water
discharges.
Alluvium Loose, unconsolidated (not cemented together into a solid rock) soil
or sediments, which has been eroded, reshaped by water in some
form, and redeposited in a non-marine setting; typically made up of
a variety of materials, including fine particles of silt and clay and
larger particles of sand and gravel. When this loose alluvial material
is deposited or cemented into a lithological unit, or lithified, it is
called an alluvial deposit.
Amphidromous fish Species that move between fresh and salt water during some part of
their life cycle, but not for breeding.
Anabranching channel A stream or river that has two or more channels at bankfull or
effective discharge flow. Unlike braided channels, anabranching
channels are separated by vegetated islands. While a braided
channel becomes a single wide channel at bankfull flow,
anabranching channels still retain multiple channels. Generally this
term is synonymous with anastomosing.

Large Wood National Manual July 2015


xxxiii
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Anadromous fish Species that incubate and hatch in freshwater, migrate to saltwater
as juveniles to grow, and return to freshwater as adults to spawn.
Analogy design method A design approach that is based on the premise that conditions in a
reference reach with similar characteristics and watershed
conditions can be copied or adapted to the project reach.
Analytical design method The use of bed resistance and sediment transport equations to
calculate channel design variables.
Anastomosed channels See Anabranching.
Annual duration gage The analysis of the recorded peak flow values that have occurred for
analysis each year in the duration of interest; typically used for the estimate
of flows with return intervals in excess of 2 years.
Annual flood The highest peak discharge that can be expected to occur on average
in a given year.
Anoxic Depleted of dissolved oxygen.
Anthropogenic constraints Constraints on a stream or river that are caused by human (i.e.,
anthropogenic) activities or constructed projects.
Areal sediment sampling See Surface sediment sampling.
Arid An area that generally has insufficient rainfall to support
conventional agriculture without supplemental irrigation.
Armor layer A streambed containing at least some sediment that is too large to
be transported by the hydraulic flow conditions; finer particles are
selectively removed leaving a layer of coarser materials.
Armor layer (sampling) Technique used to sample the upper layer of coarse surface layer
material.
Articulating concrete block A matrix of interconnected concrete block units installed to provide
(ACB) an erosion-resistant revetment for streams and rivers.
Asymptote In a curve, an asymptote is a line such that the distance between the
curve and the line approaches zero as they tend to infinity.
Attenuation The subsidence or flattening of a floodwave as it moves down the
channel.
Avulsions The rapid abandonment of, and formation of a new, river channel;
occur when bank erosion and longitudinal adjustment occur at a
large scale; typically characterized by rapid changes in channel
planform.
Band-aid solution Treatment techniques used to address small, local issues.
Bank zone The area above the toe zone, located between the average water
level and the bankfull discharge elevation.
Bankfull The water level, or stage, at which a stream, river, or lake is at the
top of its banks and any further rise would result in water moving
into the flood plain.
Bankfull depth The distance from the deepest part of the channel to the bankfull
elevation line, typically measured across a straight section (riffle) of
a channel.

Large Wood National Manual July 2015


xxxiv
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Bankfull discharge Used as a surrogate for channel-forming discharge, defined, in part,


by the visual identification of morphological bankfull indices, such as
abrupt changes in bank angle or the presence of perennial plants.
Bankfull indices Field indicators of bankfull discharge.
Bankfull width The width of channel at bankfull elevation.
Bankline migration The adjustment of planform in natural meandering channels.
Bar apex jam Wood structure composed of 10–30 logs placed in the middle of the
channel to initiate bar formation or placed on the upstream end of
an existing bar or island.
Barb A type of flow deflection structure (see Stream barb).
Base flood The flood having a 1% chance of being equaled or exceeded in any
given year.
Base flow See Low flow.
Batter pile An inclined pile that can provide downward resistance to buoyancy
when used with another (“A” frame) or with a vertical pile.
Bed control structure A type of grade control structure that is designed to provide a hard
point in the streambed that is capable of resisting the erosive forces
of the stream.
Bed zone The bottom of the channel.
Bedding layer See Filter layer.
Bedform scour Vertical channel bed movement that results from the troughs
between crests of the bedforms.
Bedrock A solid rock on the face of or beneath the Earth’s surface.
Bend scour Bed erosion along the outside of a river or streambed.
Bendway weirs A flow-changing bank stabilization technique used to protect and
stabilize stream and river banks. Flows are directed over the weir
perpendicular to the angle of the weir.
Benthic zone The ecological region at the lowest level of a body of water such as
an ocean or a lake, including the sediment surface and some sub-
surface layers. Organisms living in this zone are called benthos.
Biofilm Bacteria, fungi, and often algae that grow on submerged woody
surfaces.
Biogeomorphology The study of interactions between organisms and the development
of landforms, and are thus fields of study within geomorphology and
ichnology. Organisms affect geomorphic processes in a variety of
ways. For example, trees can reduce landslide potential where their
roots penetrate to underlying rock, and plants and their litter inhibit
soil erosion.
Biota The plants and animals of a region.

Large Wood National Manual July 2015


xxxv
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Blockage coefficient Typically expressed as percentage of effective flow area (for design
discharge) obstructed by a structure such as wood or logjam. Can be
expressed in terms of width (structure width/channel width) or area
(structure x-sectional area/channel cross-sectional area), with
structure measured orthogonal to flow.
Braided streams Wide shallow channels with multiple unvegetated bars. At low flows
they have multiple channel threads, but at a bankfull or effective
discharge the bars are submerged and flows coalesce to form a
single channel. Braided channels form in areas with high sediment
supplies and relatively steep gradients, such as downstream of
alpine glaciers. The multiple channels of braided streams tend to be
shallow and wide as opposed to the narrow and deep channels of an
anabranching or anastomosing channel.
Branch packing A soil bioengineering technique used to fill localized slumps and
gullies; involves the use of alternating layers of live cuttings and soil.
Bridge pier scour Erosion of a streambed around the piers of bridges.
Brush layering A soil bioengineering technique that provides protection against
surface erosion and shallow-seated slope failure; involves the use of
alternating layers of live cuttings and soil.
Brush mattress A streambank soil bioengineering technique that includes a layer of
live cuttings placed flat against the sloped face of the bank.
Brush revetments A soil bioengineering technique used to stabilize streambanks. Brush
and tree revetments are nonsprouting shrubs or trees installed along
the toe of the streambank to provide bank erosion protection and to
capture sediments.
Brush spur A long, box-like structure of brush that extends from within the bank
into the streambed; functions very similarly to stone stream barbs.
Brush trench A soil bioengineering technique that inserts a row of live cuttings
into a trench along the top of an eroding streambank, parallel to a
stream. The live cuttings form a fence that filters runoff and reduces
the likelihood of drilling.
Brush wattle fence See Wattle.
Bulk sediment sampling See Volumetric sediment sampling.
Burst swimming speed The highest swimming speed of a fish; generally lasts less than 20
seconds and ends in extreme fatigue.
Cable Steel aircraft cable or wire rope used to secure large wood.
Catadromous fish Species that hatch in saltwater, migrate to freshwater as juveniles to
grow, and return to saltwater to spawn.
Catchment See Drainage area.
Celerity The speed that a floodwave moves down the channel.

Large Wood National Manual July 2015


xxxvi
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Channel Convergent topography where water is conveyed either all year


(perennial) or seasonally (ephemeral). The principal part of all
streams and rivers. Channel features include bars and bedforms.
Unconfined channels include floodplains as opposed to confined
channels which do not. Channels can be alluvial or bedrock. Channel
types are defined by morphologic characteristics, bed and bank
materials, and influence of vegetation. See Classification.
Channel alignment design Techniques used to establish a stable channel planform.
Channel classification See Classification.
Channel evolution Systematic changes of a stream channel to a perturbation.
Channel evolution model A model that illustrates the stages through which a stream
(CEM) progresses when subjected to destabilizing influences.
Channel evolution model A classification system that provides a predictable sequence of
classification change in a disturbed channel system.
Channel-forming discharge Concept based on the idea that for a given alluvial stream, there
exists a single or range of discharge that, given enough time, would
produce the width, depth, and slope equivalent to those produced
by the natural flow in the stream. This discharge, therefore,
dominates channel form and process.
Channel incision The process of downcutting into a stream channel, leading to a
decrease in the channel bed elevation. Incision is often caused by a
decrease in sediment supply and/or an increase in sediment
transport capacity. A decrease in base level can cause headcutting
that migrates upstream and produces incision upstream and
initiating aggradation downstream.
Channel morphology The shapes of river channels and how they change over time. The
morphology of a river channel is a function of a number of processes
and environmental conditions, including the composition and
erodibility of the bed and banks (e.g., sand, clay, bedrock);
vegetation and the rate of plant growth; the availability of sediment;
the size and composition of the sediment moving through the
channel; the rate of sediment transport through the channel and the
rate of deposition on the floodplain, banks, bars, and bed; and
regional aggradation or degradation due to subsidence or uplift.
Channel slope The average slope of the longitudinal thalweg profile.
Channel stage classification A stream classification system based on the channel evolution
model.
Channel stages See Channel evolution model.
Channel storage Water that is temporarily stored in a natural or constructed channel
while en route to an outlet.
Channelization The alteration of an existing river or stream for a specific physical,
biologic, or aesthetic purpose, generally involving the removal of
meander bends to straighten the flow path and increase bed slope
to increase channel conveyance.

Large Wood National Manual July 2015


xxxvii
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Check dam A small dam constructed to slow stream velocity and/or prevent
degradation.
Classification The categorization of a stream reach into a specific class based on
factors and measurements such as dominant mode of sediment
transport, entrenchment ratio, and sinuosity. Streams can also be
classified by their biota, habitat conditions, baseflow levels, and
direct measures of water quality.
Clear water scour Occurs when there is insignificant transport of bed-material
sediment from the upstream into the contracted section.
Climate change A change in the statistical distribution of weather patterns when that
change lasts for an extended period of time (i.e., decades to millions
of years). Climate change may refer to a change in average weather
conditions, or in the time variation of weather around longer-term
average conditions (i.e., more or fewer extreme weather events).
Climate change is caused by factors such as biotic processes,
variations in solar radiation received by Earth, plate tectonics, and
volcanic eruptions. Certain human activities have also been
identified as significant causes of recent climate change, often
referred to as "global warming"
Coefficient of Usually expressed as R2, this commonly used measure of the
determination goodness of fit is a dimensionless ratio of the explained variation in
the dependent variable over the total variation of the dependent
variable.
Coir fascine A soil bioengineering technique used to stabilize streambanks. A
manufactured product consisting of coconut husk fibers bound
together in a cylindrical bundle held by natural or synthetic netting.
Compaction The process of densifying soil so that air is expelled and the pore
space is reduced.
Compliance Monitoring An activity often required by permits that focuses on and reports on
whether restoration activities are being implemented as designed.
Conditional Letter of Map Provides Federal Emergency Management Agency’s comment on
Amendment (CLOMA) whether a proposed project would be excluded from the Special
Flood Hazard Area.
Conditional Letter of Map Provides for a review of whether a proposed project within a Special
Revision (CLOMR) Flood Hazard Area meets the minimum flood plain management
criteria of the National Flood Insurance Program.
Confidence limits Provides a measure of the uncertainty or spread in an estimate. In
hydrologic gage analysis, confidence limits are a measure of the
uncertainty of the discharge at a selected exceedance probability.
Confluence The point where two streams or rivers merge. If they are of
approximate equal size, this point may be called a fork.
Conservation management An area having similar land use and treatment needs and
unit (CMU) management plan.
Constraints Limitations on the physical or biologic behavior and characteristics of
a stream.

Large Wood National Manual July 2015


xxxviii
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Constructed channel A ditch or reconstructed natural channel.


Construction inspector The person responsible for the day-to-day quality control inspection
required to ensure that prescribed work is installed according to the
design, industry standards, and contract requirements.
Contour fascine See Fascine.
Contract types The many methods used to direct and pay for the installation of
stream restoration or stabilization. The contract types vary primarily
by administrative burden, construction oversight, and incentive for
the contractor to control cost.
Contracting officer (CO) The person responsible for administering the contract, including
ensuring that the proper type of contract is being used and funds are
spent according to regulations.
Contracting officer’s The person responsible to the state engineer and the contracting
representative (COR) officer to see that the work is carried out as designed and in
accordance with the contract requirements.
Contraction scour Erosion of a streambed that occurs when the flow cross section is
reduced by natural features, such as stone outcrops, ice jams, or
debris accumulations, or by constructed features such as bridge
abutments.
Conveyance A measure of the flow-carrying capacity of a cross section.
Cost reimbursement A contract type whereby the contractor is paid for identified costs
contract that are defined as reimbursable. See Contract types.
Crib wall A soil bioengineering technique used to stabilize streambanks. A crib
is a hollow, box-like structure of interlocking logs or timbers. The
structure is filled with rock, soil, and live cuttings or rooted plants.
Crimping and seeding A soil bioengineering surface roughening treatment that secures
straw to the surface. This is a temporary surface treatment that
protects and promotes the establishment of permanent grasses and
vegetation.
Critical shear stress The shear stress at the initiation of particle motion.
Critical uncertainties Key questions that shape how an ecological system is actively
managed (see Adaptive management).
Cross-section area See Flow area.
Cross vane structure A structure that provides grade control and a pool for fish habitat.
Crumb test A common field test for dispersive clays.
Darting speed See Burst swimming speed.
Dead stout stakes Diagonally cut 2- by 4-inch lumber used to secure soil bioengineering
practices.

Large Wood National Manual July 2015


xxxix
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Debris Fragments of solid matter, typically rock or organic material found


moving down a river. Commonly has a negative connotation
synonymous with “refuse” or “garbage” that concerns some people
when using it to describe natural materials found in rivers;
therefore, “debris” is increasingly being replaced by “material” when
referring to wood (see Large woody debris and Large woody
material).
Deflector A structure that forms a physical barrier to protect the bank and
forces the flow to change direction either by direct impact or
deflection.
Deforestation The removal of a forest or stand of trees where the land is thereafter
converted to a non-forest use. Examples of deforestation include
conversion of forestland to farms, ranches, or urban use.
Degradation Long-term sediment removal occurring through increased erosion
from the channel bed.
Deposition The geomorphic process in which sediments, soil, and rocks are
added to a landform or land mass. Wind, ice, and water, as well as
sediment flowing via gravity, transport previously eroded sediment,
which is deposited, building up layers of sediment.
Depth The distance between the channel bottom and the water surface.
Design flow Stream restoration design should consider a variety of flow
conditions. These flows should be considered from both an
ecological and physical perspective.
Design layout The physical location of design elements in a stream restoration
project; the most common methods used to locate features on a
drawing include referencing to a baseline or centerline, creating a
grid, or using a global positioning system (GPS).
Design storm A prescribed precipitation distribution and associated recurrence
interval.
Dimensionless shear stress The ratio of the critical shear stress and the product of the grain
diameter and the submerged specific weight of the particle—also
referred to as the Shields parameter.
Direct use When humans directly use the end product of an ecosystem service,
such as consuming fish and animals, harvesting timber, or using
other forest products.
Discharge The rate of flow, often expressed in cubic feet per second, or ft3/s.
Displacement Submerged volume of an object (large wood).
Disturbances Changes to the physical or ecologic condition that are outside of the
normal range of natural variations. Disturbances can be natural or
anthropogenic.
Ditch A long, relatively narrow, constructed channel.
Dominant channel The forces at work in the watershed that cause and limit channel
processes change.
Dominant discharge See Channel-forming discharge.

Large Wood National Manual July 2015


xl
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Dormant post planting A soil bioengineering technique involving the use of large dormant
stems, branches, or trunks of live woody plant material that are
planted for bank erosion control and creation of riparian vegetation.
Downwelling The process of accumulation and sinking of higher density material
beneath lower density material.
Drag The fluid force component acting on a sediment particle, which is
parallel to the mean flow.
Drainage area The area from which surface rainfall runoff is contributed to a
specific point.
Drained soil conditions This is not a description of the water level in the soils, but rather a
description of the pore pressure condition in the soil when it is
loaded. A drained condition implies that either no significant pore
pressures are generated from the applied load or that the load is
applied so slowly that the pressure dissipates during the slowly
applied loading. See Undrained soil conditions.
Duration The length of time that water flows at a given discharge or a given
depth.
Ecological evaluation Classifying and/or assessing the relative worth, in non-monetary
terms, of different ecological resources.
Ecological stress The physical, chemical, and biological constraints on the productivity
of species as well as alteration of ecosystem function.
Ecoregion An ecologically and geographically defined area that covers a
relatively large area of land or water, and contains characteristic,
geographically distinct assemblages of natural communities and
species. The biodiversity of flora, fauna, and ecosystems that
characterize an ecoregion tends to be distinct from that of other
ecoregions.
Ecosystem A community of living organisms (plants, animals, and microbes) in
conjunction with the nonliving components of their environment
(things like air, water, and mineral soil), interacting as a system.
Effective discharge The mean of the arithmetic discharge increment that transports the
largest fraction of the annual sediment load over a period of years;
often used as a surrogate for channel-forming discharge.
Effectiveness monitoring Activity that assesses ecosystem-, natural community–, and covered
species–scale responses to the implementation of conservation
measures and monitors progress made toward achieving biological
goals and objectives.
Embankment bench A technique used to stabilize steep banks with little or no
disturbance at the top of the slope and minimal disturbance to the
streambed. A gravel bench is constructed along the toe and
protected with riprap.
Endangered Species Act A 1973 act of Congress instructing federal agencies to carry out
(ESA) programs to conserve endangered and threatened species and to
conserve the ecosystems on which these species depend.

Large Wood National Manual July 2015


xli
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Energy A property of a body or physical system that enables it to move


against a force. It is the amount of work required to move a mass
through a distance.
Engineer The person responsible for the technical requirements of project
installation; represents the owner. An engineer is trained in or
follows as a profession a branch of engineering, licensed in the state
of the proposed project.
Engineered logjams (ELJs) General term referring to a human designed and constructed
structure of inter-locking logs intended to emulate natural logjams
using scientific and engineering data to determine appropriate
placement, materials, architecture, and size to ensure the stability
and function that will achieve project goals. There is no single type
of ELJ, and structures can vary significantly in shape, architecture,
size, and function.
Engineered wood Structures ranging from a single log to hundreds of logs, woody
placement (EWP) debris, and other materials that are intended to provide restoration
function such as grade control, flow deflection, and stress-
partitioning; create pools; provide cover and hydraulic refugia for
fish; collect floating debris and sediment; create islands and side
channels; or improve floodplain connectivity by raising water
elevations. EWP includes ELJs.
Entrenchment The extent of vertical containment of a channel relative to its
adjacent floodplain.
Entrenchment ratio The flood-prone width divided by the bankfull width.
Ephemeral stream A stream or reach of a stream that flows only in direct response to
precipitation, and whose channel is above the water table at all
times. The term may be arbitrarily restricted to a stream that does
not flow continuously during periods of as much as a month.
Equilibrium bed slope The slope at which the sediment transport capacity of the reach is in
balance with the sediment transported into it. Also referred to as
equilibrium slope.
Equipment rental contracts A contract type used in instances where a fixed-price construction
contract would be impractical because of the nature of the work and
when it would not be feasible to prepare detailed drawings and
specifications. It requires substantial construction oversight. See
Contract types.
Erosion The wearing away of soil by gravity, running water, wind, or ice.
Erosion control blankets A temporary protective blanket laid on top of bare soil vulnerable to
(ECB) erosion; commonly made of mulch, wood fiber, or synthetics.
Erosion control fabric See Erosion control blankets.
Erosion stop wattle fence See Wattle.
Eutrophication An increase in the rate of supply of organic matter.

Large Wood National Manual July 2015


xlii
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Excavated bench A technique used to stabilize steep banks with little or no


disturbance at the top of the slope and minimal disturbance to the
streambed; involves shaping the upper half or more of the high bank
to allow the formation of a bench to stabilize the toe of the slope.
Extremal hypothesis A hypothesis that assumes a channel will adjust its geometry so that
the time rate of energy expenditure is minimized.
Existence value The value people place on knowing an environmental amenity exists,
even if they have no plans to personally use it.
Facet A distinct morphological segment of a longitudinal profile; riffle,
pool, run, or glide (tail-out).
Fascine A soil bioengineering technique used to provide stabilization to the
toe of streambanks. A long bundle of live cuttings bound together
into a rope or sausage-like bundles.
Federal Acquisition Regulations that govern federal contracts.
Regulations (FAR)
Filter layer A layer that prevents finer grained particles from being lost through
the interstitial spaces of the riprap material, while allowing seepage
from the banks to pass. This layer typically consists of a geosynthetic
layer or sand, gravel, or quarry spalls.
First-order stream An unbranched tributary.
Fish screen Screen that is designed to prevent fish from swimming or being
drawn into an aqueduct, cooling water intake, dam or other
diversion on a river, lake, or waterway where water is taken for
human use. They are intended to supply debris-free water without
harming aquatic life.
Fixed-price contract In most cases, considered to be the preferable type of construction
contract. However, it requires an accurate cost estimate and
construction details. See Contract types.
Flood A general term given to a relatively high flow measured in height or
discharge quantity.
Flood insurance rate map The official map of a community on which the Federal Emergency
(FIRM) Management Agency has delineated both the special hazard areas
and the risk premium zones applicable to the community.
Floodplain An area of land adjacent to a stream or river that stretches from the
banks of its channel to the base of the enclosing valley walls and
experiences flooding during periods of high discharge often absent
in incised channels.
Floodplain maps Maps developed by the National Flood Insurance Program to reduce
damages and loss of life caused by floods. The basis for flood
management, regulation, and insurance requirements by identifying
areas subject to flooding are provided.
Flood-prone width The width of the active floodplain at the floodplain elevation (twice
the maximum bankfull depth); composed of the active channel
(bankfull width) and left and right flood plain (flood-prone) widths.

Large Wood National Manual July 2015


xliii
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Floodway The channel of a river or other watercourse and the adjacent land
areas that must be reserved in order to discharge the base flood
without cumulatively increasing the water surface elevation by more
than a designated height.
Flow area The area of the cross section between ground and water surface.
Flow-changing devices A broad category of structures that can be used to divert flows away
from eroding banks.
Flow depth See Depth.
Flow duration The percentage of time that a flow level is equaled or exceeded in a
stream or river, typically represented with a flow-duration curve.
Flow-frequency analysis A consistent, statistical method for denoting the probability of
occurrence of flow magnitudes at a specific point in a stream
system.
Fluvial Term referring to the channel drainage network of a watershed and
the processes and conditions influencing the flow of water,
sediment; and organic material; includes channels of all sizes from
where they initiate to where they end. The term is used to refer to
all topics related to flowing water and is a major discipline in the
earth sciences, physical geography, and water resource engineering.
Fluvial fish Species that live in the flowing waters of rivers or streams but
migrate between rivers and tributaries for breeding, feeding, or
sheltering.
Fluvial geomorphology The study of the origin and evolution of landforms shaped by river
processes.
Food chain A succession of organisms in an ecological community that
constitutes a continuation of food energy from one organism to
another as each consumes a lower member and in turn is preyed
upon by a higher member.
Food web A combination of food chains that integrate to form a network; the
entirety of interrelated food chains in an ecological community.

Force account agreements Used when the sponsor performs the work using its own equipment
and personnel.
Formal contract Under the Federal Acquisition Regulations as of 2005, formal
contracts must be used for projects with a value greater than
$100,000.
Friction factor (f) The roughness coefficient in the Darcy-Weisbach velocity equation.
Froude number A dimensionless ratio, relating inertial forces to gravitational forces,
and representing the effect of gravity on the state of flow in a
stream.
Future without Action The option that involves allowing the site to progress without a
alternative project. The resources, both physical and ecological, that may be lost
by not implementing the project are assessed as part of this
alternative.

Large Wood National Manual July 2015


xliv
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Gabion A rock-filled wire mesh basket used to stabilize streambanks and


slopes.
Gabion grade control Grade control structures built with rock-filled wire mesh baskets.
Gage analysis The use of statistical techniques to estimate probable frequency of
flow events from recorded stream or river gage records.
General permits Permits that are issued nationwide or regionally for categories of
activities that are either similar in nature or cause only minimal
individual and cumulative adverse impacts.
General scour Streambed erosion affecting the entire channel cross section.
Geocell A product composed of polyethylene strips, connected by a series of
offset, full-depth welds to form a three-dimensional honeycomb
system.
Geogrid A geosynthetic formed by a regular network of integrally connected
elements with apertures greater than a quarter inch to allow
interlocking with surrounding soil, rock, earth, and other
surrounding materials to function primarily as reinforcement.
Geologic assessment The review of both the surface and subsurface features of geology
and their possible impacts on a stream or river.
Geomorphic analog The use of a stable stream reach as a template for restoration
design.
Geomorphic goals Goals or objectives based on concepts of landscape position,
landforms, and ongoing processes that change them.
Geomorphology The study of the origin and evolution of landforms, focusing on
linking unique landscape attributes to the physics, chemistry, and
biology of the formative processes (e.g., tectonics, earth materials,
volcanism, surface and subsurface hydrology, slope stability,
vegetation, human alterations, waves, currents, etc.). Sub-disciplines
include hillslope, fluvial, coastal, soil, submarine, and even planetary
geomorphology.
Geonet A geosynthetic consisting of integrally connected parallel sets of ribs
overlying similar sets at various angles for planar drainage of liquids
and gases.
Geosynthetic A planar product manufactured from polymeric material used with
soil, rock, earth, or other geotechnical engineering related material
as part of a human-made project structure or system.
Geotechnical analysis The evaluation of the forces involved in bank instability problems
including cohesion, friction, gravity acting on the soils in the slope,
the internal resistance of soils in the slope, and the seepage forces in
the soils in the bank.
Geotextile A permeable geosynthetic comprised solely of textiles.
Glide The downstream end of pools, just upstream of the next riffle, where
the channel slope becomes adverse as the deeper section is
intercepted by the tailing off point bar.

Large Wood National Manual July 2015


xlv
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Goals The overall desired outcome, such as restoring a channel to preflood


conditions.
Grade control See Grade stabilization techniques.
Grade stabilization Techniques used to stop channel degradation, typically
techniques accomplished by the construction of in-channel structures.
Grain Reynolds number The ratio of the product of shear velocity and grain diameter to
kinematic viscosity.
Grass-lined channel design A threshold channel design technique used where climate and soils
method can support permanent vegetation and baseflow does not exist. The
approach is similar to the allowable velocity channel design method.
Gravelometer Device used to assist with the measurement of particles sampled as
part of a pebble count.
Groundwater Water in a saturated zone or stratum beneath the land surface.
Grout See Grouted riprap.
Grouted riprap A riprap bed where the voids have been filled with concrete; often
used where the required stone size cannot be obtained or at sites
where a significant and damaging debris load is expected.
Gully/gullies Entrenched channels extending into areas with previously undefined
or weakly defined channel conditions.
Gully plug A small earthen dam constructed at one or more locations along the
gully.
Habitat A specific environment in which a particular plant or animal lives.
Habitat Unit The area of habitat types (e.g., pools or riffles) adjusted for habitat
preference (e.g., pools have high preference for coho fry in summer
but low preference for coho spawning) and by the suitability of that
habitat indexed by the habitat suitability index (HIS).
Hybrid design methods The use of a combination of analytical, as well as analogy and
hydraulic geometry, design methods to calculate design variables.
Hydraulic control structure A type of grade control structure designed to reduce the energy
slope along the degradational zone to the degree that the stream
can no longer scour the bed.
Hydraulic depth The ratio of the cross-section area of flow to the free surface or top
width.
Hydraulic geometry design Design approach based on the concept that a river system tends to
method develop in a predictable way, producing an approximate equilibrium
between the channel and the inflowing water and sediment.
Hydraulic radius The ratio of the cross-sectional area of flow to the wetted perimeter
or flow boundary.
Hydro-physiographic area A drainage basin where the combination of the mean annual
precipitation, lithology, and land use produces similar discharge for a
given drainage basin.

Large Wood National Manual July 2015


xlvi
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Hyporheic zone A region beneath and alongside a stream bed, where there is mixing
of shallow groundwater and surface water. The flow dynamics and
behavior in this zone (termed hyporheic flow or underflow) is
recognized to be important for surface water/groundwater
interactions, as well as fish spawning, among other processes. As an
innovative urban water management practice, the hyporheic zone
can be designed by engineers and actively managed for
improvements in both water quality and riparian habitat,
Incentive contracts A contract type that links the contractor’s profit to performance by
establishing reasonable and attainable targets that are clearly
communicated to the contractor. See Contract types.
Incipient motion design See Threshold channel design.
Index of Biotic Integrity (IBI) A biological assessment technique that uses fish surveys to assess
human effects on a stream and its watershed.
Individual permit A type of permit that involves the evaluation of a specific project.
Infiltration The downward movement of water into the surface of soil.
Informal contract Under the Federal Acquisition Regulations as of 2005, informal
contracts and contracting procedures can be used for projects with a
value of $100,000 or less. Informal contracts are those put in place
using simplified acquisition procedures.
Intermittent stream A stream that flows only at certain times of the year when it receives
water from springs or from some surface source such as melting
snow in mountainous areas. The term may be arbitrarily restricted to
a stream that flows continuously during periods of at least 1 month;
also may be a stream that does not flow continuously, as when
water losses from evaporation or seepage exceed the available
streamflow.
Irrigation ditch A long, narrow, constructed channel used to convey irrigation water
from its source to place of use.
Jetties A flow-changing technique used to stabilize and protect stream and
river banks; fence-like structures extending from the bank and into
the stream.
J-hook A rock structure used to provide bank stabilization.
Joint planting A streambank soil bioengineering technique that includes cuttings of
live woody plant material inserted in the voids of riprap and into the
ground below the rock.
Jumping height The maximum height obtained by a specific species and age of fish.
Older and larger fish have greater maximum jumping heights,
although some species have no jumping abilities at any age.
Key member or log A critical structural element within an engineered logjam or wood
structures.

Large Wood National Manual July 2015


xlvii
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Key piece A functional piece of natural wood—one that is large enough to and
with a shape that contributes to formation of a stable snag that
alters flows and channel form. Diameters tend to be equal or greater
than half the bankfull or effective discharge depth and have a
rootwad or multiple stems.
Labor-hour contracts A variation of the time-and-materials contract, differing only in that
materials are not supplied by the contractor. See Contract types.
Lane’s relationship A qualitative conceptual model, also known as a stream balance,
used as an aid to visually assess stream responses to changes in flow,
slope, and sediment load.
Lane’s tractive force design See Allowable shear stress design method.
method
Large wood Term most commonly used in the literature describing pieces of
wood such as branches and tree trunks, as opposed to particulates
or small fragments of wood. Some publications define “large” as any
piece of wood more than 10 centimeters (4 inches) in diameter and
1 meter (3 feet) in length. Because the word “large” is subjective
without an explicit definition, some authors have simply used “wood
debris” to describe the same thing. Some authors have thought the
word “debris” has negative connotations and prefer using words
such as “material.” This manual simply uses “large wood.”
Letter contracts Written preliminary contractual instruments that authorize the
contractor to begin work immediately.
Letter of map amendment An amendment to the currently effective Federal Emergency
(LOMA) Management Agency map establishing that a property is not located
in a Special Flood Hazard Area.
Letter of map revision An official amendment to the currently effective Federal Emergency
(LOMR) Management Agency map.
Letter of permission (LOP) A type of permit issued through an abbreviated processing
procedure.
Lift The fluid force component on sediment particles perpendicular to
the mean flow direction.
Little Underwater A technique providing both streambank stability and edge cover
Neighborhood Keepers aquatic habitat.
Encompassing Rheotactic
Salmonids (LUNKERS)
Live bed conditions Conditions that may be assumed to exist at a site if the mean
velocity upstream exceeds the critical velocity for the beginning of
motion for the median size of bed material available for transport.
Live brush sills A soil bioengineering technique that involves rows of live cuttings
inserted into an excavated trench. This treatment is intended to
promote sediment deposition and can function as erosion stops.

Large Wood National Manual July 2015


xlviii
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Live pole cuttings A soil bioengineering technique that involves the use of dormant
stems, branches, or trunks of live woody plant material inserted into
the ground that are planted for bank erosion control and creation of
riparian vegetation.
Live post planting See Dormant post planting.
Live siltation See Live brush sills.
Live stakes See Live pole cuttings.
Local scour Erosion of the streambed immediately adjacent to some obstruction
to flow.
Log crib See Crib wall.
Log-Pearson type III The most commonly used frequency distribution for peak flows in
distribution the United States; applies to nearly all series of natural floods;
commonly used for stream gage analysis.
Log vanes/step jams Single logs or small bundles of logs secured to bed. Also called log
bendway weirs (if partially spanning channel and angling upstream)
or log steps (if fully spanning channel, and usually placed
perpendicular to channel).
Log weirs/valley jams Weir-like accumulations built around one or more large logs (key
members).
Longitudinal peak stone toe A type of bank protection involving the placement of a windrow of
(LPST) stone in a peak ridge along the toe of an eroding bank.
Loose rock grade control A simple type of a grade control structure consisting of placing
structure natural stone or other nonerodible elements across the channel to
form a hard point.
Low flow A general term that refers to the average low flows in a stream. It is
typically due to soil moisture and ground water. Critical habitat
conditions often occur during low flows.
Low-flow channel A portion of a channel that conveys low or baseflows.
Maintenance Actions taken to ensure that the stream restoration project performs
as designed and is attaining project objectives.
Manning’s n An empirical factor in Manning’s equation which accounts for
frictional resistance of the flow boundary.
Meander Deviation of the stream direction from the shortest possible path
down a stream valley.
Meander geometry The five parameters commonly used in the description of meander
patterns: wavelength, radius of curvature, arc length, amplitude, and
beltwidth.
Meander length The product of the meander wavelength and the valley slope divided
by the channel slope.
Meander ratio The length of the stream divided by the length of the valley.
Mobile boundary stability The rate at which sediment enters the channel reach from upstream
equal to the capacity of the reach to transport sediment of the same
composition on downstream.

Large Wood National Manual July 2015


xlix
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Model (1D) One-dimensional models only consider forces that occur in one
direction (usually the streamwise). Velocity and other stream
properties may vary upstream and downstream, but not from bank
to bank and not from the bed to the water surface.
Model (2D) Models are usually depth-averaged. They simulate variation in the
horizontal plane, but assume no variation in the vertical.
Model – conceptual Describes the objects and relationships either with words or
diagrams.
Model – empirical Contains any empirical relationship, one based on data. An empirical
model is based, at least in part, on observed data, rather than a
thorough understanding of the underlying physical principles.
Model – lumped Describes processes on a scale larger than a point, while a
distributed model describes all processes at a point, and then
integrates processes over space and time to produce a total system
response.
Model – mathematical Formal mathematical models representing objects and interactions
quantitatively with equations.
Model – parametric Has parameters that must be estimated in some fashion.
Model – physical Three-dimensional representations, usually at some relevant scale.
Model – steady Predicts conditions that occur for a given set of boundary conditions.
For example, a flow model might predict the water surface
elevation, given a fixed channel geometry and a constant flow.
Model – stochastic Outputs are predictable only in a statistical sense. Repeated use of a
given set of model inputs produces outputs that are not the same
but follow certain statistical patterns.
Model – unsteady Predicted variations that occur with time, such as during the passage
of a storm hydrograph, by dividing such an event into a series of
steady-state time steps. Complex, unsteady models have feedback
loops that allow channel boundaries or other key variables to
respond to inputs and change between time steps.
Momentum The mass of a body times its velocity.
Monitoring The process of measuring or assessing specific physical, chemical,
and/or biological parameters of a project.
Montgomery and A classification system based on defining channel processes. It is a
Buffington classification geomorphic process-based system.
Muddying-in The practice of pouring a slurry mix of water and soil into the hole
around the cutting stem of a plant to achieve good soil-to-stem
contact.

Large Wood National Manual July 2015


l
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Multi-Criteria Decision A sub-discipline of operations research that explicitly considers


Analysis multiple criteria in decision-making environments. Whether in our
daily lives or in professional settings, there are typically multiple
conflicting criteria that need to be evaluated in making decisions.
Cost or price is usually one of the main criteria. Some measure of
quality is typically another criterion that is in conflict with the cost.
For example, in purchasing a car, cost, comfort, safety, and fuel
economy may be some of the main criteria we consider. It is unusual
to have the cheapest car to be the most comfortable and the safest.
National Environmental The federal law establishing a national policy for the environment;
Policy Act (NEPA) requires specific actions by federal agencies.
National Flood Insurance A program administered by the Federal Emergency Management
Program (NFIP) Agency providing for flood insurance, flood plain hazard mapping,
and flood plain management.
Nationwide General Permit A type of general permit issued nationally by the U.S. Army Corps of
(NWP) Engineers for specific dredge or fill activities.
National Pollutant A provision of the Clean Water Act regulating point discharges into
Discharge Elimination waters of the United States.
System (NPDES)
Natural channel A river, stream, creek, or swale that has existed long enough and
without significant alteration to establish a dynamically stable route.
Navigable waters Defined for U.S. Army Corps of Engineers regulatory purposes as
those waters that are subject to the ebb and flow of the tide and/or
are presently used, or have been used in the past, or may be
susceptible for use to transport interstate or foreign commerce.
Newbury riffle A type of constructed loose rock grade control structure.
Natural Resources Guidance provided for applying conservation technology and setting
Conservation Service(NRCS) the minimum criteria for acceptable application of the technology.
Conservation Practice State variations on these standards may be more restrictive.
Standards
NRCS Planning Process Steps used to develop an appropriate plan for natural resource
protection or improvement.
NRCS State Conservation Each state determines which NRCS National Conservation Practice
Practice Standards Standards are applicable in their state. States add the technical
detail needed to effectively use the standards at the field office level
and issue them as state conservation practice standards. Minimum
criteria may be more restrictive than the national standards.
Objectives The detailed, focused outputs or outcomes that achieve project
goals.
Open channel flow Flow where one surface is open to the atmosphere.
Ordinary high water The limit of U.S. Army Corps of Engineers jurisdiction in nontidal
waters of the United States, in the absence of adjacent wetlands;
defined as that line on the shore established by the fluctuations of
water and indicated by physical characteristics.

Large Wood National Manual July 2015


li
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Outliers Data points that depart significantly from the trend of the remaining
data.
Owner The person responsible for contracting for construction. For NRCS
Federal contracts, NRCS is considered the owner during
construction.
Partial duration gage The analysis of the recorded peak flow values above a preselected
analysis base value that have occurred for each year in the duration of
interest; typically used for the estimate of flows with return intervals
less than 2 years.
Pattern Plan view of a stream reach.
Pebble count Technique used to sample the surface layer of sediments in
gravelbed streams.
Perennial stream A stream that flows continuously; streams flowing continuously
throughout the year and that are generally lower than the water
table in the region adjoining the stream.
Performance monitoring Activity that identifies whether conservation measures are achieving
the expected outcomes or targets.
Performance of work An agreement that requires that the value of work to be performed
agreement by the sponsoring local organization be determined by negotiation
between the sponsoring local organization and NRCS and be
included in the project agreement. NRCS must estimate the cost of
the work to establish the maximum value of work before signing the
agreement.
Periphyton Algae, fungi, bacteria, protozoa, and organic matter associated with
channel substrates.
Pile A vertical element made of wood, steel, or other material that is
embedded deeply into a streambed, either by driving with a
hydraulic, diesel or vibratory hammer.
Pile foundations Used to transfer foundation forces through relatively weak soil to
stronger strata to minimize settlement. The most likely applications
for pile foundations in stream restoration and stabilization projects
are as support for bank stabilization structures (retaining wall) and
as anchors for large woody material.
Pin deflectors Variations of the permeable jetty, generally used in streams where
only a small reduction in velocity is needed. Generally wood pilings
are used for their construction.
Piston aerial sampler Device used to facilitate underwater aerial sediment sampling of fine
material.
Plan A sequence of logical steps followed to reach a goal or objective.
Planform Horizontal alignment of a channel; view is perpendicular to the
Earth’s surface.
Point bar A depositional area formed on the inside bank of a meander that
sometimes remains bare of vegetation due to the frequent
recurrence of the bankfull discharge.

Large Wood National Manual July 2015


lii
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Pool The area in a natural channel deeper and somewhat narrower than
the average channel section. Pools are stream features that have
residual depth and therefore will not drain free of water if flows are
curtailed.
Practice standards See NRCS Conservation Practice Standards.
Pressure head The potential energy of water, usually the result of its mass and the
Earth’s gravitational pull.
Productivity The density-independent survival, which, along with density-
dependent factors of the environment, determines abundance
limited by the total capacity of the environment.
Programmatic General A type of general permit issued to avoid unnecessary duplication of
Permit (PGP) regulatory control exercised by another federal, state, or local
agency.
Post Similar to a pile but placed by excavating a hole, placing the post,
and backfilling. Excavated holes are necessary to bury a tree with
attached rootwad—an element that has significantly more
resistance to pulling out or overturning than a pile driven to the
same depth.
Project agreements Any agreement(s) entered into by NRCS and sponsors, in which
detailed working arrangements are established for the installation of
cost-shared measures.
Pump intake fish screens See Fish screens.
Quality assurance (QA) Tasks or procedures undertaken to ensure that procedures are
adhered to that will assure that work will meet minimum
requirements. Quality assurance activities vary in accordance with
the complexity and hazard class of the stream restoration project.
Quality assurance plan Identifies the individuals with the expertise to perform various QA
(QAP) tasks, outline the frequency and timing of testing, estimate the
contract completion date, and be co-approved by all responsible
supervisors.
Quality control (QC) Tasks or procedures undertaken to ensure that work installed meets
the minimum requirements of the contract.
R2 The coefficient of determination in a regression analysis. This
commonly used measure of the goodness of fit is a dimensionless
ratio of the explained variation in the dependent variable over the
total variation of the independent variable.
Racking debris Wood debris in a wide range of sizes that would be mobile within
the stream or river is retained by a natural or engineered logjam
(racked members or racking logs). This is very important material
because it tends to decrease permeability of the structure and
reduce drag coefficients. It also provides almost all of the aquatic
cover and interstitial space for fish.

Large Wood National Manual July 2015


liii
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Reach A subjective term describing a segment of stream or river. Typically


defined as a length of stream or river having some defined uniform
characteristics, typically about 20 channel widths, or a segment with
uniform planform, gradient, and width measures.
Reclamation A series of activities intended to change the biophysical capacity of
an ecosystem. The resulting ecosystem is different from the
ecosystem existing prior to recovery. The term has implied the
process of adapting wild or natural resources to serve a utilitarian
human purpose, such as the conversion of riparian or wetland
ecosystems to agricultural, industrial, or urban uses.
Reconnaissance A preliminary investigation not involving detailed investigation and
relying heavily on existing data and observations.
Recurrence interval The anticipated period in years before a given flood will reoccur.
Redirective structure A flow-changing bank stabilization technique; designed to be placed
in the stream, minimize direct impact, and rely more on the
characteristics of fluid mechanics to modify the streamflow
direction.
Reference reach design An alluvial channel design approach whereby channel dimensions
method are selected from a similar stable channel.
Regime design method An alluvial channel design approach whereby channel dimensions
are selected with the aid of empirically derived equations.
Regional curves A tool frequently associated with the Rosgen geomorphic channel
design approach, but also applicable to other design methods. It
involves bankfull dimensions correlated to a drainage area. See
Hydraulic geometry design method.
Regional general permit A type of general permit issued regionally.
(RGP)
Regression equations (gage Used to transfer flood characteristics from gaged to ungaged sites
analysis) through use of watershed and climatic characteristics as predictor
variables.
Regulated stream systems Streams or rivers that are cleared of wood, dammed, channelized,
leveed, or constrained by other types of hard structures.
Rehabilitation Making the land useful again after a disturbance; it involves the
recovery of ecosystem functions and processes in a degraded
habitat.
Replacement The minimum number of spawners required to maintain a given
abundance.
Resistance Capacity of species and ecosystems to tolerate some changes in the
intensity of ecological stressors.
Resource management Sets of approved conservation practices.
systems (RMS)
Restoration The reestablishment of the structure and function of ecosystems.
Ecological restoration is the process of returning an ecosystem as
closely as possible to predisturbance conditions and functions.

Large Wood National Manual July 2015


liv
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Retard A flow-changing bank stabilization technique. A retard structure


increases flow resistance by increasing drag, thereby slowing the
velocity in the vicinity of the structure. These structures are more
porous with a high percentage of open area.
Reynolds number A dimensionless ratio, relating the effect of viscosity to inertia, used
to determine (index) whether fluid flow is laminar or turbulent.
Riffle The area in a natural channel that is wider and shallower than the
average channel section.
Riffle pool spacing The distance between the riffles and the pools in a channel.
Rigid boundary stability Attained when the interaction between flow and the material
forming the channel boundary is such that the soil boundary
effectively resists the erosive efforts of the flow.
Rigid drop grade control A complex type of grade control structure that is used for large
structure drops. These structures are frequently constructed of concrete or a
combination of sheet pile and concrete.
Riparian forest Forested or wooded area of land adjacent to a body of water such as
a river, stream, pond, lake, marshland, estuary, canal, sink, or
reservoir.
Riparian zones The areas between aquatic and upland habitats. Often defined as
the "zone of influence" between aquatic and terrestrial
environments.
Riprap Large stone used to provide immediate and permanent stream and
river bank protection.
Riprap sizing See Stone sizing.
Risk The exposure of life, property, and/or the environment to loss or
harm. The product of a likely occurrence times its consequence.
Risk analysis The assessment of the consequences of specific action or inaction to
life, property, and/or the environment.
Risk tolerance The level of risk a decision maker is willing to accept, or the risk
response determined by law or policy.
River A large natural waterway confined within a bed and banks. In the
context of this handbook, the term stream is used and encompasses
river.
River classification See Classification.
River Continuum Concept Reaches form a continuous ecological system that processes organic
material and produces a distinct pattern of biological communities.
Rolled erosion control Consist of both erosion control blankets used for temporary erosion
products protection and turf reinforcement mats for more permanent erosion
protection.
Rootwad The root systems of upended trees.
Rootwad revetments Use of locally available logs and root fans to add physical habitat to
streams in the form of coarse woody debris and deep scour pockets.

Large Wood National Manual July 2015


lv
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Rosgen classification A stream classification system based on measurements of existing


morphology.
Rosgen geomorphic channel A hybrid channel design approach that incorporates geomorphic
design method measurements, hydraulic geometry and some analytical calculations.
Rosgen stream type See Rosgen classification.
Rotary drum fish screens See Fish screens.
Run The steepest section and shortest longitudinally, starting at the
downstream end of a riffle as the channel enters the next pool.
Salmonid Family of fish that includes salmon, trout, and char. All of the species
breed in freshwater, are migratory, and spend part of their life cycle
in the ocean.
Scour Downward vertical erosion in a channel bed.
Se or SY,X The standard error of estimate, typically expressed as Se or SY,X.
This is a measure of the quality of a regression equation and is the
root mean square of the estimates. It is a measure of the scatter
about the regression line of the independent variable.
Seasonal stream An intermittent stream that flows only during a certain climatic
season, such as a winterbourne. A stream (or segments of a stream)
that normally goes dry during a year of normal rainfall. Seasonal
streams often receive water from springs and/or long-continued
water supply from melting snow or other sources.
Sediment budget analysis A quantitative sediment impact assessment of channel stability using
the magnitude and frequency of all sediment-transporting flows
done by comparing the mean annual sediment load for the project
channel to that of the supply reach.
Sediment competence The ability to move the largest particle made available to the
channel.
Sediment continuity The volume of sediment deposited in or eroded from a reach during
analysis a given period of time is computed as the difference between the
volumes of sediment entering and leaving the reach.
Sediment impact An evaluation of a designed channel’s ability to transport the
assessment inflowing water and sediment load, without excessive sediment
deposition or scouring on the channel bed.
Sediment rating curve Correlates sediment flow to discharge for a stream reach or section.
Sediment rating curve Sediment impact assessment technique used to assess the sediment
analysis transport characteristics of an existing or proposed stream project.
This approach uses sediment rating curves to compare the sediment
transport capacity of the supply reach to the existing and proposed
project reach conditions.
Sediment sampling Technique used to quantify sediment in streams and rivers.
Shear The pull of water on the wetted area in the direction of flow;
measured in units of force/area.

Large Wood National Manual July 2015


lvi
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Shear stress (average) The product of the energy slope, hydraulic radius, and unit weight of
water. Spatial and temporal variation may result in a higher or lower
point value for shear stress.
Sheet pile Flat panels of steel, concrete, vinyl, synthetic fiber, reinforced
polymer, or wood. Typical applications include toe walls, flanking
and undermining protection, grade stabilization structures, slope
stabilization, and earth retaining walls.
Shields diagram Classic method for determining critical shear stress.
Shields parameter See Dimensionless shear stress.
Sinuosity The channel centerline length divided by the length of the valley
centerline.
Skin friction The friction acting on a solid body when it is moving through a fluid.
Slash Wood debris that often is considered waste in logging or site
clearing operations that consists of a wide range of diameters and
lengths (generally small diameter). May also include dirt and rocks.
This is excellent material to supplement racking debris or for soil
erosion protection.
Slope stability See Geotechnical analysis.
Soil anchor Technique used to anchor woody material to the streambed or bank
to resist fluvial forces.
Soil bioengineering The use of live and dead plant materials in combination with natural
and synthetic support materials for slope stabilization, erosion
reduction, and vegetative establishment.
Soil cement grade control Structures constructed with a mix of Portland Cement and onsite
soils.
Specific energy The energy per unit weight of water at a given cross section with
respect to the channel bottom.
Specific force The horizontal force of flowing water per unit weight of water.
Spur dikes Short dikes that extend out perpendicular from the bank into the
channel along a reach of eroded bank.
Stability A channel is considered stable (or in dynamic equilibrium) when the
prevailing flow and sediment regimes do not lead to long-term
aggradation or degradation.
Stability – Wood Large wood placements are stable when the forces resisting motion
are greater than those acting to move the wood.
Stakeholders Individuals or groups who fund a project or are affected by the
project.
Standard individual permit A type of permit issued for activities that have more than minimal
(SP) adverse impacts on waters of the United States. The evaluation of
each permit application involves more thorough review of the
potential effects of the proposed activity.
State administrative officer The person responsible for all administrative matters for contracts
(SAO) and most agreements.

Large Wood National Manual July 2015


lvii
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

State conservation engineer The person responsible for the design and ultimately for ensuring
(SCE) proper construction of projects in a given state.
Steady state models Models that predict conditions that occur for a given set of boundary
conditions.
Stinger Metal rod used to facilitate planting live cuttings into rock riprap.
Stone sizing Technique used to determine the minimum size stone to resist
stream velocity.
Stream A small natural waterway or channel that conveys overland flow
continuously (perennial) or seasonally (ephemeral). Defined within a
bed or banks. In the context of this handbook, stream encompasses
river.
Streambank The embankments on either side of a stream or river channel.
Stream barbs A flow-changing bank stabilization technique that uses low dikes or
sill-like structures that extend from the bank towards the stream in
an upstream direction. As flow passes over the sill of the stream
barb, it discharges normal to the face of the weir.
Streambed The bottom of a stream or river.
Stream classification See Classification.
Stream corridor Includes the stream and extends in cross section from the channel’s
bankfull level towards the upland (perpendicular to the direction of
streamflow) to a point on the landscape where channel-related
surface and/or soil moisture no longer influence the plant
community.
Stream corridor restoration One or more conservation practices used to overcome resource
impairments and reach-identified purposes.
Stream order classification A stream classification system based on the degree of channel
branching. An nth order stream is formed by the intersection of two
or more (n-1) order streams.
Stream power The product of shear stress and mean velocity. A measure of the
available energy a stream has for moving sediment, rock, woody, or
other debris.
Stream setbacks A width required to allow a stream to self-adjust its meander
pattern.
Structured Decision Making An organized approach to identifying and evaluating creative options
and making choices in complex decision situations.
Substrate The base on which an organism lives; for example, the soil is the
substrate of most seed plants.
Surcharge The gravitational load acting on wood.
Surface sediment sampling Techniques used to characterize the surface of a gravel bed.
Sustained swimming speed Refers to the low swimming speeds of a fish species. In general, such
speeds can be maintained for extended time periods with little to no
fatigue.

Large Wood National Manual July 2015


lviii
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Thalweg The deepest portion of the channel; sometimes referred to as the


low-flow channel.
Threshold channel A channel in which channel boundary material has no significant
movement during the design flow. The term “threshold” is used
because the channel geometry is designed so that applied forces
from the flow are below the threshold for movement of the
boundary material.
Threshold channel design A design approach whereby a channel configuration is selected so
that the stress applied during design conditions is below the
allowable stress for the channel boundary.
Timber crib See Crib wall.
Time-and-materials Contract used to procure supplies or services on the basis of direct
contract labor and materials costs. See Contract types.
Toe zone The portion of the bank between the average water level and the
upper edge of the bottom of the channel.
Top width The width of a channel cross section at the water surface.
Tractive power design A threshold channel design technique used in the assessment of
method channels in cemented and partially lithified (hardened) soils.
Transfer methods (gage Technique used to extrapolate peak discharges upstream or
analysis) downstream from a stream gage or from gage data from a nearby
stream with similar basin characteristics.
Transition channel A stream or river that may behave as an alluvial channel in one flow
condition and as a threshold channel in another flow condition.
Tree revetments See Brush revetments.
Tributary A continuous perennial stream.
Trophic level The position an organism occupies in a food chain.
Turbidity The cloudiness or haziness of a fluid caused by large numbers of
individual particles that are generally invisible to the naked eye,
similar to smoke in air. The measurement of turbidity is a key test of
water quality.
Turf reinforcement mats Used to provide permanent erosion protection.
(TRM)
Two-stage channel design A hybrid channel design approach that incorporates a natural alluvial
method channel nested with a constructed flood plain bench.
U.S. Army Corps of Program that evaluates permit applications for most construction
Engineers Regulatory activities that occur in the nation’s waters, including wetlands.
Program
U.S. Forest Service: An aquatic framework containing standard terms and classification
Framework of Aquatic criteria for aquatic systems and their linkages to terrestrial systems
Ecological Units at all spatial scales.
Uncertainty The likelihood of a consequence occurring.

Large Wood National Manual July 2015


lix
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Undrained soil conditions This is not a description of the water level in the soils, but rather a
description of the pore pressure condition in the soil when loaded.
An undrained condition assumes pore pressures will develop due to
a change in load. The assumption is that the pore pressures that
develop are not known and thus must be implicitly considered in the
methods used to test samples for this condition. See Drained soil
conditions.
Uniform flow Occurs when the gravitational forces that are pushing the flow along
the channel are in balance with the frictional forces exerted by the
wetted perimeter that are retarding the flow.
Unsteady models Predict variations that occur with time, such as during the passage of
a storm hydrograph, by dividing such an event into a series of
steady-state time steps.
Valley slope The maximum possible slope for the channel invert; determined by
the local topography; a channel with a slope equal to the valley
slope would be straight.
Vanes Flow-changing structures constructed in the stream designed to
redirect flow by changing the rotational eddies normally associated
with streamflow. They are used extensively as part of natural stream
restoration efforts to improve instream habitat.
Vegetated gabion Incorporates topsoil into the void spaces of the gabion. Woody
plantings and/or grass are planted into or through the structure.
Vegetated geogrid See Vegetated reinforced soil slope.
Vegetated reinforced soil A soil bioengineering technique that is made up of layers of soil
slope (VRSS) wrapped in synthetic geogrid or geotextile, with live cuttings or
rooted plants installed between the wrapped soil layers.
Vegetated riprap See Joint planting.
Vegetated rock wall A mixed-construction soil bioengineering streambank stabilization
technique. The structural-mechanical and the vegetative elements
work together to prevent surface erosion and shallow mass
movement by stabilizing and protecting the toe of steep slopes.
Vegetated soil lifts See Vegetated reinforced soil slope.
Vegetated stone Combining rock with soil bioengineering treatments can achieve
benefits from both techniques.
Velocity head The kinetic energy of water.
Vertical fixed plate fish See Fish screens.
screen
Vertical traveling fish See Fish screens.
screen
Visual geomorphic A qualitative assessment that includes judgment of current
assessment conditions, expected future conditions, and the river’s anticipated
response to the designed project.

Large Wood National Manual July 2015


lx
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

Volumetric sediment The techniques generally considered to be the standard sediment


sampling sampling procedure; involves the removal of a predetermined
volume of material that is large enough to be independent of the
maximum particle size.
W-weir Technique used to provide grade control.
Waterjet See Waterjet stinger.
Waterjet stinger A device that uses high-pressure water to hydrodrill a hole in the
ground to plant unrooted cuttings.
Watershed A topographically bounded area of land that captures precipitation,
filters and stores water, and regulates its release through a channel
network into a lake, another watershed, or an estuary and the
ocean.
Wattle A soil bioengineering technique made up of rows of live stakes or
poles with live plant materials woven in a basket-like fashion. A
wattle fence can be used to deter erosion in ditches or in small dry
channel beds to resist the formation of rills and gullies.
Weir A barrier across a river designed to alter its flow characteristics.
Wetlands Defined for U.S. Army Corps of Engineers regulatory purposes as
those areas that are inundated or saturated by surface or ground
water at a frequency and duration sufficient to support a prevalence
of vegetation typically adapted for life in saturated soil conditions.
Wetted perimeter The length of cross-section boundary between water and ground.
Width-to-depth ratio The bankfull width divided by the mean bankfull depth
(dimensionless).
Wolman pebble count See Pebble count.
Wolman walk See Pebble count.
Work Force applied over a distance.

Large Wood National Manual July 2015


lxi
Bureau of Reclamation and
U.S. Army Corps of Engineers Glossary

This page intentionally left blank.

Large Wood National Manual July 2015


lxii
LIST OF ACRONYMS
°C degrees Centigrade
°F degrees Fahrenheit
1D one-dimensional
2D two-dimensional
ACHP Advisory Council on Historic Preservation
AFDM ash free dry mass
amsl above mean sea level
BFW bankfull flow width
BMPs best management practices
BSTEM Bank Stability and Toe Erosion Model
CEQA California Environmental Quality Act
CLOMR Conditional Letter of Map Revision
CPOM coarse particulate organic matter
dB decibel
DBH diameter at breast height
DF Douglas fir
DOC dissolved organic carbon
EDT Ecosystem Diagnosis and Treatment
EFC Evergreen Funding Consultants
ELJs engineered logjams
EMA Expected Moments Algorithm
ERs Engineer Regulations
ESA Endangered Species Act
FEMA Federal Emergency Management Agency
FEMAT Forest Ecosystem Management Assessment Team
FHWA Federal Highway Administration
FIRMs flood insurance rate maps
FISRWG Federal Interagency Stream Restoration Working Group
GF grand fir
GIS geographic information system
GPS global positioning system
HEC-RAS Hydrologic Engineering Centers River Analysis System
HEP Habitat Evaluation Procedure
HIS Habitat suitability index
IFB Invitation for Bids
IFIM Instream Flow Incremental Methodolog
IUCN International Union for Conservation of Nature
JHA Job Hazard Analyses
LID Low Impact Development
LiDAR Light Detection and Ranging
LOMA letter of map amendment

Large Wood National Manual July 2015


lxiii
Bureau of Reclamation and
U.S. Army Corps of Engineers List of Acronyms

LOMC Letter Of Map Change


LOMR Letter Of Map Revision
MCDA Multi-Criteria Decision Analysis
MCDM Multiple Criteria Decision Making
MH mountain hemlock
NEPA National Environmental Protection Act
NHPA National Historic Preservation Act
NIOSH National Institute of Occupational Safety and Health
NMFS National Marine Fisheries Service
NRCS Natural Resources Conservation Service
NSD Natural Systems Design
NWP Nationwide Permit
OSHA Occupational Safety and Health Administration
PHABSIM Physical Habitat Simulation Model
PP ponderosa pine
PPE personal protective equipment
PS&E plans, specifications, and estimates
QA/QC Quality control and quality assurance
RCC River Continuum Concept
RiverRAT River Restoration Analysis Tool
SAF Subalpine fir
SDM Structured Decision Making
SEPA Washington’s State Environmental Policy Act
SF silver fir
SS Sitka spruce
SWPPP stormwater pollution prevention plan
T&M Time and Material
USACE U.S. Army Corps of Engineers
USC United States Code
USDA U.S. Department of Agriculture
USFS U.S. Forest Service
USFWS U.S. Fish and Wildlife Service
USGS U.S. Geological Survey
VSP Viable Salmonid Population
WH western hemlock
WUA Weighted Useable Are

Large Wood National Manual July 2015


lxiv
LIST OF SYMBOLS

A Area of structure projected in the plane perpendicular to flow


d Width of ditch
c Cohesion of soil
CD Drag coefficient
CL Lift coefficient
Cw Coefficient that captures interaction between ditch walls and fill
d Diameter of log
D Distance from top bank to top of log buried in bank
D50 The grain diameter at which 50% of the sediment sample is finer than the rest.
D90 The grain diameter at which 90% of the sediment sample is finer than the rest.
db Representative boulder diameter
Dp Mean number of days in the month with 0.01 inch or more of precipitation
Dw Distance from top bank elevation to water table elevation
dwn Water depth at which the structure becomes neutrally buoyant
Fn Force normal (perpendicular) to bed
Fsh Safety factor with respect to horizontal movement
Fsv Safety factor with respect to vertical movement

Fb Buoyant force

FL Lift force

Ff Force of friction

Fd Drag force

Fc  Restraining force due to anchors or ballast

𝐹⃗𝑎𝑣 Vertical restraint force provided by anchors


Horizontal restraint force provided by geotechnical processes (buried posts or piles, embedded
𝐹⃗𝑔ℎ
logs)

𝐹⃗𝑔𝑣 Vertical restraint force provided by geotechnical processes (buried posts or piles, embedded logs)

𝐹⃗𝑝 Passive soil pressure force

𝐹⃗𝑠𝑜𝑖𝑙 Vertical loading on buried log due to weight of soil

g Acceleration of gravity

Large Wood National Manual July 2015


lxv
Bureau of Reclamation and
U.S. Army Corps of Engineers List of Symbols

Height of large wood structure (mean distance from stream bed to structure crest at upstream
s face)
p Rankine coefficient of passive earth pressure

l Unit vector along the axis of the buried log—positive in the direction away from the buried tip

Ls Length of large wood structure (dimension perpendicular to width, W)


L Length of log
Lc Appropriate moment arm about buried tip of horizontal log embedded in bank
Lem Embedment length—length of log that is buried in bank
Lex Exposed length of horizontal log partially buried in bank such that Lex + Lem = L
lk Length of kth log, exclusive of rootwad

Md Driving moment about buried tip of horizontal log embedded in bank

n Number of boulders used as ballast


Qx x-year return interval discharge
r DBH radius of the kth log
T Mean monthly temperature, degrees Fahrenheit
tk Thickness (measured in direction parallel to trunk) of kth rootwad
Uo Mean velocity of approach flow in the absence of large wood structure
Vw Volume of displaced water
Vd Volume of wood
Ws Width of large wood structure (dimension perpendicular to length, L)
Wbl(sub) Submerged weight of ballast
⃗⃗⃗⃗𝑏𝑙
𝑊 Weight of ballast

wk Radius of the kth rootwad


 Friction angle of soil
bl Specific weight of ballast
s Bulk unit weight of soil

structure Bulk density of large wood structure

w Unit weight of water

d Unit weight of wood

bed Coefficient of friction between large wood and bed


p Passive soil pressure
'v Vertical effective stress on buried log

Large Wood National Manual July 2015


lxvi
Chapter 1
LARGE WOOD INTRODUCTION

Complex timber revetment along South Fork Nooksack River, Washington


(Tim Abbe 2012)

AUTHORS

Tim Abbe (NSD)


Leo Lentsch (ICF International)
Martin Fox (Fox Environmental)
Chris Earle (ICF International)
This page intentionally left blank.
Has the channel been altered?
1.1 Need for and Purpose 
(e.g., armored banks, levees,
of this Manual straightening, incision)?

This national manual was developed to provide a  Is there point or non-point pollution?
broad range of resource managers (surface and  Has there been a loss of riparian
ground water, forestry, fish and wildlife, vegetation?
watershed, land, etc.) and specifically restoration
practitioners (engineers, geomorphologists,  Has there been a reduction in the
hyporheic exchange?
ecologists, landscape planners, etc.) with a basic
understanding of the role of wood in fluvial Questions that help define solutions:
aquatic and riparian ecosystems and how it  Has wood in the watershed ever been
should be maintained, reintroduced, and/or harvested?
managed. It highlights the best available science,
creative engineering, and policies associated with  How large were riparian trees under
restoring wood in rivers and underscores the old-growth conditions?
significance of wood in fluvial ecosystems. It is  What are the width, depth, and gradient
also a source of practical information on how to of the channel in which you are working?
assess the need for wood, use wood in restoration
projects, and manage wood that naturally enters  How have riparian forest conditions
rivers and streams. To this end, this national changed over time?
manual provides resource managers and  How do undisturbed historic channel
restoration practitioners with comprehensive characteristics compare to current
guidelines for the planning, design, placement, conditions?
and maintenance of large wood in rivers and
 Bed substrate 7
streams with an emphasis on restoring ecosystem
process and function.  Sinuosity

Resource managers and restoration practitioners  Anabranching (presence of


with objectives to restore ecological functions of ephemeral and perennial side
streams and rivers are faced with many questions channels)
when using large wood, such as the following:  Hydraulic geometry (unvegetated
Questions that define the problem: width and depth)

 What are current problems with the  Alluvial landforms


ecosystem?  What was undisturbed historic
 Is there degradation of: floodplain?

 Instream habitat (pools, cover,  What is current floodplain?


substrate conditions)  Has the channel experienced aggradation
 Floodplain environment (immature or incision?
forests, loss of wetlands)  What is the current instream wood
 Water quality (excess fine sediment, loading in terms of pieces and volumes?
nutrients, high temperatures,  How have infrastructure decisions
pollution) governed the management of wood?
 What are potential contributing factors?
 Has the flow been altered?

Large Wood National Manual July 2015


1-1
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

GUIDANCE

To make this national manual a practical tool that speaks to these types of questions and assists resource
managers and restoration practitioners, it includes the following subjects:
 Ecological restoration, large wood, an overview of the ecological functions of wood, and history of wood
management and restoration in streams (Chapter 1, Large Wood Introduction).
 Application of the ecological restoration process and decision support tools for projects using large wood
(Chapter 2, Large Wood and the Fluvial Ecosystem Restoration Process).
 Maintaining and restoring biological function in streams with wood (Chapter 3, Ecological and Biological
Considerations).
 Understanding the role of wood in geomorphologic and hydrologic function (Chapter 4, Geomorphology and
Hydrology Considerations).
 Large wood management considerations at a large geographic as well as long-term temporal scale (Chapter 5,
Watershed-Scale and Long-Term Considerations).
 Designing and engineering wood projects (Chapter 6, Engineering Considerations).
 Recognizing the risks of using wood for restoration (Chapter 7, Risk Considerations).
 Identifying the regulatory requirements associated with wood products, and implementing wood restoration
projects (Chapter 8, Regulatory Compliance, Public Involvement, and Implementation).
 Understanding and documenting project success (Chapter 9, Assessing Ecological Performance).

intervals (Agee 1993). The distribution of tree


1.2 Ecological Restoration species, heights, diameters, and stem densities in
An ecosystem is a complex of living organisms, distinct ecoregions often differs due to variations
their physical environment, and all their in elevation, aspect, precipitation/soil moisture,
interrelationships in a particular unit of space. An and temperature (Henderson et al. 1992; Agee
ecosystem’s abiotic (nonbiological) constituents 1993). In 1981, the map of ecoregions was
include minerals, climate, soil, water, sunlight, expanded to include the rest of North America
and all other nonliving elements; its biotic (Bailey 2009), and an explanation of the basis for
constituents consist of all its living members. the regions delineated on the map was provided
Ecosystems at any site are governed by later (Bailey 2009). In 1993, as part of USFS’s
hierarchical regional, watershed, and reach-scale National Hierarchical Framework of Ecological
processes controlling hydrologic and sediment Units (ECOMAP 1993), the ecoregions were
regimes; floodplain and aquatic habitat dynamics; adopted for use in ecosystem management
and riparian and aquatic biota. In 1978, the U.S. (Figure 1-1) (Bailey 1995). As such, resource
Forest Service (USFS) recognized these managers and restoration practitioners can
relationships and categorized ecosystems with expect ecosystems within these areas to have
similar characteristics across the United States by similar functions.
mapping them into ecoregions (USDA 1980).
For the purposes of this manual, we have used the
Ecoregions are characterized by climax species,
term “ecosystem function” to define the
tree size, and density of forest stands as
biological, geochemical, and physical processes
influenced by climate and fire disturbance
and components that occur and interact within an

Large Wood National Manual July 2015


1-2
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

ecosystem. This includes the functional processes decomposition, and trophic interactions. Studies
and mechanisms that maintain the ecological of ecosystem function have greatly improved
structure and services produced by ecosystems. human understanding of sustainable production
For example, ecosystem functions include of forage, fiber, and fuel, as well as the provision
primary productivity (production of biomass), of clean water.

Figure 1-1. Map of Ecosystem Divisions, Regions, and Providences Across North America

Source: eathsciences.org.

Rivers and streams are a defining component of manual, a fluvial ecosystem includes the river
the landscape and the foundation of fluvial corridor that extends in cross-section from the
ecosystems (Figure 1-2). For the purposes of this channel’s invert or thalweg to a point on the

Large Wood National Manual July 2015


1-3
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

landscape where channel-related surface and/or The length of a river corridor is typically
soil moisture no longer influence the plant characterized by the valley that encompasses the
community (Ward et al. 1999, 2002). Ecological channel from the headwaters to the mouth of the
attributes of rivers and streams are defined by watershed (Figure 1-2). Rarely does an alluvial
their geographic location, underlying geology, valley consist of a single channel, but usually
topography (e.g., slope), climate and hydrologic comprises a complex mosaic of both perennial
characteristics, and biological characteristics and ephemeral channels and floodplain wetlands
(i.e., aquatic, terrestrial, and subterranean (Abbe and Montgomery 2003; Montgomery and
species). Abbe 2006; Abbe and Brooks 2011).

Figure 1-2. Schematic of a Fluvial Aquatic Ecosystem

Modified from Stanford and Ward (1998).

Restoring process and function to damaged or considered restored when it contains sufficient
altered fluvial aquatic ecosystems is a basic biotic and abiotic resources to continue its
tenant of ecological restoration. Ecological development without further human assistance
restoration encompasses a set of intentional or intervention. It will sustain itself structurally
activities that initiates or accelerates the recovery and functionally, and will demonstrate resilience
of an ecosystem with respect to its health, to normal ranges of environmental stress and
integrity, and sustainability. The Society for disturbance. As a central component of these
Ecological Restoration defines ecological restoration activities, the use of wood plays a
restoration as “the process of assisting the critical role in the restoration of fluvial aquatic
recovery of an ecosystem that has been degraded, ecosystems.
damaged, or destroyed” (Society for Ecological
Restoration International 2004). An ecosystem is

Large Wood National Manual July 2015


1-4
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Although the type of material—logs, branches,


GUIDANCE
rootwads—is generally accepted by all, there are
no absolute size criteria for what is sufficiently
Adherence to four process-based principles can ensure
“large.” Minimum diameters of between 10 and
river restoration actions will be guided toward
maintaining and/or establishing sustainable 25 centimeters (4 and 10 inches) are common
ecosystems (Beechie et al. 2010): criteria in the published literature (Keller and
Swanson 1979; Bilby and Ward 1989; Beechie
1. Restoration actions should address the root causes and Wyman 1992; Montgomery et al. 1995a;
of degradation. Schuett-Hames et al. 1999; Fox and Bolton 2007).
2. Restoration actions must be consistent with the The minimum length of large wood, however, has
physical and biological potential of the site. less agreement. Bilby (1984) suggests that any
piece shorter than 2 meters (7 feet) may be
3. Restoration actions should be at a scale
unstable, and Bilby and Ward (1989) counted
commensurate with environmental problems.
none shorter than 2 meters in their study;
4. Restoration actions should have clearly articulated Montgomery et al. (1995a) counted any piece
expected outcomes for ecosystem functions. longer than 1 meter (3.3 feet); meanwhile, the
Oregon Department of Forestry (1995) requires a
length double to that of the bankfull width, and
Within this broad context, restoration actions that
the National Marine Fisheries Service (1996)
restore fluvial aquatic ecosystem function by
requires lengths of 15.25 meters (50 feet) in west
enhancing wood in streams and river channels
Washington State and 10.7 meters (35 feet) in
can include activities that range from protecting
east Washington State. Researchers such as Wohl
riparian forests and the sources of wood in
et al. (2010) recognize the importance of
channels to replacing wood in channels. As such,
reporting the specific minimum sizes measured,
any decision to place wood in channels should
the proportions relative to the low-flow and
only be made after carefully assessing the need
bankfull channel zones, and actual dimensions for
and benefits associated with that action. As stated
sorting data and enabling more universal
earlier, this manual provides guidance for
comparisons among data sets. For the purposes of
resource managers and restoration practitioners
this manual, minimum dimensions are only
faced with making those types of decisions.
provided where necessary for clarity. The most
important concept associated with the definition
1.3 Large Wood of large wood is the functional role it plays in
fluvial ecosystems. As such, the specific size of a
Wood is part of a continuum of allochthonous given piece of wood and its specific ecosystem
plant material from outside the stream itself and function can change based on the relative size and
can include leaves, twigs, branches, trunks, and location of the alluvial ecosystem in which it is
root masses. All of these provide structure at located.
various scales, nutrients (decomposing at varying
rates), and habitats for aquatic species including Wood has been used as a principal structural
material for thousands of years, including
fish.
applications in water bodies such as bulkheads,
Large wood generally refers to tree trunks or root piers, docks, and abutments. The use of wood to
masses of varying dimensions that contribute improve fish habitat in streams goes back at least
especially to the physical structure of the fluvial to the late nineteenth century (e.g., Van Cleef
system as it relates to larger organisms, especially 1885). Guidelines on using wood to enhance
fish. It is worth noting that even with widespread habitat and protect stream banks continued
usage in both scientific and agency literature, through the twentieth century (e.g., Hewitt 1934;
“large wood” has no universal definition. Tarzwell 1936; Ahmad 1951; Saunders and Smith

Large Wood National Manual July 2015


1-5
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

1962; Sedell et al. 1982; Seehorn 1985; 1.3.1.1 Tree Type


Thompson 2002, 2005). Despite this historical
context and the recognition that natural wood had The types or species of trees in a riparian forest
significant effects on river morphology (e.g., Wolff affect many functional characteristics. Evergreen
1916; Guardia 1933), research into the form and trees provide shade and perform other
function of natural wood didn’t begin until late in microclimate functions throughout the year, while
the twentieth century (e.g., Keller and Swanson deciduous trees provide few such functions when
1979; Bilby and Likens 1980; Harmon et al. 1986; leaves have fallen. It is important to realize,
Maser et al. 1988). however, that in many settings deciduous trees
may outperform evergreen trees because they
Historically, wood placements focused on simple shade when seasons are hot and then allow
structures (e.g., deflectors and steps) and a static warming when seasons are cold, thereby
simple view of stream channels (e.g., Thompson improving primary productivity during times of
2002, 2005). Recent recognition of the inherent food shortages. Trees with strong, decay-resistant
spatial and temporal physical complexity of wood provide more durable woody debris than
natural wood accumulations and their beneficial trees with weak or easily decayed wood (Harmon
influence on fluvial ecosystems has driven et al. 1986). Trees in the willow family fix
increased efforts to re-introduce wood to streams nitrogen in a form available to organisms in the
throughout the United States. Current wood forest and the stream (Wuehlisch 2011), whereas
placement strategies such as engineered logjams other tree species lack this capacity. Trees also
differ markedly from historic wood placements in vary in their response to channel disturbance; for
both the complexity of the structures and intent instance, some can survive having their roots
to restore complexity and natural process to buried by layers of sediment, while others cannot.
disturbed stream channels and floodplains (e.g.,
Abbe et al. 1997; Abbe and Brooks 2011). 1.3.1.2 Tree Size and Stand Density
Tree size affects the potential for the riparian
1.3.1 Importance of Riparian forest to provide functional large wood in the
Forests channel, to shade the channel, to provide root
reinforcement of stream banks, and to survive
Most sites that are candidates for large wood
channel disturbance. Large trees produce larger
placement are in areas where humans have
woody debris that is more likely to remain in the
played a substantial role in the history of the
stream and provide geomorphic and ecological
riparian zone, reducing the size and abundance of
functions, compared to smaller trees that may not
trees relative to what might be found in a pristine
be recruited to the stream or, if recruited, may be
setting. Human activities may have also altered
swept away during high flows (Harmon et al.
the stream channel; for example, by altering
1986; Lienkaemper and Swanson 1987). Tall
fluvial disturbance regimes or by engineered
trees can shade wider streams or provide greater
channel alterations, thereby altering interactions
shade in small streams (Beschta et al. 1987).
between the riparian forest and the stream. The
Large trees also have extensive root systems and
potential array of impairments is extensive,
may provide a barrier to trap flood-borne debris;
including all aspects of riparian forest function,
both functions can reduce the risk and rate of
and is usually highly site-specific.
channel migration or avulsion (Coho and Burges
Although riparian forest types vary widely 1994).
throughout the coterminous United States, they
Stand density affects the magnitude of the
can broadly be characterized according to the
functions described above. For instance, a less
type, size, and density of trees; the width of the
dense stand will provide less shade and less wood
forest; and the degree of channel confinement.

Large Wood National Manual July 2015


1-6
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

recruitment than a similar stand with greater tree trees; for example, debris torrents or dam-break
basal area. floods (Coho and Burges 1994). As channel
confinement reduces, the riparian forest is
1.3.1.3 Riparian Forest Width potentially much wider and is capable of
providing all of the physical and chemical
Riparian forests are commonly long and narrow. ecosystem functions described earlier. With
There are notable exceptions, the principal one further reductions in channel confinement, the
being forests on the floodplains of large rivers.
stream has the potential to shift its channel by
However, in most of the United States, riparian migration or avulsion, and to develop side
forests are narrow due to natural reasons (soil channel environments that may host valuable
moisture becomes limited with increasing
natural resources such as wetlands or sensitive
distance from the stream) or from human activity species habitat. In such settings the riparian
(removal of trees away from the stream, leaving a forest may provide the functions described above
riparian forest strip, the width of which often
to multiple channels and larger water bodies such
reflects a regulatory requirement). Many studies
as ponds or small lakes.
have attempted to describe the progressive loss of
function that occurs in progressively narrower
strips of riparian forest (reviews by FEMAT 1993; 1.3.2 Wood Loading in
Castelle et al. 1994) and have generally found that Natural Settings
different ecological functions diminish at different
rates with distance from the stream. Effects of the As described above, one common question for
forest on wind speed, for instance, continue to restoration projects is “how much wood is
accrue even over distances of hundreds of meters enough?”. There is very little to no data from
(Chen et al. 1995), while large wood recruitment around the country on natural wood loading, but
primarily occurs over distances of less than one there are several important concepts to
tree height (McDade et al. 1990; Robison and remember. Natural wood loading varies widely
Beschta 1990). and thus offers designers a great deal of flexibility
if using reference conditions, assuming they are
available. Practitioners should therefore focus on
1.3.1.4 Channel Confinement
the function of the wood, not just whether the
Stream channels can be broadly categorized quantity is representative of “natural” loading.
according to degree of channel confinement,
Fox and Bolton (2007) show that wood loading,
represented as the ratio of valley bottom width to
as measured by the number of pieces per bankfull
bankfull width (Montgomery and Buffington
channel width, increases with channel size and
1993). The riparian forest is largely confined to
drainage area (Figure 1-3). This runs counter to
the valley bottom. On the valley sides, soil
common views that wood loading diminishes in
moisture is usually reduced and vegetation
larger channels. While the spacing of wood
changes to a different type. On many sites, the
accumulations tends to increase in larger
riparian forest is confined to a fraction of the
channels, the size of the accumulations increases
valley bottom, as described above (Section 1.3.1.3,
dramatically. Some of the largest wood
Riparian Forest Width). In a tightly confined
accumulations historically were found in the
channel, the riparian forest is necessarily very
response lower reaches of large rivers such as the
narrow. Principal functions of the riparian forest
Red and Colorado rivers described earlier. Large
on these sites primarily relate to bank
logjams were once common in the distributary
reinforcement by roots and to other mechanisms
channels of large river deltas in the Pacific
by which the forest may alter the severity of
Northwest (Abbe 2000). In 2005 a logjam began
channel disturbance, either by resisting or by
forming in the main channel of the Nooksack
failing to resist peak flow events that may entrain
River delta in northwestern Washington, growing

Large Wood National Manual July 2015


1-7
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

to 14 acres by 2011. The formation of these types positive sign of passive restoration (Figure 1-4).
of logjams reflects natural processes and a

Figure 1-3. Wood Loading Tends to Increase With Channel Size When Normalized to Bankfull Width

Adapted from Fox and Bolton (2007).

Large Wood National Manual July 2015


1-8
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Figure 1-4. Example of High Wood Loading in a Large Channel (Nooksack River Delta, Northwest
Washington)

A logjam began forming about 2005 in what had been the dominant channel (initiating at point A). Most of the
46-meter-wide (150-foot-wide) channel was filled with wood by 2011, and most of the flow is in the western
channel. The logjam covers about 14 acres, and the larger logs are about 0.6 meter (2 feet) in diameter and
21 meters (70 feet) long. The logjam represents an example of passive restoration (formed naturally) that is
increasing physical, hydrological, and ecological complexity in the delta. These types of logjams are well
documented in gulf coast deltas (e.g., Clay 1949; Wadsworth 1966; Hartopo 1991; Phillips 2012).

Large Wood National Manual July 2015


1-9
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

1.3.3 Historical Instream that remnant stands of eastern old-growth mixed


conifer-hardwood or hemlock-white pine stands
Wood Conditions have a biomass of 560 to 820 tons per hectare
Historically, fish and other aquatic species (Whitney 1996). This is nearly as much as that of
adapted to stream systems where wood was the productive coniferous forests of the Pacific
abundant and distributed in the form of Northwest (Franklin and Dyrness 1973). Whitney
individuals and groups (jams) recruited from also estimated that white pine trees often grew to
riparian trees, beaver dams, and other means diameters of 2–3 meters (7–10 feet) and heights
(Sedell and Luchessa 1981). Large river systems of 45–60 meters (148–197 feet); while sycamores
contained so much wood that river navigation attained diameters greater than 4 meters
was nearly impossible, with some jams up to (13 feet), tulip trees almost 2 meters (7 feet), and
1,500 meters (4,920 feet) long (Sedell and cottonwood and oak well in excess of 2 meters
Luchessa 1981). Collins et al. (2002) determined (7 feet), which is also within the range of
that wood in some lowland Puget Sound rivers unmanaged Pacific Northwest stand
was one to two orders of magnitude greater prior characteristics. Sedell et al. (1982) reported that
to European settlement based on historical data, snags in the lower Mississippi River, pulled over a
journal accounts, and observations in undisturbed 50-year period, had diameters averaging
river reaches of the Nisqually River. These 1.5 meters (5 feet) at the base. Triska (1984)
authors found reports from early surveyors, reported diameters of 1.75 meters (5.75 feet) in
U.S. Army documents, and others referencing the the Red River of Louisiana. Based on the
vast amounts of wood in the major rivers: similarities in historic stand characteristics
between eastern and northwest forests in North
….the channels are strewn with immense trunks, America, similarities of instream wood loads are a
often two hundred feet long, with roots, tops, and
reasonable assumption. Verification may be
all …[forming] jams, which frequently block the
channels altogether (Major Hiram Chittenden supported by the same approach taken by
1907) Montgomery et al. (2003), who suggest using
references to historical records of instream wood
Snags are numerous and large, and so deeply
removal and clearing of riparian forests to
imbedded in the bottom that a steam snag-boat
would be required for five or six months to open reconstruct and evaluate how the role of wood in
a channel 100 feet wide… (U.S. Army’s Robert A. some river systems has changed in the last
Habersham 1897) 200 years.
The amount of wood was so abundant and well- The use of wood for improving instream fish
lodged into riverbeds that logging and upstream habitat goes back at least to the late nineteenth
settlement was stymied until settlers and USACE century (e.g., Van Cleef 1885; Hewitt 1934;
could pull, blast, and cut wood from rivers in the Tarzwell 1936; Thompson 2002, 2005). In
1870s to 1890s (Sedell and Luchessa 1981). reviewing historic stream improvement,
Thompson (2002, 2005), concluded there is little
In terms of large wood piece sizes, Collins et al.
evidence the constructed structures made a
(2002) found that the annual maximum diameter
measureable improvement to habitat or restoring
between 1889 and 1909 ranged from 3.6–
natural channel conditions, and in some cases had
5.3 meters (11.8–17.4 feet) (U.S. War Department
negative impacts. Thompson (2002, 2005) argues
1889–1909), based on snag boat captains’ records
that much of the recent restoration work using
and confirmed by engineers’ observations (e.g.,
wood is similar to historic structure types (such
U.S. War Department 1895).
as log steps and log crib deflectors), and thus is
Historic instream wood loads in other parts of unlikely to achieve the desired results,
North America may have been similar to those in recommending that restoration avoid static
the Pacific Northwest. Whitney (1996) reported structures and focus on restoring riparian

Large Wood National Manual July 2015


1-10
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

corridors and unsecured wood placements. This 1.3.4 Wood Recruitment


argument is challenged by Roni et al. (2014a) who
demonstrate that recent wood placements are Processes
successfully restoring habitat and natural Every stream has a unique hydrologic and
processes. While Thompson (2002, 2005) is sediment regime linked to the climate, geology,
correct that there are similarities between the relief, vegetation, and landscape disturbance
simple static structures constructed throughout within its watershed. Hydrologic regimes vary
the 1900s and some of the current restoration substantially around the country, such as the
design commonly associated with “natural range of flows that can recruit, move, and deposit
channel design” (Rosgen and Silvey 1996), there wood. In some regions these will differ by orders
is a significant difference with regard to the of magnitude. Fluvial processes can also vary. In
recent engineered logjam approach, in which some cold regions of the country, river ice can
wood structures are not just intended to emulate play a dominant role in channel morphology (e.g.,
natural structures, but restore and accommodate Pariset et al. 1966; Keller and Swanson 1979;
fluvial processes such as bank erosion, hydraulic Smith 1979; Beltaos 1983; Smith and Pearce
variability, and wood recruitment (e.g., Abbe and 2000), and, in steep terrains, debris/mud flows
Brooks 2011). Most of the current stream have pronounced effects on channel form and
restoration involving wood placement includes wood. Similarly, the frequency and magnitude of
land acquisition or conservation easements with processes delivering wood from hillslopes and
aggressive riparian reforestation. There is floodplains vary geographically, and the
certainly consensus in the scientific community characteristics of individual forest trees (size,
(including Thompson 2002, 2005; Abbe and shape, specific gravity) have a major effect on the
Brooks 2011; Roni et al. 2014a) that long-term deposition and transport, and, therefore, the
stream restoration depends on understanding distribution of, wood within a channel network.
and accommodating processes at both the reach
and watershed scale. Key to this is establishing a Geomorphically, fluvial disturbance rates in large,
geomorphic response corridor that includes unconfined rivers dictate both the size of trees
restoration of mature riparian forests and and the species recruited to the channel, where
sufficient portions of the floodplain, channel deciduous trees colonize rapidly following
migration zone (e.g., Rapp and Abbe 2003), and disturbance and are later succeeded by coniferous
adjacent hillsides to accommodate fluvial stands (Naiman et al. 1992; Fetherston et al.
processes and wood recruitment. Roni et al. 1995; Johnson et al. 2000). Recruitment also
(2014a) point out that wood placement projects comes from non-fluvial means, where Palik et al.
that did not take into account processes such as (1998) reported wood enters the channel due to
hydrology or sediment, tended to be the projects the natural mortality of trees in reaches having
that did not show improvements. narrow valleys with riparian landforms elevated
above the channel. Collins et al. (2012)
demonstrate how logjams in the Pacific
CAVEAT Northwest formed by large trees create “islands”
where the disturbance frequency associated with
Stable wood placements can be essential project channel migration (bank erosion) is reduced and
elements but should always be designed in a reach and allows for large trees to develop within
watershed context that help to restore and floodplains characterized by much smaller trees.
accommodate the natural spatial and temporal In a detailed wood budget covering
dynamics of wood and channel processes essential to
177 kilometers (110 miles) of the Roanoke River
sustain healthy streams.
in the Mid-Atlantic Coastal Plain (North Carolina),
a large low-gradient sand-bedded channel, Moulin
et al. (2011) found that bank erosion accounted

Large Wood National Manual July 2015


1-11
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

for over 70% of the wood. Of the instream wood, that promote the colonization of deciduous
75% was available for transport, over 50% of species (Naiman et al. 1992; Fetherston et al.
which was stored in logjams, most formed by 1995; Johnson et al. 2000). During periods of high
stable snags. The same researchers also looked at flows, channel avulsion, accelerated lateral
wood transport in the Roanoke River, tagging migration, and bank land sliding can topple trees
344 pieces of wood (290 with radio frequency from riparian areas (Johnson et al. 2000).
tags, 54 with aluminum tags) (Schenk et al. Deciduous trees typically are first to colonize
2014b). They found that 5% of the instream wood riparian areas following disturbances, the causes
turns over (losses from export, decomposition, of which can be both direct channel action and
and burial equal inputs from mass wasting and debris flows (Grant et al. 1984; Wilford et al.
bank erosion) and that 16% is moving through 1998) or snow avalanches (Fetherston et al. 1995;
the system. The remaining population consists of Cushman 1981). Following these disturbances,
individual snags and logjams. conifer succession may not occur for 80 years or
more (Jenkins and Hebertson 1998). Disturbance
Fire is also a dominant influence that affects
patterns can affect the characteristics of the
timber age (Henderson et al. 1992), tree diameter
riparian area and thus influence the
(Rot et al. 2000) and height (Agee 1993;
characteristics of wood loads. Wind throw, insect
Henderson et al. 1992), and recruitment to
infestations, drought, disease, ice storms, and fire
streams. Patches of timber unscathed by a fire
all affect recruitment rates, stand age, wood
(often termed fire refugia) can diversify timber
diameter, and species composition.
ages along riparian areas (Camp et al. 1996).
The ability of wood to have a significant effect on
Other important recruitment processes in the
hydraulics and stream channel morphology is
eastern, southern, and Midwest regions are linked
dependent on its stability. Wood that is easily
to severe weather, particularly hurricanes and
transported is unlikely to remain in the channel
wind storms (e.g., Frangi and Lugo 1991; Foster
for any length of time unless it encounters stable
and Boose 1992; Boose et al. 2001; Chambers et
obstructions. Unstable wood that doesn’t simply
al. 2007; Phillips and Park 2009), and ice storms
pass through the system usually ends up
(e.g., Millward et al. 2010). Recruitment processes
entangled on a pre-existing snag or log jam (flow
are discussed in Section 1.3.4.2, Wood
obstruction) or on a depositional surface such as a
Recruitment through Natural Disturbance
bar or floodplain. When wood forms a stable
Regimes. These processes drive the rate of wood
obstruction within a zone of active sediment
recruitment as well as the structure and
transport, it begins to alter channel-forming
composition of instream wood.
processes by influencing flow conditions, scour,
and deposition. Those pieces of wood that have
1.3.4.1 Riparian Contribution sufficient resistance to withstand the forces
As introduced above, geomorphic processes, imposed by peak flows are most likely to become
disturbance patterns, and regional climate local hydraulic and geomorphic controls that
differences influence the structure and define riffle formation. The effect of these key
composition of riparian forests, both spatially and pieces is further exaggerated when they trap
temporally. The effects of fluvial activity on additional wood debris that would otherwise
riparian forests are predominantly associated have passed through the channel. Key pieces
with large rivers, because smaller streams do not commonly form obstructions where they first
have the same energy and consequent rates of enter the channel, and if they do move
channel migration and bank erosion sufficient to downstream, they don’t tend to go far. The
affect large swaths of riparian forest. In contrast, presence of key members is strongly dependent
the riparian floodplains of large, unconfined on a local sources of trees capable of creating
channels are developed by fluvial disturbances stable snags, either because of their size or shape

Large Wood National Manual July 2015


1-12
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

(e.g., intact rootwad, large branches, or clump of z = perpendicular downslope distance


multiple stems attached to same rootwad). The from standing tree to nearest channel
supply of mobile wood debris to the channel will boundary
affect the development of wood accumulations or
h = tree height
jams downstream. Streams that have young
riparian areas of small trees and channels that The number of fallen trees that enter a stream, NI,
have been cleared of natural obstructions can is predicted as a function of the riparian forest
have an artificially high supply of mobile wood, width, WF, and length, LF; the density of trees; the
which can end up forming massive logjams in probability of a tree falling, PF (1); and the
depositional areas (e.g., deltas, reservoirs) or at probability of a tree falling into the stream, PS.
artificial obstructions such as bridge piers and
culverts, The restoration of key pieces or ELJs in Equation 1-3:
unconstrained stream segments will trap mobile
NI = D LF WF PF PS
debris and thus help to control the downstream
flux of wood that may pose a threat and will also Data of large wood and channel dimensions from
increase carbon sequestration. Kennard et al. western Washington were used to develop an
(1998) present a model for managing riparian empirical means of estimating the diameter of
forests based on the supply of adequate large functional wood. Using an empirical estimate:
wood to stream channels. They start with a
simple prediction of the probability of a tree Equation 1-4:
falling due to mortality or windfall.
dbh  3.06 WC  22.10
Equation 1-1:
where:
PF = (1-(1-TF)t)
dbh = tree diameter at breast height in
where: centimeters
PF = probability of a given tree falling after WC = channel bankfull width in meters
time, t, in years
Kennard et al. (1998) present the following wood
TF = tree fall rate, assumed to be 20% for depletion model for estimating the percentage of
first decade, 15% for second decade, large wood pieces remaining after time t:
and 10% thereafter (equilibrium rate
based on Murphy and Koski 1989) Equation 1-5:
The probabilistic tree-fall model of Van Sickle and y = (1-x)t
Gregory (1990) is used to predict large wood
where
recruitment to a channel.
x = annual depletion rate 0.015 (based on
Equation 1-2: values of 0.014–0.016 from Murphy
cos 1 z h  and Koski 1989)
PS 
 t = elapsed time in years

where: y = % of large wood pieces remaining


after time, t
PS = probability of a tree falling
Kennard et al. (1998) offer the following means of
estimating pool spacing based on the number of
functional large wood pieces, NF, in a channel.

Large Wood National Manual July 2015


1-13
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Equation 1-6: (Harmon et al. 1986). Source distance is


correlated to tree height (McDade et al.1990;
 1   J(K )  S Robison and Beschta 1990), but McDade et al.
G     
 N F   WC R   (1990) could not attribute 47.7% of identified
wood pieces to an adjacent source, suggesting
where: that many pieces are routed in from upstream
sources. Clearly, instream wood loads are
G = number of channel widths, WC, per dynamic and fluctuate according to various
pool natural processes at the reach and watershed
scales. The following elaborates on these
L = length of stream segment
processes.
R = recruitment factor
Geomorphic Influence
J = proportion of pools from debris jams
Channel reach morphology, such as the types
S = proportion of single large wood pools identified by Montgomery and Buffington (1997),
K = number of functional large wood also influences instream wood loads. Rot et al.
pieces in debris jams (2000) found significantly more large wood
pieces in forced pool-riffle channels than in
The above discussion can provide guidance on the bedrock or plane-bed channels, where wood
recovery of riparian areas to the point where they volume followed a similar trend. However, these
are making a geomorphic difference for both authors and others found that confinement was
passive and active approaches to restoration. significantly related to large wood volume only in
forced pool/riffle channels, where less wood was
1.3.4.2 Factors of Variability found in confined channels. They report that
confinement had no effect on large wood volume
There are several means by which wood finds its
in plane-bed channels.
way into a stream. At the reach scale, trees can fall
directly into a channel due to bole breakage or by Fox and Bolton (2007) illustrate an increase in
being uprooted. These are often the result of large wood piece numbers and volumes as
various forms of chronic tree mortality such as channels increase in width. Fox and Bolton found
stem suppression/exclusion, wind throw, disease, 0.38 pieces/meter in the smallest channels
and old age, and also the result of fluvial (>0–6 meters [0–20 feet] bankfull flow width
processes such as channel avulsion or lateral [BFW]) to 2.08 pieces/meter in the largest rivers
migration and bank erosion. At the watershed (30–100 meters [98–328 feet] BFW). This
scale, other processes such as debris flows and difference between the studies could be
snow avalanches can deliver trees into attributed to the inclusion of large rivers
downstream channels (Cushman 1981; Grant and (20–100 meters [66–328 feet] BFW) studied by
Swanson 1995). The river can also exhume buried Fox (2001), which displayed many multi-piece log
wood within floodplains (Fetherston et al. 1995). jams.
Ultimately, the quantity of wood in a stream at
any point in time is a result of input and output Fox (2001) reported that small channels are likely
balances over the previous centuries (Swanson et to obtain a significant proportion of riparian trees
al. 1982; Martin and Benda 2001). for instream wood by bole breakage and passive
tree mortality, rather than by active recruitment
Instream large wood biomass is positively such as the lateral bank avulsion common to
correlated to tree density (Bilby and Wasserman larger streams. Similarly, confined streams
1989), tree maturity (Bilby and Ward 1991; Rot draining large basins are also likely to obtain
et al. 2000), and the percentage of conifers wood passively. Fox observed that confined

Large Wood National Manual July 2015


1-14
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

reaches often had resistant banks, which likely the characteristics of forest vegetation can be
slow the rate of avulsion compared to banks grouped by a forest zone or forest series (Franklin
composed of unconsolidated material. Due to the and Dyrness 1973; Agee 1993). Ecoregions are
resistance to lateral migration, trees adjacent to characterized by climax species, tree size, and
these channels are afforded greater intervals density of forest stands as influenced by climate
between disturbances and thus have the potential and fire succession (Agee 1993). The distribution
to grow older and perhaps larger (given favorable of tree species, heights, diameters, and stem
soils and climate). As a result, confined channels densities in distinct ecoregions often differs due
often have greater potential to recruit fewer but to variations in elevation, aspect,
larger trees than unconfined channels, where the precipitation/soil moisture, and temperature
lateral migration rate within the floodplain limits (Henderson et al. 1992; Agee 1993). These in turn
tree growth. influence wood loads.

The frequency of fluvial disturbance also dictates Each region in North America provides a unique
stand age, which influences large wood size. set of characteristics (Figure 1-5). The example
Latterell and Naiman (2007) found that larger used herein is from the Pacific Northwest, where
trees are not recruited from floodplain riparian instream wood loading data was compiled for
areas, but rather from higher surfaces less prone specific forest types. Comparisons could be made
to frequent fluvial disturbance. These authors to forests with similar stand characteristics to
reported that 72% of large trees (<1-meter estimate the relationship of riparian sources to
[3-foot] diameter) entering the Queets River in potential wood loading. Seven major forest types
Washington were recruited as the river undercut compose ecoregions in the Pacific Northwest and
higher fluvial terraces. These terrace surfaces had are described below.
not been disturbed by the river since the forest
stand origin. This is supported by the research of Sitka Spruce Forests
O’Connor et al. (2003), who report that channel This forest type is generally limited to the coastal
and floodplain dynamics and morphology are west-slope of the Pacific Northwest mountain
affected by interactions involving time frames ranges, particularly the western Olympic
similar to 200–500-year floodplain half-lives in Peninsula due to the unique climate
the Queets River. characteristics found there. The elevation of these
Riparian vegetation can influence the rate of forests is typically less than 300 meters (984 feet)
lateral migration of rivers. Dense root systems above mean sea level (amsl), and normally within
can armor banks, reducing bank erosion and 20 kilometers (12 miles) of the coast; however,
processes that promote lateral river movement. sites can be found farther inland up low-elevation
For example, Collins and Sheikh (2005) found an river valleys (Agee 1993). Dominant tree species
1898 USACE report describing dense growth of are the Sitka spruce (Picea sitchensis), with co-
alder, willow, and vine maple on the shores of the dominants of western hemlock (Tsuga
White River in Washington: “This brush affords heterophylla) and western red cedar (Thuja
complete protection from washing and plicata), and, to a lesser degree, Douglas-fir
undermining effects of the current. In a majority (Pseudotsuga menziesii) (Agee 1993). The annual
of cases where the brush has been removed, the precipitation of the Sitka spruce (SS) region is
river has begun to eat into its bank.” 200–300 centimeters (79–118 inches) and
includes a component of fog-drip. The air
Regional Ecological Influence temperatures are mild year-round (Franklin and
Dyrness 1973). The large dense timber of this
Adjacent forest vegetation, as noted above, region is attributed to climate, which facilitates
influences the sizes and quantities of instream tree growth. Indeed, Edmonds et al. (1993) found
wood. Regional climatic variations that control

Large Wood National Manual July 2015


1-15
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

stem densities in this region between 476 and of the last fire in these forests has been identified
508 per hectare (>5 centimeters [>2 inches] in by some researchers as over 1,100 years ago
diameter at breast height [dbh]), and tree stem (Fahnestock and Agee 1983). Although this is not
basal areas between 77 and 94 square meters generally applicable to the entire SS forest type, it
(92 and 112 square yards) per hectare. The date suggests that stand-replacement fires are rare.

Figure 1-5. North American Forests

Source: North American Forest Commission (2011).

Western Hemlock Forests co-dominant (Agee 1993). Although Douglas-fir is


dominant in the early seral stages following fire, it
This forest type is generally found in the interior
will eventually be succeeded by western hemlock
low elevations of western Washington, Oregon,
at late succession (Agee 1993; Henderson et al.
and Northern California. The elevation is typically
1992). The western hemlock (WH) forest type has
less than 800 meters (2,623 feet) amsl, although
greater extremes of moisture and temperature
this may vary + 60 meters (197 feet) depending
than the SS forest type (Franklin and Dyrness
on aspect and local climate differences
1973). The dryer summers are reflected in the
(Henderson et al. 1992). The dominant tree
wide spectrum of plant associations across this
species is western hemlock, with Douglas-fir (DF)
zone (Zobel et al. 1976). Fire frequency intervals

Large Wood National Manual July 2015


1-16
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

are generally less than 750 years, although Mountain Hemlock Forests
ignitions from Native Americans may have
This forest type is generally found on upper
increased this frequency in some areas (Agee
elevations to the west of the Cascade crest, but
1992).
below subalpine regions. There is substantial
The physical characteristics of the timber in this overlap with the silver fir forests; however,
forest type are well documented. Spies and mountain hemlock (MH) is generally more
Franklin (1991) reported that the average stem prevalent at higher elevations. The elevation of
densities of Douglas-fir (>100 centimeters this forest type is typically between 1,000 and
[39 inches] dbh) in late-successional stands 1,375 meters (3,280 and 4,511 feet) amsl,
ranged from 18–29 trees per hectare, while although this may vary + 60 meters (197 feet)
Hershey (1995) reported 6–90 trees per hectare depending on aspect and local climate differences
with stems >54 centimeters (22 inches). (Henderson 1996). The dominant climax tree
Tappeiner et al. (1997) reported basal areas in species is mountain hemlock, with the Pacific
old-stands range between 46 and 91, with a silver fir and subalpine fir (Abies lasiocarpa) as
median of 66 (square meters per hectare). Tree co-dominants (Agee 1993). Mountain hemlock
heights for two common plant association groups has been found at elevations up to 1,800 meters
in this forest type average between 60 and (5,900 feet) in Washington where aspect, latitude,
60 meters (200 and 225 feet), with mean and local climates are favorable. Winter
maximum heights reaching 87 meters (285 feet) temperatures are cool, but summer temperatures
after about age 300 (years) (Henderson can reach extremes of 26–30 degrees Centigrade
unpublished 1996). (°C) (79–86 degrees Fahrenheit [°F]) (Arno and
Hoff 1989). Fire return intervals are estimated to
Silver Fir Forests be around 500 years (Dickman and Cook 1989).
This forest type is generally found at moderate to Subalpine Fir Forests
upper elevations on the west-slope of the
Cascades. The typical elevation is between This forest type is generally found along the
800 and 1,200 meters (244 and 366 feet) amsl, Cascade crest and the interior of the Pasayten
although this may vary + 60 meters (197 feet) Wilderness in the North Cascades at elevations
depending on aspect and local climate differences above 1,300 meters (4,265 feet) amsl (Henderson
(Henderson 1996). The dominant tree species is 1996; Agee 1993) although this may vary
Pacific silver fir (Abies amabilis), with western +60 meters (197 feet) depending on aspect and
hemlock and Douglas-fir co-dominant at lower local climate differences (Henderson 1996). The
elevations and mountain hemlock (Tsuga annual precipitation is typically between 100 and
martensiana) co-dominant at upper elevations 200 centimeters (30 and 79 inches) (Agee 1993).
(Agee 1993). Winter temperatures are moderate, The prolonged winter snow-pack (often between
but with a 1- to 3-meter (3- to 10-foot) winter 7 and 8 meters [23 and 26 feet] in wetter zones),
snow pack (Franklin and Dyrness 1973). along with the coldest winter temperatures of all
Droughts are infrequent, and summer Pacific Northwest forests, limits growth compared
precipitation usually exceeds 15 centimeters to trees in lower elevation forests (Agee 1993).
(6 inches) (Minore 1979). Fire return intervals Summer temperatures can be relatively high,
are estimated to be between 300 and 600 years, reaching 26–30°C (79–86°F) (Agee 1993).
but can be more frequent at lower elevations Mountain hemlock is often found at the lower
(100–300 years) (Agee 1993). Silver fir (SF) trees boundaries of this forest type. The dominant
seldom survive major fires (Agee 1993); thus, fire climax tree species is subalpine fir (Abies
return intervals often are points of stand origin. lasiocarpa), with co-dominants of mountain
hemlock, lodgepole pine (Pinus contorta), and
Englemann spruce (Picea englemanni) (Agee

Large Wood National Manual July 2015


1-17
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

1993). Subalpine fir (SAF) and co-dominants are Mountains (Holstine 1992), Douglas-fir has
not well-adapted to surviving fires (Agee 1993) become more prevalent in many areas (Harrod
and fire return intervals, estimated to be around pers. comm. 2000). Ponderosa pine typically can
250 years (Fahnestock 1976) or 109–137 years reach 35–45 meters (115–148 feet) in height with
(Agee 1990), often are points of stand origin. some exceeding 55 meters (180 feet) (WWPA
1995).
Grand Fir Forests
Grand fir (GF) (Abies grandis) are typically found GUIDANCE
at elevations between 1,100 and 1,500 meters
(3,610 and 4,921 feet) amsl east of the Cascade
crest, although populations of grand fir can be Other Forest Regions of North America
found at low elevations of inland western North America had some of the largest forested areas
Washington (Agee 1993). GF forests generally on earth. Forest regions across North America each
separate ponderosa pine (Pinus ponderosa) have unique attributes, some of which provide
forests (see below) from SAF forests. A mixture of distinctions for instream wood loading. The continent
species characterizes this forest type, with is surrounded by oceans and seas of various
Douglas-fir as the climax dominant. Rarely is GF temperatures and climate. The National American
Forestry Commission has identified 19 forest ecological
the late-successional dominant species.
zones of North American (Figure 1-5), with the 5 major
Hardwood species are often found as co- zones defined as (1) the tropical climate in Southern
dominants. Fire intervals are frequent, often due Mexico; (2) the mild climate with wet winters and dry
to lightning strikes, producing a return interval of summers of the Pacific zone along the coastal regions
50–100 years in drier sites. from southern Alaska to southern California; (3) the
mountainous and dry western interior of the United
Ponderosa Pine Forests States and much of northern Mexico; (4) the humid
eastern two-thirds of the United States and southern
This species is typically found in dry, lower Canada, which have a humid climate with defined
elevation (1,200–1,800 meters [3,937–5,905 feet seasonal changes; and (5) the northern two-thirds of
amsl) sites east of the Cascades (Franklin and Canada and Alaska, as well as all of Greenland, which
Dyrness 1973). Ponderosa pine (PP) forests have arctic and sub-arctic climates. The most notable
contain a large co-dominant component of forest in North America is the taiga or boreal forest,
Douglas-fir (Agee 1993). Douglas-fir is always the which is a large expanse of mainly coniferous trees
co-dominant species in this forest type and is that covers much of central and southern Canada and
Alaska. There is also a large area of redwood forests in
typically suppressed by fire (Agee 1993).
California in the United States and tropical forests in
A natural fire-recurrence interval is typically Mexico. These forests have various levels of
between 11 and 24 years (Agee 1993). Due to productivity, as indicated by their biomass, a measure
of organic carbon.
frequent burns, fires are typically of low intensity;
therefore, the older ponderosa pines are rarely
killed unless fires are fueled by excess wood
buildup in the under-story (Agee 1993). Camp et 1.3.4.3 Wood Recruitment through
al. (1996) found ponderosa pines in portions of Natural Disturbance Regimes
these forests (Swauk Late Successional Reserve)
to have ages between 13 and 597 years, with a Natural Disturbances
mean of 127 and a standard deviation of 100. Fire
Instream wood loads vary over space and time
refugia are common in this forest type, and are
due to an array of natural disturbance processes
typically found on north-aspect slopes and in
(e.g., Hickin 1984; Keller and Swanson 1979;
confined channels (Camp et al. 1996). With fire
Abbe and Montgomery 2003; Phillips and Park
suppression, beginning in 1909 in the Wenatchee
2009). All channels have been affected by

Large Wood National Manual July 2015


1-18
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

disturbance of some kind, whether historic or Bank erosion can be the dominant wood
recent. Therefore, the characterization of wood mechanism in many parts of the channel network.
from a single survey provides a temporal “snap- Erosion over-steepens adjacent hillslopes and
shot,” documenting only a single instant in the triggers landslides that deliver trees to the
patterns of fluctuation. Wood accumulations are channel. Volumes of wood recruitment are
not constant, but rather fluctuate with typically highest in larger alluvial streams prone
disturbance cycles. The accretion of wood may to channel migration. Large, low-gradient
continue over time until capacities exceed an channels characterized by high banks of
ecological or morphological threshold, some of unconsolidated fine sediments are particularly
which result in a catastrophic removal by prone to bank erosion. As discussed earlier,
disturbance. The amount of instream wood, Moulin et al. (2011) found that bank erosion
therefore, represents a time since the last accounted for over 70% of the instream wood in
disturbance and the temporal conditions during the Roanoke River in Eastern North Carolina.
the recovery period. Several natural disturbances
Wood recruitment, transport, and deposition in
responsible for wood recruitment to channel
cold regions can be directly influenced by ice
networks are discussed below.
flows and ice jams. Ice flows can entrain wood
Bank Erosion and mow down riparian vegetation (e.g., Keller
and Swanson 1979; Smith 1979; Smith and
Bank erosion occurs throughout the channel Reynolds 1983; Hickin 1984; Smith and Pearce
network of a watershed. It occurs when the 2000; Prowse 2001). Large accumulations of
erosive forces acting upon a bank (shear stress, wood and logjams also occur in many northern
pore pressure) exceed the resisting forces rivers that are subject to freeze up (e.g., Hickin
(material internal shear strength and cohesion, 1984; Makaske et al. 2002) but not to large ice
root reinforcement). Bank stability decreases flows that scour the channel (Pariset et al. 1966;
with increasing slope, height, and shear stress. It Smith 1979; Beltaos 1983; Prowse 2001).
is also directly related to material properties of
the bank that define its strength or resistance, Severe Weather and Wind Throw
such as grain size distribution, internal shear
Storms that bring severe winds and rainfall can be
strength/friction angle, cohesion, water pore
a major wood recruitment mechanism to streams
pressure, stratigraphy, shear planes, and root
throughout the United States. Severe wind
reinforcement. Processes triggering bank erosion
capable of tree “blow down” is often associated
include the following:
with major storm fronts and can be further
 Shear stress imposed on bank by high flows. exasperated by local orographic effects.
 Shear stress at bank toe that undermines and Hurricanes contribute huge quantities of wood to
over-steepens and destabilizes the bank. streams within their path, primarily impacting
states along the Gulf Coast and Eastern seaboard
 Channel incision that over-steepens and (e.g., Frangi and Lugo 1991, Foster and Boose
destabilizes adjacent banks. 1992, Boose et al. 2001, Chambers et al. 2007).
 Lateral channel migration resulting from Chambers et al. (2007) predicted that Hurricane
instream sedimentation (sand and gravel Katrina in 2005 resulted in the mortality and
bars) that directs flow against channel banks. severe structural damage to approximately
320 million large trees, equal to 50-140% of the
 Removal of riparian vegetation that reduces net annual U.S. forest carbon sink. Zeng et al.
bank strength. (2009) estimated that tropical cyclones result in
 Flow constriction due to channel obstructions the mortality and damage of 97 million trees
such as landslides, snags, and logjams. annually from 1851 to 2000 in the continental

Large Wood National Manual July 2015


1-19
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

United States, primarily in the Gulf Coast region. (Agee 1993; Henderson et al. 1992), and wood
Hurricanes and severe thunder storms often are recruitment distance is a function of height
associated with flooding which is called out as a (McDade et al. 1990). Therefore, fire affects
separate recruitment process below. Ice storms instream wood diameter and recruitment
also result in major wood inputs, particularly in patterns.
deciduous forests of the Midwest and Northeast
Fires do not burn forests evenly. The variability in
states (e.g., Millward et al. 2010).
timber age due to stand-replacement fires
Wind throw, usually associated with severe illustrates that “old-growth” forests are clearly
storms, is a significant source of large wood not homogenous in their life cycles among forest
recruitment to streams in all parts of the United zones or within basins. Forest growth frequently
States (Lienkamper and Swanson 1987; Robison is interrupted prior to the maximum life span of
and Beschta 1990; Frangi and Lugo 1991; Foster many trees in forested basins, as suggested by the
and Boose 1992; Boose et al. 2001; Chambers et heterogeneity in forest ages within forest zones.
al. 2007; Phillips and Park 2009), and ice storms This likely adds diversity in tree sizes, densities,
(e.g., Millward et al. 2010).). In old-growth and rates of stem exclusion and mortality. Patches
riparian forests, wind throw does not topple of timber unscathed by a fire (fire refugia) can
whole trees as much as it recruits a greater diversify timber ages along riparian areas (Camp
proportion of branches and treetops to the et al. 1996). Fire affects the rate of recruitment
channel than in younger riparian stands, when the regenerated forest selectively thins,
especially in areas prone to strong winds or heavy dropping the younger trees, out-competed by
snowfall (Bisson et al. 1987). However, wind larger, more dominant trees. This occurs in stands
throw accelerates mortality in riparian areas <220 years old (Rot et al. 2000). This may also
abutting newly harvested forests, disrupting the explain why measured instream large wood
rate of recruitment to streams (Grizzel and Wolff volumes increase as stands become older,
1998). A riparian stand's orientation to prevailing because the recruited trees are larger (Fox 2001).
winds and soil wetness can exacerbate wind Fox (2003) found a relationship between
throw (Bisson et al. 1987). Wind throw and instream wood loads and riparian stand age as a
subsequent recruitment to the channel is thus good indicator of succession. In that study, the
chronic or episodic, and it can be influenced by distribution of number of large wood pieces by
both natural and anthropogenic conditions. age class suggested that stem-exclusion processes
provide large initial inputs of wood numbers over
Fires the first 150 years, but they are low in volume
Although fires do not directly deliver wood to (Figure 1-6A) likely due to their small diameter.
streams, they are responsible for increasing tree Pollock and Beechie (2014) also found large
fall and slope erosion, which do deliver wood. numbers of recruited trees to the channel as
Disturbance that kills some or all the vegetation stands naturally thin through stem seclusion.
in a particular location is an intrinsic part of Wood recruitment (both piece number and
ecosystem development (Raup 1957; Oliver volume) is relatively low as stands mature over
1981). Effects will vary with climate, the next 400 years, after much of the stem-
geomorphology, topography, soils, and vegetation exclusion process has occurred but before age-
(Swanson et al. 1988). The return intervals for related mortality takes place.
fires, which vary by ecoregion (Agee 1993), affect
timber age (Henderson et al. 1992). In turn,
timber age influences mean tree diameter, which
influences the diameter of instream wood (Rot et
al. 2000). Timber age also influences tree height

Large Wood National Manual July 2015


1-20
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Figure 1-6. The Median Instream Large Wood the late-successional stands and the mortality
Volume (A) and Number of Pieces (B) According to rate decreases, resulting in a seemingly
Adjacent Riparian Stand Age Class, at the Time of paradoxical decrease of instream large wood
1999–2000 Surveys
abundance. This concept is supported by the
findings of McDade et al. (1990), who report that
approximately half of the large wood found in the
channel adjacent to second-growth forests came
from the previous old forest rather than from
newly regenerated stands.

Floods
Floods can entrain wood from areas adjacent to
stream reaches due to floodplain inundation and
an increase in fluvial forces. High flows associated
with floods increase the shear stress on and
buoyancy of instream wood and carry wood
downstream or perhaps completely out of a
system. Rootwads inhibit large wood movement
because they increase displacement and draft
(such as keel on a sailboat), and scour around the
rootwad allows logs to become embedded in the
channel (Abbe and Montgomery 1996; Abbe et al.
2003a; Abbe and Brooks 2011). Floods not only
remove wood from streams but can also recruit
new trees. Palik et al. (1998) found an average of
22 new trees per kilometer recruited into a
coastal plain stream during a large flood.

Despite the potential mobilization of wood due to


floods, Fox (2001) found that floods had little
influence on the overall instream wood loads of
natural systems. He explained this by two
Source: from Fox (2003), with age data courtesy of
observations: (1) much of the wood in these
Ian Henderson, unpublished data,
systems has previously resisted mobility during
This data also suggests that as the forests reach large floods, as broadly interpreted by the overall
late-succession at approximately 550 years, the age of pieces (as estimated by decay
mortality of the remaining older seral species classifications) found in the channel during the
becomes most prominent, combined with some surveys—even small pieces of wood in some
mortality of late-successional dominants streams had advanced decay, which suggests
associated with aging stands. This likely explains these pieces have also prevailed within the
the increases of instream large wood pieces at system despite floods; and (2) floods may replace
this age class (Figure 1-6B), as well as the fact that wood flushed from a system with newly recruited
these large trees are likely to be more stable and trees from bank avulsion, debris flows, or
resist entrainment, and so more readily upstream sources. Therefore, net loss of wood
accumulate in the channel. from floods may not occur in unmanaged basins.
However, this phenomenon may not extend to
At 800 years, younger trees are released by
heavily managed watersheds where natural
canopy openings during vertical stratification of

Large Wood National Manual July 2015


1-21
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

hydrological regimes have been altered to The loss of riparian vegetation is likely to
increase the frequency of flooding; banks have influence instream wood quantities due to the
been hardened to minimize lateral migration; and disturbance of the recruitment source. Most
instream wood is smaller, less stable, and more likely, snow avalanches occur at frequent
susceptible to mobilization. intervals in certain channels, maintaining a level
of disturbance to the channel and riparian area.
Landslides and Debris Flows This can preclude new wood recruitment to the
Landslides and debris flows are most common in adjacent and downstream channel. Fox (2001)
mountain regions. They are a natural process that reports that low-gradient channels (<6%)
occurs in every region of the United States. These impacted by snow avalanches had nearly the
hillslope or mass-wasting processes can be same number of wood pieces per 100 meters
triggered by human disturbance such as (328 feet) as channels with no recent disturbance;
deforestation, unstable earthen fill, and poor however, both the median numbers and volumes
drainage from roads and developments. of wood per 100 meters (328 feet) were lower in
Landslides and debris flows affect stream steep channels (>10% gradient) with snow
channels and influence the quantity, quality, and avalanches compared to channels without
distribution of instream wood. The often-violent disturbance. This could likely be attributed to the
mobilization of material in channels where this lack of riparian trees available for recruitment.
occurs may either transport wood out of a reach
or bring in new wood from upstream sources.
Human Influence
Debris flows tend to deposit wood on slopes of 3– The difference in the distribution and
6 degrees (approximately 5–10% gradient) (Ikeya characteristics of wood between managed and
1981; Benda and Cundy 1990; Fox 2001) and unmanaged basins has been clearly established.
remove it from streams with gradients >10% (Fox Wood can be limited due to riparian vegetation
2001). In older forests, large standing trees and modifications (Ralph et al. 1991), whether due to
instream logs can retard debris flow propagation forest practices, urban development, or
and run-out lengths compared to debris flows in agricultural practices. Unmanaged channels, often
industrialized forests (Coho and Burges 1993). defined by streams draining un-roaded and
Small, high-elevation regions of the country unlogged basins, typically have more channel
experience snow avalanches that recruit wood roughness due to instream wood than managed
into streams (Keller and Swanson 1979) and channels (Bilby and Ward 1991; Ralph et al.
influence the riparian vegetation (Fetherston et 1991), especially if the stream has been
al. 1995). Snow avalanche paths are typically less channelized. These factors, especially if peak
confined than debris flows, and they often form a flows are exacerbated due to land uses, may lead
broad fan where the channel gradient flattens, to less retention of recruited wood than in
such as at the channel bottom intersecting with streams draining unmanaged basins.
the floodplain of a larger system. Snow
avalanches are most common in small headwater Other forms of human influence on wood loading
channels within the snow zone (Keller and besides forest clearing can also result in
Swanson 1979). Due to the snow pack buffering of disruptions of the process by which wood reaches
the channel bed, substrates are often undisturbed streams and is distributed.
following a snow avalanche; however, most trees
larger than 10–15 centimeters (4–6 inches) in the
Hydromodifications
path are sheared off at the level of snow depth Disruptions to flow and subsequent transport of
(Fox 2001). wood by hydromodifications can alter wood
loads. Features such as dams, levees, road
revetments, culverts, and similar facilities can

Large Wood National Manual July 2015


1-22
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

intercept or impede the recruitment and model calibrated in southeastern Alaska


entrainment of wood. Alterations to flow regimes predicted a 70% reduction in wood 90 years after
can hinder or accelerate fluvial distribution of clear-cutting, and that full recovery exceeded
wood. 250 years (Murphy and Koski 1989).

Road Networks Logs derived from second-growth forests have


smaller diameter and lower volume than old-
Roads also influence instream wood loading and growth large wood, contributing to lower
transport. Maintenance activities often remove instream loading in logged streams (Bilby and
wood and clear jams to keep culverts and bridges Ward 1991; Ralph et al. 1991). Second-growth
free of debris and reduce structural damage wood loads tend to be composed of deciduous
during storms (Singer and Swanson 1983). Wood riparian species and small conifers that degrade
mobilized during high flows frequently becomes more easily and have less of an effect on long-
trapped on channel-spanning bridges and term channel morphology (Dominguez and
culverts, leading to road overtopping and Cederholm 2000).
eventual structure failure. Managers clear jams to
keep structures free of obstructions and reduce A comparison between unlogged, moderately
damage to river crossings. Such management may logged, and intensively logged catchments found
completely eliminate large wood or remove only that undisturbed streams contained more logs in
the largest pieces that pose the greatest hazard, the largest size categories (>50 centimeter
but which are the most important to habitat [20 inches] in diameter) than managed streams,
formation. Additionally, roads themselves can which reduce the amount of pool habitat (Ralph
encroach upon riparian forests, permanently et al. 1991). Bilby and Ward (1991) found that
removing sources of wood recruitment as long as wood-formed pools were less abundant in
the road is maintained. second-growth forests compared to old-growth
forests. This is supported by Chesney (2000), who
Forest Practices found that wood within the low-flow channel was
greater in unmanaged forests compared to
Timber harvest activities in streamside forests
streams within second-growth forests.
can directly affect wood input (Lienkaemper and
Swanson 1978; Bilby and Bisson 1998). Clear-cut Logging can have indirect effects on wood loads.
logging, often with inadequate buffer strips Harvest activities can destabilize hillslopes and
surrounding the stream channel, was a common increase the likelihood of debris avalanches
management practice throughout the Pacific (Lienkaemper and Swanson 1978). The use of
Northwest and Alaska until the late 1980s buffer strips is a common technique for reducing
(Dominguez and Cederholm 2000). The logging effects on forests and streams; however,
harvesting of streamside forests may temporarily buffer strips adjacent to clear-cuts are exposed to
reduce or eliminate large wood recruitment to the higher wind velocities, increasing the occurrence
stream (Bryant 1980), and the recovery time for of wind thrown logs to the stream channel (Reid
input to return to pre-harvest conditions will take and Hilton 1998). Higher rates of wind throw may
many decades. For example, Andrus et al. (1988) lead to rapid depletion of available wood from the
reported that in a stream flowing through a remaining adjacent forest, increasing short-term
second-growth forest 50 years after logging, 86% large wood input, but decreasing long-term input.
of the instream wood was remnants of the pre-
existing forest. The old wood accounted for 93% Urbanization
of the pools. The results indicate that some
The paucity of large wood in streams within
second growth stands will take much longer than
developed and developing parts of the world is
50 years for the new forest trees to make a
common, even where vast forests once covered
significant contribution of large wood. A decay

Large Wood National Manual July 2015


1-23
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

the landscape (e.g., Wiltshire and Moore 1983; have crews remove wood from streams (also
Petts et al. 1989). Navigation and conveyance known as “stream cleaning”), particularly
interests motivated widespread removal of between the 1950s and 1970s (Bisson et al.
instream wood obstructions and riparian trees. 1987). Wood was eradicated so successfully from
Data from Horner et al. (1997) show a clear many streams (Reeves et al. 1991) that there are
general trend that the more urbanization, the less still consequences to fish habitat (Bisson et al.
large wood in the encompassed stream. 1987). The removal of wood from rivers was a
major endeavor to promote navigation and log
transport to mills, particularly in Pacific
GUIDANCE Northwest streams. The removal of hundreds to
thousands of snags per year by the USACE
Common Means by Which Large Wood Is Lost continued in the region’s rivers through at least
From the Urban Channel 1960 (Collins et al. 2003).
 Peak Flows. As the magnitude of channel flow
increases due to proliferations in impervious 1.3.5 Wood Management
surface, the peak discharges of annual and multi-
year floods increase typically two- to five-fold Not all forms of human influence have led to the
(Hollis 1975) and the duration of flood flows may depletion of wood from our streams. Because of
increase more than ten-fold (Barker et al. 1991). the correlations wood has to channel morphology,
The consequences are high rates of wood aquatic habitat, and salmonid production, and due
depletion through entrainment. The scour of bank to the paucity of instream wood stemming from
vegetation that may normally assist in the
past land-use practices, wood placement projects
stabilization of wood further compounds wood
depletion. have become a common method for restoring or
enhancing salmonid habitat (Kauffman et al.
 Channel Incision. With increased flows and 1997). Resource managers have been successful
sediment transport comes channel down-cutting. at inducing salmonid response by placing wood in
The immediate consequences of such a process is
streams (House and Boehne 1986; Cederholm et
a deep and narrow channel that vertically strands
wood that once was in contact with the bed, and al. 1988; Nickelson et al. 1992; Murphy 1995;
further increases erosion due to less resistance Riley and Fausch 1995; Solazzi et al. 2000; Roni
(Booth 1990). and Quinn 2001). As a long-term approach, many
researchers have advocated the maintenance of
 Human Removal. Wood has been removed for wood loads by restoring natural riparian
various reasons from the urban landscape. Stream
processes (Sedell and Luchessa 1981; Elmore and
beautification and tidiness, the perception of better
fish passage, better safeguards against avulsion and Beschta 1988; Cederholm et al. 1997b; Roni and
lateral migration, and improved water craft Quinn 2001).
navigation, for example, compelled humans to
remove wood from streams. 1.3.5.1 Forest Characteristics
Various forest characteristics, perhaps
Although not necessarily an artifact of independent of large-scale climatic or
urbanization, the presence of humans has disturbance-related factors, will influence the
implications for instream wood loads. Across number, volume, and size of instream wood. Rot
much of North America, particularly the Pacific et al. (2000) found the diameter of instream large
Northwest, wood has been extirpated from our wood increased with riparian stand age, and that
streams, and the riparian sources have been stand age and mean stem diameter were
compromised in their ability to recruit wood. A correlated. Tree age varies considerably within
common practice to improve fish passage and older forests. For example, Tappeiner et al. (1997)
flow conveyance in the Pacific Northwest was to found age in old-growth stands ranged between

Large Wood National Manual July 2015


1-24
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

50 and 414 years at one site. They saw median 1.3.5.2 Instream Wood Quantities
age differences of 187 years across ten sites in the
same region. Timber on the Olympic Peninsula, The composition and character of riparian
Washington, often older than 700 years vegetation can dictate the species composition,
(Henderson unpublished data), can produce very numbers, size, and volume of large wood
large-diameter instream wood. Within streams recruited to the channel, and lateral and vertical
draining old-growth forests, McHenry et al. distribution of that large wood within the channel
(1998) found a mean large wood diameter of (Grette 1985; Bisson et al. 1987; Bilby and
0.3 meter (1 foot) and diameters up to at least Wasserman 1989; Bilby and Ward 1991; Ralph et
2.5 meters (8 feet). However, because most wood al. 1991; Bryant and Sedell 1995; Bilby and Bisson
pieces could not be attributed to an adjacent 1998; Fox and Bolton 2007). Factors that
source (McDade et al. 1990), upstream riparian influence the spatial distribution of instream
areas and basin processes may provide a better wood include both the regional context and the
predictor of instream wood quantities than local geomorphic setting.
adjacent riparian areas. Regional factors influence the quantities of wood
In much of the forestry literature, riparian forests in a system but do not appear to vary their spatial
are characterized with general forest attributes organization. Fox (2003) found that forest regions
only. However, significant distinctions are likely did not have a pronounced effect on the grouping
to exist between upland and riparian stands. or clustering of large wood pieces, which were
Naiman et al. (1998) reported that the basal area proportionally the same in streams of similar
of riparian forests is generally as great as or widths regardless of forest type.
greater than that of upland forests; riparian Fox and Bolton (2007) counted pieces of wood in
forests have relatively high rates of biomass 150 sites totaling nearly 38 kilometers (24 miles)
production in comparison with upland forests, of streams draining unmanaged Pacific Northwest
likely influenced by moisture, nutrients, and forests. Sampled stream gradients ranged
temperature gradients. They also often promote between 0.04 and 49% and represented a diverse
deciduous seral species regeneration in response array of channel types, confinement classes,
to channel-associated disturbances (Naiman et al. bedforms, dominant water origins, disturbance
1998). Collins et al. (2003) tallied the occurrence histories (fire, debris flows, snow avalanches, and
of tree species along the major rivers of western floods), basin sizes, elevations, and forest types
Washington as reported in surveyors’ notes from common in the Pacific Northwest. These authors
the mid- to late nineteenth century; they found an quantified wood loads within forest types and
average of 84% hardwood species by stem count channel sizes based on statistically discrete
and about 55% by biomass, particularly from the groupings, where they found similarities between
presence of red alder (Alnus rubra). This the SS/WH and SF/MH ecoregions, and between
contrasted to the dominance of Douglas-fir and the SAF and GF ecoregions. These large wood
western hemlock on adjacent upland terraces, quantities are provided in Table 1-1, using data
together with a significant component of riparian only from fully unmanaged watersheds. The
western red cedar. Finally, Gregory et al. (1991) watersheds in this data set are characterized by
and Pollock et al. (1998) found that microclimate forests that are all loosely termed as “old-growth”
gradients also contribute to greater plant and and also meet the following criteria: (1) no part of
animal species diversity in riparian forests than in the basin upstream of the survey site was ever
upland forests. Riparian forest structure and logged according to forest practices commonly
characteristics are therefore apparently different employed since European settlement; and (2) the
from, and generally more productive than, typical basin upstream of the survey site contains no
upland forests. roads or human-made modifications to the
landscape that potentially could affect the

Large Wood National Manual July 2015


1-25
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

hydrology, slope stability, or other factors channel substrate, and flow regimes may account
affecting the natural processes of wood for major differences independent of channel size.
recruitment and transport in streams. Some of The size of wood and whether or not it includes a
these basins may be managed to remain pristine, rootwad (e.g., Abbe and Montgomery 1996; Abbe
however, which may also include fire et al. 2003b; Abbe and Brooks 2011; Moulin et al.
suppression. It is assumed that these forest 2011) directly influence piece stability. Abbe and
conditions incorporate the range of variability Montgomery (1996) discuss the importance of
and disturbance frequencies to which many large trees in the formation of key pieces or snags
aquatic species have adapted. that initiate logjams in large rivers. Fox and
Bolton (2007) identified individual logs (i.e., key
Estimations of wood loading around the world
pieces) that exhibited indicators of long-term
vary from less than 1 to 2,000 megagrams per
stability (persisting through at least moderate
hectare of channel with no strong correlation to
floods) and related them to channel size (Table
region or channel size (Moulin et al. 2011) (Figure
1-2).
1-7). Cordova et al. (2007) report that average
wood loading in pieces per kilometer range from Fox and Bolton (2007) suggest that minimum
a high of 362 in the Northwest to 326 in the piece volumes used to define a key piece should
Midwest, to 161 in the Northeast to 61 in the consider the role rootwads play in achieving
Southeast. Recent surveys by Krause and Roghair stability. In channels greater than 30 meters
(2014) found the average piece count in six North (98 feet) BFW, more than 91% of all key pieces
Carolina streams measured in 2007/2008 and had rootwads attached. Therefore, in order to
2012/2013 ranged from 206 to 170 pieces per meet the objective of defining a key piece, not
kilometer, respectively. Wood loading per unit only do the prescribed minimum volumes need to
channel area tends to decrease with increasing be met, but also rootwads must be considered in
channel size, but differences in forest trees, the definition.

Large Wood National Manual July 2015


1-26
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Figure 1-7. Wood Loading in Streams Throughout the United States and Other Regions Typically Range
from 1 to 2,000 Megagrams per Hectare

Although there is a generally reduction in wood loading with increasing channel size, there is significant
variance due to tree size and the size, slope, and substrate of channels. (Data compiled from Keller and
Swanson 1979; Bryant 1983; Wallace and Benke 1984; Hauer 1989; Shields and Smith 1992; Keller and
MacDonald 1995; Lisle 1995; Richmond and Faush 1995; Gippel et al. 1996; Piégay and Marston 1998; Piégay
et al. 1999; Cordova et al. 2007; Baillie et al. 2008; Magilligan et al. 2008; Moulin et al. 2011.)

Large Wood National Manual July 2015


1-27
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Table 1-1. Distributions of Wood1

BFW Class
Region (meters) 75th Percentile Median 25th Percentile
Number of Pieces per 100 meters of channel length
Western Washington 0–6 >38 29 <26
>6–30 >63 52 <29
>30–100 >208 106 <57
Alpine >0–3 >28 22 <15
>3–30 >56 35 <25
>30–50 >63 34 <22
DF/PP Forest Zone 0–6 >29 15 <5
>6–30 >35 17 <5
Volume (cubic meters per 100 meters of channel length)
Western Washington 0–30 >99 51 <28
>30–100 >317 93 <44
Alpine >0–3 >10 8 <3
>3–50 >30 18 <11
DF/PP Forest Zone 0–30 >15 7 <2
Number of Key Pieces per 100 meters of channel length
Western Washington 0–10 >11 6 <4
>10–100 >4 1.3 <1
Alpine >0–15 >4 2 <0.5
>15–50 >1 0.3 <0.5
DF/PP Forest Zone 0–30 >2 0.4 <0.5
Source: Fox and Bolton (2007, Table 4).
1 Number, volume (cubic meters), and number of key pieces, all per 100 meters of channel by Forest Regions

in Washington State and bankfull width class. Wood includes pieces exceeding 10 centimeters (4 inches) in
diameter and 2 meters (6 feet) in length.

Large Wood National Manual July 2015


1-28
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Table 1-2. Minimum Wood Piece Volume Required to Qualify as a Key Piece (by Bankfull Width Class)

BFW Class Minimum Volume


(meters) (cubic meters)
0–5 1.001
5–10 2.501
10–15 6.001
15–20 9.001
20–30 9.75
30–50 10.502
50–100 10.752
1 Current WFPB (1997) definition.
2 Piece must have an attached rootwad.

1.3.5.3 Stability Factors relate to instream wood, this range can be used to
set management targets in the Pacific Northwest
Factors inducing wood stability are seemingly for riparian recruitment objectives, regulation,
dependent on the interaction of pieces within habitat restoration, enhancement, and evaluation.
groups and how groups are assembled during For restoration and enhancement of instream
fluvial processes. Fox (2003) and Parrish and wood loads, streams should be managed to meet
Jenkins (2012) found that the stability of wood this natural distribution at a basin scale, where
increases with jam size due to a larger matrix of restoring the natural heterogeneity of wood loads
pinned logs. However, Fox suggests that the is the primary objective. Streams in a degraded
percentage of stable logs decreases as channel state (e.g., below the median) should be managed
size increases because much wood is loosely for wood inputs exceeding the median of this
assembled as pieces become stranded on gravel range. The top of these distributions, the 75th
bars as flows recede. Conversely, gravel bars and percentile and above, should be used as an
highly sinuous channels were less commonly interim management ‘‘target’’ until the basin-
observed in small streams; thus, accumulations of scale wood loads achieve the central tendencies of
wood along the banks and channel margins may natural and unmanaged wood-loading ranges.
require greater proactive fluvial force to impinge
wood because there are fewer collection points The precise quantities and volumes of wood
for wood during flow recession. needed by salmonids for successful production
are not well understood. Statistically sound
1.3.6 Wood Performance studies to link instream wood loads to salmonid
production would be expensive and have high
Standards levels of uncertainty owing to the multiple
The percentile distributions for large wood variables influencing salmon production (Roni et
quantity, volume, and key-piece quantity (Table al. 2003). However, historic salmonid populations
1-1) represent the range of conditions found in were much higher than those found today, and, as
streams draining unmanaged forests that are noted earlier, unmanaged forests offer the best
subject to a natural rate of disturbance (except source of information on wood loads as one
fire suppression) in the Pacific Northwest. component of habitat to which salmonids have
Assuming these data include both favorable and adapted. In degraded streams, where
unfavorable salmonid habitat conditions as they management is needed to restore favorable

Large Wood National Manual July 2015


1-29
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

conditions, wood loads are often no longer found the percentage distribution of large wood shifts to
in the upper distribution of these ranges, or the larger group size classes, as depicted by the shift
distribution is centered around a lower mean. In in the median in Figure 1-8. These data together
these cases, merely managing for the mean or with observations of Abbe and Montgomery
median will not restore the natural ranges of (2003) support the theories and hypotheses of
heterogeneity. Therefore, for management Keller and Swanson (1979) and Swanson et al.
purposes intending to restore natural wood- (1982) that wood becomes more clumped (i.e.,
loading conditions, establishing instream wood organized into larger jams) with increasing
targets based on the upper portion of the channel size.
distribution observed in natural systems (i.e., the
The lateral channel position of wood is also an
75th percentile) rather than the lower portion of
important design consideration. Wood placed too
the distribution is reasonable as well as prudent
high on the bank may serve to resist channel
to restore natural ranges.
migration, but fail to provide habitat at lower
The reported wood loading ranges of Fox and flows. Wood organization in unmanaged systems
Bolton (2007) are not likely representative of all may also provide a reference for conditions to
streams across North America due to differences which salmonids have adapted. Fox (2003) also
in wood source characteristics and loading looked at lateral distribution of wood as broken
mechanisms. For example, wood loads in the into four zones: Zone 1 is the wetted low-flow
sparsely forested regions of the western desert channel, Zone 2 is above the wetted low-flow
are likely to be much lower than those of the channel but below the horizontal axis of the
densely forested Pacific Northwest; therefore, it bankfull channel, Zone 3 is above the high-flow
may be unrealistic to apply these wood loads as a channel but within the vertical confines of
performance standard everywhere. In this regard, bankfull, and Zone 4 is laterally beyond the
performance standards could be formulated in a bankfull width. Wood in these four zones
similar manner using reference site surveys, river provides different purposes, from summer
snagging records, old forest characteristics, and rearing habitat in Zone 1 to stability functions
other information as available. However, it can be when wood extends far into Zone 4. Distributions
acknowledged that restoration endeavors that of wood from small groups of wood (less than
aim to create favorable habitat conditions in a 10 pieces per group) are presented in Figure 1-9A
degraded system may benefit from using overly and for large groups (10 or more pieces per
conservative wood loading conditions. Assuming group) in Figure 1-9B.
the wood conditions in target restoration reaches
Restoration projects involving wood as a
are far below the median range (and hence need
restoration tool often utilize ELJs, where all
enhancement), a reach or more with higher-than-
aspects of the design are carefully planned using
expected wood loads may help restore
the principles of physics, hydraulics, biology,
heterogeneity and provide ecological elements
safety factors, and other considerations to ensure
that are in short supply.
the project meets the intended project objectives.
The questions “how many jams,” “how much
1.3.7 Wood Distribution wood should be placed within them,” and other
within Channel Networks specifics are valid points to consider. Some
questions are best answered on a hydraulic and
Fox (2003) found that channel size (as geomorphic basis; however, replicating the
represented by bankfull width) is a significant natural range and heterogeneity of conditions to
geomorphic influence on group size distribution. which salmonids are accustomed will provide
Figure 1-8 illustrates that the percentage of wood greater certainty in ecological success. Therefore,
allocated to larger group sizes increases with it may be more prudent to couple the wood loads
channel size. With each greater channel size class,

Large Wood National Manual July 2015


1-30
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

from Table 1-1 and distribute them in Figure 1-9. Comparison of the Mean Percent Large
proportions reported in Figure 1-8. For example, Wood Volume by Four Lateral Zone Distributions1
place the targeted number and volume of wood
into jam grouping percentages provided in the
appropriate channel size of that figure, combined
with the lateral distributions provided in Figures
1-9A and B.

Figure 1-8. The Percent Distribution of Large Wood


to Group Size Class According to Five Bankfull
Width Classes

1 Between (A) small groups (<10 pieces per group)


and (B) large groups (≥10 pieces per group)
according to five BFW classes. Zone 1 is the wetted
low-flow channel, and Zone 2 is above the wetted
low-flow channel but below the horizontal axis of
the bankfull channel. Zone 3 is above the high-flow
channel but within the vertical confines of bankfull,
and Zone 4 is laterally beyond the BFW. The
numbers in parentheses are the standard deviations,
and n= the number of large wood groups. From Fox
(2003).

When restoration projects involving the artificial


placement of instream wood are warranted, the
reference conditions of instream wood in natural
systems can offer guidance for restoring the
The vertical bars represent the median values. From heterogeneity and structure of wood in adversely
Fox (2003). impacted systems.

The following steps provide an example of how to


use such comparisons to proceed with a
restoration endeavor, based on the findings of Fox
and Bolton (2007) and Fox (2003):

Large Wood National Manual July 2015


1-31
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

1. Through monitoring and assessment, persisted in the system for decades. Replicating
determine the current status of instream natural jams may serve as a better template for
wood in a potential restoration project reach. restoration than jams designed to merely remain
stable and un-deformable through large floods at
2. Based on natural distributions of large wood
the expense of ecological functions.
piece numbers and volumes, assess if wood
additions are warranted, and how much more
is needed to attain natural loads. Tables 1-1 1.3.8 Wood Longevity
and 1-2 provide a summary of natural large
The longevity of wood is another concern.
wood distributions based on Fox and Bolton
Observations on the longevity of natural instream
(2007).
wood are briefly discussed here, and the longevity
3. Organize spatial wood distributions according of wood placements is discussed in more detail in
to those found in natural systems. Figure 1-8 Chapter 6, Engineering Considerations. Wood
provides the natural distribution of wood to deposited in saturated or anaerobic conditions
various group sizes based on Fox (2003), within a stream bed will essentially last forever
enabling a comparison to the existing (e.g., Gastaldo and Demko 2011). In settings
organization of wood in the stream targeted where it is subject to wetting and drying, wood is
for restoration. The filling of voids in this subject to rapid decay. The rates of wood
distribution within the project area can then decomposition vary by species, submergence,
be facilitated in order to mimic a more natural burial, and climatic conditions (e.g., Graham and
spatial distribution. Cromack 1982; Melillo et al. 1983; Means et al.
1986; Sollins et al. 1987; Spänhoff et al. 2001;
4. Organize lateral wood distributions according
Scherer 2004; Beets et al. 2008; Guyette et al.
to those found in natural systems. Figures
2008) and is covered in Chapter 6. Wood, or
1-9A and 1-9B provide the natural
evidence of wood, can be found in fluvial
distribution of wood according to lateral
sediments deposited since trees appeared about
channel zones based on Fox (2003), enabling
360 million years ago in the Devonian Period.
a comparison to the existing organization of
During this time they have not only left abundant
wood in the stream targeted for restoration.
evidence of their presence in the geologic record,
The filling of voids in this distribution within
but they have played an important role in the
the project area can then be facilitated in
evolution of landscapes and biota. The geologic
order to mimic a more natural distribution.
record shows that logjams began to have a
Other design objectives to consider are the notable influence on river channel morphology
replications of habitat features useful to within the Pennsylvanian subperiod of the
salmonids and in short supply within the reach of Carboniferous Period 323.2 million years ago
interest. Restoring specific habitats while (e.g., Gastaldo and Degges 2007; Gibling et al.
maintaining certain engineering standards may 2010).
be challenging but valuable objectives. For
Using known decay rates of the wood, estimates
example, Parrish and Jenkins (2012) found that
can be made on how long it will last or what its
many natural jams consisted of numerous racked
effective size will be after a given time (Figure
members that allowed flow to pass through the
1-10A) (Abbe 2000; Abbe et al. 2003b; Abbe and
interior, which provided excellent cover and pool
Brooks 2011). Examples of buried wood found
habitat for fish. Despite not having buried
exposed in eroding river banks have shown that
members or rock ballast (commonly used in
natural logjams can last hundreds to thousands of
ELJs), these jams were highly stable and had
years (Figure 1-10B).

Large Wood National Manual July 2015


1-32
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Figure 1-10. (A) Example of Decay Curves for Three Common Pacific Northwest Tree Species; (B) Example of
Ancient Logjam More than 120 Years Old Exposed in the Right Bank of South Fork Nooksack River,
Washington

Sources: (A) Abbe (2000) and Abbe and Brooks (2011); (B) Abbe and Brooks (2011).

Large Wood National Manual July 2015


1-33
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Abbe (2000) and Montgomery and Abbe (2006) formation has largely been overlooked. For most
dated Pacific Northwest logjam ages ranging from of the industrial revolution, right up to the late
several decades to over a thousand years. Guyette twentieth century, science ignored the possible
et al. (2008) radiocarbon dated 200 tree boles role of wood on fluvial systems. It has only been
exposed in eroding banks of eight streams in in the last several decades, long after the
north Missouri and found that oak trees have alteration of river valleys across the Northern
been accumulating in alluvial sediments since the Hemisphere, that scientific research began to
late Pleistocene, 14,000 years ago. The median recognize that wood could influence fluvial
age of oak boles was 3,515 years B.P.[1] Wood ecosystem processes. Research on wood has
samples from buried logjams exposed along the increased exponentially in the last decade when it
montane Queets River in the Olympic Peninsula in has become more evident that wood in streams
Washington were considerably younger with has influenced our landscape for millions of years.
radiocarbon ages of 0 to 1400 years B.P. (Abbe
In terms of evaluating fluvial aquatic ecosystem
2000; Hyatt and Naiman 2001; Montgomery and
conditions and developing restoration strategies,
Abbe 2006). Samples of buried logs exposed in
the introduction of large wood and natural wood
the banks of the Ducktrap River, a low-gradient
has impacts on the local geology that affect the
coastal stream in Maine had radiocarbon dates of
evaluation. Research and the observed results of
1180 and 1650 years B.P. (Magilligan et al. 2008).
wood reintroduction have clearly demonstrated
Brooks and Brierly (2002) dated wood in the
the beneficial role of large wood in creating and
Thurra River of southeast Australia as tens of
sustaining healthy river ecosystems. Large wood
thousands of years old. These observations and
influences channels of all sizes by introducing
the success of engineered wood placements used
physical complexity to the system. Wood
over the last several decades indicate wood
accumulates in any river or stream that has
placements may have a positive role in carbon
riparian forests, from New England (Figure 1-11)
sequestering.
to the arid west (Figure 1-12).

1.4 Ecological Functions of Figure 1-11. Naturally Occurring Snag Embedded in


Channel Thalweg, Androscoggin River near Bethel,
Wood Maine

Large wood can be found in nearly all streams


and rivers where trees populate riparian areas. It
has multiple functions in these systems, such as
trapping sediment, forming pools, providing
shade and cover for aquatic organisms, and
diversifying flows. It is a key ecosystem
component for stream organisms, particularly for
fish. Large wood has been part of virtually all
forested freshwater systems for many thousands
of years, and its role is significant for life stages of
many aquatic species.

The historic reduction in wood found in streams


has been grossly under-appreciated; likewise, the
magnitude of the effect wood has on habitat

[1]B.P. rabiocarbon dating reference “before present”


referring to time prior to 1950.

Large Wood National Manual July 2015


1-34
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Figure 1-12. Lush Riparian Areas Even Occur in Arid


Regions Where They Deliver Wood to Streams, GUIDANCE
North Central Oregon
The Function of Wood
Associated With Aquatic Fluvial Ecosystems
 Shade.
 Hydraulic influence raising local water elevations,
scouring pools, and creating low-velocity refugia.
 Channel grade control.
 Reduction in rate of water flow and increase in
residence time.
 Retention and storage of sediment and flotsam
(small wood and organic material).
 Retention of nutrients.
 Side channel formation.

The recognition of the importance of wood has  Increased floodplain connectivity.


led to its increased use in river restoration  Maintaining biological structure and ecosystem
projects and changes in river management, such productivity.
as leaving natural wood in place and protecting
 Maintaining channel and floodplain physical
riparian forests, which are necessary for wood complexity.
recruitment. Stable wood traps small wood, which
creates large amounts of complex environments  Providing complex cover for aquatic organisms.
and increased surface area for biologic activity  Increased hyporheic exchange.
that supports more complex food webs.
 Improved water quality.
The role of wood in aquatic fluvial ecosystems can  Increased recharge and aquifer storage
be categorized into two basic functions: biological  Creating habitat for fish and macroinvertebrates.
and physical. The biological functions include how
wood provides a unique growth medium and
nutrient source for invertebrates and also its use
in creating vegetation and other ecological niches
1.4.1 Biological Functions
for habitat. The physical functions include how Large wood is a key structural element in forested
wood influences channel structure and energy stream ecosystems worldwide (Maser and Sedell
dissipation, hydrology, sediment and organic 1994; Nagayama and Nakamura 2010). Wood
debris transport, substrate conditions, and serves as a food resource for microbes, fungi, and
channel and floodplain morphology. These macroinvertebrates. In addition, a primary
functions are closely linked and are influenced by ecological role of large wood and accumulations
other factors such as flood events, fires, and of large wood (wood jams) is associated with its
human development. influence on the physical environment of streams
and the creation of habitats for aquatic species
(Roni et al. 2014a).

In most cases, restoration of large wood is


undertaken to achieve some biological goal.
Hence, the inherent assumption of restoration of

Large Wood National Manual July 2015


1-35
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

large wood is that habitat features in streams biological functions associated with wood and
associated with word wood are positively related fluvial ecosystems.
to the survival, persistence, and abundance of
desired species and communities and ecological 1.4.1.1 Habitat Formation
functions (Whiteway et al. 2010). While
intuitively appealing, the relationship between Habitat consists of elements of the environment
individual habitat attributes and fish survival or that affect the persistence and performance of a
species in a specific location (Whittaker et al.
abundance can be difficult to prove in a
quantifiable and statistically meaningful way 1973; Hall et al. 1997). The quality and quantity of
(Conquest and Ralph 1998; Bradford et al. 2005). habitat across the life history of the species shape
biological performance in terms of abundance,
Consequently, some researchers have reasonably
questioned the benefits of stream restoration persistence, and fitness (Southwood 1977).
activities (Thompson 2006; Stewart et al. 2009), Habitats for species can overlap but are usually
separated temporally, spatially, or in terms of
or called for a better accounting of the costs and
function. For example, large wood can be an
benefits of restoration investments (Bernhardt et
al. 2007). Benefits are challenging to detect, in element of habitat for both juvenile salmonids
part, because of the number of confounding and benthic insect life stages, but the nature of
that habitat differs; wood generally provides
factors affecting fish abundance in any year or
over time, especially at a population scale for far- cover for juvenile salmonids while it provides a
ranging anadromous species such as salmon substrate on which benthic insects move and feed.
(Rose 2000).
1.4.1.2 Aquatic Food Webs
When scientists documented the decline in
salmonid populations in the Northwestern United A food chain is the linkage between primary
States and correlated that decline with stream resources (plants, detritus) and secondary
simplification following wood removal , efforts to consumers (e.g., insects and fish) (Pianka 1994).
replace large wood in streams received national A network of linked food chains forms a food web,
attention from researchers, resource managers, and stream food webs are among the most
and restoration practitioners. Efforts highlighted complex. Like most ecosystems, aquatic foodwebs
the importance of woody debris in forming begin with the capture of energy from the sun
salmonid habitat in fluvial ecosystems, and that is fixed by terrestrial and aquatic plants via
restoration efforts using wood became widely photosynthesis. This energy is stored in the tissue
accepted (Bisson et al. 1987; Kauffman et al. of the plant where it is available to secondary
1997). consumers.

1.4.1.3 Biogeochemical Functions


CROSS-REFERENCE
Large wood plays a key role in nutrient cycling
A detailed description of biological functions related to in streams (Bilby and Bisson 1998). In general,
wood and fluvial ecosystems is provided in Chapter 3, wood itself is a poor carbon source. The
Ecological and Biological Considerations. amount of nitrogen and phosphorous relative
to carbon is low, and the lignin in wood is
particularly difficult for many organisms to
Large wood structures not only provide cover for
break down (Webster and Benfield 1986). In
fish, mammals, and birds, they also increase
invertebrate and aquatic plant productivity that temperate ecosystems, few macroinvertebrates
enhances the ecosystem food web (Coe et al. or fish eat wood directly, but there is a suite of
2009). The following provides a discussion of the microbes and fungi that break down wood,
which, in turn, form food for benthic

Large Wood National Manual July 2015


1-36
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

invertebrates and other biota (Webster and upstream of it, resulting in an expanse of slower
Benfield 1986; Findlay et al. 2002; Spänhoff and and higher water extending upstream from the
Cleven 2010). The stream macroinvertebrates obstruction. The backwater effect can result in
that do eat wood tend to eat smaller particles, higher water surface elevations along the banks
and/or they ingest wood as a byproduct of and, in unconstrained reaches, enhanced
feeding on microbial biofilms on wood surfaces floodplain connectivity with an increased volume
(Johnson et al. 2003; Coe et al. 2009). The rate of water spilling out onto the floodplain. The
of wood decay by microbes and fungi varies by ability of large wood to alter water levels and
influence habitat varies based on local conditions,
species. As a rule, trees with more nitrogen per
including the volume of assembled wood and its
unit of carbon (such as alders maples, and
size relative to channel morphology.
poplars) decay faster that those with lower
nitrogen–to- carbon ratios (such as oaks, firs,
1.4.1.5 Hyporheic Zone
and spruce) (Spänhoff and Meyer 2004).
The hyporheic zone is the water-saturated
Large wood can enhance stream nutrient cycling
sediment volume below the stream bed and
in multiple ways. First, large wood retains leaf
adjacent stream banks where mixing between
litter and fine particulate organic matter. The
surface water and groundwater occurs (Bencala
breakdown of this organic matter by microbes
2005). It may extend 30 meters (98 feet) or more
and fungi creates an elevated demand for
into the adjacent floodplain (Hinkle et al. 2001;
nutrients, especially nitrogen and phosphorous.
Boulton et al. 2010). Definitions may vary with
This elevated demand increases the rate at which
the scale and intent of a given study and include
nutrients are taken up from the water column and
hydrological, hydrogeological, biological, and
increases the retention of nutrients in the stream
physiochemical criteria (Environmental Agency
(Mulholland et al. 2009). Second, when channel-
2009; Boulton et al. 2010).
spanning wood and wood jams retain a
combination of organic material and fine Although the hyporheic zone may only extend as
inorganic material they can create areas of little as 5 centimeters (2 inches) into the
saturated sediment behind and around the wood streambed it is extensive because it extends from
where oxygen can be locally depleted. Under the uppermost headwaters through the
these anaerobic conditions available nitrogen can lowermost reaches of rivers and into the
be converted to nitrogen gas through a process estuarine zone (Krause et al. 2014). The
referred to as denitrification (Steinhart et al. cumulative effect of large-scale wood placement
2000). This conversion is highly variable across can improve water quality by trapping sediment
streams and across regions but it can be an and increasing hyporheic flow (e.g., Lautz et al.
important loss of nitrogen from these systems, 2006; Mutz et al. 2007; Wondzell et al. 2009).
especially in areas of the northeastern and Increasing hyporheic exchange moderates water
midwestern United States where excess nitrogen temperatures (Hester and Gooseff 2010) and
pollution is a particular concern. improves water quality by increasing uptake of
phosphate (Warren et al. 2007) and buffering
1.4.1.4 Wetted Area of the Channel pollutants (Hester and Gooseff 2010).

Large wood creates bedform roughness The hyporheic zone is both a physical space and a
(resistance to flow, or drag) that effectively slows biological habitat for microbes, invertebrates,
flow down, consequently raising the water insect eggs and pupae, fish eggs, and fish embryos
surface level. This may facilitate a hydraulic (the hyporheos). In the hyporheic zone surface
“backwater effect,” whereby the water level water and solutes exchange into and out of the
immediately upstream of the obstruction is stream bed having mixed with groundwater to
raised, which in turn raises the level of water varying extents. Numerous biogeochemical

Large Wood National Manual July 2015


1-37
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

reactions occur in this zone, and it can influence connectivity and propagate upstream to degrade
mineralization, major ions, and nutrient and an entire drainage network.
contaminant components in the stream system
This following provides a discussion of the
(Bencala 2005; Gandy and Jarvis 2006;
physical functions associated with wood in fluvial
Mulholland and Webster 2010; Krause et al.
ecosystems.
2014).

Hyporheic flow also has localized influences on


stream temperature and dissolved oxygen. All of CROSS-REFERENCE
these aspects argue for hyporheic zone
A detailed description of physical functions related to
consideration in restoration and large wood
wood in fluvial ecosystems is provided in Chapter 4,
placement projects (Hester and Gooseff 2010). Geomorphology and Hydrology Considerations, and
Krause et al. (2014) point out that much of the Chapter 5, Watershed-Scale and Long-Term
research on large wood and its hydrological, Considerations.
ecological, and biogeochemical roles has focused
on headwater and upland streams (e.g., Tonina
and Buffington 2009; Buffington and Tonina 1.4.2.1 Grade Control
2009). Their review considers these influences
For millions of years wood in streams has been
and previous studies from the perspective of
responsible for controlling much of the grade in
lowland rivers. Wondzell (2011) evaluates data
alluvial systems of all sizes, whether as individual
from a fifth-order mountainous stream and shows
tree trunks spanning a channel (Figure 1-13), as
that the size of hyporheic exchange flows relative
logjams, or in beaver dams. A majority of channels
to stream discharge was large only in very small
in a small drainage network could easily be
streams and at low discharge. In the larger
impounded by a single tree, particularly in the
streams and at higher flows this ratio was small.
old-growth forests that dominated landscapes
across North America. In larger channels, logjams
1.4.2 Physical Functions were a common obstruction that controlled a
river’s morphology (Wolff 1916; Abbe and
Forest cover, or canopy, within a watershed
Montgomery 1996; Abbe and Brooks 2011) and
directly affects the hydrology and supply of
even impounded large rivers (Guardia 1933).
sediment to a river. Riparian forests along a river
aid in stabilizing the bank, which influences bank
Figure 1-13. Large Trees Can Play a Major Role in
resistance to erosion, which in turn affects the the Morphology of Rivers, Such as this 2.4-Meter
hydraulic geometry of a river. Increases in bank Douglas Fir Across Carbon River, Washington
strength result in narrower, deeper channels.
Conversely, decreases in bank strength result in
wider, shallower channels (Eaton and Lawrence
2006). Stable instream wood provides hydraulic
and morphologic complexity to a channel. Wood
defines water surface profiles, flow energy
expenditure, sediment transport, channel
morphology, and aquatic habitat. The removal of
large wood increases the river’s energy to move
sediment, erodes its bed and banks, simplifies
channel morphology, and severely degrades fish
Caddo Lake in Texas, one of the largest natural
habitat. Impacts such as channel incision are not
lakes in the southern United States, was formed
limited to the channel but affect floodplain
by a logjam in the Red River (Veatch 1906).

Large Wood National Manual July 2015


1-38
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Natural log “steps” are a familiar feature in small 1.4.2.2 Riparian Forests
streams throughout North America where log
length exceeds the channel width (e.g., Marston The principal physical functions of a riparian
1982). The presence of larger trees can extend the forest are mediation of microclimate and shade,
influence of wood into large channels, whether as generating effects on channel form by root
single pieces (Figure 1-13) or logjams (Figure reinforcement and recruitment of large wood, and
1-14). Within steep channels large boulders and resulting mediation of channel disturbance
logs both create stable obstructions. In most regime.
lower gradient alluvial rivers, large snags were The riparian forest affects stream microclimate by
naturally the principal flow obstructions. Where attenuating wind, shading the stream surface, and
snags and large riparian trees have been in many cases buffering the stream from
removed, human structures such as bridge piers microclimatic conditions in nonforest areas (such
may be the only obstructions. With conversion of as logged or developed lands) located farther
riparian areas to younger forests and fewer from the stream. Chen et al. (1995), studying
natural obstructions to trap and moderate the microclimate in a forest adjacent to the edge of a
movement of mobile wood, the accumulation of recent clearcut, found that the forest attenuated
wood at human structures becomes a greater risk. variation in soil and air temperature, soil
The removal of large trees that once lined rivers moisture, relative humidity, solar radiation, and
throughout the country has contributed to the wind speed, relative to the adjacent clearcuts.
much lower volumes of wood currently found in Brosofske et al. (1997) and Anderson et al. (2007)
rivers. Most of the rivers in the Mississippi corroborated these findings for the riparian areas
watershed once were lined with massive trees of small streams in western Washington and
such as American sycamores and cottonwoods Oregon, finding that forested stream buffer strips
that often attained diameters well over 2 meters moderate microclimate above the stream.
(6.6 feet). In every region of the country, the Similarly, Danehy and Kirpes (2000) found
largest trees were usually found in riparian areas increased variation in relative humidity in
where there is abundant moisture and nutrients riparian areas of harvested forests along eastern
(e.g., Muir 1878). These streamside trees were Washington streams. These studies examined
also the first to be removed for timber, relatively small (second- and third-order)
agriculture, and development. streams; riparian forest effects on a microclimate
would presumably be reduced on larger streams.
Figure 1-14. Logjam Deflecting the Hoh River in
Northwest Washington The potential for a riparian forest to provide
shade to the stream surface, and thereby to
moderate stream temperatures, has been studied
extensively, and a variety of models exist to
provide estimates of stream temperature as a
function of riparian shade (e.g., Program
SSSHADE [Bartholow 1988]). In general, the
potential of riparian shade to affect stream
temperature depends upon the fraction of water
surface receiving shade, especially during the
The logjam is approximately 70 meters (230 feet)
warmest part of the day; the temperature of the
wide and forms a 2.8-meter (10-foot) deep pool. The
logjam creates a hardpoint that allows riparian trees stream when it enters the shaded reach; and the
to mature. importance of other factors influencing stream
temperatures (e.g., stream gradient, relative
humidity, ambient air temperature, channel
morphology, and groundwater or hyporheic flow

Large Wood National Manual July 2015


1-39
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

inputs) (Beschta et al. 1987; Sedell and Swanson 1.4.2.3 Channel Features and
1984; Sullivan et al. 1990). Overall loss of riparian Characteristics
canopy cover is also associated with increased
stream temperatures, as is forest clearing at the Large wood plays an important role in increasing
basin-wide scale (Pollock et al. 2009). channel length and creating side channels,
thereby increasing overall channel complexity.
Riparian forests can influence channel form when This decreases the radius of curvature, traps
their roots stabilize streambanks and when large nutrients, provides complex channel features, and
wood from the forest enters the channel. Root increases floodplain connectivity by raising
stabilization of streambanks is effective at streambeds and water levels (Abbe and
retarding erosion, although the magnitude of Montgomery 2003; Stock et al. 2005; Abbe and
effect depends heavily upon soil pore water Brooks 2011). Wood is also a critical factor in
pressure, reaching a minimum value in saturated how floodplain forests develop (Collins et al.
soils (Pollen-Bankhead and Simon 2010). Gibling 2012; Wohl 2013). Channel conditions and wood
and Davies (2012) provide evidence that riparian loading are closely linked to the flow regime and
forest has been affecting channel form for almost sediment supply, and the characteristics of
as long as there have been trees, with broad sand- disturbances such as storms, floods, and human
bed rivers of the early Paleozoic era (circa modifications (Keeton et al. 2007). River
400 million years ago) giving way to well-defined morphology is the cumulative result of numerous
channels constrained by roots and logjams by the variables and how they change over time. Where
later Paleozoic (250 million years ago). Triska trees are large enough to create stable flow
(1984) relates the reverse of this process on the obstructions, wood becomes one of the dominant
Red River in Louisiana; during presettlement time variables controlling channel form (e.g., Abbe and
the river channel consisted of over Montgomery 1996, 2003; Abbe and Brooks 2011).
225 kilometers (140 miles) of debris jams derived In a study of streams in northern New York with
from floodplain hardwood forests, but since then bankfull widths of 2 to 16 meters (6.6 to 53 feet),
removal of debris dams to support navigation and Keeton et al. (2007) found a direct relationship
flood control has reduced the stream’s average between forest age, basal tree area, and instream
width from 185 to 40 meters (607 to 131 feet), wood volumes. Old-growth forests
and produced a greatly simplified floodplain with (205–410 years old) had instream wood loading
little in the way of riparian tree cover. Similar volumes five times those found in mature forests
changes have been described for lowland rivers in (85–145 years old): 200 cubic meters (262 cubic
western Washington (Collins et al. 2002) and yards) per hectare versus 34 cubic meters
Oregon (Sedell and Froggatt 1984). (45 cubic yards) per hectare, respectively. They
These studies also show that riparian forests also found that the presence of large logs
mediate the channel disturbance regime. Streams (>30 centimeters [12 inches] in diameter) was
with frequent and substantial inputs of large directly linked to the number of debris dams that
wood, either from catastrophic inputs (e.g., debris were primarily responsible for wood and
torrents and dam-break floods in tributary sediment retention (Figure 1-15). An aquatic
channels) or from episodic channel processes fluvial ecosystem can quickly respond to human
(e.g., bank cutting or channel avulsion) are more actions that alter a channel’s morphology, flow
likely to develop woody debris jams either within regime, or riparian forests.
or along the channel. These wood jams protect Removing wood from a river can lead to rapid
the forest from channel migration for long enough channel incision and floodplain disconnection
to allow development of large trees that will, (Figure 1-16; Veatch 1906; Guardia 1933; Brooks
when recruited to the channel, continue to and Brierly 2002; Abbe and Brooks 2011). Human
produce debris jams (Collins et al. 2012). development of the landscape has had a major

Large Wood National Manual July 2015


1-40
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

impact on the quantity of instream wood, from in a large enough area, this strategy has the
both direct removal and the deforestation of potential benefit of improving downstream flood
riparian areas. protection by lowering peak stage and discharge
(Anderson 2006). This strategy can involve the
Figure 1-15. Relationship Between Large Logs placement of large “key” logs, engineered logjams
(>30 Centimeters) and Debris Dams in Adirondack (ELJs), or beaver dams in portions of the drainage
Streams with Bankfull Widths of 2 to 16 Meters,
Northern New York network with relatively undeveloped floodplains.

1.4.2.5 Hydraulic Influence


Wood placements can be used to create pools by
generating different hydraulic conditions ranging
from plunging flow (log steps), vortex flows
associated with channel obstructions (flow
deflectors), or constriction scour associated with
narrowing the cross-sectional area. Wood can
also be used to develop and enhance riffles by
sorting bed material and setting up hydraulic
gradients that drive hyporheic flow. Complex
Source: Keeton et al. (2007). wood placements such as ELJs have been
repeatedly demonstrated to provide excellent
Figure 1-16. Removal of Wood Leads to Channel bank protection while also enhancing habitat by
Incision, Converting Alluvial Pool-Riffle Channels to creating pools and cover.
Bedrock and Damaging Habitat and Infrastructure,
Such as this Bridge Failure in the Mashel River,
Western Washington
1.5 History of the Use of
Wood for Restoration
in Streams
Wood has been humanity’s primary building
material throughout history. Timber cribbing and
piles have been used for centuries in rivers to
build bridge abutments, small dams, flood walls,
and bulkheads. The historical application of large
wood for river restoration did not begin until well
after the impacts of deforestation, agriculture, and
development. Beginning in the 1930s, coincident
with efforts to improve soil conservation, the use
1.4.2.4 Water and Sediment of large wood was focused on minor
Retention and Floodplain “improvements,” such as log weirs and timber
Connectivity cribbing to create overhanging cover for
enhancing trout habitat (Tarzwell 1934; Saunders
Placing a series of channel-spanning logs or and Smith 1955). For most of the twentieth
logjams can slow the movement of water and century, large wood was removed from streams
sediment and increase floodplain connectivity and rivers with the intent to improve navigation,
that sustains productive side channels, wetlands, reduce local flooding, or improve fish passage
hyporheic flow, and riparian forests. When done (White and Brynildson 1967). The practice of

Large Wood National Manual July 2015


1-41
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

wood removal has occurred all around the world and in-channel bridge spans, and was a threat to
(e.g., Ruiz-Villanueva et al. 2014a) and severely structures built along the banks when it deflected
impacted the hydraulic, geomorphic, and ecologic flows or created unpredictable hydraulic
role wood has played for millions of years. conditions.

The American Fisheries Society published As a result, across much of North America,
guidelines on wood in 1983, citing the potentially particularly the Pacific Northwest, wood has been
beneficial habitat that small wood placements greatly reduced in many of our streams and
could create. However, the guidelines, which are rivers. The consequences of these actions include
still available through the agency’s website, increased magnitude and frequency of flows,
continued to encourage the removal of wood which has increased channel incision, resulting in
occupying significant portions of the river channel even more severe and detrimental hydraulic
(American Fisheries Society 1983). conditions that damage habitat and infrastructure
(Figure 1-16). The alterations of ecosystem
While the physical and biologic effects of wood
functions mean long-term impacts on water
are remarkably similar across diverse ecological
quality and ecosystem structure, but they also
regions (Figure 1-1), the policies regarding wood
significantly affect the human infrastructure built
vary markedly across the country. In the Pacific
around an entirely different river than once
Northwest millions of dollars are spent annually
existed.
on reintroducing large wood to restore salmonid
habitat. But large wood is still considered a When the United States began its westward
nuisance across much of the country and is expansion, wood was commonly present in river
regularly cleared. The removal of instream wood systems, which created obstacles for those
is based more on tradition and misconceptions, pioneers. The U.S. Army was tasked with clearing
not science. Large wood removal should be wood from rivers to improve navigation and
carefully considered because leaving the wood development (Gillespie 1881; Ruffner 1886; Dacy
not only improves aquatic and riparian habitat, 1921; McCall 1984; Collins et al. 2002).
but can provide real benefits such as preventing
At the same time, recognition of wood’s role in
channel incision that can threaten infrastructure,
defining the geomorphology and ecology of fluvial
lowering groundwater tables, and exacerbating
systems appeared in some of the classic textbooks
downstream flooding.
in geology and physical geography. Lyell (1830)
Over the last 150 years there was a concerted described the formation of massive logjams and
resource management directive that cleared the lakes they created in the Red River valley of
wood from streams and rivers in an effort to Louisiana. Davis (1901) clearly describes the
enhance navigation and increase flood geomorphic effect of wood in the Red River as not
conveyance, while many land and resource just “dividing the current into many small
management practices diminished sources of channels,” but in aiding in “building of the flood
wood available within streams, riparian corridors, plain” (Davis 1901:279–280). Veatch (1906) and
and watersheds. For example, a common practice Guardia (1933) describe how removal of Red
to “improve” fish passage and flow conveyance in River logjams led to channel incision and
the Pacific Northwest was to have crews remove disconnected large areas of floodplain. Similar
wood from streams (also known as “stream logjam–dominated systems were described in the
cleaning”), particularly between the 1950s and Colorado River of Southeast Texas (Clay 1949)
1970s (Bisson et al. 1987). Wood was eradicated and occurred in many lowland alluvial rivers. The
so successfully from many streams (Reeves et al. geomorphic role of wood was described by Muir
1991) that consequences to fish habitat still exist (1878) in how giant Sequoia trees impound the
(Bisson et al. 1987). Wood was often considered a streams of the high Sierra to trap water and
nuisance when it impinged on undersize culverts create lush bogs. Russell (1909) presents similar

Large Wood National Manual July 2015


1-42
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

observations of large trees impounding the salmonid density by 167% and biomass by 162%.
Teanaway River of central Washington State, a In a similar review of 24 stream restoration
river that experienced 2 meters (6.6 feet) of projects, Miller et al. (2010) found that wood
channel incision after large wood was historically restoration projects had the largest and most
removed (Stock et al. 2005). Wolf (1916) clearly consistent benefits to macroinvertebrate
noted the role of large snags in deflecting the communities. Efforts to stabilize wood began
course of the White River of western Washington without much scientific basis regarding the
and trapping large quantities of sediment and hydraulic forces the placements would be
organic debris. Despite these observations, there subjected to, or how the stabilizing method would
was almost no scientific research conducted on perform, which could explain the failure of some
the role of large wood for most of the twentieth projects (Frissell and Nawa 1992; Abbe et al.
century, coincidently during a time when streams 1997). As an example, cable earth anchors were a
were being aggressively cleared and simplified popular stabilizing method that had limited
(Sedell and Luchessa 1982; Sedell and Frogatt success. This method involves attaching a log with
1984; Abbe 2000; Collins et al. 2002). some length of cable (typically 3 to 30 meters
[3 to 98 feet]) to an existing structure (e.g., tree)
After a long hiatus, scientific recognition about
or some sort of buried anchor, either a simple
the beneficial role of wood in river ecology and
dead weight (e.g., boulder) or a mechanism
morphology began to be published in the last
intended to maximize resistance (e.g., duckbill
40 years (Zimmerman et al. 1967; Heede 1972;
anchor). If the log began to move (float, vibrate),
Keller and Swanson 1979; Keller and Tally 1979;
so would the cable, creating a situation that could
Triska and Cromack 1979; Marston 1982; Sedell
quickly damage the bank (e.g., acting similar to a
et al. 1984; Harmon et al. 1986; Hogan 1987;
backpacker’s cable saw). In many cases, the forces
Linkaemper and Swanson 1987; Abbe and
on the log were simply too great for the anchor or
Montgomery 1996; Gippel et al. 1996; Wallerstein
bank erosion exposed the anchor (Figure 1-17).
et al. 1997; Montgomery et al. 2003; Wondzell
and Bisson 2003; Montgomery et al. 2003; Abbe Research demonstrated the key role that the size
and Brooks 2011; Collins et al. 2012). The listing and shape of trees entering the channel plays and
of Pacific Northwest salmon as threatened or how it affects river morphology (Abbe and
endangered in the 1990s began to change Montgomery 1996; Abbe 2000; Abbe and
perceptions about large wood and drive more Montgomery 2003). Replicating the massive trees
aggressive efforts to restore large wood to that once existed throughout North America is
streams after over 150 years of removal. one of the principal challenges faced in
restoration, particularly in creating stable wood
Large wood reintroduction as part of
structures. Restoration designs must rely on
rehabilitating streams began in the 1980s in
engineering designs that can emulate the natural
U.S. National Forests of the Pacific Northwest
role of old-growth timber. It is this premise under
(e.g., House and Boehne 1985). Early wood
which ELJs were developed, not only to
placements typically entailed placing log “dams”
demonstrate the physical significance wood plays
across relatively small channels and often
in defining channel morphology and habitat, but
resulted in significant biological benefits
how wood can be used to protect infrastructure
(e.g., Wallace et al. 1995). Unstable or simple
by limiting bank erosion and channel incision
wood placements along the banks of channels
(Abbe et al. 1997, 2003a, b, c; Abbe and Brooks
tended to have little or no benefit (Frissell and
2011). In the 18 years since the first ELJ
Nawa 1992; Beamer and Henderson 1998; Peters
prototype was built in 1995, there have been
et al. 1998). After assessing 211 restoration
hundreds of ELJs and thousands of wood
projects involving instream structures, Whiteway
placements in the Pacific Northwest. Wood
et al. (2010) found the projects increased
stability has been a critical issue in many

Large Wood National Manual July 2015


1-43
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

restoration programs and a variety of techniques centuries where it is submerged (Figure 1-18)
have been developed to increase their design life (Abbe and Brooks 2011).
and ensure wood remains in the original location
(Abbe et al. 2003a; Abbe and Brooks 2011). Figure 1-18. A Buried Log More than 500 Years Old
Forming Grade Control, Coal Creek, 2004, Ozette
Figure 1-17. Stable Wood Bifurcates Flow Leading River Tributary, Washington
to Anabranching Channels when Undisturbed, and
Creates a Complex and Productive Habitat

The fate of wood in rivers is integrally tied to how


riparian forests are managed. Large mature trees
are essential in providing wood large enough to
influence habitat formation. Concerns about wood
Photo credit: Ken DeCamp stability and life expectancy should be
The extensive application of ELJs in the Pacific anticipated, but can always be addressed with
Northwest has led to a general standard of good science and engineering. There are
practice that has greatly improved structure situations where large wood is not appropriate or
stability. Stable large wood placements in the where it can pose unacceptable hazards, but it is
Pacific Northwest are now common, and clear that instream wood is beneficial and should
numerous ELJs have successfully weathered be an integral part of watershed management
severe floods, including events equal to or throughout the country.
exceeding the 100-year flood (Abbe and Brooks
2011). Observations of natural wood and ELJs are
also demonstrating that wood can last for many
decades under various conditions, and even for

Large Wood National Manual July 2015


1-44
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

1.6 References
Abbe, T. B. 2000. Patterns, Mechanics, and Geomorphic Effects of Wood Debris Accumulations in a
Forest River System. Ph.D. dissertation. University of Washington, Seattle, WA. 222 pp.

Abbe, T. B., and A. P. Brooks. 2011. Geomorphic, Engineering, and Ecological Considerations when
Using Wood in River Restoration. Pages 419–451 in A. Simon, S. J. Bennett, and J. M. Castro
(eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools.
Geophysical Monograph Series 194. Washington, D.C.: American Geophysical Union.

Abbe, T. B., and D. R. Montgomery. 1996. Large Woody Debris Jams, Channel Hydraulics and Habitat
Formation in Large Rivers. Regulated Rivers: Research and Management 12:201–221.

Abbe, T. B., and D. R. Montgomery. 2003. Patterns and Processes of Wood Debris Accumulation in
the Queets River Basin, Washington. Geomorphology 51:81–107.

Abbe, T. B., D. R. Montgomery, and C. Petroff. 1997. Design of Stable In-Channel Wood Debris
Structures for Bank Protection and Habitat Restoration: An Example from the Cowlitz River, WA.
Pages 809–816 in S. S. Y. Wang, E. J. Langendoen, and F. D. Shields, F.D. (eds.), Proceedings of the
Conference on Management of Landscapes Disturbed by Channel Incision. University of
Mississippi, Oxford, MS.

Abbe, T. B., J. Carrasquero, M. McBride, A. Ritchie, M. McHenry, and K. Dublanica. 2003a.


Rehabilitating River Valley Ecosystems: Examples of Public, Private, and First Nation Cooperation
in Western Washington. Proceedings of the Georgia Basin/Puget Sound 2003 Research
Conference, Vancouver, B.C., March 31–April 1, 2003, T. Droscher (ed.). Puget Sound Action
Team, Olympia, WA.

Abbe, T. B, A. P. Brooks, and D. R. Montgomery. 2003b. Wood in River Rehabilitation and


Management. Pages 367–389 in S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology
and Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society.

Abbe, T. B., G. Pess, D. R. Montgomery, and K. L. Fetherston. 2003c. Integrating Engineered Log Jam
Technology into River Rehabilitation. In D. R. Montgomery, S. Bolton, D. Booth, and L. Wall
(eds.), Restoration of Puget Sound Rivers. Center for Water and Watershed Studies.

Agee, J. K. 1990. The historical role of fire in Pacific Northwest forests. Pages 25–38 in J. Walstad, S.
R. Radosevich, and D. V. Sandberg (eds.), Natural and Prescribed Fire in Pacific Northwest Forests.
Corvallis: Oregon State University Press.

Agee, J. K. 1992. The Historical Role of Fire in Pacific Northwest Forests. Pages 25–38 in J. Walstad, S.
R. Radosevich, and D. V. Sandberg (eds.), Natural and Prescribed Fire in Pacific Northwest Forests.
Corvallis: Oregon State University Press.

Agee, J. K. 1993. Fire Ecology of Pacific Northwest Forests. Washington, D.C.: Island Press

Ahmad, M. 1951. Spacing and Projection of Spurs for Bank Protection. Civil Engineering and Public
Works Review. March:172–174; April:256–258.

Allan, J. D., M. S. Wipfli, J. P. Caouette, A. Prussian, and J. Rodgers. 2003. Influence of Streamside
Vegetation on Inputs of Terrestrial Invertebrates to salmonid Food Webs. Canadian Journal of
Fisheries and Aquatic Sciences 60(3):309–320.

Large Wood National Manual July 2015


1-45
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

American Fisheries Society. 1983. Stream Obstruction Removal Guidelines. Stream Renovation
Guidelines Committee. The Wildlife Society and American Fisheries Society. Published by AFS,
Washington D.C. 9 pp.

Anderson, D. B. 2006. Quantifying the Interaction between Riparian Vegetation and Flooding: from
Cross-Section to Catchment Scale. University of Melbourne.

Anderson, N. H., R. J. Steedman, and T. Dudley. 1984. Patterns of Exploitation by Stream


Invertebrates of Wood Debris (Xylophagy). Verhandlungen der Internationalen Vereinigung für
theoretische und angewandte Limnologie 22:1847-1852.

Anderson, P. D., D. J. Larson, and S. S. Chan. 2007. Riparian Buffer and Density Management
Influences on Microclimate of Young Headwater Forests of Western Oregon. Forest Science
53(2):254–269.

Andrus, C. W., B. A. Long, and H. A. Froehlich. 1988. Woody debris and its contribution to pool
formation in a coastal stream 50 years after logging: Canadian Journal of Fish and Aquatic
Science 45:2080–2086.

Bailey, R. G. 1995. Description of Ecoregions of the United States, 2nd Edition, USDA, US Forest Service.
Washington, D.C. Miscellaneous Publication No. 1391.

Bailey, R. G. 2009. Ecosystem Geography – From Ecoregions to Sites. New York, NY: Springer Science
and Business Media.

Baillie, B. R., L. G. Garret, and A. W. Evanson. 2008. Spatial Distribution Influence of LWD in an Old-
growth Forest River System, New Zealand. Forest Ecology and Management 256:20–27.

Barker, B. L., R. D. Nelson, and M. S. Wigmosta. 1991. Performance of detention ponds designed
according to current standards. Puget Sound Water Quality Authority, Puget Sound Research '91:
Conference Proceedings. Seattle, Washington.

Bartholow, J. 1988. Stream Segment Shade Model (SSSHADE) Version 1.4. Temperature Model
Technical Note #3. U.S. Fish and Wildlife Service, Fort Collins, CO.

Baxter, C. V., K. D. Fausch, and W. Carl Saunders. 2005. Tangled Webs: Reciprocal Flows of
Invertebrate Prey Link Streams and Riparian Zones. Freshwater Biology 50(2):201–220.

Beamer, E. M., and R. A. Henderson. 1998. Juvenile Salmonid use of Natural and Hydromodified
Stream Bank Habitat in the Mainstem Skagit River, Northwest Washington. Miscellaneous Report.
Skagit System Cooperative. La Connor, WA.

Beechie, T. J., and K. Wyman. 1992. Stream Habitat Conditions, Unstable Slopes and Status of Roads in
Four Small Watersheds of the Skagit River. Skagit System Cooperative, Fisheries services for the
Swinomish Tribal Community, Upper Skagit and Sauk-Suiattle Indian Tribes.

Beets, P. N., I. A. Hood, M. O. Kimberley, G. R. Oliver, S. H. Pearce, and J. F. Gardner. 2008. Coarse
Woody Debris Decay Rates for Seven Indigenous Tree Species in the Central North Island of New
Zealand. Forest Ecology and Management 256:548–557.

Beltaos, S. 1983. River Ice Jams: Theory, Case Studies, and Applications. Journal of Hydraulic
Engineering 109(10):1338–1359.

Large Wood National Manual July 2015


1-46
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Bencala, K. E. 2005. Hyporheic Exchange Flows. Encyclopedia of Hydrological Sciences, M. G.


Anderson and J. J. Mcdonnell (eds.). Wiley-Blackwell. 3,456 pp.

Benda, L. and T. W. Cundy. 1990. Predicting Deposition of Debris Flow in Mountain Channels.
Canadian Geotechnical Journal 27:409–417.

Benke, A. C., and J. B. Wallace. 2010. Influence of Wood on Invertebrate Communities in Streams and
Rivers. American Fisheries Society Symposium 37:149–177.

Benke, A. C., R. L. Henry III, D. M. Gillespie, and R. J. Hunter. 1985. Importance of Snag Habitat for
Animal Production in Southeastern Streams. Fisheries 10:8–12.

Berg, N. A., A. Carlson, and D. Azuma. 1998. Function and Dynamics of Woody Debris in Stream
Reaches in the Central Sierra Nevada, California. Canadian Journal of Fisheries and Aquatic
Sciences 55:1807–1820.

Beschta, R. L., R. E. Bilby, L. B. Brown, L. B. Holtby, and T. D. Hofstra. 1987. Stream Temperature and
Aquatic Habitat: Fisheries and Forestry Interactions. Pages 191-232 in E. O. Salo and T. W.
Cundy (eds.), Streamside Management: Forestry and Fishery Interactions. College of Forest
Resources, University of Washington, Seattle, WA. 471p.

Bilby, R. E. 1984. Removal of Woody Debris May Affect Stream Channel Stability. Journal of Forestry,
609–613. October.

Bilby, R. E., and P. A. Bisson. 1998. Function and Distribution of Large Woody Debris. Pages 324–346
in R. J. Naiman and R. E. Bilby (eds.), River Ecology and Management: Lessons from the Pacific
Coast Ecoregion. New York, NY: Springer-Verlag.

Bilby, R. E., and G. E. Likens. 1980. Importance of Debris Dams in the Structure and Function of
Stream Ecosystems. Ecology 61:1107–1113.

Bilby, R. E., and J. W. Ward. 1989. Changes in Characteristics and Function of Woody Debris With
Increasing Size of Streams in Western Washington. Transactions of the American Fisheries
Society 118:368–378.

Bilby, R. E. and J. W. Ward. 1991. Characteristics and Function of Large Woody Debris in Streams
Draining Old-Growth, Clear-Cut, and Second-Growth Forests in Southwestern Washington.
Canadian Journal of Fisheries and Aquatic Sciences 48:2499–2508.

Bilby, R. E., and L. J. Wasserman. 1989. Forest Practices and Riparian Management in Washington
State: Data Based Regulation Development. In R. E. Gresswell, B. A. Barton, and J. L. Kershner
(eds.), Practical Approaches to Riparian Management. U.S. Bureau of Land Management, BLM MT
PT 89 001 4351, Billings, Montana.

Bisson, P. A., R. E. Bilby, M. D. Bryant, C. A. Dolloff, G. B. Grette, R. A. House, M. L. Murphy, K. V. Koski,


and J. R. Sedell. 1987. Large Woody Debris in Forested Streams in the Pacific Northwest: Past,
Present, and Future. Pages 143–190, in E. O. Salo and T. W. Cundy (eds.), Streamside
Management: Forestry and Fishery Interactions. College of Forest Resources, University of
Washington, Seattle, Washington.

Boose, E. R., K. E. Chamberlin, and D. R. Foster. 2001. Landscape and Regional Impacts of Hurricanes
in New England. Ecological Monographs 71:27–48.

Large Wood National Manual July 2015


1-47
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Booth, D. B. 1990. Stream-Channel Incision Following Drainage-Basin Urbanization. Water Resources


Bulletin 26:407–417.

Boulton, A. J., T. Datry, T. Kasahara, M. Mutz, and J.A. Stanford. 2010. Ecology and Management of the
Hyporheic Zone – Groundwater Interactions of Running Waters and Their Floodplains. Journal
of the North American Benthological Society 29:26–40.

Braun, A., K. Auerswald, and J. Geist. 2012. Drivers and Spatio-Temporal Extent of Hyporheic Patch
Variation: Implications for Sampling. PLOS One 7:e42046.

Brooks, A. P., and G. J. Brierly. 2002. Mediated Equilibrium: The Influence of Riparian Vegetation and
Wood on the Long-Term Evolution and Behavior of a Near-Pristine River. Earth Surface
Processes and Landforms 27:343–367.

Brosofske, K. D., J. Chen, R. J. Naiman, and J. F. Franklin. 1997. Harvesting Effects on Microclimatic
Gradients from Small Streams to Uplands in Western Washington. Ecological Applications
7(4):1188–1200.

Bryant, M. D. 1980. Evolution of large, Organic Debris after Timber Harvest: Maybeso Creek, 1949 to
1978. USDA Forest Service, General Technical Report, PNW-101.

Bryant, M. D. 1983. The Role and Management of Woody Debris in West Coast Salmonid Nursery
Stream. North American Journal of Fisheries Management 3(3):322–330.

Bryant, M. D., and J. R. Sedell. 1995. Riparian Forests, Wood in the Water, and Fish Habitat
Complexity. Pages 202–224 in N. B. Armantrout and R. J. Wolotira, Jr. (eds.), Conditions of the
World's Aquatic Habitats. Proceedings of the World Fisheries Congress Theme 1. Oxford and IBH
Publishing Co. Pvt. Ltd., New Delhi.

Buffington, J. M. and D. Tonina. 2009. Hyporheic Exchange in Mountain Rivers II: Effects of Channel
Morphology on Mechanics, Scales, and Rates of Exchange. Geography Compass 3:1038–1062.

Camp, A., C. Oliver, P. Hessburg, and R. Everett. 1996. Predicting Late-Successional Fire Refugia Pre-
Dating European Settlement in the Wenatchee Mountains. USDA PNW, Wenatchee For. Sci. Lab.,
Univ. of Washington, Seattle. Elsevier Science Publishers B.V. Forest Ecology and Management
95:63–77.

Castelle, A. J., A. W. Johnson, and C. Conolly. 1994. Wetland and Stream Buffer Size Requirements—A
Review. Journal of Environmental Quality 23(5):878–882.

Cederholm, C. J., R. E. Bilby, P. A. Bisson, T. W. Bumstead, B. R. Fransen, W. J. Scarlett, and J. W. Ward.


1997b. Response of Juvenile Coho Salmon and Steelhead to the Placement of Large Woody
Debris in a Coastal Washington Stream. Transactions of the American Fisheries Society. 118:368–
378.

Cederholm, C. J., W. J. Scarlett, N. P. and Peterson. 1988. Low-Cost Enhancement Technique for
Winter Habitat of Juvenile Coho Salmon. North American Journal of Fisheries Management
8:438–441.

Chambers, J. Q., J. I. Fisher, H. Zeng, E. L. Chapman, D. B. Baker, and G. C. Hurtt. 2007. Hurricane
Katrina’s Carbon Footprint on U.S. Gulf Coast Forests. Science 318 (5853):1107.

Large Wood National Manual July 2015


1-48
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Chen, J., J. F. Franklin, and T. A. Spies. 1995. Growing-Season Microclimatic Gradients from Clearcut
Edges into Old-Growth Douglas-Fir Forests. Ecological Applications 5(1):74–86.

Chesney, C. 2000. Functions of Wood in Small, Steep Streams in Eastern Washington: Summary of
Results for Project Activity in the Ahtanum, Cowiche, and Tieton Basins. Washington Department
of Natural Resources. Prepared for the Timber/Fish/Wildlife Monitoring Advisory Group and
the Northwest Indian Fisheries Commission. TFW Effectiveness Monitoring Report: TFW-MAGl-
00-002.

Chin, A., M. D. Daniels, M. A. Urban, H. Piegay, K. J. Gregory, W. Bigler, A. Z. Butt, J. L. Grable, S. V.


Gregory, M. Lafrenz, L. R. Laurencio, and E. Wohl. 2008. Perceptions of Wood in Rivers and
Challenges for Stream Restoration in the United States. Environmental Management 41:893–903.

Clay, C. 1949. The Colorado River Raft. The Southwestern Historical Quarterly 102 (4):400–426.

Coe, H. J., P. M. Kiffney, G. R. Press, K. K. Kloehn, and M. L. McHenry. 2009. Periphyton and
Invertebrate Response to Wood Placement in Large Pacific Coastal Rivers. River Research and
Applications 25(8):1025–1035.

Coho, C., and S. J. Burges. 1993. Dam-Break Floods in Low Order Mountain Channels of the PNW.
Water Resources Series Tech Rep no. 138. Dept. Civil Engineering, Univ. of Washington, Seattle. 68
pp.

Coho, C., and S. J. Burges. 1994. Dam Break Floods in Low Order Mountain Channels of the Pacific
Northwest. TFW SH9 93 001. Timber Fish and Wildlife, Department of Natural Resources,
Olympia. 70 pp.

Collins, B. D., and A. J. Sheikh. 2005. Historical Reconstruction, Classification, and Change Analysis of
Puget Sound Tidal Marshes. University of Washington (Seattle, WA) and the Nearshore Habitat
Program, Washington State Dept. of Natural Resources, Olympia, WA. See more at:
http://www.eopugetsound.org/science-review/3-tidal-wetlands#sthash. T4OyhfFd.dpuf

Collins, B. D., D. R. Montgomery, and A. D. Haas. 2002. Historical Changes in the Distribution and
Functions of Large Wood in Puget Lowland Rivers. Canadian Journal of Fisheries and Aquatic
Sciences 59:66–76.

Collins, B. D., D. R. Montgomery, and A. J. Sheikh. 2003. Reconstructing the Historical Riverine
Landscape of the Puget Lowland. Pages 79–128 in D. R. Montgomery, S. M. Bolton, D. B. Booth,
and L. Wall (eds.), Restoration of Puget Sound Rivers. University of Washington Press, Seattle.

Collins, B. D., D. R. Montgomery, K. L. Fetherston, and T. B. Abbe. 2012. The Floodplain Large-Wood
Cycle Hypothesis: A Mechanism for the Physical and Biotic Structuring of Temperate Forested
Alluvial Valleys in the North Pacific Coastal Ecoregion. Geomorphology 139/140:460–470.

Compton, J. E., M. R. Church, S. T. Larned, and W. E. Hogsett. 2003. Nitrogen Export from Forested
Watersheds in the Oregon Coast Range: The Role of N2-Fixing Red Alder. Ecosystems 6(8):773–
785.

Cordova, J. M., E. J. Rosi-Marshall, A. M. Yamamuro, and G. A. Lamberti. 2007. Quantity, Controls, and
Functions of Large Woody Debris in Midwestern USA Streams. River Research and Applications
23:21–23.

Large Wood National Manual July 2015


1-49
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Crook, D., and A. Robertson. 1999. Relationships between Riverine Fish and Woody Debris:
Implications for Lowland Rivers. Marine and Freshwater Research 50:941–953.

Cushman, M. J. 1981. The Influence of Recurrent Snow Avalanches on Vegetation Patterns in the
Washington Cascades. Ph.D. dissertation. University of Washington, Seattle, Washington.

Dacy, G. H. 1921. Pulling the Mississippi’s Teeth. Scientific American 75(4):60, 70.

Danehy, R. J., and B. J. Kirpes. 2000. Relative Humidity Gradients across Riparian Areas in Eastern
Oregon and Washington Forests. Northwest Science 74(3):224–233.

Davis, W. M. 1901. Physical Geography. Boston, MA: Ginn and Company.

Dickman, A., and S. Cook. 1989. Fire and Fungus in a Mountain Hemlock Forest. Canadian Journal of
Botany 67:2005–2016.

Doloff, C. A., and M. L. Warren, Jr. 2003. Fish Relationships With Large Wood in Small Streams.
American Fisheries Symposium 37:179–193.

Dominguez, L. G., and C. J. Cederholm. 2000. Rehabilitating Stream Channels Using Large Woody
Debris with Considerations for Salmonid Life History and Fluvial Geomorphic Processes. Pages
545–563 in E. E. Knudsen. C. R. Steward, D. D. MacDonald, J. E. Williams, and D. W. Reiser (eds.),
Sustainable Fisheries Management: Pacific Salmon. Lewis Publishers, New York.

Edmonds, R. L., T. B. Thomas, and K. P. Maybury. 1993. Tree Population Dynamics, Growth, and
Mortality in old-Growth Forests in the Western Olympic Mountains, Washington. Canadian
Journal of Forest Research 23:512–519.

Elmore, W., and R. L. Beschta. 1988. The Fallacy of Structures and the Fortitude of Vegetation. Proc.
of Calif. Riparian Systems Conference. Davis, Calif.

Environmental Agency. 2009. The Hyporheic Handbook. a Handbook of the Groundwater-Surface


Water Interface and Hyporheic Zone for Environmental Managers. Science Report SC050070. 264
pp. Available: http://www.hyporheic.net/SCHO1009BRDX-e-e.pdf. Accessed: June 13, 2014.

Fahnestock, G. R. 1976. Fires, Fuel, and Flora as Factors in Wilderness Management: The Pasayten
Case. Tall Timbers Fire Ecology Conf. 15:33–70.

Fahnestock, G. R., and J. K. Agee. 1983. Biomass Consumption and Smoke Production by Prehistoric
and Modern Forest Fires in Western Washington. Journal of Forestry 81:653–657.

Fetherston, K. L., R. J. Naiman, and R. E. Bilby. 1995. Large Woody Debris, Physical Process, and
Riparian Forest Development in Montane River Networks of the Pacific Northwest.
Geomorphology 13:133–144. Elsevier Science B.V.

Findlay, S., J. Tank, S. Dye, H. M. Valett, P. J. Mulholland, W. H. McDowell, S. L. Johnson, S. K. Hamilton,


J. Edmonds, W. K. Dodds, and W. B. Bowden. 2002. A Cross System Comparison of Bacterial and
Fungal Biomass in Detritus Pools of Headwater Streams. Microbial Ecology 43(1):55–66.

Flebbe, P. A. 1999. Trout Use of Wood Debris and Habitat in Wine Spring Creek, North Carolina.
Forest Ecology and Management 114:367–376.

Large Wood National Manual July 2015


1-50
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Forest Ecosystem Management Assessment Team (FEMAT). 1993. Forest Ecosystem Management:
An Ecological, Economic, and Social Assessment. Report of the Forest Ecosystem Management
Assessment Team. July.

Foster, D. R., and E. R. Boose. 1992. Patterns of Forest Damage Resulting from Catastrophic Wind in
Central New England, USA. Journal of Ecology 80:79–98.

Fox, M. J. 2001. A New Look at the Quantities and Volumes of Instream Wood in Forested Basins within
Washington State. Master of Science thesis. College of Forest Resources, University of
Washington.

Fox, M. J. 2003. Spatial Organization, Position, and Source Characteristics of Large Woody Debris in
Natural Systems. Ph.D. dissertation. College of Forest Resources, University of Washington.
Seattle, Washington.

Fox, M. J. and S. Bolton. 2007. A Regional and Geomorphic Reference for Quantities and Volumes of
Instream Wood in Unmanaged Forested Basins of Washington State. North American Journal of
Fisheries Management 27:342–359.

Frangi, J. L., and A. E. Lugo. 1991. Hurricane Damage to a Flood Plain Forest in the Luquillo
Mountains of Puerto Rico. Biotropica 23(4a):324–335.

Franklin, J. F., and C. T. Dyrness. 1973. Natural Vegetation of Oregon and Washington. USDA Forest
Service. Gen. Tech. Rep. PNW-8.

Frissell, C. A., and R. K. Nawa. 1992. Incidence and Causes of Physical Failure of Artificial Habitat
Structures in Streams of Western Oregon and Washington. North American Journal of Fisheries
Management 12 182–197.

Gandy, C. J., and A. P. Jarvis. 2006. Attenuation of Nine Pollutants in the Hyporheic Zone. Environment
Agency, Bristol, England, June. 33 pp.

Gastaldo, R. A., and C. W. Degges. 2007. Sedimentology and Paleontology of a Carboniferous Log Jam.
International Journal of Coal Geology 69:103–113.

Gastaldo, R. A., and T. M. Demko. 2011. The Relationship Between Continental Landscape Evolution
and the plant-Fossil Record: Long Term Hydrologic Controls on Preservation. Pages 249–285 in
P. A. Allison and D. J. Bottjer (eds.), Taphonomy: Process and Bias Through Time. Aims & Scope
Topics in Geobiology Volume 32. Springer Netherlands.

Gibling, M. R., and N. S. Davies. 2012. Palaeozoic Landscapes Shaped by Plant Evolution. Nature
Geoscience 5(2):99–105.

Gibling, M. R., A. R. Bashforth, H. J. Falcon-Lang, J. P. Allen, and C. R. Fielding. 2010. Log Jams and
Flood Sediment Buildup Caused Channel Abandonment and Avulsion in the Pennsylvanian of
Atlantic Canada. Journal of Sedimentary Research 80:268–287.

Gillespie, Major G. L. 1881. Report of the Chief of Engineers, U.S. Army. Appendix OO 10, 2603–2605.

Gippel, C. J., I. C. O’Neill, and B. L. Finlayson. 1996. Distribution and Hydraulic Significance of Large
Woody Debris in a Lowland Australian River. Hydrobiologia 318:179–194.

Graham, R., and K. Cromack. 1982. Mass, Nutrient Content, and Decay Rate of Dead Boles in Rain
Forests of Olympic National Park. Canadian Journal of Forest Research 12(3):511–521.

Large Wood National Manual July 2015


1-51
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Grant, G. E., and F. J. Swanson. 1995. Morphology and Processes of Valley Floors in Mountain
Streams, western Cascades, Oregon. Pages 83–101 in J. D. Costa, A. J. Miller, K. W. Potter, and P.
R. Wilcock (eds.). Natural and Anthropogenic Influences in Fluvial Geomorphology. Geophysical
Monograph 89. American Geophysical Union, Washington DC.

Grant, G. E., M. J. Crozier, and F. J. Swanson. 1984. An Approach to Evaluating Off-Site Effects of
Timber Harvest Activities on Channel Morphology. Proceedings of the Symposium on the Effects
of Forest and Land Use on Erosion and Slope Stability. Environment and Policy Institute, E-West
Center, University of Hawaii, Honolulu 177–186.

Gregory, S. V., F. J. Swanson, W. A. McKee, and K. W. Cummins. 1991. An Ecosystem Perspective of


Riparian Zones. BioScience 41(8):540–551.

Gregory, S. V., K. L. Boyer, and A. M. Gurnell (eds.). 2003. The Ecology and Management of Wood in
World Rivers. Bethesda, MD: American Fisheries Society.

Grette, G. B. 1985. The role of Large Organic Debris in Juvenile Salmonid Rearing Habitat in Small
Streams. MS thesis, University of Washington, Seattle, WA.

Grizzel, J. D., and N. Wolff. 1998. Occurrence of Windthrow in Forest Buffer Strips and its Effect on
Small Streams in Northwest Washington. Northwest Science 72:214–223.

Grizzel, J., M. McGowan, D. Smith, and T. Beechie. 2000. Streamside Buffers and Large Woody Debris
Recruitment: Evaluating the Effectiveness of Watershed Analysis Prescriptions in the North
Cascades Region. TFW-MAGI-00-003. Washington State Timber, Fish & Wildlife.

Guardia, J. E. 1933. Some Results of the Log Jams in the Red River. The Bulletin of the Geographical
Society of Philadelphia 31(3):103–114.

Guyette, R. P., D. C. Dey, and M. C. Stambaugh 2008. The Temporal Distribution and Carbon Storage
of Large Oak Wood in Streams and Floodplain Deposits. Ecosystems 11:643–653.

Hafs, A. W., L. R. Harrison, R. M. Utz, and T. Dunneda. 2014. Quantifying the Role of Woody Debris in
Providing Bioenergetically Favorable Habitat for Juvenile Salmon. Ecological Modelling 286:30–
38.

Harmon, M. E., J. F. Franklin, F. J. Swanson, P. Sollins, S. V. Gregory, J. D. Lattin, N. H. Anderson, S. P.


Cline, N. G. Aumen, J. R. Sedell, G. W. Lienkaemper, K. Cromack, Jr., and K. W. Cummins. 1986.
Ecology of Coarse Woody Debris in Temperate Ecosystems. Advances in Ecological Research
15:133–302.

Hartopo, 1991. The Effect of Raft Removal and Dam Construction on the Lower Colorado River, Texas.
Unpublished M.S. Thesis, Texas A & M University.

Hauer, F. R. 1989. Organic Matter Transport and Retention in a Blackwater Stream Recovering from
Flow Augmentation and Thermal Discharge. Regulated Rivers: Research and Management 4:371–
380.

Hedman, C. W., D. H. Van Lear, and W. T. Swank. 1996. In-Stream Large Woody Debris Loading and
Riparian Forest Seral Stage Associations in the Southern Appalachian Mountains. Canadian
Journal of Forest Research 26:1218–1227.

Large Wood National Manual July 2015


1-52
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Heede, B. H. 1972. Influences of a Forest on the Hydraulic Geometry of Two Mountain Streams.
Water Resources Bulletin 8:523–530.

Henderson, J. 1996. Unpublished Data Regarding Tree Height vs. Age for Two Common Plant
Association Groups. USDA Forest Service, Pacific Northwest Region, Mount Lake Terrace, WA.

Henderson, J. A., R. D. Lesher, D. H. Peter and D. C. Shaw. 1992. Field Guide to the Forested Plant
Associations of the Mt. Baker-Snoqualmie National Forest. USDA Forest Service, Pacific Northwest
Region. Tech paper R6 ECOL TP 028-91.

Hershey, K. 1995. Characteristics of Forests at Spotted Owl Nest Sites in the Pacific Northwest. M.S.
thesis, Oregon State University, Corvallis.Hewitt, E. R. 1934. Hewitt’s Handbook of Stream
Improvement. The Marchbanks Press, New York.

Hester, E. T., and M. N. Gooseff. 2010. Moving Beyond the Banks: Hyporheic Restoration is
Fundamental to Restoring Ecological Services and Functions of Streams. Environmental Science
and Technology 44:1521–1525.

Hester, E. T., M. W. Doyle, and G. C. Poole. 2009. The Influence of in-Stream Structures on Summer
Water Temperatures via Induced Hyporheic Exchange. Limnology and Oceanography 54:355–
367.

Hewitt, E. R. 1934. Hewitt’s Handbook of Stream Improvement. New York: The Marchbanks Press.

Hickin E. J. 1984. Vegetation and River Channel Dynamics. Canadian Geographer 28(2):111–126.

Hinkle, S. R., J. H. Duff, F. J. Triska, A. Laenen, E. B. Gates, K. E. Bencala, D. A. Wentz, and S. R. Silva.
2001. Linking Hyporheic Flow and Nitrogen Cycling near the Willamette River – A Large River In
Oregon, USA. Journal of Hydrology 244:157–180.

Hogan, D. L. 1987. The influence of large organic debris on channel recovery in the Queen Charlotte
Islands, British Columbia, Canada. Pages 343–353 in R. L. Beschta, T. Blinn, G. E. Grant, F. J.
Swanson, and G. G. Ice (eds.), Erosion and Sedimentation in the Pacific Rim. IAHS Publication
No.165.

Hollis, G. E. 1975. The Effects of Urbanization on Floods of Different Recurrence Intervals. Water
Resources Research 11:431–435.

Holstine, C. 1992. An Historical Overview of the Wenatchee National Forest, Washington. Rep. 100-80.
Archaeological and historical Services. Eastern Washington University, Cheney.

Horner, R. R., D. B. Booth, A. Azous, and C. W. 1997. Watershed Determinants of Ecosystem


Functioning. Pages 251–274 in L. A. Roesner (ed.), Effects of Watershed Development and
Management on Aquatic Ecosystems, American Society of Civil Engineers, New York, NY.

House, R. A., and P. L. Boehne. 1985. Evaluation of Instream Enhancement Structures for Salmonid
Spawning and Rearing in a Coastal Oregon Stream. North American Journal of Fish Management
5:283–295.

House, R. A., and P. L. Boehne. 1986. Effects of Instream Structures on Salmonid Habitat and
Populations in Tobe Creek, Oregon. North American Journal of Fisheries Management 6:283–295.

Hyatt, T. L., and R. J. Naiman. 2001. The Residence Time of Large Woody Debris in the Queets River,
Washington, USA. Ecological Applications 11(1):191–202.

Large Wood National Manual July 2015


1-53
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Ikeya, H. 1981. A Method for Designation Forested Areas in Danger of Debris Flows. In Erosion and
Sediment Transport in Pacific Rim Steeplands. Edited by T. R. H. Davies and A. J. Pearce.
International Association of Hydrological Sciences, Publication 132:576–588.

Johnson, L. B., D. H. Breneman, and C. Richards. 2003. Macroinvertebrate Community Structure and
Function Associated with Large Wood in Low Gradient Streams. River Research and Applications
19:199–218.

Johnson, S. L., F. J. Swanson, G. E. Grant, and S. M. Wondzell. 2000. Riparian Forest Disturbances by a
Mountain Flood—The Influence of Floated Wood. Hydrological Processes 14:3031–3050.

Kauffman, J. B., R. L. Beschta, N. Otting, and D. Lytjen. 1997. An Ecological Perspective of Riparian
and Stream Restoration in the Western United States. Fisheries (Bethesda) 22:12–24.

Keeton, W. S., C. E. Kraft, and D. R. Warren. 2007. Mature and Old-Growth Riparian Forests:
Structure, Dynamics and Effects on Adirondack Stream Habitats. Ecological Applications 17:852–
868.

Keller, E. A. and A. MacDonald. 1995. River Channel Change: The Role of Large Woody Debris. Pages
217–236 in A. Gurnell and G. Petts (eds.), Changing River Channels. John Wiley and Sons,
Chichester. 217-235.

Keller, E. A., and F. J. Swanson. 1979. Effects of Large Organic Material on Channel Form and Fluvial
Processes. Earth Surface Processes 4:361–380.

Keller, E. A., and T. Tally. 1979. Effects of Large Organic Debris on Channel Form and Fluvial
Processes in the Coastal Redwood Environment. Pages 169–197 in D. D. Rhodes and G. P.
Williams (eds.), Adjustments of the Fluvial System. Proceedings of the 10th Annual Binghamton
Geomorphology Symposium. Kendal-Hunt. Dubuque, IA.

Kennard, P., G. Pess, T. Beechie, B. Bilby, and D. Berg. 1998. Riparian-in-a-Box: A Manager’s Tool to
Predict the Impacts of Riparian Management on Fish Habitat. Pages 483-490. in M. K. Brewin
and D. M. A. Monita (eds.), Forest-Fish Conference: Land Management Practices Affecting Aquatic
Ecosystems. Proceedings of Forest-fish conference, May 1-4, 1996, Calgary, Alberta. Natural
Resources Canada. North For. Cent., Edmonton, Alberta Inf. Rep. NOR-X-356.

Koehn, J. D., W. G. O’Connor, P. D. Jackson. 1994. Seasonal and size-Related Variation in Microhabitat
Use of a Small Victorian Stream Fish Assemblage. Australian Journal of Marine and Freshwater
Research 45:1353–1366.

Krause, C., and C. Roghair. 2014. Inventory of Large Wood in the Upper Chattooga River Watershed,
2007–2013. U.S. Forest Service Southern Research Station, Center for Aquatic Technology
Transfer. Blacksburg, VA.

Krause, S., M. J. Klaar, D. M. Hannah, J. Mant, J. Bridgeman, M. Trimmer, and S. Manning-Jones. 2014.
The Potential of Large Woody Debris to Alter Biogeochemical Processes and Ecosystem Services
in Lowland Rivers. Wiley Interdisciplinary Reviews (WIREs): Water 1:263–275.

Lancaster, S. T., S. K. Hayes, and G. E. Grant. 2001. Modeling Sediment and Wood Storage and
Dynamics in Small Mountainous Watersheds. Geomorphic Processes and Riverine Habitat, Water
Science and Application Volume 4:85–102. American Geophysical Union.

Large Wood National Manual July 2015


1-54
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Lautz, L. K., D. I. Siegel, and R. L. Bauer. 2006. Impact of Debris Dams on Hyporheic Interaction along
a Semi-Arid Stream. Hydrological Processes 20:183–196.

Lee, P. C., C. Smyth, and S. Boutin. 2004. Quantitative Review of Riparian Buffer Width Guidelines
from Canada and the United States. Journal of Environmental Management 70:165–189.

Lemly, A. D., and R. H. Hilderbrand. 2000. Influence of Large Woody Debris on Stream Insect
Communities and Benthic Detritus. Hydrobiologia 421:179–185.

Leopold, L. B. 1973. River Channel Change with Time: An Example. Geological Society of America

Lienkaemper, G. W., and F. J. Swanson. 1987. Dynamics of Large Woody Debris in Streams in Old-
Growth Douglas-Fir Forests. Canadian Journal of Forest Research 17:150–156.

Lisle, T. 1995. Effects of Coarse Woody Debris and its Removal on a Channel Affected by the 1980
Eruption of Mount St. Helens, Washington. Water Resources Research 31:1797–1808.

Lockaby, B. G., J. A. Stanturf, and M. G. Messina. 1997. Effects of Silvicultural Activity on Ecological
Processes in the Floodplain Forests of the Southern United States: A Review of Existing Reports.
Forest Ecology and Management 90:93–100.

Lyell, C. 1830. Principles of Geology, Volume I. London, UK: John Murray. Published in 1990 by
University of Chicago Press. Chicago, IL.

Magilligan, F. J., K. H. Nislov, G. B. Fisher, J. Wright, G. Mackey, and M. Laser 2008. The Geomorphic
Function and Characteristics of Large Woody Debris in Low Gradient Rivers, Coastal Maine, USA.
Geomorphology 97:467–482.Makaske, B., D. G. Smith, and H. J. Berendsen. 2002. Avulsions,
Channel Evolution and Floodplain Sedimentation Rates of the Anastomosing Upper Columbia
River, British Columbia, Canada. Sedimentology 49(5):1049–1071.

Makaske, B., D. G. Smith, and H. J. Berendsen. 2002. Avulsions, Channel Evolution and Floodplain
Sedimentation Rates of the Anastomosing Upper Columbia River, British Columbia, Canada.
Sedimentology 49(5):1049–1071.

Marston, R. A. 1982. The Geomorphic Significance of Log Steps in Forested Streams. Annals of the
Association of American Geographers 72:99–108.

Martin, D. J., and L. E. Benda. 2001. Patterns of Instream Wood Recruitment and Transport at the
Watershed Scale. Transactions of the American Fisheries Society 130:940–958.

Maser, C., and J. M. Trappe (eds.). 1984. The Seen and Unseen World of the Fallen Tree. Gen. Tech. Rep.
PNW-164. Portland, OR: U.S. Forest Service, Pacific Northwest Forest and Range Experiment
Station.

Maser, C., R. F. Tarrant, J. M. Trappe, and J. F. Franklin (eds.). 1988. From the Forest to the Sea: A Story
of Fallen Trees. General Tech. Report PNW-GTR-229. USFS. 153 pp.

McCall, E. 1984. Conquering the Rivers. Louisiana State University Press. Baton Rouge, LA.

McDade, M. H., F. J. Swanson, W. A. McKee, J. F. Franklin, and J. Van Sickle. 1990. Source Distances for
Coarse Woody Debris Entering Small Streams in Western Oregon and Washington. Canadian
Journal of Forest Research 20:326–330.

Large Wood National Manual July 2015


1-55
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

McHenry, M. L., E. Shott, R. H. Conrad, and G. B. Grette. 1998. Changes in the Quantity and
Characteristics of LWD in Streams of the Olympic Peninsula, Washington, USA (1982-1993).
Canadian Journal of Fisheries and Aquatic Sciences 55(6):1395–1407.

Means, J. E., K. Cromack Jr., and P. C. MacMillan, 1986, Comparison of Decomposition Models Using
Wood Density of Douglas-Fir Logs. Canadian Journal of Forestry Research 15:1092–1098.

Melillo, J. M., R. J. Naiman, J. D. Aber, and K. N. Eshleman. 1983. The Influence of Substrate Quality
and Stream Size on Wood Decomposition Dynamics. Oecolgia (Berlin) 58:281–285.

Mellina, E. and S. G. Hinch. 2009. Influences of Riparian Logging and in-Stream Large Wood Removal
on Pool Habitat and Salmonid Density and Biomass: A Meta-Analysis. Canadian Journal of Forest
Research 39:1280–1301.

Miller, D., C. Luce, and L. Benda. 2003. Time, Space, and Episodicity of Physical Disturbance in
Streams. Forest Ecology and Management 178(1):121–140.

Miller, S. W., P. Budy, and J. C. Schmidt. 2010. Quantifying Macroinvertebrate Responses to In-Stream
Habitat Restoration. Applications of Restoration Ecology 18:8–19.

Millward, A. A., C. E. Kraft, and D. R. Warren. 2010. Ice Storm Damage Greater Along the Terrestrial-
Aquatic Interface in Forested Landscapes. Ecosystems 13:249–260.

Montgomery, D. R., and T. B. Abbe. 2006. Influence of Logjam-Formed Hard Points on the Formation
of Valley-Bottom Landforms in an Old-Growth Forest Valley, Queets River, Washington, USA.
Quaternary Research 65:147–155.

Montgomery, D. R., and J. M. Buffington. 1993. Channel Classification, Prediction of Channel Response,
and Assessment of Channel Condition. TFW-SH10-93-002. Washington State Timber, Fish &
Wildlife.

Montgomery, D. R., and J. M. Buffington. 1997. Channel-Reach Morphology in Mountain Drainage


Basins. Geological Society of America Bulletin 109:596–611.

Montgomery, D. R., T. B. Abbe, J. M. Buffington, N. P. Peterson, K. M. Schmidt, and J. D. Stock. 1995a.


Distribution of Bedrock and Alluvial Channels in Forested Mountain Drainage Basins. Nature
381:587–589.

Montgomery, D. R., J. M. Buffington, R. D. Smith, K. M. Schmidt, and G. Pess. 1995b. Pool Spacing in
Forest Channels. Water Resources Research 31:1097–1105.

Montgomery, D. R., B. D. Collins, J. M. Buffington, and T. B. Abbe. 2003. Geomorphic Effects of Wood
in Rivers. Pages 21–47 in S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology and
Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society.

Moore, M. K. 1977. Factors Contributing to Blowdown in Streamside Leave Strips on Vancouver


Island. Land Management Report No. 3. Victoria, BC: Province of British Columbia Ministry of
Forests, Information Division.

Moulin, B., E. R. Schenk, and C. R. Hupp. 2011. Distribution and Characterization of In-channel Large
Wood in Relation to Geomorphic Patterns on a Low-gradient River. Earth Surface Processes and
Landforms 36:1137–1151.

Large Wood National Manual July 2015


1-56
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Muir, J. 1878. Forests of California, the New Sequoia. Harper’s New Monthly Magazine LVII
(CCCXLII):813–827.

Mulholland, P. J., and J. R. Webster. 2010. Nutrient Dynamics in Streams and the Role of J-NABS.
Journal of the North American Benthological Society 29:100–117.

Mulholland, P. J., and 33 others. 2009. Nitrate Removal in Stream Ecosystems Measured by 15N
Addition Experiments: Denitrification. Limnology and Oceanography 54:666–680.

Murphy, M. L. 1995. Forestry Impacts on Freshwater Habitat of Anadromous Salmonids in the


Pacific Northwest and Alaska—Requirements for Protection and Restoration. U.S. Department of
Commerce Coastal Ocean Program, NOAA. Decision Analysis Series No. 7, 156 pp.

Murphy, M. L., and K. V. Koski. 1989. Input and Depletion of Woody Debris in Alaska Streams and
Implications for Streamside Management. North American Journal of Fisheries Management
9(4):427–436.

Mutz, M., E. Kalbus, and S. Meinecke. 2007. Effect of Instream Wood on Vertical Water Flux in Low-
Energy Sand Bed Flume Experiments. Water Resources Research 43:W10424.

Nagayama, S. and F. Nakamura. 2010. Fish Habitat Rehabilitation Using Wood in the World.
Landscape and Ecologic Engineering 6:289–305.

Naiman, R. J., T. J. Beechie, L. E. Benda, P. A. Bisson, L. H. MacDonald, M. D. O’Conner, P. L. Olsen, and


E. A. Steel. 1992. Fundamental elements of ecologically healthy watersheds in the Pacific
Northwest coastal ecoregion. Pages 127–188 in R. J. Naiman (ed.), Watershed Management:
Balancing Sustainability and Environmental Change. Springer: New York.

Naiman, R. J., K. L. Fetherston, S. McKay, and J. Chen. 1998. Riparian Forests. Pages 289-323 in R. J.
Naiman and R. E. Bilby (eds.), River Ecology and Management: Lessons from the Pacific Coastal
Ecoregion. Springer-Verlag: New York.

Naiman, R. J., E. V. Balian, K. K. Bartz, R. E. Bilby, and J. J. Latterell. 2002. Dead Wood Dynamics in
Stream Ecosystems. Pages 23–48 in W. F. Laudenslayer Jr., P. J. Shea, B. E. Valentine, C. P.
Weatherspoon, and T. E. Lisle (eds.), Proceedings of the Symposium on the Ecology and
Management of Dead Wood in Western Forests. Gen. Tech. Rep. PSW-GTR-181. US Forest Service,
Pacific Southwest Forest and Range Experiment Station.

National Marine Fisheries Service. 1996. Making Endangered Species Act Determinations of Effect for
Individual or Grouped Actions at the Watershed Scale. Environmental and Technical Services
Division, Habitat Conservation Branch.

Neumann, R. M., and T. L. Wildman. 2002. Relationships Between Trout Habitat Use and Woody
Debris in Two Southern New England Streams. Ecology of Freshwater Fish 11:240–250.

Nickelson, T. E., M. F. Solazzi, S. L. Johnson, and J. D. Rodgers. 1992. Effectiveness of Selected Stream
Improvement Techniques to Created Suitable Summer and Winter Rearing Habitat for Juvenile
Coho Salmon (Oncorhynchus kisutch) in Oregon Coastal Streams. Canadian Journal of Fisheries
and Aquatic Sciences 49:790–794.

North American Forest Commission. 2011. Forests of North America. Vector Digital Data. Food and
Agriculture Organization of the United Nations. Commission for Environmental Cooperation.
Montreal, Quebec, CA.

Large Wood National Manual July 2015


1-57
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Oliver, C. D. 1980/1981. Forest Development in North America Following Major Disturbances. Forest
Ecology and Management 3:153–168.

Oregon Department of Forestry. 1995. A Guide to Placing Large Wood in Streams. Salem, OR, Forest
Practices Section. 13 pp.

Palik, B., S. W. Golladay, P. C. Goebel, and B. W. Taylor. 1998. Geomorphic Variation in Riparian Tree
Mortality and Stream Coarse Woody Debris Recruitment from Record Flooding in a Coastal Plain
Stream. Ecoscience 5:551–560.Pariset, E., R. Hausser, and A. Gagnon. 1966. Formation of Ice
Covers and Ice Jams in Rivers. Journal of the Hydraulics Division 92(6):1–24.

Pariset, E., R. Hausser, and A. Gagnon. 1966. Formation of Ice Covers and Ice Jams in Rivers. Journal
of the Hydraulics Division 92(6):1–24.

Parrish, R. M. and P. B. Jenkins. 2012. Natural Log Jams in the White River: Lessons for Geomimetic
Design of Engineered Log Jams. U.S. Fish and Wildlife Service, Leavenworth, WA.

Pearsons, T. D., and H. W. Li. 1992. Influence of Habitat Complexity on Resistance to Flooding and
Resilience of Stream Fish Assemblages. Transactions of the American Fisheries Society 121:427–
436.

Peters, P. J., B. R. Missildine, and D. L. Low. 1998. Seasonal Fish Densities near River Banks Stabilized
with Various Stabilization Methods. First Year Report of the Flood Technical Assistance Project.
U.S. Fish and Wildlife Service, North Pacific Coast Ecoregion. Western Washington Office,
Aquatic Resources Division. Lacey, WA. 34 pp.

Petts, G. E., A. L. Roux, and H. Moller (eds.). 1989. Historical Changes of Large Alluvial Rivers, Western
Europe. Chichester: John Wiley.

Phillips, J. D. 2012. Log-jams and Avulsions in the San Antonio River Delta, Texas. Earth Surface
Processes and Landforms 37:936–950.

Phillips, J. D., and L. Park. 2009. Forest Blowdown Impacts of Hurricane Rita on Fluvial Systems.
Earth Surface Processes and Landforms 34:1069–1081.

Piégay, H., A. and R. A. Marston. 1998. Distribution of Coarse Woody Debris Along the Concave Bank
of a Meandering River (the Ain River, France). Physical Geography 19(4):318–340.

Piégay, H., A. Thevenet, and A. Citterio. 1999. Input, Storage and Distribution of LWD Along a
Mountain River Continuum, the Drôme River, France. Catena 35:19–39.

Pollen-Bankhead, N., and A. Simon. 2010. Hydrologic and Hydraulic Effects of Riparian Root
Networks on Streambank Stability: Is Mechanical Root-Reinforcement the Whole Story?
Geomorphology 116(3):353–362.

Pollock, M. M., and T. J. Beechie. 2014. Does Riparian Forest Restoration Thinning Enhance
Biodiversity? The Ecological Importance of Large Wood. JAWRA Journal of the American Water
Resources Association 50(3):543–559. Online publication date: June 1, 2014.

Pollock, M. M., R. J. Naiman, and T. A. Hanley. 1998. Plant Species Richness in Riparian Wetlands—A
Test of Biodiversity Theory. Ecology 79:94–105.

Large Wood National Manual July 2015


1-58
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Pollock, M. M., T. J. Beechie., M. Liermann, and R. E. Bigley. 2009. Stream Temperature Relationships
to Forest Harvest in Western Washington. Journal of the American Water Resources Association
45(1):141–156.

Prowse, T. D. 2001. River Ice Ecology. 1: Hydrologic, Geomorphic, and Water Quality Aspects. Journal
of Cold Regions Engineering 15(1):1–16.

Ralph, S. C., G. C. Poole, L. L. Conquest, and R. J. Naiman. 1991. Stream Channel Morphology and
Woody Debris in Logged and Unlogged Basins of Western Washington. Canadian Journal of
Fisheries and Aquatic Sciences 51:37–51.

Raup, H. M. 1957. Vegetation Adjustment to the Instability of Sites. Proceedings and Papers of the 6th
Technical Meeting of the International Union for Conservation of Nature and Natural Resources.
Edinburgh. Pages 36–48.

Reeves, G. H., J. D. Hall, T. D. Roelofs, T. L. Hickman, and C. O. Baker. 1991. Rehabilitating and
Modifying Stream Habitats. Pages 519–557 in Influences of Forest and Rangeland Management
on Salmonid Fishes and Their Habitats. American Fisheries Society Special Publication 19.

Reid, L. M., and S. Hilton. 1998. Buffering the Buffer. Pages 71–80 in R. R. Ziemer (ed.), Proceedings of
the Conference on Coastal Watersheds: The Caspar Creek Story; held May 6, 1998, in Ukiah,
California. USDA Forest Service, Pacific Southwest Research Station, General Technical Report
PSW-GTR-168.

Richmond, A. D., and K. D. Fausch. 1995. Characteristics and Function of Large Woody Debris in
Subalpine Rocky Mountain Streams in Northern Colorado. Canadian Journal of Fisheries and
Aquatic Sciences 52:1789–1802.

Riley, S. C. and K. D. Fausch. 1995. Trout Population Response to Habitat Enhancement in Six
Northern Colorado Streams. Canadian Journal of Fisheries and Aquatic Sciences. 52:34–53.

Robison, E. G. and R. L. Beschta. 1990. Identifying Trees in Riparian Areas that can Provide Coarse
Woody Debris to Streams. Forest Science 36:790–801.

Roni, P., M. Liermann, and A. Steel. 2003. Monitoring and Evaluating Fish Response to Instream
Restoration. In D. Montgomery, S. Bolton, D. Booth, and L. Wall (eds.), Restoration of Puget Sound
Rivers. Center for Water and Watershed Studies. University of Washington Press: Seattle.

Roni, P., T. Beechie, G. Pess, and K. Hanson. 2014a. Wood Placement in River Restoration: Fact,
Fiction, and Future Direction. Canadian Journal of Fisheries and Aquatic Sciences 72(3):466–478.

Rosenfeld, J. S., and L. Huato. 2003. Relationship Between Large Woody Debris Characteristics and
Pool Formation in Small Coastal British Columbia Streams. North American Journal of Fisheries
Management 23:928–938.

Rosgen, D., and H. L. Silvey. 1996. Applied River Morphology. Wildland Hydrology. Pagosa Springs, CO.

Rot, B. 1993. Windthrow in Stream Buffers on Coastal Washington Streams. ITT-Rayonier Inc. 49 pp.

Rot, B. W., R. J. Naiman, and R. E. Bilby. 2000. Stream Channel Configuration, Landform, and Riparian
Forest Structure in the Cascade Mountains, Washington. Canadian Journal of Fisheries and
Aquatic Sciences 57:699–707.

Large Wood National Manual July 2015


1-59
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Ruffner, E. H. 1886. The Practice of the Improvement of the Non-Tidal Rivers of the United States, with
an Examination of the Results Thereof. New York, NY: John Wiley and Sons.

Ruiz-Villanueva, V., A. Díez-Herrero, J. M. Bodoque, and E. Bladé. 2014. Large Wood in Rivers and its
Influence on Flood Hazard. Cuadernos de Investigación Geográfica 40:229–246.

Russell, I.C. 1909. Rivers of North America. New York, NY: G.P. Putnam and Sons.

Saunders, J. W. and M. W. Smith. 1955. Physical Alteration of Stream Habitat to Improve Brook Trout
Production. Canadian Fish Culturist 16:185–188.

Saunders, J. W. and M. W. Smith. 1962. Physical Alteration o f Stream Habitat to Improve Brook
Trout Production. Transactions of the American Fisheries Society 82:185–188.

Schenk, E. R., J. W. McCargo, B. Moulin, C. R. Hupp, and J. M. Richter. 2014a. The Influence of Logjams
on Largemouth Bass (Micropterus salmoides) Concentrations on the Lower Roanoke River, a
Large Sand-bed River. River Research and Applications. www.wileyonlinelibrary.com, DOI:
10.1002/rra.2779

Schenk, E. R., B. Moulin, C. R. Hupp, J. M. Richter. 2014b. Large Wood Budget and Transport
Dynamics on a Large River Using Radio Telemetry. Earth Surface Processes and Landforms
39:487–498.

Scherer, R. 2004. Decomposition and Longevity of In-Stream Woody Debris: A Review of Literature
from North America. Pages 127–133 in Forest Land–Fish Conference–Ecosystem Stewardship
through Collaboration. Proceedings of Forest-Land-Fish Conference II.

Schuett-Hames, D., A. E. Pleus, J. Ward, M. Fox, and J. Light. 1999. TFW Monitoring Program Methods
Manual for the Large Woody Debris Survey. Prepared for the Washington State Dept. of Natural
Resources under the Timber, Fish, and Wildlife Agreement. TFW-AM9-99-004. DNR #106.
March.

Sedell, J. R., and J. L. Frogatt. 1984. Importance of Streamside Forests to Large Rivers: The Isolation
of the Willamette River, Oregon, U.S.A., from its Floodplain by Snagging and Streamside Forest
Removal. Verhandlungen-Internationale Vereinigung für Theorelifche und Angewandte
Limnologie 22:1828–1834.

Sedell, J. R., and K. J. Luchessa. 1981. Using the Historical Record as an Aid to Salmonid Habitat
Enhancement. Symposium on Acquisition and Utilization of Aquatic Habitat Inventory
Information. October 23–28, Portland, OR.

Sedell, J. R., and K. J. Luchessa. 1982. Using the Historical Record as an Aid to Salmonid Habitat
Enhancement. Pages 222–245 in N. B. Armantrout (ed.). Acquisition and Utilization of Aquatic
Habitat Inventory Information. Proceedings of a Symposium October 28–30, 1981. Billings, MT:
The Hague Publishing.

Sedell, J. R., and F. J. Swanson. 1984. Ecological Characteristics of Streams in Old-Growth Forests of
the Pacific Northwest. Pages 9–16 in W. R. Meehan, T. R. Merrell Jr., and T. A. Hanley (eds.), Fish
and Wildlife Relationships in Old-Growth Forests. Juneau, AK: American Institute of Fisheries
Research Biologists.

Sedell, J. R., F. H. Everest, and F. J. Swanson. 1982. Fish Habitat and Streamside Management: Past
and Present. Pages 244–255 in Proceedings of the 1981 Convention of the Society of American

Large Wood National Manual July 2015


1-60
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Foresters, September 27–30, 1981. Society of American Foresters, Publication 82–01, Bethesda,
Maryland.

Sedell, J. R., F. J. Swanson, and S. V. Gregory. 1984. Evaluating Fish Response to Woody Debris. Pages
191–221 in T. J. Hassler (ed.). Proceedings of the Pacific Northwest Streams Habitat Management
Workshop. American Fisheries Society. Humboldt State University. Arcata, CA.

Seehorn, M. E. 1985. Fish Habitat Improvement Handbook: Atlanta, Georgia. U.S. Forest Service,
Southern Region. Technical Publication R8-TP-16. 30 pp.

Senter, A. E., and G. B. Pasternack. 2010. Large Wood Aids Spawning Chinook Salmon (Oncorhynchus
Tshawytscha) in Marginal Habitat on a Regulated River in California. River Research and
Applications 27:550–565.

Shields, F. D., Jr., and R. H. Smith. 1992. Effects of Large Woody Debris Removal on Physical
Characteristics of a Sand Bedded River. Aquatic Conservation: Marine and Freshwater Ecosystems
2:145–163.

Singer, S., and M. L. Swanson. 1983. The Soquel Creek Storm Damage Recovery Plan with
Recommendations for Reduction of Geologic Hazards in Soquel Village, Santa Cruz County,
California. Unpublished USDA Soil Conservation Service report to the Santa Cruz County Board
of Supervisors.

Smith, D. G. 1979. Effects of Channel Enlargement by River Ice Processes on Bankfull Discharge in
Alberta, Canada. Water Resources Research, 15(2):469–475.

Smith, D. G., and C. M. Pearce. 2000. River Ice and its Role in Limiting Woodland Development on a
Sandy Braid-Plain, Milk River, Montana. Wetlands, 20(2):232–250.

Smith, D. G., and D. M. Reynolds. 1983. Tree Scars to Determine the Frequency and Stage of High
Magnitude River Ice Drives and Jams, Red Deer, Alberta. Canadian Water Resources Journal
8(3):77–94.

Smock, L. A., G. M. Metzler and J. E. Gladden. 1989. Role of Debris Dams in the Structure and
Functioning of Low Gradient Headwater Streams. Ecology 70:764–775.

Society for Ecological Restoration. 2004. The SER International Primer on Ecological Restoration.
Available: <http://www.ser.org>.

Solazzi, M. F., T. E. Nickelson, S. L. Johnson, and J. D. Rodgers. 2000. Effects of Increasing Winter
Rearing Habitat on Abundance of Salmonids in Two Coastal Oregon Streams. Canadian Journal of
Fisheries and Aquatic Sciences 57:906–914.

Sollins, P., S. P. Cline, T. Verhoeven, D. Sachs, and G. Spycher. 1987. Patterns of Log Decay in Old-
Growth Douglas-Fir Forests, Canadian Journal of Forest Research 17:1585–1595.

Spänhoff, B., and E. Cleven. 2010. Wood in Different Stream Types: Epixylic Biofilm and Wood-
Inhabiting Invertebrates in a Lowland Versus an Upland Stream. Annales De Limnologie-
International Journal of Limnology 46(3):169–179.

Spänhoff, B., and E. I. Meyer. 2004. Breakdown Rates of Wood in Streams. Journal of the North
American Benthological Society 23(2):189–197.

Large Wood National Manual July 2015


1-61
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Spänhoff, B., C. Alecke, and E. Irmgard Meyer. 2001. Simple Method for Rating the Decay Stages of
Submerged Woody Debris. Journal of the North American Benthological Society 20(3):385–394.

Spies, T. A., J. F. Franklin, and T. B. Thomas. 1988. Coarse Woody Debris in Douglas-Fir Forests of
Western Washington and Oregon. Ecology 69:1689–1702.

Spies, T. A., and J. F. Franklin. 1991. The Structure of Natural Young, Mature, and Old-Growth
Douglas Fir Forests in Oregon and Washington. Pages 91–109 in L. F. Ruggiero, K. B. Aubrey, A.
B. Carey, and M. H. Huff (technical coordinators), Wildlife and Vegetation of Unmanaged Douglas
Fir Forests. USDA Forest Service. General Technical Report PNW-GTR-285.

Stanford, J. A. and J. V. Ward. 1988. The Hyporheic Habitat of River Ecosystems. Nature 335:64–66.

Steinhart, G. S., G. E. Likens, and P. M. Groffman. 2000. Denitrification in Stream Sediments in Five
Northeastern (USA) Streams. Verhandlungen Internationale Vereinigung für Theorertische und
Angewandte Limnologie 27:1331–1336.

Stock, J. D., D. R. Montgomery, B. D. Collins, W. E. Dietrich, and L. Sklar. 2005. Field Measurements of
Incision Rates Following Bedrock Exposure: Implications for Process Controls on the Long
Profiles of Valleys Cut by Rivers and Debris Flows. Geological Society of America Bulletin
117(11/12):174–194.

Sullivan, K. J., J. Tooley, K. Doughty, J. E. Caldwell, and P.A. Knudsen. 1990. Evaluation of Prediction
Models and Characterization of Stream Temperature Regimes in Washington. TFW-WQ3-90-006,
Timber Fish & Wildlife, Department of Natural Resources, Olympia, WA.

Sundbaum, K. and I. Naslund. 1998. Effects of Woody Debris on the Growth and Behavior of Brown
Trout in Experimental Stream Channels. Canadian Journal of Zoology 76:56–61.

Swanson, F. J., S. V. Gregory, J. R. Sedell, and A. G. Campbell. 1982. Land-Water Interactions: The
Riparian Zone. Pages 267–291 on R. L. Edmonds (ed.), Analysis of Coniferous Forest Ecosystems in
the Western United States. US/IBP Synthesis Series, Hutchinson Ross Publishing Company:
Stroudsburg, PA.

Swanson, F. J., T. K. Kranz, N. Caine, and R. G. Woodmansee. 1988. Landform Effects on Ecosystem
Patterns and Processes. BioScience 38:92–98.

Tappeiner, J. C., D. Huffman, D. Marshall, T. A. Spies, and J. D. Bailey. 1997. Density, Ages, and Growth
Rates in Old-Growth and Young-Growth Forests in Coastal Oregon. Paper 3166 of the Forest
Research Laboratory, Oregon State University, Corvallis.

Tarzwell, C. M. 1934. The Purpose and Value of Stream Improvement Method. Stream Improvement.
Bulletin R-4. Presented at the Annual Meeting of the American Fisheries Society. Ogden, UT.

Tarzwell, C. M. 1936. Experimental Evidence of the Value of Trout Stream Improvements.


Transactions of the American Fisheries Society 66:177–187.

Thompson, D. M. 2002. Long-term Effect of Instream Habitat-improvement Structures on Channel


Morphology along the Blackledge and Salmon Rivers, Connecticut, USA. Environmental
Management 29(1):250–265.

Thompson, D. M. 2005. The History of the Use and Effectiveness of Instream Structures in the United
States. Geological Society of America Reviews in Engineering Geology XVI:35–50.

Large Wood National Manual July 2015


1-62
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Tonina, D., and J. M. Buffington. 2009. Hyporheic Exchange In Mountain Rivers I: Mechanics and
Environmental Effects. Geography Compass 3:1063–1086.

Triska, F. J. 1984. Role of Large Wood in Modifying Channel Morphology and Riparian Areas of a
Large Lowland River under Pristine Conditions: A Historical Case Study. Verhandlungen-
InternationaleVereinigung für Theorelifche und Angewandte Limnologie 22:1876–1892.

Triska, F. J., and K. Cromack, Jr. 1979. The Role of Wood Debris in Forests and Streams. In R.H.
Waring, Forests: Fresh Perspectives from Ecosystem Analysis. Pages 171–190 in Proceedings of
the 40th Annual Biology Colloquium. Corvallis, OR: Oregon State University Press. Corvallis, OR.

Tufekcioglu, A., J. W. Raich, T. M. Isenhart, and R. C. Schultz. 2003. Biomass, Carbon and Nitrogen
Dynamics of Multi-Species Riparian Buffers within an Agricultural Watershed in Iowa, USA.
Agroforestry Systems 57(3):187–198.

U.S. Department of Agriculture (USDA). 1980. Ecoregions of the United States. U.S. Forest Service,
Washington, D.C. Miscellaneous Publication No. 1391

Valett, H. M., C. L. Crenshaw, and P. F. Wagner. 2002. Stream Nutrient Uptake, Forest Succession, and
Biogeochemical Theory. Ecology 83:2888–2901.

Van Cleef, J. S. 1885. How to Restore Our Trout Streams. Transactions of the American Fisheries
Society 14:50–55.

Van Sickle, J., and S. V. Gregory. 1990. Modeling Inputs of Large Woody Debris to Streams from
Falling Trees. Canadian Journal of Forest Research 20(10):1593–1601.

Veatch, A. C. 1906. Geology and Underground Water Resources of Northern Louisiana and Southern
Arkansas. Washington D.C. United States Geological Survey Professional Paper 46.

Wadsworth, A. H., Jr. 1966. Historical Deltation of the Colorado River, Texas. Pages 99–105 in Deltas
in Their Geologic Framework. American Association of Petroleum Geologists.

Wallace, J. B., and A. C. Benke. 1984. Quantification of Wood Habitat in Subtropical Coastal Plain
Streams. Canadian Journal of Fisheries and Aquatic Sciences 41:1643–1652.

Wallace, J. B., J. R. Webster, and J. L. Meyer. 1995. Influence of Log Additions on Physical and Biotic
Characteristics of a Mountain Stream. Canadian Journal of Fisheries and Aquatic Sciences
52:2120–2137.

Wallerstein, N., C. R. Thorne, and M. W. Doyle. 1997. Spatial Distribution and Impact of Large Woody
Debris in Northern Mississippi. Pages 145–150 in C. C. Wang, E. J. Langendoen, and F. D. Shields
(eds.), Proceedings of the Conference on Management of Landscapes Disturbed by Channel
Incision. University of Mississippi. Oxford, MI.

Ward, J. V., K. Tockner, and F. Schiemer. 1999. Biodiversity of Floodplain River Ecosystems: Ecotones
and Connectivity. Regulated Rivers Research and Management 15:125–139.

Ward, J. V., K. Tockner, D. B. Arscott, and C. Claret. 2002. Riverine Landscape Diversity. Freshwater
Biology 47:517–539.

Warren, D. R., E. S. Bernhardt, R. O. Hall Jr., and G. E. Likens. 2007. Forest Age, Wood and Nutrient
Dynamics in Headwater Streams of the Hubbard Brook Experimental Forest. N.H. Earth Surface
Processes & Landforms 32(8):1154–1163.

Large Wood National Manual July 2015


1-63
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 1. Large Wood Introduction

Webster, J. R., and E. F. Benfield. 1986. Vascular Plant Breakdown in Freshwater Ecosystems. Annual
Review of Ecology and Systematics 17(1):567–594.

White, R. J., and O. M. Brynildson. 1967. Guidelines for Management of Trout Stream Habitat in
Wisconsin. Wisconsin Department of Natural Resources Technical Bulletin 39. Madison, WI.

Whiteway, S. L., P. M. Biron, A. Zimmermann, O. Venter, and J. W. A. Grant. 2010. Do In-Stream


Restoration Structures Enhance Salmonid Abundance? A Meta-Analysis. Canadian Journal of
Fisheries and Aquatic Sciences 67:831–841.

Whitney, G. G. 1996. From Coastal Wilderness to Fruited Plain: A History of Environmental Change in
Temperate North America from 1500 to the Present. Cambridge University Press: Cambridge, UK.

Wilford, D., Maloney, D., Schwab, J., and Geertsema, M. 1998. Tributary Alluvial Fans. B.C. Ministry of
Forests Extension Note 30.

Wiltshire, P. E. J., and P. D. Moore. 1983, Paleovegetation and Paleohydrology in Upland Britain.
Pages 433–451 in K. J. Gregory (ed.), Background to Paleohydrology. John Wiley: Chichester, UK.

Wohl E., D. A. Cenderelli, K. A. Dwire, S. E. Ryan-Burkett, M. K. Young, and K. D. Fausch. 2010. Large
in-Stream Wood Studies: A Call for Common Metrics. Earth Surface Processes and Landforms
35:618–625.

Wohl, E. 2013. Floodplains and Wood. Earth-Science Reviews 123:194–212.

Wolff, H. H. 1916. The Design of a Drift Barrier Across the White River, near Auburn, Washington.
Transactions of the American Society of Civil Engineers 16:2061–2085.

Wondzell, S. M. 2011. The Role of the Hyporheic Zone across Stream Networks. Hydrological
Processes 25:3525–3532.

Wondzell, S.M., and P. A. Bisson. 2003. Influence of Wood on Aquatic Biodiversity. Pages 249–263 in
S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology and Management of Wood in
World Rivers. American Fisheries Society Symposium 37. Bethesda, MD: American Fisheries
Society.

Wondzell, S. M., J. LaNier, R. Haggerty, R. D. Woodsmith, and R. T. Edwards. 2009. Changes in


Hyporheic Flow Following Experimental Removal of a Small, Low-Gradient Stream. Water
Resources Research 45:W05406, 13 pp.

Wuehlisch, G. Von. 2011. Evidence for nitrogen-Fixation in the Salicaceae Family. Tree Planters’
Notes 54(2):38–41.

Zeng, H., J. Q. Chambers, R. I. Negron-Juarez, G. C. Hurtt, D. B. Baker, and M. D Powell. 2009. Impacts
of Tropical Cyclones on U.S. Forest Tree Mortality and Carbon Flux from 1851 to 2000.
Proceedings of the National Academy of Sciences 106(19), 7888–7892.

Zimmerman, R. C., J. C. Goodlett, and G. H. Comer. 1967. The Influence of Vegetation on Channel Form
of Small Streams, Symposium on River Morphology. International Association of Science
Hydrology Publication, Gentbrugge, Belgium 75:255–275.

Zobel, D. B., A. McKee, G. M. Hawk, and C. T. Dyrness. 1976. Relationships of Environment to


Composition, Structure, and Diversity of Forest Communities of the Central Western Cascades of
Oregon. Ecological Monographs 46:135–156.

Large Wood National Manual July 2015


1-64
Chapter 2
LARGE WOOD AND THE FLUVIAL
ECOSYSTEM RESTORATION PROCESS

Placement of large wood in the Lower Elwha River, Elwha River Ecosystem Restoration Project,
Lower Elwha Klallam Tribe, Port Angeles, Washington. Source: Tim Abbe.

AUTHORS

Leo D. Lentsch (ICF International)


Tim Abbe (NSD)
This page intentionally left blank.
as the supply and movement of sediment from
2.1 Introduction hillslopes, large wood recruitment, shading of
the channel, and the volume of water (Roni
This chapter provides a general overview of the 2005; Roni et al. 2002). Many processes that
ecological restoration planning and decision- create in-channel features operate on time
making process and how it applies to the scales of decades or longer (e.g., channel
overall planning and implementation of projects migration). Interrupting these processes (e.g.,
that use large wood to restore process and by stabilizing banks or by constructing roads
function to fluvial aquatic ecosystems. It and levees) can lead to loss of ecosystem
describes the important components to functions for decades or even centuries
consider by proposing a 12-element planning (Beechie and Bolton 1999). As such, most
framework that leads to successful restoration anthropogenic activities tend to disrupt natural
projects. Inherent to the restoration process is processes that form habitat (e.g., delivery of
the recognition that suitable solutions may wood, water, sediment, and nutrients).
include a wide range of design elements, from
simple changes in resource management It is important to note that most habitat
practices to major structural alterations, the enhancement efforts by themselves tend to be
selection of which depends on the nature of relatively short lived (less than a decade) if the
each individual project. To this end, an underlying ecological process that has been
integrated approach to the planning and disrupted is not corrected (Roni 2005). To this
decision-making process provides the end, restoration of watershed processes should
foundation for selecting and using appropriate occur in conjunction with site-specific habitat
tools and procedures for placing wood in enhancement.
streams.

Anthropogenic activities have degraded aquatic GUIDANCE


ecosystems around the world. As a result, large
efforts to restore function to these ecosystems Restoration actions, intended to offset the effects of
have occurred for a variety of economic, anthropogenic activities, can affect species habitat
through two major pathways:
cultural, and environmental reasons (Roni
2005; NRC 1992). In North America, hundreds 1. Some habitat restoration approaches focus on
of millions of dollars are invested annually by restoring natural processes (e.g., road removal,
federal, state, and local agencies for restoring or riparian replanting) and thus affect ecosystem
improving fish habitat alone. Millions more are functions by influencing the underlying watershed
invested by local programs to restore function processes (e.g., sediment supply, delivery of
organic material).
to aquatic ecosystems for other social and
economic values such as flood control. For these 2. Other techniques focus on manipulating or
efforts to be successful, an understanding as enhancing habitats for organisms at specific sites
well as consideration of the immutable controls (e.g., wood placement for cover). Restoration
(e.g., geology, climate) and processes (e.g., actions should be at a scale commensurate with
delivery of wood, water, and sediment) that environmental problems. (Roni 2005)
affect and create fluvial aquatic ecosystems
needs to be incorporated into the restoration Resource managers and restoration
planning process. practitioners should be mindful of the broader
Land use and other anthropogenic activities can watershed context and recognize that coupling
affect ecosystem functions by disrupting the site-specific enhancement efforts with
processes that form and/or sustain them, such restoration of basic watershed processes will be
the most efficient course for habitat restoration.

Large Wood National Manual July 2015


2-1
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

It is also important to recognize that there are and Rinne 1990; NRC 1992; Kauffman et al.
many scientific, societal, and economic factors 1997; Beechie and Bolton 1999).
to consider when planning a restoration project.
For example, cost, cost–benefit (e.g., fish/dollar,
CROSS-REFERENCE
area restored/dollar), habitat quality, location,
access, land ownership, endangered species,
As discussed in Chapter 1, Large Wood Introduction,
and other factors often must be considered
ecological restoration is an “intentional activity that
when planning restoration projects. Focusing initiates or accelerates the recovery of an ecosystem
on restoring watershed and ecosystem with respect to its health, integrity and
processes rather than focusing solely on site- sustainability” (Society for Ecological Restoration
specific habitat enhancement activities ensures 2002).
that the naturally diverse and dynamic
conditions to which a variety of species are Ecological restoration planning is commonly an
adapted are maintained, and, in the long run, iterative process where initial design concepts
may be the most efficient and cost-effective must be carefully assessed, adjusted, and
course of action. reevaluated through consideration of a variety
of planning elements (Figure 2-1). The planning
This manual, and specifically this chapter, is not
process not only can result in design changes,
intended to replace any existing planning
but in modified goals and objectives based on
guidelines previously adopted by federal
the site information and constraints
agencies, such as the Natural Resource
encountered.
Conservation Service’s Stream Restoration
Design Handbook (NEH 654); U.S. Army Corps of In general, ecological restoration is based on the
Engineers’ (USACE’s) Engineer Regulations particular site—its location, upstream
(ERs) 1105-2-100, 1165-2-501, and 1165-2-100; watershed conditions, and downstream
and U.S. Environmental Protection Agency’s development. However, one of the key planning
Handbook for Developing Watershed Plans to elements will be describing the historical
Restore and Protect our Water (2008) and A changes a site has undergone—documenting
Quick Guide to Developing Watershed Plans to the role of both natural and human disturbance.
Restore and Protect Our Waters (2013). Rather, If possible, a reference reach should be
this manual addresses how the use of large identified, one that has a similar drainage area
wood can be considered in concert with these and valley confinement but has not been
restoration planning processes. Furthermore, subjected to human disturbance. Documenting
the use of additional tools is discussed—such as the reference reach’s ecology will help the
the Project Screening Matrix and River planning team evaluate the extent of the subject
Restoration Analysis Tool (RiverRAT) reach’s problems and establish reasonable
(Skidmore et al. 2011), Structured Decision objectives. However, because reference reaches
Making (SDM), and Multi-Criteria Decision are difficult to identify and quantitatively
Analysis (MCDA)—along with how these tools describe, often conditions reflecting pre-human
can be applied to alluvial ecological restoration disturbance have to be modeled based on
projects. historic records of channel planforms, riparian
floodplain vegetation, flow records, and changes
to the watershed. Depending on the magnitude
2.2 Ecological Restoration and complexity of impacts and project
Process constraints, considerable effort could be
required for data gathering and analysis.
The need for a holistic approach for conducting
restoration activities is well established (Heede

Large Wood National Manual July 2015


2-2
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Issues such as land ownership, local knowledge Within the scientific community, there is
of river restoration, available restoration general agreement on the fundamental
funding, politics, recreation, flow regulation, ecological restoration planning elements that
uncertainty in upstream watershed should be considered in restoring streams
disturbances, and permitting can all influence (Society for Ecological Restoration 2004; EPA
ecological restoration planning. How rivers are 2000). For the purposes of this manual, we
supposed to look and function are ideals that describe 12 principal elements that are typically
vary from one person to another. Philosophies sequenced within three general phases of an
and approaches to river restoration vary within ecological restoration planning framework
the restoration profession starting with the (Figure 2-1). The Project Screening Matrix and
state to which a site should be restored and the River Restoration Analysis Tool provides a good
conditions needed to achieve the desired goals. example of the practical application of these
concepts to river restoration projects
Project designs must also address
(Skidmore et al. 2011). Consideration of these
environmental and ecological factors, as well as
elements will help guide aquatic ecological
satisfy the immediate river restoration need.
restoration projects using large wood.
Rivers in urban areas present unique challenges
for restoration, particularly how existing
infrastructure is protected, either through
improvements or how wood is used.

Figure 2-1. Phases and Considerations Associated With Ecological Restoration Projects Using Large Wood

Large Wood National Manual July 2015


2-3
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

2.2.1 Define the Problem basis for most of the comprehensive restoration
efforts that are currently underway.
and Develop Goals
Many perceived or actual river problems are
2.2.1.1 Defining the Problem associated with the three fundamental types of
Roni (2005) pointed out that most aquatic matter conveyed by streams: water, sediment,
ecosystem restoration efforts are largely in and wood. Changes in flow regimes, or the
response to a whole host of impacts (i.e., supply of sediment and wood, can result in major
problems caused) on aquatic ecosystems that changes to a river. Changes can be a result of
occurred following European settlement of localized river modifications such as a new
North America. Improving the navigation in culvert or bridge crossing, or floodplain
aquatic environments through dredging and modification, or a more system-wide alteration.
snagging (removal of wood) has simplified many They might be due to urbanization that
rivers (Sedell and Froggatt 1984; Collins et al. increased impermeable surface area that
2003). increased runoff, leading to more frequent high
flows, which, in turn, increased sediment
transport capacity and led to channel incision.
CROSS-REFERENCE Biological and ecological impacts are sometimes
associated with other factors such as changes in
As described in Chapter 1, Large Wood Introduction, water chemistry; low-flow regimes; or
forest practices have negatively affected many vegetation on the banks, floodplain, and riparian
streams by increasing fine and coarse sediment, zones.
altering stream hydrology, disrupting delivery of
woody and organic debris, and simplifying habitat
(Salo and Cundy 1987; Murphy 1995). CROSS-REFERENCE

Chapter 3, Ecological and Biological Considerations,


Agricultural activities have had detrimental and Chapter 4, Geomorphology and Hydrology
effects on aquatic ecosystems through dredging, Considerations, assess these factors in detail.
draining, filling, pollution, and channelization of
waterways (NRC 1992). Water diversions for a
large variety of uses have led to altered Bank and meander migration, scour, and
hydrologic patterns, reduced stream flows, deposition are natural river processes that are
higher water temperatures, reduced ability to not necessarily negative themselves and actually
transport sediment, and other deleterious effects contribute to healthy ecosystems. For example,
(Orth 1987; Hill et al. 1991). Mining has resulted cottonwood regeneration in Midwestern rivers is
in the removal of substrates as well as releases of very much dependent on channel migration and
toxic substances (Nelson et al. 1991). Residential new formation of sand bars. Natural meander
development, industrialization, and urbanization migration rates vary across hydrophysiographic
have led to a suite of problems for aquatic areas, so that a particular rate may or may not
ecosystems such as filling and channelization, constitute a problem. In some areas, very small
changes in hydrology from increased impervious rates, perhaps a fraction of a meter per year,
surface area, pollutants from point and nonpoint might signal a problem, while in other areas
sources, elimination of riparian zones, and many meters of movement in a single event
simplification of habitat (Booth 1990). All of might be normal.
these factors have contributed to the
Often, any adjustment is viewed as a problem
degradation of aquatic ecosystems that are the
because it causes an unwanted impact on
anthropogenic land use or structures. People

Large Wood National Guidelines Manual June 2015


2-4
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

tend to view static conditions as desirable when continuing to provide the desired habitat
in reality such situations are rare. This conditions that sustain a healthy ecosystem.
perception is magnified in systems constrained Many mistakes have been made in the past due
by development, and there are economic to the lack of recognition and understanding of
incentives for a static state. Although there are natural disturbance and the consequences of
changes that can signal problems, the problems human actions (Wohl 2013; Reid and Dunne
may have as much to do with the constraints on 1996).
the system as the processes that created the
problem. 2.2.1.2 Develop Goals
Understanding the characteristics and variability The perceived success or failure of many river
of natural processes and how human changes restoration projects can be heavily dependent
affect these processes is critical in all river upon describing the problem and defining
restoration endeavors. Thousands of miles of project goals and objectives. Achieving project
streambanks have been artificially hardened to objectives depends on understanding the
create a static condition that is very unnatural problem and why restoration is needed. Once
and can lead to local impacts that are established, the defined parameters can help
compounded downstream. This is especially true delineate key metrics for success, data collection,
regarding wood. Having the right expertise to assessment methodologies, and finally the design
describe and diagnose a system is essential to itself. Having vague and ambiguous objectives
restoring and managing streams. for the project can lead to problems. Narrowing
the objectives reduces ambiguity for the team
It is important to recognize that short-term
members. Objectives should be specific, realistic,
changes in sediment storage, channel shape, and
achievable, and measurable.
planform are both inevitable and acceptable in
natural channels with unprotected banks. A key Clear ecological objectives that are achievable
to preventing problems or developing self- and that identify the constraints and capabilities
sustaining solutions is to provide the channel of the river and its associated riparian area will
system with adequate space and time for lead to better designs that perform as intended.
adjustment. Define the “geomorphic response Some objectives may, at first glance, appear to be
corridor” and build your restoration plan around realistic, but may need to be redefined if
this. The area encompassing both the 500-year preliminary design information indicates that
flood inundation zone and channel migration the costs will be too high, that intended results
zone provides a good proxy for defining the may not be achievable, or that site constraints
geomorphic response corridor (Rapp and Abbe may significantly alter or preclude
2003; Abbe and Brooks 2011). implementation of the final design. Ecological
objectives should address the maintenance or
The term stability with respect to channels or
rehabilitation of environmental quality by
wood should only be used when clearly defined.
designing and constructing river restoration
Is it morphologic stability in a channel that
projects that have the following traits:
regularly moves its position? Is it spatial stability
where a channel rarely moves? Numerous  Focus on ecosystem function and how wood
factors contribute to channel stability, such as placement will influence hydraulics and
variability of flow regime, fluctuations in wood habitat formation to achieve restoration
and sediment loading, bank materials, and the goals.
influence of riparian vegetation. Effective  Address the needs of endangered and/or
restoration projects build a robust system that imperiled species and their habitats.
can experience variations in flow, sediment, and
wood, and will change through time while

Large Wood National Manual July 2015


2-5
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

 Incorporate wood placements designed to 550 years in order to not limit potential
sustain and accommodate natural processes wood recruitment opportunity.
such as channel migration and still achieve
 Intensive thinning of stands through riparian
restoration goals if the channel moves.
management will likely reduce the
 Incorporate engineered structures that look short-term amount of wood delivered to the
like natural structures. stream.
 Provide desirable river and riparian habitat,  Riparian management objectives developed
including overhanging root cover and large for the purpose of maintaining instream
woody material. wood loads should not focus merely on
stands adjacent to stream reaches in need of
 Maintain or improve water quality.
wood, but on basin-wide riparian areas.
 Are economical to design and build. Restoration project objectives should also
address and/or consider infrastructure
The restoration of riparian areas is a critical
constraints such as the following:
component of restoring natural basin processes
that will establish and maintain natural delivery o Infrastructure that has adversely
of large wood to the watershed. The following affected the river should be replaced in a
riparian management recommendations for manner that sustains natural processes
Puget Sound streams are based on the findings of characteristic of a restored state (e.g.,
Fox (2003) and serve as solid examples of undersized culverts replaced with larger
attainable ecological objectives for riparian crossings to accommodate sediment and
areas. wood transport; levees set back to re-
establish natural unconfined condition).
 Riparian areas should be managed for a
o Infrastructure that does not adversely
diversity of tree species. Managing stand
affect river morphology and processes
attributes to the potentials of each forest
should be protected (e.g., road or levee
zone will promote riparian characteristics
at margin of channel migration zone).
and wood loads assumed to provide
favorable habitat.  Restoration actions should provide
downstream flood benefits by
 Maintaining stem densities and species
diversity along a gradient from the stream o Trapping sediment and wood in
channel will provide heterogeneity in acceptable reaches of the river.
riparian stand characteristics and resemble o Limiting or reversing channel incision to
natural structure. maintain floodplain connectivity.
o Raising the water levels where possible
 Stream buffer widths should consider the to increase flood storage.
potential for disturbances such as debris
o Not increasing flood risks to
flows and snow avalanches, which often infrastructure or developed areas (“no
influence stand attributes for at least rise” to the 100-year flood stage).
20 meters (66 feet) on each side of the o Not increasing erosion risk to
channel but can alter trees beyond 65 meters infrastructure or developed areas.
(213 feet). o Having a desired factor of safety for
 To provide stream channels with the full structural stability and demonstrating a
potential of large wood that riparian areas low risk of failure.
can deliver, riparian stands should use o Not increasing the quantity of woody
management trajectories to at least material moving downstream that could
pose a threat to culverts or bridges.

Large Wood National Manual July 2015


2-6
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

GUIDANCE

Examples of Where Project Objectives and Constraints Are Typically Incompatible

Objective: Allow natural channel migration over time.

Challenge: Road crossing must be maintained where current bridge constricts the river.

Potential Solution: This situation would likely require a wider bridge span that accommodates channel migration
and changes in river planform.

Objective: Add wood material to river channel downstream of a dam to restore supply cut off by an upstream
impoundment.

Challenge: Will wood simply pass through the system or will it benefit habitat restoration? Will mobile wood
threaten downstream bridges?

Potential Solution: This situation could be solved by knowing where added wood will move and ensuring there are
stable wood structures, natural or engineered, in the river downstream of the dam that will trap mobile wood,
create habitat (pools, bars, islands), and not threaten bridges.

Examples of Where Project Objectives Are Mutually Supportive

Objective: Provide instream structure, increase pool frequency and aquatic cover, and protect riparian areas while
they are reforested.

Challenge: Local landowners are concerned about eroding banks.

Potential Solution: In addition to habitat-focused wood placements, include ELJs or complex timber revetments
that protect landowners while also creating instream structure, pools, and cover.

Objective: Protect a buried pipeline exposed by channel incision.

Challenge: Pipeline engineers unfamiliar with fluvial geomorphology are considering reburying the pipeline, but
that will only allow the incision to proceed upstream and further degrade the stream. Engineers are also
concerned about the integrity and longevity of wood. Regulatory agencies require fish passage for any grade
control.

Potential Solution: Use engineered wood placement downstream of the pipeline to restore channel grade and
rebury the pipeline to the desired depth. Wood placements should not be simple weirs, but rather complex broad
structures well buried into the streambed that ensure fish passage and cannot be scoured or flanked by the
stream. Wood burial ensures structural integrity and longevity. The complexity and quantity of wood adds
structural redundancy, increasing the factor of safety. Restoration design can incorporate rock within the buried
wood matrix to further increase the factor of safety if needed.

2.2.2 Assess Site Conditions processes and functions that create habitat
(Beechie and Bolton 1999; Roni et al. 2002). This
An initial watershed or ecosystem assessment of also provides information on opportunities for
current and historical conditions as well as habitat enhancement. The assessment of
disrupted processes is necessary to identify watershed conditions and processes is a critical,
restoration opportunities that are consistent obligatory step in developing an effective
with reestablishing the natural watershed restoration plan.

Large Wood National Manual July 2015


2-7
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

An initial assessment is needed to provide the In the case of wood, the assessment should
process-based framework to define past and address such questions as: how did wood
present watershed dynamics, develop integrated loading and functions change over time, what
solutions, and assess the consequences and was the cause of the change, and how have the
success of past restoration activities. Data changes in wood affected the conditions and
collection and assessment creates the foundation processes that create and sustain habitat?
for analysis and design and is an essential step in
Once the underlying problem is defined,
the design process, whether planning the
opportunities and constraints can be identified
treatment of a single reach or attempting to
at the site. Once the objectives have been
develop a comprehensive plan for an entire
established, measureable metrics for
watershed. This assessment generally includes
determining project success should be defined,
compiling historical photos, records, and data;
such as increasing the number of pools or
conducting preliminary topographic and
decreasing the D50. A basic understanding of the
bathymetric surveys; conducting preliminary
watershed and site-specific conditions is
field investigations of habitat, hydraulic, and
necessary prior to identifying opportunities and
geomorphic conditions; and determining how
constraints and defining risks and uncertainties.
the system has changed and the nature of both
This information requires an initial assessment
natural and anthropogenic disturbances.
of existing information on site conditions.

GUIDANCE

An Assessment Allows You To


 Document the baseline biological conditions.
 Identify the dominant fluvial processes (channel morphology, sediment flux), riparian conditions, and geologic
characteristics of the river system.
 Identify the natural state of the stream and the character, frequency, and magnitude of natural disturbances
and how they influenced the system (e.g., beaver dams, fires, channel avulsions, and large floods).
 Determine the types of human disturbances and how they may have affected the system (e.g., deforestation,
channel straightening, channel clearing, flow impoundment or regulation, and changes in flow and sediment
regimes).
 Determine if there is a problem. If so, is it an anthropogenic problem; a problem associated with the system;
an existing or potential problem associated with past, current, or future land use, floodplain, or riparian zone
changes; or a combination of factors?
 Identify the factors that influence the issues of concern, as well as potential mitigation strategies.

Large Wood National Manual July 2015


2-8
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

CROSS-REFERENCE 2.2.3 Identify Opportunities


and Constraints
As described in Chapter 4, Geomorphology and
Hydrology Considerations, and Chapter 3, Ecological One of the most important tasks in any
and Biological Considerations, fluvial ecological restoration project is identifying opportunities
restoration is an interdisciplinary, comprehensive and constraints. In general, opportunities
effort that focuses not only on reversing past represent situations where the project can best
damages but also on establishing a self-sustaining achieve its goals and deliver additional benefits
ecosystem with conditions that support the fluvial in terms of improving habitat or alleviating
ecosystem processes, function, and structure—
stakeholder concerns. Constraints involve
including the species that depend on them. As such, it
is important to understand the existing limitations the project must accommodate. They
environmental baseline conditions on the site, as well may be spatial (e.g., the need for temporary
as natural or human-made perturbations that may water crossings, ensuring the project has no
have caused them (e.g., altered flow regimes and adverse impact on local property or
infrastructure). infrastructure) or temporal (regulatory work
windows limiting in-water construction).
Knowledge of dominant processes and how However, with a thorough understanding of a
human activity has changed them allows the site’s history and the geomorphic processes
prediction of a system’s response to particular influencing the reach, constraints can be turned
alternatives and how to best achieve project into opportunities. For example, restoring pool
goals within site and stakeholder constraints. It frequency, cover, and channel length and
is imperative to accurately determine the natural providing better riparian conditions in an over-
characteristics of the system, which can be very widened river often involve placement of ELJs to
difficult given not only the magnitude of create stable forested islands and pools with
historical changes, but that the system has complex cover, and increase channel length by
responded to cumulative impacts for over creating anabranching channel patterns. If
100 years. placement of ELJs can be arranged to protect an
eroding bank threatening a road or property, the
The initial assessment provides the foundation project can achieve its goals while also benefiting
for understanding how future changes— the local stakeholder. This type of opportunity
including alternative management, design, or can build community relationships, speed up
mitigation strategies—would affect the system. implementation, and demonstrate how wood is
Solutions can be developed once the assessment not just for fish but also can solve infrastructure
has been completed and will address the goals problems.
and objectives of the project. The solutions might
be self-sustaining, or may require periodic 2.2.3.1 Site Limitations
maintenance, or only be temporary. In some
cases, the best solution might be a river rules Constraints are often associated with limitations
concept that simply provides adequate space for at the site. Access, staging of materials, and
the river to adjust to change. water crossings are all critical issues to contend
with during the design process. Work sites may
Stream evaluations can be performed at varying be limited by a variety of constraints such as
levels of detail. The appropriate level depends on property ownership, state and federal
the status of the study, the perceived significance regulations, and habitat.
of potential problems, the scale of the project,
associated risks, and available resources.

Large Wood National Manual July 2015


2-9
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Depth of Alluvium
GUIDANCE
Many types of wood placements involve partial
Examples of How Wood Influences Fluvial Systems burial or placement of piles, posts, or cribbing,
all of which require an understanding of
1. Channel grade: wood can naturally account for
much of the head loss in channels and bed geotechnical conditions and the depth of
material storage; loss of wood can trigger alluvium. The characteristics and depth of
channel incision, loss of alluvium, and alluvium influence scour depth estimates.
conversion of step-pool and pool-riffle Changes in subsurface conditions, such as the
channels to bedrock and plane-bed channels presence of fine sediments or shallow bedrock,
(e.g., Buffington and Montgomery 1999a–c). can directly influence design and construction
methods. Bedrock canyons are an example of
2. Water surface profile: wood can increase water natural channels where wood doesn’t tend to
surface elevations at local (< channel width) to
accumulate because they act like a log flume. But
reach (>10 channel widths) scales that increase
there are many examples in more unconfined
the spatial extent and duration of inundation,
systems where wood created stable obstructions
enhancing aquatic habitat and floodplain
connectivity (e.g., Brummer et al. 2006). on bedrock that trapped bed material to form an
alluvial channel. These cases depended on large
3. Grain size: stress-partitioning by wood reduces trees capable of crossing the channel and
the median grain size (D50) of bed material; withstanding flows until alluvium built up. Once
reductions in the quantities of functional wood the alluvial channel formed it stabilized more
will increase the D50 (e.g., Manga and Kirchner wood, which contributed to trapping more
2002). sediment. When wood is removed from these
4. Channel complexity: functional wood increases systems they quickly revert to bedrock (Figure
morphologic and textural variability; 2-2). Without large timber, restoring an alluvial
reductions in wood lead to channel channel on bedrock can require substantial
simplification (e.g., Lisle 1995.); Buffington and engineering to replace the function wood once
Montgomery 1999a–c; Abbe and Montgomery had in trapping bed material. This underscores
2003). the importance of function, not material. If
materials other than wood offer the most secure
5. Hyporheic flow: functions 1–4 above all
and inexpensive means of restoring the function
contribute to hyporheic groundwater
of wood that is no longer available or affordable
exchange, which enhances aquatic ecosystem
and water quality (e.g., Poole et al. 2006; (i.e., large trees) and it will facilitate the
Hester et al. 2009). accumulation of wood that would not otherwise
form, then that is the most reasonable
6. Aquatic cover: wood accumulations (i.e., alternative. A no-action alternative that leaves an
logjams) provide complex aquatic cover not unnatural bedrock channel untreated represents
created by any other material (e.g., Abbe and a long-term loss of habitat with consequences
Brooks 2011). upstream and downstream; therefore, if your
7. Pool frequency: functional wood increases pool goal is to restore wood, keep an open mind to
frequency and size distributions (e.g., using a variety of materials, focusing on
Montgomery et al. 1995). restoring function that wood once provided and
creating the conditions that can sustain wood.
8. Aquatic and riparian species habitat: see
Chapter 3, Ecological and Biological
Considerations.

Large Wood National Manual July 2015


2-10
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Figure 2-2. Gravel Patch on Incising Bedrock successfully used to protect infrastructure in
Channel, Rickreall Creek, Oregon ways that enhance, not degrade, aquatic and
riparian habitat (Figure 2-3).

GUIDANCE

Engineered Logjams Have Been Successfully Used to…


 Improve channel alignment and reduce wood
loading at bridge crossings (e.g., Abbe et al.
2003c).
 Raise a streambed and protect buried pipelines
(Abbe et al. 2009).
 Protect roads (Abbe et al. 1997, 2003a, 2003b;
Abbe and Brooks 2011)
Infrastructure
Roads, bridges, culverts, power lines, cable Figure 2-3. Woodward Creek Pipeline Crossing
crossings, and buried pipelines all may be found Wood Placement, Washington
within a project reach. Wood has historically
been viewed as a threat to infrastructure. Mobile
wood certainly can accumulate at bridges and
culverts where blockages can increase risks of
flooding and channel avulsions. But these issues
are primarily due to infrastructure design that
fails to assess and accommodate wood transport.
The supply of mobile wood or flotsam has almost
always been altered by human actions. The
removal of large riparian trees, removal of snags
and logjams, and channel straightening increase
the supply of small mobile wood and the river’s
capacity to transport wood. Thus, in many
watersheds throughout the country, instream
storage of wood has decreased, reducing the Flood Regulations
system’s ability to moderate downstream Many streams are covered under flood
delivery and increase the risks to inadequately regulations, most commonly the federal flood
designed infrastructure. This underscores an insurance program overseen by the Federal
important point of understanding the function of Emergency Management Agency (FEMA) and
wood in restoration: stable wood structures that administered by local and county public works
control hydraulics create habitat, trap wood, and departments. In areas without flood insurance
thus reduce risks to downstream infrastructure. rate maps (FIRMs), there may not be any
For these same reasons, stable wood can trigger regulatory requirements, but in areas with flood
local channel response and increase water insurance, the project may need to provide a
elevations that could affect infrastructure within “zero rise” analysis, demonstrating the project
or upstream of the project reach, highlighting the will not raise the elevation of the 100-year flood
importance of the location and size of wood event. Typically, this analysis requires the
placements. Wood placements have been approval of a licensed professional with

Large Wood National Manual July 2015


2-11
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

expertise in flood modeling and the FEMA safety is through education and by
process (i.e., Certified Floodplain Manager). In demonstrating the importance and value of
systems affected by channel incision since the wood placements. Complex wood placements
last FEMA mapping, water elevations may be have been installed in urban streams, community
raised, but a rise in the regulatory flood surface parks, and many other settings without
would not take place. If restoration results in compromising safety. Early on in the project it is
unavoidable increases in flood elevations but not advisable to bring in a recognized professional
in any damages, it is possible to remap with expertise and experience in wood
regulatory flood zones through a Letter Of Map placement and assessment to present
Change (LOMC), Letter Of Map Amendment information to the public about historical
(LOMA), or Letter of Map Revision (LOMR). impacts and how wood can have positive effects
on and benefits for the community (Figure 2-5).
Acceptable Level of Channel Dynamics Figure 2-5. Public Meeting
Streams are subject to horizontal and vertical
changes in channel position. It is important to
understand how dynamic a channel is prior to
installing wood. A channel can move away from
wood placements, thereby diminishing the
beneficial effects of the wood. The most serious
type of channel movement is at-grade control
structures. If the channel moves around the
wood, it can quickly cut back down and fail to
achieve the desired goals (Figure 2-4). In large
rivers, channel migration may require numerous
wood placements to achieve the desired goals.
Figure 2-4. Eroding River Bank, Nisqually River,
Washington Recreational User Safety
Boating, swimming, and inner tubing are
common recreational activities in many streams.
Wood placements typically impose little risk to
safety during recreational flows but the design
should account for recreational use and should
carefully consider potential hazards the wood
may pose. It is important to document the
current presence and role of wood in the system
as part of assessing recreational safety.
Recreational use in no way precludes the
placement of wood, but designers should be
forthcoming with how particular wood
placements will influence conditions in the river.
Community Safety Some states or local governments may have
specific guidelines or requirements pertaining to
Wood placements should not introduce any
recreational safety, such as placing warning
significant risks to community safety, such as
signs in locations that can be easily seen by
elevating flood levels or creating erosion
boaters floating downstream (Figure 2-6).
hazards. The best way to improve community

Large Wood National Manual July 2015


2-12
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Figure 2-6. A Warning Sign on Wood Placement, changes in riparian vegetation, shorter winters
South Fork Nooksack River, Washington (less snowpack, later onset of snowfall, and
earlier snow melt), and changes in rainfall
intensity are all important considerations.
Evaluating climate change is similar to
considering changes in a watershed that could
influence flow, sediment, and natural wood on
land development, land clearing, road
construction, flow regulation, dams, or dam
removal. It is important to evaluate key project
goals and how those may change under
predicted climate change impacts. However, in
general, adding large wood to streams may be
one of the most important actions available for
building resilience to the impacts of a warming
climate. For example, much of the western
2.2.3.2 Understanding Regulatory United States will be subjected to more severe
Perspectives in Design flood peaks and lower summer base flows as a
Local, state, and federal regulatory requirements result of climate change.
can influence design, cost, and scheduling of any Figure 2-7. Excavation and Dewatering During
restoration project. In many cases, agencies that Construction of an Engineered Logjam in Elwha
permit instream work do not directly regulate River, Washington
the design of large wood placements, their role is
typically limited to determining the influence the
proposed actions may have on the stream's
condition if allowed to proceed. As such, in
addition to knowing which permits are required
and typical processing time (some permits take
months), it is also valuable to understand the
perspectives of individual regulatory personnel
who would be issuing the permits. For example, Pit is 5 meters (16 feet) deep and pumped to
in some districts USACE does not consider wood dewater. The river is flowing from bottom to top at
as fill, while in others it does. Permission to the left side of the photo.
perform in-water construction can also vary Meanwhile, beaver dams have been recognized
significantly depending on local regulations. as a significant means of buffering the impact of
Construction may require temporary channel increasing peak flow magnitudes and
crossings, temporary dams to isolate project diminishing base flows (Beechie et al. 2012).
area, and dewatering (Figure 2-7). Therefore, higher wood loading helps attenuate
higher peak flows (decreasing magnitude and
2.2.3.3 Considering Climate Change increasing duration of hydrograph) and retain
more water in floodplain areas that can better
Restoration projects are increasingly required to
supplement base flow conditions. Wood
consider the potential effects of climate change
restoration can be a critical part of protecting
in the design of large wood installations (Figure
infrastructure that will be subjected to
2-8). The warming climate could result in a
increasing peak flows by securing unstable wood
variety of changes influencing a project. Changes
(i.e., flotsam) found in many watersheds due to
in the magnitude and timing of peak flows,
immature riparian vegetation.

Large Wood National Manual July 2015


2-13
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Figure 2-8. Recession of Honeycomb Glacier in


North Cascades of Washington Is an Example of
2.2.4 Define Risks and
How Warming Climate Affects Hydrology (less Uncertainties
snow, more rain, and more variable flows)

CROSS-REFERENCE

As described in Chapter 7, Risk Considerations,


restoration work has inherent risks and uncertainty,
and it is the responsibility of the design team to
define those aspects of the project that carry
uncertainty and risk. For example, there will be
uncertainty regarding the weather and the timing and
magnitude of flows that affect a project after
construction.

Aspects of design such as revegetation may be


particularly susceptible to floods shortly after
construction before the plants have had the time
to become established.

2.2.4.1 Habitat Recovery


The primary goal of restoration is the recovery
of habitat that has been degraded over time.
Every project involves some uncertainty in
achieving habitat recovery goals. Therefore,
designs should clearly define how they might
From Mauri S. Pelto, North Cascade Glacier Retreat: change the system to achieve restoration goals
http://www.nichols.edu/departments/glacier/bill.
and include quantitative metrics for measuring
htm.
success. As an example, how will increasing the
Increasing the extent of forest cover is typically a number of pools or channel length within a
component of using wood in river restoration, project reach help in achieving recovery goals?
particularly for cases in which wood is being Defining how the project goals and outcomes will
used to restore anabranching channel systems. achieve specific goals helps to understand the
Increasing summer shade and floodplain risks.
wetland together with reducing channel widths
helps reduce water losses due to bed infiltration 2.2.4.2 Infrastructure
and evaporation. Wood also helps attenuate the
adverse impacts of increased sediment supply To reduce risk to infrastructure, existing
resulting from larger magnitude storm events. infrastructure should be clearly mapped, both
Therefore, wood is one of most efficient and overwater structures and subsurface
direct means of preparing watersheds for the infrastructure such as pipelines and other
adverse impacts that have already begun to utilities. Ideally, defining the risk to
occur as a result of climate change. infrastructure would be straightforward; for
example, “the proposed project will not cause
flooding or erosion that threatens structures.” As
stated earlier, well designed wood placements
can not only diminish risks but can be designed

Large Wood National Manual July 2015


2-14
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

to protect infrastructure while also restoring proposed project would have no adverse impacts
habitat. Additional effort may be needed in the on private property.
design phase to ensure existing infrastructure
Property boundaries can influence work areas
will not be exposed to increased risk. The design
and access routes. Ownership of stream- and
team should identify the presence of
riverbeds can depend on local, state, and federal
infrastructure and work closely with public
laws. Most large channels fall under state or
agencies and utility companies that have existing
federal jurisdiction, but it is important to identify
facilities or rights-of-way. As an example, bridge
ownership ahead of time, particularly in systems
crossings in the project reach may benefit from
where stream or river channels move over time.
improved flow and wood conveyance (Figure 2-
Restoration, especially in the case of wood
9).
placement, will benefit from support of the local
Figure 2-9. Bridge Improvements Done to Improve community and landowners. Any project will
Wood Conveyance as Part of a Stream Restoration gain community support when local residents
Project (State Route 7 over Ohop Creek, near
Eatonville, Washington) feel involved, respected, and invested in project
success. For this reason, public meetings and
community education are well worth the time
and investment.

CROSS-REFERENCE

Chapter 8, Regulatory Compliance, Public


Involvement, and Implementation provides an
indepth examination of public meetings and
community education.

2.2.4.4 Public Safety


Streams are inherently dangerous environments.
Accidents happen even in standing water bodies.
Restoration involving wood placements should
clearly describe the context in which the wood is
being placed. For example, many streams
naturally have accumulations of wood and
varying degrees of bank erosion. Trees fall into
rivers during storms, while naturally occurring
bank erosion can create potential safety hazards
that are simply part of the river ecosystem.
Complex wood placements have been used
successfully to protect public facilities (Abbe et
al. 2003a; Abbe and Brooks 2011). Wood
placements should emulate the placement and
2.2.4.3 Private Property conditions that the public would encounter in a
Many projects involve multiple land owners, natural environment. Challenges tend to occur in
including private property owners. The highly altered or managed systems where the
restoration design should demonstrate that the public has grown accustomed to unobstructed
and simplified channels such as those confined

Large Wood National Manual July 2015


2-15
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

by levees and revetments. The best strategy to which defines the type of snag they will form
educate the public on the potential dangers once in a stream or river channel. Designing
associated with restoration is through a public wood structures can be as simple as placing
education campaign. This can be accomplished “key” pieces into a channel to create functional
by working with boating and fishing clubs and wood, or it can be as complex as creating
school programs, and by placing interpretative structures composed of tens to hundreds of logs
signage at boat ramps, parks, and other facilities. arranged in interlocking patterns.

A key piece is one that is likely to create a stable


2.2.5 Develop Design snag and typically has a basal diameter equal to
Considerations or greater than half the bankfull depth of the
river. The presence of a rootwad significantly
increases the stability of a snag as will multiple
CROSS-REFERENCE stems. The most common challenge is the
acquisition and transport of large trees with
As described in Chapter 6, Engineering attached rootwads. When limited to trees
Considerations, there are numerous and excellent smaller than the desired key piece size, stable
publications available regarding design criteria. structures can be designed using a variety of
approaches, such as piles (driven or vibrated
into streambed), posts (placed in excavated or
While every project will have its own unique site
drilled holes and backfilled), cribbing with
conditions, each project should evaluate the
ballast, and rock ballast collars. Wood can also be
stability of the proposed structures relative to
stabilized with other materials such as rock, steel
flow conditions, sediment transport, and bed
piles, and concrete jacks. Regardless of the
scour at the site. Designing wood structures
materials used or structural configuration, the
should include the same design criteria and
design should always try to emulate the function
engineering principles as designing a rock
and appearance of natural wood structures.
revetment or groin, a bridge pier, or a floodwall.
Timber has been successfully used for centuries
to construct spur dikes, flood walls, bulkheads,
2.2.5.2 Wood Structure Stability
bridge abutments, and even dams. Additional For wood to be functional and safe, the designer
analysis would be needed for creating more must create a structure capable of controlling
complex structures, assessing geomorphic and hydraulics and channel morphology. The size
ecological response, and using more variable and shape of natural and engineered wood
natural materials (e.g., range of wood material structures may change over time with the
shapes, sizes, and conditions). accumulation or export of mobile debris, but the
core of the structure must remain stable.
2.2.5.1 Wood Structures
The general public has had negative CROSS-REFERENCE
preconceptions about wood in streams or rivers:
that it floats away or quickly rots. However,
Chapter 6, Engineering Considerations, addresses
wood is a very diverse and extremely versatile wood stability in detail.
material. There are upward of 90,000 species of
trees and woods, with specific gravities ranging
from 0.15 to 1.4 (anything over 1.0 will sink). A log weir that forms a plunge pool is only
Some woods are extremely resistant and can last functional if it remains in place. Stability of wood
for centuries while others will break down in a is simply based on the premise that the forces
few years. Trees vary widely in sizes and shapes, resisting motion must be greater than those

Large Wood National Manual July 2015


2-16
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

acting to move the wood. The forces acting on


GUIDANCE
the wood are drag and buoyancy, both of which
can either move or break the wood. Drag is
proportional to the square of the flow velocity Structural Elements Providing Stability
and cross-sectional area of the wood normal to As previously described, there are key factors
the flow, and a drag coefficient related to the contributing to the stability of wood placements:
structure’s permeability, shape, and size relative
 Size, shape (large rootwad, multiple stems),
to the flow area. Buoyancy is proportional to the effective density, and strength of log(s).
volume of water the wood displaces, which will
be a function of the wood’s specific gravity and  Displacement (submerged volume) of wood.
volume. Resistance is provided by gravitational  Buried surface area of wood and skin friction.
forces acting on the wood. In the case of a large
 Surcharge (gravitational load acting on wood).
snag that is only partially submerged, the
gravitational forces can far exceed the buoyant  Frictional resistance of surface where the wood is
forces, thus holding the snag in place (Abbe and situated.
Montgomery 1996). Once a snag begins to  Longevity/preservation of the wood.
become embedded in alluvium its stability
increases quickly (Abbe et al. 2003b; Abbe and Specific structural elements providing stability
include:
Brooks 2011). Dense or saturated wood with a
specific gravity greater than 1.0 (i.e., density  Key pieces.
greater than 1,000 kilograms per cubic meter)  Piles and posts.
will sink, and with a net downward normal force
it will encounter some frictional resistance from  Ballast (e.g., alluvium or rock).
the streambed. Understanding the physical  Interlocking architecture of the structure.
attributes of snags (shape, stem, and rootwad
 Mechanical attachments (e.g., cable lashing or
size; relative density) that define stability in a
steel pins securing logs to one another).
particular channel can be used to develop
specifications for key pieces needed for  Racked logs and debris that reduce permeability
restoration in relatively undeveloped areas. For and drag coefficient of the structure and prevent
scour from undermining the core of the
many locations it will not be possible to secure
structure.
key pieces, or circumstances simply require
more certainty. In these cases engineered
solutions are available to stabilize wood, such as Pile embedment should be well below estimates
using ballast or piling. The greater the of maximum probable scour. Artificial armor
gravitational force, the more stable the wood, layers such as scour aprons consisting of
just as in the case of a simple gravity dam. This is immobile clasts (more than D90) can be used to
the basic approach to using rock or alluvial limit scour depths. For ELJs, the racked logs that
surcharge to stabilize wood. accumulate on the upstream side of the structure
can be an important element contributing to
Scour is one of the principle reasons for wood
stability. The drag imposed on a logjam is
placement failures and occurs when constricting
dependent on the structure’s shape, size and
or deflecting flow, and with plunging flow for
permeability. When permeability is reduced to a
structures acting as weirs. Proper embedment is
point that the flow going thru the structure
the best strategy for preventing scour damage,
approaches zero, the structure behaves as a bluff
but this can be expensive if it requires extensive
body, and the drag associated with interstitial
excavation.
flow (skin friction and wake interference) is
reduced (e.g., Li and Shen 1973; Shields and

Large Wood National Manual July 2015


2-17
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Alonso 2012). The racked material also defines occurred to the system. The objective of the
the location of scour around the structure. historical analysis is to understand how previous
Racked material creates a buffer around the modifications continue to guide current process
stabilizing core of a structure (e.g., location of and form. Historical analysis identifies the
piles or timber cribbing). The larger the pile of attributes that may be permanently lost from
racked material, the farther the scour hole is those that could and should be recreated, thus
from the core. Vegetation is another factor narrowing the focus for identifying realistic
contributing to long-term stability, adding root restoration options. Historical analysis for wood
cohesion to alluvium in burying the wood and placement projects includes assessing the role
increasing surcharge as trees grow. Trees wood might have had in the river system,
growing on a large wood structure also can including wood sources, rates of delivery, and
increase retention of debris and other racked accumulation and movement of wood in the
material delivered during floods. Eventually the river. The historical wood-loading analysis is
trees protected by the large wood structure will compared to current conditions and used to
mature to create a sustainable source of large develop appropriate strategies for wood
stable wood within the system. placement. Sources of information include air
photographs, maps, surveys, and ground
2.2.6 Conduct Site Surveys photographs. Geographic information system
(GIS) software is used to assess data at multiple
spatial scales to determine historical landscape
CROSS-REFERENCE change. The historical data sources provide input
for the assessment of historical planform
As described in Chapter 3, Ecological and Biological dynamics and mapping of channel migration
Considerations, Chapter 4, Geomorphology and zones to determine the role of wood that
Hydrology Considerations, and Chapter 9, Assessing influenced geomorphic processes, and examines
Ecological Performance, a detailed analysis of site- the extent to which wood was available for
specific conditions is needed to develop the final habitat.
design features of an ecological restoration project.
2.2.6.2 Assess Current Site
Conditions
Comprehensive evaluations of river systems can
require both extensive resources and expertise Although wood placement projects are typically
across a wide range of disciplines. It is important site specific, a site assessment begins at the
to have adequate expertise and to identify and watershed level to determine how existing or
address the most important issues. For example, future land use changes could affect the wood
it is not uncommon for assessments to focus on placement, helping to ensure long-term
hydrology, hydraulics, and biological sustainability. Because wood placement projects
characteristics. While these might be vitally often occur in disturbed river systems, a
important in developing an appropriate solution, watershed assessment is necessary to
the most critical basic information is first-hand understand if channel adjustments at the site are
knowledge of the river system and an a response to local disturbances or indicative of
assessment of the past, current, and future broader watershed-scale alterations. Watershed
equilibrium state of the river system. supplies of sediment, wood, and runoff should be
assessed, at least qualitatively, to determine if
2.2.6.1 Assess Site-Specific History the assumptions used in the planning process
will be sustainable or if the changes in the
A site condition analysis typically begins with an watershed are likely to result in a new condition
investigation of the historical changes that have that is incompatible with planning assumptions.

Large Wood National Manual July 2015


2-18
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Assessing the magnitude, frequency, and are available, and the project team must decide
duration of physical processes gives context to whether to develop a streamflow gaging network
the temporal and spatial scales of system or use hydrologic modeling to develop synthetic
adjustment, the system’s direction of change, and runoff curves for use in developing design flows.
the predicted timeframe for the system to regain For projects where knowledge of groundwater
an energy balance and stability. GIS is a powerful and hyporheic flow is important, piezometers
tool for a watershed-scale assessment. It allows can be installed to measure groundwater flux
for a thorough analysis of physical processes and the interaction of surface and sub-surface
through subbasin delineation, slope calculation, flow.
relating sediment load and caliber with lithology,
Reach assessments are performed in the field at
identification of mass wasting rates, and other
an appropriate scale to describe the physical and
analyses.
ecological conditions of the channel and
The focus of the current conditions assessment floodplain. The physical reach assessments
can then be scaled down from the watershed to typically include field mapping and topographic
the river reach and local site. Streams are open surveying components focused on collecting
systems and are continually adjusting their form information related to channel morphology,
to altering energy inputs and materials. It is sediment transport, and geotechnical issues that
important to understand how adjustments in could affect the project. The texture of alluvium
flow, sediment supply, boundary sediment and substrate forming the channel bed, bars,
texture and cohesion, large wood inputs, and banks, and floodplain is mapped as part of facies
riparian vegetation interrelate; and how they units or quantitatively measured and then
collectively determine channel form and habitat. interpreted to understand sediment transport
This understanding is critical for predicting dynamics. Channel bed and bank parameters
channel response and developing sustainable that contribute to habitat or that can be evidence
wood placement designs. of stability or instability are assessed, and field
estimates of appropriate roughness coefficients
A solid understanding of river hydrology is
needed for future hydraulic modeling are made.
critical because knowledge of the streamflow
Geologic controls are observed, and the
regime is integral to nearly every aspect of wood
geotechnical properties of streambanks and
placement objectives. Because hydrologic
other landforms can be rapidly assessed in the
information is needed to determine base flow
field or studied more quantitatively using bank
conditions and provide input for hydraulic
stability modeling and other tools. Features that
modeling, design calculations, habitat modeling,
indicate previous or potential channel dynamics,
and flood risk assessment, hydrologic analysis
such as levees, side channels, crevasse channels,
should be one of the first studies performed.
and riparian buffer widths, are measured. Reach
Ideally, an active river gage exists with many data is also collected on any evidence of
years of data from which statistical analyses of disturbances, perturbations, or hazards, such as
flow records can be performed. This information bridges, diversions, beaver dams, landslides, and
helps calculate flood-frequency return intervals nearby infrastructure or property potentially at
and flow duration curves, and provides an risk. The field mapping of infrastructure is
understanding of the timing and movement of compared with available infrastructure maps to
water through the watershed and trends in verify the presence and location of utilities,
runoff related to land use changes. If a nearby pipes, and property lines. Measurements and
river gage is not available, then gages elsewhere observations made during the reach analysis are
in the watershed can be analyzed and scaled used to refine channel migration zone mapping
accordingly to provide a surrogate flow record and identify zones where the channel could and
for the site. Often times, however, no gage data should be allowed to migrate, versus areas

Large Wood National Manual July 2015


2-19
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

where migration, erosion, and inundation are not measurements of sediment transport are made
likely or desired. to calibrate and verify model calculations.

Detailed topographic surveys are performed to Riparian restoration and management are
provide the elevations needed for hydraulic and critical to the success of any wood project,
sediment transport modeling. The field-surveyed particularly in the long term. Trees help stabilize
elevations are often combined with Light banks and provide a future source of wood to the
Detection and Ranging (LiDAR) or channel. For systems with migrating channels or
photogrammetry-acquired elevations to create a that are subject to avulsions, future channel
composite elevation surface for the site with response would largely depend on riparian
continuous coverage of ground and bathymetric conditions. Planting plans should incorporate
surfaces. multispecies combinations that provide short-
and long-term benefits. Willow and cottonwood
Many wood placement projects occur in river
can provide short-term benefits because they are
reaches where the channel has become
fast growing, but plans should include
hydrologically disconnected from its floodplain
successional climax species that provide the
(such as from incision, leveeing, or channel
large, long-lived trees needed along the
pattern simplification from an anabranching
streambanks.
pattern into single thread channel), and a project
objective may be using wood to increase the Wood placement projects are designed to be
frequency of floodplain inundation. A current compatible with and augment the prevailing
conditions assessment usually includes studies processes that sustain the current channel
to determine the presence of a floodplain and morphology, or are designed to substantially
how current floodplain connectivity compares to alter processes that will lead to new desired
prior conditions. Height above the water surface channel morphology, typically with increased
analysis is often performed to map all elevations complexity and habitat structure. Therefore, the
at the site relative to the river’s water surface data sources and analyses described above are
elevation to show landform elevations in relation used to better understand the site’s alluvial
to the river channel. Hydraulic modeling is channel behavior and dynamic equilibrium
performed to quantify the flow magnitude energy states. Before designing a wood
required for water to spill out onto the placement project, it is necessary to know if the
floodplain. Both analyses are important for observed channel morphology reflects natural
identifying opportunities and constraints for scour and fill events or if there has been more
increasing floodplain inundation. long-term aggradation and degradation
indicative of channel adjustment response to
In addition to calculating water surface
disturbance. These assessments determine if the
elevations needed for evaluating floodplain
observed channel pattern (e.g., straight, sinuous,
inundation and current flood risk, one-
anabranching, braided) will be sustained by
dimensional (1D) or two-dimensional (2D)
prevailing processes and the extent to which
hydraulic modeling is performed to simulate
wood placement could accomplish desired
velocity vectors, shear stresses, and flow depths
morphologic change.
in the channel prior to proposed wood
placement. The hydraulic models can be coupled
2.2.6.3 Assess Future Site
with sediment transport calculations, either
within the model itself in a mobile bed analysis
Conditions
or in user-created spreadsheets, to evaluate Assessing future conditions is necessary if the
sediment flux in the reach. This type of wood placement project is expected to be self-
evaluation is not commonly done because of sustaining and resilient. Many of the analyses
time and budget constraints. Typically field undertaken in the historical and current

Large Wood National Manual July 2015


2-20
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

conditions analysis provide the information presented graphically and descriptively,


needed to assess the trajectory of the system and explaining how they would achieve project goals
determine if future conditions will support the and their associated risks and costs. Clear
project goals. The evaluation of watershed land metrics for success should be established for
use trends and associated channel morphologic evaluating alternatives that are agreed upon by
response is used to predict future change using decision makers and stakeholders so that
appropriate channel evolution models and alternatives are assessed as objectively as
knowledge of alluvial channel response to possible. Traditional metrics include project cost,
perturbation. Numerical physical and habitat stability factor of safety, and design life. In
modeling tools can also be used to evaluate addition to these, wood projects should consider
certain scenarios and test how the wood metrics that measure habitat enhancement, such
placement would perform. Considerations as increases in channel length, increases in pool
should be made on how riparian conditions frequency and cover, or increases in inundated
would change and how they may influence the area. Other metrics could include channel
project, in both positive and negative ways. This response associated with each alternative, or the
is especially true in the restoration of urban relative uncertainty in each alternative. Project
creek corridors where restoration of riparian maintenance is another important consideration:
areas leads to future wood recruitment that will the project require maintenance and, if so,
could raise water levels or create concerns for what kind and to what extent? Other metrics
downstream culverts. Beavers and their should cover how well the alternative works
influence on the channel should also be within the project constraints, such as flooding
considered in riparian planting plans. Beavers risks or risks to infrastructure. For more
are coming back to many areas in the United complex and costly projects, it may be valuable
States, including urban streams. to weight assessment metrics and conduct a
quantitative benefit-cost assessment.
Climate change predictions also must be factored
into the design. Although the exact implications
of climate change on streamflow characteristics 2.2.8 Prepare Monitoring
are not certain, existing climate change models and Adaptive Management
that predict regional watershed response are
available for review. This information provides
Plan
estimates about runoff patterns and sediment Monitoring of wood placement project metrics is
loads and helps determine adjustments that may necessary for assessing whether project goals
be needed to adequately design the structures. and objectives are being achieved and to
determine which adaptive management actions
2.2.7 Prepare and Evaluate must be implemented.

Alternative Restoration
Concepts CROSS-REFERENCE

Every restoration project should include a set of Chapter 9, Assessing Ecological Performance, provides
alternatives, the first of which is a “no-action” a detailed discussion about adaptive management.
scenario. Typically there would be at least two
additional alternatives that vary in magnitude or
in the manner in which they address the Monitoring not only evaluates project successes
problem; these additional alternatives also or failures, it also enables appropriate post-
would achieve the desired goals within the project adjustments to be made, and is critical
constraints of the project. Alternatives are for expanding scientific knowledge of physical

Large Wood National Manual July 2015


2-21
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

processes that allow future restoration projects monitoring surveys. Monitoring is recommended
to be more advanced and successful. The at least once per year for the first 5 years, with
monitoring elements should be developed from reduced monitoring frequency thereafter (e.g.,
the project’s objectives and evaluation criteria. once every 2 to 3 years until year 10).
As discussed in the last section, clear metrics
During the design phase and prior to
should be established with stakeholders that can
construction, the project team and stakeholders
be measured through time to assess how well a
should develop criteria for adaptive
project is performing. Metrics should be realistic
management. Clear threshold criteria should be
and affordable. Simple presence or absence
agreed upon so that necessary remedial actions
monitoring can be valuable for wood placements
can be taken quickly if monitoring shows that
that simply require noting global positioning
anticipated performance of the wood placement
system (GPS) coordinates through time and
features is not meeting expectations. For
tracking flow events.
example, threshold criteria could document that
A monitoring plan is typically prepared after the additional wood will be added or additional
alternatives development and evaluation stage, anchoring implemented if a certain percentage of
and must incorporate the previously defined the structure is lost, or additional plantings will
risks and uncertainties. be added if survival rates of revegetation are too
low. Additionally, flood risk criteria can
Physical attribute monitoring for wood
document that any racked debris will be
placement projects typically includes
removed once a certain volume is accumulated.
measurements to assess structural integrity and
The adaptive management plan must document
the level to which a feature is providing desired
how decisions will be made regarding whether
functionality. Measurements can include
or not maintenance work is needed, how the
identifying the number of wood pieces from the
work will be paid for, and who will perform the
original structure that are still intact,
work. The adaptive management plan must also
accumulated debris, scour and fill associated
specify how monitoring and maintenance work
with the structure, sediment sorting and
will be reported and the timing for delivering
diversification, channel planform change, and
reports to the stakeholders.
hydraulic attributes. Natural wood accumulation
is common for large wood placements and can
provide an idea of how structures change 2.2.9 Prepare Detailed
through time; it can also provide information on Design Plans
the amount of wood that is naturally moving
through the system. Ecological monitoring often Detailed design drawings serve multiple
includes surveys of adult and juvenile fish; purposes for a large wood project. Initially
temperature monitoring; and assessment of concept design plans are developed to convey
depths, cover, shading, and other important the major elements of one or more alternatives
habitat elements. that would achieve the project objectives. Often
design plans are put forth as a percentage of
The frequency of monitoring and total period completion to give stakeholders a general idea of
over which monitoring occurs is largely design development status. When the design
dependent on project budgets and availability of team is ready to lay out basic concept ideas on
staff or volunteers to perform the monitoring paper they are typically referred to as a 10%
work. Immediately following construction, an as- design. Concept plans include enough detail to
built survey with established photo points show the general layout the project would have
should be established to document that the wood on the landscape and principal construction
placements were built as designed and to activities such as excavation or grading, instream
provide a comparison condition for future construction, relative size of structures, and how

Large Wood National Manual July 2015


2-22
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

the site would look after construction. A 30% the detailed design plans. Although writing of
design refers to the stage when a preferred environmental documents and permit
concept alternative is selected and incorporated applications usually begins at the 30 to 60%
into the engineering plan sheets design set, design level, early and continual involvement of
referring to the relative percentage of a complete the regulatory agencies in the review and
design package. This initial design set is often advancement of the entire planning and design
presented to stakeholders, landowners, and process will greatly increase the likelihood of the
regulatory agencies to get initial approval and wood placement project receiving timely
begin the permitting process. Final permits environmental clearance with necessary permits.
typically require a more detailed design set, such Generally, design is not regulated by
as 60% that includes all construction aspects of environmental review agencies but rather
the project such as water crossings, erosion indirectly regulated by professional licensure
control, access and staging, structure details requirements of due diligence. As such,
(e.g., pile driving, pumping), traffic management, representatives from the regulatory agencies are
and other actions that could affect fish and often invited to be part of the project
wildlife and local communities. In the final stages stakeholder team or technical advisory group to
of the project, designs serve as the basis for cost allow them to give feedback on project
estimates and material acquisition, and they opportunities and constraints, and have a vested
provide contractors the necessary information to interest in the project’s success. Their early input
construct the project. A final design package not will help identify future obstacles that could
only includes detailed spatial plans (map or plan derail the project entirely or necessitate costly
view, cross-sections, and profiles) but details of design changes. During the design stage, an open
subsurface conditions, water, and specifications. dialogue must be maintained with all regulatory
The design engineer prepares a cost estimate agencies.
based on the plans. The completed package is Project schedules must allow ample time for
referred to as plans, specifications, and estimates agency review and comments, and be flexible to
(PS&E). A 90% design usually represents a allow for design changes that will arise based on
completed design package ready for final review. agency feedback and permitting needs. The
design and permitting process should be
Given the variability of fluvial environments,
considered an iterative stage of the project
flow and groundwater conditions, nonuniform
because neither component can be successfully
characteristics of large wood, and other common
completed without direct interaction with the
materials, the detailed design plans should build
other. Many wood placement projects have
in flexibility for field adjustments during the
encountered serious challenges by advancing the
construction process. The level of detail
design too far before introducing it to the
provided should be tailored to the specific design
regulatory agencies and asking for their
phase and intent of that phase.
approval.

2.2.10 Complete
Environmental and Regulatory
Requirements
Environmental compliance and documentation is
performed in tandem with the development of

Large Wood National Manual July 2015


2-23
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

GUIDANCE

Common Design Phases and Additional Information Provided


 30% Conceptual Design Level. The primary purpose of the 30% design is to convey the concept and objectives
to key stakeholders. It includes enough information to convey the general intent.
 60% Preliminary Design Level. The primary purpose of the 60% design level is to support project permits and
develop a planning-level cost estimate. It includes key design information (number of structures, sections,
profiles, typical details) to evaluate construction impacts from the project. Many large wood projects proceed
to construction at this stage using engineer-led field construction. The viability of this method is dependent
on the experience of the owner, designer, and contractor and findings from the risk assessment. The 60%
plans are typically used for permitting because no significant changes in design are expected from this point
forward.
 90% Final Design Level. The primary purpose of the 90% design set is to develop an accurate engineer’s cost
estimate and finalize design details and specifications to ensure constructability and compliance with project
permits. Technical specifications are often included in this phase to accompany the design plans and provide
additional details to project sponsors.
 100% Bid-Ready Package. This design level includes all pertinent information for a contractor to develop an
accurate construction estimate and construct the project elements.

Technical specifications from the 90% design level are expanded into full contractual documents. Through the
design process additional information is often provided to the project sponsor and stakeholders:.
 Selection Criteria. Selection criteria for the contractors (e.g., restoration experience and required equipment).
 Quality Control/Quality Assurance (QA/QC) and Liability. Beginning with initial concepts, it is expected that
experts are reviewing and signing off on designs, including licensed professionals who stamp the designs. Any
plans focused on restoration of ecological communities should include reviews by appropriate fisheries,
wildlife, and plant experts prior to submitting design plans for permitting. Most construction design plans
require the stamp of a civil engineer with expertise in stream restoration. A licensed geologist with expertise
in fluvial or coastal geomorphology should also stamp any plans that involve alteration of the landscape and
natural processes such as surface and subsurface flow, erosion, and sedimentation. Some plan sheets may also
require stamping by a land surveyor, landscape architect, or structural engineer.
Construction Checklists and Inspection Reports. Preparing construction checklists helps ensure projects are
successfully completed:
o Proper safety equipment is used and procedures are followed.
o Permit conditions are met throughout construction.
o Structure locations are accurately staked.
o Builder meets specifications in the plans for materials and completed structures.
o Contractor invoices reflect materials and work completed.
o Photo points documenting before and after conditions are established.

be implemented, construction techniques, and


Most wood placement projects will require
fish and wildlife exclusion and avoidance
instream work and the need for dewatering or
measures taken before permits are issued.
site isolation to contain turbidity and limit
Environmental compliance will often require a
disruption to fish, wildlife, and vegetation
risk level assessment and a site-specific
communities. Detailed plans will likely be
stormwater pollution prevention plan (SWPPP)
required by the regulatory agencies that identify
that identifies an effective combination of
the best management practices (BMPs) that will
erosion control, sediment control, and non-

Large Wood National Manual July 2015


2-24
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

stormwater BMPs that must be implemented to consideration should include construction


reduce construction effects on receiving water management and oversight.
quality. The SWPPP must define a program of
regular inspections of the BMPs and in some 2.2.11.1 Construction Management
cases sampling of water quality parameters. The
SWPPP also should demonstrate compliance Prebid Meeting
with all applicable local and regional erosion and
This meeting should be required for all potential
sediment control standards, identification of
bidders to increase the odds of receiving
responsible parties, a detailed construction
qualified bids and reduce the potential for
timeline, and a BMP monitoring and
change orders.
maintenance schedule.

Wood placement projects are often designed Contractor Selection


with the intention of altering channel hydraulics
Due to the inherent complexity and uncertainty
to promote beneficial physical processes and
of site conditions and wood materials in many
enhance habitat conditions. The extent to which
restoration projects, it is highly advised that
the wood placement project creates a large
contractor selection not be limited to low bids,
enough flow obstruction to appreciably increase
but be based on contractor experience,
water surface elevations, and thus potentially
qualifications, and construction approach. This
increase flood risk, depends on the scale of the
will increase the odds of receiving qualified bids
project and wood placement locations. Larger
and reduce the potential for quality control
projects with the potential to elevate water
issues and change orders.
surface elevations and increase flood risk in the
regulatory floodway may be required by the
local jurisdiction and FEMA to document with a
2.2.11.2 Construction Oversight
hydraulic modeling study, approved by a Large wood is not uniform and is available in a
licensed engineer, that the project will not create variety of quality grades. An initial task for
a net-rise in the 100-year flood base elevations. project sponsors and designers should always be
If modeling shows the wood placement project to inspect the large wood delivered to the site to
will create a net-rise, then the project team must ensure it meets the species, size, and quality
advocate with agencies and neighbors to specifications provided in the bid documents.
approve the net-rise because the base level Any pieces not meeting the specifications should
elevations used to assess flood risk will increase. be tagged and removed from the site
This will typically require completion of immediately. Extreme care should also be taken
Conditional LOMR and LOMR documents. in working around wood placements as the size
of pieces and ability of heavy equipment to move
2.2.11 Implement the Project wood can create safety hazards for unaware
laborers and observers. Safety protocols should
The project implementation stage is the most be clearly established and followed on job sites.
critical component to a successful large wood
project and achieving the desired project Daily Logs
outcomes. Despite all the planning and design
efforts to get to this stage, if a project encounters Daily logs should be kept by the sponsor and/or
challenges during construction that compromise designated field engineer/scientist when on site
the design intent or intended outcomes, the to document field conditions, construction
project may be viewed as a failure, making it progress, compliance with design plans, and
difficult to pursue future large wood projects in conversations with the contractor. Daily logs
that watershed. Critical implementation should include photos and plan mark-ups

Large Wood National Manual July 2015


2-25
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

documenting whether the project is being of the key project elements should be approved
constructed as designed or if changes are and in place as preparation of the plan has
needed. These logs can serve as key information already been completed. As discussed above, the
to the sponsor and/or designer during any monitoring and adaptive management plan will
disputes and can limit the potential for have clear criteria stating which elements will be
unnecessary change orders that increase costs. monitored, the frequency of monitoring, and
whether performance standards have been met.
Change As the monitoring is implemented and reports
are written, the stakeholders will use the
Some amount of change from the design plans
approved adaptive management plan to
during the construction process should be
determine any remedial work that must be
expected on every large wood project for the
performed. The length of monitoring and
reasons previously mentioned. Projects where
adaptive management will vary between
some amount of uncertainty is anticipated
projects based on budget constraints, but the
should build that expectation into the design
longer the monitoring periods the greater the
plans and contract documents, and discussions
probability the project will achieve its objectives.
should be held with contractors to reduce the
potential for costly change orders during the
construction process. Common strategies to add 2.3 Restoration Decision
flexibility in contract documents include bidding
items lump sum and creating force account items Making
for miscellaneous items (e.g., setting up bid item
At each step in the ecological restoration
for contractor to lock into cost of machine and
planning process critical decisions need to be
operator time).
made that will influence the outcome of the
project. Historically, even well-intentioned
2.2.12 Monitor and resource management has resulted in degraded
Implement Adaptive ecosystems across the United States (Noss and
Peters 1995). To address the past impacts as
Management Measures well as prevent potential future degradation of
Once environmental documentation is approved, ecosystems, habitat enhancement or
permits are received, and construction is improvement projects have become the
completed, the monitoring and adaptive mitigation action of choice to offset many habitat
management phase of the project begins. deficiencies. Unfortunately, restoration projects
are often planned and implemented without
proper consideration of their landscape context
CROSS-REFERENCE as well as the ecosystem processes and structure
(Beechie et al. 2010). Failure to recognize these
Chapter 9, Assessing Ecological Performance, provides broader scale concerns may lead to poor project
a detailed discussion about monitoring and
selection and increased potential for failure.
implementing adaptive management measures.
The integration of socioeconomic factors into
restoration plans has been a critical component
of successful programs. To this end, over the last
Although completion of the environmental
20 years, researchers and resource managers
documentation and permitting process may
have emphasized the use of decision support
introduce new requirements that necessitate
tools for implementing ecosystem restoration
modification of the monitoring and adaptive
efforts (Wyant et al. 1995; EPA 1995; Pastoroka
management plan, by this stage in the project, all
et al. 1997; Linkov et al. 2005; Linkov and

Large Wood National Manual July 2015


2-26
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Moberg 2012; Suding and Hobbs 2009; Beechie many of whom should have experience and
et al. 2010; USACE IWR 2010; Gregory et al. expertise in the use of wood in restoration
2012; Convertino et al. 2013; IUCN 2014). This projects.
section describes the importance of the project
planning team and scaling project size, along
GUIDANCE
with a number of the tools that have been
developed to assist in quantifying and
understanding the resource as well as the Professionals and Experts to Involve
socioeconomic tradeoffs associated with those
 Geologists with specific expertise in fluvial
decisions. Specifically, it highlights the value of
geomorphology, hydrology, sediment transport,
taking an SDM approach that considers MCDA and wood.
tools in ecological restoration planning for
improving the manner in which restoration  Engineers with expertise in river system design
and construction of wood structures.
decisions are made as well as the success of
restoration projects.  Hydrologists with specific expertise in flow
characterization and fluvial geomorphology.
2.3.1 Planning Team  Fisheries biologists and/or aquatic ecologists with
regional expertise.
Composition
 Riparian plant ecologists or foresters with
A strong multidisciplinary team is necessary regional expertise.
given the amount and range of infrastructure,
 Regulatory specialists with local and regional
public and private stakeholder involvement,
expertise.
regulatory issues, and potential liability.
Professional engineers should be aware that  Wetland scientists.
foundational due diligence includes obtaining  Landscape architects.
substantial information from other professional
disciplines. Designs compiled by an individual or  Resource economists.
a limited team may represent a breach of the  Community facilitators, planners, or watershed
ethical canon, “Engineers shall perform services coordinators.
only in areas of their competence.” The project
team should comprise more than those just
working on technical aspects of the design. As It is crucial to include the stakeholders
appropriate, it should also include throughout the ecological restoration planning
representatives from key regulatory agencies, efforts. Stakeholders are often the individuals or
landowners, or local municipalities. groups who may fund the project, affect the river
Understanding local knowledge and politics directly, or be affected by actions taken on the
provides value in defining project constraints, river. A trained facilitator may be needed to
speeding up stakeholder approval, and guide the development of goals and objectives
ultimately reducing costs. and to ensure that all stakeholders, challenges,
other opportunities, and constraints are fully
While the exact makeup of the project team can recognized. Once agreement is reached on the
vary, it should typically include engineering goals and objectives, the team can start on the
professionals, geomorphological professionals, design process and develop design alternatives.
landscape architects, and experts as well as
individuals with a variety of ecological expertise. 2.3.2 Scaling the Process
As such, the design team should consider the
The scale of ecological restoration projects can
experts listed in the Guidance box that follows,
range from simply stabilizing a streambank at a

Large Wood National Manual July 2015


2-27
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

specific location to restoring self-sustaining 2.3.3 Integrating


ecological functions to an entire watershed.
Restoration projects vary in scale from an entire Socioeconomics into the
watershed to a valley reach to a small section of Restoration Process
channel. Regardless of size and focus, every
restoration undertaking should consider the Large wood projects result in both costs and
cumulative impacts of every project element and benefits to private resources owners and to the
how the project would not only achieve the wider public. Consideration of these costs and
desired goals, but sustain those conditions. benefits is therefore a crucial component of
restoration planning and will influence large
The larger and more complex the project, the wood project decision making. Comparing the
greater the effort will be required for design and costs and benefits of large wood projects is
permitting. Projects with greater visibility challenging because of the differing manner in
typically require more detailed design and which costs and benefits of projects are realized.
review. A project design can be directly affected Costs are mostly centralized and incur to those
by a wide range of site conditions such as implementing the project. Benefits, on the other
infrastructure (above and below ground), earth hand, are more diffuse in nature for several
materials, threatened and endangered species, reasons. First, benefits of large wood projects
archaeological and cultural resources, land accrue to a much larger segment of society than
ownership, or wetland delineations. Site the project costs. Second, while some benefits of
conditions should be assessed early on in the large wood projects, such as increased fish
project to guide how the greatest ecological populations and increased recreational
uplift can be accomplished within site opportunities, may be measured by market
constraints—or how the project can redefine or activity, other benefits of large wood projects,
eliminate constraints by improving such as many ecosystem services, are less
infrastructure, securing easements, or avoiding directly tied to or have no ties to market activity.
sensitive areas. These benefits require methods other than
analyzing market data to estimate their value.
The scale of restoration also pertains to how
This section provides a brief introduction to the
much intervention is needed. Passive restoration
costs and benefits of restoration planning. In
focuses on allowing natural processes to proceed
relation to restoration planning costs, this
without human intervention. This strategy
section describes the major factors that influence
would focus on making sure a system is capable
project costs. In considering restoration planning
of re-establishing the desired conditions in a
benefits, the section describes a framework for
reasonable timeframe if left alone, and “design”
estimating the benefits of restoration projects,
would focus on new management guidelines that
the types of values that result from projects, and
encourage and protect natural recovery.
common methods used to estimate the monetary
Examples include systems with mature riparian
value of these benefits.
vegetation where wood is naturally being
recruited. Many other sites will require direct
intervention. As such, scaling the process to fit
2.3.3.1 Costs of Restoration
the size and complexity of the project is The costs of restoration activities are an
essential. important factor that can influence the scale and
scope of projects, and help determine which
projects are undertaken. It is important to
consider project costs in the early stages of
planning, as costs drive project decisions. As
noted by Evergreen Funding Consultants (EFC

Large Wood National Manual July 2015


2-28
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

2003), this early consideration of costs helps to Cost estimates for large wood projects are
ensure that project plans present realistic commonly estimated on a scale of dollars per
descriptions of associated costs and thus stream mile. EFC prepared a primer on various
increases the likelihood that funding will be types of restoration project costs for the Puget
available for proposed actions. In addition, a Sound Shared Strategy (2003), including large
careful consideration of costs at the planning wood projects. This primer can be used to obtain
stage helps to ensure that the funding sources rough estimates of project costs, based on
will be available when they are needed over the different factors that could influence costs. As
course of the project. noted by EFC in the primer, the level of
predictability of costs for large wood projects is
The costs of large wood projects can be broken
generally good because there is a large amount
down into several categories, a helpful tool for
of certainty as to how the main factors of large
estimating project costs, because various factors
wood projects affect costs.
will affect these cost categories in different ways.
Understanding the categories likely to affect In addition to the factors discussed above, EFC
restoration costs, and how these categories mentioned two other factors that can affect large
would be affected by the specific aspects of a wood project costs. First, risks associated with
project, increases the ability of planners to the project can increase costs. Second, risks can
accurately estimate project costs. result from hazards that may be introduced by
large wood projects, such as trapping
recreational river users, jamming downstream
GUIDANCE culverts, and changing channel and floodplain
characteristics (which can increase erosion and
Major Cost Categories for Restoration Projects flood risks). Risks of large wood projects are
 Construction higher in more heavily populated areas, such as
projects on streams that traverse urban or
 Design
suburban areas. Risks related to large wood
 Permitting projects can be mitigated through design and
 Appraisal planning, but usually at an additional cost. The
remoteness of a project site can be a good
 Basic monitoring indication of risk-related costs, with more
 Routine maintenance remote sites posing fewer risks.
 Reestablishing the site to prior conditions
 Project management
 General administration and enforcement
 Longer-term monitoring and maintenance

By considering each of these categories separately,


one can assess how different factors of a given
restoration project would likely affect each cost
category. (EFC 2003)

Large Wood National Manual July 2015


2-29
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

CAVEAT

Factors Likely to Affect Restoration Planning Costs

Project scope influences project costs in that larger projects tend to have smaller costs on a per-stream-mile basis.
This is due to economies of scale that come into play when a larger stream area is the focus of the restoration
activity. In other words, the fixed costs of restoration projects can be spread out for larger projects, reducing the
implementation costs as compared to fixed costs incurred by a project.

Treatment intensity of the large wood project also affects costs, with more intensive levels of restoration having
higher costs. This is because higher levels of treatment intensity generally require more materials, equipment, and
labor than projects with a lower level of treatment intensity. One way that treatment intensity can be measured is
by the density of wood used in the project. Large wood projects with less wood density have smaller materials
costs than projects with greater levels of wood density.

Stream size is one of the biggest factors affecting restoration project costs, with larger streams generally having
higher project costs than smaller streams. The reason for this positive correlation between stream size and project
costs is that large wood projects on larger streams generally require more planning, design, materials, permitting,
equipment, and labor than projects on smaller streams.

Access can have a significant influence on restoration project costs because the ease of access will determine the
type of equipment and the amount of labor needed. For example, some large wood projects may require the use
of helicopters to get the material to the project site, which increases costs over projects where material can be
directly hauled to the site by ground transport.

Material availability can affect costs based on whether materials are purchased or obtained through some other
means. If the materials are purchased, the quantity and quality of timber that is acquired for the project would
have a direct impact on the resulting project costs.

Contract type can influence project costs based on whether labor and equipment are rented by the hour or based
on some other type of arrangement. Other variations in contract type include whether the contract is for
construction or for equipment rental. Construction contracts are generally more expensive than equipment rental
contracts, but the arrangements for liability are different between these two contract types, which also affects
costs. For construction contracts, the contractor usually assumes the liability, whereas liability is not assumed for
equipment rental contracts.

Amount of time needed for the project affects costs because longer projects require more labor hours. Time may
also be needed to acquire necessary permits.

2.3.3.2 Economic Benefits of Missing or incomplete data and uncertainties


Restoration about the cause-and-effect relationships
between restoration and improvements create
Although project costs are often relatively challenges when quantifying changes to
straightforward to estimate, benefits of ecosystem services. Additionally, the process of
restoration projects are more challenging and monetizing these changes is often challenging
often require specialized estimation techniques. because it is difficult to identify the actual values
Estimating the benefits of restoration projects that society places on ecosystem services.
usually starts with a process of quantifying the
change in ecosystems, and the services provided The following discussion identifies an overall
by them, that would result from a project. Next, framework and specific steps for estimating the
monetary values are ascribed to these changes. benefits of restoration projects. This is followed

Large Wood National Manual July 2015


2-30
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

by descriptions of the potential benefits of helps to avoid the double-counting of ecosystem


restoration planning. Lastly, the methods and services and thus artificially inflating their value.
approaches used to value the ecosystem service
benefits of restoration projects are discussed.
GUIDANCE
Framework for Valuing the Benefits of
Restoration Projects EPA’s Framework Steps for Valuing the
Benefits of Restoration Projects
A framework for valuing the benefits of
restoration planning involves a systematic and 1. Identify management actions and ecosystem
scientific process of quantifying ecological services of strategic importance.
impacts and ascribing monetary values to them. a. Identify ecosystem production functions
The U.S. Environmental Protection Agency’s and restoration activities that affect
Science Advisory Board (2009) developed such a ecosystem services.
framework in a report titled Valuing the
b. Of the affected ecosystem services,
Protection of Ecological Systems and Services.
identify the ones that directly affect the
This framework consists of three steps, as shown
well-being of society and are of the
in the Guidance box that follows. greatest importance to society.
In Step 1 of the framework, the scope of the 2. Quantify final ecosystem services.
benefit estimation process is narrowed down to
only focus on ecosystem services that would be a. Develop ecological production functions
affected by the proposed restoration project. that specify the cause and effect
relationships among restoration activities
Restricting this focus should not be interpreted
and ecosystem services.
as deducing that other unaffected ecosystem
services do not have value or are less important. b. Quantify changes in ecosystem services in
Because these other services will not be affected units of measurement that can be linked to
by the restoration project, their value will not be societal well-being.
changed by the restoration project, and knowing 3. Value final ecosystem services.
their value will thus not affect decision making
related to the restoration project. Identifying key a. For each ecosystem service, determine
ecosystem services can be accomplished by whether it provides value to the resource
making a list of the restoration activities that a owner, the wider public, or both.
project will entail, and then making a list of the b. Focus valuation efforts on the types of
ecosystem services affected by these activities. benefits from each ecosystem service that
are the most important to the private
Step 2 of the framework involves quantifying the resource owner and the public.
final ecosystem services that were developed in
Step 1. This set of ecosystem services is called c. Use a range of economic valuation
“final” because the framework makes a methods to capture the different types of
values of ecosystem services.
distinction between the ecosystem services that
are directly valued by people and those that
provide an input to another good or service, and Lastly, Step 3 of the framework involves
are thus valued only indirectly. Another way to ascribing a monetary value to the quantified
think of this distinction is to focus only on changes in ecosystem services identified in
ecosystem services that are directly enjoyed, Step 1 and quantified in Step 2. The steps of
consumed, or used by human beings. The focus quantifying and monetizing ecosystem services
only on the end products of ecosystem services are challenging, and require changes in

Large Wood National Manual July 2015


2-31
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

ecosystem services to be expressed in examples of indirect use values of ecosystem


measurable units that are meaningful to people services include flood protection, waste
and reflect the sense of the well-being people assimilation, and carbon sequestration.
receive from ecosystem services. For the
In addition to the current use of ecosystem
purpose of valuation, it is often important to
services, another type of value is that which
make a distinction between benefits that accrue
people place on the option to use ecosystem
to private owners of a resource and those that
services in the future. This type of value is called
accrue to the general public. Ecosystem services
option value, an example of which is a wilderness
can then be valued by estimating the well-being
area that one hopes to visit one day, or a species
people receive from their use, by the reduced
of bird one hopes to someday see.
costs to society for not having to provide these
services by other means, or by the avoided Another type of value, non-use, does not involve
damages that could result if these ecosystem any actual direct or indirect use by people. One
services were lost (U.S. Environmental type of non-use value, existence value, is the
Protection Agency, 2010). value people place on knowing an environmental
amenity exists, even if they have no plans to
Values of Ecosystem Services personally use it. Existence values are commonly
Ecosystem services that are enhanced or identified with rare landscapes or with
improved by restoration projects provide value threatened or endangered species. Bequest value,
to private resource owners and the public in another non-use value, refers to the value people
different ways. At the highest level, such values place on knowing that resources will be available
can be broken down into use values and non-use for use by their children or future generations.
values, with total economic value representing
the sum of these two components. Methods for Estimating Economic Values
of Ecosystem Services
A variety of methods exist for estimating the
CROSS-REFERENCE
value of ecosystem services, with different
Chapter 3, Ecological and Biological Considerations, methods being more suited to estimating the
provides an indepth examination of using models to benefits of different ecosystem services. The
quantify the biological benefits of restoration. objective of all of these methods is to estimate
the net benefits that ecosystem services provide,
or their benefits to society over and above the
Use values are the most straightforward manner
cost required to obtain them. The measure of net
in which ecosystem services provide value to
benefit is also called consumer surplus and can be
society. Direct use refers to when humans
thought of as the difference between what a
directly use the end product of an ecosystem
person is willing to pay to receive a good or
service, such as consuming fish and animals,
service and what they must actually pay for it.
harvesting timber, or utilizing other forest
Economic valuation methods measure consumer
products. Humans also use ecosystem services
surplus either by estimating willingness to pay
indirectly, such as when an ecosystem service is
and total cost for a good or service and then
an input into something else that human beings
taking the difference of the two, or by measuring
directly use. For example, ecosystems provide
consumer surplus directly.
habitat for plants and animal species that are
then used by people, either consumptively or Economic valuation methods can be classified at
non-consumptively (such as through wildlife the highest level into revealed preference
viewing). Habitat provision is thus an indirect methods and stated preference methods.
use value provided by ecosystem services. Other

Large Wood National Manual July 2015


2-32
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Revealed Preference Methods value of a related non-market ecosystem service.


The most common application of this method
Revealed preference methods rely on actual
involves using property market data to estimate
market data to estimate the value of ecosystem
the value of ecosystem services associated with
services. The most straightforward approach of
private properties. In other words, hedonic
revealed preference methods is the market price
pricing involves using statistical techniques to
method, which uses market data for a good or
infer the value of ecosystem services by
service as a means for measuring consumer
comparing the value of properties that include
surplus. The market price method can be used
these services with similar properties that do not
when the products of ecosystem services are
include them. Hedonic pricing has been used
directly traded in market, such as the markets
extensively to value the impact of open space
for timber, food, or fuel, or for ecosystem
and other environmental amenities such as air
services that produce goods and services sold
and water quality on properties. A drawback of
directly in markets. Despite the simplicity of
this method is that it can only be used to value
using the market price method, its application
ecosystem services that would affect property
for valuing ecosystem services is limited because
values.
it can only be used for ecosystem services with a
direct tie to market activity. This method also Another revealed preference approach, the
does not provide a direct measure of consumer travel cost method, is commonly used to value
surplus, which must be inferred or estimated ecosystem services related to recreation and
from the existing market data. aesthetic enjoyment. This method is based on the
premise that the value that people place on a
Other revealed preference methods can be used
recreational resource is equal to the amount of
to estimate the values of ecosystem services that
time and money they had to spend to travel to
are less directly tied to market activity. For
the resource and to use it. Information on travel
example, some valuation methods value
costs is most often collected through surveys of
ecosystem services by estimating the cost of
visitors to a specific recreation site. The travel
replacing the services provided by ecosystems
cost method has been used extensively to value
by other means. This method, the replacement
parks, open space, and other sites used for
cost method, can be used to value ecosystem
outdoor recreation. This method, however, is
services such as water quality that could be
limited in its applicability because it can only be
provided by other means if they were not
used to measure resources related to recreation
provided by ecosystems. Another, the damage
and visitation by people.
cost method, values ecosystem services by
estimating the costs of damages that could result Stated Preference Methods
if an ecosystem service were lost. For example,
the water retention benefits of a restoration As opposed to revealed preference methods that
project could be valued by estimating the costs use actual market data, stated preference
that the project would help to avoid from floods methods use public opinion surveys to ask
that would be more likely to occur without water people about their values for ecosystem services.
retention services. One drawback of these Stated preference methods are commonly used
approaches, however, is that they estimate costs to value ecosystem services that do not have
of replacing lost services or avoiding damages, clear ties to market activity. Additionally, stated
which may not represent the full value that preference methods are the only valuation
societies place on ecosystem services. method that can capture the non-use component
of the value of ecosystem services. One common
Another revealed preference approach is the stated preference method, contingent valuation,
hedonic pricing method. This method uses the is conducted by creating a hypothetical market
value of related market goods to estimate the

Large Wood National Manual July 2015


2-33
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

for a good or service, and then asking people to value, because transferring a function allows the
state their willingness to pay for a change in the researcher to customize the variables in the
level of provision of it. Another stated preference function to match the current analytical context.
method, choice modeling, asks people to select
their preferred option among repeated choices 2.3.4 Using Structured
among alternatives with differing levels of
provision of attributes. Both of these approaches
Decision Making
can be used to estimate the consumer surplus Structured decision making is a general concept
associated with ecosystem services, and can be that applies to a carefully organized analysis of
used to estimate the value of proposed programs problems used to reach decisions that are
or policies that have not yet occurred. Despite focused on achieving clearly defined
these advantages, however, the hypothetical fundamental objectives (Gregory et al. 2012;
nature of stated preference methods can be Clemen and Reilly 2001; Kirkwood 1997; Keeney
controversial and has led to criticism of these and Raiffa 1993). Based in decision theory and
methods and the results obtained by them. Also, risk analysis, SDM encompasses a simple set of
due to the need to collect survey data, stated concepts and helpful steps, rather than a rigidly
preference methods are time consuming and prescribed approach, for problem solving
expensive to conduct. (Figure 2-10). Key SDM concepts include making
An alternate approach to conducting original decisions based on clearly articulated objectives,
revealed and stated preference studies is to use addressing uncertainty, and responding
benefit transfer methods, which customize, or transparently to legal mandates and public
adapt, the results of previous studies to fit a new preferences or values in decision making. As
context. Given the time and expense needed to such, SDM integrates science and policy. Every
conduct primary studies, benefit transfer decision consists of several primary elements:
methods are widely used by government management objectives, decision options, and
agencies and other researchers to value predictions of decision outcomes (Table 2-1). By
ecosystem services. Conducting a benefit analyzing the components within a
transfer initially involves doing a comprehensive comprehensive decision framework, it is
literature search to learn if similar ecosystem possible to improve the quality of decision
services have been valued in other studies. Any making. The core SDM concepts and steps are
identified similar studies are then evaluated for applicable to all types of decisions, from those
their quality and suitability for a benefit transfer associated with minor restoration projects to
exercise. If suitable source data can be found, the complex public sector decisions involving
next step is to transfer the original value to the multiple decision makers, scientists, and other
new study context. The transfer is accomplished stakeholders. The key component for success is
either through a direct transfer of the estimated the ability to integrate quantifiable information
value, or a transfer of the function used to at critical steps in the process.
estimate the original value. Many researchers
prefer to transfer the benefit as opposed to the

Large Wood National Manual July 2015


2-34
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Figure 2-10. The Structured Decision Making Process

Source: Modified from USFWS 2008.

Large Wood National Manual July 2015


2-35
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Table 2-1. Steps in Structured Decision Making

Step Important Considerations


1. Problem Definition What specific decision has to be made? What is the spatial and
temporal scope of the decision? Will the decision be iterated over
time?
2. Establishing Objectives What are the management objectives? Ideally, these are stated in
quantitative terms that relate to metrics that can be measured.
Setting objectives falls in the realm of policy, and should be
informed by legal and regulatory mandates, as well as stakeholder
viewpoints.
3. Defining/Understanding Alternatives What are the different management actions from which we can
choose? This element requires explicit articulation of the
alternatives available to the decision maker. The range of
permissible options is often constrained by legal or political
considerations, but structured assessment may lead to creative new
alternatives.
a. Uncertainty Because we rarely know precisely how management actions will
affect natural systems, decisions are frequently made in the face of
uncertainty. Uncertainty makes choosing among alternatives far
more difficult. A good decision-making process will confront
uncertainty explicitly, and evaluate the likelihood of different
outcomes and their possible consequences. Scientific uncertainty
will exist in the flow alteration–ecological response relationships, in
part because of the confounding of hydrologic alteration with other
important environmental determinants of river ecosystem condition
(e.g., temperature).
b. Risk Tolerance Identifying the uncertainty that impedes decision making, then
analyzing the risk that uncertainty presents to management is an
important step in making a sound decision. Understanding the level
of risk a decision maker is willing to accept, or the risk response
determined by law or policy, will make the decision-making process
more objectives-driven, transparent, and defensible.
c. Linked Decisions Many important decisions are linked over time. The key to
effectively addressing issues associated with linked decisions is to
isolate and resolve the near-term issues while sequencing the
collection of information needed for future decisions.
4. Quantifying Consequences What are the consequences of different management actions? To
what degree would each alternative lead to successfully reaching a
given objective? In SDM, we predict the consequences of the
alternative actions with an appropriately chosen model. Depending
on the information available or the quantification desired for a
structured decision process, consequences may be modeled with
highly scientific computer applications, or with personal judgment
elicited carefully and transparently. Ideally, models are quantitative,
but they need not be; what is most important is that they link
actions to consequences.

Large Wood National Manual July 2015


2-36
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Step Important Considerations


5. Understanding Tradeoffs If there are multiple objectives, how do they trade off with each
other? In most complex decisions, the best we can do is to choose
intelligently between less-than-perfect alternatives. Numerous tools
are available to help determine the relative importance or weights
among conflicting objectives; this information is used to compare
alternatives across multiple attributes to find the best compromise
solutions.
6. Decide and Take Action For those decisions that are iterated over time, actions taken early
on may provide a learning opportunity that improves management
later, provided that an appropriate monitoring program is in place
to provide the feedback. Adaptive management is a special case of
structured decision making for decisions that are iterated or linked
over time.
Source: Gregory et al. (2012).

2.3.4.1 Decision Support Tools made available practical methods for applying
scientific decision theoretical approaches to
Decision making for ecosystem restoration multi-criteria problems. For example, in 2010,
projects can be a complex and challenging USACE recognized its value in achieving its
process, characterized by trade-offs between environmental mission by considering a broad
socio-political, environmental, and economic range of criteria (USACE IWR 2010). MCDA
impacts. The adherence to appropriate techniques were identified as excellent ways to
environmental policies, land-use planning, and help USACE planners and project managers
other regulatory decision-making challenges balance their decisions based on social equality,
involves multiple selection criteria such as cost, environmental soundness, and economic
benefit, environmental impact, safety, and risk. viability. To this end, MCDA techniques are
As such, mangers have often used cost-benefit tools USACE can use to improve the
analyses, occasionally in concert with transparency of the decision-making process.
comparative risk assessment, to choose MCDA provides a proven mathematical means
between competing project alternatives. for comparing criteria with differing units such
Additionally, some selection criteria cannot as habitat units, cultural resources, public
easily be condensed into simple values, which sentiment, and total cost. The stakeholders,
complicates the integration of resource and both those in support and against a project, can
socioeconomic values inherent in making provide input into the criteria used to evaluate
comparisons and trade-offs. Furthermore, plans. The plans and their effects are plainly
environmental concerns often involve ethical described in the decision matrix, allowing the
and moral principles that may not be related to stakeholders and project team an greater
any economic use or value. To this end, this understanding of the problems associated with
manual presents two decision support tools a particular plan. As such, MCDA is a valuable
that enhance the decision-making process. tool that can be applied to many complex
decisions. It is most applicable to solving
Multi-Criteria Decision Analysis problems that are characterized as a choice
MCDA is a sub-discipline of operations research among alternatives and has all the
that explicitly considers multiple criteria in characteristics of a useful decision support tool:
decision-making environments (Clemen and it helps us focus on what is important, logical
Reilly 2001; Kirkwood 1997; Keeney and Raiffa and consistent, and is easy to use.
1993). Considerable research on MCDA has

Large Wood National Manual July 2015


2-37
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

GUIDANCE

The International Society on Multiple Criteria Decision Making (MCDM)

MCDM identifies software packages that can assist with MCDA (http://www.mcdmsociety.org/soft.html):
 1000Minds software for MCDM, prioritization, and resource allocation. Internet-based and free for academic
use.
 BENSOLVE Free MatLab implementation of Benson’s algorithm to solve linear vector optimization problems.
 Decisionarium, global space for decision support (for academic use).
 DEXi, program for qualitative multi-attribute decision modeling, developed at the Jožef Stefan Institute,
Ljubljana, Slovenia.
 D-Sight, visual and interactive tool for multicriteria decision aid problems based on the PROMETHEE methods
and Multi-Attribute Utility Theory.
 GUIMOO, Graphical User Interface for Multi Objective Optimization from INRIA.
 IDS Intelligent Decision System for Multiple Criteria Decision Analysis under Uncertainty (using the Evidential
Reasoning Approach).
 IDSS Software: MCDM software of the Laboratory of Intelligent Decision Support Systems (University of
Poznan, Poland).
 IND-NIMBUS: implementation of the interactive NIMBUS method that can be connected with different
simulation and modeling tools.
 Interalg free solver, which includes global nonlinear multiobjective optimization with user-defined accuracy.
 IRIS and VIP, IRIS: Interactive Robustness analysis and parameters’ Inference software for multicriteria Sorting
problems and VIP (Variable Interdependent Parameters) Analysis software.
 MACBETH for MCDA, Measuring Attractiveness by a Categorical Based Evaluation TecHnique in MultiCriteria
Decision Aid.
 MakeItRational, AHP based decision software.
 modeFRONTIER, commercial software developed by ESTECO Spa dedicated to multi-objective optimization
and multi-disciplinary design, providing an easy coupling to almost any Computer Aided Engineering (CAE)
tool.
 Collection of Multiple Criteria Decision Support Software by Dr. Roland Weistroffer.
 WWW-NIMBUS for solving nonlinear (and even nondifferentiable) multiobjective optimization problems in an
interactive way. Operates via the Internet, free for academic use.
 ParadisEO-MOEO, module specifically devoted to multiobjective optimization in ParadisEO, software
framework for the design and implementation of metaheuristics, hybrid methods as well as parallel and
distributed models from INRIA.
 Priority Estimation Tool, open-source (free) software for AHP-based decision making.
 PROMETHEE-GAIA software.
 MCDA software by Quartzstar Ltd.: OnBalance for evaluation decisions and HiPriority for resource allocation.
 RGDB, Graphic tool that helps to select preferable rows from relational databases.
 Accord by Robust Decisions implementing the Bayesian Team Support technique.
 TransparentChoice - Strategic decision-making software, MCDM software that allows multi-disciplinary teams
to collaborate on complex decisions.
 VISA, Web based Multi-Criteria Decision Making Software.

Large Wood National Manual July 2015


2-38
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Within the context of SDM, MCDA helps natural the appropriate information in making river
resource decision makers talk about the restoration decisions (Skidmore et al. 2011).
restoration project in a way that allows them to RiverRAT includes a suite of resources to guide
consider both natural and socioeconomic more efficient, consistent, and comprehensive
values. It provides a tool for decision makers to reviews of stream management and restoration
consider and assess the complex trade-offs proposals. Such resources help determine the
among project alternatives. In effect, it helps depth of review required, ensure that a project
decision makers think, re-think, query, adjust, proposal is complete, and guide reviewers
decide, rethink some more, test, adjust, and through a thorough and scientifically sound
finally decide. To this end, the typical elements project review. The RiverRAT Science
of the MCDA tool integrate well with SDM. Document and its appendices provide a
comprehensive synthesis of science behind
 Define the Decision
stream management and restoration project
 Identify Decision Criteria development.
 Build a Decision Framework The ultimate, long-term goals of RiverRAT
 Rate the Alternatives include:

Two common rating scales that that are  Enabling consistent, comprehensive,
used in MCDA are: transparent, and documented project
reviews.
o Relative Scale:
 Facilitating improved project planning and
Each alternative is rated relative to the design.
others in satisfying a particular interest.
For example, among four alternatives,  Encouraging projects that are attuned to
assign each a 1, 2, 3, or 4 depending on their watershed and geomorphic context.
which satisfies the interest: the best = 4;  Improving the science and technology of
second best = 3; third best = 2; and the stream restoration and management.
worst at satisfying the interest = 1.
The RiverRAT tools, the supporting Science
o Ordinal Scale:
Document, and the detailed technical
Using a scale of your choosing (e.g., a appendices, are available to the public at
five–point scale, or a ten-point scale) www.restorationreview.com. For example, the
assign each alternative a rating for how Project Screening Matrix and River Restoration
well it satisfies a particular interest. For Analysis Tool is a good example of the practical
example, a five-point scale might be: 5 = application of river restoration concepts to
excellent; 4 = good; 3 = satisfactory; 2 = individual projects (Skidmore et al. 2011).
below average; 1 = poor.
 Weight Decision Maker/Stakeholder CROSS-REFERENCE
Interests
 Score the Alternatives Figure 7-3 in Chapter 7, Risk Considerations, shows
the RiverRAT Project Risk Screening Matrix.
 Discuss Results, Re-Score, Discuss Again,
Decide

River Restoration Analysis Tool


RiverRAT was developed to address the failure
by resource managers to consider and integrate
Large Wood National Manual July 2015
2-39
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

2.4 References
Abbe, T. B., and A. P. Brooks. 2011. Geomorphic, Engineering, and Ecological Considerations when
Using Wood in River Restoration. Pages 419–451 in A. Simon, S. J. Bennett, and J. M. Castro
(eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools.
Geophysical Monograph Series 194. Washington, D.C.: American Geophysical Union.

Abbe, T. B., and D. R. Montgomery. 1996. Large Woody Debris Jams, Channel Hydraulics and Habitat
Formation in Large Rivers. Regulated Rivers Research and Management 12:201–221.

Abbe, T. B., and D. R. Montgomery. 2003. Patterns and Processes of Wood Debris Accumulation in
the Queets River Basin, Washington. Geomorphology 51:81–107.

Abbe, T. B., D. R. Montgomery, and C. Petroff. 1997. Design of Stable In-Channel Wood Debris
Structures for Bank Protection and Habitat Restoration: An Example from the Cowlitz River, WA.
Pages 809–816 in S. S. Y. Wang, E. J. Langendoen, and F. D. Shields (eds.), Proceedings of the
Conference on Management of Landscapes Disturbed by Channel Incision. University of
Mississippi, Oxford, MS.

Abbe, T. B., J. Carrasquero, M. McBride, A. Ritchie, M. McHenry, and K. Dublanica. 2003a.


Rehabilitating River Valley Ecosystems: Examples of Public, Private, and First Nation Cooperation
in Western Washington. Proceedings of the Georgia Basin/Puget Sound 2003 Research
Conference, Vancouver, B.C., March 31–April 1, 2003, T. Droscher (ed.). Puget Sound Action
Team, Olympia, WA.

Abbe T. B, A. P. Brooks, and D. R. Montgomery. 2003b. Wood in River Rehabilitation and


Management. Pages 367–389 in S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology
and Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society.

Abbe, T. B., G. Pess, D. R. Montgomery, and K. L. Fetherston. 2003c. Integrating Engineered Log Jam
Technology into River Rehabilitation. In D. R. Montgomery, S. Bolton, D. Booth, and L. Wall
(eds.), Restoration of Puget Sound Rivers. Center for Water and Watershed Studies.

Abbe, T. B., C. Miller, and A. Michael. 2009. Self-Mitigating Protection for Pipeline Crossings in
Degraded Streams: A Case Study from Woodward Creek, Washington. 9th International Right of
Way Symposium. 2009. Portland, OR.

Beechie, T. J., and S. Bolton. 1999. An Approach to Restoring Salmonid Habitat-Forming Processes in
Pacific Northwest Watersheds. Fisheries 24:6–15.

Beechie, T. J., D. A. Sear, J. D. Olden, G. R. Pess, J. M. Buffington, H. Moir, P. Roni, and M. M. Pollock.
2010. Process-Based Principles for Restoring River Ecosystems. Bioscience 60:209–222.

Beechie, T. J., H. Imaki, J. Greene, A. Wade, H. Wu, G. Pess, P. Roni, J. Kimball, J. Stanford, P. Kiffney,
and N Mantua. 2012. Restoring Salmon Habitat for a Changing Climate. River Research and
Applications 29:939–960.

Booth, D. B. 1990. Stream-Channel Incision Following Drainage-Basin Urbanization. Water Resources


Bulletin 26:407–417.

Large Wood National Manual July 2015


2-40
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Brummer, C., T. B. Abbe, J. R. Sampson, and D. R. Montgomery. 2006. Influence of Vertical Channel
Change Associated with Wood Accumulations on Delineating Channel Migration Zones,
Washington State, USA. Geomorphology 80:295–309Clemen, R. T., and T. Reilly. 2001. Making
Hard Decisions with Decision Tools. Pacific Grove, CA: Duxbury Thomson Learning.

Buffington, J. M., and D. R. Montgomery. 1999a. A Procedure for Classifying Textural Facies in Gravel-
Bed Rivers. Water Resources Research 35(6):1903-1914.

Buffington, J. M., and D. R. Montgomery. 1999b. Effects of Hydraulic Roughness on Surface Textures
of Gravel-Bed Rivers. Water Resources Research 35(11):3507–3521.

Buffington, J. M., and D. R. Montgomery. 1999c. Effects of Sediment Supply on Surface Textures of
Gravel-Bed Rivers. Water Resources Research 35(11):3523–3530.

Clemen, R. T., and T. Reilly. 2001. Making Hard Decisions with Decision Tools. South-Western Cengage
Learning. 733 pp.

Collins, B. D., D. R. Montgomery, and A. J. Sheikh. 2003. Reconstructing the Historical Riverine
Landscape of the Puget Lowland. Pages 79–128 in D. R. Montgomery, S. M. Bolton, D. B. Booth,
and L. Wall (eds.), Restoration of Puget Sound Rivers. University of Washington Press, Seattle.

Convertino, M., K. M. Baker, J. T. Vogel, C. Lu, B. Suedel, and I. Linkov. 2013. Multi-Criteria Decision
Analysis to Select Metrics for Design and Monitoring of Sustainable Ecosystem Restorations. U.S.
Army Research. Paper 190. Available: <http://digitalcommons.unl.edu/usarmyresearch/1905>.

Evergreen Funding Consultants (EFC). 2003. A Primer on Habitat Project Costs. Available:
http://www.evergreenfc.com/section_services/resources/primer.pdf.

Fox, M. J. 2003. Spatial Organization, Position, and Source Characteristics of Large Woody Debris in
Natural Systems. Ph.D. dissertation. College of Forest Resources, University of Washington.
Seattle, Washington.

Gregory, R., L. Failing, M. Harstone, G. Long, T. McDaniels, and D. Ohlson. 2012. Structured Decision
Making: A Practical Guide to Environmental Management Choices. ISBN: 978-1-4443-3341-1.
Wiley-Blackwell. 312 pp.

Heede, B. H. and J. N. Rinne, 1990. Hydrodynamic and Fluvial Morphological Processes: Implications
for Fisheries Management and Research. North American Journal of Fisheries Management,
10:249–268.

Hester, E. T., M. W. Doyle, and G. C. Poole. 2009. The Influence of In-Stream Structures on Summer
Water Temperatures via Induced Hyporheic Exchange. Limonology and Oceanography 54:355–
367.

International Union for Conservation of Nature (IUCN). 2014. IUCN Releases an Economic Framework
for Analyzing Forest Landscape Restoration Decisions. Available:
http://cmsdata.iucn.org/downloads/flr_economic_analysis_tutorial___july_2014_1.pdf.

Kauffman, J. B., R. L. Beschta, N. Otting, and D. Lytjen. 1997. An Ecological Perspective of Riparian
and Stream Restoration in the Western United States. Fisheries (Bethesda) 22:12–24.

Keeney, R. L., and H. Raiffa. 1993. Decisions with Multiple Objectives Preferences and Value Tradeoffs.
Cambridge University Press.

Large Wood National Manual July 2015


2-41
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Kirkwood, C. W. 1997. Strategic Decision Making. Multi-Objective Decision Analysis with Spreadsheets.
Wadsworth.

Li, R., and H. W. Shen. 1973. Effect of Tall Vegetation on Flow and Sediment. Journal of
the Hydraulic Division, ASCE 99(5):793–814.

Linkov. I., and E. Moberg. 2012. Multi-Criteria Decision Analysis: Environmental Applications and Case
Studies. CRC Press. ISBN: 978-1-4398-5318-4.

Linkov, I., A. Varghese, S. Jamil, T. P. Seager, G. Kiker, and T. Bridges. 2005. Multi-Criteria Decision
Analysis: A Framework for Structuring Remedial Decisions at Contaminated Sites. In I. Linkov
and A. Bakr Ramadan (eds.), Comparative Risk Assessment and Environmental Decision Making.
NATO Science Series Volume 38 2005ISBN: 978-1-4020-1895-4 (Print) 978-1-4020-2243-2
(Online).

Lisle, T. 1995. Effects of Coarse Woody Debris and its Removal on a Channel Affected by the 1980
Eruption of Mount St. Helens, Washington. Water Resources Research 31:1797–1808.

Manga, M., and J. W. Kirchner. 2002. Stress Partitioning in Streams by Large Woody Debris. Water
Resources Research 36:2373–2379.

Montgomery, D. R., J. M. Buffington, R. D. Smith, K. M. Schmidt, and G. Pess. 1995. Pool Spacing in
Forest Channels. Water Resources Research 31:1097–1105.

Murphy, M. L. 1995. Forestry Impacts on Freshwater Habitat of Anadromous Salmonids in the


Pacific Northwest and Alaska—Requirements for Protection and Restoration. U.S. Department of
Commerce Coastal Ocean Program, NOAA. Decision Analysis Series No. 7, 156 pp.

Noss, F., and R. L. Peters. 1995. Endangered Ecosystems – A Status Report on America’sVanishing
Habitat and Wildlife. 133 Pages. Defenders of Wildlife, 1101 Fourteenth Street, NW, Suite 1400,
Washington, DC 20005.

Pastoroka, R. A., A. MacDonaldb, J. R. Sampsona, P. Wilberc, D. J Yozzod, and J. P Titred. 1997. An


Ecological Decision Framework for Environmental Restoration Projects. Ecological Engineering
9 (1–2):89–107.

Pelto, M. S. 2011. North Cascade Glacier Retreat. Nichols College, Dudley, MA. Available:
http://www.nichols.edu/departments/glacier/bill.htm.

Poole, G. C., J. A. Stanford, S. W. Running, and C. A. Frissell. 2006. Multiscale Geomorphic Drivers of
Groundwater Flow Paths: Subsurface Hydrologic Dynamics and Hyporheic Habitat Diversity.
Journal of the North American Benthological Society 25:288–303.

Rapp, C., and T. Abbe. 2003. A Framework for Delineating Channel Migration Zones. Washington State
Department of Ecology Publication Number 03-06-027. Final Draft.

Reid, L. M., and T. Dunne. 1996. Rapid Evaluation of Sediment Budgets. Reiskirchen, Germany: Catena
Verlag (GeoEcology paperback). 164 pp.

Sedell, J. R., and J. L. Frogatt. 1984. Importance of Streamside Forests to Large Rivers: The Isolation
of the Willamette River, Oregon, U.S.A., from its Floodplain by Snagging and Streamside Forest
Removal. Verhandlungen-Internationale Vereinigung für Theorelifche und Angewandte
Limnologie 22:1828–1834.

Large Wood National Manual July 2015


2-42
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

Shields, F. D., Jr., and C. V. Alonso. 2012. Assessment of Flow Forces on Large Wood in Rivers. Water
Resources Research 48(4):W04156.

Skidmore, P. B., C. R. Thorne, B. L. Cluer, G. R. Pess, J. M. Castro, T. J. Beechie, and C. C. Shea. 2011.
Science Base and Tools for Evaluating Stream Engineering, Management, and Restoration
Proposals. U.S. Department of Commerce. NOAA Tech. Memo. NMFS-NWFSC-112.

Society for Ecological Restoration. 2002. SER International Primer on Ecological Restoration. Science
& Policy Working Group, Version 2, October. Available: http://www.ser.org/resources/
resources-detail-view/ser-international-primer-on-ecological-restoration.

Society for Ecological Restoration. 2004. The SER International Primer on Ecological Restoration.
Available: <http://www.ser.org>.

Suding, K. N., and R. J. Hobbs. 2009. Threshold Models in Restoration and Conservation: A
Developing Framework. Trends in Ecology & Evolution 24 (5):271–279.

U.S. Army Corp of Engineers Institute for Water Resources (USACE IWR). 2010. IWR Planning Suite
MCDA Module User’s Guide. U.S. Army Corp of Engineers Institute for Water Resources.

U.S. Environmental Protection Agency (EPA). 1995. A Decision-Making Guide for Restoration in
Ecological Restoration. EPA 841-F-95-007 (November)

U.S. Environmental Protection Agency (EPA). 2000. Principles for the Ecological Restoration of
Aquatic Resources. EPA841-F-00-003. Available: <http://www.epa.gov/owow/wetlands/
restore/.

U.S. Environmental Protection Agency (EPA). 2008. Handbook for Developing Watershed Plans to
Restore and Protect our Water.

U.S. Environmental Protection Agency. 2009. Valuing the Protection of Ecological Systems and
Services. May. Available: http://yosemite.epa.gov/sab%5CSABPRODUCT.NSF/
F3DB1F5C6EF90EE1852575C500589157/$File/EPA-SAB-09-012-unsigned.pdf. Accessed:
October 8, 2014.

U.S. Environmental Protection Agency. 2013. A Quick Guide to Developing Watershed Plans to Restore
and Protect Our Waters. Available: http://water.epa.gov/polwaste/nps/upload/
watershed_mgmnt_quick_guide.pdf.

U.S. Fish and Wildlife Service (USFWS). 2008. SDM Fact Sheet. Available:
http://www.fws.gov/science/doc/structured_decision_making_factsheet.pdf. Accessed: May 15,
2015.

Wohl, E. 2013. Floodplains and Wood. Earth-Science Reviews 123:194–212.

Wyant, J. G., R. A. Meganck, and S. H. Ham. 1995. A Planning and Decision-Making Framework for
Ecological Restoration. Environmental Management 19(6):789–796.

Large Wood National Manual July 2015


2-43
Bureau of Reclamation and Chapter 2. Large Wood and the Fluvial
U.S. Army Corps of Engineers Ecosystem Restoration Process

This page intentionally left blank.

Large Wood National Manual July 2015


2-44
Chapter 3
ECOLOGICAL AND BIOLOGICAL CONSIDERATIONS

Photo credit: Ken DeCamp

AUTHORS

Willis McConnaha (ICF International)


Dana Warren (Oregon State University)
Jordan Rosenfeld (British Columbia Ministry of Environment)
Leo Lentsch (ICF International)
Tom Stewart (ICF International)
This page intentionally left blank.
3.1 Introduction GUIDANCE

Large wood can be found in nearly all streams


This chapter provides a basic understanding of the
with forested riparian areas. Large wood has following:
several ecological functions in streams, such as
trapping sediment, creating structure,  Ecological functions of large wood in streams.
providing shade and cover for aquatic  Assessing the need for wood placement.
organisms, and diversifying flows. Wood is a
 Natural sources of wood.
key habitat component for many stream
organisms, particularly for fish such as  Scale and the River Continuum Concept.
salmonids.  Hyporheic zone.
Resource managers using wood to restore  Wood as habitat for invertebrates and
ecological functions of streams are faced with terrestrial species.
many questions, such as:  Importance of assessing wood placement.
 What are the biological purposes in
restoring large wood?
In most cases, restoration of large wood is
 What ecological functions are enhanced by undertaken to achieve some biological goal.
restoring large wood? Hence, the inherent assumption of restoration
of large wood is that habitat features in streams
 What is the potential contribution of large
associated with wood are positively related to
wood to achievement of management goals?
the survival, persistence, and abundance of
 How does wood restoration relate to fish desired species and communities and ecological
habitat? functions (Whiteway et al. 2010). The
 How do we manage riparian forests to relationship between individual habitat
maintain wood supply to the channel? attributes and fish survival or abundance can be
difficult to prove in a quantifiable and
The answers to these questions depend on how statistically meaningful way (Conquest and
the quantity and quality of habitat relates to fish Ralph 1998; Bradford et al. 2005).
abundance, distribution, and persistence over Consequently, some researchers have
time and how these are affected by large reasonably questioned the benefits of stream
instream wood. These issues are addressed in restoration activities (Thompson 2006; Stewart
this chapter. The chapter discusses the et al. 2009), or called for a better accounting of
ecological and other biological considerations the costs and benefits of restoration
associated with large wood in streams. The investments (Bernhardt et al. 2007). Benefits
discussion favors the role of wood in salmonid are challenging to detect, in part, because of the
ecosystems, reflecting the preponderance of number of confounding factors affecting fish
research and the key role of wood in salmonid abundance in any year or over time, especially
ecosystems. However, many of the principles at a population scale for far-ranging
derived from work on salmonid ecosystems are anadromous species such as salmon (Rose
applicable to other systems and species. 2000).

The observed response of fish and other


organisms to wood enhancement also reflects a
suite of watershed level conditions that can
obscure the effects of site-specific wood
restoration. Engineering solutions that do not

Large Wood National Manual July 2015


3-1
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

account for species habitat needs, stream salmonids although there were appreciable
dynamics, disturbance regimes, and watershed differences among species. In should be noted,
characteristics are often unsuccessful (Beschta however, that increases in abundance at project
1997). Nagayama and Nakamura (2010) sites by themselves can be misleading. The most
reviewed the success of wood enhancement relevant metric is the whether or not those
worldwide and observed that wood restoration increases in abundance cascade to a population
has localized effects but may not be sufficient to level increase.
recovery fish populations at the watershed
scale. They found ample examples of
restoration projects that have failed because of 3.2 Ecological Functions
physical failure or sediment accumulation. They of Large Wood
conclude that, “restoration projects should be
aimed at restoring natural processes of wood Large wood is a key structural element in
recruitment and routing, which can provide fish forested stream ecosystems worldwide (Maser
and other organisms with sustainable wood and Sedell 1994; Nagayama and Nakamura
habitats at the watershed scale over the long 2010). Wood serves as a food resource for
term” (Nagayama and Nakamura 2010). In microbes, fungi, and macroinvertebrates. In
other words, large wood enhancement should addition, a primary ecological role of large
be viewed as an interim restoration measure wood and accumulations of large wood (wood
until natural processes of wood recruitment jams) is associated with its influence on the
recover to natural levels. physical environment of streams and the
creation of habitats for aquatic species (Roni et
On the whole, however, the bulk of evidence
al. 2014a).
supports the notion that the addition of large
wood and large wood structures can, in many The influence of wood on stream habitat and
cases, provide habitat features believed to be stream ecosystem processes is affected by
conducive to fish production and that stream size, wood stability, stream gradient,
restoration generally (though not always) and the underlying geology. In low-gradient
results in greater abundance and/or biomass of (blackwater) systems typical of the
fish at life stage and population scales. For southeastern United States where the bottom of
example, Roni et al. (2014b) reviewed 409 the stream is composed of fine, unconsolidated
published studies evaluating specific sediments, large wood and wood jams can
restoration actions in terms of fish response provide a stable substrate that can enhance
and found generally positive, though variable, invertebrate abundance, productivity, and
fish response to restoration actions. Many of diversity (Smock et al. 1989; Johnson et al.
these studies (209) focused on placement of 2003; Stewart et al. 2012). As the stream
logs and instream structures; and the bulk of gradient increases, stream power also
the studies demonstrated a positive biological increases. In alluvial systems, wood becomes
response in terms of increased abundance of increasingly important in pool formation and in
juvenile or adult salmonids, a minority of the retention of sediments, particulate organic
studies showed no response, and only a few matter, and the inorganic bedload in streams
studies found a negative response to placement (Wallace et al. 1995a; Roni et al. 2008).
of instream structures. In a meta-analysis of
Wood and wood jams also enhance habitat for
published studies of large wood structures in
fish by increasing the complexity of the stream
streams, Whiteway et al. (2010) found that
environment, providing habitat for multiple life
wood structure provided key habitat elements
stages and species. This increase in habitat
including pools and cover; most studies also
complexity can occur even when wood has no
reported increases in density and biomass of

Large Wood National Manual July 2015


3-2
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

direct pool forming function (Berg et al. 1998; abundance, persistence, and fitness (Southwood
Flebbe 1999). The physical and visual isolation 1977). Habitats for species can overlap but are
from competitors and predators that a complex usually separated temporally, spatially, or in
tangle of wood provides may be as or more terms of function. For example, large wood can
important in enhancing habitat for fish than be an element of habitat for both juvenile
pool formation (although these factors often go salmonids and benthic insect life stages, but the
hand-in-hand). Numerous field studies have nature of that habitat differs; wood generally
attributed increases in fish abundance provides cover for juvenile salmonids while it
associated with wood additions to increases in provides a substrate on which benthic insects
habitat complexity and visual isolation from move and feed.
predators and competitors that accompanies
The abundance and persistence of a species in
the increase in wood. In experimental channels
an environment reflect the quality and quantity
the presence of even a single piece of wood has
of habitat and food resources experienced along
been found to dramatically reduce aggressive
spatial-temporal pathways defined by the
interactions among individuals and enhance
species’ life history, as well as predation and
growth of both dominant and subordinate
competition. Habitat along the life history
individuals (Sundbaum and Naslund 1998). The
pathway consists of patches arrayed across
potential for wood to create complex habitat
space and time that are linked by the life history
that increases local fish abundance extends
trajectory of the species (Fausch et al. 2002). In
beyond small streams and salmonids fish
freshwater, these patches are often formed and
(where most research on this topic has
maintained as a result of instream wood.
focused); wood additions to larger river
Habitat patches are distributed across the
systems increase habitat diversity at multiple
riverscape, varying in quality and quantity,
spatial scales and have also been found to
resulting in the heterogeneous distribution and
elevate local fish abundances (Pess et al. 2012).
performance of individuals and the population
(Townsend 1989; Pickett and Rogers 1997).
3.2.1 Habitat Formation
Physical features of the stream that are
The flow obstruction created by large wood is perceived as habitat by biota form as the result
effective at increasing the range of physical of a hierarchy of controls, ranging from regional
habitat through diversification in flow depths, to watershed to reaches and channel units
velocities, substrate size, and bed morphology. (Frissell et al. 1986; Montgomery and
Large wood can transform an otherwise planar Buffington 1998). Large wood contributes to
morphology reach with relatively uniform the formation and maintenance of habitat types
hydraulics into a reach where pool scour, and survival factors at reach and channel-unit
sediment sorting, and bar formation can scales. Formation of geomorphic channel units
directly create new habitat; given the right that constitute habitat for salmonid life stages is
conditions, large wood can transform stream dependent on flow, channel form, riparian
morphology into more complex channel and conditions, and structural elements, including
floodplain features that provide reach-scale large wood (Montgomery and Buffington 1998).
habitat enhancements. Habitat controls operate at the reach or channel
Habitat consists of elements of the environment unit scale while in turn being constrained by the
that affect the persistence and performance of a larger watershed context of controls that affect
species in a specific location (Whittaker et al. local environmental conditions.
1973; Hall et al. 1997). The quality and quantity It is important to view wood in the context of
of habitat across the life history of the species the entire life span experience of a species and
shape biological performance in terms of conditions encountered across its life history.

Large Wood National Manual July 2015


3-3
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Anadromous salmonids, for example, spend events greater than the 20-year flood) because
only a portion of their life history in freshwater, much of the flood water is out on the floodplain.
but the success of the population may be But large wood can increase the duration of
affected by spawning habitat and juvenile smaller floods (i.e., 1 to 2-year events) where
survival in streams, which are often closely tied most of the flow is still contained within the
to habitat conditions associated with large channel (Rutherford et al. 2007). Large wood of
wood in streams. The biological value of a given size will have a greater effect on a small
restoration of large wood in streams depends stream. Rutherford et al. (2007) report large
on the bio-physical context and the array of wood generally will not affect small flood
factors across a range of environments that events when the projected area of the large
potentially affect the success of taxa of interest wood is less than 10% of the area of the cross-
and their associated biological communities. section. The “projected” area is the area of the
large wood in a two-dimensional cross-section
The ability of large wood to form habitat varies
perpendicular to the channel (direction of flow).
considerably with the specific characteristics of
A large wood structure needs to be very large to
the channel type and of the wood pieces or jams
occupy 10% of the cross-section of a third order
themselves, including their size, position along
or higher stream.
the bank or within the channel, orientation to
flow, and porosity. Some of the ways in which
3.2.1.2 Hydraulic Diversity
large wood influences physical habitat are
discussed below. The presence of large wood will create highly
three-dimensional flow patterns in surface
3.2.1.1 Wetted Area of the waters including hydraulic refugia for fish
Channel (Daniels and Rhoads 2004). The hydraulics
associated with a piece of wood or logjam will
Large wood creates bedform roughness vary with the complexity of the wood
(resistance to flow, or drag) that effectively structure’s composition, including its size,
slows flow down, consequently raising the position, and orientation to flow. The flow
water surface level. This may facilitate a pattern associated with large wood jams is
hydraulic “backwater effect,” whereby the often analogous to the flow pattern
water level immediately upstream of the encountered at bridge abutments and piers,
obstruction is raised, which in turn raises the depending on whether the structure is bank-
level of water upstream of it, resulting in an attached or isolated in the channel. Unlike
expanse of slower and higher water extending abutments or piers, however, wood structures
upstream from the obstruction. The backwater typically have a level of porosity that has an
effect can result in higher water surface important controlling influence on the flow field
elevations along the banks and, in and the diversity of hydraulics generated
unconstrained reaches, enhanced floodplain (Manners et al. 2007).
connectivity with an increased volume of water
spilling out onto the floodplain. The ability of The flow obstruction created by the wood
large wood to alter water levels and influence creates steep hydraulic gradients in all
habitat varies based on local conditions, dimensions where flow depths and velocities
including the volume of assembled wood and its can rapidly change from a local maximum to
size relative to channel morphology. zero over a short area. Pressure gradients
created by the structure can generate
Though not uniform to all systems, research has downwelling, horseshoe vortices, separation
shown that large wood in the channel has a zones, wake eddies, and levels of turbulent
small to insignificant effect on the duration or scour, nutrient mixing, and oxygenation that
frequency of large flood events (approximately

Large Wood National Manual July 2015


3-4
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

would not occur in channels with otherwise island that bifurcates flow. Because large wood
subcritical reach average conditions. The structures are often fixed in location for long
manner in which the wood structure influences periods, bedforms created are often stable
flow also changes with discharge, creating features relative to ones not linked to flow
variability in hydraulic patterns over the obstructions that are more prone to migration,
entirety of the hydrograph. such as bar-pool morphology in a meandering
channel.
3.2.1.3 Substrate Composition
Overall, large wood can be quite effective at CROSS-REFERENCE
sorting sediment and channel substrate,
creating a diversity in sediment texture Chapter 4, Geomorphology and Hydrology
available as habitat for aquatic life. Local areas Considerations, provides an indepth examination of
of flow, convergence, and divergence are the ability of large wood to significantly alter channel
typically associated with large wood that results morphology.
in spatially variable shear stress with
corresponding variability in sediment texture. Large wood often creates and maintains pools
Fine sediment can be scoured away to expose important to the different life stages of aquatic
coarser substrate suitable for spawning, while organisms. Channels located in forested reaches
in other areas sediment deposition and a (particularly in old growth forests) have
reduction in sediment texture can occur. significantly more pools per unit length than in
unforested reaches. The specific pool spacing
Research has shown that up to 60% of the total
for a given wood frequency can be quite
bankfull shear stress in a channel can be spent
variable due to regional and site-specific
on form drag caused by large wood (Manga and
differences in channel type and wood
Kirchner 2000). This means less shear stress is
characteristics (Montgomery et al. 2003).
available for transporting sediment, and stream
Research has shown that as wood loading
competence declines (Montgomery et al. 2003).
increases there is an increase in pool frequency
Consequently, the median surface grain size of
that begins to level off at wood loadings of
the bed near large wood can be up to 90% finer
about 0.03 piece per square meter (Buffington
than what it would be in a wide, planar channel
and Montgomery 1999b; see Figure 4-14a).
without large wood (Buffington and
Montgomery 1999b). Large wood can be Depending largely on its orientation and
effective at promoting deposition of gravel in position above the bed, the type of obstruction
reaches otherwise too coarse or armored to formed by wood can create many different pool
provide spawning habitat for salmonids. types, including plunge, underscour, eddy, and
dammed pools (Montgomery et al. 2003). The
3.2.1.4 Channel Morphology importance of wood size and its ability to create
deep pools is illustrated by research that shows
The shape and characteristic features of the the number of stream pools with residual
stream channel (channel morphology) affect the depths >0.5 meter (1.6 feet) increases rapidly
quantity and quality of habitat for fish and other with riparian forest stand age, diminishing only
species. The ability of large wood to after stands reach ages of more than 200 years
significantly alter channel morphology at the (Rot et al. 2000; see Figure 4-13a).
unit and reach scales is well-documented.
Morphologic effects can range from a single Sediment bars typically form in conjunction
rootwad partially embedded in a channel with the pools created by large wood. Wood can
causing enough of a flow obstruction to scour a act as a dam that impounds water and forces
pool, to a large logjam capable of creating an upstream sediment deposition, similar to the

Large Wood National Manual July 2015


3-5
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

process of sedimentation in a reservoir behind jam located in the low-flow channel will
a dam. Bar formation also occurs in flow increase the sinuosity of the low-flow channel
separation and deposition areas downstream of as flow is forced around the obstruction. Wood
the zone of flow convergence where pools are jams of sufficient size that occupy enough of the
scoured. Flow acceleration at large wood channel width to significantly constrict flow can
accumulations can also create riffle habitats as deflect flow into the opposite bank and cause
part of the bar-unit complexes. Much like pool bank erosion and undercut bank habitat much
types associated with large wood, the type (e.g., the way a point-bar develops in association
bank-attached, mid-channel) and size of bar with outer cut-bank erosion. Strategic
formed can be quite variable (see Chapter 4). placement of multiple wood jams (often on
Wood accumulations of sufficient size and alternating sides of the channel) can promote
stability can create a large enough flow enough flow redirection, bank erosion, and
obstruction with subsequent sediment channel migration to increase the overall
deposition to create new bars or enlarge channel morphology’s sinuosity.
existing bars. Racking of additional woody
In forested channels, wood can be a primary
material on jams formed at bar apexes can
driver in bifurcating or splitting flow, creating
enlarge the jam and enhance its hydraulic
channel avulsions, and creating anabranching
influence and stability to a level where enough
rivers that may otherwise be braided or single
sediment accretion occurs to ultimately form
thread meandering channels (Collins et al.
vegetated channel islands that support new
2012; see Figure 4-8). Using channel bank
riparian habitat (Abbe and Montgomery 1996).
length as a metric for edge habitat, it is
Large wood has the ability to not only create apparent that an unconfined anabranching
localized habitat unit features, but to also channel reach has significantly more habitat
transform channel morphologic types. In low- than incised and leveed reaches of the same
order headwater streams, large wood can river (Chapter 4). Large wood is also important
create step-pool morphology with plunging for forming and sustaining side channels that
flow important for oxygenation, and trap can be wetted at low-flow or only during floods
enough sediment to develop an alluvial bed in (Abbe 2000).
what would otherwise be a bedrock channel.
Likewise, large wood can form pool-riffle 3.2.2 Aquatic Food Webs
morphology in reaches that would otherwise be
plane bed or bedrock (Montgomery et al. 2003). A food chain is the linkage between primary
In fact, it can be rare to observe pools and bars resources (plants, detritus) and secondary
in moderate-gradient (i.e., >0.01) cobble and consumers (e.g., insects and fish) (Pianka 1994).
gravel-bed forest channels not formed or A network of linked food chains forms a food
influenced by wood (Montgomery et al. 2003). web, and stream food webs are among the most
complex. Like most ecosystems, aquatic
3.2.1.5 Planform Change foodwebs begin with the capture of energy from
the sun that is fixed by terrestrial and aquatic
Planform refers to the shape of the channel as plants via photosynthesis. This energy is stored
viewed from above, including sinuosity, side in the tissue of the plant where it is available to
channels, oxbows, and other features affecting secondary consumers.
the type and amount of habitat for species. The
flow obstructions created by large wood The food chain that supports fish and
accumulations can dramatically alter channel invertebrate production in streams is based on
planform, increasing channel length and two photosynthetic energy sources: terrestrial
sinuosity. The obstruction created by a wood organic matter (leaves, twigs, branches, and

Large Wood National Manual July 2015


3-6
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

large wood) that enters streams from the can grow may limit the potential production of
riparian forest and upstream watershed algae in the stream. While instream wood often
(Cummins 1974), and algal production in the has limited direct influence on light or on
stream itself (Cummins et al. 1984). Leaves and nutrients, it can be a particularly important
woody material are largely composed of substrate on which algae can grow in sand-bed
cellulose that is broken down and made streams or in systems with unconsolidated
available to other organism by bacteria and streambed material. Wood addition can
fungi (Webster and Benfield 1986). Leaves and indirectly influence nutrients by changing water
detritus are rapidly colonized by bacteria and transport times, which changes the ability of
fungi that begin to break them down. Aquatic microorganisms in the stream to remove and
invertebrates like mayflies, stoneflies, midge regenerate nutrients (Ensign and Doyle 2005).
larvae, and scuds (freshwater shrimp) shred Also, the source of wood in streams can alter
leaves and feed on terrestrial detrital inputs. light availability in forested streams—either
Much of the carbon they assimilate comes from because wood is provided by riparian trees that
aquatic bacteria and fungi colonizing the are cut or pulled down and therefore create
detritus rather than the detritus itself. Grazing canopy gaps or because wood placement
invertebrates—commonly mayfly and requires the removal of riparian vegetation
chironomid nymphs—also feed on algae on around the placement areas in order to bring in
stream rocks. These algal communities often logs from outside the system.
occur as thin (almost invisible) algal layers on
While production of algae depends on light and
rock or wood, but they may support
nutrient availability in stream water (e.g.,
considerable grazer production because of high
nitrogen and phosphorous), production of
algal turnover rates (McNeely and Power 2007;
invertebrates that feed on detritus (insect
Coe et al. 2009). These aquatic invertebrates (as
shredders, collector-gatherers, and filter-
well as terrestrial insects that fall onto the
feeders) depends strongly on retention of
stream surface) are the primary food source for
detritus in the stream (Cummins 1974; Wallace
stream fishes like juvenile salmonids in
et al. 1997). Deposition and storage of organic
headwater streams, although many fish feed
matter takes place in slow-moving backwaters,
directly on algae (e.g., stonerollers), detritus
stream margins, above dammed pools (Wallace
(suckers), or other fish (e.g., pike, bass, adult
et al. 1995b), in openings between rocks, and on
trout).
the downstream side of obstructions like
Production of algae depends largely on light and boulders and large wood. Simplified stream
nutrient (i.e., nitrogen and phosphorous) channels with minimal structure tend to act as
availability in the stream. These are the two flumes that transport material downstream
main “bottom-up” factors influencing stream with limited local benefit. In contrast, the
primary production (in contrast to “top-down” physical complexity associated with boulders
effects of grazing invertebrates and fish). In and large wood greatly increases retention of
forested headwaters light may be particularly organic detritus, and therefore populations of
limiting, with nutrient effects on primary invertebrates that feed on it (Bilby 1981). Wood
production manifesting only after light in particular can greatly enhance organic matter
limitation has been alleviated (Sabater et al. deposition by trapping fine branches and leaves
2005; Bernhardt et al. 2007; Ambrose et al. to form debris dams that enhance invertebrate
2004). In mid-order streams and streams with production (Wallace et al. 1997).
more limited riparian shading, nutrients are
In streams with anadromous species such as
commonly the key factors limiting primary
salmon, an additional source of energy and
production. And in other systems the
nutrients is the carcasses of spawned adult fish.
availability of stable substrates on which algae

Large Wood National Manual July 2015


3-7
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

This is particularly the case in systems with fish (Schenk et al. 2014). The multiple habitats
Pacific salmon die soon after their single affecting species interactions include backwater
spawning event (semelparity). Because salmon pools, side channels, and eddies, and have
acquire most of their adult biomass during their structural and hydraulic diversity near the
ocean residency, salmon carcasses potentially stream margins (Naiman et al. 2002c).
supply many tons of marine-derived nutrients Sometimes changing the position or removing
and biomass to otherwise nutrient poor large wood will decrease the habitats in which
systems (Cederholm et al. 1999; Wipfli et al. predators and prey can hide. Research has
2003) Large wood traps carcasses of spawning found that the volume of large wood in streams
salmon, thereby preventing their export and can be associated with the density of fish
retaining marine-derived nutrients in populations (Murphy et al., 1986). Invertebrate
headwater reaches where it can increase local predator biomass increases when there is large
production of fish and other biota (Cederholm wood present in streams, and, in general, the
et al. 1999). Marine-derived nutrients have invertebrate communities of predators are
been found to enhance biofilm development, more productive per unit of biomass following
macro-invertebrate production, and overall the introduction of woody debris (Naiman et al.
stream productivity (Wipfli et al. 1998); and 2002c).
may be transferred to wildlife and terrestrial
vegetation (Quinn et al. 2009), including even 3.2.3 Biogeochemical
wine grapes (Merz and Moyle 2006). Large
instream wood traps carcasses, allowing them
Functions
to be processed locally by birds, insects, fish, Large wood plays a key role in nutrient cycling
bacteria, and fungi (Cederholm et al. 1999). in streams (Bilby and Bisson 1998). In general,
Nutrients from salmon carcasses can also wood itself is a poor carbon source. The amount
enhance decomposition of detritus, resulting in of nitrogen and phosphorous relative to carbon
a synergistic effect of large wood to trap and is low, and the lignin in wood is particularly
hold both forms of organic input (Bretherton et difficult for many organisms to break down
al. 2011). (Webster and Benfield 1986). In temperate
Atlantic salmon (Salmo salarin) in eastern ecosystems, few macroinvertebrates or fish eat
North America and northern Europe may wood directly, but there is a suite of microbes
spawn multiple times (iteroparity); therefore, and fungi that break down wood, which, in turn,
there is less mortality on spawning grounds form food for benthic invertebrates and other
than there is in Pacific salmon streams. biota (Webster and Benfield 1986; Findlay et al.
However, transfer of nutrients and energy from 2002; Spanhoff and Cleven 2010). The stream
marine to freshwater via Atlantic salmon macroinvertebrates that do eat wood tend to
carcasses has been demonstrated (Jonsson and eat smaller particles and/or they ingest wood as
Jonsson 2003; Williams et al. 2009), and it is a byproduct of feeding on microbial biofilms on
reasonable to assume that trapping of Atlantic wood surfaces (Johnson et al. 2003; Coe et al.
salmon carcasses by large wood could be 2009). The rate of wood decay by microbes and
important for enhancing nutrient transfer from fungi varies by species. As a rule, trees with
marine to freshwater systems. more nitrogen per unit of carbon (such as
alders maples, and poplars) decay faster that
Wood also affects predator-prey dynamics. For those with lower nitrogen to carbon ratios
example, large wood creates cover for prey and (such as oaks, firs, and spruce) (Spanhoff and
provides substrate for algae, microorganisms, Meyer 2004). As a broad generalization,
and invertebrates. Wood also creates cover for hardwoods decay faster than softwoods
predatory fish that eat invertebrates or other (Webster and Benfield 1986). Slower decay can

Large Wood National Manual July 2015


3-8
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

influence wood persistence, which in turn Studies removing instream wood have found
influences wood function. In many eastern variable effects on nitrogen processing. Webster
United States streams, for example, large wood et al. (2000) found that wood removal reduced
from American chestnut (Castanea dentata) nitrogen processing, while Warren et al. (2014)
remains highly functional even though large found that wood removal enhanced nitrogen
chestnut trees have been essentially lost from processing. The authors of the latter study
eastern forests for over 80 years (Hedman et al. attributed their result to changes in substrate
1996). Nutrient availability in the stream also composition and algal production associated
influences wood persistence; if nutrients are with the removal of sediment from rocks
added to a system biological breakdown of the around and upstream of the dam. This
wood can occur much more rapidly (Spanhoff conclusion was based in part on a study in
and Meyer 2004). northern Michigan sand-bed streams that found
that modification of stream substrates by scour
Large wood can enhance stream nutrient
around added wood can enhance nutrient
cycling in multiple ways. First, large wood
uptake by exposing stable substrates for algal
retains leaf litter and fine particulate organic
growth (Holleine et al. 2007). These results
matter. The breakdown of this organic matter
highlight how the overall role of wood in stream
by microbes and fungi creates an elevated
nutrient dynamics will vary depending upon the
demand for nutrients, especially nitrogen and
physical characteristics of the system and how a
phosphorous. This elevated demand increases
given wood structure modifies those (i.e., does
the rate at which nutrients are taken up from
it expose via scour large stable substrates
the water column and increases the retention of
where algae can grow or does it cover them by
nutrients in the stream (Mulholland et al. 2009).
enhancing sediment deposition around the
Second, when channel-spanning wood and
dam?). Studies exploring empirical
wood jams retain a combination of organic
relationships between natural wood addition
material and fine inorganic material they can
processes and phosphorous cycling (rather than
create areas of saturated sediment behind and
experimental manipulations) have found
around the wood where oxygen can be locally
significant associations between wood and
depleted. Under these anaerobic conditions
phosphorous demand, suggesting that as wood
available nitrogen can be converted to nitrogen
loading increases the capacity of streams to
gas through a process referred to as
process phosphorous also increases (Valett et
denitrification (Steinhart et al. 2000). This
al. 2002; Warren et al. 2007).
conversion is highly variable across streams
and across regions but it can be an important
loss of nitrogen from these systems, especially 3.3 Hyporheic Zone
in areas of the northeastern and Midwestern
United States where excess nitrogen pollution is Much of our consideration and scientific study
a particular concern. The creation of pools and of streams focuses on the visible components of
the modification of stream flow that directs water, channel, wood, and biota. The hyporheic
water into subsurface areas also leads to zone extends the river below what we see to
increased uptake and retention of nutrients include the “sponge” of saturated substrate
(Ensign and Doyle 2005). Wood directly where water can regularly exchange between
supports fungi and bacteria, and wood is often a surface and subsurface flows (Stanford and
surface on which algae grow in streams. Ward 1993). The hyporheic zone is defined as
Collectively the biofilm that lives on wood can the water-saturated sediment volume below the
be an important area for nutrient uptake as well stream bed and adjacent stream banks where
(Sobota et al. 2007). mixing between surface water and groundwater
occurs (Bencala 2005). The hyporheic zone has

Large Wood National Manual July 2015


3-9
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

several ecological functions and is the source of interstitial spaces of the substrate (Lautz et al.
summer base flow in many systems. Depending 2007). In a sand-bed flume the introduction of
on geological and soil conditions, the hyporheic wood produced irregular bedforms, increased
zone may extend only a few centimeters or flow resistance, and increased vertical water
30 meters (100 feet) or more into the adjacent flow across the streambed, which caused
floodplain (Hinkle et al. 2001; Boulton et al. surface water to mix deeper into the hyporheic
2010). zone (Mutz et al. 2007). However, Stofleth et al.
(2007) found that hyporheic storage was an
Although the hyporheic zone may be shallow, it
insignificant percentage (less than 0.5%) of
can also be extensive because it extends from
total hydraulic retention in sand-bed streams
the uppermost headwaters through the
and that it did not increase with the addition of
lowermost reaches of rivers and into the
flow obstructions. Their findings suggest that
estuarine zone (Krause et al. 2014). The
hyporheic zone biogeochemical processing in
hyporheic zone is a hydraulic feature but it is
these lowland streams may not be significant.
also a biological habitat for microbes,
Lautz and Fanelli (2008) found primarily anoxic
invertebrates, insect eggs and pupae, fish eggs,
zones in pools upstream of log restoration
and fish embryos and is therefore a key
structures and oxic zones downstream in a
consideration in the biological and ecological
turbulent riffle.
function of streams (Stanford and Ward 1995).
In the hyporheic zone, surface water and Hyporheic flow and the exchange with surface
solutes exchange into and out of the stream bed, water are complex, and wood can enhance that
having mixed with groundwater to varying complexity. Hester and Doyle (2008)
extents. Numerous biogeochemical reactions investigated instream geomorphic structures
occur in this zone, and it can influence such as debris dams and wood-associated steps.
mineralization, major ions, nutrients, and They found that hyporheic exchange flow was
contaminants (Bencala 2005; Gandy and Jarvis influenced most strongly by structure size,
2006; Mulholland and Webster 2010; Krause et background groundwater discharge, and
al. 2014). Hyporheic flow also has localized sediment hydraulic conductivity with lesser
influences on stream temperature and influences from geomorphic structure type,
dissolved oxygen. depth to bedrock, and channel slope. Debris
dams can also exchange seasonal variations in
The key ecological role of the hyporheic zone
hyporheic flow and associated nutrient
argues for its consideration in stream
processing within this section of the stream
restoration and large wood placement projects
(Claussens et al. 2010). Sawyer et al. (2011,
(Hester and Gooseff 2010). Large wood can
2012) found downwelling water upstream, and
affect formation and maintenance of the
upwelling water downstream, of channel-
hyporheic zone and the flux of water between
spanning logs with distinctive temperature
the stream and the hyporheic zone, although
effects. In a meadow stream Sawyer and
the effect varies between streams of different
Cardenas (2012) found that large wood
geology. Lautz et al. (2006) showed that log
addition increased hyporheic flow, and that
dams in a semi-arid stream increased hyporheic
hyporheic return flow locally stabilized stream
interactions by slowing stream velocity,
water diel temperature fluctuations, although
increasing flow complexity, and diverting water
only at a local scale (creating refuge habitat
to the subsurface. Debris dams slowed water
rather than whole-stream effects on
upstream causing localized fine sediment
temperature). But they also found that the
deposition so that sediments immediately
nature of hyporheic exchange could limit the
downstream contained less fine sediment and
influence of wood on flowpaths. The influence
had higher capacity to allow water through
of wood on hyporheic exchange (as with

Large Wood National Manual July 2015


3-10
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

assessments of biota) also warrants a long-term typically have a relatively high gradient and
perspective. Wondzell et al. (2009), for ample alluvium, and, west of the Cascade Range,
example, investigated the responses of a small high precipitation with dense forests. In these
low-gradient stream to large wood removal. streams, wood creates pools and steps, traps
They found that hyporheic exchange flow sediment and organic debris, and provides
declined in the first few years. Subsequently, cover and protection from predators.
however, the decline reversed as pool-riffle
Wood in this region plays an important role in
patterns developed and enhanced hyporheic
creating and enhancing habitat for salmonids in
exchange flow.
particular. A number of studies have
documented benefits of wood additions for
3.4 Regional Differences juvenile coho salmon (Oncorhynchus kisutch)
and juvenile Chinook salmon (Oncorhynchus
in Large Wood Ecology tshawytscha), which may rear in streams for
one to two years before migrating to the ocean
The biological and physical roles of large wood
or estuary to grow and mature (see reviews by
in streams discussed above apply generally to a
(Smokorowski and Pratt 2007; Roni et al. 2008;
wide range of geographies and stream types.
and Nagayama and Nakamura 2010). Benefits
However, there are important regional
to fish were largely associated with increasing
differences in wood effects based on differences
pool habitat or habitat complexity overall (as
in geology, climate, and species. Large wood
noted above and elsewhere in this manual).
provides different ecological functions in steep
Stream wood in the Pacific Northwest is often
gradient and gravel-bed streams typical of the
quite large because the region has retained a
western United States than in low-gradient,
good deal of its old-growth forest relative to
sand-bed rivers more typical of the southern
other regions of the country. In addition,
United States. Similarly, wood may also provide
climate conditions in the coastal and western
different ecological functions for warmwater
Cascade mountain ranges are such that tree
versus coldwater fish communities. Effective
productivity is often quite high, which leads to
application of large wood restoration means
relatively rapid development of larger trees.
understanding the diverse and context-specific
Although a good deal of old-growth forest
functions of large wood across diverse
remains in the Pacific Northwest compared
landscapes.
with other regions of the country, much of the
region has undergone (and continues to
3.4.1 Western United States undergo) extensive forest management. Early
The majority of research on stream wood and forest management often used splash dams to
wood additions has been conducted in the move wood downstream. These dams scoured
Pacific Northwest of North America. Large away not only wood but also much of the
wood plays a dominant role in the physical and stream substrate. Wood addition in these areas
biological nature of streams of western North functions not only to enhance pools and
America (Bilby and Bisson 1998). In particular, increase habitat complexity, but it is also added
wood provides habitat characteristics that are in many cases to help promote channel
beneficial to multiple life stages of anadromous aggradation and the development of spawning
and non-anadromous salmonids. As a result, the habitat for anadromous salmon.
addition of wood is the predominant habitat Anadromous salmon are key species in many
restoration action undertaken by federal, state, western streams. They are not only a dominant
tribal, and local agencies and stewardship fish species for management, but, as discussed
groups intent on enhancing these species in previously, they also have a key ecological role
particular (Roni et al. 2014a). Western streams

Large Wood National Manual July 2015


3-11
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

and provide an important nutrient subsidy to found little evidence for a long-term positive
the aquatic ecosystems in which they spawn effect on habitat—attributed to a limited long-
(Bilby et al. 1998; Stockner 2003; Wipfli and term geomorphic effect. Wood was likely to
Baxter 2010). Large wood retains salmonid have been more important in this region in the
carcasses so that they can be processed locally past when larger trees were adding larger
rather than flushed downstream. In the absence wood. When remnant old-growth forests do
of wood, carcasses may be flushed from a remain along streams in the northeastern and
system during high flow, thereby removing the central Appalachian mountain regions, wood
nutrient and carbon subsidies that they provide. size and total wood volumes can rival those in
Wood is instrumental in keeping carcasses in a old-growth forests from other regions (Keeton
stream and thereby maintaining their subsidy et al. 2007; Warren et al. 2009).
function in the ecosystem.
Stream wood and wood jams are important in
these mountain ecosystems in carbon retention,
3.4.2 Northeastern United and a good deal of the early and classic work on
States the importance of particulate organic matter
retained by wood has been conducted in these
In the northeastern United States and uplands streams (Fisher and Likens 1972; Bilby 1981;
of the mid-Atlantic region, studies evaluating Wallace et al. 1997)
wood function have been more variable. Wood
can be an important explanatory factor in Atlantic salmon have been largely extirpated
accounting for variability in stream trout from the rivers and streams in the northeastern
abundance (Kratzer and Warren 2013). When United States where they were historically
wood was assessed as a habitat feature in abundant. Based on work in the Pacific
Appalachian mountain streams, its use was Northwest, wood and wood accumulations
disproportionate to its availability (Flebbe were likely to have been important in carcass
1999). However, other studies have found retention functions for Atlantic salmon and
mixed results in assessing the influence of wood other anadromous stream species in these
jams and large wood structures on stream regions. Although Atlantic salmon are often
salmonids in the northeastern United States capable of repeat spawning, stress associated
(Warren and Kraft 2003; Thompson 2006), with spawning migration can elevate mortality
especially when long-term processes are rates.
considered (Warren and Kraft 2002). The
variable function of wood in these systems may 3.4.3 Midwestern and
be attributed in part to the forest management
history of this region (Williams et al. 2009).
Southeastern United
Unlike in the Pacific Northwest, very few areas States
of old-growth forest remain in the northeastern
In the Midwest, stream gradients are lower
United States, especially near streams. The large
across much of the region. In addition to its
wood in northeastern streams is therefore
function in creating pools and enhancing
generally smaller than in other regions. The
habitat complexity, wood also plays a key role
amount of wood exceeding 30 centimeters
as a stable substrate in many low-gradient
(12 inches) in diameter (“large logs”) has been
streams that dominate the coastal plains of the
tied to stream pool habitat in this region, but
southeast and regions across the Midwest
wood volume using a smaller size threshold
(Smock et al. 1989; Stewart et al. 2012).
(10–centimeter [4-inch] diameter) has not
(Keeton et al. 2007). Assessments of historic Many of the streams in the Midwest and
wood addition structures by Thompson (2006) southeast drain relatively low-gradient

Large Wood National Manual July 2015


3-12
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

landscapes and harbor largely warmwater fish streams. They assessed the abundance of fish
communities. The historic reference condition before wood addition and up to 6 years
and function of large wood in low-gradient and afterward and found significant increases in fish
warmwater streams is generally less well abundance. This work was then followed up by
documented here than in higher gradient White et al. 2011, who found that the effects of
systems in the northeast and northwest, but wood loading on fish persisted. Fish abundance
tends to indicate a lesser control by wood over at the treatment sites remained well above
channel structure and bedform in very low- those in the reference sites 20 years after wood
gradient streams than in gravel-bed channels addition. Wood is also important in
(Wohl and Merritts 2007; Walter and Merritts mountainous regions of the southwestern
2008). Sand-bed rivers of the southeastern United States. Wood in these systems functions
United States contrast strongly with gravel-bed to create and enhance habitat and as in other
rivers of the Pacific Northwest; however, large regions it can be important for litter retention
wood in sand-bed rivers was found to be in streams (Trotter 1990).
significant as both structure for fish and as a
relatively rare stable substrate in a sand-bed
environment. Benke et al. (1985) found that 3.5 Considering the Need
filter-feeding invertebrates like caddisfly for Wood Placement
(Trichoptera spp.) and blackfly (Simuliidae
spp.) larvae only occurred on stable substrate When restoration of wood in streams is
and were consequently severely habitat limited, undertaken to achieve a biological goal, the
achieving their highest densities on large wood focus is frequently an increase in the abundance
(“snags”); although large wood only accounted of desired fish species. For instream wood
for 6% of substrate by area, it accounted for restoration to be effective in achieving this goal,
50% of invertebrate biomass, highlighting the wood, habitat diversity, and cover should be
importance of large wood on the productive identified as limiting factors for fish and their
capacity of sand-bed rivers, above and beyond associated biological communities. An intuitive
its role in providing cover for fish. (if not always practical) measure of success of
stream restoration programs in this context is
3.4.4 Mountain West and generally a clear and persistent increase in the
abundance of desired species over time.
Southwestern United
Decisions to invest in restoration of large wood
States (or indeed any environmental attribute) need to
In the Mountain West, the frequency of fires is be based on clear assessment of factors limiting
higher than in the northeastern, midwestern, or a desired species or process. Restoration of
Pacific Northwest regions of the country. Wood large wood at specific sites can have limited
loading occurs as a result of individual mortality biological value if other factors are more
of trees but is also often the result of these large limiting at a watershed scale (Nagayama and
disturbance events. (Richmond and Fausch Nakamura 2010). The factors that limit
1995). Much of the research on wood function population productivity need to be understood,
and biota in western systems has focused on at least in a qualitative sense, before designing a
fish (Schmetterling and Pierce 1999; O'Connor restoration project. A number of models are
and Rahel 2009; White et al. 2011). One of the discussed in Section 3.5.2, Linking Habitat to
longest running assessments of stream wood Fish Population Dynamics, which can assist in
addition was done in a system in Colorado, the assessment of needs for wood enhancement
where Gowan and Faush (1996) added channel- relative to other possible limiting factors.
spanning wood to a series of headwater

Large Wood National Manual July 2015


3-13
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Habitat factors that limit fish production are densities of juveniles with a scarcity of larger
often described as “bottlenecks” that constrain subadults and adults may indicate juvenile
particular life stages in specific locations rearing habitat limitation, even if there is
(Kennedy et al. 2008). Habitat bottlenecks and abundant suitable adult habitat (Armstrong and
limiting factors operate hierarchically. Factors Nislow 2006; Rosenfeld 2014). However,
such as temperature operate at watershed or inferences with respect to habitat limitation
sub-watershed scales and exercise a pervasive should be made with care because factors
impact on growth and survival of biota, while unrelated to habitat—such as low marine
factors such as large wood operate at more survival of adult salmon, leading to low
localized scales. Consideration of this hierarchy spawner returns—can also lead to under
is key to understanding the potential recruitment (low egg deposition and juvenile
effectiveness of restoration measures, many of density), even when spawning habitat is
which are designed in response to localized abundant and not limiting the population.
conditions that are controlled by larger scale
In general, restoration that does not directly
factors.
address the factors limiting the growth,
Water quality parameters represent systemic abundance, or survival of fish will be ineffective
factors that control the effectiveness of in meeting management goals geared toward
restoration efforts. For example, if stream pH or increasing the overall target fish population. For
water temperatures are marginal or this reason, management interventions that
unfavorable for a target fish species (i.e., too increase available habitat for multiple life
high or too low for salmonids), then investing in history stages are ideal, particularly if there is
restoration actions may be unwise until the uncertainty concerning which life stage is most
systemic issues are addressed. Similarly, limiting. This is one of the advantages of large
extremely high nutrient loads leading to high wood addition because the creation of
levels of primary production and nuisance algal complexity usually increases overall habitat
species (eutrophication) and low oxygen could diversity for multiple life stages. For instance,
also limit the target species, and may need to be carefully planned large wood additions could
addressed before large wood can effectively enhance spawning habitat by trapping gravel
enhance habitat. wedges above, or depositing gravel bars below,
engineered jams, while simultaneously
If water quality in a stream is within the optimal
increasing the availability of low velocity
range for a target species, then availability of
marginal juvenile rearing and overwintering
suitable habitat, including refuges from
habitat, creating deeper scour holes for larger
predation and high flow events, may become
fish, and increasing organic matter retention
the dominant factors limiting population size
that may also benefit overall prey production.
(Breau et al. 2011; Reeves et al. 1989; Nickleson
Nevertheless, limiting habitat factors should
and Lawson 1998). Limiting bottlenecks will be
still be assessed to the extent possible because
present when habitat limitation only affects
creation of one type of habitat may result in
particular life history stages. For instance, lack
reduced abundance of another habitat type that
of adequate spawning habitat may prevent
could negatively impact species with
sufficient egg production to saturate available
contrasting habitat requirements. For instance,
rearing habitat, resulting in juvenile rearing
if riffle habitat is very limited, riffle-dependent
habitat that is under-seeded (below capacity).
species could be negatively impacted by
Very low densities of juveniles, despite back-flooding from large wood jams or other
abundant suitable rearing habitat, can be used restoration structures, potentially reducing the
as a diagnostic to infer spawning habitat (i.e., availability of scarce riffle habitat and any
recruitment) limitation; similarly, very high species dependent on it.

Large Wood National Manual July 2015


3-14
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Distinguishing between the effects of quantity, and connectivity are reflected in the
restoration on habitat quantity and quality is survival and abundance of fish at life stage and
useful for understanding population responses population scales (Schlosser and Angermeier
to habitat change. As discussed in the following 1995; Hayes et al. 1996). Sustained production
section, these habitat measures are related by a species requires a network of complex and
respectively to the biological capacity and interconnected habitat patches spanning the
productivity of fish populations. Restoration species’ life history (Williams 2006).
can increase the number of fish by either
The quality of habitat in each patch and
increasing the area of available habitat for a
competition for limited resources result in an
limiting life stage, or by improving the quality of
overall survival that reflects both density-
available habitat, so that fish will experience
independent and density-dependent survival
higher growth and survival and more will
factors. Density-independent factors refer to
recruit to the next age class.
attributes, such as temperature, that affect
If a management intervention such as large survival regardless of fish abundance
wood restoration does not increase either the (notwithstanding the relationship between
area of habitat limiting a life stage or the habitat pathogens, temperature, and fish density in
quality for a key life stage (leading to better disease outbreaks). Other survival factors, like
growth and survival), then there will be no food and cover, are consumable and in limited
population response. In principle, population supply, and their per capita availability declines
responses to restoration should be greatest as density increases. The area of habitat and
when habitats are present in an optimal ratio food availability determine the capacity of the
where no single habitat becomes a severely environment to sustain a given species
limiting bottleneck (Reeves et al. 1989; (Chapman 1966). As density increases,
Rosenfeld 2014). Although the diversity of density-dependent factors limit survival and/or
habitats associated with large wood restoration growth, and abundance approaches carrying
often has the capacity to help minimize habitat capacity of a species in a particular habitat
bottlenecks, there are some geomorphic (Hayes et al. 1996). Carrying capacity is linked
contexts where large wood may not be the most to survival because it relates to the quantity of
effective restoration intervention (e.g., steep available resources such as food and space that
colluvial boulder channels where wood has place an upper bound on density of a life stage
minimal impact on channel structure). in any particular environment (Chapman 1966).

For salmon and other fish, the population


3.5.1 Fish Population dynamics of density dependence are often
Dynamics and depicted by stock-recruitment relationships
(Hilborn and Walters 1992). There are several
Instream Wood forms of these functions, but all portray the
Fluvial environments are a mosaic of channel relationship between the number of spawners
units with differing conditions that are (stock) and resulting progeny (recruits). Figure
perceived in unique ways by species and life 3-1 shows the stock-recruitment relationship
stages as habitat patches of varying quality area developed by Beverton and Holt (1957) that is
(Winemiller et al. 2010). As discussed above, commonly used in salmonid fisheries
large wood can play an important role in management. Because of it tractable
determining the array of habitat patches across mathematics, the Beverton-Holt relationship is
the riverscape. Connectivity between habitat used in many of the models that relate habitat
patches in time and space allows a species to to potential fish production. The relationship
complete its life history. Habitat quality, has two parameters: productivity or survival

Large Wood National Manual July 2015


3-15
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

(progeny/spawner), at low abundances where Productivity is the density-independent


density effects are absent, and carrying survival, which, along with density-dependent
capacity, the maximum abundance of the factors of the environment, determines
species possible given the available habitat; abundance limited by the total capacity of the
density-dependent effects typically limit environment. Replacement is the minimum
populations as they approach habitat carrying number of spawners required to maintain a
capacity. Productivity and capacity can be given abundance. Under steady-state
related to the quality and quantity of habitat, environmental conditions, the population
respectively (Reisenbichler 1989; Hayes et al. abundance equilibrates at Neq, the point where
1996). The diagonal line in Figure 3-1 is the abundance crosses the replacement line.
replacement line where the number of
spawners equals the number of progeny 3.5.2 Linking Habitat to Fish
(productivity = 1.0). At abundances greater
than replacement, the population will increase,
Population Dynamics
while it will decrease when abundance falls Analytical models provide managers the ability
below replacement. As long as the productivity to link characteristics of the environment to fish
is greater than 1.0 (replacement), the population dynamics in order to assess limiting
population abundance will increase, although at factors and evaluate alternative restoration
a declining rate due to density-dependent strategies. Models serve a role in planning by
survival factors (e.g., increased competition for helping managers address issues such as
food and space) that are in limited supply identifying factors currently limiting species
(Figure 3-1). Under steady-state conditions, the and populations of concern, determining
abundance of a fish population will equilibrate whether restoration of large wood is likely to
where the ratio between spawners and progeny address these factors, understanding how other
is one designated Neq in Figure 3-1. Neq is factors may augment or limit the value of large
calculated from the productivity and carrying wood restoration and quantifying reasonable
capacity and thus serves as a useful summary expectations of the biological benefits of
metric of the population response to both restoring wood. Models can thus help managers
habitat quality (productivity) and habitat invest restoration dollars wisely to meet
quantity (capacity). management objectives. Models also have a key
Figure 3-1. Features of a Beverton-Holt role in adaptive management where they can
Production Function generate testable hypotheses and synthesize
information from monitoring and research
programs. However, models are never a
substitute for monitoring, assessment, and
research and, to be effective, the limitations of
each model need to be understood by
practitioners and users of model results. Models
can structure information from monitoring
programs, restoration assessments, and
research in a form that is useful to managers
and funding agencies and can increase the value
added from assessment and monitoring
programs. By providing an explicit, logical
framework that incorporates available data and
knowledge, models can provide accountability
for restoration priorities and funding decisions.

Large Wood National Manual July 2015


3-16
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Roni et al. (2014b) conclude from their review programs test the fundamental assumptions in
of the effectiveness of stream restoration the models and reflect changes in habitat
actions that models can “help to set realistic conditions over time, leading to a refined basis
expectations for restoration outcomes and help for making decisions regarding investments in
managers choose among alternative restoration restoration of large wood or other measures.
scenarios.” A key factor in the successful
application of models to assess habitat
GUIDELINES
restoration needs is to understand the model’s
purpose, limitations, and data requirements.
Objectives of Species-Habitat Models
The value of the results generated by a model
reflects the quality of the data used in the model 1. Formalize our current understanding of the
and the validity of its underlying relationships. habitat requirements of a species.
To be effective, models need to be continually
2 Understand how environmental factors affect
tested against empirical and experimental data the distribution and abundance of a species.
to provide an adaptive platform to guide
restoration based on the available science 3. Predict further distributions of a species.
(Boisclair 2001). 4. Identify weaknesses in our understanding.
In the context of planning and prioritizing 5. Generate hypotheses about the species.
habitat restoration actions, including the (Morrison et al. 1998)
addition of large wood, the role of models is to
evaluate factors potentially limiting fish
Species-habitat models are usually not
production in the candidate restoration stream
characterized as statistical models in the sense
for various stages in a species’ life history. For
of a regression model that attempts to find the
example, observers might examine a portion of
most parsimonious relationship between
stream and conclude that it lacks large wood
variables with no necessary mechanistic
and recommend investments in large wood
relationship (Hilborn and Mangel 1997).
structures and other measures with the
Instead, species-habitat models are often
expectation of improving habitat for salmon or
mechanistic and reflect a hypothesis concerning
other species. Evaluation of the effects of the
attributes of potential importance to fish
action in the context of the entire watershed
production based on the available literature
and species’ life history within a modeling
(“scientific model” in the sense of Hilborn and
framework, however, might indicate that the
Mangel 1997). For instance, these types of
biological value of investments in large wood
models may include relationships between fish
would be limited because of other limiting
abundance (density) and habitat type, or
factors such as downstream fish passage
relationships between growth or survival and
impediments, high temperature, or other
habitat type (Rosenfeld and Boss 2001;
limiting factors. In this case, the value of the
Railsback et al. 2003; Rosenfeld 2003). They can
model is to identify factors limiting the
be deterministic or include statistical
population, and suggest the appropriate order
confidence in attributes and relationships.
for restoration actions based on the available
Models differ in regard to the complexity of
science and conditions within the system.
hypotheses they can create and their ability to
Habitat models create a set of working compare restoration alternatives in biologically
hypotheses for limiting physical and biological meaningful terms. However, the reliability of
elements of the environment that can form a model predictions is only as good as the field
basis for evaluation of habitat hypotheses and data and knowledge that were used to build
management actions. Research and monitoring them, and users should carefully consider

Large Wood National Manual July 2015


3-17
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

potential biases when applying them to 2008) to complex depictions of species


different stream or river systems. Several types distribution across the landscape (Boyce and
of species-habitat model are discussed below McDonald 1999; Manly 2002). The Habitat
with application to evaluating restoration Evaluation Procedure (HEP) is a widely used
measures, including the addition of large wood. habitat suitability procedure developed by the
U.S. Fish and Wildlife Service to “quantify the
3.5.2.1 Habitat Association impacts of changes made through land and
Models water development projects” (Stiehl 1998).
Typically, these models start with development
The most basic habitat models are based on of a habitat suitability index (HSI). This is a
observed differences in fish density among dimensionless index of habitat suitability from
different habitat types (Nickelson et al. 1993; 0 (entirely unsuitable) to 1 (ideal habitat
Rosenfeld 2003), or simple relationships condition) for life stages or species. Suitability
between fish abundance and total habitat area relationships are developed for species
(Sharma et al. 2005; Rosenfeld and Hatfield occurrence and habitat conditions; for example,
2006). For instance, the potential increase in the abundance of large wood and the density of
coho smolt production associated with coho fry (McMahon 1983). Suitability
construction of side-channel habitat can be relationships may be developed for multiple
estimated based on average coho smolt suitability attributes believed to be important to
production from side-channel habitat reported species occurrence and performance and then
in the primary literature (i.e., 0.37 smolt m-2) integrated to create a suitability model that is
(Roni et al. 2010), if there is confidence that applied to data for a particular location or
recruitment of juvenile coho is not limited at an scenario. Suitability attributes are analogous to
earlier stage (i.e., that side-channel habitat is at the survival factors of habitat discussed above
capacity). Similarly, an average smolt and relate to habitat quality, growth and
production of 0.39 smolts m-2 from pool habitat survival experienced in different habitats. The
(Sharma et al. 2005) can be used to estimate the quantity of habitat is assessed as Habitat Units.
effects of increasing pool habitat through large
wood or other restoration. This approach,
however, does not take into account the larger GUIDELINES
context of limiting factors that may affect the
value of restoration and must be tempered by Objectives of Species-Habitat Models
recognition of the limitations and uncertainties Habitat units are the area of habitat types (e.g.,
associated with extrapolating abundance or pools or riffles) adjusted for habitat preference (e.g.,
production estimates from one stream to pools have high preference for coho fry in summer
another. While such estimates based on simple but low preference for coho spawning) and by the
areal production are not a substitute for full suitability of that habitat indexed by the HSI. Habitat
assessment of restoration needs, they may be Units are thus the quantity (acres or square meters)
appropriate for basic assessment or project of habitat adjusted for species preference and
planning. suitability.

Habitat suitability models are an additional Like all models, habitat suitability models need
class of model that have been used to assess to be used with caution because territoriality
habitat conditions for many species, especially and other factors can cause habitat selection to
wildlife, although models have been developed generate misleading indices of habitat quality
for fish as well. These models provide flexibility (Van Horne 1983; Garshelis 2000).
and can range from simple tools for collecting
expert knowledge (Railsback and Kadvany

Large Wood National Manual July 2015


3-18
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

The Instream Flow Incremental Methodology production. The key uses gradient, summer
(IFIM) uses habitat suitability relationships to water temperature, and area of key habitat in a
assess the effects of incremental changes in stream to identify which life stage habitat limits
discharge on habitat availability (Jowett et al. coho abundance. The procedure is not spatially
2008). IFIM is a set of procedures developed to explicit and analyzes an entire stream or
examine the impacts of alternative flow defined area of management interest.
regulations in streams (Bovee et al. 1998). IFIM
Nickelson et al. (1993) expanded on the key of
typically combines a hydraulic model, like the
Reeves et al. (1989) through development of an
Physical Habitat Simulation Model (PHABSIM)
analytical technique to calculate coho carrying
that predicts changes in velocity and depth with
capacity in Oregon streams. They associated the
increasing discharge, with a biological model
quantity of different stream habitat types with
that predicts how habitat quality changes with
knowledge of life stage habitat needs to
altered depths and velocities. The biological
calculate the habitat type and season limiting
models are usually habitat suitability curves for
the capacity of the stream for coho production.
velocity, depth, and substrate that rate habitat
The habitat bottleneck for a species is identified
quality for the target species between 0 and 1
as the life stage habitat in shortest supply
for a range of velocities and depths (equivalent
relative to habitat for other life stages. The
to the HSIs described above). The result of IFIM
habitat bottleneck therefore constrains overall
is an estimation of the amount of available
production of the species in the environment.
habitat in a stream for a target fish species,
Nickelson and Lawson (1998) combined their
expressed as the product of area and habitat
habitat-based model with fish population
suitability termed the Weighted Useable Area
modeling to examine the impacts of habitat
(WUA). WUA is analogous to the Habitat Units
change on population viability and extinction
in HEP. Habitat suitability curves can be
probabilities of Oregon coho.
generated for different life stages of the target
species, allowing assessment of how available Cramer and Ackerman (2009) developed the
habitat changes with discharge for different Unit Characteristic Method to assess limiting
taxa or life history stages. Similarly, flow can be factors for steelhead based on life stage
held constant, and PHABSIM or other hydraulic carrying capacity. Their method includes
models can be used for predicting the effects of consideration of food, addressed as primary
altered channel structure, such as that productivity, based on alkalinity and turbidity.
associated with restoration, on habitat Alkalinity can be used as a general indicator of
suitability and availability. Although IFIM and stream productivity (Ptolemy 1993) and has
PHABSIM are widely used, they do have been used in other models to address food
limitations and potential biases that must be availability, including the Ecosystem Diagnosis
recognized (Mathur et al. 1985; Rosenfeld and and Treatment model discussed below.
Ptolemy 2012).
3.5.2.3 Life Cycle Habitat
3.5.2.2 Habitat Capacity Models Models
Reeves et al. (1989) developed a knowledge- Life cycle models mechanistically link life stages
based key that identifies potential physical such that fish are moved from one life stage to
limitations on streams and the carrying capacity the next based on life history. Movement
of the stream for coho salmon. This between life stages reflects limitations on
dichotomous key assists managers in productivity and capacity using the features of
identifying factors limiting coho abundance and stock recruitment. The Beverton-Holt function
capacity and evaluates the need for habitat (Figure 3-1) is frequently used because of its
restoration or augmentation to optimize coho tractable mathematics, its ability to be

Large Wood National Manual July 2015


3-19
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

disaggregated into life stage functions The Ecosystem Diagnosis and Treatment (EDT)
(Moussalli and Hilborn 1986), and the ability to is a population equilibrium model that is
relate habitat attributes to the productivity and commonly used in the Columbia River Basin,
capacity parameters of stock-recruit functions Puget Sound, and the California Central Valley
(Bradford et al. 2005; Sharma et al. 2005). to identify habitat-limiting factors and develop
Productivity and capacity can be input as values restoration strategies (Blair et al. 2009). The
to life cycle models based on hypotheses, mechanism of EDT is the derivation of the
empirical measurements, or other models. population parameters of the Beverton-Holt
Productivity and capacity can also be related to relationship (Moussalli and Hilborn 1986) as a
quality and quantity of habitat (Hayes et al. function of habitat at reach and life stage scales.
1996), making it possible to model changes in These life stage estimates of productivity and
fish populations due to habitat conditions or capacity are integrated across the life history to
changes in habitat. Life cycle habitat models estimate population-level performance. Input to
discussed here evaluate potential fish the model is a reach-scale description of
performance as a function of habitat in terms of environmental conditions (e.g., temperature,
population productivity, capacity, and habitat types, and large wood) that is evaluated
abundance, which are parameters of the Viable as habitat for salmonid taxa and life stages
Salmonid Population (VSP) concept (McElhany using the VSP metrics. The model assesses the
et al. 2000) used to characterize salmonid potential diversity in fish production by
populations under the federal Endangered evaluating habitat along thousands of spatial-
Species Act. temporal pathways across the riverscape.
Potential fish production along each pathway is
The SHIRAZ model (Scheuerell et al. 2006) is an
estimated and then integrated at the population
example of a time series salmonid habitat
level; variation in performance across the
model. SHIRAZ uses a set of relationships
pathways reflects variation in the environment
between environmental attributes
and the potential life history response of the
(temperature, flow, sediment, and habitat area)
population. EDT is frequently used in a
and productivity and capacity of eggs and fry to
diagnostic mode to evaluate limiting factors at
evaluate habitat quality at the reach-scale;
attribute, reach, and life stage scales. To do this
productivity and capacity for other life stages
the model compares fish performance under a
are input as empirical values derived from the
modeled condition (e.g., the current condition)
literature (e.g., Bartz et al. 2006), expert
to performance under a reference condition
hypotheses, or observations. Input to the model
that could represent historic or future
is reach-level environmental conditions
conditions (Lichatowich et al. 1995). The result
(temperature, habitat, etc.). SHIRAZ has been
is a “blue print” for restoration actions and
used to evaluate habitat restoration alternatives
priorities.
and potential population responses to
restoration (Battin et al. 2007). Stochasticity is There are three general limitations of all habitat
included in the model by assigning statistical models discussed here. First, they are restricted
distributions to parameters or by randomly by the availability of life stage-survival
assigning parameter values across model relationships for particular taxa and life stages,
simulations (Monte Carlo approach). The model which may be lacking for important species.
evaluates potential fish performance over time This limitation applies to habitat suitability
based on a historic or simulated time series of curves used in HEP and IFIM as well as species-
annual input parameters for flow, temperature, habitat relationships used in SHIRAZ and EDT.
and channel structure. Second, habitat rating models require
information on environmental conditions at a
relatively fine scale, such as stream reaches.

Large Wood National Manual July 2015


3-20
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Such information is also often lacking (Pess et assessment programs. Testing of the model
al. 2002). When suitable fine-scale empirical components and their predictions should lead
data on habitat conditions are not available, to a refinement of the tools over time. This
extrapolations from other areas are made or creates a working hypothesis that can guide
information is derived from other models. development of restoration priorities and
Increasingly, GIS, remote sensing, and other optimize investment based on the information
techniques are being used to describe available. In this way a model becomes a vehicle
environmental conditions and can be used to for navigating restoration through uncertainty
parameterize habitat models (e.g., Benda et al. and change.
2007) While, on the one hand, the ability to
incorporate a wide range of information is a 3.5.3 Fish Assemblages and
strength of these models, on the other hand the
robustness of the conclusions must be
Large Wood
tempered by the uncertainty of input Fish assemblage responses to wood as habitat is
information. Third, habitat rating models are evident across the range of ecological regions in
notoriously difficult to validate (Morrison et al. the United States. The response of fluvial fish
1998). Habitat suitability relationships used in assemblages to wood as habitat, based on the
HEP and IFIM are dimensionless indices mesohabitat and microhabitat functions
reflecting habitat preferences that cannot be outlined above, can be judged in two main ways.
measured in the field. Life cycle habitat models First, the fish assemblage in a reach can be
such as SHIRAZ and EDT evaluate habitat using compared before and after either wood addition
the VSP parameters of productivity, abundance, or wood removal, preferably with appropriate
and biological diversity. These are intuitively control reaches without these treatments, in
attractive measures of fish performance that order to control large-scale variability (e.g.,
are routinely used in fisheries management but strong year classes for spawning fish at the
are difficult to measure except in situations basin-wide scale). Although there were nearly
where a long series of fish life stage abundance 1,200 published studies on the functions and
can be generated. Even when such data are dynamics of wood in rivers in the twentieth
available, the VSP parameters are also typically century (Gregory 2003, as cited by Nagayama
highly variable and reflect variation in survival and Nakamura 2010), relatively few published
conditions in the ocean as well as freshwater, studies have dealt with fish assemblage
and change in a VSP parameter is difficult to responses to stream rehabilitation involving
ascribe to specific habitat changes (McElhany et wood installation. Nagayama and Nakamura
al. 2000). While the overall output parameters (2010) conducted a comprehensive literature
are difficult to test, the individual components search and found 14 published studies
of habitat models, such as the relationship involving projects in fluvial habitats in the
between large wood and fry survival, can and United States. Six of these studies occurred in
should be tested through evaluation and the Marine West Coast Forest ecological region,
monitoring programs. Long-term data sets of and three each occurred in the Northwestern
biomass or production can also form the basis Forested Mountains and Eastern Temperate
for model validation by comparing model Forests ecological regions. Salmonids were the
predictions to observed fish abundance and focal species in 11 of the studies, with other
production (e.g., Beecher et al. 2010). fishes being examined in six of the studies
Consequently, the effectiveness of conclusions (including three at the level of the assemblage).
based on habitat models regarding limiting In general, these studies showed positive
factors and restoration priorities can be changes in focal fish abundance following
evaluated using data collected through rehabilitation at various scales.

Large Wood National Manual July 2015


3-21
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

In contrast, removal of large wood from and enhances vertical mixing of water (Mutz et
streams has been shown to have negative al. 2007). Researchers also find that removing
effects on fish assemblages. In a meta-analysis logs and branches in rivers can decrease
of riparian logging and wood removal from invertebrate density, richness, and biodiversity
37 studies (primarily in the Marine West Coast (Benke et al. 1985).
Forest and Northwestern Forested Mountains
Invertebrates are key prey for fish in streams
ecological regions), Mellina and Hinch (2009)
and, by consuming biofilms and periphyton,
found that thorough removal of instream large
play an important role in transferring energy to
wood following logging generally gave negative
higher trophic levels. Invertebrate consumption
responses in salmonid density and biomass.
varies by species, but detritivore assimilation of
The second means of assessing the response of leaf litter was often thought to be much higher
fish assemblages to wood as habitat is to than wood (Hutchens et al. 1997). However,
compare the assemblages in different areas recent research found that wood biofilms—
(e.g., reaches of the same river) based on the bacteria and fungi that grow on submerged
extent (quality/quantity) of wood. In Ozark woody surfaces—are an important source of
headwater streams (Great Plains ecological nutrition for invertebrates (Eggert and Wallace
region), Mitchell et al. (2012) found that large 2007). Specifically, invertebrates ingest and
wood volume was uncorrelated with overall assimilate wood biofilm at higher assimilation
fish abundance, biomass, or functional feeding efficiencies than they assimilate leaves for some
guilds, but also found that creek chub and detritivore species (Eggert and Wallace 2007).
southern redbelly dace had higher biomass with
Accumulations of organic material around wood
greater debris accumulation (as might be
also provides important habitat for terrestrial
facilitated by wood).
species. Wood accumulations that form during
high flow in spring or winter in many regions
3.5.4 Wood as Habitat for are often left dry in summer as water levels
Aquatic Invertebrates recede. During the summer wood in and around
streams can provide habitat for a range of
and Terrestrial Species terrestrial and amphibian species (Howey and
Aquatic invertebrates are very important in Dinkelacker 2009; Pittman and Dorcas 2009;
processing wood debris in forested streams and Wojan et al. 2014).
are key components of aquatic food webs. Large
The velocity refuge that aquatic fish enjoy can
wood in streams provides a physical habitat for
be used by terrestrial animals also, as small
all parts of the food web, from bacteria to
invertebrates to fish species (Cummins et al. animals can use the calmer pools for foraging
1984). The more complex the woody surface, and bathing. If the large wood spans the
the greater the resource availability and stream channel, it can be used for crossing as
associated invertebrate species richness well.
(Treadwell et al. 2007). If logs have holes,
hollows, or branches, there are several shapes 3.5.5 Assessing the
and sizes of habitats that can be formed
(Phillips 2003). Wood is especially influential in
Effectiveness of Wood
sand-bed rivers, where it provides a stable Restoration
substrate for important benthic invertebrate
Following implementation of wood restoration
species (Benke et al. 1985; Phillips 2003; Smock
projects, assessment of the effectiveness of the
et al. 1989). In sand-bed streams, wood also
projects to achieve environmental and species
creates geomorphic complexity and structure
goals is an essential component of adaptive

Large Wood National Manual July 2015


3-22
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

learning and fiscally responsible management. programs. While intensive biological


Assessment can provide new scientific monitoring may not be essential at all
understandings of how wood restoration alters restoration sites, it is necessary for a subset of
the environment, and how it provides habitat actions to ensure the effectiveness of
elements and controls fish production. These restoration techniques, and to learn from past
insights should guide funding and prioritization experience (Bernhardt et al. 2007).
for habitat restoration and provide
accountability for restoration investments.
3.6 Scale and the River
CROSS-REFERENCE Continuum Concept
The scale at which we consider streams for
Chapter 9, Assessing Ecological Performance, details
wood addition is generally the segment or reach
the processes for monitoring and assessing project
success, including a thorough discussion of Adaptive
scale, which ranges from a as little as 10 linear
Management. meters (33 linear feet) of stream to as much as a
kilometer or more. Delineation of reaches in a
stream network is usually based on geomorphic
Three levels of restoration effectiveness can be
characteristics such as tributary confluences or
distinguished. First, the most basic level of
valley form. Practically, however, reaches may
assessment is the accounting of location, cost,
also be delineated based on management
design, and expected habitat changes as well as
concerns or features such as bridges, roads, or
the biological rationale for the restoration; this
dams. Scaling up from reaches, streams are
level of assessment should be included with all
broadly categorized into size classes based on
restoration projects (Kondolf 2000; Bernhardt
where they are in the system and how many
et al. 2007). The second level evaluates the
feeder streams they have. The headwaters are
physical changes in habitat produced at various
smaller systems higher in the basins with few
intervals post-construction (e.g., 1, 2, 5, 10, or
or no feeder streams. And mainstems generally
20 years) to assess the persistence of physical
refer to the larger channels lower in the basin
restoration effects with respect to the
that carry the accumulated flow from the
expectations and purpose of the restoration.
headwaters. The River Continuum Concept
The third level addresses the biological effects
(RCC) proposes that these reaches form a
of restoration (i.e., increases in fish and/or
continuous ecological system that processes
benthic invertebrate abundance) at various
organic material and produces a distinct
intervals post-construction (Gowan and Fausch
pattern of biological communities (Vannote et
1996). It is not practical or necessary to
al. 1980). Leaves and other organic material are
perform all three of these levels of assessment
degraded and processed along the stream
for every large wood restoration project. The
continuum, and downstream reaches are
cost and effective time period to produce useful
supported to increasing degrees by “leakage” of
results increase across these three levels and
material and nutrients from upstream. While
dictate the need for a strategic approach to
acknowledged as a simplification of a complex
restoration assessment. All restoration projects
and dynamic system that includes
should include an assessment of physical
discontinuities and floodplain interactions
habitat immediately post-construction;
(Junk et al. 1989), the RCC is a reasonable
decisions to invest in detailed or long-term
framework within which to consider processes
biological monitoring should depend on the
that occur across a gradient of stream sizes in
resources available to the responsible agency
forested landscapes. The concept envisions an
and their capacity for using monitoring for
idealized riverine system as a continuum of
adaptive management of ongoing restoration

Large Wood National Manual July 2015


3-23
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

reaches grading from small headwater streams 2000; Flores et al. 2011; Wellnitz et al. 2014).
to alluvial middle sections, and deep, stable By reducing stream energy upstream of the dam
lower mainstem reaches. The role of the and (often) dissipating energy in a cascade, a
riparian zone and large wood in structuring dam’s alteration of stream energy allows
habitats and biological communities changes deposition and retention of finer material and
moving downstream between these areas. reduces bedload movement (Wallace et al.
1995a; Lemly and Hilderbrand 2000). This in
Small headwaters generally have narrow
turn alters invertebrate communities around
channels (less than 5 meters [15 feet]). Due to
the dam itself (Wallace et al. 1995a).
their size and their placement in the landscape,
these smaller tributary streams are strongly In the RCC framework, mid-level reaches are
influenced by the adjacent terrestrial ecosystem characterized as having a moderate gradient
and riparian vegetation (Vannote et al. 1980; with an increasing width that allows sunlight to
Wallace et al. 1997). Headwater reaches are reach the stream. Mid-level reaches may be
often heavily shaded by riparian vegetation and constrained within a narrow valley, are more
support only limited photosynthesis within the typically alluvial, and move back and forth
stream (Fisher and Likens 1973). As a result across the floodplain in response to high flow
headwater reaches are generally detrital-based events. The influence of the riparian zone on
and dependent on the breakdown of leaves food production and habitat formation in
from riparian forests (Vannote et al. 1980; mid-level reaches is less than in head-water
Wallace et al. 1997). Large wood has important reaches. Alluvial reaches have an abundance of
direct influences on habitat in headwater gravel and rock and unconfined valley form,
streams associated with pool formation or step resulting in complex channels, side channels,
formation in steeper streams (Bilby and Ward and gravel bars. Lateral movement of the
1989). stream undercuts riparian trees, which
increases the availability of large wood
Because stream channels are narrow and have
(Lienkaemper and Swanson 1987; Latterell and
limited power, the export of individual large
Naiman 2007). Sunlight associated with a wider
wood pieces is often low, but these large stable
channel allows greater photosynthesis in the
wood pieces often collect smaller wood pieces
form of algae on rock and on wood substrates
and organic material and form a debris dam.
(Coe et al. 2009), and algae production is
The role of large wood in forming debris dams
enhanced by nutrient additions, which can
and other retention features is particularly
come from the weathering of bedrock in the
important in these small headwater ecosystems.
system, from natural processes in the
Indeed, some of the earliest work on wood
watershed forest (e.g., the capture and
function in streams focused on the role of wood
conversion of nitrogen by plants like alder), by
in carbon retention (Bilby and Likens 1980;
processes within the stream (e.g., the release of
Bilby 1981; Bilby and Ward 1989), and
nutrients by carcasses of anadromous salmon),
subsequent studies have gone on to
or from human processes occurring in the
demonstrate that this carbon retention has
watershed (e.g., urbanization or agriculture).
important implications for stream biota and
With an increase in stream algal production, the
nutrient cycling (Wallace et al. 1997; Hall et al.
aquatic insect community in these areas is
2000; Warren et al. 2007).
dominated by species that scrape the biofilm
Debris dams created by large wood in from wood and rocks or collect pieces of leaves
headwater streams not only increase detrital or organic matter transported from upstream
food retention, they also change the substrate areas (Vannote et al. 1980).
composition of the streambed around the wood
(Wallace et al. 1995a; Lemly and Hilderbrand

Large Wood National Manual July 2015


3-24
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

A key biological function of wood in mid-level interactions decrease, as does predation risk
reaches is to structure point bars, pools, and (Sundbaum and Naslund 1998).
side-channel units (Abbe and Montgomery
Wood is particularly important in mid-level
1996). Because large wood does not typically
reaches in systems with anadromous salmon,
span the stream channel in mid-level reaches,
where wood creates holding pools and shelter
wood is more susceptible to movement by high
for juvenile salmon and migrating adults. In
flow events, creating a highly dynamic system
addition, the bedload material (i.e., gravel)
of diverse habitats. The threshold size of
retained by stable large wood and wood jams
functional wood tends to increase as stream
can be particularly important spawning habitat.
size increases (Abbe and Montgomery 1996a;
The addition of wood and other structural
Merten et al. 2010). In mid-level streams,
elements can be vitally important in retaining
smaller wood pieces are highly mobile and
and re-establishing stable spawning substrates
often function within the matrix of a debris jam
for anadromous salmon and trout (Roni et al.
that is held in place by one or two large logs
2006).
(Warren and Kraft 2008). Collins et al. (2012)
describe the role of large wood in alluviating In lower areas of river systems, channel width
reaches to create point bars, and form increases, gradient declines, depth generally
secondary channels and islands resulting in a increases, and the biological community
patchwork of habitats and features. Large wood becomes more dependent on transfer of large
jams comprising many pieces of large and small wood, nutrients, and organic material from
stream wood can also span the stream channel, upstream areas (Vannote et al. 1980). In larger
even if there is no single channel-spanning stream reaches, large “key pieces” of wood can
piece within it. become quite stable and serve to anchor bars
and other features (Abbe and Montgomery
In mid-sized streams much of the interest in
1996; Collins et al. 2012). These large pieces are
large wood has focused on its role in pool
often derived from lateral erosion and channel
formation and the associated ecological benefits
avulsion from upstream reaches. The presence
(see discussion above regarding habitat and
of a rootwad is particularly important in
reviews by Smokorowski and Pratt 2007; Roni
anchoring these key pieces and maintaining
et al. 2008; and Nagayama and Nakamura
their stability over time (Abbe and Montgomery
2010). In mid-sized streams individual pieces of
2003). In large alluvial systems, anchored large
wood and wood jams can form scour pools or
wood leads to a more complex channel form
dammed pools, or wood can enhance the size
and stabilizes floodplains (Abbe and
and complexity of existing pool habitat. Scour
Montgomery 2003; Collins et al. 2012). Large
pools are the most common type of pool formed
wood provides important habitat for many
by large wood in streams this size. Dammed
large river fish, and ecologists continue to
pools can occur in association with wood jams
advocate for wood augmentation in large
in mid-sized streams, but they are less common
streams to enhance fish habitat (Koehn and
than in smaller headwater streams. Pools
Nicol 2014).
provide holding areas for adult and juvenile fish
and feeding stations for drift feeding species Less work has been done on the dynamics and
such as coho salmon (Berg et al. 1998; Warren functions of large wood in larger river
and Kraft 2003). Enhancing complexity and ecosystems than has been done in first- through
providing areas of visual isolation can be fifth-order streams, but wood is also a key
particularly important in increasing the habitat feature in larger rivers (see Section
carrying capacity of a given pool. When fish are 3.2.1, Habitat Formation). In higher order
visually isolated, aggressive intra-species streams, the role of wood in pool formation is
dependent on the size of the wood, and the size

Large Wood National Manual July 2015


3-25
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

and energy of the stream (Abbe and are dominated by unconsolidated and fine
Montgomery 1996). In large alluvial systems, material, wood provides key stable habitat for
wood and wood jams are important in creating many invertebrate species in marsh
and enhancing scour pool habitat (Latterell et ecosystems. Wood also provides a food and
al. 2006; Pess et al. 2012). In larger streams, the habitat resource for the wood-boring isopods
presence of a rootwad is particularly important that occur in estuarine and marine ecosystems,
in reducing transport, and large trees with large which are absent from freshwater systems
rootwads tend to be stable and can be (Maser and Sedell 1994). Wood-borers in these
instrumental in developing mid-channel islands ecosystems physically degrade wood much
on large alluvial streams (Abbe and faster than any freshwater species, and wood
Montgomery 1996). In many lower gradient persistence in marine systems is therefore quite
rivers or in very wide rivers, there is not short compared to freshwater habitats.
enough energy associated with scour around
There has also been key research on the role of
the wood to create a pool. Channel-spanning
large wood in lakes that is directly relevant to
dams are very rare, and when present they are
the role of wood in low-gradient rivers where it
often removed to allow navigation by boats
may contribute minimally to controlling
(large and small).
channel structure but may still provide
While the RCC is a useful conceptual framework structure, cover, and habitat for fish and other
for stream ecology, it has been criticized as a biota. Large wood in lakes provides no
simplification (Statzner and Higler 1985). Most geomorphic effect other than to create habitat
steams do not fit the idealized RCC stream, and structure and heterogeneity, as well as
few streams conform to the totality of the providing substrate for periphyton and benthic
concept. Dams, tributary confluences, and valley invertebrates. However, studies have shown
constrictions can reset the continuum (Stanford that the presence and abundance of small
and Ward 1993). Systems in which beaver forage fish are closely related to the abundance
occur are often a series of ponds, meadows, and of littoral wood. Helmuss and Sass (2008)
streams that create discontinuities in the RCC showed that removal of 70% of large wood
(Burchsted et al. 2010). Thus, streams are more from the littoral zone of a small temperate lake
often a series of continua while still retaining resulted in a four-fold decline of yellow perch,
aspects of the generalizations in the RCC. Also, the most abundant fish in the lake, likely as a
the RCC does not address the fundamental role result of decreased refuge from predatory bass
of floods and floodplains (Sedell and Froggatt (Sass et al. 2006). Similarly, other studies have
1984; Junk et al. 1989; Sedell et al. 1989). shown that reduction in littoral large wood and
Nonetheless, the RCC is a useful framework for habitat complexity in lakes reduces abundance
thinking about stream ecosystems and how and diversity of fish. These effects can also
they may respond to changes in various habitat reasonably be expected in low-gradient or
and community features, and it continues to be warmwater rivers where the function of wood
a fundamental component of stream ecosystem relates primarily to providing cover and
theory and a useful template for contextualizing structure, rather than influencing channel
restoration of large wood. bedform, sediment storage, or transport as it
does in steeper gradient streams. Thus,
Ultimately, most streams in North America
although the function of large wood may vary
enter into ocean ecosystems (except in the great
across steep and low-gradient landscapes and
basin region in western North America) (Maser
between cold- and warmwater fish
and Sedell 1994). Large wood also functions to
communities, it generally plays an important
create habitat in intertidal and estuarine areas.
ecological role across most systems.
As in low-gradient streams where substrates

Large Wood National Manual July 2015


3-26
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

3.7 Uncertainties and Research Needs


1. The amount of wood needed to achieve management objectives—how much is enough?
2. Because so much of the scientific literature reflects conditions for salmonids in high-rainfall
areas of the Pacific Northwest, significant uncertainties exist regarding the role of wood in other
environments and for other species.

3.8 Key Points


1. Large wood is an essential element of aquatic ecosystems and creates essential habitat features
for many fish and invertebrate species.
2. The importance of wood as a habitat-forming element is highest in low-order streams
(headwater streams) and decreases as stream order increases (larger rivers).
3. Wood is a dynamic habitat feature that changes over time, reflecting patterns of recruitment and
decomposition.
4. Successful restoration of large wood in streams requires attention to systemic factors
determining the supply and movement of large wood throughout the watershed including
especially the integrity of riparian vegetation.
5. Large wood is one element of a hierarchy of factors controlling conditions in streams, hence,
successful restoration requires consideration of the wood in the context of other factors limiting
achievement of ecological and species goals.

3.9 References
Abbe, T. B. 2000. Patterns, Mechanics, and Geomorphic Effects of Wood Debris Accumulations in a
Forest River System. Ph.D. dissertation. University of Washington, Seattle, WA. 222 pp.

Abbe, T. B., and D. R. Montgomery. 1996. Large Woody Debris Jams, Channel Hydraulics and Habitat
Formation in Large Rivers. Regulated Rivers: Research and Management 12:201–221.

Abbe, T. B., and D. R. Montgomery. 2003. Patterns and Processes of Wood Debris Accumulation in
the Queets River Basin, Washington. Geomorphology 51:81–107.

Ambrose, H. E., M. A. Wilzbach, and K. W. Cummins. 2004. Periphyton Response to Increased Light
and Salmon Carcass Introduction in Northern California Streams. Journal of the North American
Benthological Society 23(4):701–712.

Armstrong J. D., and K. H. Nislow. 2006. Critical Habitat During the Transition from Maternal
Provisioning in Freshwater Fish, with Emphasis on Atlantic Salmon (Salmo salar) and Brown
Trout (Salmo trutta). Journal of Zoology 269,403–413.

Bartz, K. K., K. M. Lagueux, M. D. Scheuerell, T. Beechie, A. D. Haas, and M. H. Ruckelshaus. 2006.


Translating Restoration Scenarios into Habitat Conditions: An Initial Step in Evaluating
Recovery Strategies for Chinook Salmon (Oncorhynchus tshawytscha). Canadian Journal of
Fisheries and Aquatic Sciences 63(7):1578–1595.

Large Wood National Manual July 2015


3-27
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Beecher, H. A., B. A. Caldwell, S. B. DeMond, D. Seiler, D., and S. N. Boessow. 2010. An Empirical
Assessment of PHABSIM Using Long-Term Monitoring of Coho Salmon Smolt Production in
Bingham Creek, Washington. North American Journal of Fisheries Management 30(6):1529–
1543. doi:10.1577/M10.

Bencala, K. E. 2005. Hyporheic Exchange Flows. Encyclopedia of Hydrological Sciences, M. G.


Anderson and J. J. Mcdonnell (eds.). Wiley-Blackwell. 3,456 pp.

Benda, L. E., D. Miller, K. Andras, P. E. Bigelow, G. H. Reeves, and D. Michael. 2007. NetMap: A New
Tool in Support of Watershed Science and Resource Management. Forest Science 53(2):206–219.

Benke, A. C., R. L. Henry III, D. M. Gillespie, and R. J. Hunter. 1985. Importance of Snag Habitat for
Animal Production in Southeastern Streams. Fisheries 10:8–12.

Berg, N. A., A. Carlson, and D. Azuma. 1998. Function and Dynamics of Woody Debris in Stream
Reaches in the Central Sierra Nevada, California. Canadian Journal of Fisheries and Aquatic
Sciences 55:1807–1820.

Bernhardt, E. S., E. B. Sudduth, M. A. Palmer, J. D. Allan, J. L. Meyer, G. Alexander, J. Follastad-Shah, B.


A. Hassett, R. Jenkinson, R. Lave, J. Rumps, and L. Pagano. 2007. Restoring Rivers One Reach at a
Time: Results from a Survey of U.S. River Restoration Practitioners. Restoration Ecology
15(3):482–493.

Beschta, R. L. 1997. Restoration of Riparian and Aquatic Systems for Improved Aquatic Habitats in
the Upper Columbia River Basin. Pages 475-491 in D. J. Stouder and P. A. Bisson (eds.). Pacific
Salmon and Their Ecosystems: Status and Future Options. New York: Chapman Hall.

Beverton, R. J. H., and S. J. Holt. 1957. On the Dynamics of Exploited Fish Populations. U.K. Ministry of
Agriculture, Fisheries Investigation Service 2(19):553.

Bilby, R. E. 1981. Role of Organic Debris Dams in Regulating the Export of Dissolved and Particulate
Matter from a Forested Watershed. Ecology 62(5):1234–1243.

Bilby, R. E., and P. A. Bisson. 1998. Function and Distribution of Large Woody Debris. Pages 324–346
in R. J. Naiman and R. E. Bilby (eds.), River Ecology and Management: Lessons from the Pacific
Coast Ecoregion. New York, NY: Springer-Verlag.

Bilby, R. E., and G. E. Likens. 1980. Importance of Debris Dams in the Structure and Function of
Stream Ecosystems. Ecology 61:1107–1113.

Bilby, R. E., and J. W. Ward. 1989. Changes in Characteristics and Function of Woody Debris With
Increasing Size of Streams in Western Washington. Transactions of the American Fisheries
Society 118:368–378.

Bilby, R. E., B. R. Fransen, P. A. Bisson, and J. K. Walter. 1998. Response of Juvenile Coho Salmon
(Oncorhynchus kisutch) and Steelhead (Oncorhynchus mykiss) to the Addition of Salmon
Carcasses to Two Streams in Southwestern Washington, U.S.A. Canadian Journal of Fisheries and
Aquatic Science 55:1909–1918.

Blair, G. R., L. C. Lestelle, and L. E. Mobrand. 2009. The Ecosystem Diagnosis and treatment Model: A
Tool for Assessing Salmonid Performance Potential Based on Habitat Conditions. Pages 289–309
in E. E. Knudsen and J. J. Michael, Jr., Pacific Salmon Environment and Life History Models:

Large Wood National Manual July 2015


3-28
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Advancing Science for Sustainable Salmon in the Future. Bethesda, MD: American Fisheries
Society.

Boisclair, D. 2001. Fish Habitat Modeling: From Conceptual Framework to Functional Tools.
Canadian Journal of Fisheries and Aquatic Sciences 58:1–9.

Boulton, A. J., T. Datry, T. Kasahara, M. Mutz, and J.A. Stanford. 2010. Ecology and Management of the
Hyporheic Zone – Groundwater Interactions of Running Waters and Their Floodplains. Journal
of the North American Benthological Society 29:26–40.

Bovee, K. D., B. L. Lamb, J. M. Bartholow, C. B. Stalnaker, J. Taylor, and J. Henriksen. 1998. Stream
Habitat Analysis using the Instream Flow Incremental Methodology. Washington, DC: U.S.
Geological Survey.

Boyce, M. S., and L. L. McDonald. 1999. Relating Populations to Habitats Using Resource Selection
Functions. Trends in Ecology and Evolution 14(7):268–272.

Bradford, M. J., J. Korman, and P. S. Higgins. 2005. Using Confidence Intervals to Estimate the
Response of Salmon Populations (Oncorhynchus spp.) to Experimental Habitat Alterations.
Canadian Journal of Fisheries and Aquatic Sciences 62(12):2716–2726.

Braun, A., K. Auerswald, and J. Geist. 2012. Drivers and Spatio-Temporal Extent of Hyporheic Patch
Variation: Implications for Sampling. PLOS One 7:e42046.

Bretherton, W. D., J. S. Kominoski, D. G. Fischer, and C. J. LeRoy. 2011. Salmon Carcasses Alter Leaf
Litter Species Diversity Effects on In-stream Decomposition. Canadian Journal of Fisheries and
Aquatic Sciences 68(8):1495–1506.

Buffington, J. M., and D. R. Montgomery. 1999b. Effects of Hydraulic Roughness on Surface Textures
of Gravel-Bed Rivers. Water Resources Research 35(11):3507–3521.

Buffington, J. M. and D. Tonina. 2009. Hyporheic Exchange in Mountain Rivers II: Effects of Channel
Morphology on Mechanics, Scales, and Rates of Exchange. Geography Compass 3:1038–1062.

Burchsted, D., M. Daniels, R. Thorson, and J. Vokoun. 2010. The River Discontinuum: Beaver
Modifications to Baseline Conditions for Restoration of Forested Headwaters. BioScience
60(11):908–922.

Cederholm, C. J., M. D. Kunze, T. Murota, and A. Sibatani. 1999. Pacific salmon Carcasses: Essential
Contributions of Nutrients and Energy for Aquatic and Terrestrial Ecosystems. Fisheries
24(10):6–15.

Chapman, D. W. 1966. Food and Space as Regulators of Salmonid Populations in Streams. American
Naturalist 100:345–357.

Claussens, L., C. L. Tague, P. M. Groffman, and J. M. Melack. 2010. Longitudinal and Seasonal Variation
of N Uptake in an Urbanizing Watershed: Effect of organic Matter, Stream Size, Transient Storage
and Debris Dams. Biogeochemistry 98:45–62.

Coe, H. J., P. M. Kiffney, G. R. Press, K. K. Kloehn, and M. L. McHenry. 2009. Periphyton and
Invertebrate Response to Wood Placement in Large Pacific Coastal Rivers. River Research and
Applications 25(8):1025–1035.

Large Wood National Manual July 2015


3-29
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Collins, B. D., D. R. Montgomery, K. L. Fetherston, and T. B. Abbe. 2012. The Floodplain Large-Wood
Cycle Hypothesis: A Mechanism for the Physical and Biotic Structuring of Temperate Forested
Alluvial Valleys in the North Pacific Coastal Ecoregion. Geomorphology 139/140:460–470.

Conquest, L. L., and S. C. Ralph. 1998. Statistical Design and Analysis Considerations for Monitoring
and Assessment. Pages 455-475 in R. J. Naiman and R. E. Bilby (eds.), River Ecology and
Management: Lessons from the Pacific Coastal Ecoregion. New York: Springer.

Cramer, S. P., and N. K. Ackerman. 2009. Prediction of Stream Carrying Capacity for Steelhead; The
Unit Characterization Method. Pages 255–288 in Pacific Salmon Environment and Life History
Models. Bethesda, MD: American Fisheries Society.

Cummins, K. W. 1974. Structure and Function of Stream Ecosystems. BioScience 24(11):631–641.

Cummins, K. W., G. W. Minshall, J. R. Sedell, C. E. Cushing, and R. C. Petersen. 1984. Stream Ecosystem
Theory. Internationale Vereinigung fur theoretische und angewandte Limnologie, Verhandlungen
22:1818–1827.

Daniels, M. D., and Rhoads, B. L. 2004. Effect of Large Woody Debris Configuration on Three-
Dimensional Flow Structure in Two Low-Energy Meander Bends at Varying Stages. Water
Resources Research (40):W11302.

Eggert, S. L., and J. B. Wallace. 2007. Wood Biofilm as a Food Resource for Stream Detritivores.
Limnology and Oceanography 52(3):1239–1245.

Ehrman, T. P., and G. A. Lamberti. 1992. Hydraulic and Particulate Matter Retention in a 3rd-Order
Indiana Stream. Journal of the North American Benthological Society. 11:341–349.

Ensign, S. H., and M. W. Doyle. 2005. In-channel Transient Storage and Associated Nutrient
Retention: Evidence from Experimental Manipulations. Limnology and Oceanography
50(6):1740–1751.

Environmental Agency. 2009. The Hyporheic Handbook. a Handbook of the Groundwater-Surface


Water Interface and Hyporheic Zone for Environmental Managers. Science Report SC050070. 264
pp. Available: http://www.hyporheic.net/SCHO1009BRDX-e-e.pdf. Accessed: June 13, 2014.

Fausch, K. D., C. E. Torgersen, C. V. Baxter, and H. W. Li. 2002. Landscapes to Riverscapes: Bridging
the Gap Between Research and Conservation of Stream Fishes. BioScience 52(6):483–498.

Findlay, S., J. Tank, S. Dye, H. M. Valett, P. J. Mulholland, W. H. McDowell, S. L. Johnson, S. K. Hamilton,


J. Edmonds, W. K. Dodds, and W. B. Bowden. 2002. A Cross System Comparison of Bacterial and
Fungal Biomass in Detritus Pools of Headwater Streams. Microbial Ecology 43(1):55–66.

Fisher, S. G., and G. E. Likens. 1972. Stream Ecosystem—Organic Energy Budget. Bioscience 22(1):33-
35.

Flebbe, P. A. 1999. Trout Use of Wood Debris and Habitat in Wine Spring Creek, North Carolina.
Forest Ecology and Management 114:367–376.

Flores, L., A. Larranaga, J. Diez, and A. Elosegi. 2011. Experimental Wood Addition in Streams: Effects
on Organic Matter Storage and Breakdown. Freshwater Biology 56(10):2156–2167.

Large Wood National Manual July 2015


3-30
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Frissell, W. J., W. J. Liss, C. E. Warren, and M. D. Hurley. 1986. A Hierarchical Framework for Stream
Habitat Classification: Viewing Streams in a Watershed Context. Environmental Management
10(2):199–214.

Gandy, C. J., and A. P. Jarvis. 2006. Attenuation of Nine Pollutants in the Hyporheic Zone. Environment
Agency, Bristol, England, June. 33 pp.

Garshelis, D. L. 2000. Delusions in Habitat Evaluation: Measuring Use, Selection, and Importance.
Pages 11–165 in L. Boitani and T. K. Fuller (eds.), Research Techniques in Animal Ecology. New
York: Columbia University Press.

Gowan, C., and K. D. Fausch. 1996. Long-Term Demographic Responses of Trout Populations to
Habitat Manipulation in Six Colorado Streams. Ecological Applications 6(3):931–946.

Grizzel, J., M. McGowan, D. Smith, and T. Beechie. 2000. Streamside Buffers and Large Woody Debris
Recruitment: Evaluating the Effectiveness of Watershed Analysis Prescriptions in the North
Cascades Region. TFW-MAGI-00-003. Washington State Timber, Fish & Wildlife.

Hall, L. S., P. R. Krausman, and M. L. Morrison. 1997. The Habitat Concept and a Plea for Standard
Terminology. Wildlife Society Bulletin 25(1):173–182.

Hall, R. O., J. B. Wallace, and S. L. Eggert. 2000. Organic Matter Flow in Stream Food Webs with
Reduced Detrital Resource Base. Ecology 81(12):3445–3463.

Harmon, M. E., J. F. Franklin, F. J. Swanson, P. Sollins, S. V. Gregory, J. D. Lattin, N. H. Anderson, S. P.


Cline, N. G. Aumen, J. R. Sedell, G. W. Lienkaemper, K. Cromack, Jr., and K. W. Cummins. 1986.
Ecology of Coarse Woody Debris in Temperate Ecosystems. Advances in Ecological Research
15:133–302.

Hayes, D. B., C. P. Ferreri, and W. W. Taylor. 1996. Linking Fish Habitat to Their Population
Dynamics. Canadian Journal of Fisheries and Aquatic Sciences 53(S1):383–390.

Hedman, C. W., D. H. Van Lear, and W. T. Swank. 1996. In-Stream Large Woody Debris Loading and
Riparian Forest Seral Stage Associations in the Southern Appalachian Mountains. Canadian
Journal of Forest Research 26:1218–1227.

Helmus, M. R. and G. G. Sass. 2008. The Rapid Effects of a Whole-Lake Reduction of Coarse Woody
Debris on Fish and Benthic Macroinvertebrates. Freshwater Biology 53:1423–1433.

Hester, E. T., and M. W. Doyle. 2008. In-Stream Geomorphic Structures as Drivers of Hyporheic
Change. Water Resources Research 44:W03427.

Hester, E. T., and M. N. Gooseff. 2010. Moving Beyond the Banks: Hyporheic Restoration is
Fundamental to Restoring Ecological Services and Functions of Streams. Environmental Science
and Technology 44:1521–1525.

Hilborn, R., and M. Mangel. 1997. The Ecological Detective. Princeton, NJ: Princeton University Press.

Hilborn, R., and C. J. Walters. 1992. Quantitative Fish Stock Assessment. London: Chapman and Hall.

Hoellein, T. J., J. L. Tank, E. J. Rosi-Marshall, S. A. Entrekin, and G. A. Lamberti. 2007. Controls on


Spatial and Temporal Variation of Nutrient Uptake in Three Michigan Headwater Streams.
Limnol. Oceanogr.52:1964–1977.

Large Wood National Manual July 2015


3-31
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Howey, C. A. F., and S. A. Dinkelacker. 2009. Habitat Selection of the Alligator Snapping Turtle
(Macrochelys temminckii) in Arkansas. Journal of Herpetology 43(4):589–596.

Hutchens, J. J., E. F. Benfield, and J. R. Webster. 1997. Diet and Growth of a Leaf-Shredding Caddisfly
in Southern Appalachian Streams of Contrasting Disturbance History. Hydrobiologia 346:193–
201.

Johnson, L. B., D. H. Breneman, and C. Richards. 2003. Macroinvertebrate Community Structure and
Function Associated with Large Wood in Low Gradient Streams. River Research and Applications
19:199–218.

Jonsson, B., and N. Jonsson. 2003. Migratory Atlantic Salmon as Vectors for the Transfer of Energy
and Nutrients Between Freshwater and Marine Environments. Freshwater Biology 48:21–27.

Johnston, N. T., S. A. Bird, D. L. Hogan, and E. A. MacIsaac. 2011. Mechanisms and Source Distances
for the Input of Large Woody Debris to Forested Streams in British Columbia, Canada Canadian
Journal of Forest Research. 41(11):2231–2246.

Jowett, I. G., J. W. Hayes, and M. J. Duncan. 2008. A Guide to Instream Habitat Survey Methods and
Analysis. NIWA Science and Technology Series No. 54. Available:
http://www.niwa.co.nz/sites/niwa.co.nz/files/a_guide_to_instream_habitat_survey_methods_an
d_analysis.pdf.

Junk, W. J., P. B. Bayley, and R. E. Sparks. 1989. The Flood Pulse Concept in River Floodplain Systems.
Pages 110-127 in D. P. Dodge (ed.), Proceedings of the International Large River Symposium.

Keeton, W. S., C. E. Kraft, and D. R. Warren. 2007. Mature and Old-Growth Riparian Forests:
Structure, Dynamics and Effects on Adirondack Stream Habitats. Ecological Applications 17:852–
868.

Kennedy, B. P., K. H. Nislow, and C. L. Folt. 2008. Habitat-Mediated Foraging Limitations Drive
Survival Bottlenecks for Juvenile Salmon. Ecology 89(9):2529–2541.

Koehn, J. D., and S. J. Nicol. 2014. Comparative Habitat Use by Large Riverine Fishes. Marine and
Freshwater Research 65(2):164–174.

Kondolf, G. M. 2000. Some Suggested Guidelines for Geomorphic Aspects of Anadromous Salmonid
Habitat Restoration Proposals. Restoration Ecology 8:48–55.

Kratzer, J. F., and D. R. Warren. 2013. Factors Limiting Brook Trout Biomass in Northeastern
Vermont Streams. North American Journal of Fisheries Management 33(1):130–139.

Krause, S., M. J. Klaar, D. M. Hannah, J. Mant, J. Bridgeman, M. Trimmer, and S. Manning-Jones. 2014.
The Potential of Large Woody Debris to Alter Biogeochemical Processes and Ecosystem Services
in Lowland Rivers. Wiley Interdisciplinary Reviews (WIREs): Water 1:263–275.

Lancaster, S. T., S. K. Hayes, and G. E. Grant. 2001. Modeling Sediment and Wood Storage and
Dynamics in Small Mountainous Watersheds. Geomorphic Processes and Riverine Habitat, Water
Science and Application Volume 4:85–102. American Geophysical Union.

Latterell, J. J., and R. J. Naiman. 2007. Sources and Dynamics of Large Logs In a Temperate Floodplain
River. Ecological Applications 17:1127–1141.

Large Wood National Manual July 2015


3-32
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Latterell, J. J., J. S. Bechtold, T. C. O'Keefe, R. Van Pelt, and R. J. Naiman. 2006. Dynamic Patch Mosaics
and Channel Movement in an Unconfined River Valley of the Olympic Mountains Freshwater
Biology 51(3):523–544.

Lautz, L. K., and R. M. Fanelli. 2008. Biogeochemical Hotspots in the Streambed around Restoration
Structures. Biogeochemistry 91:85–104.

Lautz, L. K., D. I. Siegel, and R. L. Bauer. 2006. Impact of Debris Dams on Hyporheic Interaction along
a Semi-Arid Stream. Hydrological Processes 20:183–196.

Lautz, L. K., R. M. Fanelli, N. Kranes, and D. I. Siegel. 2007. Abstract. Sediment distribution around
Debris Dams: Impacts on Streambed Hydrology, Biogeochemistry and Temperature Dynamics in
Small Streams. Geological Society of America Annual Meeting (28–31 October 2007), Paper No.
180-4.

Lemly, A. D., and R. H. Hilderbrand. 2000. Influence of Large Woody Debris on Stream Insect
Communities and Benthic Detritus. Hydrobiologia 421:179–185.

Lichatowich, J. A., L. E. Mobrand, L. Lestelle, and T. Vogel. 1995. An Approach to the Diagnosis and
Treatment of Depleted Pacific Salmon Populations in Freshwater Ecosystems. Fisheries
20(1):10–18.

Lienkaemper, G. W., and F. J. Swanson. 1987. Dynamics of Large Woody Debris in Streams in Old-
Growth Douglas-Fir Forests. Canadian Journal of Forest Research 17:150–156.

Manga, M., and J. W. Kirchner. 2000. Stress Partitioning in Streams by Large Woody Debris. Water
Resources Research 36:2373–2379.

Manly, B. F. 2002. Resource Selection by Animals. Springer, New York.

Manners, R. W., M. W. Doyle, and M. J. Small. 2007. Structure and Hydraulics of Natural Woody
Debris Jams. Water Resources Research 43, doi:10.1029/2006WR004910.

Maser, C., and J. R. Sedell. 1988. From the Forest to the Sea: the Ecology of Wood in Streams, Rivers,
Estuaries and Oceans. Delray Beach, FL: St. Lucie Press.

Maser, C. and J. R. Sedell, J.R. 1994. From the Forest to the Sea: The Ecology of Wood in Streams,
Rivers, Estuaries, and Oceans. St. Lucie Press. 200 pp.

Mathur, D., W. H. Bason, E. J. Purdy, and C. A. Silver. 1985. A Critique of Instream Flow Incremental
Methodology. Canadian Journal of Fisheries and Aquatic Science 42(4):825–831.

McDade, M. H., F. J. Swanson, W. A. McKee, J. F. Franklin, and J. Van Sickle. 1990. Source Distances for
Coarse Woody Debris Entering Small Streams in Western Oregon and Washington. Canadian
Journal of Forest Research 20:326–330.

McElhany, P., M. H. Ruckelshaus, M. J. Ford, T. C. Wainwright, and E. P. Bjorkstedt. 2000. Viable


Salmonid Populations and the Recovery of Evolutionary Significant Units. U.S. Department of
Commerce, Seattle, WA. NOAA Tech. Memo NMFS-NWFSC-42.

McMahon, T. E. 1983. Habitat Suitability Index Models: Coho Salmon. U.S. Fish and Wildlife Service.

McNeely, C., and M. E. Power. 2007. Spatial Variation in Caddisfly Grazing Regimes Within a
Northern California Watershed. Ecology 88(10):2609–2619.

Large Wood National Manual July 2015


3-33
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Merten, E., J. Finlay, L. Johnson, R. Newman, R., H. Stefan, and B. Vondracek. 2010. Factors
Influencing Wood Mobilization in Stream. Water Resources Research 46:W10514.

Merz, J., and P. B. Moyle. 2006. Salmon, Wildlife and Wine: Marine-Derived Nutrients in Human--
Dominated Ecosystems of Central California. Ecological Applications 13(3):999–1009.

Miller, D., C. Luce, and L. Benda. 2003. Time, Space, and Episodicity of Physical Disturbance in
Streams. Forest Ecology and Management 178(1):121–140.

Montgomery, D. R., and J. M. Buffington. 1998. Channel Processes, Classification and Response. Pages
13–42 in R. J. Naiman and R. E. Bilby (eds.), River Ecology and Management: Lessons from the
Pacific Coastal Ecoregion. New York: Springer.

Montgomery, D. R., B. D. Collins, J. M. Buffington, and T. B. Abbe. 2003. Geomorphic Effects of Wood
in Rivers. Pages 21–47 in S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology and
Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society.

Moore, M. K. 1977. Factors Contributing to Blowdown in Streamside Leave Strips on Vancouver


Island. Land Management Report No. 3. Victoria, BC: Province of British Columbia Ministry of
Forests, Information Division.

Morrison, M. L., B. G. Marcot, and R. W. Mannon. 1998. Wildlife-Habitat Relationships. Concepts and
Applications. Madison, WI: University of Wisconsin Press.

Moussalli, E., and R. Hilborn. 1986. Optimal Stock Size and Harvest Rate in Multistage Life History
Models. Canadian Journal of Fisheries and Aquatic Sciences 43(1):135–141.

Mulholland, P. J., and J. R. Webster. 2010. Nutrient Dynamics in Streams and the Role of J-NABS.
Journal of the North American Benthological Society 29:100–117.

Mulholland, P. J., and 33 others. 2009. Nitrate Removal in Stream Ecosystems Measured by 15N
Addition Experiments: Denitrification. Limnology and Oceanography 54:666–680.

Murphy, M. L., and K. V. Koski. 1989. Input and Depletion of Woody Debris in Alaska Streams and
Implications for Streamside Management. North American Journal of Fisheries Management
9(4):427-436.

Murphy, M. L., J. Heifetz, S. W. Johnson, K. V. Koski, and J. F. Thendinga. 1986. Effects of Clear-Cut
Logging with and without Buffer Strips on Juvenile Salmonids in Alaskan Streams. Canadian
Journal of Fisheries and Aquatic Sciences 43:1521–1533.

Mutz, M., E. Kalbus, and S. Meinecke. 2007. Effect of Instream Wood on Vertical Water Flux in Low-
Energy Sand Bed Flume Experiments. Water Resources Research 43:W10424.

Nagayama, S., and F. Nakamura. 2010. Fish Habitat Rehabilitation Using Wood in the World.
Landscape and Ecological Engineering 6(2):289–305.

Naiman, R. J., K. L. Fetherston, S. McKay, and J. Chen. 1998. Riparian Forests. Pages 289-323 in R. J.
Naiman and R. E. Bilby (eds.), River Ecology and Management: Lessons from the Pacific Coastal
Ecoregion. Springer-Verlag: New York.

Naiman, R. J., R. E. Bilby, D. E. Schindler, and J. M. Helfield. 2002c. Pacific Salmon, Nutrients, and the
Dynamics of Freshwater and Riparian Ecosystems. Ecosystems 5:399–417.

Large Wood National Manual July 2015


3-34
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Nickelson, T. E., and P. W. Lawson. 1998. Population Viability of Coho Salmon, Oncorhynchus kisutch,
in Oregon Coastal Basins: Application of a Habitat-Based Life Cycle Model. Canadian Journal of
Fisheries and Aquatic Sciences 55:2383–2392.

Nickelson, T. E., M. F. Solazzi, S. L. Johnson, and J. D. Rodgers. 1993. An Approach to Determining


Stream Carrying Capacity and Limiting Habitat For Coho Salmon (Oncorhynchus kisutch). Paper
read at Proceedings of the Coho Workshop, May 1992, Nanaimo, B.C.

O’Connor J. E., M. A. Jones, and T. L. Haluska. 2003. Flood Plain and Channel Dynamics of the Quinault
and Queets Rivers, Washington, U.S.A. Geomorphology 52:31–59.

O’Connor, M., and G. Watson, 1998. Geomorphology of Channel Migration Zones and Implications for
Riparian Forest Management. Available: http://www.oei.com/Reports/Geomorph_of_CMZ.htm.

O'Connor, R. R., and F. J. Rahel. 2009. A Patch Perspective on Summer Habitat Use by Brown Trout
Salmo trutta in a High Plains Stream in Wyoming, USA. Ecology of Freshwater Fish 18(3):473–
480.

Oliver, C. D. 1980. Forest Development in North America Following Major Disturbances. Forest
Ecology and Management 3:153–168.

Pess, G. R., D. R. Montgomery, E. A. Steel, R. E. Bilby, B. E. Feist, and H. M. Greenberg. 2002. Landscape
Characteristics, Land Use, and Coho Salmon (Oncorhynchus kisutch) Abundance, Snohomish
River, Washington, USA. Canadian Journal of Fisheries and Aquatic Sciences 59:613–623.

Pess, G. R., M. C. Liermann, M. L. Mchenry, R. J. Peters, and T. R. Bennett. 2012. Juvenile Salmon
Response to the Placement of Engineered Log Jams (Eljs) in the Elwha River, Washington State,
USA. River Research and Applications 28(7):872–881.

Phillips, E. C. 2003. Habitat Preference of Aquatic Macroinvertebrates in an East Texas Sandy


Stream. Journal of Freshwater Ecology 18(1):1–11.

Pianka, E. R. 1994. Evolutionary Ecology. Harper-Collins.

Pickett, S. T. A., and K. H. Rogers. 1997. Patch Dynamics: The Transformation of Landscape Structure
and Function. Pages 101–127 in J. A. Bissonette (ed.), Wildlife and Landscape Ecology. New York:
Springer-Verlag.

Pittman, S. E., and M. E. Dorcas. 2009. Movements, Habitat Use, and Thermal Ecology of an Isolated
Population of Bog Turtles (Glyptemys muhlenbergii). Copeia(4):781–790.

Ptolemy, R. A. 1993. Maximum Salmonid Densities in Fluvial Habitats in British Columbia. Paper
read at Proceedings of the Coho Workshop, May 26–28, 1992, at Nanaimo, B.C.

Quinn, T. P., S. M. Carlson, S. M. Gende, and H. B. Rich. 2009. Transportation of Pacific Salmon
Carcasses from Streams to Riparian Forests by Bears. Canadian Journal of Fisheries and Aquatic
Science 87:195–203.

Railsback, S. F., and J. Kadvany. 2008. Demonstration Flow Assessment: Judgment and Visual
Observation in Instream Flow Studies. Fisheries 33:217–227.

Railsback, S. F., H. Stauffer, and B. Harvey. 2003. What can Habitat Preference Models Tell Us? Tests
using a Virtual Trout Population. Ecological Applications 13(6):1580–1594.

Large Wood National Manual July 2015


3-35
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Reeves, G. H., F. H. Everest, and T. E. Nickelson. 1989. Identification of Physical Habitats Limiting the
Production of Coho Salmon in Western Oregon and Washington. Portland, OR: USDA Forest
Service.

Reisenbichler, R. R. 1989. Utility of Spawner-Recruit Relations for Evaluating the Effect of Degraded
Environment on the Abundance of Chinook Salmon, Oncorhynchus tshawytscha. Pages 21–32 in
C. D. Levings, L. B. Holtby, and M. A. Henderson (eds.), Proceedings of the National Workshop on
Effects of Habitat Alteration on Salmonid Stocks: Canadian Special Publication on Fisheries and
Aquatic Sciences 105.

Richmond, A. D., and K. D. Fausch. 1995. Characteristics and Function of Large Woody Debris in
Subalpine Rocky Mountain Streams in Northern Colorado. Canadian Journal of Fisheries and
Aquatic Sciences 52:1789–1802.

Robison, E. G. and R. L. Beschta. 1990. Identifying Trees in Riparian Areas that can Provide Coarse
Woody Debris to Streams. Forest Science 36:790–801.

Roni, P., T. Bennett, S. Morley, G. R. Pess, K. Hanson, D. Van Slyke, and P. Olmstead. 2006.
Rehabilitation of Bedrock Stream Channels: The Effects of Boulder Weir Placement on Aquatic
Habitat and Biota. River Research and Applications 22(9):967–980.

Roni, P., K. Hanson, and T. Beechie. 2008. Global Review of the Physical and Biological Effectiveness
of Stream Habitat Rehabilitation Techniques. North American Journal of Fisheries Management
28(3):856–890.

Roni, P., G. Pess, T. Beechie, and S. Morley. 2010. Estimating Changes in Coho Salmon and Steelhead
Abundance from Watershed Restoration: How Much Restoration is Needed to Measurably
Increase Smolt Production? North American Journal of Fisheries Management 30(6):1469–1484.

Roni, P., T. J. Beechie, G. R. Pess, and K. M. Hanson. 2014a. Wood Placement in River Restoration:
Fact, Fiction and Future Direction. Canadian Journal of Fisheries and Aquatic Sciences.

Roni, P., G. R. Pess, and T. J. Beechie. 2014b. Fish-Habitat Relationships and Effectiveness of Habitat
Restoration. National Marine Fisheries Service. Seattle, WA. NOAA Technical Memorandum
NMFS-NWFSC-127.

Rose, K. A. 2000. Why are Quantitative Relationships Between Environmental Quality and Fish
Populations so Elusive? Ecological Applications 10(2):367–385.

Rosenfeld, J. 2003. Assessing the Habitat Requirements of Stream Fishes: an Overview and
Evaluation of Different Approaches. Transactions of the American Fisheries Society 132:953–968.

Rosenfeld, J. S. 2014. Modelling the Effects of Habitat on Self-thinning, Energy Equivalence, and
Optimal Habitat Structure for Juvenile Trout. Canadian Journal of Fisheries and Aquatic Sciences.
71(9):1395–1406.

Rosenfeld, J. S., and S. Boss. 2001. Fitness Consequences of Habitat Use for Juvenile Cutthroat Trout:
Energetic Costs and Benefits in Pools and Riffles. Canadian Journal of Fisheries and Aquatic
Sciences 58(3):585–593.

Rosenfeld, J., and T. Hatfield. 2006. Information Needs for Assessing Critical Habitat of Freshwater
Fish. Canadian Journal of Fisheries and Aquatic Sciences 63:683–698.

Large Wood National Manual July 2015


3-36
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Rosenfeld, J. S., and R. Ptolemy. 2012. Modelling Available Habitat Versus Available Energy Flux: Do
PHABSIM Applications that Neglect Prey Abundance Underestimate Optimal Flows for Juvenile
Salmonids? Canadian Journal of Fisheries and Aquatic Sciences 69(12):1920–1934.

Rot, B. 1993. Windthrow in Stream Buffers on Coastal Washington Streams. ITT-Rayonier Inc. 49 pp.

Rot, B. W., R. J. Naiman, and R. E. Bilby. 2000. Stream Channel Configuration, Landform, and Riparian
Forest Structure in the Cascade Mountains, Washington. Canadian Journal of Fisheries and
Aquatic Sciences 57:699–707.

Rutherford, I., B. Anderson, and A. Ladson. 2007. Managing the Effects of Riparian Vegetation on
Flooding. In S. Lovett and P. Price (eds.), Principles for Riparian Lands Management. Land &
Water Australia, Canberra.

Sabater, S., V. Acuña, A. Giorgi, E. Guerra, I. Muñoz, and A. M. Romaní, 2005. Effects of Nutrient Inputs
in a Forested Mediterranean Stream Under Moderate Light Availability. Archiv für Hydrobiologie
163:479–496.

Sass, G. G., J. F. Kitchell, S. R. Carpenter, T. R. Hrabik, A. E. Marburg, and M. G. Turner. 2006. Fish
Community and Food Web Responses to a Whole-Lake Removal of Coarse Woody Habitat.
Fisheries 31:321–330.

Sawyer, A. H., and M. B. Cardenas. 2012. Effect of Experimental Wood Addition on Hyporheic
Exchange and Thermal Dynamics in a Losing Meadow Stream. Water Resources Research
48:W10537.

Sawyer, A. H., M. B. Cardenas, and J. Buttles. 2011. Hyporheic Exchange due to Channel-Spanning
Logs. Water Resources Research 47(8):W08502.

Sawyer, A. H., M. B. Cardenas, and J. Buttles. 2012. Hyporheic Temperature Dynamics and Heat
Exchange Near Channel-Spanning Logs. Water Resources Research 48:W01529.

Schenk, E. R., J. W. McCargo, B. Moulin, C. R. Hupp, and J. M. Richter. 2014. The Influence of Logjams
on Largemouth Bass (Micropterus salmoides) Concentrations on the lower Roanoke River, a
Large Sand-Bed River. River Research and Applications 2014(DOI: 10.1002/rra.2779).

Scheuerell, M. D., R. Hilborn, M. H. Ruckelshaus, K. K. Bartz, K. M. Lagueux, A. D. Haas, and K. Rawson.


2006. The Shiraz Model: A Tool for Incorporating Anthropogenic Effects and Fish–Habitat
Relationships in Conservation Planning. Canadian Journal of Fisheries and Aquatic Sciences
63(7):1596–1607.

Schlosser, I. J., and P. L. Angermeier. 1995. Spatial Variation in Demographic Processes of Lotic
Fishes: Conceptual Models, Empirical Evidence, and Implications for Conservation. Pages 392–
401 in J. L. Nielsen and D. A. Powers (eds.), Evolution and the Aquatic Ecosystem: Defining Unique
Units in Population Conservation. Bethesda, MD: American Fisheries Society.

Schmetterling, D. A., and R. W. Pierce. 1999. Success of Instream Habitat Structures After a 50-Year
Flood in Gold Creek, Montana. Restoration Ecology 7(4):369–375.

Sedell, J. R., and J. L. Frogatt. 1984. Importance of Streamside Forests to Large Rivers: The Isolation
of the Willamette River, Oregon, U.S.A., from its Floodplain by Snagging and Streamside Forest
Removal. Verhandlungen-Internationale Vereinigung für Theorelifche und Angewandte
Limnologie 22:1828–1834.

Large Wood National Manual July 2015


3-37
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Sedell, J. R., J. E. Richey, and F. J. Swanson. 1989. The River Continuum Concept: A Basis for the
Expected Ecosystem Behavior of Very Large Rivers? Canadian Special Publication of Fisheries
and Aquatic Sciences 106:49–55.

Sharma, R., A. B. Cooper, and R. Hilborn. 2005. A Quantitative Framework for the Analysis of Habitat
and Hatchery Practices on Pacific Salmon. Ecological Modeling 18:231–250.

Smock, L. A., G. M. Metzler and J. E. Gladden. 1989. Role of Debris Dams in the Structure and
Functioning of Low Gradient Headwater Streams. Ecology 70:764–775.

Smokorowski, K. E., and T. C. Pratt. 2007. Effect of a Change in Physical Structure and Cover on Fish
and Fish Habitat in Freshwater Ecosystems - A Review and Meta-Analysis. Environmental
Reviews 15:15–41.

Sobota, D. J., S. V. Gregory, S. V., and S. L. Johnson. 2007. Influence of Wood Decomposition on
Nitrogen Dynamics in Stream Ecosystems: Interactive Effects of Substrate Quality and Nitrogen
Loading. North American Benthological Society 55th Annual Meeting. Available:
https://nabs.confex.com/nabs/2007/techprogram/P1365.HTM.

Southwood, T. R. E. 1977. Habitat, the Template for Ecological Strategies? Journal of Animal Ecology
46:337–365.

Spanhoff, B., and E. Cleven. 2010. Wood in Different Stream Types: Epixylic Biofilm and Wood-
Inhabiting Invertebrates in a Lowland Versus an Upland Stream. Annales De Limnologie-
International Journal of Limnology 46(3):169–179.

Spanhoff, B., and E. I. Meyer. 2004. Breakdown Rates of Wood in Streams. Journal of the North
American Benthological Society 23(2):189–197.

Stanford, J. A. and J. V. Ward. 1988. The Hyporheic Habitat of River Ecosystems. Nature 335:64–66.

Stanford, J. A., and J. V. Ward. 1993. An Ecosystem Perspective of Alluvial Rivers: Connectivity and
the Hyporheic Corridor. Journal of the North American Benthological Society 12(1):48–60.

Stanford, J. A., and J. V. Ward. 1995. The Serial Discontinuity Concept: Extending the Model to
Floodplain Rivers. Regulated Rivers: Research and Management:159–168.

Statzner, B., and B. Higler. 1985. Questions and Comments on the River Continuum Concept.
Canadian Journal of Fisheries and Aquatic Science 42:1038–1044.

Steinhart, G. S., G. E. Likens, and P. M. Groffman. 2000. Denitrification in Stream Sediments in Five
Northeastern (USA) Streams. Verhandlungen Internationale Vereinigung für Theorertische und
Angewandte Limnologie 27:1331–1336.

Stewart, G. B., H. R. Bayliss, D. A. Showler, W. J. Sutherland, and A. S. Pullin. 2009. Effectiveness of


Engineered In-Stream Structure Mitigation Measures to increase Salmonid Abundance: A
Systematic Review. Ecological Applications 19(4):931–941.

Stewart, P. M., S. Bhattarai, M. W. Mullen, C. K. Metcalf, and E. G. Reategui-Zirena. 2012.


Characterization of Large Wood and its Relationship to Pool Formation and Macroinvertebrate
Metrics in Southeastern Coastal Plain Streams, USA. Journal of Freshwater Ecology 27(3):351–
365.

Stiehl, R. B. 1998. Habitat Evaluation Procedures Workbook. Fort Collins, CO: U.S. Geological Survey.

Large Wood National Manual July 2015


3-38
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Stockner, J. G. (ed.). 2003. Nutrients in Salmonid Ecosystems: Sustaining Production and


Biodiversity. Bethesda, MD: American Fisheries Society.

Stofleth, J. M., F. D. Shields, Jr., and G. A. Fox. 2007. Hyporheic and Total Transient Storage in Small,
Sand-Bed Streams. Hydrological Processes 22:1885–1894.

Sundbaum, K. and I. Naslund. 1998. Effects of Woody Debris on the Growth and Behavior of Brown
Trout in Experimental Stream Channels. Canadian Journal of Zoology 76:56–61.

Thompson, D. M. 2006. Did the Pre-1980 Use of In-Stream Structures Improve Streams? A
Reanalysis of Historical Data. Ecological Applications 16(2):784–796.

Tonina, D., and J. M. Buffington. 2009. Hyporheic Exchange In Mountain Rivers I: Mechanics and
Environmental Effects. Geography Compass 3:1063–1086.

Townsend, C. R. 1989. The Patch Dynamics Concept of Stream Community Ecology. Journal of the
North American Benthological Society 8(1):36–50.

Treadwell, S., J. Koehn, S. Bunn, and A. Brooks. 2007. Wood and Other Aquatic Habitat. Chapter 7 in
S. Lovett and P. Price (eds.), Principles for Riparian Lands Management. Land and Water
Australia, Canberra.

Trotter, E. H. 1990. Woody Debris, Forest-Stream Succession, and Catchment Geomorphology.


Journal of the North American Benthological Society 9(2):141–156.

Valett, H. M., C. L. Crenshaw, and P. F. Wagner. 2002. Stream Nutrient Uptake, Forest Succession, and
Biogeochemical Theory. Ecology 83:2888–2901.

Van Horne, B. 1983. Density as a Misleading Indicator of Habitat Quality. Journal of Wildlife
Management 47:893–901.

Van Sickle, J., and S. V. Gregory. 1990. Modeling Inputs of Large Woody Debris to Streams from
Falling Trees. Canadian Journal of Forest Research 20(10):1593–1601.

Vannote, R. L., G. W. Minshall, K. W. Cummins, J. R. Sedell, and C. E. Cushing. 1980. The River
Continuum Concept. Canadian Journal of Fisheries and Aquatic Science 37(1):130–137.

Wallace, J. B., J. R. Webster, and J. L. Meyer. 1995a. Influence of Log Additions on Physical and Biotic
Characteristics of a Mountain Stream. Canadian Journal of Fisheries and Aquatic Sciences
52:2120–2137.

Wallace, J. B., M. R. Whiles, S. Eggert, T. F. Cuffney, G. H. Lugthart, and K. Chung. 1995b. Long-Term
Dynamics of Coarse Particulate Organic-Matter in 3 Appalachian Mountain Streams. Journal of
the North American Benthological Society 14(2):217–232.

Wallace, J. B., S. L. Eggert, J. L. Meyer, and J. R. Webster. 1997. Multiple Trophic Levels of a Forest
Stream Linked to Terrestrial Litter Inputs. Science 277:102–104.

Walter, R. C., and D. J. Merritts. 2008. Natural Streams and the Legacy of Water-Powered Mills.
Science 319:299–304.

Warren, D. R., and C. E. Kraft. 2003. Brook Trout (Salvelinus fontinalis) Response to Wood Removal
from High-Gradient Streams of the Adirondack Mountains (NY, USA). Canadian Journal of
Fisheries and Aquatic Sciences 60(4):379-389.

Large Wood National Manual July 2015


3-39
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Warren, D. R., and C. E. Kraft. 2008. Dynamics of Large Wood in an Eastern US Mountain Stream.
Forest Ecology and Management 256(4):808–814.

Warren, D. R., E. S. Bernhardt, R. O. Hall Jr., and G. E. Likens. 2007. Forest Age, Wood and Nutrient
Dynamics in Headwater Streams of the Hubbard Brook Experimental Forest. N.H. Earth Surface
Processes & Landforms 32(8):1154–1163.

Warren, D. R., C. E. Kraft, W. S. Keeton, J. S. Nunery, and G. E. Likens. 2009. Dynamics of wood
recruitment in streams of the northeastern U.S. Forest Ecology and Management 258:804-813.

Warren, D. R., J. D. Dunham, and D. Hockman-Wert. 2014. Geographic Variability in Elevation and
Topographic Constraints on the Distribution of Native and Nonnative Trout in the Great Basin.
Transactions of the American Fisheries Society 143:205–218.

Webster, J. R., and E. F. Benfield. 1986. Vascular Plant Breakdown in Freshwater Ecosystems. Annual
Review of Ecology and Systematics 17(1):567–594.

Webster, J. R., J. L. Tank, J. B. Wallace, J. L. Meyer, S. L. Eggert, T. P. Ehrman, B. R. Ward, B. L. Bennet, P.


F. Wagner, and M. E. McTammy. 2000. Effects of Litter Exclusion and Wood Removal on
Phosphorus and Nitrogen Retention in A Forest Stream. Verhandlungen der Internationale
Vereinigung für Limnologie 27:1337–1340.

Wellnitz, T., S. Y. Kim, and E. Merten. 2014. Do Installed Stream Logjams Change Benthic Community
Structure? Limnologica 49:68–72.

White, S. L., C. Gowan, K. D. Fausch, J. G. Harris, and W. C. Saunders. 2011. Response of Trout
Populations in Five Colorado Streams Two Decades After Habitat Manipulation. Canadian
Journal of Fisheries and Aquatic Sciences 68(12):2057–2063.

Whiteway, S. L., P. M. Biron, A. Zimmermann, O. Venter, and J. W. A. Grant. 2010. Do In-Stream


Restoration Structures Enhance Salmonid Abundance? A Meta-Analysis. Canadian Journal of
Fisheries and Aquatic Sciences 67:831–841.

Whittaker, R. H., S. A. Levin, and R. B. Root. 1973. Niche, Habitat and Ecotope. American Naturalist
107(955):321–338.

Williams, K. L., S. W. Griffiths, K. H. Nislow, S. McKelvey, and J. D. Armstrong. 2009. Response of


Juvenile Atlantic Salmon, Salmo salar, to the Introduction of Salmon Carcasses in Upland
Streams. Fisheries Management and Ecology 16(4):290–297.

Williams, R. N. (ed.). 2006. Return to the River: Restoring Salmon Back to the Columbia River. New
York: Elsevier.

Winemiller, K. O., A. S. Flecker, and D. J. Hoeinghaus. 2010. Patch Dynamics and Environmental
Heterogeneity in Lotic Ecosystems. Journal of the North American Benthological Society 29:84–
99.

Wipfli, M. S., and C. V. Baxter. 2010. Linking Ecosystems, Food Webs, and Fish Production: Subsidies
in Salmonid Watersheds. Fisheries 35(8):373–387.

Wipfli, M. S., J. Hudson, and J. P. Caouette. 1998. Influence of Salmon Carcasses on Stream
Productivity: Response of Biofilm and Benthic Macroinvertebrates in Southeastern Alaska, U.S.A.
Canadian Journal of Fisheries and Aquatic Science 55:1503–1511.

Large Wood National Manual July 2015


3-40
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

Wipfli, M. S., J. Hudson, and J. P. Caouette. 2003. Marine Subsidies in Freshwater Ecosystems: Salmon
Carcasses Increase the Growth Rates of Stream-Resident Salmonids. Transactions of the
American Fisheries Society 132:371–381.

Wohl, E., and D. J. Merritts. 2007. What is a Natural River? Geography Compass 1(4):871–900.

Wojan, C., A. Devoe, E. Merten, and T. Wellnitz. 2014. Web-building Spider Response to a Logjam in a
Northern Minnesota Stream. American Midland Naturalist 172(1):185–190.

Wondzell, S. M. 2011. The Role of the Hyporheic Zone across Stream Networks. Hydrological
Processes 25:3525–3532.

Wondzell, S. M., J. LaNier, R. Haggerty, R. D. Woodsmith, and R. T. Edwards. 2009. Changes in


Hyporheic Flow Following Experimental Removal of a Small, Low-Gradient Stream. Water
Resources Research 45:W05406, 13 pp.

Large Wood National Manual July 2015


3-41
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 3. Ecological and Biological Considerations

This page intentionally left blank.

Large Wood National Manual July 2015


3-42
Chapter 4
GEOMORPHOLOGY AND HYDROLOGY CONSIDERATIONS

Point Bar Structure, Salmon River, Near Welches, Oregon. Photo credit: Brian Bair.

AUTHOR

Tim Abbe (NSD)


Brendan Belby (ICF International)
Doug Shields (Shields Engineering, LLC)
This page intentionally left blank.
and lateral bank erosion that results in hundreds
4.1 Introduction of millions of dollars in damage to infrastructure
Geomorphology is the study of landscape throughout the country. It is important to
morphology and the processes that create it. understand the processes by which instream
Hydrology is the study of the water cycle from wood can be a benefit or hazard. Properly
precipitation to how water moves through the designed and placed, engineered wood
landscape (both as surface and subsurface water), placements can be important in managing
and back to the atmosphere. With respect to unstable wood debris that could otherwise be a
wood, both of these topics are critically problem.
important, influencing the type of trees growing
Geomorphic and hydrologic assessments are
within a watershed and the fate of wood within
essential for understanding the past and present
the channel network. Most important is
conditions within a stream, including how wood
understanding that trees and wood influence
naturally influenced the system and how it can be
geomorphology and hydrology. Vegetation can
used to restore it. Wood can be used to have
play an important landscape morphology by
either minor or significant hydraulic and
influencing surface water runoff and erosion.
geomorphic effects depending on the severity of
Forests moderate runoff, diminish downstream
disturbance and site constraints.
flooding, and stabilize stream banks and
hillslopes. Because precipitation provides the
water responsible for surficial erosion, sediment 4.2 Geomorphology
yield should increase with increasing
precipitation. But in reality, erosion and sediment A well-founded understanding of the physical and
yield tends to decrease with increasing biological processes influencing landscape
precipitation due to vegetation (Langbien and development is critical to stream restoration and
Schumm 1958, Figure 4-1) management. Geomorphology is the study of the
Earth’s surface, the processes that formed it, and
Within channel networks, wood can trap how it will change in the future. Fluvial
significant quantities of sediment and organic geomorphology focuses on streams: the flow of
matter, define the morphology of channels and water through a channel network; the movement
floodplains, reduce rates of bank erosion, limit of sediment and woody debris; the factors
long-term rates of incision that influence valley controlling channel form, the stability of stream
formation, and provide habitat resilience to beds and banks, and the rate and magnitude to
extreme events. There are also instances where which channels move; and how large wood and
wood can result in localized erosion by deflecting logjams influence flow conditions to alter the
flow toward a bank. channels and floodplains (Keller and Swanson
Human development since the industrial 1979; Abbe and Montgomery 1996; O’Connor et
revolution has had a dramatic impact on forests al. 2003; Montgomery and Abbe 2006; Collins et
and wood in streams. Establishing and protecting al. 2012; Wohl 2013). To sustain the functions
mature riparian forests along streams provides a wood provides to a fluvial system it is imperative
passive, long-term means of restoring wood and to restore mature riparian forests (Abbe and
the functions it once provided, but that can take Montgomery 1996; Abbe 2000; Keeton et al.
over a century to establish and start to reverse 2007; Collins et al. 2012; Wohl 2013).
the geomorphic changes impacting the stream. One of the most important contributions
Re-introducing wood to a stream can be an geomorphology brings to any project is an
essential part of rehabilitating a stream’s understanding of how and why a stream looks
morphology and ecosystem. Wood can also be a and functions the way it does over time.
cost-effective means for treating channel incision Geomorphologists explain the processes and

Large Wood National Manual July 2015


4-1
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

factors influencing landscape evolution: erosion knock a stream out of equilibrium or alter the
and sedimentation. The climate, intensity and magnitude and rate of morphologic adjustment
magnitude of precipitation, geologic and soil include changes to the flow regime (either
characteristics, topographic relief, vegetation, and increased peak flows due to development or
development all influence the flow of surface and reductions due to dams), changes in riparian
subsurface water and the movement of sediment vegetation or wood loading, or changes to the
through a watershed. The magnitude and rate of stream’s base level (e.g., influence of sea level
changes to the landscape depend on how these fluctuations, tectonics, and landslides on stream
factors change over time. The characteristics of outlet elevation). Streams, by definition, are
any landscape, including stream channels, is the dynamic and subject to spatial and temporal
cumulative result of geomorphic processes over changes (Figure 4-2). Understanding fluvial
time. Morphological changes to stream channels geomorphology of the system you are working in
can occur on timescales ranging from hours to is essential to defining how the system has been
thousands of years. Stream channel morphology impacted, what the current processes are that
will adjust to processes acting on it. If the restoration design must take into account,
processes remain relatively constant over time defining restoration targets, understanding how
(such as periods of climatic stability), channel watershed and external factors will influence the
morphology adjusts to reflect the process regime site and developing designs that accommodate
and can remain relatively constant. This state is these factors, and predicting how the site will
referred to as equilibrium. A sediment budget evolve under different restoration scenarios.
provides a simple illustration of morphologic Channel dynamics influence wood recruitment
equilibrium. If the input of sediment into a stream and the structure of riparian forests. Bedrock
reach is equal to the output, the sediment supply channels tend to have relatively low wood
and sediment transport capacity are in a state of quantities due to low rates of wood recruitment
equilibrium, and no morphologic adjustments are and high transport capacities. Wood stability
likely. In cases where the sediment transport increases when it becomes embedded, so wood is
capacity extends the sediment supply, the output most prevalent in alluvial channels. Alluvial
of sediment exceeds the input. The difference is channel dynamics tend to increase proportional
the erosion within the reach that is enlarging the to discharge and inversely to the grain size of
channel. Conversely, if the sediment output is less their beds and banks. Wood input (recruitment
than the input (the supply exceeds the transport rate) increases with the rate at which adjacent
capacity), sediment storage within the reach is riparian forests are eroded. Therefore, wood
reducing the channel area and altering its loading tends to increase in higher order (larger,
morphology. In stream restoration the application lower gradient) channels, a trend explained in
of a “reference reach” is based on assumption that more detail in Section 1.3.2, Wood Loading in
the reference reach reflects an equilibrium state Natural Settings, above.
under the same set of processes and conditions
In many situations human-induced changes may
that affect the project site. The presence of an
have so altered fluvial processes that finding a
equilibrium state is limited to relatively low relief
reference condition that reflects the natural or
watersheds not subject to major physical or
pre-disturbance condition of the project site will
biologic disturbances. Even in these areas the
not be possible. Understanding what geomorphic
term “dynamic equilibrium” is far more applicable
changes a site has undergone and why is the first
because it refers to a range of morphologic
step in restoring or rehabilitating streams. The
conditions a stream will experience over time,
next step is defining the desired state of the
Any changes to a stream in dynamic equilibrium
stream and understanding whether and how
represent the variance about a mean and do not
wood can be used. Wood can be an essential
reflect long-term adjustments. Factors that will
element in rehabilitating channel form and

Large Wood National Manual July 2015


4-2
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

process, but not in all settings. The role of wood Wallerstein and Thorne 2004) and effectively
will vary by geology, hydrology, location within reduces the shear stress available for sediment
the channel network, the local climate, size and transport (e.g., Manga and Kirchner 2000), it
type of riparian trees, and historic development. follows that wood removal can lead to export of
this sediment, channel incision, and subsequent
Wood is naturally found in alluvial streams
widening (e.g., Guardia 1933; Baker 1979;
throughout the United States, in essentially any
Hartopo 1991; Abbe 2000; Brooks et al. 2003;
location where trees grow along the channel or
Wallerstein and Thorne 2004; Stock et al. 2005;
upstream. Wood provides geomorphic and
Abbe and Brooks 2011; Daley 2012; Phillips
ecologic functions throughout a watershed, from
2012).
the headwaters to estuaries (e.g., Keller and
Swanson 1979; Maser et al. 1988; Maser and
Sedell 1994; Abbe 2000; Abbe and Montgomery GUIDANCE
2003). There are types of channels where wood
isn’t typically found or has relatively little effect, Basic Geomorphic Questions that Apply to Any Project
such as bedrock canyons or stable confined
channels subject to deep flows. Within these  Is the channel profile stable, incising, or
locations there is little local recruitment, and aggrading?
wood entering from upstream tends to pass  What are local and temporal variances in channel
through due to deep fast flows. profile versus long-term trends?
The connection between large wood and channel  Is the channel moving laterally?
processes, substrate, and morphology has been
well documented. Numerous studies—such as,  Is the channel hydraulic geometry stable or does it
change over time or within reaches of similar
but not limited to Baker (1979), Keller and
discharge?
Swanson (1979), Abbe and Montgomery (1996,
2003), Buffington and Montgomery (1999b),  What is the natural variability of stream
Manga and Kirchner (2000), Baillie and Davies morphology over time?
(2002), Stewart and Martin (2005), Magilligan et
 What are the time scales and rates over which the
al. (2008), Montgomery and Abbe (2006),
morphologic change occurs?
Brummer et al. (2006), and Cordova et al.
(2007)—have shown that large wood promotes  Are the current hydrologic, hydraulic, and
in-channel sediment storage as the logs deflect sediment conditions representative of future
flow and increase channel roughness. Large wood conditions?
promotes heterogeneity in channel form by  How have historic disturbances altered
creating flow divergence and changing local base development of the alluvial landscape? Can the
level. These processes lead to sediment system truly be “restored” or simply rehabilitated?
deposition in both upstream and downstream
 What physical controls did wood impose on the
eddies (e.g., Abbe and Montgomery 1996).
system under undisturbed natural conditions (e.g.,
When streams are straightened and confined the pool formation, channel grade, anabranching and
resulting increase in energy can trigger incision, side channels, sediment retention, water surface
profile)?
which then leads to a sequence of morphological
stages that have been described in channel
evolution models (e.g., Schumm et al. 1984; The degree to which wood can influence stream
Schumm 1999; Simon 1989, 1994; Doyle and channels is demonstrated in southeastern
Shields 2000; Simon and Rinaldi 2006). Because Australia where floodplain forests have been
wood can be the dominant grade control in many intact for hundreds of thousands of years. Erskine
streams (e.g., Keller and Tally 1979; Abbe 2000; and Webb (2003) indicated that streams in

Large Wood National Manual July 2015


4-3
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

southeastern Australia that had a history of de- public works and flood control districts
snagging were significantly more incised, had responsible for managing most of our waterways.
higher flow velocities, and were wider than This chapter and the manual as a whole
adjacent streams that were left undisturbed. demonstrate that the perception that wood is bad
Brooks and Brierly (2002) describe the complex is fundamentally flawed. Many of the problems
and stable morphology of the sand-bedded attributed to wood have more to do with
Thurra River in southeastern Australia that is inadequately designed infrastructure, such as an
characterized by high wood loading. Brooks et al. undersized culvert or bridge, and encroachment
(2003) show a dramatic difference in morphology of human development within the floodplain and
between the Thurra and Cann rivers, sites with channel migration zone. The problems also have
similarly sized adjacent catchments, but different to do with the fact that humans have changed the
riparian conditions. The Thurra valley was character of wood entering our streams. Where
preserved and the Cann River was historically we still have riparian forests, the old-growth trees
cleared of riparian vegetation and instream wood. that are inherently stable have been replaced by
Brooks and Brierly (2004) go on to describe how dense forests of small trees, and the wood
the Cann River experienced a 150-fold increase in entering the river is much smaller and simply
the rate of channel migration and a 700% flows downstream to cause problems. Restoring
increase in channel capacity. Examples exist all and using wood can range from moderate
across the United States of channel incision where enhancements of edge habitat in highly
riparian areas were cleared, wood was removed, constrained reaches to valley-altering placements.
channels were straightened, dams have
In the right placements large wood can contribute
impounded bed replenishing sediment, and
to flood and erosion protection, but when
development has increased peak flows. In most
inappropriately placed or managed it can also
cases of channel incision, wood can play a
adversely impact flooding. Riparian forests and
fundamental role in the restoration of fluvial
instream wood can have a significant effect on
processes. Applications of engineered log jams
increasing local flood storage and decreasing the
have been shown to effectively treat channel
celerity or velocity of flood waves (Anderson
incision (e.g., Daley 2012).
2006; Thomas and Nisbet 2006). Instream wood
Instream wood, as with the addition of any adds roughness that slows flow velocities and
channel roughness, will tend to increase water raises water levels, which increases overbank
elevations. Ice jams can have a dramatic influence connectivity (e.g., Brummer et al. 2006).
on river stage and flooding (e.g., Pariset et al. Therefore, the defining attributes of forested
1966; Beltaos 1983; Smith and Reynolds 1983; rivers are trees and wood spreads out a flood
Prowse 2001). Logjams form similar but more hydrograph, increasing the duration of flood
permanent blockages, and historic accounts inundation while reducing maximum flood stage
recognized how logjams obstructed flow to in downstream reaches (e.g., Anderson 2006;
impound rivers to create lakes (Lyell 1830; Thomas and Nisbet 2006). This provides a very
Guardia 1933; Harvey et al. 1988; Barrett 1996). important ecological enhancement by increasing
The widespread presence of wood in rivers of the the duration of floodplain inundation and water
United States led to aggressive wood removal by retention in reaches with high wood loading,
the federal government with the intent of which in turn sustains important aquatic refugia,
improving navigation and drainage (e.g., Ruffner traps organics and fine sediments, and recharges
1886; Collins et al. 2002). shallow groundwater. Logjams form preferred
habitat for largemouth bass (Micropterus
Channel clearing has had dire consequences to
salmoides) in the low-gradient sand-bedded rivers
the geomorphology and ecology of streams that
of the mid-Atlantic coastal plain (Schenk et al.
largely remain unrecognized, particularly by the
2014a), flathead catfish (Pylodictis olivaris) in

Large Wood National Manual July 2015


4-4
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

rivers of the Great Plains region (Paukert and Demko 2011). Some fraction of the wood will end
Makinster 2008), and smallmouth bass up in lakes or the ocean, and will continue to play
(Micropterus dolomieu) and rock bass a fundamental ecological role (Harmon et al.
(Ambloplites rupestris) in the Great Lakes (e.g., 1986; Maser et al. 1988; Maser and Sedell 1994).
Bovee et al. 1994). Wood can also store significant
Just as sediment can accumulate into distinct
quantities of sediment and organic matter that
bedforms depending on flow conditions and the
would otherwise move downstream where it
characteristics of the grain size distribution, wood
could aggravate flooding. Wood restoration
can accumulate into distinctive deposits. Studies
projects can therefore be the principal means of
have documented unique types of wood
restoring both channel and floodplain habitat and
accumulations or jams in the Pacific Northwest
can indirectly contribute to reducing downstream
(Abbe 2000; Abbe and Montgomery 2003),
flooding. Complex wood structures have been
northern New York (Kraft and Warren 2003;
successfully applied to protect banks in a manner
Keeton et al. 2007), northern Michigan (Morris et
that enhances instream and riparian habitat (e.g.,
al. 2010), South Carolina (Wohl et al. 2011), and
Abbe et al. 1997, 2003b, 2003c; Brooks et al.
the Italian Alps and Chilean and Argentinean
2004, 2006; Abbe and Brooks 2011).
Andes (Comiti et al. 2006, 2008). Unique natural
4.2.1 Wood Structures wood accumulations occur in different parts of a
channel network and have different geomorphic
Trees enter streams through a variety of effects (Figure 4-3A). Each of these natural
mechanisms, such as landslides, avalanches, accumulations have inherent physical complexity
windstorms, fires, and bank erosion. due to the size and shape of the trees forming the
snags. Natural accumulations also include the
accumulation of mobile wood debris, referred to
CROSS-REFERENCE
as “racking” material. The recognition that natural
wood accumulations or logjams can influence
How wood enters and behaves in streams is described channel morphology, limit channel incision, and
in Chapter 1, Large Wood Introduction. protect floodplain areas led to the concept of ELJs
(Figure 4-3B) (Abbe et al. 1997, 2003b, 2003c;
Once in a channel, wood can remain exactly Brooks et al. 2004, 2006; Abbe and Brooks 2011).
where it entered and stay for centuries, or it can Wood placement in streams for improving habitat
move downstream depending on its size and the is not new (e.g., Van Cleef 1885; Hewitt 1934;
transport capacity of the channel. Wood that does Thompson 2002, 2005), but emulating natural
not move far from where it entered will complexity is unique to recent efforts to re-
accumulate mobile wood and form logjams. Wood introduce wood to streams. Structure and channel
that moves downstream but ends up embedded in complexity are defining characteristics of
the channel creates snags that can also initiate engineered logjams that makes some of the
logjams. The geologic record shows that wood current work on wood placement unique from all
entering a channel network can end up preserved historical efforts. Complexity is used to refer to
in alluvial sediments. Much of the wood deposited the architecture of individual structures designed
on bars and floodplains will decompose or be to include complex shapes and assortments of
consumed within decades by fungi or termites wood, as well as collect and shed wood debris
(Hyatt and Naiman 2001; Scherer 2004; Latteral through time. Complexity also refers to the spatial
and Naiman 2007). Wood buried in the channel or arrays of ELJs that are used to rehabilitate fluvial
floodplain that remains within the water table or processes and morphology, such as restoring
anaerobic conditions can persist for very long anabranching, channel systems, floodplain
periods (Guyette and Stambaugh 2003; connectivity, and diverse riparian forest
Montgomery and Abbe 2006; Gestaldo and communities.

Large Wood National Manual July 2015


4-5
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-1. Although Precipitation Increases Surficial Runoff, Erosion Rates Diminish (as measured by
sediment yield) due to the Influence of Vegetation

Adapted from Langbein and Schumm (1958).

Large Wood National Manual July 2015


4-6
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-2. Illustration of Several Basic Fundamentals of Fluvial Geomorphology, including Spatial and
Temporal Change over Time, the Importance of Sediment Budgets, and the Role of Wood

Large Wood National Manual July 2015


4-7
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-3. (A) General Distribution of Natural Wood Accumulation Types Within a Watershed; (B)
Application of Four of Those Types to Engineered Logjam Structures

Sources: (A) Abbe and Montgomery (2003); (B) Abbe et al. (2003).

Large Wood National Manual July 2015


4-8
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

John Muir (1878) wrote of “sequoia stream-


4.2.2 Big Trees making,” the process by which the falling of a
single giant sequoia (Sequoiadendron giganteum)
Wood is a common component of the particulate in the Sierra Nevada of California would impound
matter in streams throughout the world. streams and capture sediment, a process that
created a series of bogs that merged together to
provide the ideal growing conditions for these
CROSS-REFERENCE
trees in an otherwise dry and steep landscape.
Veatch (1906) noted how large trees would form
A detailed description of wood in streams is provided
in Chapter 3, Ecological and Biological Considerations. “planters” in the bed of the Red River, Louisiana,
which would then create logjams capable of
impounding and redirecting the river. Eighteen
In many areas wood constitutes the largest feet of bed aggradation behind logjams initiated
individual particles found in the stream by large fallen trees was observed in the Middle
(Figure 4-4). The size and shape of a tree is Fork Teanaway River of Central Washington State
critical with regard to its stability in a channel and (Russell 1898). Recent research has
how it can influence channel morphology (Abbe demonstrated that following logging and channel
and Montgomery 1996, 2003; Abbe et al. 2003a; clearing, alluvial channels in West Fork Teanaway
Montgomery et al. 2003; Abbe and Brooks 2011). River disappeared and the river cut 1–2 meters
Keeton et al. (2007) found the number of stable (3.3–6.6 feet) into the underlying bedrock at a
debris jams increased as a function of large logs in mean rate of 30 millimeters (1.2 inches) per year,
Adirondack streams of northern New York, with 600 times faster than the geologic incision rate
much greater frequency in old-growth forests. (Stock et al. 2005). Hazard (1948) wrote of fallen
Therefore, the size of riparian trees directly cedar trees blocking the lower Quinault River, and
affects the magnitude of the geomorphic effects Wolff (1916) noted that “a close study of
and extent of the channel network affected. conditions shows that in every instance the
Almost all of the old-growth riparian forests in current was first deflected by an accumulation of
the United States have been cleared, but historical drift, the huge timber of this section serving
accounts not only describe the large trees that readily in its formation.” From the few records of
once existed, but the effect they had on altering what rivers were once like, we can get a glimpse
rivers. The eastern sycamore (Platanus of historic wood loading (Figure 4-6).
occidentalis), native to stream banks throughout Today, to restore wood within the full range of
much of the East and Midwest (Figure 4-5), environments it once influenced, we need to
historically formed giant snags even in the largest engineer solutions using smaller trees to provide
rivers such as the Mississippi (Dacy 1921). When the function large trees once provided.
George Washington visited the Ohio Valley in Engineered wood placements have been
1771 he noted sycamores with basal diameters successfully used to reverse incision of rivers
over 4 meters (13 feet) (Federal Writers’ Project impacted by historic clearing and splash damming
1952). American chestnuts (Castanea dentata) in (Abbe and Brooks 2011). In low-order streams a
the Appalachian Mountains reached diameters of single piece of wood can have dimensions easily
5 meters (16 feet) (Grimm 1967). The old-growth exceeding those of the channel itself and create
bald cypress (Taxodium distichum) forests that steps that can account for the majority of the
included trees 4.5 meters (15 feet) in diameter, vertical drop of a channel (e.g., Keller and Tally
currently cover only 0.01% of the 131 million 1979; Montgomery et al. 1995b, 1996b;
acres in the southeastern United States that they Montgomery and Buffington 1997; Abbe 2000;
did just 200 years ago (Stahle et al. 2006). Abbe and Montgomery 2003). In larger order
channels a piece of relatively large wood can form

Large Wood National Manual July 2015


4-9
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

the nucleus of much larger accumulations (i.e., R = hydraulic radius = A/P, m


logjams) that can redirect currents, alter channel
A = cross-sectional area of flow—square meters
planform, or even completely block the channel
(e.g., Abbe and Montgomery 1996, 2003). It is P = wetted perimeter of cross-section—meters
now well recognized that wood can be the
principal control in channel avulsions and S = energy gradient ~ channel gradient
anabranching (Figure 4-7) (Hickin 2004; Sedell n = Manning’s roughness coefficient
and Frogatt 1984; Abbe and Montgomery 1996;
Makaske et al. 2002; Abbe et al. 2003a; Because discharge, Q, is the product of mean
Montgomery and Abbe 2006; Sear et al. 2010; velocity, U, and cross-sectional area of the flow, A,
Phillips 2012) and defining the structure of substituting the Manning’s expression for U
floodplain forests (Figure 4-8) (Collins et al. yields:
2012). Equation 4-2:

4.2.3 Hydraulic Influence of 𝐴 𝑅 2/3 𝑆 1/2


𝑄=
𝑛
Wood
Observations described above are largely due to
the hydraulic influence of wood, particularly in For a wide rectangular channel cross-section, R
reducing the stream’s erosive capacity by can be approximated by average channel depth, h.
partitioning shear stress. Wood induces Because the cross-section area, A, is simply width.
geomorphic change by controlling the hydraulic w, multiplied by depth, h:
processes that erode, transport, and deposit Equation 4-3:
sediments. A detailed discussion regarding the
forces acting on wood is provided in Chapter 6, ℎ 𝑤 ℎ2/3 𝑆 1/2
𝑄=
Engineering Considerations, but because wood can 𝑛
influence sediment transport and bank erosion, a 𝑤 ℎ5/3 𝑆 1/2
brief discussion is included here. =
𝑛
Stable wood adds roughness to a channel that can Solving for depth gives:
range from bed texture to obstructions that
Equation 4-4:
occlude the entire channel. The simplest hydraulic
expression that accounts for frictional energy loss 𝑄𝑛 3/5
due to channel roughness is the commonly used ℎ=( )
(𝑤𝑆 0.5 )
Manning’s equation that incorporates a roughness
coefficient, the Manning’s n: Based on this simple expression we see that water
depth increases with roughness when other
Equation 4-1:
variables are held constant.
𝑅 2/3 𝑆 1/2
𝑈=
𝑛
where

U = mean flow velocity—meters/second

Large Wood National Manual July 2015


4-10
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-4. Wood is Typically the Largest Bed Material Entering Streams and Tends to Get Larger in Lower
Elevations of a Watershed (Larger Channels), the Inverse of Rock Particles

(A) A plot of particle size illustrating the range of tree size. (B) Looking at the size of snags relative to
different channel dimensions.

Large Wood National Manual July 2015


4-11
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-5. Big Trees Were Historically Common Along Streams Throughout the United States

(A) Western Red Cedar, Washington. (B) American Sycamore, Indiana. (C) Bald Cypress, Arkansas (Stahle et
al. 2006). (D) Fremont Cottonwood, Arizona.

Large Wood National Manual July 2015


4-12
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-6. (A) Snags and Logjams, Were Common Throughout Much of the Missouri and other Midwestern
Rivers, as Depicted in this Illustration by George Catlin in 1832; (B) Undated Photo, Circa Early 1900s, of a
River on the Olympic Peninsula of Washington Loaded with Sitka Spruce (Picea sitchensis) Snags

Large Wood National Manual July 2015


4-13
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-7. (A) Historic Changes to the Upper Willamette Transforming the Natural Anabranching
Morphology into a Single-Thread Channel; (B) Lower Taiya River, a Wood-Rich Anabranching River in
Southeastern Alaska

Sources: (A) from Sedell and Frogatt 1984; (B) from Abbe et al. 2003b.

Large Wood National Manual July 2015


4-14
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-8. Comparison of an Alluvial River with Wood (Hoh River, Washington) to One Where Wood Has
Been Removed (Cowlitz River, Washington)

Source: Collins et al. (2012).

Manning’s n tends to diminish with increasing blockage coefficients rise above 0.1 (wood is
water depth, except in cases where wood extends obstructing 10% of channel cross-sectional area),
through the range of water depths and is even there is a substantial (non-linear) decrease in
suspended over the current channel, in which conveyance and an associated increase in water
cases there may be no change or even an increase elevations (depths).
in n. When the roughness begins to reduce the
Manning’s n is the sum of all the factors
channel’s cross-sectional area, w, the increase in
contributing roughness or frictional energy loss:
water depth is even greater. Wood has to obstruct
10% or more of a channel before it has an Equation 4-5:
appreciable effect on conveyance and stage
𝑛 = 𝑛𝐺𝑆 + 𝑛𝐵𝐹 + 𝑛𝑊𝐷 + 𝑛𝑃𝐹 + 𝑛𝑜𝑡ℎ𝑒𝑟𝑠
(Gippel et al. 1992, 1994; Gippel 1995). When

Large Wood National Manual July 2015


4-15
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

where: wood structure. Secondary types of drag that are


usually not considered are skin friction (or
n = Manning’s roughness coefficient
surface) drag and wake drag. Skin drag is due to
nGS = roughness contribution from grain size friction water encounters along the surface area
of bed material or skin of the wood. An object in flowing water
also creates a wake that dissipates energy and
nBF = roughness contribution from bedforms thus contributes to drag. When flow moves
nWD = roughness contribution from wood through an array of objects (pieces of wood
comprising a logjam), the wakes begin to interfere
nPL = roughness contribution from channel with one another. The closer the spacing of the
planform = f (sinuosity, width variance) array, the greater the wake interference and the
greater the reduction in the water’s local
nothers = roughness contribution from other
approach velocity, until the array (logjam)
factors
behaves like a bluff body defined by its form drag.
Shields and Gippel (1995) and Buffington and
Because drag is a measure of the energy losses, it
Montgomery (1999a, 1999b, 1999c) demonstrate
can be used to illustrate how wood reduces a
how the roughness contribution of wood
stream’s energy available to do work such as
partitions roughness, alters flow conditions, and
sediment transport and bank erosion. Shear
creates more complex bed textures. Examining
stress is commonly defined by the depth slope
flow around a snag illustrates some of the basic
product:
hydraulic effects wood can impose and how it can
change channel bed topography and substrate o =  g h S
characteristics (Figure 4-9). Unobstructed flow
vectors are dominated by a downstream where:
horizontal component. As flow approaches a bluff o = shear stress (Pa),
body such as a snag or logjam, the downstream
horizontal component diminishes while the  = fluid density (kg/m3),
downward and lateral components increase. This
g = gravitational acceleration (m/s2),
generates vortex flow and bed scour immediately
upstream of the obstruction. Flow around a bluff h = fluid depth (m), and
body creates three distinct domains of flow:
(1) accelerated downstream flow in the S = energy gradient (~water surface slope).
constricted unobstructed portion of the channel, This expression represents the total shear stress
(2) an eddy of lower velocity recirculating flow acting on the bed and is thus the sum of the shear
immediately downstream (leeward) of the stresses acting on different objects within the
obstruction, and (3) a “vortex street” separating stream channel, principally sediment and wood.
the two domains (Figure 4-9). Wood imposes So the more wood engages flow, the greater the
fluid resistance that is referred to as drag and is shear stress acting on wood and the less shear
proportional to the fluid velocity squared in stress there is to transport sediment or erode the
turbulent flow that occurs in streams. There are channel. This concept is referred to as stress
several different types of drag associated with partitioning and is key to understanding some of
instream wood, The most important is form (or the hydraulic and geomorphic effects of wood.
pressure) drag, which results from the shape and
relative size (to the wetted cross-section) of the

Large Wood National Manual July 2015


4-16
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-9. Flow Around a Stable Snag

Scour around the rootwad and sedimentation within the eddy increases passive earth resistance stabilizing
the snag. The ratio of the snag (or logjam) width, WW, to the channel width, WC, is defined as the blockage
coefficient, B. The upstream width of the snag or logjam defines the downstream flow separation envelope
and recirculation zone (eddy). The length of the obstruction has no effect on the re-circulation zone. The
approach velocity and shape and permeability of the structure affect vortex development at flow separation.
In cases where vortex diameter, DA, approaches 0.5WW, turbulent exchange in the re-circulation zone limits
sediment accumulation. Therefore, conditions where WW>>2DA best promote the development of bars and
islands. The deposition of bed material increases resisting forces by adding surcharge (vertical load) and
passive earth pressure (lateral load) that counteract buoyant and drag forces to help stabilize the wood (e.g.,
Abbe 2000; Abbe et al. 2003b; Abbe and Brooks 2011). The re-circulation zone creates hydraulic refugia for
fish, and sediment deposition can create areas more suitable for spawning.

The hydraulic effects of wood are also reflected in allows wood to become more embedded within
the flow patterns within the stream. Individual the stream (increasing stability) or remove
snags or logjams create obstructions that alter sediment that provided resistance (decreasing
streamlines and increase turbulence. stability). If the wood forms a large enough
Downstream velocities are reduced immediately obstruction it creates an eddy of recirculating
upstream of an obstruction, then accelerate as lower velocity flow (where sediment can settle
they move around. The vortices that form can out) and increasing stability by adding surcharge
contribute to bed scour that can affect the wood’s (burial) and passive earth pressures (buttressing)
stability (both positively and negatively). Scour (Abbe and Montgomery 1996; Abbe 2000; Abbe

Large Wood National Manual July 2015


4-17
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

et al. 2003b; Brooks et al. 2004, 2006; Abbe and Brooks 2011). For snags that aren’t large enough,
Brooks 2011). Natural snags are often observed the increase in buoyancy associated with
embedded in a river channel with their tips settlement can destabilize them, and they can
pointed downstream (Figure 4-10). move down the river by spinning 360o before
stopping again and the process repeating itself.
Scour at the upstream end of a snag allows the
Engineered solutions can be to simply install
rootwad or basal end of the snag to settle into the
posts or piles on either side of a snag immediately
stream bed (Figures 4-9 and 4-10). Embedment
downstream of its rootwad, thereby providing the
dramatically increases the resisting forces acting
passive earth pressure resistance of a large
on the snag and its stability (Abbe and
buried rootwad.
Montgomery 1996; Abbe et al. 2003b; Abbe and
Figure 4-10. Process by Which a Snag Becomes Imbedded in a Channel Bed

Downward and lateral acceleration of water velocity upstream of snag (see Figure 4-9) produces vortices that
scour the bed around the rootwad (A, B). If the snag has net downward gravitational force, the rootwad
settles into the stream bed (B, C). Adapted from Abbe (2000).

Large Wood National Manual July 2015


4-18
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Manga and Kirchner (2000) provide an ρ = density of water (1,000 kilograms per
examination of partitioning shear stress to cubic meter)
demonstrate how wood reduces the available U = design flow in channel; suggested V100
energy for sediment transport, thereby
CD is dependent on Reynolds number (Re =
diminishing the median grain size of a channel
bed: u*D50/), Froude number (Fr = U/(gh)0.5), the
object’s shape, and the object’s orientation. Drag
Equation 4-6: coefficients for wood have been estimated in
𝜏0 = 𝜏𝐺𝑆 + 𝜏𝐵𝐹 + 𝜏𝐿𝑊 + 𝜏𝑜𝑡ℎ𝑒𝑟𝑠 several laboratory and field studies (e.g., Shields
and Gippel 1995; Hygelund and Manga 2003;
where: Manners et al. 2007; Shields and Alonso 2012).
Measurements of CD found that the extent to
0 = total bed shear stress, gRS
which wood obstructs or blocks flow can
 = water density
influence drag. Shields and Gippel (1995) used
g = gravity
experimental data (Gippel el al. 1992, 1994) to
R = hydraulic radius
derive an expression for the apparent drag
S = slope of energy grade line
coefficient based on blockage:
GS = grain stress that is effective shear
stress available to sediment transport Equation 4-8:
BF = stress component due to wood 𝐶𝐷
LW = stress component due to wood 𝐶𝐷𝑎 =
(1 − 𝐵)2
others = stress component due to wood
where:
Manga and Kirchner (2000) present three
different arguments for how wood reduces the u* = shear velocity = U  (ln(h/0.258 D90))1
shear stress available to do work on the bed and (Wilcock et al. 1996)
banks of a channel based on drag force it imposes:  = dynamic viscosity (0.00089 N s/m2)
k = von Karman’s constant = 0.4
1. Drag associated with water velocity D90 = grain size for which 90% of bed is finer
2. Drag inferred from water slopes D50 = median grain size of the bed
CDa = apparent drag coefficient due to
3. Drag from water steps
blockage effect
All three of these effects are visually apparent in B = blockage coefficient = ALW/Ac
steep channel wood accumulations, such as those ALW = area of wood projected normal to flow
illustrated in Figure 4-10 where the wood is Ac = cross-section area of flow
clearly slowing down flow velocities, reducing
The blockage coefficient, B, can be reduced to the
water slopes, and forming distinct steps in the
ratio of the submerged height of wood, H, to
channel profile. The force of water acting on an
average water depth, h. For a submerged log, H is
obstruction such as wood is expressed as
equivalent to log diameter (Manga and Kirchner
Equation 4-7: 2000):
𝐹 1 Equation 4-9:
= 𝜌 𝐶𝐷 𝑈 2
𝐴𝐿𝑊 2
𝐻
𝐵 =
where: ℎ

F = drag force per unit area Dividing the drag force imposed on wood by the
CD = drag coefficient channel bed area provides an estimate of the
ALW = area of projection normal to flow (e.g., reach-averaged resistance or drag due to wood:
ELJ width * depth)

Large Wood National Manual July 2015


4-19
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Equation 4-10: where ℎ̅ = average water depth.


𝐻 2 The slope reduction between a series of log steps
𝜏𝐿𝑊 = 𝜌 𝐶𝐷𝑎 𝑈
2𝐿 provides a simple illustration. By dissipating
where L = distance between wood. energy in vertical drops, the steps reduce the
energy throughout the channel length. Sediment
The role of submerged wood in stress partitioning storage between the steps reflects the reduction
can then be examined using a force balance in transport capacity resulting from the steps. In
equation using only bed texture and large wood their example, Manga and Kirchner (2000) found
roughness (Manga and Kirchner 2000). The first the water surface slope between wood is about
term in the equation below is equivalent to GS, half the reach-averaged slope, so the large wood
the second term to LW, and the third to 0. must be responsible for about half the resistance
to flow. Using a form of the Bernoulli equation,
Equation 4-11:
assuming h=0.36m, Sb=0.0035, and Sw=0.0018,
𝐻 2 then LW = 6.0 Pa.
𝜌𝐶𝐵 𝑈 2 + 𝜌𝐶𝐷𝑎 𝑈 =𝜌𝑔ℎ𝑆
2𝐿
Manga and Kirchner (2000) go on to show the
GS LW 0 effect of wood by looking at water surface steps.
where CB = drag coefficient for the channel bed. The magnitude of energy dissipation over a step
can be estimated by defining energy per unit
Another approach examined by Manga and volume in cross-section of flow.
Kirchner (2000) uses water surface slope
Equation 4-14:
measurements to infer drag. They measured a
reach averaged bed slope, Sb, of 0.00346, but the 1
average water surface slope, Sw, between large 𝐸𝑚 = 𝜌𝑔(𝑧 + ℎ) + 𝜌 𝛼 𝑈 2
2
wood was considerably less, between 0.0009 and
0.0021. Therefore, in the sub-reach scale between
𝜌𝛼𝑞 2
large wood the flow is not uniform (bed and = 𝜌𝑔(𝑧 + ℎ) +
2ℎ2
water slopes are not equal). The energy gradient,
S, can be estimated using energy arguments
where:
(Robert 1997):
z= bed elevation above arbitrary datum
Equation 4-12:
q= discharge per unit width
S = Sw – Fr2 (Sw – Sb) h= local water depth
ℎ̅ = average water depth
where:
= coefficient between 1 and 1.4 that accounts
Fr = Froude number = U/(gh)0.5 for the fact that the square of velocity is
Sb = average bed slope somewhat less than the mean of the
Sw = average slope between wood squared velocities of the individual water
parcels comprising the flow (Richards
In the example from Manga and Kirchner (2000), 1982)
Fr = 0.19. The average bed stress, GS = ghSw. The
additional resistance provided by wood causes a Manga and Kirchner (2000) show that if the water
reduction in slope of the water surface and is surface elevation drops by h (from h+h/2 to
expressed by h-h/2), the energy, Em, will be reduced by Em,
Equation 4-13:

LW = g ℎ̅ (Sb – Sw)

Large Wood National Manual July 2015


4-20
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Equation 4-15: The difference between the average bed slope, Sb,
and average water slope between wood, Sw, is
𝜌𝛼𝑞2 Δℎ equal to ratio h/L, so Equation 4-16 is
Δ𝐸𝑚 ≈ −𝜌𝑔∆ℎ +
ℎ̅ℎ2 equivalent to

Loss of Gain of Equation 4-17:


potential energy kinetic energy
LW = g (Sb-Sw).
If gh >> U2
(equivalent to the condition that is Fr2
The effect of wood on reducing sediment
small), then change in kinetic energy will be small
transport capacity is directly proportional to the
compared to change in potential energy (Manga
relative head loss (h/z), where z is total
and Kirchner 2000). The effect of large wood in
vertical drop channel and h is vertical drop due
terms of an average shear stress is the sum of
to wood. Wood can account for more 90% of the
energy loss over the cross-section, and averaging
vertical drop in natural channels (Keller and Tally
over the area of the bed yields
1979; Abbe 2000; Abbe and Montgomery 2003).
Equation 4-16: Energy dissipation by wood affects flow
conditions, channel form, and sediment storage
Δℎ
𝜏𝐿𝑊 ≈ 𝜌𝑔ℎ̅ (Figure 4-11). These effects can occur throughout
𝐿
much of a channel network, from steep headwater
channels to large low-gradient channels.
where L is the spacing between large wood and
the respective water surface steps.

Figure 4-11. Natural Log Steps Influencing Water Elevations and Distribution of Shear Stress
in Fisher Creek in the North Cascades, Washington

The wood impounds the channel into a series of steps that lower the water surface slope, which partitions
shear stress and lowers the shear stress available for sediment transport. This increases sediment storage
and reduces the median grain size of the bed.

Large Wood National Manual July 2015


4-21
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

4.2.4 Channel Morphology The presence and age of riparian forests have a
significant effect on stream channel morphology.
The physical effect of wood is clearly evident Lunetta et al. (1997) found the percentage of
when looking at the variations in bed topography forced pool-riffle reaches went from 100% for
and texture variability of channels with and channels with late seral stage riparian buffers (30
without wood (Figure 4-12). Montgomery and meters [98 feet] on each bank) to 35% in non-
Buffington (1997) found that channel morphology forest lands (urban, agriculture, rangeland).
at the reach scale is not just controlled by Ditching, diking, and dredging in floodplains
discharge and sediment supply, but also wood. In primarily found in urban and agricultural regions
their channel classification they demonstrate how was associated with 73% of the coho salmon
wood can “force” changes in bed morphology such rearing habitat losses in the Skagit River system
as transforming a plane bed into a pool-riffle (Beechie and Wyman 1992). Rot et al. (2000)
channel type. Montgomery and Buffington (1993) show that the number of stream pools with
stratify specific channel morphologies into three residual depths > 0.5 meter (1.6 feet) increases
basic reach categories as a function of reach rapidly with riparian forest stand age,
average slope (Table 4-1). diminishing only after stands reach ages of more
than 200 years (Figure 4-13A). Hilderbrand et al.
Low-gradient channels (“response” reaches) are
(1997) found that pool area increased after wood
particularly susceptible to morphologic alteration
placement in low-gradient streams of southwest
due to changes in discharge and sediment load,
Virginia.
which can result from land development (e.g.,
Hammer 1972; Leopold 1973; Graf 1975; Dunne
and Leopold 1978; Booth 1990, 1991; Booth and
Reinelt 1993; Moscrip and Montgomery 1997).

Table 4-1. Channel Reach Classification

Reach Category Channel Reach Slope (S) Typical Channel Morphology (Pacific Northwest)
Source  0.20 Headwater colluvial channels prone to debris flows

Transport 0.04  S < 0.20 Cascade and step-pool

Response S < 0.04 Plane-bed,1 forced pool-riffle2 (0.01 < S < 0.04)
Riffle dominated pool-riffle1 (0.01  S < 0.02)
Pool-riffle3 (S < 0.01)
Source: Montgomery and Buffington (1997).
1 Low large wood loading

2 High large wood loading

3 Independent of large wood loading, but large wood loading will control pool frequency and the

morphologic complexity of the channel (e.g., Buffington and Montgomery 1999b, c; Abbe 2000).

Large Wood National Manual July 2015


4-22
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-12. Examples of Alluvial (Gravel-Bed) Stream Channels With Low Wood Loading (A) and High
Wood Loading (B)

Source: from Buffington and Montgomery (1999b, Figure 8a, page 3515 and 8b, page 3516).

Large Wood National Manual July 2015


4-23
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-13. (A) Correlation Between Percent of Large Wood Pools (with residual depth > 0.5 meter [1.6
feet]) Formed by Wood as a Function of Riparian Forest Stand Age; (B) Frequency of Textural Patches as a
Function of Wood Pieces per Reach for Streams Draining the West Slope of the Olympic Mountains in
Northwestern Washington

Source: (A) Rot et al. (2000, Figure 6, page 704); (B) Buffington and Montgomery (1999b, Figure 9, page
3518).

Large Wood National Manual July 2015


4-24
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Urbanization tends to increase peak flows in a (eddy) downstream of the obstruction, where
basin (James 1965; Hollis 1975) by removing flow is constricted around the obstruction; (ii) the
vegetation (decreasing evapotranspiration and streamline zone of principal flow past the
interception) and primarily by decreasing soil obstruction; and (iii) a shear layer separating
permeability through compaction and impervious (i) and (ii) sometimes referred to as the von
surfaces. The hydrologic effects of urbanization Karman vortex street. All of these flow patterns
are usually assessed by estimating the percentage result in a complex assemblage of dramatically
of impervious surface of a drainage area (e.g., different velocities and depths within a very small
Dunne and Leopold 1978). An increase in the area, each usually associated with different
frequency of peak flows (decrease in recurrence substrate textures. Mapping of textural patches
interval of a particular discharge) goes on to within a channel (Buffington and Montgomery
directly alter channel morphology, primarily 1999c) can provide valuable insight into the
through increases in depth (i.e., incision) and hydraulic characteristics of a channel.
width (Hammer 1972; Leopold 1973; Booth 1990,
These physical responses translate into extremely
1991) and stream ecology (Booth and Reinelt
beneficial habitat for different salmonid species
1993; Luchetti and Fuerstenberg 1993). The
and life stages. When the flow obstructions are
frequency of peak flows increases significantly
formed by snags (fallen trees) and logjams, they
when a catchment is urbanized. Moscrip and
also provide intricate cover and shade. Buffington
Montgomery (1997) report that flows with a 10-
and Montgomery (1999a, 1999b) demonstrate
year recurrence interval prior to urbanization
how the presence of wood can dramatically alter
occurred with a 1- to 4-year recurrence interval
the texture and topographic complexity of a
after urbanization of 14% or more of catchments
channel (Figure 4-12) and offer quantitative
in the Puget Sound lowlands. These results were
means of assessing stream condition.
consistent with predictions by Booth (1990) that
urbanization would transform 10-year flows into As the frequency of functional wood (stable wood
2- to 5-year flows within the Puget Sound region. impinging on flow) increases in a reach, the rate
Wood can be an important element in moderating by which the number of textural patches in the
these increases in peak flows. reach increase is initially exponential, then
gradually diminishes (~20 pieces/reach in
As discussed above, stable wood in stream
Figure 4-13B). As wood loading increases there is
channels can have significant hydraulic effects by
a decrease in pool spacing (inverse of pool
increasing boundary roughness and forming flow
frequency) that approaches a constant value at
obstructions. The presence of flow obstructions is
wood loading of about 0.03 piece per square
probably the single most effective means of
meter (Figure 4-14A). Wood increases the
increasing the diversity and range of physical
complexity of channel topography, bed textures,
habitat. As flow approaches an obstruction, its
and substrate material (organic and inorganic).
downstream horizontal velocity diminishes and
Channel complexity increases ecological
its vertical velocity accelerates part-way down the
productivity and resilience (e.g., Power et al.
water column before decelerating to zero close to
1995; Power and Dietrich 2002). Stability of a
the bed where the flow can be directed upstream
piece of wood is dependent on its size relative to
(Abbe and Montgomery 1996; Abbe 2000). The
the channel’s hydraulic geometry, its density, and
horizontal component of flow normal to the
its shape (Abbe 2000; Abbe et al. 2003b; Abbe
original streamlines then accelerates as flow is
and Brooks 2011). Size (length and diameter) and
constricted around the obstruction. Vortices are
density will affect a log’s weight and buoyancy
generated directly upstream of the obstruction
under particular flow conditions and the
that can scour the bed. Flow “separates” as it
resistance it may encounter with the channel bed,
moves past the obstruction, forming three distinct
banks, or pre-existing obstructions. Shape can
flow regions (Abbe 2000): (i) a recirculation zone

Large Wood National Manual July 2015


4-25
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

have a pronounced effect on how the weight of a with numerous forested islands into single, wide
log is distributed and the frictional resistance the meandering channels lacking smaller channels or
log encounters within the channel (Abbe 2000; highly braided and dynamic wide channel
Abbe et al. 2003b, 2003c; Abbe and Brooks 2011). networks (Figure 4-17; Abbe et al. 1997). Adding
Measurements of key, racked, and loose pieces of logjams can not only increase the number of pools
wood in five different channel reaches of the within a channel segment, but can also increase
Queets River system in northwestern Washington the range of pool depths within the system, as
(west slope of the Olympic Peninsula) provide an seen after ELJs were constructed in the lower
empirical means for estimating the size log (i.e., Elwha River (Figure 4-18A). The same channel
tree) necessary to form key members, based on segment where the ELJs were installed also
the average bankfull width and depth of the experienced a significant reduction in median
channel (Figure 4-14B). In many relatively small grain size, consistent with the stress partitioning
channels where the key piece size can be obtained done by the wood (Figure 4-18B). This reduction
for creating functional wood (e.g., Abbe and in grain size can then be used to demonstrate how
Montgomery 1996, 2003), simply adding wood to the ELJs could alter channel morphology.
the channel can result in a significant
Eaton et al. (2010) examined thresholds between
improvement in habitat (Figure 4-15).
single thread, anabranching (or anastomosing),
One of the principal means by which wood and braided channels. Using Elwha River data
increases physical complexity is in splitting flow shows how the reduction in grain size increases
to create islands and multi-thread channel the dimensionless formative discharge and
systems referred to as anabranching or pushes the Elwha channel from a single thread to
anastomosing channel patterns. Anabranching is anabranching form (Figure 4-19). This is exactly
the most effective means of adding channel length what happened where the ELJs were installed. By
and edge habitat to a river. In addition to island splitting flow and raising water elevations, wood
formation, logjams also create secondary can have a dramatic effect on the quantity and
channels by raising water elevations high enough quality of aquatic habitat. Using channel bank
for overbank flows to carve new floodplain length as a metric for edge habitat, we can see
channels. This process has been observed in a that in an unconfined anabranching channel reach
range of different physiographic regions there is significantly more habitat than in incised
throughout North America (e.g., Hickin 1984; and leveed reaches of the exact same river
Abbe and Montgomery 1996, 2003; Webster et al. (Figure 4-20). Hydraulic modeling of the Lower
2002; Phillips 2012; Wohl 2013). Measurements White River in Washington shows that over a
from the Queets River show how channel wide range of flow discharge, the wood-rich
meanders with logjams have significantly smaller anabranching reach of the river has far more bank
radii of curvature than those without (Figure 4- length than reaches with a single-thread channel
16A). Decreasing the radius of curvature of a constrained by levees or incision downstream of a
meander can raise water elevations through dam.
super-elevation around the bend. Using the data
Logjams can create major blockages that are very
on channel curvature we can see how water
effective at increasing the frequency of overbank
elevations at the logjam meanders can be 0.3–1
inundation. A large logjam in the Deschutes River
meter (1–3 feet) higher than the unobstructed
near Olympia, Washington, raised low-flow water
meanders (Figure 4-16B). As discussed earlier,
elevations over 1.2 meters (3.9 feet) (Figure
wood can be a primary driver in bifurcating or
4-21A). During high flows the relative effect of the
splitting flow and creating anabranching rivers
logjam diminished because flow was already out
that may otherwise be braided or single-thread
on the floodplain (Figure 4-21A). The logjam also
meandering channels. Historic channel clearing
showed a temporal hysteresis with respect to
transformed many complex anabranching rivers

Large Wood National Manual July 2015


4-26
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

water levels. During low flows the wood settles time the hydrograph begins to wane, much of the
into the channel and creates a denser or lower wood has become buoyant, and the permeability
porosity obstruction, so that during the rising of the obstruction has increased; thus, it has a
limb of a hydrograph the logjam has a greater diminished effect on water elevations during the
effect on water elevations (Figure 4-21B). By the receding limb of the hydrograph (Figure 4-21B).
Figure 4-14. (A) Threshold of Effective Wood Loading Based on Pool Frequency as a Function of Wood
Loading per Square Meter of Channel Bed; (B) Size of Functional Wood in Queets River Basin

(A) defined by Buffington and Montgomery (1999b, Figure 2, page 3511); (B) source = Abbe (2000) and Abbe
and Montgomery (2003).

Large Wood National Manual July 2015


4-27
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-15. Conceptual Illustration of How Wood Introduces Physical Complexity to a Simplified Channel

This complexity creates the greatest ecological diversity and resilience, and supports a much more productive
food web (e.g., Power et al. 1995; Power and Dietrich 2002).

Large Wood National Manual July 2015


4-28
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-16. (A) Role of Natural Logjams in Reducing the Radius of Curvature of Channel Meanders in the
Queets River, Washington; (B) Based on Assumptions for Channel Sizing Relative to Drainage Area, the
Super Elevation Associated with Smaller Radii of Curvature Results in an Increase in Water Elevations of
0.35–1.0 meters (1.1–3.3 feet), Demonstrating Another Way Logjams Increase Floodplain Connectivity and
Drive Side Channel Formation

Source: (A) Abbe (2000) and Abbe and Montgomery (2003; (B) Abbe (2000).

Large Wood National Manual July 2015


4-29
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-17. Wood Forces Channel Complexity Such as Anabranching (a); the Removal of Wood Can
Transform These Multi-Thread Systems Into a Wide Single-Thread Channel (b); Observations of the Upper
Cowlitz River in Washington Show the Loss of Vegetated Island Coincident With Increasing Channel Width
(c)

Source: Abbe et al. (1997)

Large Wood National Manual July 2015


4-30
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-18. Geomorphic Changes in Lower Elwha River, Washington, Associated with ELJ Placement

For years prior to dam removal. (A) Pool depths at seven ELJs increased. (B) Median grain size (D50)
diminished from 90 to 19 millimeters (3.5 to 0.8 inches), a 79% decrease due to stress partitioning of the
wood (data from Mike McHenry, Elwha Tribe).

Large Wood National Manual July 2015


4-31
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-19. Predicting Channel Planform Morphology Based on Formative Discharge (Q*), Median Grain
Size (D50), and Channel Slope

The stress partitioning imposed by Elwha River ELJs effectively pushes river from single thread to
anabranching (Eaton et al. 2010).

Figure 4-20. Illustration From White River in Western Washington Showing the Difference in Cumulative
Bank Length (2x channel length) for Unconfined Anabranching Reach With Numerous Logjams Versus
Confined Reaches

For identical flows in the same river, the wood-dominated anabranching reach has 2 to 5 times the amount of
habitat as measured by channel length.

Large Wood National Manual July 2015


4-32
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-21. (A) Hydrograph Showing the Influence of a Large Channel Spanning Logjam in the Deschutes
River, South of Olympia, Washington1; (B) Hysteresis Curve Showing How the Logjam Has the Most
Significant Effect on Head (Dz) During Rising Limb of Hydrograph2

1 As discharge increases (lower curve) the relative effect (head differential) of the logjam dimensions. This is
because the logjam obstructs most of the bankfull cross-section and as flow increases it spreads out across
the floodplain.
2 This is because logjam permeability increases with rising flood and therefore there is higher conveyance on

the descending limb of the hydrograph.


Source: Data provided by Thurston County Public Works pressure transducer stage gages installed
downstream and upstream of logjam.

Large Wood National Manual July 2015


4-33
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

The hydraulic effect of obstructions was modeled 4.2.5 Wood and Channel
to simulate these effects (Figure 4-22A)
(Brummer et al. 2006). These large logjams can Incision
occur in surprising places without adverse Starting in headwater channels, wood can play a
impacts, such as a channel-spanning logjam in the fundamental role in dissipating energy, capturing
Upper Yakima River right off Interstate 90 in sediment, and limiting down-cutting or incision.
Washington State (Figure 4-22B). In smaller channels only a small portion of a log
may be inside the wetted channel, but it can still
Figure 4-22. (A) Dimensionless Plot of How Wood
be effective (Figure 4-23).
Obstructing 80% of the Ozette River, Washington,
Increases Water Elevations Using a 1D Hydraulic Figure 4-23. Wood in Steep (S=0.18) Headwater
Model1; (B) Channel Spanning Logjam on Upper Channel of Olympic Peninsula, Washington
Yakima River, West of Easton, Kittitas County,
Washington2

1 As discharge, Q, increases relative to the bankfull


discharge, Qbf, there is a substantial increase in
water elevation, Dz, relative to bankfull depth, Dbf.
As flows overtop the banks, the relative effect
dimensions. (Adapted from Brummer et al. 2006) Logs create sediment traps and surface roughness
2 Logjam raises water elevations to feed extensive
that dissipates energy of floods and debris flows.
side channel network. Logjam is located just off Logs are buried in alluvium and above the
Interstate-90 and never caused flood problems due
to intact floodplain it retains. Circles highlight
channel. As sediment accumulates, logs
logjams; flow is left to right. (1998 photo) previously located above the channel can be
incorporated into the stream bed. Without wood
this would be an actively incising bedrock
channel.

As wood traps bed material and aggrades the


channel, wood that was previously suspended
above the channel can become engaged with flow

Large Wood National Manual July 2015


4-34
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

and continue the process of channel aggradation. of the wood was associated with sediment
With sediments and wood creating a mantle storage. Through cosmogenic dating of sediment
above the underlying bedrock or glacial deposits, in one of the same Maine rivers (the Ducktrap),
wood effectively retards incision and stabilizes Fisher et al. (2010) found that wood
the landscape. Removal of wood can dramatically accumulations increased the residence time of
increase the rate of incision, and in a few decades sediment stored in the channel.
the channel can cut down what would have
otherwise taken thousands of years (e.g., Veatch
1906; Guardia 1933; Wadsworth 1966; Brooks GUIDANCE
and Brierly 2002; Stock et al. 2005). Steep
headwater channels can be subject to extreme
Potential Impacts of Channel Incision
events such as debris flows or rock falls. In these
systems, logs tend to easily span the channel  Transforms alluvial beds to bedrock.
width, and the diameter of the tree is an
 Disconnects the stream from its floodplain.
important factor; for the log to function it must
withstand the forces the stream imposes. Because  Destabilizes its banks and adjacent hillslopes.
forest management and harvest can directly
 Negatively impacts water quality.
influence the size of riparian trees growing along
headwater channels, policy can have significant  Increases downstream flood peaks.
geomorphic consequences. Modeling a channel
 Delivers large quantities of sediment to
spanning log as a cylindrical beam shows that
downstream reaches (which are often in
diameter plays a critical role in the forces it can
developed areas).
withstand without breaking (Figure 4-24) (Abbe
2000). This type of analysis demonstrates that  Compromises the integrity of bridge abutments,
trees can grow to sizes that are capable of pipelines, and road embankments.
withstanding extreme forces. By doing so, they  Increases the shear stresses acting on the bed due
can effectively diffuse debris flows near their to flow confinement and lack of wood.
initiation points, minimizing their inertia and
distance traveled. This can reduce or limit
downstream consequences to habitat and human The consequences of wood removal in relatively
communities. small headwater streams can be seen in many
urban stream corridors. Where mature riparian
Baker (1979) showed how logjam removal forests and wood was left intact it can provide
resulted in short-term increases in sediment resilience to major increases in peak flows that
supply and the transformation of channel occur as a result of urbanization (Figure 4-26A).
substrate from alluvium to bedrock (Figure 25). Similar streams in the same region where wood
By storing alluvial sediments, wood creates a was removed have experienced incision of 6–18
protective barrier that slows the process of meters (20–60 feet) (Figure 4-26B).
channel downcutting. Stock et al. (2005)
document how wood removal in the Teanaway There are several well-established mechanisms
River in central Washington not only led to loss of initiating channel incision, such as a reduction in
alluvial channels, but approximately 2 meters sediment supply (e.g., downstream effect of
(7 feet) of bedrock incision in 100 years. Cordova dams), an increase in peak flows (e.g.,
et al. (2007) found that 50% of the wood found in urbanization or climate change), or
low-gradient streams of the upper Midwest were channelization (e.g., straightening and
responsible for sediment storage. In a similar confinement of flood flows by levees). The role of
assessment of low-gradient coastal rivers in wood removal as a trigger of incision has been
Maine, Magilligan et al. (2008) found that 5–20% recognized but under-appreciated, even in

Large Wood National Manual July 2015


4-35
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

restoration. Recent geomorphic analysis of direct threat to grade control structures that are
several Pacific Northwest rivers has not sufficiently keyed into the banks. Bank
demonstrated that larger rivers that have not erosion that cuts around a constructed grade
been dammed, that experienced a significant control structure can re-initiate incision. There
increase in peak flows, or that have been should be enough roughness built into the stream
channelized, are incising—despite experiencing valley to prevent incision from getting around the
an increase in sediment supply as a result of structure. Because of its size, wood is an ideal
industrial logging. The most significant material for creating complex grade control
disturbance in these systems has been the loss of structures that extend beyond the channel to
wood. The result of wood removal in the South reinforce banks and floodplain areas that may be
Fork Nooksack River has been incision that has subject to erosion (Figure 4-31). However, single
left areas occupied by the river just decades ago log weirs should be avoided; they are subject to
well above the 100-year flood elevations today undercutting and have no redundancy should the
(Figure 4-27) (Abbe et al. 2013). Bank log fail. The more logs used, the stronger the
stratigraphy can provide direct evidence of structure and greater the factor of safety.
incision by revealing old alluvial channels once Whether using a step-pool or reinforced riffle
occupied by the river sitting on top of underlying design, it is important to minimize the magnitude
geologic material (Figure 4-28). Given the size of individual drops and thus create broad crested
trees once attained, it should not be surprising structures (Figure 4-32). This typically increases
that they were capable of trapping bed material the cost, but greatly increases the structure’s
and aggrading the channels of relatively large stability and enhances fish passage. In steep
rivers. A single native old-growth tree can create step-pool or cascade channels this may entail
a 3-meter-high, 30-meter-wide (10-foot-high, placing wood through the length of the stream.
98-foot-wide) impoundment across a river
In montane rivers natural logjams can create
(Figure 4-29).
steps several meters high and have a dramatic
A geomorphic assessment should clearly describe effect on floodplain morphology by creating
and quantify the processes and rates of landscape terrace surfaces with slopes several times lower
evolution and predict what a project will be than the valley grade (Figure 4-33, A and B)
subjected to and how it will influence the (Montgomery and Abbe 2006). Observations from
evolution of the site (e.g., Schumm et al. 1984; the Queets River in Washington showed how
Schumm 1999; Simon 1989, 1994; Doyle and logjams aggraded channels to elevations higher
Shields 2000; Wallerstein and Thorne 2004; than surfaces that had previously been well above
Brummer et al. 2006; Simon and Rinaldi 2006). flood stage. The lower slopes between wood steps
Channel incision begins a long-term channel reflect how the wood is partitioning shear stress
evolution process (Figure 4-30) that can result in and storing sediment that would otherwise route
many decades before the restoration of some through the reach. In restoration sites where a
form of equilibrium. Therefore, designers should large portion of the valley can be restored,
be well aware of what stage (I–VI) of channel constructing channel-spanning structures should
evolution their system is in. For restoration sites certainly be considered. Even in a highly
in the early stages of downcutting (stages II–III), constrained urban setting it can be possible to
it may be possible to quickly reverse the process include wood where there is sufficient freeboard.
to re-establish the undisturbed condition (stage
I). Channel widening (stages IV–V) can pose a

Large Wood National Manual July 2015


4-36
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-24. Log Strength Can Be Critical in Headwater Channels Where They Are Subjected to Severe
Forces Imposed by Debris Flows

Plot illustrates an example of how large logs need to be to overcome the impact of a large boulder moving at
9 meters (30 feet) per second, as a function of log length (assuming it spans the channel). Plot illustrates that
logs 0.5 meter (20 inches) in diameter can withstand this impact for a 10-meter (33-foot) wide channel (Abbe
2000).

Large Wood National Manual July 2015


4-37
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-25. Wood Stores Sediment thus Reducing Sediment Transport Capacity by Obstructing Flow and
Increasing Roughness, Thereby Increasing Sediment Storage Within a Channel1

1 This process can transform a channel from bedrock to alluvium, which not only slows down long-term
incision rates, but increases ecological productivity. Logjam sediment storage in Hehe Creek, Oregon
Cascades. The logjam stored 1,100 cubic meters of alluvium. Within a year after removal of the logjam, 97% of
the sediment had been eroded and the channel reverted to bedrock. The log seen laying on the stream bed in
1977 was suspended 2 meters above the channel in 1978. (Adapted from Baker 1979)

Large Wood National Manual July 2015


4-38
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-26. (A) Wood in Taylor Creek (Seattle) Is Trapping Sediment and Dissipating Flood Energy1; (B) Coal
Creek in Nearby Bellevue also Experienced Increased Peak Flows due to Urbanization but Was also
Historically Cleared and Lacks Mature Riparian Conditions and Is Undergoing Incision2

1 Taylor Creek lies entirely within the city and has experienced a dramatic increase in peak flows due to
urbanization. The 100-year flood flow prior to development now occurs annually. Segments of the creek with
mature riparian forests and instream wood have demonstrated resilience to the increased flows unlike
segments without trees and wood.
2 Incision of 3–12 meters (10–40 feet) is common in creeks of the Puget Sound region and requires costly

engineering solutions to protect pipelines, bridges, road embankments, and homes.

Large Wood National Manual July 2015


4-39
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-27. Historic Channel Incision in the South Fork Nooksack River, Washington

The river has no dams and this reach has not been channelized; the only disturbances have been channel
clearing and clearcut logging of valley bottom and hillslopes. Incision was determined by mapping channel
planform (A) and determining elevations of abandoned channel beds (B). Hydraulic modeling shows that
channels occupied as recently as the 1980s are no longer inundated in a 100-year flood.

Large Wood National Manual July 2015


4-40
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-28. Eroding Bank Along the Hoh River, Washington, Showing a Snag Pointing in Flow Direction of a
Relic Channel With its Invert Perched Over 2.4 Meters (8 Feet) Above the Current River Bed

The old alluvial floodplain surface is now a terrace due to historic incision of the river.

Figure 4-29. A Single 2.5-Meter (8.2-Foot) Diameter Old Growth Douglas Fir (Pseudotsuga menziesii)
Impounding the Carbon River in Mt. Rainier National Park, Washington

Large Wood National Manual July 2015


4-41
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-30. Conceptual Channel Evolution Model of Stream Experiencing Incision due to Channelization

Removing wood from a channel increases the effective shear stress available for sediment transport and
erosion. A loss of instream wood can trigger long-term incision that is difficult to reverse (from Doyle and
Shields 2000, adapted from Simon 1994).

Large Wood National Manual July 2015


4-42
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-31. Channel Incision Poses a Serious Threat to Infrastructure Such as Pipelines, Bridge Abutments
and Piers, Water Intakes, and Road Grades1

1A geomorphic assessment is essential in evaluating channel incision, determining causal mechanisms, and
predicting the consequences. Complex wood assemblages offer a natural means of controlling incision that
can also improve fish passage.
The standard of practice is to bury pipeline only 1.5 meters (5 feet) below the streambed), bridge piers, and
road embankments (A). Emulating natural wood accumulations, engineered placements can create complex
grade control that can reverse channel incision to protect infrastructure while also improving fish passage
and floodplain connectivity (B).

Large Wood National Manual July 2015


4-43
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-32. Geomorphologists Offer Direct Design Input on the Role of Wood and Bed Material on Channel
Morphology that Is Essential in Stream Restoration and Providing Sustainable Solutions for Protecting
Infrastructure1

Here (Woodward Creek in southwestern Washington), wood and large rock was used to create a cascading
channel step to treat incision threatening a gas pipeline and restore fish passage,
The pipeline was exposed in 2007 after the channel incised approximately 2.4 meters (8 feet) (A). To
safeguard the pipeline and improve fish passage, a log and rock grade control riffle was designed (B) and
constructed in 2008 (C). The bottom photo shows the structure in 2009 after it had been subjected to a
25-year flood (D). This type of approach shows how wood and stream restoration can also benefit
infrastructure (Abbe et al. 2009).

Large Wood National Manual July 2015


4-44
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-33. Natural Logjam Influence on Channel Aggradation and Terrace Construction in 4th Order Alta
Creek (A) and 6th Order Mainstem Queets River (B)

Source: Abbe (2000), Abbe and Montgomery (2003), and Montgomery and Abbe (2006).

Large Wood National Manual July 2015


4-45
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

4.2.6 Wood and Bank Erosion Small trees are easily flushed downstream by the
river and do not retard erosion (Figure 4-35B).
Trees further slow the water and their roots hold The role of large trees is evident in the forest
the underlying soil together (e.g., Tsukamoto structure of undisturbed rivers where channel
1987; Sidle 1991) and increase the strength of migration erodes valley hill slopes. At such sites,
river banks to resist erosion (Eaton et al. 2004; logjams form at the toe of the hill slope that
Eaton 2006; Simon and Collison 2002; Simon et redirects the river and halts further erosion
al. 2000; Konsoer 2014). The cohesion provided (Figure 4-36). Where large trees were removed,
by riparian vegetation directly influences bank erosion tends to proceed, even into hill slopes
strength and hydraulic geometry, effectively that rise far above the river (Figure 4-37). An
reducing channel width (Eaton et al. 2004; Eaton analysis of the Hoh River found almost four times
2006). When banks erode, fallen trees can form as much land was eroded by channel migration
logjams that further diffuse the river’s energy and outside Olympic National Park in logged lands
even protect some areas of floodplain from versus unlogged areas (Figure 4-38).
erosion (Abbe et al. 2003a; Konsoer 2014).
Bank erosion rates along forested banks can be
Functional wood can play an influential role in the 50 to 90% lower than along unforested banks
rate of bank erosion along rivers and (Thorne and Furbish 1995; Micheli et al. 2003;
demonstrates how both the restoration of mature Abbe et al. 2003a; Konsoer 2014). Abbe et al.
riparian forest buffers and engineered wood (2003a) found erosion rates were dependent on
placements can be used to protect banks and tree size, which was attributed to larger trees
enhance habitat. A geomorphic analysis of being more likely than smaller trees to form
channel migration along the Hoh and Queets stable roughness elements with a longer
rivers of northwestern Washington found that residence time along an eroding bank. Konsoer
erosion rates were lower in areas with larger (2014) found that tree snags along a bank were
trees (Figure 4-34A) (Abbe et al. 2003a). The the primary roughness element and responsible
analysis showed a statistically significant for major changes in flow patterns along eroding
difference between areas with trees less than banks. Flow patterns and erosion rates were
53 centimeters (21 inches) in diameter versus compared to two similar meander bends of the
those with greater diameters (Figure 4-34B). The Wabash River in Illinois, one with a smooth bank
median normalized erosion rate for the areas along agricultural land, one along forested land
with larger trees was 5 meters (16 feet) per year, (Figure 4-39). Rougher banks have much more
and the rate for areas with small trees was pronounced secondary flow vortices that slow
11 meters (36 feet) per year (Abbe et al. 2003a). near-bank velocities and push the primary
These differences are consistent with work by current farther from the bank (Thorne and
Micheli et al. (2003) who found erosion rates Furbish 1995; Meile et al. 2011; Konsoer 2014).
along meanders of the Sacramento River in The smoother bank eroded 17 times faster than
California were twice as high in agricultural areas the rough bank (Konsoer 2014). Increasing bank
as they were along riparian forests (forest erosion roughness increases the width of slower near-
rates ranged from 2.5–6 meters [8–20 feet] per bank velocities, reducing shear stresses acting on
year versus agricultural rates of 6–11 meters the bank and creating more refuge and cover for
[20–36 feet] per year. The differences in rates fish. Roughened bank treatments can offer greater
found along the Hoh and Queets rivers (Abbe et erosion protection and fish benefits than
al. 2003a) are due to the role of key pieces of traditional methods (Figure 4-40). Complex
wood falling into the river as the bank erodes. placements of wood that increase river bank
Stable snags create roughness along the bank that roughness and lower shear stress can be an
partitions shear stress and deflects flow away effective means of bank protection (e.g., Abbe and
from the bank (Figure 4-35A). Brooks 2011; Abbe et al. 2011).

Large Wood National Manual July 2015


4-46
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-34. (A) Forest Areas With Larger Trees Erode More Slowly Than Areas With Smaller Trees Along
the Hoh and Queets Rivers; (B) Breaking Data Into Two Categories greater and less than 53 Centimeters (21
Inches), There Is a Statistically Significant Difference, With Areas With Larger Trees Eroding at less than Half
the Rate of Smaller Trees

Source: Abbe et al. (2003a).

Large Wood National Manual July 2015


4-47
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-35. (A) Erosion Into Mature Forests Along the Hoh River Recruits Large Snags That Form Stable
Obstructions (Key Pieces) in the Channel That Slow Erosion Rates; (B) Areas of Industrial Forest or
Agriculture (Trees Less Than 21 Inches) Erode at Over Twice the Rate

Source: Abbe et al. (2003a).

Large Wood National Manual July 2015


4-48
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-36. Illustration of How Large Wood Influenced Channel Process and Morphology on the South Fork
Hoh River, Washington, from 1993 to 2013

From 1990 to 2006 the river migrated north about 32 meters (106 feet) into mature timber, a rate of about
2.1 meters per year (7 feet per year). This recruited trees that obstructed the channel and halted further
erosion. The logjam moved the river south, stabilized the toe of the embankment, and established new
floodplain forest.

Large Wood National Manual July 2015


4-49
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-37. Clearing Mature Riparian Forests Eliminates Functional Wood Recruitment and Alters
Processes and Channel Form

This site in the South Fork Hoh River, Washington (just downstream of the site shown in Figure 4-36), the
river migrated about 47 meters (153 feet) to the northwest from 1990 to 2006 into the adjacent valley
hillslope that had been clear cut. The average rate was 3 meters per year (10 feet per year). Because the
erosion destabilized the hillslope, erosion impacted a much greater area, extending 167 meters (550 feet)
into the valley margin. From 1990 to 2013 the rate of head scarp retreat was about 7.3 meters per year (24
feet per year).

Large Wood National Manual July 2015


4-50
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-38. Outside Olympic National Park Almost All Old Growth Forest Within the River Valley and
Adjacent Hillslopes Has Been Cut

The area outside the park has experienced much more erosion and expansion of historic channel migration
zone than the areas within the park (U.S. Bureau of Reclamation 2005).

Figure 4-39. Flow Velocity Fields Around Two Bends of the Lower Wabash River, Illinois

Cross-section MB 150 is downstream of meander apex with relatively smooth bank, HSB72 is cross-section in
similar location of bend where there are snags along the bank. The roughness created by the snags along
HSB72 slows down velocities near the bank. Erosion rates at HSB72 are 17 times less than MB 150 (Konsoer
2014)

Large Wood National Manual July 2015


4-51
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-40. Illustration of How Rougher Banks Reduce River Velocities Near the Bank

Traditional bank protection tends to create a smooth bank where high flow velocities hug the bank (solid
lines above). Where banks are roughened, near-bank velocities are reduced (dashed lines), diminishing the
risk of erosion and improving salmon habitat.

Large Wood National Manual July 2015


4-52
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

4.2.7 Sediment Storage channel-spanning wood accumulation


(historically referred to as “rafts”) in the lower
Sediment storage behind large wood Colorado River of Texas was by Spanish explorers
accumulations was clearly illustrated in the in 1690 (Clay 1949). Prior to removal of this
patterns of channel erosion that followed wood wood by the U.S. Army in 1927, the shoreline at
removal (e.g., Veatch 1906; Guardia 1933; Baker the river’s confluence into Matagorda Bay
1979; Sedell and Luchessa 1981; Hickin 1984; exhibited no protruding delta. Directly after
Sedell and Frogatt 1984; Hartopo 1991; Barrett removing the jam, a pronounced delta began to
1996; Abbe 2000; Wallerstein and Thorne 2004; extend into and eventually across the bay
Stock et al. 2005; Phillips 2012). Examples of (Wadsworth 1966; Hartopo 1991).
channel incision in large low-gradient rivers Approximately 11 by 81 cubic meters (14 by
include the Cann River in Victoria, Australia 106 cubic yards) of sediment was introduced to
(Brooks and Brierly 2004), the Red River of Matagorda Bay (Gulf of Mexico) over a 29-year
Louisiana (Veatch 1906; Guardia 1933; Triska period after raft removal (Abbe 2000).
1984), and the Colorado River near the town of
Matagorda, Texas (Wadsworth 1966; Kanes 1970; These are examples from relatively large rivers,
Hartopo 1991). but the same issues impact streams of all sizes
throughout a watershed. In the Puget Sound small
The Cann River is a sand-bedded river draining streams that were relatively stable for several
about 650 square kilometers (250 square miles). thousand years have been subject to dramatic
After wood removal, the Cann River experienced incision in the twentieth century following
an 860-fold increase in sediment transport and historic logging that removed riparian trees and
evacuated a volume of sediment representing instream wood (e.g., Baker 1979; Stock et al.
1,500 years of floodplain sedimentation (Brooks 2005), and upland development that increased
and Brierly 2004). These authors estimated that peak flows (e.g., Booth 1990). Incision sent head
to aggrade the river to its pre-existing state would cuts and gullies up these drainages (such as
take 31,000 years, assuming restoration of Figure 4-26B) and large quantities of sediment
instream wood and riparian vegetation. downstream that often required construction of
expensive sediment retention ponds throughout
The Red River drains 110,190 square kilometers
the region. Restoration of wood can provide a
(43,545 square miles) of Arkansas and Louisiana.
much more economical, environmentally
Logjams were documented in the earliest records
beneficial, and sustainable solution to limit
of European exploration, and Native Americans
channel incision, reduce erosive power (sediment
could not recall a time when jams did not block
transport capacity), and trap sediment (e.g.,
the Red River (Lowrey 1968; Triska 1984; Barrett
Lester and Wright 2009; Abbe et al. 2009; Abbe
1996). After large logjams were removed from
and Brooks 2011; Daley 2012).
1873–1892, the river experienced 5 meters (16
feet) of incision (Veatch 1906). The river
transformed from a network of narrow 4.2.8 Water Quality
anabranching channels to a larger single thread The ability of large wood to trap sediment is also
channel. Barrett (1996) estimated floodplain applicable to other suspended materials in the
sedimentation rates were reduced almost 10-fold channel. Ehrman and Lamberti (1992) compared
from 2.0–3.0 to 0.3–0.4 centimeters per year channels with and without large wood, and found
(0.8–1.2 to 0.1–0.2 inches per year) after logjam that channels with large wood retained water
removal in the Red River. 1.5–1.7 times longer and reduced coarse
The Colorado River of Texas has a drainage area particulate organic matter (CPOM) transport by
of 110,190 square kilometers (43,545 square 35% when compared to channels without large
miles). The first recorded observation of a large wood. Similarly, Jacobson et al. (1999) found that

Large Wood National Manual July 2015


4-53
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

large wood trapped CPOM, which was then exclusion decreases phosphorus retention and
incorporated into the benthic biomass, creating that large wood exclusion further impedes a
islands of organic matter in the channel that stream’s natural ability to absorb nutrients. Using
became focal points for decomposition and solute injection techniques Valett et al. (2002)
secondary production. Because decomposition found that phosphorus uptake in channels with
(Sinsabaugh et al. 1994) and invertebrate grazing high large wood volumes, frequent debris dams,
(Lampert 1978) of CPOM releases dissolved and fine-grained sediments was significantly
organic carbon (DOC) into the water, it would greater than in channels in younger forests
follow that CPOM retention would create higher without these characteristics. Finally,
DOC concentrations in the stream. corroborating this finding, Ensign and Doyle
(2005) conducted phosphorus injections in
streams both before and after the removal of
CROSS-REFERENCE large wood and CPOM in the channels, and found
that phosphate uptake decreased by up to 88%
Wood influences on hydraulics, water levels, substrate,
after large wood removal. These studies show
channel morphology, and hyporheic flow has direct
implications for water quality, a topic that is also
that large wood increases water retention and
addressed in Chapter 3, Ecological and Biological provides a substrate for biofilm growth; both
Considerations. these factors contribute to higher phosphorus
retention in streams that have large volumes of
large wood.
A litter exclusion study by Meyer et al. (1998) at
Coweeta, North Carolina, showed that DOC Given these factors, it would seem that the
contribution from the in-channel leaf pack presence of large wood acts to reduce phosphorus
contributed 30% of the total export of the stream; and sediment export while having little effect on
the remainder of the DOC was imported into the DOC concentrations other than at sites where
channel from the landscape. Once in the channel, wood increases turbulence and plunging flow.
however, DOC is labile and will likely be taken up The removal of upstream snags, especially those
quickly in a reach with a high volume of large associated with major channel formations (pools,
wood. It has been shown that certain labile large flow divergences), would exacerbate any
dissolved organic compounds are quickly taken potential water quality problems. Of course there
up in channels that have flow obstructions (Hall would be other ramifications if the reach was de-
and Meyer 1998; Wiegner et al. 2005). Therefore, snagged, including the degradation of fish habitat
it would seem that the presence of large wood (Lehane et al. 2002; Mossop and Bradford 2004),
and associated CPOM would not significantly reduced carcass retention (Johnston et al. 2004;
impact stream DOC concentrations because there Minakawa and Gara 2005), lower macro-
are the counteracting dynamics of DOC export invertebrate populations (Johnson et al. 2003),
from CPOM decomposition and DOC uptake from and increased risk to downstream infrastructure
large wood–induced water retention. because logjams trap mobile debris. From a water
quality perspective the presence of large wood is
Water retention from large wood and the clearly beneficial.
presence of CPOM in the channel also play
important roles in the fate of nutrients in the
stream channel. In a classic paper by Mulholland 4.3 Hydrology
et al. (1985) it was suggested that leaf litter in
streams promotes nutrient retention as the leaf Hydrology is an earth science discipline focused
pack acts as a substrate for nutrient-hungry on the properties, origin, circulation, and
microbes. Additionally, a more recent study by distribution of water in the environment,
Webster et al. (2000) suggests that litter including fluxes in streamflow, interflow, and

Large Wood National Manual July 2015


4-54
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

groundwater discharge. Understanding the Large wood and other riparian vegetation (flow
timing, rate, and mechanism of movement of obstructions) create a hydraulic “backwater
water through watersheds and its role in effect” whereby the water level immediately
geomorphic processes is important for large upstream of the obstruction is raised, which in
woody material design as it affects erosion, turn raises the level of water upstream of it, and
sedimentation, riparian plant growth, and other so forth, resulting in a curve of slower and higher
key processes. Furthermore, knowledge of how water extending upstream from the obstruction.
hydrological processes, namely streamflow Backwater curves indicate water storage created
hydrographs and flood wave dynamics discussed by the obstruction. Backwater effects typically
herein, are affected by riparian vegetation and extend farther upstream in lower slope channels.
large wood in the channel is important to When the velocity upstream of the obstruction is
understanding the tradeoffs between enhanced slowed, then it is not routed downstream as
ecological benefit and altered levels of flood quickly as it otherwise would be, and the water
protection. already downstream will drain away as the water
level drops (Rutherford et al. 2007). Therefore, an
4.3.1 Effects of Riparian obstruction in the channel has the effect of
altering hydrology by slowing a flood wave by
Vegetation and Wood increasing storage, depth, and duration of the
on Hydrology wave upstream with a decrease in flood depth
downstream. As Figure 4-41 illustrates, the effect
Floods are often described as traveling as a of individual plants and logs on the flood stage
slow-moving wave that increases in size as it largely depends on biomass, plant flexibility, the
travels down the watershed from the addition of level of streamlining and lying down of
many smaller waves derived from the upstream stems/leaves under flow pressure, and whether
network. Quick traveling small waves typically the plant is submerged or emergent. Typically,
create a flood wave with a higher peak discharge and especially for grasses, willows, and other
but shorter duration than slower waves flexible riparian plants, flow resistance decreases
(Anderson 2005, 2006). with increasing discharge and stage. However,
Floodplains, riparian vegetation, and large wood flow resistance can increase in situations where
in the channel play a key role in flood wave elevated flood stage results in increased flow
dynamics by storing flood water, at least through the tree canopy. At the cross-section
temporarily, and slowing the pace of the flood scale the presence of riparian vegetation results
wave as it moves down the watershed. Ultimately, in increased stage levels compared to a condition
the size of the flood wave’s peak discharge at a without riparian vegetation (Figure 4-41).
particular location in the watershed is related to Therefore, placement of wood in the channel and
how quickly smaller waves from tributaries join increased riparian vegetation can result in higher
together and the volume of water stored during water surface elevations along the banks and, in
the flood (Anderson 2005, 2006). The effect of unconstrained reaches, enhanced floodplain
riparian vegetation and large wood on flood connectivity from an increased volume of water
hydrology depends largely on the scale spilling out onto the floodplain.
considered, network geometry, channel
morphology, and the flood magnitude.

Large Wood National Manual July 2015


4-55
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-41. Conceptual Diagram of the Effect of Riparian Vegetation on Discharge at the Scale of a Plant, a
Cross-Section, a Reach, and a Catchment

Source: Anderson (2006).

Large Wood National Manual July 2015


4-56
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

cross-section and into the reach scale versus a


GUIDANCE
delaying of the flood wave and reduction in its
peak discharge downstream at the watershed
Riparian vegetation and large wood in a channel
scale. When considered at the watershed scale,
network directly influence flood hydrographs
(Anderson 2006; Thomas and Nisbet 2006; Rutherford the presence of riparian vegetation and large
et al. 2007). wood actually enhances flood protection for
 Increase roughness that dissipates flow energy downstream locations compared to a situation
(e.g., stress partitioning). devoid of riparian vegetation and large wood
 Reduce channel and floodplain conveyance (Anderson 2005).
 Diminish flow velocity and the flood peak celerity ,
 increase stage and flood storage which can Research has also shown the influence of channel
contribute to attenuating flood waves moving shape on the ability of riparian vegetation and
through the watershed, lowering flood stages large wood to attenuate flood peaks (see
downstream. Figure 4-42). In a scenario with a 50-kilometer
(31-mile) reach, riparian vegetation and large
wood delays the flood wave peak discharge
When considered at the reach scale, defined as
between 5 and 10 hours depending on cross-
approximately 10 kilometers (6 miles) in
section shape, with wider and shallower cross-
Anderson’s (2006) research, the combined
sections with substantial floodplain roughness
backwater effects of increased wood and riparian
providing the greatest attenuation and narrower
vegetation result in increased water storage and a
and deep cross-sections the least (Anderson
slower flood wave celerity (velocity of the wave
2006). If the width/depth ratio of the stream is
traveling through the reach) and increased time
greater than 17, vegetation is not likely to
to peak discharge compared to the no vegetation
appreciably affect flooding because the cross-
condition (Figure 4-41). The time rate of flood
section is too wide and shallow (Masterman and
wave diffusion is greater in the no vegetation
Thorne 1992). Also, more flood attenuation is
condition compared to the densely vegetated
observed in the case of the small input discharge
channel. Yet, the peak discharge of the flood wave
compared to the large input discharge (Figure
at the end of the reach with vegetation is less than
4-42).
the peak of the wave in the no vegetation reach;
thus, the rate of diffusion per channel length with The following discussion focuses specifically on
vegetation is equal to or greater than the reach large wood in the channel without considering the
without vegetation (Anderson 2006). This result role of riparian vegetation. The backwater effects
is explained by differences in flood wave celerity. and flood storage created by multiple wood
In Anderson’s (2006) research, the hydrograph in structures are additive and can result in
reaches void of vegetation generally moves three appreciable flood wave attenuation accompanied
times faster than with vegetation; thus, diffusion by local increases in flood height. Though not
can occur only a third of the time, and at the end uniform to all systems, research has shown that
of the reach the no-vegetation condition has more large wood in the channel has a small to
discharge than the vegetated condition. insignificant effect on the duration or frequency
of large flood events (approximately events
At the catchment scale, the riparian vegetation
greater than the 20-year flood) because much of
and large wood have the effect of pushing water
the flood water is out on the floodplain, but can
out onto floodplains, and delaying and reducing
increase the duration of smaller floods (i.e., 1- to
the magnitude of the flood peak but increasing
2-year events) where most of the flow is still
the duration of higher flows (Figure 4-41).
contained to the channel (Rutherford et al. 2007).
Therefore, a tradeoff exists whereby the presence
Large wood of a given size will have a greater
of riparian vegetation and large wood in the
effect on a small stream.
channel can lead to increased flood height at the

Large Wood National Manual July 2015


4-57
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Figure 4-42. Sample of Simulated Waves Computed for Different Channel Shapes

Figure shows the input hydrograph and hydrographs at 10 kilometers (6 miles) downstream of the input
hydrograph for channels with vegetation, and clear of vegetation. (From Anderson 2006 as cited in
Rutherford et al. 2007).

CAVEAT

Large wood generally will not affect small flood events when the following is true (Rutherford et al. 2007):

• The projected area of the large wood is less than 10% of the area of the cross-section. The projected area is
the area of the large wood in a two-dimensional cross-section across the stream. A large wood structure needs
to be very large to occupy 10% of the cross-section of a third-order or higher stream.

• The large wood is angled at 40° to the flow (i.e., with the upstream end of the large wood against the bank).

• The large wood is submerged in a backwater at higher flows. That is, the level of the flood could be
hydraulically controlled by some feature downstream. For example, a bridge crossing downstream may
constrict the flood flow. This constriction will then produce a backwater upstream. If the large wood falls
within that backwater, then it will have no hydraulic effect on flow at all during that flood. As the flood level
falls, however, the large wood will eventually produce its own shorter backwater. The same principle applies
to a backwater produced by large wood: if additional large wood falls within that backwater, it will have no
hydraulic effect on flow. A rule of thumb for this effect is that large wood that is five to six diameters upstream
of other large wood of similar (or larger) size will not affect flood level, because it will be within the backwater
of the existing large wood.

 Several large wood structures in line will not produce any more afflux than a single large wood structure, so
long as each structure is located within two times the diameter of the next structure up or downstream.
Therefore, up to six structures can be placed parallel to each other in a line. In general, any piece of wood will
add little extra afflux (i.e., rise in water level) if it is placed within four log diameters of the next piece.

Large Wood National Manual July 2015


4-58
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

The research into the effects of large wood on


flood hydrology shows that adding wood to the
stream in many situations has a negligible effect
on the stage or duration of local flooding
(Rutherford et al. 2007). However, the local
increase in flood stage created by the addition of
large wood into the project reach may elevate
flood risk for a particular recurrence interval
beyond an acceptable level even though a net
benefit in flood protection is gained for locations
downstream. Hydraulic modeling would need to
be performed to fully evaluate the extent to which
placement of large wood elevates flood stage at
the project site. Similarly, an unsteady flow
hydraulic model could be used to also evaluate
how placement of the large wood may increase
flood wave diffusion and attenuate flooding
downstream.

Beyond the question of flood risk, current


research is showing how riparian vegetation and
the addition of large wood in the channel alters
hydrology by slowing water and enhancing
hyporheic flow and pushing it out onto
floodplains and backwater areas in unconfined
reaches where benefits of floodplain connectivity,
storage, and recharge and elevation of
groundwater levels can be attained.

Large Wood National Manual July 2015


4-59
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Geomorphic and Hydrologic Analysis Checklist


Watershed
 Drainage area
 Relief
 Annual and monthly precipitation
 Land cover
 Sediment sources
 Presence of dams or diversions influencing flow, water, and wood

Project Site
 Topography and bathymetry
 High-resolution digital elevation mapping using LiDAR and ground surveys
 Precipitation and flow data
 Valley
 Gradient
 Extent of alluvial valley bottom
 Extent of active floodplain (e.g., 100- and 500-year flood inundation areas)
 Presence of relic channels and wetlands
 Presence of terraces
 Valley perimeter and geologic composition
 Infrastructure/development
 Channel
 Gradient
 Unvegetated width
 Sinuosity or total channel length for anabranching systems
 Map presence of side channels or anabranches
 Pool location, frequency, and size
 Grain size distribution of channel substrate, differentiating surface and subsurface in gravel bedded
channels
 Bed texture mapping
 Wood
 Stable snags
 Location
 Size

Large Wood National Manual July 2015


4-60
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

 Effect (e.g., pool, bar)


 Logjams
 Location
 Size
 Formative mechanism (e.g., snag, rock, trees, bridge pier)
 Effect (e.g., pool, bar)
 Infrastructure within floodplain (e.g., bridges, roads)

Historic Change
 Historic documentation
 Historic accounts and photos
 Geo-rectified maps and aerial photos
 Previous studies
 Watershed (e.g., extent of forest clearing, development, dams)
 Hydrology (e.g., are there trends of increasing or decreasing peak and base flows?)
 Floodplain (e.g., how much of floodplain has been disconnected?)
 Original old-growth riparian forest conditions at project site
 Tree diameters and heights
 Stem densities
 Channel
 Location, date, and extent of modifications
(e.g., clearing, straightening, levees, revetments, bridges, wood clearing, splash damming, gravel mining)
 Channel patterns
 Sinuosity, unvegetated width, anabranching
 Incision or aggradation
 Channel migration
 Historic channel change mapped
 Erosion rates computed
 Is erosion linked to peak flows or floodplain conditions?
 Historic wood loading (from historic evidence or using applicable reference studies)

Problem Definition
 Summarized historic geomorphic change at site
 Simplification of channel?
 Shortening of channel length?

Large Wood National Manual July 2015


4-61
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

 Floodplain disconnection?
 Channel incision?
 How has wood supply and loading changed?
 How has flow regime changed?
 How has sediment supply and transport capacity changed?
 Extent of habitat impacts; for example:
 Total reduction in channel length?
 Reduction in number of pools?
 Loss of large riparian trees?
 Change in substrate?
 Changes to extent and rate of channel migration?
 Reduction in flood inundation frequency?
 Increase in peak flow magnitude and frequency?
 Alteration of natural flow regime?
 Increase or reduction in sediment supply?
 Historic and current threats to infrastructure
 Future change prediction under a no-action scenario
 Problem summary

Project Geomorphic Goals Defined


 Full or partial restoration
 Objectives and numerical metrics described; for example:
 Restore channel gradient
 Increase sediment storage
 Increase instream wood quantities
 Increase number of pools and cover
 Increase channel length
 Increase floodplain connectivity (inundation frequency)
 Protect property or infrastructure
 Identify area of restored habitats

Large Wood National Manual July 2015


4-62
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

4.4 Uncertainties and Research Needs


1. Regional descriptions and databases of wood loading, functions, and longevity in streams are
lacking throughout the United States, particularly with regards to undisturbed native forests.
Much of the present state of knowledge comes primarily come from the Pacific Northwest
(Oregon, Washington and Southeast Alaska), and Rocky Mountains (Idaho and Colorado Front
Range). Recent work has included the Upper Midwest and New England, but much more is
needed. Regions particularly unrepresented include the arid southwest (Nevada, Utah, New
Mexico, Arizona, and Southern California), the Sierra Nevada (eastern California), the Great
Plains (Eastern Colorado, Kansas, the Dakotas, Nebraska, Missouri), the lower Midwest (Iowa,
Illinois, Indiana, Ohio), the South (Arkansas, Oklahoma, Texas, Tennessee, Kentucky, the
Carolinas, Georgia, Alabama, Louisiana, and Florida), the Mid-Atlantic (Pennsylvania, Maryland,
West Virginia, Virginia), and the Alaskan interior.
2. Better predictive models and empirical databases are needed for predicting pre-European
settlement wood loadings for each major ecoregion, or a basis for estimating ecologically
appropriate loads for future land use and climatic conditions.
3. Little remains known regarding wood in rivers subject to annual ice cover and break-up. It
appears that processes that influence ice floes and ice jam formation also influence wood
accumulation. Large wood accumulations are found in many rivers that ice over but are rare in
some channels, possibly due to the scouring effect of ice floes.
4. More information is needed regarding the linkages between particular types of wood loading
(size, individual pieces, logjam types) to:
a. the geomorphic functions the wood provides
b. the characteristics of riparian forests
c. the disturbance processes responsible for wood recruitment to a channel
5. More data on the porosity and permeability of logjams and how these parameters influence
hydraulics in and around a logjam.
6. More quantitative data is needed linking key attributes of wood to unique channel types, and
models to predict channel response to particular wood loading scenarios.
7. More quantitative data is needed about water and sediment storage due to wood at both the
reach and watershed scales.
8. There has been substantial research recently into how beavers, through building dams of wood
debris, alter fluvial processes. The differences and linkages between the characteristics and
functions of beaver and non-beaver wood placements need to be described.
9. Key factors influencing wood longevity need to be determined and practical models for
predicting wood longevity need to be developed.
10. More experimentation is done to show how other materials such as rock or concrete “snags” can
be used to restore wood function with the longevity and stability factor of safety that is being
requested in many locations that are constrained by existing development.

Large Wood National Manual July 2015


4-63
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

11. More studies are needed on the antecedent conditions and hydrograph characteristics (e.g.,
gradual vs. rapid increase in flow) that influence wood stability and transport.
12. More data is needed on wood stability and transport in deformable channels composed of
different sediment gradations.
13. The effect of wood on long-term channel incision and landscape evolution needs further study.
14. More information is needed on the extent to which wood altered fluvial systems within the
geological record. There has been significant research into the sedimentology and stratigraphy
of fluvial systems that includes information on wood and logjams, but much of this information
has never been evaluated in the context of current stream management and restoration. This
research may offer valuable insights into not only wood longevity, but also into the long-term
influence of wood in fluvial environments.
15. Geological literature and additional research could provide information on wood and its effects
through past periods of climate change and major disturbances.
16. More physical and numerical modeling is needed on how the density of different wood
placements (from random to fixed orientations) along a channel margin influences bank erosion
rates.
17. More modeling is needed for flow separation around different logjam configurations (forming
bluff bodies), particularly with regards to turbulence, eddy formation, sediment retention, and
scour.
18. More modeling is needed to predict wood transport and deposition under different flood
hydrograph scenarios, particularly for the National Flood Insurance Program.
19. Qualitative and quantitative channel evolution models could specifically address wood inputs to
managers, stakeholders, and communities guidance on how their streams will look under
different management scenarios.
20. Guidance is needed on qualifications expected for professional geologists and others to provide
expertise in geomorphology sufficient to stamp design plans and reports.

4.5 Key Points


1. Trees have influenced watershed processes and the morphology of fluvial systems for the last
370 million years of earth history.
2. Trees grow along stream banks in almost every region of the country, including the arid
southwest, the Great Plains, the tropics of Hawaii and Puerto Rico, and the Alaska interior.
3. Wood enters streams as whole trees or fragments where it becomes part of the physical matter
or sediment comprising the stream bed. Larger or key pieces of wood can form stable
obstructions (i.e., snags) close to where they entered the stream or after becoming embedded in
the channel farther downstream. Smaller pieces move downstream, some accumulating on
obstructions to form logjams, some deposited on floodplains by high flows, and some exiting the
system.
4. Wood naturally forms distinct types of accumulations depending on the size and shape of the
wood and the channel, the channel gradient, and the channel substrate. Wood accumulations are
found in all parts of a channel network from headwater tributaries to the largest rivers. Reach-

Large Wood National Manual July 2015


4-64
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

scale wood density (quantity per channel area or length) generally tends to diminish in larger
channels, but when normalized to channel size, wood loading increases with increasing channel
size.
5. The geomorphic effects of wood include, but not limited to, the following.
a. It adds physical complexity, creating more variation in channel geometry (widths and
depths).
b. It increases pool frequency by creating hydraulic steps, flow deflectors, and flow
constrictions.
c. It partitions shear stress to reduce a stream’s energy available for sediment transport, bank
erosion, or channel incision.
d. It traps and stores sediment within a stream valley.
e. It increases channel length by:
1) Splitting flow into multiple channels, creating islands and anabranching or
anastomosing channel patterns.
2) Reducing channel radius of curvature and increasing sinuosity.
f. It increases water elevations locally and on a reach scale; thereby, it can:
1) Create and sustain ephemeral and perennial side channels and floodplain wetlands.
2) Increase the frequency of overbank inundation and water storage within the floodplain.
3) Raise groundwater tables.
4) Increase growth rates of riparian vegetation.
5) Increase hyporheic exchange, increasing upwelling and downwelling within the system
and influencing water quality and temperature.
6. Wood loading tends to be more evenly distributed throughout the length of smaller channels
(those with widths less than or similar to height of riparian trees) and more concentrated in
intermittent accumulations (i.e., logjams) in large channels (those with widths greater than tree
height).
7. Stream management and restoration should focus on wood function in the fluvial system (how it
will influence hydraulics, sediment retention, channel form and dynamics) and only use
empirical data/models of regional wood loading for the context of placements.
8. Sediment and wood budgets provide an accounting of material inputs and outputs to a project
reach that can provide insight into future evolution of the channel. Practitioners should
understand how different project design alternatives will influence sediment and wood budgets.
9. Stable instream wood can reduce the median grain size of channel substrate by reducing the
shear stress available for sediment transport. It can increase the residence time of alluvial
sediment within its valley and control channel gradient.
10. The loss of functional instream wood can trigger channel incision that may be difficult to
reverse. This process disconnects streams from their floodplains, which can severely affect the
stream’s ecology as well as increase downstream flood peaks and lower base flows.

Large Wood National Manual July 2015


4-65
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

11. Riparian forests are the primary source of instream wood and thus are an essential element for
ensuring that stream restoration is sustained over the long term. Riparian forests also help to
limit erosion, provide critical ecosystem functions and diffuse flood peaks moving down a
channel network.
12. Stable wood (natural or engineered) can be critical in protecting riparian areas from erosion
and thus allowing trees to mature so they can sustain the supply of large key pieces in the
system.
13. Wood accumulations along stream banks can be effective in reducing bank erosion rates. Even
along large rivers, erosion rates into forested banks are one-half to less than one-tenth those
observed along banks without trees or only small trees.
14. Wood longevity/preservation in streams varies widely, depending on the depositional setting.
Anaerobic conditions (such as remaining saturated) increase preservation so wood that remains
saturated can last for thousands of years or longer. Wood subject to periodic drying will be
much more susceptible to decay (i.e., fungal or insects) and be gone in several years.
15. Geomorphologists should provide input on wood longevity and whether some situations
warrant the use of other materials (e.g., rock) that can provide the same physical function and
retention as small wood in a way that better ensures long-term recovery of the system and the
required factors of safety needed for some projects. In situations where other materials are used
to provide the hydraulic function wood once provided, it is essential to show how the project
will restore the many other functions that wood provides by increasing wood retention (both
natural inputs and wood placements).
16. Fluvial geomorphology provides an understanding of the processes that shape and change a
stream and thus is essential to all stream management and restoration. Geomorphology should
provide every project with an explanation of the following.
a. What factors influence the morphology and dynamics of the stream?
1) Hydrology/flow regime
2) Geology and topography
3) Sediment (characteristics, supply, transport)
4) Vegetation
5) Wood
6) Watershed disturbance regimes (magnitude and frequency)
b. What did the stream look like prior to human disturbance?
c. Did the stream channel move around historically or in its natural state (time period
undisturbed by human development)? What is the stream corridor, including its floodplain
and channel migration zone? Is there evidence the stream channel cut down (incised) or
rose (aggraded), and how has this affected fluvial processes and morphology?
d. Was the stream ever in a state of dynamic equilibrium; if not, what state of channel
evolution is it in?
e. How much wood was naturally in-channel, and how did wood influence channel
morphology, hydraulics, and substrate?

Large Wood National Manual July 2015


4-66
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

f. What changed and how did the stream respond? How much did the loss of wood contribute
to the current state of the stream? What is the present state of the channel evolution?
g. What will happen to the stream if nothing is done?
h. How can wood be used to rehabilitate the stream given the current context of the
watershed, changes in flow regime, and sediment supply?
i. What will be the channel’s response to wood placement, and what will the benefits and risks
be?
j. How will a design respond to channel scour or aggradation? How will it perform if the
channel moves away?
k. How will watershed development and climate change influence flow regime, sediment,
riparian conditions, and instream wood? How will the project design accommodate and
possibly moderate changes?
l. How long will wood placement function as intended within the stream? How long will it take
to restore riparian forests to sustain long-term wood functions?
17. Stream management and restoration should include both stable and dynamic wood placements
to restore natural processes to channels depleted in wood.

4.6 References
Abbe, T. B. 2000. Patterns, Mechanics, and Geomorphic Effects of Wood Debris Accumulations in a
Forest River System. Ph.D. dissertation. University of Washington, Seattle, WA. 222 pp.

Abbe, T. B., and A. P. Brooks. 2011. Geomorphic, Engineering, and Ecological Considerations when
Using Wood in River Restoration. Pages 419–451 in A. Simon, S. J. Bennett, and J. M. Castro
(eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools.
Geophysical Monograph Series 194. Washington, D.C.: American Geophysical Union.

Abbe, T. B., and D. R. Montgomery. 1996. Large Woody Debris Jams, Channel Hydraulics and Habitat
Formation in Large Rivers. Regulated Rivers: Research and Management 12:201–221.

Abbe, T. B., and D. R. Montgomery. 2003. Patterns and Processes of Wood Debris Accumulation in
the Queets River Basin, Washington. Geomorphology 51:81–107.

Abbe, T. B., D. R. Montgomery, and C. Petroff. 1997. Design of Stable In-Channel Wood Debris
Structures for Bank Protection and Habitat Restoration: An Example from the Cowlitz River, WA.
Pages 809–816 in S. S. Y. Wang, E. J. Langendoen, and F. D. Shields, F.D. (eds.), Proceedings of the
Conference on Management of Landscapes Disturbed by Channel Incision. University of
Mississippi, Oxford, MS.

Abbe, T. B., J. Carrasquero, M. McBride, A. Ritchie, M. McHenry, and K. Dublanica. 2003a.


Rehabilitating River Valley Ecosystems: Examples of Public, Private, and First Nation Cooperation
in Western Washington. Proceedings of the Georgia Basin/Puget Sound 2003 Research
Conference, Vancouver, B.C., March 31–April 1, 2003, T. Droscher (ed.). Puget Sound Action
Team, Olympia, WA.

Large Wood National Manual July 2015


4-67
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Abbe, T. B, A. P. Brooks, and D. R. Montgomery. 2003b. Wood in River Rehabilitation and


Management. Pages 367–389 in S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology
and Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society.

Abbe, T. B., G. Pess, D. R. Montgomery, and K. L. Fetherston. 2003c. Integrating Engineered Log Jam
Technology into River Rehabilitation. In D. R. Montgomery, S. Bolton, D. Booth, and L. Wall
(eds.), Restoration of Puget Sound Rivers. Center for Water and Watershed Studies.

Abbe, T. B., C. Miller, and A. Michael. 2009. Self-Mitigating Protection for Pipeline Crossings in
Degraded Streams: A Case Study from Woodward Creek, Washington. 9th International Right of
Way Symposium. 2009. Portland, OR.

Abbe. T., J. Bjork, A. Zehni, T. Nelson, and J. Park. 2011. New Innovative, Habitat-Creating Bank
Protection Method. Pages 2011–2021 in Proceedings of ASCE World Environmental and Water
Resources Congress.

Abbe. T., M. Ericsson, and L. Embertson, L. 2013. Geomorphic Assessment of the Larson Reach of the
South Fork Nooksack River, NW Washington. Report submitted to Lummi Indian Nation.

Ahmad, M. 1951. Spacing and Projection of Spurs for Bank Protection. Civil Engineering and Public
Works Review. March:172–174; April:256–258.

Agee, J. K. 1990. The historical role of fire in Pacific Northwest forests. Pages 25–38 in J. Walstad, S.
R. Radosevich, and D. V. Sandberg (eds.), Natural and Prescribed Fire in Pacific Northwest Forests.
Corvallis: Oregon State University Press.

Agee, J. K. 1992. The Historical Role of Fire in Pacific Northwest Forests. Pages 25–38 in J. Walstad, S.
R. Radosevich, and D. V. Sandberg (eds.), Natural and Prescribed Fire in Pacific Northwest Forests.
Corvallis: Oregon State University Press.

Agee, J. K. 1993. Fire Ecology of Pacific Northwest Forests. Island Press, Wash. D.C.

Anderson, B. 2005. Will Replanting Vegetation along River Banks Make Floods Worse? 8th
International River Symposium, Brisbane, Australia.

Anderson, D. B. 2006. Quantifying the Interaction between Riparian Vegetation and Flooding: from
Cross-Section to Catchment Scale. University of Melbourne.

Andrus, C. W., B. A. Long, and H. A. Froehlich. 1988. Woody debris and its contribution to pool
formation in a coastal stream 50 years after logging: Canadian Journal of Fish and Aquatic
Science 45:2080–2086.

Arno, S. F., and R. J. Hoff. 1989. Silvics of Whitebark Pine (Pinus albicaulis). USDA For. Serv. Gen Tech.
Rep. INT-253.

Babakaiff, S., D. Hay, and C. Fromuth. 1997. Rehabilitating Stream Banks. In P. A. Slaney and D.
Zaldokas (eds.), Fish Habitat Rehabilitation Procedures, Watershed Restoration Program. Ministry
of Environment, Lands and Parks, Vancouver, BC.

Baillie, B. R., and T. R. Davies. 2002. Influence of Large Woody Debris on Channel Morphology in
Native Forest and Pine Plantation Streams in the Nelson Region, New Zealand. New Zealand
Journal of Marine and Freshwater Research. 36:763–774.

Large Wood National Manual July 2015


4-68
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Baillie, B. R., L. G. Garret, and A. W. Evanson. 2008. Spatial Distribution Influence of LWD in an Old-
growth Forest River System, New Zealand. Forest Ecology and Management 256:20–27.

Baker, C. O. 1979. The Impacts of Logjam Removal on Fish Populations and Stream Habitat in Western
Oregon. M.S. Thesis, Oregon State University, Corvallis, OR.

Barker, B. L., R. D. Nelson, and M. S. Wigmosta. 1991. Performance of detention ponds designed
according to current standards. Puget Sound Water Quality Authority, Puget Sound Research '91:
Conference Proceedings. Seattle, Washington.

Barrett, M. L. 1996. Environmental Reconstruction of a 19th-Century Red River Raft Lake: Caddo
Lake, Louisiana and Texas. Gulf Coast Association of Geological Societies Transactions 46 471–
471.

Beechie, T. J., and K. Wyman. 1992. Stream Habitat Conditions, Unstable Slopes and Status of Roads in
Four Small Watersheds of the Skagit River. Skagit System Cooperative, Fisheries services for the
Swinomish Tribal Community, Upper Skagit and Sauk-Suiattle Indian Tribes.

Beets, P. N., I. A. Hood, M. O. Kimberley, G. R. Oliver, S. H. Pearce, and J. F. Gardner. 2008. Coarse
Woody Debris Decay Rates for Seven Indigenous Tree Species in the Central North Island of New
Zealand. Forest Ecology and Management 256:548–557.

Beltaos, S. 1983. River Ice Jams: Theory, Case Studies, and Applications. Journal of Hydraulic
Engineering 109(10):1338–1359.

Benda, L. and T. W. Cundy. 1990. Predicting Deposition of Debris Flow in Mountain Channels.
Canadian Geotechnical Journal 27:409–417.

Bilby, R. E. 1984. Removal of Woody Debris May Affect Stream Channel Stability. Journal of Forestry,
609–613. October.

Bilby, R. E., and P. A. Bisson. 1998. Function and distribution of large woody debris. Pages 324–346
in R. J. Naiman and R. E. Bilby (eds.), River Ecology and Management. New York, Springer.

Bilby, R. E., and G. E. Likens. 1980. Importance of debris dams in the structure and function of stream
ecosystems. Ecology 61:1107–1113.

Bilby, R. E. and J. W. Ward. 1991. Characteristics and Function of Large Woody Debris in Streams
Draining Old-Growth, Clear-Cut, and Second-Growth Forests in Southwestern Washington.
Canadian Journal of Fisheries and Aquatic Sciences 48:2499–2508.

Bilby, R. E., and L. J. Wasserman. 1989. Forest Practices and Riparian Management in Washington
State: Data Based Regulation Development. In R. E. Gresswell, B. A. Barton, and J. L. Kershner
(eds.), Practical Approaches to Riparian Management. U.S. Bureau of Land Management, BLM MT
PT 89 001 4351, Billings, Montana.

Bisson, P. A., R. E. Bilby, M. D. Bryant, C. A. Dolloff, G. B. Grette, R. A. House, M. L. Murphy, K. V. Koski,


and J. R. Sedell. 1987. Large Woody Debris in Forested Streams in the Pacific Northwest: Past,
Present, and Future. Pages 143–190, in E. O. Salo and T. W. Cundy (eds.), Streamside
Management: Forestry and Fishery Interactions. College of Forest Resources, University of
Washington, Seattle, Washington.

Large Wood National Manual July 2015


4-69
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Boose, E. R., K. E. Chamberlin, and D. R. Foster. 2001. Landscape and Regional Impacts of Hurricanes
in New England. Ecological Monographs 71:27–48.

Booth, D. B. 1990. Stream-Channel Incision Following Drainage-Basin Urbanization. Water Resources


Bulletin 26:407–417.

Booth, D. 1991. Urbanization and the Natural Drainage System: Impacts, Solutions, and Prognoses.
The Northwest Environmental Journal 7, 93-118.

Booth, D., and L. E. Reinelt. 1993. Consequences of Urbanization on Aquatic Systems—Measured


Effects, Degradation Thresholds, and Corrective Strategies. Watershed ’93, American Water
Resources Association, pages 545–550.

Bovee, K. D., T. J. Newcomb, and T. G. Coon. 1994. Relations Between Habitat Variability and
Population Dynamics of Bass in the Huron River, Michigan. Biological Report 21. National
Biological Survey, U.S. Department of the Interior. Washington D.C.

Braudrick, C. A., and G. E. Grant. 2000. When do Logs Move in Rivers? Water Resources Research
36(2):571–583.

Brooks, A. P., and G. J. Brierly. 2002. Mediated Equilibrium: The Influence of Riparian Vegetation and
Wood on the Long-Term Evolution and Behavior of a Near-Pristine River. Earth Surface
Processes and Landforms 27:343–367.

Brooks, A., and G. J. Brierly. 2004. Framing Realistic River Rehabilitation Targets in Light of Altered
Sediment Supply and Transport Relationships: Lessons from East Gippsland, Australia.
Geomorphology 58:107–123.

Brooks, A. P., G. J. Brierly, and R. G. Millar. 2003. The Long-Term Control of Vegetation and Woody
Debris on Channel and Flood-Plain Evolution: Insights from a Paired Catchment Study in
Southeastern Australia. Geomorphology 51:7–30.

Brooks, A. P., P. Gehrke, J. D. Jansen, and T. B. Abbe. 2004. Experimental Reintroduction of Woody
Debris on the Williams River, NSW: Geomorphic and Ecological Responses. River Research and
Applications 20:513–536.

Brooks, A. P., T. Howell, T. B. Abbe, and A. H. Arthington. 2006. Confronting Hysteresis: Wood Basin
River Rehabilitation in Highly Altered Riverine Landscapes of South-Eastern Australia.
Geomorphology 79(3/4):395–422.

Brummer, C., T. B. Abbe, J. R. Sampson, and D. R. Montgomery. 2006. Influence of Vertical Channel
Change Associated with Wood Accumulations on Delineating Channel Migration Zones,
Washington State, USA. Geomorphology 80:295–309.

Bryant, M. D. 1980. Evolution of large, Organic Debris after Timber Harvest: Maybeso Creek, 1949 to
1978. USDA Forest Service, General Technical Report, PNW-101.

Bryant, M. D. 1983. The Role and Management of Woody Debris in West Coast Salmonid Nursery
Stream. North American Journal of Fisheries Management 3(3):322–330.

Bryant, M. D., and J. R. Sedell. 1995. Riparian Forests, Wood in the Water, and Fish Habitat
Complexity. Pages 202–224 in N. B. Armantrout and R. J. Wolotira, Jr. (eds.), Conditions of the

Large Wood National Manual July 2015


4-70
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

World's Aquatic Habitats. Proceedings of the World Fisheries Congress Theme 1. Oxford and IBH
Publishing Co. Pvt. Ltd., New Delhi.

Buffington, J. M., and D. R. Montgomery. 1999a. A Procedure for Classifying Textural Facies in Gravel-
Bed Rivers. Water Resources Research 35(6):1903-1914.

Buffington, J. M., and D. R. Montgomery. 1999b. Effects of Hydraulic Roughness on Surface Textures
of Gravel-Bed Rivers. Water Resources Research 35(11):3507–3521.

Buffington, J. M., and D. R. Montgomery. 1999c. Effects of Sediment Supply on Surface Textures of
Gravel-Bed Rivers. Water Resources Research 35(11):3523–3530.

Clay, C. 1949. The Colorado River Raft. The Southwestern Historical Quarterly 102 (4):400–426.

Camp, A., C. Oliver, P. Hessburg, and R. Everett. 1996. Predicting Late-Successional Fire Refugia Pre-
Dating European Settlement in the Wenatchee Mountains. USDA PNW, Wenatchee For. Sci. Lab.,
Univ. of Washington, Seattle. Elsevier Science Publishers B.V. Forest Ecology and Management
95:63–77.

Cederholm, C. J., W. J. Scarlett, N. P. and Peterson. 1988. Low-Cost Enhancement Technique for
Winter Habitat of Juvenile Coho Salmon. North American Journal of Fisheries Management
8:438–441.

Cederholm, C. J., R. E. Bilby, P. A. Bisson, T. W. Bumstead, B. R. Fransen, W. J. Scarlett, and J. W. Ward.


1997b. Response of Juvenile Coho Salmon and Steelhead to the Placement of Large Woody
Debris in a Coastal Washington Stream. Transactions of the American Fisheries Society. 118:368–
378.

Chambers, J. Q., J. I. Fisher, H. Zeng, E. L. Chapman, D. B. Baker, and G. C. Hurtt. 2007. Hurricane
Katrina’s Carbon Footprint on U.S. Gulf Coast Forests. Science 318 (5853):1107.

Chesney, C. 2000. Functions of Wood in Small, Steep Streams in Eastern Washington: Summary of
Results for Project Activity in the Ahtanum, Cowiche, and Tieton Basins. Washington Department
of Natural Resources. Prepared for the Timber/Fish/Wildlife Monitoring Advisory Group and
the Northwest Indian Fisheries Commission. TFW Effectiveness Monitoring Report: TFW-MAGl-
00-002.

Clay, C. 1949. The Colorado River Raft. The Southwestern Historical Quarterly 102 (4):400–426.

Coho, C., and S. J. Burges. 1993. Dam-Break Floods in Low Order Mountain Channels of the PNW.
Water Resources Series Tech Rep no. 138. Dept. Civil Engineering, Univ. of Washington, Seattle. 68
pp.

Collins, B. D., and A. J. Sheikh. 2005. Historical Reconstruction, Classification, and Change Analysis of
Puget Sound Tidal Marshes. University of Washington (Seattle, WA) and the Nearshore Habitat
Program, Washington State Dept. of Natural Resources, Olympia, WA. See more at:
http://www.eopugetsound.org/science-review/3-tidal-wetlands#sthash. T4OyhfFd.dpuf

Collins, B. D., D. R. Montgomery, and A. D. Haas. 2002. Historical Changes in the Distribution and
Functions of Large Wood in Puget Lowland Rivers. Canadian Journal of Fisheries and Aquatic
Sciences 59:66–76.

Large Wood National Manual July 2015


4-71
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Collins, B. D., D. R. Montgomery, and A. J. Sheikh. 2003. Reconstructing the Historical Riverine
Landscape of the Puget Lowland. Pages 79–128 in D. R. Montgomery, S. M. Bolton, D. B. Booth,
and L. Wall (eds.), Restoration of Puget Sound Rivers. University of Washington Press, Seattle.

Collins, B. D., D. R. Montgomery, K. L. Fetherston, and T. B. Abbe. 2012. The Floodplain Large-Wood
Cycle Hypothesis: A Mechanism for the Physical and Biotic Structuring of Temperate Forested
Alluvial Valleys in the North Pacific Coastal Ecoregion. Geomorphology 139/140:460–470.

Comiti, F., A. Andreoli, M. A. Lenzi, and L. Mao. 2006. Spatial Density and Characteristics of Woody
Debris in Five Mountain Rivers of the Dolomites (Italian Alps). Geomorphology 78:44–63.

Comiti, F., A. Andreoli, L. Mao, and M. A. Lenzi. 2008. Wood Storage in Three Mountain Streams of the
Southern Andes and its Hydro-Morphological Effects. Earth Surface Processes and Landforms
33:244–262.

Copeland, R. R. 1983. Bank Protection Techniques Using Spur Dikes. Paper No. HL-83-1. Hydraulics
Laboratory. U.S. Army Waterways Experiment Station. Vicksburg, MS.

Cordova, J. M., E. J. Rosi-Marshall, A. M. Yamamuro, and G. A. Lamberti. 2007. Quantity, Controls, and
Functions of Large Woody Debris in Midwestern USA Streams. River Research and Applications
23:21–23.

Costa, J. E. 1984. Physical Geometry of Debris Flows. Pages 268–317 in J. E. Costa and P. J. Fleisher
(eds.), Developments and Applications of Geomorphology. Springer-Verlag, Berlin.

Cushman, M. J. 1981. The Influence of Recurrent Snow Avalanches on Vegetation Patterns in the
Washington Cascades. Ph.D. dissertation. University of Washington, Seattle, Washington.

Dacy, G. H. 1921. Pulling the Mississippi’s Teeth. Scientific American 75(4):60, 70.

Daley, J. S. 2012. Taming the Hungry Beast: The Effectiveness of Engineered Log Jams in an Incising
Gravel Bedded River. B.S. Honors Thesis, School of Earth and Environmental Science, University
of Wollongong, Australia. Available: http://ro.uow.edu.au/thsci/38.

Daniels, M. D., and B. L. Rhoads. 2004. Spatial Pattern of Turbulence Kinetic Energy and Shear Stress
in a Meander Bend with Large Woody Debris. Pages 87–97 in S. J. Bennett and A. Simon (eds.),
Riparian Vegetation and Fluvial Geomorphology, Water Science and Application 8. American
Geophysical Union, Washington D.C.

Dickman, A., and S. Cook. 1989. Fire and Fungus in a Mountain Hemlock Forest. Canadian Journal of
Botany 67:2005–2016.

Dominguez, L. G., and C. J. Cederholm. 2000. Rehabilitating Stream Channels Using Large Woody
Debris with Considerations for Salmonid Life History and Fluvial Geomorphic Processes. Pages
545–563 in E. E. Knudsen. C. R. Steward, D. D. MacDonald, J. E. Williams, and D. W. Reiser (eds.),
Sustainable Fisheries Management: Pacific Salmon. Lewis Publishers, New York.

Doyle, M. W., and F. D. Shields, Jr. 2000. Incorporation of Bed Texture into a Channel Evolution
Model. Geomorphology 34:291–309.

Dunne, T., and L. B. Leopold. 1978. Water in Environmental Planning. New York, NY: W.H. Freeman &
Co.

Large Wood National Manual July 2015


4-72
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Eaton, B. C. 2006. Bank Stability Analysis for Regime Models of Vegetated Gravel Bed Rivers. Earth
Surface Processes and Landforms 31:1438–1444.

Eaton, B. C., M. Chuch, and R. G. Millar. 2004. Rational Regime Model of Alluvial Channel Morphology
and Response. Earth Surface Processes and Landforms 29:511–529.

Eaton, B.C., R. C. Millar, and S. Davidson. 2010. Channel Patterns: Braided, Anabranching and Single
Thread. Geomorphology 120:353–364.

Edmonds, R. L., T. B. Thomas, and K. P. Maybury. 1993. Tree Population Dynamics, Growth, and
Mortality in old-Growth Forests in the Western Olympic Mountains, Washington. Canadian
Journal of Forest Research 23:512–519.

Ehrman, T. P., and G. A. Lamberti. 1992. Hydraulic and Particulate Matter Retention in a 3rd-Order
Indiana Stream. Journal of the North American Benthological Society. 11:341–349.

Elmore, W., and R. L. Beschta. 1988. The Fallacy of Structures and the Fortitude of Vegetation. Proc.
of Calif. Riparian Systems Conference. Davis, Calif.

Ensign, S. H., and M. W. Doyle. 2005. In-channel Transient Storage and Associated Nutrient
Retention: Evidence from Experimental Manipulations. Limnology and Oceanography
50(6):1740–1751.

Erskine, W. D., and A. A. Webb. 2003. Desnagging to Resnagging: New Directions in River
Rehabilitation in Southeastern Australia. River Research and Applications. 19:233–249.

Fahnestock, G. R. 1976. Fires, Fuel, and Flora as Factors in Wilderness Management: The Pasayten
Case. Tall Timbers Fire Ecology Conf. 15:33–70.

Fahnestock, G. R., and J. K. Agee. 1983. Biomass Consumption and Smoke Production by Prehistoric
and Modern Forest Fires in Western Washington. Journal of Forestry 81:653–657.

Fetherston, K. L., R. J. Naiman, and R. E. Bilby. 1995. Large Woody Debris, Physical Process, and
Riparian Forest Development in Montane River Networks of the Pacific Northwest.
Geomorphology 13:133–144. Elsevier Science B.V.

Federal Highway Administration (FHWA). 1985. Design of Spur-Type Streambank Stabilization


Structures. Federal Highway Administration Report No. FHWA/RD-84/101. U.S. Department of
Transportation. Washington D.C. 112 pp.

Federal Writers’ Project. 1952. West Virginia A Guide to the Mountain State. U.S. History Publishers.

Fisher, G. B., F. J. Magilligan, J. M. Kaste, and K. H. Nislow. 2010. Constraining the Timescales of
Sediment Sequestration Associated with Large Woody Debris using Cosmogenic 7Be. Journal of
Geophysical Research 115 (F3), DOI: 10.1029/2009JF001352.

Flynn, K.M., Kirby, W.H., and Hummel, P.R. 2006. User’s Manual for Program PeakFQ Annual Flood-
Frequency Analysis Using Bulletin 17B Guidelines: U.S. Geological Survey, Techniques and Methods.
Book 4, Chapter B4. 42 pp.

Foster, D. R., and E. R. Boose. 1992. Patterns of Forest Damage Resulting from Catastrophic Wind in
Central New England, USA. Journal of Ecology 80:79–98.

Large Wood National Manual July 2015


4-73
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Fox, M. J. 2001. A New Look at the Quantities and Volumes of Instream Wood in Forested Basins within
Washington State. Master of Science thesis. College of Forest Resources, University of
Washington.

Fox, M. J. 2003. Spatial Organization, Position, and Source Characteristics of Large Woody Debris in
Natural Systems. Ph.D. dissertation. College of Forest Resources, University of Washington.
Seattle, Washington.

Fox, M. J. and S. Bolton. 2007. A Regional and Geomorphic Reference for Quantities and Volumes of
Instream Wood in Unmanaged Forested Basins of Washington State. North American Journal of
Fisheries Management 27:342–359.

Frangi, J. L., and A. E. Lugo. 1991. Hurricane Damage to a Flood Plain Forest in the Luquillo
Mountains of Puerto Rico. Biotropica 23(4a):324–335.

Franklin, J. F., and C. T. Dyrness. 1973. Natural Vegetation of Oregon and Washington. USDA Forest
Service. Gen. Tech. Rep. PNW-8.

Gastaldo, R. A., and C. W. Degges. 2007. Sedimentology and Paleontology of a Carboniferous Log Jam.
International Journal of Coal Geology 69:103–113.

Gastaldo, R. A., and T. M. Demko. 2011. The Relationship Between Continental Landscape Evolution
and the plant-Fossil Record: Long Term Hydrologic Controls on Preservation. Pages 249–285 in
P. A. Allison and D. J. Bottjer (eds.), Taphonomy: Process and Bias Through Time. Aims & Scope
Topics in Geobiology Volume 32. Springer Netherlands.

Gibling, M. R., A. R. Bashforth, H. J. Falcon-Lang, J. P. Allen, and C. R. Fielding. 2010. Log Jams and
Flood Sediment Buildup Caused Channel Abandonment and Avulsion in the Pennsylvanian of
Atlantic Canada. Journal of Sedimentary Research 80:268–287.

Gippel, C. J. 1995. Environmental Hydraulics of Large Woody Debris in Streams and Rivers. Journal of
Environmental Engineering 121:388–395.

Gippel, C. J., I. C. O’Neill, and B. L. Finlayson. 1992. The Hydraulic Basis of Snag Management. Center
for Environmental Applied Hydrology, University of Melbourne, Melbourne, Victoria, Australia,
115 pp.

Gippel, C. J., I. C. O’Neill, B. L. Finlayson, and I. Schnatz, I. 1994. Hydraulic Guidelines for
Reintroduction and Management of Large Woody Debris in Degraded Lowland Rivers. Pages 225–
239 in Proceedings of the Conference on Habitat Hydraulics. International Association for
Hydraulic Research.

Gippel, C. J., I. C. O’Neill, and B. L. Finlayson. 1996. Distribution and Hydraulic Significance of Large
Woody Debris in a Lowland Australian River. Hydrobiologia 318:179–194.

Gotvald, A. J., N. A. Barth, A. G. Veilleux, and C. Parrett. 2012. Methods for Determining Magnitude and
Frequency of Floods in California, Based on Data Through Water Year 2006. U.S. Geological Survey
Scientific Investigations Report 2012–5113. Available: http://pubs.usgs.gov/sir/2012/5113/.

Graf, W. L. 1975. The impact of Suburbanization on Fluvial Morphology. Water Resources Research
11:690–692.

Large Wood National Manual July 2015


4-74
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Graham, R., and K. Cromack. 1982. Mass, Nutrient Content, and Decay Rate of Dead Boles in Rain
Forests of Olympic National Park. Canadian Journal of Forest Research 12(3):511–521.

Grant, G. E., and F. J. Swanson. 1995. Morphology and Processes of Valley Floors in Mountain
Streams, western Cascades, Oregon. Pages 83–101 in J. D. Costa, A. J. Miller, K. W. Potter, and P.
R. Wilcock (eds.). Natural and Anthropogenic Influences in Fluvial Geomorphology. Geophysical
Monograph 89. American Geophysical Union, Washington DC.

Grant, G. E., M. J. Crozier, and F. J. Swanson. 1984. An Approach to Evaluating Off-Site Effects of
Timber Harvest Activities on Channel Morphology. Proceedings of the Symposium on the Effects
of Forest and Land Use on Erosion and Slope Stability. Environment and Policy Institute, E-West
Center, University of Hawaii, Honolulu 177–186.

Gray, D. H. 1974. Reinforcement and Stabilization of Soil by Vegetation. Journal of Geotechnical


Engineering, 100(GT6):695–699.

Gray, D. H. and D. Barker. 2004. Root-soil Mechanics and Interactions. Pages 113–123 in S. J. Bennett
and A. Simon (eds.). Riparian Vegetation and Fluvial Geomorphology, Water Science and
Application 8, American Geophysical Union, Washington D.C.

Gregory, S. V., F. J. Swanson, W. A. McKee, and K. W. Cummins. 1991. An Ecosystem Perspective of


Riparian Zones. BioScience 41(8):540–551.

Grette, G. B. 1985. The role of Large Organic Debris in Juvenile Salmonid Rearing Habitat in Small
Streams. MS thesis, University of Washington, Seattle, WA.

Grimm, W. C. 1967. Familiar Trees of America. New York: Harper & Row.

Grizzel, J. D., and N. Wolff. 1998. Occurrence of Windthrow in Forest Buffer Strips and its Effect on
Small Streams in Northwest Washington. Northwest Science 72:214–223.

Guardia, J. E. 1933. Some Results of the Log Jams in the Red River. The Bulletin of the Geographical
Society of Philadelphia 31(3):103–114.

Gurnell, A. M., G. E. Petts, N. Harris, J. V. Ward, K. Tockner, P. J. Edwards, and J. Kollman. 2000. Large
Wood Retention in River Channels: The Case of the Fiume Tagliamento, Italy. Earth Surface
Processes and Landforms 25:255–275.

Guyette, R. P., and M. Stambaugh. 2003. The Age and Density of Ancient and Modern Oak Wood in
Streams and Sediments. The International Association of Wood Anatomists (IAWA) Journal
24:345–353.

Guyette, R. P., D. C. Dey, and M. C. Stambaugh 2008. The Temporal Distribution and Carbon Storage
of Large Oak Wood in Streams and Floodplain Deposits. Ecosystems 11:643–653.

Hall, R. O., and J. L. Meyer. 1998. The Trophic Significance of Bacteria in a Detritus-Based Stream
Food Web. Ecology 79:1995–2012.

Hammer, T. R. 1972. Stream Channel Enlargement due to Urbanization. Water Resources Research
8:1530–1540.

Harmon, M. E., J. F. Franklin, F. J. Swanson, P. Sollins, S. V. Gregory, J. D. Lattin, N. H. Anderson, S. P.


Cline, N. G. Aumen, J. R. Sedell, G. W. Lienkaemper, K. Cromack, Jr., and K. W. Cummins. 1986.

Large Wood National Manual July 2015


4-75
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Ecology of Coarse Woody Debris in Temperate Ecosystems. Advances in Ecological Research


15:133–302.

Harrod, R. 2000. Ecologist with the Wenatchee National Forest Service, Wenatchee, WA. Personal
Communication.

Hartopo, 1991. The Effect of Raft Removal and Dam Construction on the Lower Colorado River, Texas.
Unpublished M.S. Thesis, Texas A & M University.

Harvey, M. D, D. S. Biedenharn, and P. Combs. 1988. Adjustments of Red River Following Removal of
the Great Raft in 1873 [abs.]. Eos, Transactions of the American Geophysical Union 69(18):567.

Hauer, F. R. 1989. Organic Matter Transport and Retention in a Blackwater Stream Recovering from
Flow Augmentation and Thermal Discharge. Regulated Rivers: Research and Management 4:371–
380.

Hazard, J. T. 1948. Our Living Forests, the Story of Their Preservation and Multiple Use. Seattle, WA:
Superior Publishing.

Hedman, C. W., D. H. Van Lear, and W. T. Swank. 1996. In-stream LWD Loading and Riparian Forest
Serial Stage Associations in the Southern Appalachian Mountains. Canadian Journal of Forest
Research 26(7):1218–1227.

Henderson, J. 1996. Unpublished Data Regarding Tree Height vs. Age for Two Common Plant
Association Groups. USDA Forest Service, Pacific Northwest Region, Mount Lake Terrace, WA.

Henderson, J. A., R. D. Lesher, D. H. Peter and D. C. Shaw. 1992. Field Guide to the Forested Plant
Associations of the Mt. Baker-Snoqualmie National Forest. USDA Forest Service, Pacific Northwest
Region. Tech paper R6 ECOL TP 028-91. Hershey, K. 1995. Characteristics of Forests at Spotted
Owl Nest Sites in the Pacific Northwest. M.S. thesis, Oregon State University, Corvallis.Hewitt, E. R.
1934. Hewitt’s Handbook of Stream Improvement. The Marchbanks Press, New York.

Hickin E. J. 1984. Vegetation and River Channel Dynamics. Canadian Geographer 28(2):111–126.

Hilderbrand, R. H., A. D. Lemly, C. A. Dollof, and K. L. Harpster. 1997. Effects of Large Woody Debris
Placement on Stream Channels and Benthic Macroinvertebrates. Canadian Journal of Fisheries
and Aquatic Sciences 54:931–939.

Hollis, G. E. 1975. The Effects of Urbanization on Floods of Different Recurrence Intervals. Water
Resources Research 11:431–435.

Holstine, C. 1992. An Historical Overview of the Wenatchee National Forest, Washington. Rep. 100-80.
Archaeological and historical Services. Eastern Washington University, Cheney.

Horner, R. R., D. B. Booth, A. Azous, and C. W. 1997. Watershed Determinants of Ecosystem


Functioning. Pages 251–274 in L. A. Roesner (ed.), Effects of Watershed Development and
Management on Aquatic Ecosystems, American Society of Civil Engineers, New York, NY.

House, R. A., and P. L. Boehne. 1986. Effects of Instream Structures on Salmonid Habitat and
Populations in Tobe Creek, Oregon. North American Journal of Fisheries Management 6:283–295.

Hyatt, T. L., and R. J. Naiman. 2001. The Residence Time of Large Woody Debris in the Queets River,
Washington, USA. Ecological Applications 11(1):191–202.

Large Wood National Manual July 2015


4-76
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Hygelund, B., and M. Manga. 2003. Field Measurements of Drag Coefficients for Model Large Woody
Debris. Geomorphology 51:175–185.

Ikeya, H. 1981. A Method for Designation Forested Areas in Danger of Debris Flows. In Erosion and
Sediment Transport in Pacific Rim Steeplands. Edited by T. R. H. Davies and A. J. Pearce.
International Association of Hydrological Sciences, Publication 132:576–588.

Interagency Advisory Committee on Water Data (IACWD). 1982. Guidelines for Determining Flood
Flow Frequency. Bulletin 17B of the Hydrology Subcommittee, Office of Water Data Coordination,
U.S. Geological Survey, Reston, Virginia. 183 p.

Jacobson, P. J., K. M. Jacobson, P. L. Angermeier, and D. S. Cherry. 1999. Transport, Retention, and
Ecological Significance of Woody Debris within a Large Ephemeral River. Journal of the North
American Benthological Society 18:429–444.

James, L. D. 1965. Using a Digital Computer to Estimate the Effects of Urban Development on flood
Peaks. Water Resources Research 1:223–234.

Jenkins, M. J., and E. G. Hebertson. 1998. Using Vegetative Analysis to Determine the Extent and
Frequency of Avalanches in Little Cottonwood Canyon, Utah. Department of Forest Resources,
Utah State University. WestWide Avalanche Network, UT.

Johnson, L. B., D. H. Breneman, and C. Richards. 2003. Macroinvertebrate Community Structure and
Function Associated with Large Wood in Low Gradient Streams. River Research and Applications
19:199–218.

Johnson, S. L., F. J. Swanson, G. E. Grant, and S. M. Wondzell. 2000. Riparian Forest Disturbances by a
Mountain Flood—The Influence of Floated Wood. Hydrological Processes 14:3031–3050.

Johnston, N. T., E. A. MacIsaac, P. J. Tschaplinski, and K. J. Hall. 2004. Effects of the Abundance of
Spawning Sockeye Salmon (Oncorhynchus nerka) on Nutrients and Algal Biomass in Forested
Streams. Canadian Journal of Fisheries and Aquatic Sciences 61:384–403.

Kanes, W. H. 1970. Facies and Development of the Colorado River Delta in Texas. Pages 78–106 in J.
P. Morgan and R. H. Shaver (eds.), Deltaic Sedimentation Modern and Ancient. Special Publication
No.15. Society of Economic Paleontologists and Mineralogists. Tulsa, Oklahoma.

Kauffman, J. B., R. L. Beschta, N. Otting, and D. Lytjen. 1997. An Ecological Perspective of Riparian
and Stream Restoration in the Western United States. Fisheries (Bethesda) 22:12–24.

Keeton, W. S., C. E. Kraft, and D. R. Warren. 2007. Mature and Old-Growth Riparian Forests:
Structure, Dynamics and Effects on Adirondack Stream Habitats. Ecological Applications 17:852–
868.

Keller, E. A. and A. MacDonald. 1995. River Channel Change: The Role of Large Woody Debris. Pages
217–236 in A. Gurnell and G. Petts (eds.), Changing River Channels. John Wiley and Sons,
Chichester. 217-235.

Keller, E. A., and F. J. Swanson. 1979. Effects of Large Organic Material on Channel Form and Fluvial
Processes. Earth Surface Processes 4:361–380.

Large Wood National Manual July 2015


4-77
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Keller, E. A., and T. Tally. 1979. Effects of Large Organic Debris on Channel Form and Fluvial
Processes in the Coastal Redwood Environment. Pages 169–197 in D. D. Rhodes and G. P.
Williams (eds.), Adjustments of the Fluvial System. Proceedings of the 10th Annual Binghamton
Geomorphology Symposium. Kendal-Hunt. Dubuque, IA.

Kennard, P., G. Pess, T. Beechie, B. Bilby, and D. Berg. 1998. Riparian-in-a-Box: A Manager’s Tool to
Predict the Impacts of Riparian Management on Fish Habitat. Pages 483-490. in M. K. Brewin
and D. M. A. Monita (eds.), Forest-Fish Conference: Land Management Practices Affecting Aquatic
Ecosystems. Proceedings of Forest-fish conference, May 1-4, 1996, Calgary, Alberta. Natural
Resources Canada. North For. Cent., Edmonton, Alberta Inf. Rep. NOR-X-356.

Klingeman, P. C., S. M. Kehe, and Y. A. Owusu. 1984. Streambank Erosion Protection and Channel
Scour Manipulation Using Rockfill Dikes and Gabions. Water Resources Research Institute,
Oregon State University. Salem, OR.

Konsoer, K. M. 2014. Influence of Riparian Vegetation on Near-Bank Flow Structure and Rates of
Erosion on a Large Meandering River. Ph.D. Dissertation, University of Illinois, Urbana-
Champaign. 236 p.

Kraft, C. E., and D. R. Warren. 2003. Development of Spatial Pattern in Large Woody Debris and
Debris Dams in Streams. Geomorphology 51:127–139.

Krause, C., and C. Roghair. 2014. Inventory of Large Wood in the Upper Chattooga River Watershed,
2007–2013. U.S. Forest Service Southern Research Station, Center for Aquatic Technology
Transfer. Blacksburg, VA.

Lampert, W. 1978. Release of Dissolved Organic-Carbon by Grazing Zooplankton. Limnology and


Oceanography 23(4):831–834.

Langbien, W. B., and S. B. Schumm. 1958. Yield of Sediment in Relation to Mean Annual Precipitation.
American Geophysical Union Transactions 39:1076–1084.

Latterell, J. J., and R. J. Naiman. 2007. Sources and Dynamics of Large Logs In a Temperate Floodplain
River. Ecological Applications 17:1127–1141.

Latterell, J. J., J. S. Bechtold, T. C. O'Keefe, R. Van Pelt, and R. J. Naiman. 2006. Dynamic Patch Mosaics
and Channel Movement in an Unconfined River Valley of the Olympic Mountains Freshwater
Biology 51(3):523–544.

Lehane, B. M., P. S. Giller, J. O’Halloran, C. Smith, J. Murphy. 2002. Experimental Provision of Large
Woody Debris in Streams as a Trout Management Technique. Aquatic Conservation-Marine and
Freshwater Ecosystems 12:289–311.

Leopold, L. B. 1973. River Channel Change with Time: An Example. Geological Society of America
Bulletin 84:1845–1860.Lester, R. E., and W. Wright. 2009. Reintroducing Wood to Streams in
Agricultural Landscapes: Changes in Velocity Profile, Stage and Erosion Rates. River Research
and Applications 25(4):276–392.

Li, R., and H. W. Shen. 1973. Effect of Tall Vegetation on Flow and Sediment. Journal of
the Hydraulic Division, ASCE 99(5):793–814.

Lienkaemper, G. W., and F. J. Swanson. 1987. Dynamics of Large Woody Debris in Streams in Old-
Growth Douglas-Fir Forests. Canadian Journal of Forest Research 17:150–156.

Large Wood National Manual July 2015


4-78
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Lisle, T. 1995. Effects of Coarse Woody Debris and its Removal on a Channel Affected by the 1980
Eruption of Mount St. Helens, Washington. Water Resources Research 31:1797–1808.

Lowrey, W. M. 1968. The Red. Pages 53–73 in E. A. Davis, The Rivers and Bayous of Louisiana.
Louisiana Education Research Association, Baton Rouge, LA.

Luchetti, G. and R. Fuerstenberg. 1993. Management of Coho Salmon Habitat in Urbanizing


:andscapes of King County, Washington, USA. Pages 308–317 In L. Berg and P. Delaney (eds.),
Proceedings of the 1992 Coho Salmon Workshop. Canada Department of Fisheries and Oceans,
Vancouver, B.C., Canada.

Lunetta, R. S., B. L. Cosentino, D. R. Montgomery, E. M. Beamer, and T. J. Beechie. 1997. GIS-Based


Evaluation of Salmon Habitat in the Pacific Northwest. Photogrammetric Engineering & Remote
Sensing 63(10):1219–1229.

Lyell, C. 1830. Principles of Geology, Volume I. London, UK: John Murray. Published in 1990 by
University of Chicago Press. Chicago, IL.

Magilligan, F. J., K. H. Nislov, G. B. Fisher, J. Wright, G. Mackey, and M. Laser 2008. The Geomorphic
Function and Characteristics of Large Woody Debris in Low Gradient Rivers, Coastal Maine, USA.
Geomorphology 97:467–482.Makaske, B., D. G. Smith, and H. J. Berendsen. 2002. Avulsions,
Channel Evolution and Floodplain Sedimentation Rates of the Anastomosing Upper Columbia
River, British Columbia, Canada. Sedimentology 49(5):1049–1071.

Manga, M., and J. W. Kirchner. 2000. Stress Partitioning in Streams by Large Woody Debris. Water
Resources Research 36:2373–2379.

Manners, R. W., M. W. Doyle, and M. J. Small. 2007. Structure and Hydraulics of Natural Woody
Debris Jams. Water Resources Research 43, doi:10.1029/2006WR004910.

Manners, R. B. and Doyle, M. W. 2008. A Mechanistic Model of Woody Debris Jam Evolution and its
Application to Wood-based Restoration and Management. River Research and Applications
24:1104-1123.

Martin, D. J., and L. E. Benda. 2001. Patterns of Instream Wood Recruitment and Transport at the
Watershed Scale. Transactions of the American Fisheries Society 130:940–958.

Maser, C. and J. R. Sedell, J.R. 1994. From the Forest to the Sea: The Ecology of Wood in Streams, Rivers,
Estuaries, and Oceans. St. Lucie Press. 200 pp.

Maser, C., R. F. Tarrant, J. M. Trappe, and J. F. Franklin (eds.). 1988. From the Forest to the Sea: A Story
of Fallen Trees. General Tech. Report PNW-GTR-229. USFS. 153 pp.

Masterman, R., and C. R. Thorne. 1992. Predicting Influence of Bank Vegetation on Channel Capacity.
Journal of Hydraulic Engineering 118:1052–1058.

McDade, M. H., F. J. Swanson, W. A. McKee, J. F. Franklin, and J. Van Sickle. 1990. Source Distances for
Coarse Woody Debris Entering Small Streams in Western Oregon and Washington. Canadian
Journal of Forest Research 20:326–330.

McHenry, M. L., E. Shott, R. H. Conrad, and G. B. Grette. 1998. Changes in the Quantity and
Characteristics of LWD in Streams of the Olympic Peninsula, Washington, USA (1982-1993).
Canadian Journal of Fisheries and Aquatic Sciences 55(6):1395–1407.

Large Wood National Manual July 2015


4-79
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Means, J. E., K. Cromack Jr., and P. C. MacMillan, 1986, Comparison of Decomposition Models Using
Wood Density of Douglas-Fir Logs. Canadian Journal of Forestry Research 15:1092–1098.

Meile, T., J. Boillat, and A. Schleiss. 2011. Flow Resistance Caused by Large-Scale Bank Roughness in
a Channel. Journal of Hydraulic Engineering 137(12):1588–1597.

Melillo, J. M., R. J. Naiman, J. D. Aber, and K. N. Eshleman. 1983. The Influence of Substrate Quality
and Stream Size on Wood Decomposition Dynamics. Oecolgia (Berlin) 58:281–285.

Meyer, J. L., J. B. Wallace, and S. L. Eggert. 1998. Leaf Litter as a Source of Dissolved Organic Carbon
in Streams. Ecosystems 1:240–249.

Micheli, E. R., J. W. Kirchner, and E. W. Larsen. 2003. Quantifying the Effect of Riparian Forest Versus
Agricultural Vegetation on River Meander Migrations Rates, Central Sacramento River,
California, USA. River Research and Applications 19:1–12.

Millward, A. A., C. E. Kraft, and D. R. Warren. 2010. Ice Storm Damage Greater Along the Terrestrial-
Aquatic Interface in Forested Landscapes. Ecosystems 13:249–260.

Minakawa, N., and R. I. Gara. 2005. Spatial and Temporal Distribution of Coho Salmon Carcasses in a
Stream in the Pacific Northwest, USA. Hydrobiologia 539:163–166.

Minore, D. 1979. The Wild Huckleberries of Oregon and Washington: A Dwindling Resource. USDA
Forest Service Research Paper 143.

Montgomery, D. R. 1999. Process Domains and the River Continuum. Journal of the American Water
Resources Association 35:397–410.

Montgomery, D. R., and T. B. Abbe. 2006. Influence of Logjam-Formed Hard Points on the Formation
of Valley-Bottom Landforms in an Old-Growth Forest Valley, Queets River, Washington, USA.
Quaternary Research 65:147–155.

Montgomery, D. R., and J. M. Buffington. 1993. Channel Classification, Prediction of Channel Response,
and Assessment of Channel Condition. TFW-SH10-93-002. Washington State Timber, Fish &
Wildlife.

Montgomery, D. R., and J. M. Buffington. 1997. Channel-Reach Morphology in Mountain Drainage


Basins. Geological Society of America Bulletin 109:596–611.

Montgomery, D. R., J. M. Buffington, R. D. Smith, K. M. Schmidt, and G. Pess. 1995b. Pool Spacing in
Forest Channels. Water Resources Research 31:1097–1105.

Montgomery, D. R., T. Abbe, N. P. Peterson, J. M. Buffington, K. M. Schmidt, and J. D. Stock 1996b.


Distribution of Bedrock and Alluvial Channels in Forested Mountain Drainage Basins. Nature
381:587–589.

Montgomery, D. R., B. D. Collins, J. M. Buffington, and T. B. Abbe. 2003. Geomorphic Effects of Wood
in Rivers. Pages 21–47 in S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology and
Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society.

Morris, A. E. L., P. C. Goebel, and B. J. Palik. 2010. Spatial Distribution of Large Wood Jams in Streams
Related to Stream-Valley Geomorphology and Forest Age in Northern Michigan. River Research
and Applications 26:835–847.

Large Wood National Manual July 2015


4-80
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Moscrip, A. L., and D. R. Montgomery. 1997. Urbanization, Flood Frequency, and Salmon Abundance
in Puget Lowland Streams. Journal of the American Water Resources Association 33(6):1289–
1297.

Mossop, B., and M. J. Bradford. 2004. Importance of Large Woody Debris for Juvenile Chinook
Salmon Habitat in Small Boreal Forest Streams in the Upper Yukon River Basin, Canada.
Canadian Journal of Forest Research-Revue Canadienne De Recherche Forestiere 34:1955–1966.

Moulin, B., E. R. Schenk, and C. R. Hupp. 2011. Distribution and Characterization of In-channel Large
Wood in Relation to Geomorphic Patterns on a Low-gradient River. Earth Surface Processes and
Landforms 36:1137–1151.

Muir, J. 1878. Forests of California, the New Sequoia. Harper’s New Monthly Magazine LVII
(CCCXLII):813–827.

Mulholland, P. J., J. D. Newbold, J. W> Elwood, L. A. Ferren, and J. R. Webster. 1985. Phosphorus
Spiraling in a Woodland Stream - Seasonal-Variations. Ecology 66:1012–1023.

Murphy, M. L. 1995. Forestry Impacts on Freshwater Habitat of Anadromous Salmonids in the


Pacific Northwest and Alaska—Requirements for Protection and Restoration. U.S. Department of
Commerce Coastal Ocean Program, NOAA. Decision Analysis Series No. 7, 156 pp.

Murphy, M. L., and K. V. Koski. 1989. Input and Depletion of Woody Debris in Alaska Streams and
Implications for Streamside Management. North American Journal of Fisheries Management
9(4):427–436.

Mutz, M., E. Kalbus, and S. Meinecke. 2007. Effect of Instream Wood on Vertical Water Flux in Low-
Energy Sand Bed Flume Experiments. Water Resources Research 43:W10424.

Naiman, R. J., T. J. Beechie, L. E. Benda, P. A. Bisson, L. H. MacDonald, M. D. O’Conner, P. L. Olsen, and


E. A. Steel. 1992. Fundamental elements of ecologically healthy watersheds in the Pacific
Northwest coastal ecoregion. Pages 127–188 in R. J. Naiman (ed.), Watershed Management:
Balancing Sustainability and Environmental Change. Springer: New York.

Naiman, R. J., K. L. Fetherston, S. McKay, and J. Chen. 1998. Riparian Forests. Pages 289-323 in R. J.
Naiman and R. E. Bilby (eds.), River Ecology and Management: Lessons from the Pacific Coastal
Ecoregion. Springer-Verlag: New York.

Naiman R. J., S. E. Bunn, and C. Nilsson. 2002b. Legitimizing Fluvial Ecosystems as Users of Water.
Environmental Management 30:455–467.

National Marine Fisheries Service. 1996. Making Endangered Species Act Determinations of Effect for
Individual or Grouped Actions at the Watershed Scale. Environmental and Technical Services
Division, Habitat Conservation Branch.

Nichols, R. A. and S. G. Sprague. 2003. The Use of Long-Line Cabled Logs for Stream Bank
Rehabilitation. Pages 422–442 in D. R. Montgomery, S. M. Bolton, D. B. Booth, and L. Wall (eds.),
Restoration of Puget Sound Rivers. University of Washington Press: Seattle.

Nickelson, T. E., M. F. Solazzi, S. L. Johnson, and J. D. Rodgers. 1992. Effectiveness of Selected Stream
Improvement Techniques to Created Suitable Summer and Winter Rearing Habitat for Juvenile
Coho Salmon (Oncorhynchus kisutch) in Oregon Coastal Streams. Canadian Journal of Fisheries
and Aquatic Sciences 49:790–794.

Large Wood National Manual July 2015


4-81
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

North American Forest Commission. 2011. Forests of North America. Vector Digital Data. Food and
Agriculture Organization of the United Nations. Commission for Environmental Cooperation.
Montreal, Quebec, CA.

O’Connor, J. E., M. A. Jones, and T. L. Haluska. 2003. Flood Plain and Channel Dynamics of the
Quinault and Queets Rivers, Washington, U.S.A. Geomorphology 51:31–59.

Oliver, C. D. 1980/1981. Forest Development in North America Following Major Disturbances. Forest
Ecology and Management 3:153–168.

Oregon Department of Forestry. 1995. A Guide to Placing Large Wood in Streams. Salem, OR, Forest
Practices Section. 13 pp.

Palik, B., S. W. Golladay, P. C. Goebel, and B. W. Taylor. 1998. Geomorphic Variation in Riparian Tree
Mortality and Stream Coarse Woody Debris Recruitment from Record Flooding in a Coastal Plain
Stream. Ecoscience 5:551–560.Pariset, E., R. Hausser, and A. Gagnon. 1966. Formation of Ice
Covers and Ice Jams in Rivers. Journal of the Hydraulics Division 92(6):1–24.

Parrish, R. M. and P. B. Jenkins. 2012. Natural Log Jams in the White River: Lessons for Geomimetic
Design of Engineered Log Jams. U.S. Fish and Wildlife Service, Leavenworth, WA.

Paukert, C. P., and A. S. Makinster. 2008. Longitudinal Patterns in Flathead Catfish Relative
Abundance and Length at Age Within a Large River: effects of an urban gradient. River Research
and Applications. Available: www.interscience.wiley.com. DOI: 10.1002/rra.1089.

Petts, G. E., A. L. Roux, and H. Moller (eds.). 1989. Historical Changes of Large Alluvial Rivers, Western
Europe. Chichester: John Wiley.

Phillips, J. D. 2012. Log-jams and Avulsions in the San Antonio River Delta, Texas. Earth Surface
Processes and Landforms 37:936–950.

Phillips, J. D., and L. Park. 2009. Forest Blowdown Impacts of Hurricane Rita on Fluvial Systems.
Earth Surface Processes and Landforms 34:1069–1081.

Piégay, H., A. and R. A. Marston. 1998. Distribution of Coarse Woody Debris Along the Concave Bank
of a Meandering River (the Ain River, France). Physical Geography 19(4):318–340.

Piégay, H., A. Thevenet, and A. Citterio. 1999. Input, Storage and Distribution of LWD Along a
Mountain River Continuum, the Drôme River, France. Catena 35:19–39.

Pollock, M. M., and T. J. Beechie. 2014. Does Riparian Forest Restoration Thinning Enhance
Biodiversity? The Ecological Importance of Large Wood. JAWRA Journal of the American Water
Resources Association 50(3):543–559. Online publication date: June 1, 2014.

Pollock, M. M., R. J. Naiman, and T. A. Hanley. 1998. Plant Species Richness in Riparian Wetlands—A
Test of Biodiversity Theory. Ecology 79:94–105.

Power, M. E., and W. E. Dietrich. 2002. Food Webs in River Networks. Ecological Research 17:451–
471.

Power, M. E., A. Sun, G. Parker, W. E. Dietrich, and J. T. Wootton. 1995. Hydraulic Food Chain Models.
BioScience 45:159–167.

Large Wood National Manual July 2015


4-82
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Prowse, T. D. 2001. River Ice Ecology. 1: Hydrologic, Geomorphic, and Water Quality Aspects. Journal
of Cold Regions Engineering 15(1):1–16.

Quinault Indian Nation (QIN). 2008. Salmon Habitat Restoration Plan for the Upper Quinault River.
Quinault Indian Nation Department of Fisheries. Taholah, Washington. Prepared by T. Abbe and
others.

Ralph, S. C., G. C. Poole, L. L. Conquest, and R. J. Naiman. 1991. Stream Channel Morphology and
Woody Debris in Logged and Unlogged Basins of Western Washington. Canadian Journal of
Fisheries and Aquatic Sciences 51:37–51.

Rapp, C., and T. Abbe. 2003. A Framework for Delineating Channel Migration Zones. Washington State
Department of Ecology Publication Number 03-06-027. Final Draft.

Raup, H. M. 1957. Vegetation Adjustment to the Instability of Sites. Proceedings and Papers of the 6th
Technical Meeting of the International Union for Conservation of Nature and Natural Resources.
Edinburgh. Pages 36–48.

Reeves, G. H., J. D. Hall, T. D. Roelofs, T. L. Hickman, and C. O. Baker. 1991. Rehabilitating and
Modifying Stream Habitats. Pages 519–557 in Influences of Forest and Rangeland Management
on Salmonid Fishes and Their Habitats. American Fisheries Society Special Publication 19.

Reid, L. M., and S. Hilton. 1998. Buffering the Buffer. Pages 71–80 in R. R. Ziemer (ed.), Proceedings of
the Conference on Coastal Watersheds: The Caspar Creek Story; held May 6, 1998, in Ukiah,
California. USDA Forest Service, Pacific Southwest Research Station, General Technical Report
PSW-GTR-168.

Richards, K. 1982. Rivers: Form and Process in Alluvial Channels. New York: Methuen. 382 pp.

Richmond, A. D., and K. D. Faush. 1995. Characteristics and Function of LWD in Subalpine Rocky
Mountains Streams in Northern Colorado. Canadian Journal of Fisheries and Aquatic Sciences
52:1789–1802.

Riley, S. C. and K. D. Fausch. 1995. Trout Population Response to Habitat Enhancement in Six
Northern Colorado Streams. Canadian Journal of Fisheries and Aquatic Sciences. 52:34–53.

Robert, A. 1997. Characteristics of Velocity Profiles Along Riffle-Pool Sequences and Estimates of
Bed Shear Stresses. Geomorphology 19:89–98.

Robison, E. G. and R. L. Beschta. 1990. Identifying Trees in Riparian Areas that can Provide Coarse
Woody Debris to Streams. Forest Science 36:790–801.

Roni, P., and T. P. Quinn. 2001. Density and Size of Juvenile Salmonids in Response to Placement of
Large Woody Debris in Western Oregon and Washington Streams. Canadian Journal of Fisheries
and Aquatic Sciences 58:282–292.

Roni, P., M. Liermann, and A. Steel. 2003. Monitoring and Evaluating Fish Response to Instream
Restoration. In D. Montgomery, S. Bolton, D. Booth, and L. Wall (eds.), Restoration of Puget Sound
Rivers. Center for Water and Watershed Studies. University of Washington Press: Seattle.

Roni, P., T. Beechie, G. Pess, and K. Hanson. 2014a. Wood Placement in River Restoration: Fact,
Fiction, and Future Direction. Canadian Journal of Fisheries and Aquatic Sciences 72(3):466–478.

Rosgen, D., and H. L. Silvey. 1996. Applied River Morphology. Wildland Hydrology. Pagosa Springs, CO.

Large Wood National Manual July 2015


4-83
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Rot, B. W., R. J. Naiman, and R. E. Bilby. 2000. Stream Channel Configuration, Landform, and Riparian
Forest Structure in the Cascade Mountains, Washington. Canadian Journal of Fisheries and
Aquatic Sciences 57:699–707.

Ruffner, E. H. 1886. The Practice of the Improvement of the Non-Tidal Rivers of the United States, with
an Examination of the Results Thereof. New York, NY: John Wiley and Sons.

Russell, I. C. 1898. Rivers of North America. New York: G.P. Putnams Sons. 327 pp.

Rutherford, I., B. Anderson, and A. Ladson. 2007. Managing the Effects of Riparian Vegetation on
Flooding. In S. Lovett and P. Price (eds.), Principles for Riparian Lands Management. Land &
Water Australia, Canberra.

Sauer, V. B. 1974. Flood Characteristics of Oklahoma Streams Techniques for Calculating Magnitude
and Frequency of Floods in Oklahoma, with Compilations of Flood Data Through 1971. U.S.
Geological Survey Water-Resources Investigations Report 73–52. 307 p.

Sauer, V. B., and D. P. Turnipseed. 2010. Stage Measurement at Gaging Stations: U.S. Geological Survey
Techniques and Methods Book 3, Chapter A7.

Sear, D. A., C. E. Millington, D. R. Kitts, and R. Jeffries. 2010. Logjam Controls on Channel:Floodplain
Interactions in Wooded Catchments and Their Role in the Formation of Multi-Channel Patterns.
Geomorphology 116:305–319.

Schenk, E. R., J. W. McCargo, B. Moulin, C. R. Hupp, and J. M. Richter. 2014a. The Influence of Logjams
on Largemouth Bass (Micropterus salmoides) Concentrations on the Lower Roanoke River, a
Large Sand-bed River. River Research and Applications. www.wileyonlinelibrary.com, DOI:
10.1002/rra.2779

Schenk, E. R., B. Moulin, C. R. Hupp, J. M. Richter. 2014b. Large Wood Budget and Transport
Dynamics on a Large River Using Radio Telemetry. Earth Surface Processes and Landforms
39:487–498.

Scherer, R. 2004. Decomposition and Longevity of In-Stream Woody Debris: A Review of Literature
from North America. Pages 127–133 in Forest Land–Fish Conference–Ecosystem Stewardship
through Collaboration. Proceedings of Forest-Land-Fish Conference II.

Schumm, S. A. 1999. Causes and Controls of Channel Incision. Pages 19–34 in S. E. Darby and A.
Simon (eds.), Incised River Channels. Chichester, UK: Wiley.

Schumm, S. A., M. D. Harvey, and C. C. Watson. 1984. Incised Channels: Morphology, Dynamics and
Control. Water Resources Publication. Littleton, CO.

Sedell, J. R., and J. L. Frogatt. 1984. Importance of Streamside Forests to Large Rivers: The Isolation
of the Willamette River, Oregon, U.S.A., from its Floodplain by Snagging and Streamside Forest
Removal. Verhandlungen-Internationale Vereinigung für Theorelifche und Angewandte
Limnologie 22:1828–1834.

Sedell, J. R., and K. J. Luchessa. 1981. Using the Historical Record as an Aid to Salmonid Habitat
Enhancement. Symposium on Acquisition and Utilization of Aquatic Habitat Inventory
Information. October 23–28, Portland, OR.

Large Wood National Manual July 2015


4-84
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Sedell, J. R., F. H. Everest, and F. J. Swanson. 1982. Fish Habitat and Streamside Management: Past
and Present. Pages 244–255 in Proceedings of the 1981 Convention of the Society of American
Foresters, September 27–30, 1981. Society of American Foresters, Publication 82–01, Bethesda,
Maryland.

Shields, F. D., Jr., and C. V. Alonso. 2012. Assessment of Flow Forces on Large Wood in Rivers. Water
Resources Research 48(4):W04156.

Shields, F. D., Jr., and C. J. Gippel. 1995. Prediction of Effects of Woody Debris Removal on Flow
Resistance. Journal of Hydraulic Engineering 121 (4):341–354.

Shields, F. D., Jr., and R. H. Smith. 1992. Effects of Large Woody Debris Removal on Physical
Characteristics of a Sand Bedded River. Aquatic Conservation: Marine and Freshwater Ecosystems
2:145–163.

Sidle, R. C. 1991. A Conceptual Model of Changes in Root Cohesion in Response to Vegetation


Management. Journal of Environmental Quality 20:43–52.

Simon, A. 1989. A Model of Channel Response in Disturbed Alluvial Channels. Earth Surface
Processes and Landforms 14:11–26.

Simon, A. 1994. Gradation Processes and Channel Evolution in Modified West Tennessee Streams:
Process, Response and Form. U.S. Geological Survey Professional Paper 1470. Washington D.C.

Simon, A., and A. J. C. Collison. 2002. Quantifying the Mechanical and Hydrological Effects of Riparian
Vegetation on Stream-Bank Stability. Earth Surface Processes and Landforms 27(5):527–546.

Simon, A., and M. Rinaldi. 2006. Disturbance, Stream Incision, and Channel Evolution: The Roles of
Excess Transport Capacity and Boundary Materials in Controlling Channel Response.
Geomorphology 79:361–383.

Simon, A., A. Curini, S. E. Darby, and E. J. Langendoen. 2000. Bank and Near-Bank Processes in an
Incised Channel. Geomorphology 35(3):193–217.

Singer, S., and M. L. Swanson. 1983. The Soquel Creek Storm Damage Recovery Plan with
Recommendations for Reduction of Geologic Hazards in Soquel Village, Santa Cruz County,
California. Unpublished USDA Soil Conservation Service report to the Santa Cruz County Board
of Supervisors.

Sinsabaugh, R. L., M. P. Osgood, and S. Findlay. 1994. Enzymatic Models for Estimating
Decomposition Rates of Particulate Detritus. Journal of the North American Benthological Society.
13:160–169.

Smith, D. G. 1979. Effects of Channel Enlargement by River Ice Processes on Bankfull Discharge in
Alberta, Canada. Water Resources Research, 15(2):469–475.

Smith, D. G., and C.M. Pearce. 2000. River Ice and its Role in Limiting Woodland Development on a
Sandy Braid-Plain, Milk River, Montana. Wetlands, 20(2):232–250.

Smith, D. G., and D. M. Reynolds. 1983. Tree Scars to Determine the Frequency and Stage of High
Magnitude River Ice Drives and Jams, Red Deer, Alberta. Canadian Water Resources Journal
8(3):77–94.

Large Wood National Manual July 2015


4-85
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Smith, J. D.. 2004. The Role of Riparian Shrubs in Preventing Floodplain Unraveling Along the Clark
Fork of the Columbia River in the Deer Lodge Valley, Montana. Pages 71–85 in S. J. Bennett. and
A. Simon (eds.), Riparian Vegetation and Fluvial Geomorphology, Water Science and Application 8.
American Geophysical Union, Washington, D.C.

Solazzi, M. F., T. E. Nickelson, S. L. Johnson, and J. D. Rodgers. 2000. Effects of Increasing Winter
Rearing Habitat on Abundance of Salmonids in Two Coastal Oregon Streams. Canadian Journal of
Fisheries and Aquatic Sciences 57:906–914.

Sollins, P., S. P. Cline, T. Verhoeven, D. Sachs, and G. Spycher. 1987. Patterns of Log Decay in Old-
Growth Douglas-Fir Forests, Canadian Journal of Forest Research 17:1585–1595.

Spänhoff, B., C. Alecke, and E. Irmgard Meyer. 2001. Simple Method for Rating the Decay Stages of
Submerged Woody Debris. Journal of the North American Benthological Society 20(3):385–394.

Spies, T. A., and J. F. Franklin. 1991. The Structure of Natural Young, Mature, and Old-Growth
Douglas Fir Forests in Oregon and Washington. Pages 91–109 in L. F. Ruggiero, K. B. Aubrey, A.
B. Carey, and M. H. Huff (technical coordinators), Wildlife and Vegetation of Unmanaged Douglas
Fir Forests. USDA Forest Service. General Technical Report PNW-GTR-285.

Stahle, D. W., M. K. Cleaveland, R. D. Griffin, M. D. Spond, F. K. Fye, R. B. Culpepper, and D. Patton.


2006. Decadal Drought Effects on Endangered Woodpecker Habitat. Eos, Transactions American
Geophysical Union 87(12):121–125.

Stewart, T. L., and J. F. Martin. 2005. Energy Model to Predict Suspended Load Deposition Induced by
Woody Debris: Case Study. Journal of Hydraulic Engineering-ASCE 131:1011–1016.

Stock, J. D., D. R. Montgomery, B. D. Collins, W. E. Dietrich, and L. Sklar. 2005. Field Measurements of
Incision Rates Following Bedrock Exposure: Implications for Process Controls on the Long
Profiles of Valleys Cut by Rivers and Debris Flows. Geological Society of America Bulletin
117(11/12):174–194.

Subramanya, K., 2008. Engineering Hydrology. New York: McGraw-Hill. 434 pp.

Swanson, F. J., S. V. Gregory, J. R. Sedell, and A. G. Campbell. 1982. Land-Water Interactions: The
Riparian Zone. Pages 267–291 on R. L. Edmonds (ed.), Analysis of Coniferous Forest Ecosystems in
the Western United States. US/IBP Synthesis Series, Hutchinson Ross Publishing Company:
Stroudsburg, PA.

Swanson, F. J., T. K. Kranz, N. Caine, and R. G. Woodmansee. 1988. Landform Effects on Ecosystem
Patterns and Processes. BioScience 38:92–98.

Tappeiner, J. C., D. Huffman, D. Marshall, T. A. Spies, and J. D. Bailey. 1997. Density, Ages, and Growth
Rates in Old-Growth and Young-Growth Forests in Coastal Oregon. Paper 3166 of the Forest
Research Laboratory, Oregon State University, Corvallis.

Thomas, H., and T. R. Nisbet. 2006. An Assessment of the Impact of Flood Plain Woodland on Flood
Flows. Water and the Environment Journal 21(2):114–126.

Tarzwell, C. M. 1936. Experimental Evidence of the Value of Trout Stream Improvements.


Transactions of the American Fisheries Society 66:177–187.

Large Wood National Manual July 2015


4-86
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Thompson, D. M. 2002. Long-term Effect of Instream Habitat-improvement Structures on Channel


Morphology along the Blackledge and Salmon Rivers, Connecticut, USA. Environmental
Management 29(1):250–265.

Thompson, D. M. 2005. The History of the Use and Effectiveness of Instream Structures in the United
States. Geological Society of America Reviews in Engineering Geology XVI:35–50.

Thorne, S. D., and D. J. Furbish. 1995. Influences of Coarse Bank Roughness on Flow Within a Sharply
Curved River Bend. Geomorphology 12(3):241–257.

Triska, F. J. 1984. Role of Large Wood in Modifying Channel Morphology and Riparian Areas of a
Large Lowland River under Pristine Conditions: A Historical Case Study. Verhandlungen-
InternationaleVereinigung für Theorelifche und Angewandte Limnologie 22:1876–1892.

Tsukamoto, Y. 1987. Evaluation of the Effect of Tree Roots on Slope Stability. Bulletin of the
Experimental Forests. 23:65–124.

Turnipseed, D. P., and V. B. Sauer. 2010. Discharge Measurements at Gaging Stations: U.S. Geological
Survey Techniques and Methods Book 3, Chapter A8, U.S. Geological Survey.

U.S. Bureau of Reclamation. 2005. Watershed Conditions and Seasonal Variability for Select Streams
within WRIA 20, Olympic Peninsula, Washington. Available:
http://www.ecy.wa.gov/programs/eap/wrias/planning/docs/opendraft_wria20_final4.pdf.

Valett, H. M., C. L. Crenshaw, and P. F. Wagner. 2002. Stream Nutrient Uptake, Forest Succession, and
Biogeochemical Theory. Ecology 83:2888–2901.Van Cleef, J. S. 1885. How to Restore Our Trout
Streams. Transactions of the American Fisheries Society 14:50–55.

Van Sickle, J., and S. V. Gregory. 1990. Modeling Inputs of Large Woody Debris to Streams from
Falling Trees. Canadian Journal of Forest Research 20(10):1593–1601.

Veatch, A. C. 1906. Geology and Underground Water Resources of Northern Louisiana and Southern
Arkansas. Washington D.C. United States Geological Survey Professional Paper 46.

Veilleux, A. G., T. A. Cohn, K. M. Flynn, R. R. Mason, and P. R. Hummel, P.R. 2013. Fact Sheet 2013-
3108: Estimating Magnitude and Frequency of Floods Using the PeakFQ 7.0 Program. 2327-6932,
U.S. Geological Survey.

Viessman, W. J., and G. L. Lewis. 2003. Introduction to Hydrology. Prentice Hall. 612 pp.Wadsworth,
A. H., Jr. 1966. Historical Deltation of the Colorado River, Texas. Pages 99–105 in Deltas in Their
Geologic Framework. American Association of Petroleum Geologists.

Wallace, J. B., and A .C. Benke. 1984. Quantification of Wood Habitat in Subtropical Coastal Plain
Streams. Canadian Journal of Fisheries and Aquatic Sciences 41:1643–1652.

Wallerstein, N. P., and C. R. Thorne. 2004. Influence of Large Woody Debris on Morphological
Evolution of Incised, Sand-Bed Channels. Geomorphology 57:53–73.

Washington Forest Practices Board. 1997. Board Manual: Standard Methodology for Conducting
Watershed Analysis. Under Chapter 222-22 WAC. Version 4.0. Olympia, Washington.Webster, J.
R., and 9 others. 2000. Effects of Litter Exclusion and Wood Removal on Phosphorus and
Nitrogen Retention in a Forest Stream. Verhandlungen der Internationale Vereinigung für
Limnologie 27:1337–1340.

Large Wood National Manual July 2015


4-87
Bureau of Reclamation and Chapter 4. Geomorphology and Hydrology
U.S. Army Corps of Engineers Considerations

Webster, J. R., J. A. Stanford, J. L. Chaffin, and Field Ecology Class. 2002. Large Wood Jam in a Fourth
Order Rocky Mountain Stream. Verhandlungen der Internationale Vereinigung für Limnologie
28:1–4.

Western Wood Products Association (WWPA). 1995. Ponderosa Pine Species Facts. Available:
www.wwpa.org/ppine.htm.

Whitney, G. G. 1996. From Coastal Wilderness to Fruited Plain: A History of Environmental Change in
Temperate North America from 1500 to the Present. Cambridge University Press: Cambridge, UK.

Wiegner, T. N., L. A. Kaplan, J. D. Newbold, and P. H. Ostrom. 2005. Contribution of Dissolved Organic
C to Stream Metabolism: A Mesocosm Study Using C-13-Enriched Tree-Tissue Leachate. Journal
of the North American Benthological Society 24:48–67.

Wilcock, P. R., A. F. Barta, C. C. Shea, G. M. Kondolf, W. V. Graham Matthew, and J. Pitlick. 1996.
Observations of Flow and Sediment Entrainment on a Large Gravel-Bed River. Water Resources
Research 32:2897–2909.

Wilford, D., Maloney, D., Schwab, J., and Geertsema, M. 1998. Tributary Alluvial Fans. B.C. Ministry of
Forests Extension Note 30.

Wiltshire, P. E. J., and P. D. Moore. 1983, Paleovegetation and Paleohydrology in Upland Britain.
Pages 433–451 in K. J. Gregory (ed.), Background to Paleohydrology. John Wiley: Chichester, UK.

Wohl, E. E. 2001. Virtual Rivers: Lessons from the Mountain Rivers of the Colorado Front Range. New
Haven, CT: Yale University Press.

Wohl, E. 2013. Floodplains and Wood. Earth-Science Reviews 123:194–212.

Wohl E., D. A. Cenderelli, K. A. Dwire, S. E. Ryan-Burkett, M. K. Young, and K. D. Fausch. 2010. Large
in-Stream Wood Studies: A Call for Common Metrics. Earth Surface Processes and Landforms
35:618–625.

Wohl, E., L. E. Polvi, and D. Cadol. 2011. Wood Distribution Along Streams Draining Old-Growth
Forests in Congaree National Park, South Carolina, USA. Geomorphology 126:108–120.

Wolff, H. H. 1916. The Design of a Drift Barrier Across the White River, near Auburn, Washington.
Transactions of the American Society of Civil Engineers 16:2061–2085.

Zeng, H., J. Q. Chambers, R. I. Negron-Juarez, G. C. Hurtt, D. B. Baker, and M. D Powell. 2009. Impacts
of Tropical Cyclones on U.S. Forest Tree Mortality and Carbon Flux from 1851 to 2000.
Proceedings of the National Academy of Sciences 106(19), 7888–7892.

Zobel, D. B., A. McKee, G. M. Hawk, and C. T. Dyrness. 1976. Relationships of Environment to


Composition, Structure, and Diversity of Forest Communities of the Central Western Cascades of
Oregon. Ecological Monographs 46:135–156.

Large Wood National Manual July 2015


4-88
Chapter 5
WATERSHED-SCALE AND LONG-TERM CONSIDERATIONS

Photo credit: Ken DeCamp

AUTHORS

Jock Conyngham (Environmental Laboratory, ERDC, USACE)


Judsen Bruzgul (ICF International)
Jim MacBroom (Milone & MacBroom, Inc.)
Rebecca Manners (University of Montana)
Roy Schiff (Milone & MacBroom, Inc.)
Ellen Wohl (Colorado State University)
Katy Maher (ICF International)
This page intentionally left blank.
fragmentation of supply and transport processes
5.1 Introduction and by transportation infrastructure, channel
armoring, leveeing, and dam construction.
Purpose Alteration to wood supply has been particularly
River restoration has received valid criticism for acute for large trees that form the key elements
focusing on active, engineered, and structure- that create relatively stable features (depending
driven approaches (Bernhardt et al. 2007; on channel dimensions relative to wood
Kondolf 2000) that incur high financial costs, dimensions). In many areas of the country wood
significant risks of failure, and uncertain biotic supply, both in terms of overall volume and,
responses relative to more passive or assisted increasingly, recruitment of large, key-sized
restoration methods. The relatively high cost material, is only now becoming available to the
profile alone of these active techniques means fluvial system at landscape and sub-landscape
that they cannot address the high degree and scales, and restoration needs to focus
wide distribution of riverine alteration and increasingly on the ability of stream crossings
damage in the United States and elsewhere. As and related infrastructure to convey that
true in most forms of restoration, but even more material. However, an approximation of
so because of the normally shorter lifespans of normative supply will not occur without a
placed wood (compared to stone) and its geographically, economically, and politically
interactions with mobile wood elements, large broader acceptance of the value of riparian and
wood-based projects must anticipate various stream corridors as supply zones for wood,
forms of and alterations to background load zones of desirable physical dynamism (as
inputs. While weighing or integrating active, currently documented by geomorphic channel
assisted, or passive techniques can be migration and geotechnical hazard zone
challenging, it is particularly important with delineation (e.g. FEMA 1999), and areas for flood
wood-based projects. storage, recreation, and concentrated ecosystem
services and values.
Addressing these points requires focusing on
large-scale and long-term issues with wood
supply and dynamics, including: basin-scale 5.3 Flood Dynamics and
large wood recruitment and supply issues, wood Response
management at reservoirs and re-operations for
wood recruitment and routing, effects of climate
change, effects of stochastic flooding and storms
5.3.1 Pulsed Stochastic
on pulsed colluvial and alluvial recruitment of Inputs as a Large Wood
wood, planning and infrastructure design for Recruitment
large wood conveyance during peak flows, and
large wood management and utilization in flood Mechanism
response. Recruitment of wood to channels and floodplains
is highly variable through time and space within
5.2 Corridor and Basin an individual drainage basin and between
drainage basins. Forest dynamics, hillslope
Management dynamics, river-network dynamics, biota, and
Concepts channel dynamics interact to govern
mechanisms, rates, and quantities of wood
Much of the need for this manual stems from the recruitment (Figure 5-1). Forest dynamics
truncation of wood supply to U.S. rivers by large- include individual tree mortality and mass
scale forest clearing and development. There has mortality caused by fires, insects, and blow
also been elimination, reduction, or

Large Wood National Manual July 2015


5-1
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

downs (Gregory et al. 1993; Marcus et al. 2011; primarily to beavers, which are capable of
Wohl 2013a). Hillslope dynamics primarily recruiting wood to streams by chewing down
refers to slope instability in the form of trees (Kreutzweiser et al. 2005). Channel
avalanches, landslides, and debris flows that dynamics includes bank erosion that undermines
introduce wood to valley bottoms (May and and recruits trees, and floodplain erosion that
Gresswell 2003a, b; Comiti et al. 2006; Wohl et exhumes buried wood and returns it to the
al. 2009; Rigon et al. 2012). River-network active channel (Downs and Simon 2001; Kukulak
dynamics describes tributary inputs of wood to a et al. 2002; Wyżga and Zawiejska 2005; Guyette
main valley (Benda et al. 2003a). Biota refers et al. 2008).

Figure 5-1. Influences on Wood Recruitment to River Corridors

Inset photographs illustrate (clockwise starting from upper right) beaver-felled trees in Colorado; a logjam at
the mouth of a tributary along the Upper Rio Chagres, Panama; trees leaning over the Snake River in
Wyoming as a result of bank erosion; abundant downed wood and secondary channels in Colorado; a small
landslide along the Upper Rio Chagres; and standing dead trees following a forest fire in Colorado.
(Photographs by Ellen Wohl)

Large Wood National Manual July 2015


5-2
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

CASE STUDIES

Episodic Mass Large Wood Recruitment

Blowdown along Glacier Creek, Colorado (Wohl 2013a): On November 21, 2011, a microburst knocked down trees
over a 33-hectare (82-acre) area along Glacier Creek in Rocky Mountain National Park with old-growth subalpine
spruce-fir forest. Many of the trees remained partly attached to the bank via rootwads partly anchored in the soil.
Other trees formed a bridge, with the trunk above peak-flow water levels but large branches oriented down into the
channel. During the next 2 years, these relatively stable downed trees acted as key pieces for new logjams by
trapping smaller wood in transport down the creek. Jam frequency along this portion of Glacier Creek increased from
approximately 1 jam per 100 meters (330 feet) to 1 jam per 54 meters (177 feet) by July 2013. The ratio of tree
length (averaging 16 meters [53 feet]) to channel width (averaging 12 meters [39 feet]) allowed downed trees to
effectively block a significant portion of the channel and form in situ log jams.

Trees blown down across Glacier


Creek (at left). Small landslide along
tributary to Rio Chagres (at right)
has created a large jam at the
tributary junction. The mainstem
flow is right to left. Person at upper
left of jam for scale.

Landsliding in the Upper Rio Chagres, Panama (Wohl et al. 2009): The Upper Rio Chagres drains 414 square
kilometers (160 square miles) of mountainous terrain covered by old-growth rainforest in central Panama. An
intense convective storm on July 10, 2007, created widespread rainfall over the basin that triggered flooding and
numerous landslides. Transport capacity is very high within the Chagres catchment, where peak unit discharge can
reach 41 cubic meters per second per square kilometer. However, landslides introduced such large masses of wood
that enormous logjams formed at sites of reduced transport capacity such as tributary junctions or bends on the
main channel. Trees in the watershed can attain a height of 30 meters (90 feet) and a diameter of 2.2 meters (7.2
feet), and key pieces in these jams were greater than 20 meters (66 feet) in length. The ratio of piece length to
channel width averaged 0.1–0.2 at sites of jam formation. Although some of the jams stored substantial volumes of
sediment upstream (1,100–8,200 cubic meters [1,440–10,725 cubic yards]), the jams broke apart and disappeared
within 3 years due to the combined effects of subsequent high flows and extremely fast rates of wood decay.

Low frequency, episodic processes such as fire, (Montgomery 1999). Ecologists define a
landslides, or channel avulsion, in particular, can disturbance as any relatively discrete event in
create relatively large inputs of sediment and time that disrupts ecosystem, community, or
wood to river networks. Understanding of population structure and changes resources,
geomorphic process domain can provide substrate availability, or the physical
understanding of where within a landscape these environment (White and Pickett 1985). A flood is
processes are likely to occur, as well as the an obvious example of a disturbance in a river
magnitude and frequency of associated wood environment. Disturbance regime refers to the
recruitment. Geomorphic process domains are spatial pattern and statistical distribution of
spatially distinct portions of the landscape that disturbances in terms of magnitude, frequency,
reflect spatial variability in geomorphic and duration of associated changes in the
processes and temporal patterns of disturbances physical environment (Montgomery 1999).
that influence ecosystem structure and dynamics

Large Wood National Manual July 2015


5-3
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

5.3.1.1 Retention of Pulsed Large segments may retain less wood than otherwise
Wood Inputs comparable wider, lower gradient segments
(Wohl 2011b; Wohl and Cadol 2011). The
Large, pulsed inputs of sediment and wood to a presence of a floodplain in a wider valley
channel can remain in place, translate segment facilitates overbank flows that limit
downstream as a relatively discrete mass, or increases in depth and velocity during higher
diffuse along a greater channel length with time, discharges. Shallower, slower overbank flows
although studies published thus far focus on can also carry wood onto the floodplain,
downstream movement of sediment pulses increasing wood retention within the valley
rather than wood (e.g., Lisle et al. 2001; Sklar et segment.
al. 2009). The retention of wood following
individual or mass recruitment also varies Existing wood loads at the time of pulsed
through time and space as a function of valley recruitment are important because they can
geometry, existing wood loads at the time of create congestion within the channel and
recruitment, discharges of water and sediment, floodplain, forming obstacles in the form of
and channel characteristics. immobile ramped pieces or logjams that trap
wood in transport (Bocchiola et al. 2006; Moulin
The most relevant aspects of valley geometry are et al. 2011; Wohl and Beckman 2014a). Even if
the ratio of active channel width to valley- all wood pieces are mobile, the volume of wood
bottom width and the channel gradient. These in transport relative to channel dimensions can
typically correlate: a valley bottom that is much create different modes of transport. Braudrick et
wider than the active channel commonly has a al. (1997) observed congested, semi-congested,
relatively low gradient, whereas steeper valley and uncongested wood transport during physical
segments have narrowly confined active experiments in a flume. During uncongested
channels. Steep, narrow channels may have transport, logs move without piece-to-piece
limited transport capacity because of large ratios interactions, whereas logs interact with one
of wood piece length to channel width. Physical another during semi-congested transport and
experiments in flumes (Braudrick and Grant move as a single mass during congested
2000; Welber et al. 2013) and field studies (Haga transport. Wood can move farther during
et al. 2002; Warren and Kraft 2008; Merten et al. congested transport (Bocchiola et al. 2008).
2010) indicate that wood transport scales with During a 10-year study of the mobility of
the ratios of piece length/channel width and individual wood pieces within mountainous
piece diameter/flow depth. As these ratios channels in the Colorado Front Range, Wohl and
increase, wood mobility declines. Depth Goode (2008) found that pieces within jams
increases rapidly with discharge in narrow typically had longer residence times than
reaches, however, as do hydraulic forces acting isolated pieces.
on wood, so that narrower, steeper valley

Large Wood National Manual July 2015


5-4
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

CASE STUDIES

Case Studies of Spatial Distribution of Large Wood

Longitudinal Segregation and Aggregation of Large Wood (Kraft and Warren 2003; Morris et al. 2010): Two years
after extensive wood deposition from an ice storm in the eastern Adirondack Mountains of New York, Neighbor K
statistics Case Studies
indicated of Spatial
that individual Distribution
pieces of wood wereof LW
aggregated at spatial extents ranging from 0 to 40 meters
(0 to 131 Longitudinal
feet) and were segregatedand
Segregation (regularly spaced)ofatLW
Aggregation distances ranging
(Kraft and from2003;
Warren, 100 to 300 meters
Morris (328 toTwo
et al., 2010): 984years after extensive
feet) alongwood deposition
channels from
draining 6 toan icesquare
130 storm kilometers
in the eastern
(2 toAdirondack Mountains
50 square miles). Meanofchannel
New York, Neighbor
widths variedKfrom
statistics indicated that
individual pieces of wood were aggregated at spatial extents ranging from 0 to 40 m and were
4 to 13 meters (13 to 43 feet). Spatial segregation of jams occurred in response to stream features that created segregated (regularly space
at distances ranging
stable accumulation points. from 100 to 300 m along channels draining 6 to 130 km2. Mean channel widths varied from 5 to 13 m
Spatial segregation of jams occurred in response to stream features that created stable accumulation points.
700
alternating anastomosing single channel
fan anastomosing occasional anastomosing
& small gorges avg 85 avg 13
avg 405 avg 192
600 avg 79

500
Wood load (m3/ha)

North St. Vrain Creek


400

300

gorge
200
avg 49

100

0
0 500 1000 1500 2000 2500 3000 3500 4000 4500 5000 5500 6000 6500 7000 7500 8000 8500 9000

16 Distance downstream (m)


avg 0.2
North St. Vrain jams avg 1.4
14
avg 0.8
North St. North
Vrain Creek,
St. Vrain Colorado
Creek,(WohlColoradoand (Wohl
Cadol 2011):
and Cadol, A longitudinally continuous survey
2011): A longitudinally of instream
continuous wood
survey of instream wood distribution
Volume of wood in jam (m3)

12
along
distribution along9 km of North
9 kilometersavg 1.5 St. Vrain
(5.6 miles)Creek
of included
North St. diverse
Vrain stand
Creek ages
included of riparian
diverse forest
stand and
ages ofvalley geometry.
riparian forest The surveyed portio
of the
10
creek drains 15-82 km 2. Channel width varies from 7 to 20 m and tree heights are typically < 20 m. Individual wood
and valley geometry. The surveyed portion of the creek drains 15 to 82 square kilometers (6 to 32 square miles).
pieces
Channel width 8 are highly
varies from 7aggregated
avg 1.1
to 20 meters at length
(23 to 66scales
feet)ofand1 totree
150heights
m. Localarevalley and less
typically channel
thangeometry
20 metersexert
(66 a stronger influence o
longitudinal patterns of wood distribution than either time since last forest
feet). Individual6 wood pieces are highly aggregated at length scales of 1 to 150 meters (3.3 to 492 feet). Local disturbance or progressive downstream trends
associated with increasing drainage area. Wood loads and average jam size are greater in valley segments with greater
valley and channel
4 geometry exert a stronger influence on longitudinal patterns of wood distribution than since
valley-bottom
avg 0.3
width and lower gradient. Red highlights in figure above indicate relatively steep, narrow portions of the
either the last forest disturbance or progressive downstream trends associated with increasing drainage area.
channel, which have lower wood loads.
2

Wood loads and average jam size are greater in valley segments with greater valley-bottom width and lower
0
gradient. Red highlights
0 500 in 1000the1500
figure
2000 above indicate
2500 3000 relatively
3500 4000 4500 5000steep, narrow
5500 6000 6500 portions
7000 7500 of the8500
8000 channel,
9000 which have
Cumulative distance downstream (m)
lower wood loads.

Sediment discharge influences wood retention 2004). Weakened wood pieces can be broken by
and redistribution along a channel by influencing hydraulic forces or by the impact of very coarse
processes such as abrasion, breakage, and burial sediment, although, again, no data are available
of wood pieces (Webb and Erskine 2003; Young on rates or magnitudes of these processes. Burial
et al. 2006; Merten et al. 2013). Although few can protect wood from hydraulic forces and
data exist for abrasion rates for instream wood, abrasion, but can also promote wood decay if the
observations suggest that abrasion can wood is buried in anoxic conditions.
significantly erode logs close to the streambed in
Water discharge influences wood retention and
channels with high rates of flux for sand-sized
redistribution by creating hydraulic forces that
and coarser particles (Spänhoff and Meyer

Large Wood National Manual July 2015


5-5
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

can mobilize wood, as well as sufficient transport consistent wood load through time and minimal
capacity to keep the wood moving. Discharge- development of logjams, and an episodic end-
stage relationships within a channel are likely to member in which episodic mass recruitment via
create numerous thresholds at which differently landslides or blowdowns results in formation of
sized and oriented pieces of wood are mobilized transient logjams, so that wood loads are highly
(MacVicar and Piégay 2012; Kramer and Wohl spatially and temporally variable.
2014; Schenk et al. 2014a), although few studies
In summary, the ability of any river segment to
address this phenomenon.
retain pulsed stochastic inputs of large wood
Channel characteristics that influence wood reflects the pre-event wood load, the presence of
mobility include what Braudrick and Grant channel and valley features that can enhance
(2001) referred to as debris roughness. A rough wood storage (e.g., large protruding clasts,
channel, in this context, is one with at least some meander bends, abrupt expansions or
large clasts that protrude well above the bed and constrictions, floodplains), and the sequence of
in some cases above the water surface. These water and sediment fluxes following large wood
clasts can effectively trap wood in transport recruitment. Marcus et al. (2002) described river
(Figure 5-2). Bends in the channel and segments as being either supply-limited or
downstream variations in channel width can also transport-limited with respect to wood. Wohl
enhance wood retention. Wood can be and Jaeger (2009) built on this idea to develop a
preferentially deposited either on point bars at conceptual model of wood distribution
the inside of bends (Daniels and Rhoads 2004) throughout a network. Lower order streams are
or along the top of the bank/edge of the transport-limited for wood and have high loads
floodplain along the outside of bends (Piégay of randomly distributed wood pieces (Figure
1993; Johnson et al. 2000) (Figure 5-3). 5-4). Moderate order streams have sufficient
transport capacity to move wood into jams that
The few stochastic models of instream wood
form at sites with local limitations on transport
loads through time reflect the influences on large
capacity, such as abrupt channel expansions or
wood retention of these disparate mechanisms
bends. Higher order streams are supply-limited
(e.g., Meleason et al. 2007). The model of Eaton
for wood and have lower wood loads. The
et al. (2012), for example, includes parameters
specific portions of a network that fit into these
for wood piece dimension, channel and flow
three general categories will depend on factors
dimensions, wood load, and wood decay and
such as peak discharge per unit drainage area,
breakage. Conceptual models of instream wood
rates of downstream increase in channel width,
loads can also implicitly incorporate factors that
and the size and abundance of wood pieces
influence recruitment and retention. Based on
recruited to the channel. Subsequent research
four field sites in Costa Rica and Panama, Wohl et
supports the idea that smaller watersheds can be
al. (2012) differentiate a steady-state end-
transport-limited for wood (Fremier et al. 2010).
member with gradual recruitment of wood
through individual tree fall that creates relatively

Large Wood National Manual July 2015


5-6
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Figure 5-2. Examples of Protruding Boulders Helping to Trap Wood along Streams

Upper left is on North St. Vrain Creek in Colorado and lower right is along Atlas Creek in Canada. In each case,
the yellow arrow indicates flow direction. (Photographs by Ellen Wohl)

Large Wood National Manual July 2015


5-7
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Figure 5-3. Wood Deposited Along the Top of Bank at the Outside of a Meander Bend on the Dall River in
Central Alaska

Yellow arrow indicates flow direction.

Large Wood National Manual July 2015


5-8
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Figure 5-4. Conceptual Illustration of Downstream Trends in Total Wood Load and Logjams along a River
Network

After Wohl and Jaeger (2009: Figure 7).

5.3.1.2 The Role of Floods greater hydraulic forces and transport capacity,
floods can also limit the stability of deliberately
Floods are likely to be the critical intervals for introduced wood such as ELJs. During a flood,
large wood management. Floods can result in wood can also form temporary debris dams that,
greater wood recruitment, particularly rainfall- when they break, release a surge of water and
generated floods that destabilize hillslopes and sediment downstream (Mao and Comiti 2010).
promote mass wood recruitment. Floods can also
result in substantially higher rates of wood Flood duration may be as important as flood
transport. Videos posted on the internet of flash magnitude in governing the balance between
floods in environments as diverse as Costa Rica wood recruitment and transport. Successive
and Nevada indicate that these floods can have a rainfall-generated floods within the Upper Rio
leading front of coarse wood, analogous to the Chagres drainage of Panama had very different
coarse sediment concentrated on the leading effects on instream wood load because of
edge of many debris flows. By generating much differences in duration (see Case Study above).

Large Wood National Manual July 2015


5-9
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

A flood in July 2007 resulted from intense wood within the channel and across the
rainfall that triggered widespread landsliding floodplain. This is one of the rationales
within the catchment, resulting in the formation commonly used for removing all wood
of channel-spanning logjams, each of which immediately after a flood. As exemplified by the
trapped a large volume of sediment and organic September 2013 floods along the Colorado Front
matter upstream (Wohl et al. 2009). A flood in Range, recruitment of wood into river corridors
December 2010 also resulted from intense is viewed as moving communities out of
rainfall that triggered dozens of landslides and compliance with FEMA requirements to return
even greater wood recruitment to the river river corridors to pre-flood conditions in order
network. The 2010 flood, however, lasted 2 days to qualify for federal financial assistance.
rather than the 5 hours of the 2007 flood, and Roughness is likely to decrease after a flood if
the enormous volumes of wood recruited during the net effect is removal of pre-existing wood
the longer flood were transported through the naturally present in the channel or deliberately
river network and into Lake Alhajuela, the placed there as part of river restoration.
reservoir behind Madden Dam (Wohl and Ogden Roughness may not change significantly after a
2013). flood if existing wood is transported
downstream but new wood is deposited during
Depending on the levels of wood recruitment
the waning stages of the flood.
versus transport, hydraulic roughness levels
after a flood can increase, decrease, or remain
relatively constant. Roughness is likely to
increase if the flood leaves large amounts of

CASE STUDY

Case Study of Flood Large Wood Recruitment and Retention

Jökulhlaup Flooding and Logjams along Dinwoody Creek, Wind River Range of Wyoming (Oswald and Wohl 2008):
A jökulhlaup (glacier outburst flood) from Grasshopper Glacier in September 2003 created anomalously high flows
along snowmelt-dominated Dinwoody Creek. High flows resulted in extensive bank erosion and large amounts of
large wood recruitment. Recruited trees formed channel-spanning logjams at longitudinal intervals of tens to
hundreds of meters. Channel aggradation upstream from each jam facilitated extensive overbank flooding and
floodplain deposition. Logjams likely persist for decades in this relatively dry region, and floodplain stratigraphy
indicates repeated episodic overbank deposition of the type observed following the 2003 flood.

Logjams formed along Dinwoody


Creek (right) and 1 meter (3.3
feet) of overbank deposition
(left) facilitated by log jams.

Large Wood National Manual July 2015


5-10
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

GUIDANCE

Effects of Individual Pieces of Wood and Accumulations Within Channels


 Increase hydraulic resistance; the magnitude of this effect depends on the abundance and spatial distribution
of wood relative to channel width, flow depth, and other sources of hydraulic resistance (e.g., grains,
bedforms, bends) (Shields and Smith 1992; Manga and Kirchner 2000; Mutz 2003; Daniels and Rhoads 2007;
Manners et al. 2007; Dunkerley 2014).
 Create local flow separation, with associated scour of the bed and banks, storage of finer sediment and
particulate organic matter, and enhanced diversity of aquatic habitat (Carlson et al. 1990; Nakamura and
Swanson 1993; Smith et al. 1993; Thompson 1995; Hart 2002; Brooks et al. 2003; Faustini and Jones 2003;
Ryan et al. 2014).
 Increase pool volume and fish abundance (Richmond and Fausch 1995; Buffington et al. 2002; Senter and
Pasternack 2010; Schenk et al. 2014b).
 Increase hyporheic exchange (Lautz et al. 2006; Hester and Doyle 2008; Lautz and Fanelli 2008; Wondzell et al.
2009; Sawyer et al. 2011).
 Increase nutrient retention and uptake (Buckley and Triska 1978; Bilby and Likens 1980; Munn and Meyer
1990; Raikow et al. 1995; Beckman and Wohl 2014).
 Alter bedform dimensions (e.g., wood can result in taller and more widely spaced steps in step-pool channels)
(Curran and Wohl 2003; MacFarlane and Wohl 2003).
 Alter substrate type, as in studies from the Pacific Northwest documenting forced alluvial reaches that would
likely have a bedrock bed if no wood were present to trap sediment (Montgomery et al. 1995a.
 Initiate and stabilize bars and alter channel planform (Gurnell et al. 2012; Mikuś et al. 2013).
 Create sufficient obstruction to flow to increase bank erosion, channel avulsion, and channel-floodplain
connectivity (Harwood and Brown 1993; Jeffries et al. 2003; Montgomery and Abbe 2006; Sear et al. 2010;
Wohl 2011b; Phillips 2012; Umazano et al. 2014).

Effects of Individual Pieces of Wood and Accumulations Within Channels on Floodplains


 Create preferred habitat for organisms, including macroinvertebrates during periods of flooding (Benke 2001)
and small mammals and birds during other periods (Harmon et al. 1986).
 Provide germination sites for riparian plants, particularly species that disperse via hydrochory (water transport
of propagules) (Pettit et al. 2005; Pettit and Naiman 2006).
 Enhance nutrient storage and uptake (Naiman et al. 2010).
 Provide more erosionally resistant “hard points” within the floodplain that influence rates of bank erosion and
lateral channel migration, as well as the age and species diversity of riparian forests (Montgomery and Abbe
2006; Collins et al. 2012).

corridor are the same as those associated with


5.3.1.3 Geomorphic and Ecological gradual recruitment of individual wood pieces.
Effects of Pulsed Inputs of Large Wood that remains within a river corridor has
Wood the net effect of creating a more physically
Many of the geomorphic and ecological effects heterogeneous environment that can increase
associated with pulsed inputs of wood to a river biodiversity because of habitat abundance and

Large Wood National Manual July 2015


5-11
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

diversity. By helping to retain nutrients, wood constrain NRV for instream wood, but these have
can also increase animal production (Huryn and only been published for relatively small rivers in
Wallace 1987; Gowan and Fausch 1996; the Pacific Northwest (Fox and Bolton 2007) and
Nagayama et al. 2012) and improve water the Colorado Front Range (Wohl 2011a). In the
quality. These well-documented effects can absence of quantitative data, numerical
provide a strong rationale for retaining or re- simulations may be used (Gregory et al. 2003b),
introducing wood to river corridors. With the although existing models are predominantly
exception of unmanaged rivers flowing through region specific (e.g., Beechie et al. 2000; Bragg
old-growth forest, it is reasonable to assume that 2000; Welty et al. 2002; Eaton et al. 2012).
any river within the United States historically
Although development of wood budgets must
had much greater wood loads within the river
start with an evaluation of the NRV of wood
corridor than are present today (Wohl 2014).
loadings in a watershed, or for a region, one
must evaluate the numerous management
5.3.2 Large-Scale and Long- activities and policies that have in the past, or
Term Considerations will in the future, impact the spatial and
temporal distribution of wood in a watershed. As
The existence of diverse sources of variability in the natural range provides an idea of how much
large wood recruitment and retention, either wood may be recruited and stored within a river
within a limited river segment or across a river system, a consideration of management actions
network, implies that “snapshot” assessments of will help to identify the limitations. For example,
wood load and wood mobility within a river at a infrastructure or safety considerations often
moment in time can be misleading. It is more require the removal of large pulses of wood (e.g.,
appropriate to think about wood in the context Benda et al. 2003b) (also see Chapter 7, Risk
of a wood budget (Benda and Sias 2003) that Considerations). The following discussion covers
varies through time and space, and in terms of a two impacts on stream systems: dams and land-
historical or natural range of variability. Natural use/land-cover.
range of variability (NRV) describes the range of
temporal and spatial variations in river
5.3.2.1 Dams
parameters such as flow regime, channel
geometry, or wood load (Wohl 2011c). Wood Dams interrupt the movement of water,
loadings depend on the recruitment, storage, and sediment, and organic matter through a
transport of wood; and a budget may be watershed. At the most fundamental level, dams
constructed using an approximation of the alter a wood budget when the upstream
following (Martin and Benda 2001). reservoir traps the large wood recruited and
transported by the river. Changes to the flow
Equation 5-1: regime and sediment supply downstream,
∆S = (I∆x – O∆x + Qi – QO – D)∆t however, also have the potential to greatly
impact a river’s wood budget. Channel mobility
where ΔS is a change in storage within a channel is often reduced and, as a result, the recruitment
with a length of Δx over the time interval Δt; I is and regeneration of riparian forests (Scott et al.
lateral wood recruitment, O is loss of wood from 1996; Kloehn et al. 2008). Channel pattern and
the active channel to overbank deposition during other geomorphic variables can also shift
flood events, abandonment of jams, and burial; (Walter and Merrits 2008). Lassettre and Piegay
D is in situ decay; and Qi and Qo are fluvial (2008) documented an increase in wood
transport of wood into and out of the segment. recruitment on the Ain River downstream from a
Large compilations of field data on relatively dam in response to floodplain afforestation as a
unmanaged rivers within a region can be used to result of a change from a braided to meandering

Large Wood National Manual July 2015


5-12
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

channel pattern. Various studies have noted (Major et al. 2012; East et al. 2014) and many
shifts in the composition of the riparian more removals are planned (e.g., Gosnell and
vegetation community (Nilsson and Berggren Kelly 2010). The majority of the removals have
2000; Friedman et al. 2005). For example, occurred in the northeastern region of the
cottonwood-dominated forests have declined United States, where there is a high density of
along many western rivers, at least in part dams. Various organizations have begun to track
because of dams (Rood and Mahoney 1990). In removals, and the associated dialogue (e.g.,
some systems, dense stands of shrubby tamarisk American Rivers, www.americanrivers.org/
took the place of cottonwood (Merritt and Poff initiatives/dams and the Clearinghouse for Dam
2010). These changes can exacerbate reductions Removal, a collaboration of California
in channel mobility (Dean et al. 2011; Manners et Universities, http://library.ucr.edu/wrca/
al. 2014) but may also alter the potential for the collections/cdri/).
recruitment of jam-forming logs or the wood
Dam removals have the potential to greatly affect
storage capacity on the floodplains.
the wood budget, both in the short term, as the
The magnitude and length-scale of the impact of flux of sediment and wood that accumulated
a dam on the wood budget varies greatly and within the reservoir passes downstream, and the
depends on the type of dam (i.e., its intended long term, as the river responds to the new,
purpose) (Magilligan and Nislow 2005), the more natural flow regime and sediment supply.
hydroclimatic region (Graf 1999), and the Between 2011 and 2013, two large dams on the
geologic setting (Grant et al. 2003; Schmidt and Elwha River in Washington State were removed.
Wilcock 2008). In the western mountains and Prior to the removal, Brenkman et al. (2012)
plains regions of the U.S., dams generally have a documented that the number of jams below the
more profound impact on the hydrologic regime dams was one to two orders of magnitude less
(Graf 1999); however, the direction of these than upstream. East et al. (2014) noted that new
changes is relatively constant across dam sites wood was present after dam removal, most of
and include an increase in the minimum flows which had eroded from within former reservoir
and a decrease in maximum flows (Magilligan deposits. Little additional anecdotal evidence
and Nislow 2005). Inputs from tributaries, or exists on how large wood loadings change after
changes in the geomorphic character of the the removal of a dam.
downstream channel, can offset the impact of a
Dam operations are also being re-evaluated in
dam (Williams and Wolman 1984; Schmidt and
order to meet a growing demand for increasingly
Wilcock 2008; Draut et al. 2011). One must also
limited water supplies (Watts et al. 2011) as well
consider where within the watershed the dam is
as downstream ecosystem needs (Whiting
located with respect to the dominant controls on
2002). Increasingly, new dam operation schemes
wood recruitment and transport (see Figures 5-1
mimic the natural flow regime, or at least a
and 5-4). The direct impact of these downstream
reduced and/or simplified version of it (Poff and
changes on a wood budget have rarely been
Zimmerman 2010). These flow regimes can
quantified.
restore lateral dynamism that recruits wood but
There has been a growing movement to remove can also reintroduce natural processes that may
dams (Doyle et al. 2003). Motivations for push a wood budget closer to its natural range.
removal include an aging infrastructure, public For example, the restoration of flow patterns to
safety, and an enhanced environmental match the seed release and germination needs of
awareness of the importance of free-flowing native riparian plant species on regulated rivers
rivers (Pohl 2002). To date, removals have in Alberta, Canada and Nevada promoted new
predominately occurred on smaller dams, but in recruitment of cottonwood and willow (Rood et
recent years large dams have been removed al. 2003) (Figure 5-5). Furthermore, some dam

Large Wood National Manual July 2015


5-13
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

operators have now modified their management wood to a channel; and (2) the larger
practices to include downstream placement or watershed—important for determining the flow
disposal for wood deposited in reservoir regime and sediment supply, and indirectly
impoundments. critical for the wood budget and long-term
Figure 5-5. Impact that Reoperation of Dams, to health of a river system.
Include More Natural Elements of the In evaluating the impact of land-use/land-cover
Hydrograph, Can Have on a Riparian Ecosystem
within the river corridor on the wood budget, it
is necessary to remain conscientious of the
dominant recruitment mechanisms. For example,
in steep landscapes and in headwater reaches,
bank erosion is limited and landslides deliver the
majority of wood to the channel (May and
Gresswell 2003a, b). In other settings, landslides
are not important (Kasprak et al. 2012), and
instead wood recruitment depends on bank
erosion or tree-fall in proximity to the channel.
For this latter scenario, hillslope land-use and
management history is not as important to
consider. Instead, the state of the river bank and
floodplain, including any past engineering
actions, are important.

Numerous studies have quantified the impact of


forest management history on wood loadings
and shown repeatedly that old-growth forests
generally have greater jam frequency or wood
volume than previously cleared ones (Bilby and
Ward 1991; Collins and Montgomery 2002;
Collins et al. 2002; Warren et al. 2009). Forest
Photos show the Truckee River in 1977 (a) with management in the western United States has
only relatively old stands of cottonwood and no
also included fire suppression and fuels
new recruitment and in 1997 (b) after a series of
years (1987 and 1995) with flood hydrographs that management, resulting in the alteration of fire
met the recruitment needs of cottonwood and regimes (Dwire and Kauffman 2003). Prior to
willow. Photos were taken from approximately the European settlement, fire was the most
same location, however, not matched exactly important disturbance process in many of these
because of a shift in the channel. (Figure taken from western systems, and represents an important
Rood et al. 2003.) mechanism for wood delivery (Naiman et al.
2000). Changes to the fire regime, therefore,
5.3.2.2 Land-use/Land-cover altered the frequency and magnitude of wood
The mosaic of land-uses and land-covers within a recruitment processes.
watershed impacts the timing, rate, and spatial
Human development of a river corridor impacts
distribution of biophysical fluxes into a river
the wood budget in a number of ways.
channel. The focus herein is on two regions
Conversion of forested hillslopes or floodplains
within a watershed: (1) the river corridor,
removes the local wood source, although where
including the hillslope adjacent to the channel or
a riparian buffer is maintained, this impact may
floodplain, the floodplain, and the channel
be mitigated (Robison and Beschta 1990). Even a
bank—important for the direct contribution of
single Forest Service road paralleling a channel

Large Wood National Manual July 2015


5-14
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

in a well-forested watershed can significantly (Sleeter et al. 2012). The study highlights the fact
reduce wood supply (Meredith et al. 2014). that these trends were highly variable and will
Engineering actions that stabilize banks in order likely continue to as a result of changing climate,
to protect infrastructure and valuable crop land population trends, and local sociopolitical
reduce recruitment potential (Angradi et al. pressures.
2004) and make streams more efficient at
transporting wood, reducing storage potential.
5.3.2.3 Thresholds and Alternate
Roads reduce the infiltration capacity on the Stable States
road surface itself and intercept surface flow and In ecosystems where wood decays relatively
throughflow by cutslopes, resulting in increased slowly and hydrologic variability is limited—
surface runoff and leading to periodic mass mostly cold-temperate and boreal regions—the
failure from the adjacent hillslopes (Wemple and limited studies that have considered the
Jones 2003; Arnáez et al. 2004; Goode et al. existence of alternate stable states suggest that
2012), potentially increasing recruitment rates. river corridors (channel-floodplain systems) can
Removal of forests also reduces the infiltration assume alternate stable states in relation to
capacity of the landscape (Matheussen et al. wood load. The slowness of wood decay is
2000). With the addition of urban or suburban important because it means that, despite natural
infrastructure and roads, the entire drainage fluctuations in volume of large wood recruitment
network can shift. Often, the hydrograph through time and space, the river corridor
becomes “flashier,” with a more rapid storm to always contains some wood. As long as sufficient
stream signal, increased flood peak magnitude, wood is present to obstruct transport of newly
and reduced base flows. River channels, as a recruited wood, a positive feedback can develop
result, can incise and become disconnected from in which stable wood traps wood in transport,
their floodplains (Walsh and Roy 2005). enhancing bank erosion, channel avulsion,
formation of multiple, subparallel channels, and
The Puget Sound region of Washington provides overbank flows, and thus further increasing
a good example of how changes in land-use can wood recruitment and retention (Figure 5-6)
alter a wood budget. Urbanization increased the (Collins et al. 2012; Wohl and Beckman 2014b).
proportion of the basin May et al. (1997) studied These river corridors can become stabilized in a
from 0% to 60%, reducing the volume of large wood-poor condition if people remove instream
wood from approximately 1,200 cubic meters and floodplain wood and reduce recruitment
per kilometer to near zero, and the number of through processes such as timber harvest,
pieces declined by 75% (May et al. 1997). channelization, inadequate stream crossings, and
bank stabilization. River corridors with
Land-uses and land-cover are continually
extremely high peak discharge per unit drainage
shifting, especially in the western United States
area and/or high rates of wood decay, such as
The USGS undertook a study of land changes
those in the tropics and subtropics, may have
across the United States between 1973 and
such continually changing wood loads that they
2000. Currently, only the results from the
never achieve the positive feedbacks that create
Western region have been published and tell a
persistent, wood-rich conditions (Benke and
story of losses in forest cover as a result of
Wallace 1990; Wohl et al. 2009, 2011, 2012).
logging, fire, urbanization, and other land uses

Large Wood National Manual July 2015


5-15
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Figure 5-6. Conceptual Illustration of Wood-Related Feedback

Inset photographs illustrate (clockwise starting from upper right) fine sediment and organic matter
deposited along the margin of a stream in the backwater created by a channel-spanning logjam; multiple,
subparallel channels (flow direction indicated by white arrows) along the floodplain of North St. Vrain Creek
in Colorado; bank erosion opposite a tree that fell into the stream; trees gradually falling into a stream as a
result of bank erosion; a channel-spanning logjam; and a spring-head channel fed by hyporheic return flow
along the floodplain of Cony Creek in Colorado. (Photographs by Ellen Wohl)

hydraulics and slows water on meander bends


5.4 Large Wood and and can reduce erosion (Daniels and Rhoads
2004); and influences sediment storage, spacing
River Crossing of riffle-pool sequences, and vertical channel
stability (Thompson 1995). Also, large wood
Interaction forms physical holding locations for fish and
The generation, storage, and transport of large serves as the base of the aquatic food web (Allan
wood from trees falling into a river influence 1995). Consequently, the known benefits of large
channel equilibrium and stability and improves wood have led to a rise in the use of engineered
instream habitat (Lassettre and Harris 2001; log jams and other large wood installations to
Brooks et al. 2006b). For example, wood can restore instream habitat.
create stable step-pools in steep mountain
Wholesale woody debris removal following large
channels (Wohl and Merritt 2008); influences
floods to protect bridges, culverts, and other

Large Wood National Manual July 2015


5-16
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

infrastructure removes an important mechanism


for long-term channel bed and bank stability and
5.5 Large Wood’s Impact
future instream habitat. Effective river on Bridges and
management must strike a balance between the
removal of wood that can be problematic and the
Culverts
wood that could be of benefit is left in the
channel and on the floodplain. Consideration of 5.5.1 National Overview
the diversity, quantity, retention, and transport
A national survey by the Federal Highway
of large wood is an important aspect of proper
Administration (Diehl 1997) found that floating
flood recovery.
debris contributed to more than one-third of
Removal of large wood is typically performed to bridge failures in the United States Roughly two-
increase conveyance for water, sediment, thirds of bridge failures are attributed to
additional large wood, and ice during the next hydraulic problems such as floods and scour,
flood to protect infrastructure, buildings, and both of which are linked to large wood
unmovable improved property where vertical accumulation and structure clogging. There are
instability and risk of rapid channel migration approximately 485,000 bridges over water in the
(i.e., avulsion) exist (Schiff et al. 2014). Common United States, with 17,000 listed as being scour
scenarios for removal of large wood include critical.
clogged bridges and culverts, filled channels and
The National Bridge Inspections and Safety
floodplains, and avulsed channels. Too often
programs (initiated in the 1980s after many
wood is removed without addressing the stream
bridge failures) highlighted the known risks of
crossing characteristics that induced deposition.
bridge failure and the level of national
Many rivers have a history of bridge and culvert investment needed to keep bridges safe. The
failures and backwater flooding due to large Federal Highway Administration considers the
wood jams. Disproportionate or extreme flood accumulation of wood at bridges and culverts to
damage is often highly correlated to stream be a major safety and maintenance problem.
crossings. For example 200 state bridges, 280
Guidance exists on the contributions of large
local bridges, and 960 local culverts were
wood to scour at bridges and culverts (Bradley et
damaged in Vermont alone during Tropical
al. 2005; Lagasse et al. 2010; Arneson et al.
Storm Irene in 2011 (Pealer 2012). Many
2012). Diehl (1997) provides detailed research
structure failures are observed to be caused by
on the formation of log jams at bridges. Lagasse
wood; therefore, many communities shy away
et al. (2012) identify locations to place structures
from leaving wood in a channel during flood
to limit conflicts with wood.
recovery or from implementing wood-based
restoration projects. A need exists to understand
the budget of large wood at the watershed level,
5.5.1.1 Large Wood and Bridge
the risk large wood elements and jams pose in Scour
rivers, and the important roles large wood plays The most frequent bridge failure is due to scour
toward channel stability and habitat. A better at piers or abutments. Large wood
understanding of large wood benefits and accumulations create scour by restricting the
mitigation will help encourage stream crossing waterway cross-sectional area, leading to higher
designs for wood conveyance and reduce velocities and scour-inducing turbulence. Log
infrastructure risks while maintaining the jams can also redirect flowing water toward the
important natural functions of large wood. river bed where vulnerable piles and footings
are located.

Large Wood National Manual July 2015


5-17
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Bridge failures have been attributed to debris Sediment, tree branches, large wood, and trash
jams against piers that create high lateral forces often intermittently block bridge and culvert
and induce scour (Wipf et al. 2012). The trapping openings during floods leading to contraction or
problems include accumulations at individual local scour and eventual structure failure. The
bridge piers that increase local pier scour and clogged structure then releases the accumulated
span-wide channel blockages that increase material, and the true cause of the failure may
contraction and local abutment scour. Log jams not be properly recorded.
at bridge piers begin at the water surface and
then expand upstream and laterally. The width 5.5.1.4 Forest Roads
of pier log jams is often influenced by a few long
logs. Some log jams extend from pier to pier In the Pacific Northwest, Furniss et al. (1998)
while others ultimately fill the entire bridge conducted data collection and evaluations on the
opening. performance of logging road culverts with
respect to large wood, sediment, flood flow
Lagasse et al. (2010) details the procedure by conveyances, and fish passage. Numerous culvert
which debris contributes to bridge scour based failure mechanisms were identified including
on field data and laboratory tests. Expanded debris flows with channel scour and deposition,
protocols for predicting scour depths at piers lodgment of large wood that plugged culverts,
and abutments and due to contraction are sediment deposition from mass flows, and
contained in Hydraulic Engineering Circular 18 hydraulic exceedance that overtopped roads.
(Arneson et al. 2012). Sediment and large wood accounted for over
90% of the failures, while failure due to
5.5.1.2 Large Wood and Hydraulic hydraulic conveyance alone was less than 10%
Capacity (Table 5-1).

Large wood jams at bridges reduce the cross- Table 5-1. Culvert Failure Data
sectional flow area and reduce the structure's
hydraulic capacity to pass flood flows raising Percent of Total Failure Type
upstream flood levels (i.e., backwatering). Large 36 Sediment
wood also increases the risk of overtopping and 26 Debris torrent
eroding the abutments and embankment. 17 Woody debris
12 Woody debris/sediment
5.5.1.3 Regional Data 9 Hydraulic exceedance
In a recent study of 691 bridge failures in the
United States, 52% were hydraulic failures, of 5.5.1.5 Woody Debris at Dams
which 40% were due to scour that largely
Large wood transport to, and accumulation at,
originated from large wood accumulation (Cook
dams is a continuing problem because dams
2014). In New York, of 92 bridge failures
block the movement of both bedload and
between 1987 and 2011, 20% were due to scour
entrained material. Most hydroelectric dams
and 28% were due to floods. Hamill (1999)
have to use trash racks at turbine intakes and
studied bridge failure and found that floods
often use floating booms to trap wood.
leading to scour and large wood piled against the
structure were the most common cause of The increase in removal of obsolete dams over
damages. Large wood and clogging is often cited the past 20 years has allowed more wood to
to contribute to most bridge and culvert damage. move downstream and has also revealed large
quantities of wood previously submerged in
The effects of large wood on structures is likely
impoundments. For example, after draining the
higher than reported (Agrawal et al. 2007).

Large Wood National Manual July 2015


5-18
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

former Carbonton Dam in North Carolina, an Additional sources include islands such as
upstream bridge within the impounded area was observed on large rivers with multiple channel
found to be fully blocked by the submerged large paths such as the Connecticut and Penobscot
wood. The removal of the Great Works Dam from Rivers. Large loads of wood can originate from
the Penobscot River in Maine in 2011 exposed colluvial inputs from valley wall mass wasting.
tree remnants originating from upstream natural Landslides in the Catskill Mountains of New York
transport and sawed timber that was the along Stony Clove Creek and Westkill Creek and
product of logging drives in the nineteenth in the Berkshire Mountains along Cold River and
century. Large wood was also observed in the North River were the source of significant
pool upstream of recently removed large dams quantities of both large wood and sediment. The
such as Elwha Dam on the Elwha River, Condit valley wall erosion originated from large floods
Dam on the White Salmon River in Washington, that scoured the bottom of the valley wall and
and Milltown Dam on the Clark Fork River in the overlying material slide down the slope and
Montana. into the river.

Decisions regarding large wood management at


dams and dam removals must rely on 5.6.2 Wood Transport to
site-specific analyses of constraints and Bridges
opportunities, but it is clear that dropping wood
immediately downstream of existing structures Trees that fall into or across river channels tend
can represent a financially and ecologically to align parallel to the flow, with the stump at the
effective strategy, and that allowing exposed upstream end. Observations of floating wood in
wood to route itself passively following a dam eastern rivers reveals that smaller wood and logs
removal can help attenuate sediment pulses and without stumps tend to float with random
provide valuable habitat complexity. orientation to the flow and can pass wide bridge
spans and large culverts. Montgomery and
Piégay (2003) showed that trees with spreading
5.6 Watershed-Scale branches tend to form bed snags, while
Risk to Structures cylindrical-shaped conifers are more readily
transported downstream. For narrow streams,
the channel width can be used to estimate the log
5.6.1 System-scale and Local length that can readily be transported
Large Wood Sources downstream (Lagasse et al. 2010).

The primary sources of large pieces of wood in In smaller streams and during low flows, the
rivers are forested riverbanks, the top of banks, stumps on floating logs drag on the river bed,
and adjacent floodplain edges (Diehl 1997). and the tree may become grounded on existing
Unstable channels that are widening, degrading, sediment bars. Sediment then accumulates
or actively migrating are continually behind the tree, and the shape and depth of the
undermining their banks and causing trees to fall bar changes. Grounded large wood creates
into the channel from within the meander belt channel roughness and hydraulic diversity that
(Williams 1986) or material contribution zones improves habitat. As sedimentation takes place
(Smith et al. 2008). Other trees fall due to age, and the level of embeddedness increases,
wind, or ice storms. A single alluvial meander grounded wood can remain in place for decades
bend on the Pomperaug River with a tall failing until it decays into smaller pieces or a large flood
bank in Southbury, Connecticut, has been mobilizes the wood again.
observed to contribute trees with intact root
masses annually over the past 15 years.

Large Wood National Manual July 2015


5-19
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

5.6.3 Critical Wood Size (Rosgen and Silvey 1996), can have a strong
influence on sediment deposition and flow
The risk of a large wood blockage is a function of patterns. Conversely, bedform morphology helps
the structure size and alignment, large wood determine stability and transport characteristics
dimensions, channel width, and watershed wood of individual and massed wood elements; the
yield (Lagasse et al. 2010). Large tree trunks and ratios of rootwad and crown diameters relative
branches that are longer than the channel width to depth to riffle crests and other bedforms at
tend to get caught on the banks or on meander various flood stages are important determinants.
bends and remain in place unless a large flood
takes place. Wood that is shorter than the 5.6.5 Floodplain Wood
channel width will tend to regularly get
transported downstream during frequent floods. Large wood can both establish and isolate
If the wood is very small, it will likely be floodplains from channels. Large deposits of
transported through bridges and culverts unless wood of in-channel wood can elevate the channel
they are very undersized. Undersized structures bed and establish more frequent floodplain
are common. A review of over 3,000 culverts in connection. In appropriate areas, more
Vermont indicated that over 25% of structures floodplain connection is beneficial in spreading
were less than half the channel bankfull width floodwaters, depositing sediment, and creating
(Schiff et al. 2008b). Higher risks are associated areas for nutrient uptake. Frequent floodplain
with medium-sized wood that is shorter than the inundation is also important for diversifying
channel width and so can readily be transported riparian habitat for birds and wildlife that rely
downstream but is prone to accumulation at on river corridors.
undersized structures (Gurnell et al. 2002).
Large wood deposits that increase the
Once a jam begins with blocked key pieces of inundation frequency of floodplains in developed
large wood, smaller brush, leaves, and bedload areas can lead to property and infrastructure
sediment add to the mass and can clog a damages. During floods they can disconnect
structure rapidly. In this way, a small number of secondary channels, raising floodwater levels.
large logs near a structure and an abundance of During Tropical Storm Irene, wood inputs from
small wood delivered from upstream can lead to landslides on the Cold and Chickley Rivers
structure clogging and failure. blocked the channel around the Charlemont
Island in the Deerfield River leading to flood
5.6.4 Bed Forms damages in the area. Along Bushkill Creek in
New York, wood preferentially accumulated
At the reach scale, wood in rivers influences the along the top of banks during Irene, limiting
shape and size of the channel cross section, lateral flows from the channel into the floodplain
pattern, and profile. Pools can be formed and confining flood flows within the channel that
upstream of large wood elements or jams that increased risks downstream.
create local dams. Scour holes are common
along, under, or just downstream of large wood. Those issues noted, wood and woody vegetation
Wood also influences sediment bars and flood on floodplains perform many functions. They
benches located next to the wetted channel. trap wood and mobile ice during overbank flows
Wood deposits on floodplains and inside that might otherwise clog stream crossings. They
channels affects overbank flows approaching attenuate floodplain flow velocities and, at wide
bridges and culverts, while woody debris along distribution, can reduce the risk of channel
the banks affects lateral floodplain connectivity. avulsions. They induce sediment and propagule
Excess wood in channels prone to deposition, deposition of floodplains, creating in many
such as braided and anastomosed stream types settings optimal conditions for self-revegetation.

Large Wood National Manual July 2015


5-20
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Finally, they create microtopographic diversity hydraulics, confine floodplain flow, encroach on
on floodplains with consequent beneficial effects the channel, and create higher scour potential.
on habitat diversity. Large wood that is cut and disposed of along
road embankments across floodplains can lead
5.6.6 Spoil Piles to increased downstream jamming at structures.

Wood that has been cleared from channels


approaching bridges and culverts and disposed
of along the banks can interfere with structure

CASE STUDY

Wood Recruitment and Structure Damage During Tropical Storm Irene

Tropical Storm Irene moved across eastern New York, western Massachusetts, and Vermont on August 28, 2011,
damaging hundreds of miles of highways and hundreds of bridges. The storm caused flooding that inundated
entire valley bottoms and caused channel erosion and floodplain enlargement. Landscape changes also included
many landslides on forested valley walls that generated high volumes of large wood inputs to channels that
blocked bridges and culverts leading to structural failures.

In New York, a 500-year-frequency flood with unit discharges of 860 cubic feet per second per square mile of
watershed along Westkill Creek in the Catskill Mountains caused 1.5 meters (5 feet) of bed down-cutting (i.e.,
degradation) and landslides along 610 meters (2,000 feet) of channel. Large wood clogged structures, bridges
failed on Route 42, and sections of the highway were washed away. Large and small wood combined with
sediment that obstructed the channel and blocked bridges in the towns of Windham, Maplecrest, and Shandakan.
Wood jams contributed to bridge scour and overtopping that damaged numerous structures in the region.

In Massachusetts, four landslides along the Route 2 Mohawk Trail left large wood and debris in the Cold River that
contributed to road embankment loss. The wood also caused floodplain blockages along the Deerfield River in
Charlemont. The National Guard was mobilized and cleared all of the debris from bridges in Colrain and Buckland.

The upper reaches of Esopus Creek in New York's Catskill Mountains had a unit discharge of 460 cubic feet per
second per square mile of watershed. The flood caused channel widening and degradation that generated large
volumes of wood from the undermined banks and valley walls. Floodplains were obstructed, channels were
jammed, and bridges were blocked and destroyed by scour and overtopped due to wood and sediment loading.
Headwater channels such as McKinley Hollow were filled with wood and sediment leading to structure failure and
road washout.

A series of three landslides up to 18 meters (60 feet) high along Stony Clove Creek in the Catskill Mountains
delivered colluvium and large wood into the channel. Large wood and sediment had to be removed from the
downstream channel and bridge at Route 214 in Phoenicia, and the pier at Bridge Street was damaged. Again, all of
the large wood was removed from the channel.

Large Wood National Manual July 2015


5-21
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

volume of wood generated in the flood for


5.7 Structure restoration purposes such as engineered log
Vulnerability and jams or bank revetments.

Design Part of the solution is to increase new structure


resiliency by proper sizing to pass wood
Recommendations downstream with reduced risk of blockage and
damages. This approach has been shown to be
5.7.1 Vulnerability more cost-effective in the long run (Gillespie et
al. 2014).
Bridge and culvert vulnerability to large wood
jams and debris is a function of the watershed Proper structure sizing includes consideration of
and channel characteristics and the structure transport of water, woody debris, sediment, and
geometry. The debris size and volume depend on ice both in the channel and floodplains.
the watershed forest type and age and,
particularly, the floodplain area within
30.5 meters (100 feet) of the channel.
5.7.3 Improved Bridge and
Culvert Design to Pass
In high risk areas, the channel banks and
floodplain can be inventoried to identify the size
Large Wood
and volume of potential wood load. The size of Removal of large wood and sediment often
living and downed trees is compared to the begins at clogged bridges and culverts following
channel bankfull width to assess the quantity large floods. Material should be removed
and size of wood that could be transported. throughout the entire structure and in the
Trees that are longer than the bankfull width are adjacent bankfull channel entering and leaving
less likely to be transported to structures. the structure. The cover over the footings should
Denser, and often older, forests tend to produce be reviewed at structures to identify if the
more wood at their edges near channels than structure remains stable following the flood. (See
younger forests in the early stages of succession. Table 5-2.)
Bridge and culvert clear spans should be When removing large wood and sediment from
compared with the size and amount of potential bridges and culverts, an acceptable profile needs
wood load. Proposed structures should at least to be established that will not lead to an erosion
span the entire bankfull channel (i.e., structure face moving upstream (i.e., headcutting) or
width >1.0 channel bankfull width) with increase the risk of avulsion. Uniform and
adequate under clearance for crowns and gradual slope transitions in and out of the
rootwads. structure should be established.

Replacement of undersized structures that block


5.7.2 Increasing Structure sediment and large wood should be considered
Resiliency to reduce the risks during future floods
(Lassettre and Kondolf 2012).
Many communities that see natural wood and log
jams in rivers express concern about
downstream bridges and culverts. Following
Tropical Storm Irene, there was a rush to clear
every piece of wood out of channels to ensure
remaining structures would not clog in the next
flood. The flood recovery was marked by a
strong reluctance to allow the use of the large

Large Wood National Manual July 2015


5-22
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Table 5-2. Debris Countermeasures for Culverts and Bridges


Measure Culverts Bridges
Structural Measures
Debris Deflector Structure that deflects the major portion of the debris away Structure placed upstream of the bridge piers to deflect and
from the culvert entrance. Normally “V”-shaped in plan with guide debris through the bridge opening. Normally V-shaped
the apex upstream. in plan with the apex upstream.
Debris Rack Structure placed across the stream channel to collect the
debris before it reaches the culvert entrance. Usually vertical
and at right angles to the streamflow, but may be skewed
with the flow or inclined with the vertical.
Debris Riser Closed-type structure placed directly over the culvert inlet to
cause deposition of flowing debris and fine detritus before it
reaches the culvert inlet. Usually built of metal pipe.
Debris Crib Open crib-type structure placed vertically over the culvert Walls built between open-pile bents to prevent debris lodging
inlet in log-cabin fashion to prevent inflow of coarse bed load between the bents. Typically constructed out of timber or
and light floating debris. metal material.
Debris Fin Walls built in the stream channel upstream of the culvert. Walls built in the stream channel upstream of the bridge to
Purpose is to align the debris with the culvert so that the align large floating trees so that their length is parallel to the
debris would pass through the culvert without accumulating flow, enabling them to pass under the bridge without
at the inlet. incident. Also referred to as a “pier nose extension.”
Debris Dam/Basin Structure placed across well-defined channels to form basin, Structure placed across well-defined channels to form basin,
which impedes the stream flow and provides storage space which impedes the streamflow and provides storage space
for deposits of detritus and floating debris. for deposits of detritus and floating debris.
River Training Structure placed in the river flow to create counter-rotating
Structures stream-wise vortices in the wake to modify the near-bed flow
pattern to redistribute flow and sediment transport within
the channel cross section.
Flood Relief Overtopping or flow through structure that diverts excess
Sections flow and floating debris away from the bridge structure and
through the structure.

Large Wood National Manual July 2015


5-23
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Measure Culverts Bridges


Debris Sweeper Polyethylene device that is attached to a vertical stainless
steel cable or column affixed to the upstream side of the
bridge pier. Device travels vertically along the pier as the
water surface rises and falls. It is also rotated by the flow,
causing the debris to be deflected away from the pier and
through the bridge opening.
Booms Logs or timbers that float on the water surface to collect
floating drift. Drift booms require guides or stays to hold
them in place laterally.
Design Features Structural features that can be implemented in the design of a
proposed bridge structure. The first feature is freeboard,
which is a safety precaution providing additional space
between the maximum water surface elevation and the low
chord elevation of the bridge. The second feature is related to
the type of piers and the location and spacing of the piers.
Ideally, the pier should be a solid wall type aligned with the
approaching flow. It should also be located and spaced such
that the potential for debris accumulation is minimized. The
third feature involves the use of special superstructure
design, such as thin decks, to prevent or reduce the debris
accumulation on the structure when the flood stage rises
above the deck. The last feature involves providing adequate
access to the structure for emergency and annual
maintenance.
Combination Combination of two or more debris-control structures at one
Devices site to handle more than one type of debris and to provide
additional insurance against the culvert inlet from becoming
clogged.

Large Wood National Manual July 2015


5-24
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Measure Culverts Bridges


Nonstructural Measures
Emergency and Although not always feasible for remote culverts or culverts Emergency maintenance could involve removing debris from
Annual with small drainage areas, maintenance could be a viable the bridge piers and/or abutments; placing riprap near the
Maintenance option for larger culverts with fairly large drainage basins. piers, abutments, or where erosion is occurring due to flow
Emergency maintenance could involve removing debris from impingement created by the debris accumulation; and/or
the culvert entrance and/or an existing debris-control dredging of the channel bottom. Annual maintenance could
structure. Annual maintenance could involve removing debris involve debris removal and repair to any existing structural
from within the culvert, at the culvert entrance, and/or measures.
immediately upstream of the culvert, or repairing any
existing structural measures.
Source: FHWA (2005).

Large Wood National Manual July 2015


5-25
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

The hydraulic sizing of bridges and culverts is structures. If past flood damages have occurred
well established (e.g., FHWA 1985b; VTrans due to clogging or are suspected due to a high
2001), yet current design guidelines (e.g., UNH possible wood load during a flood, the bridge or
2009; MassDOT 2010) are now turning to culvert should be designed to fill to 80% of the
geomorphic principles to both naturalize stream opening height (i.e., Hw/D < 0.8) during clear
crossings and make them less prone to flood flow to allow vertical space in the structure to
damages. Structures commonly fail due to pass sediment, large wood, and ice. Post-flood
geomorphic incompatibility (Schiff et al. 2008b) evaluations of failed structures indicate that
such as stream instability (FHWA 2012) and structures that were filled or overtopped during
clogging with sediment and large wood (Furniss a flood were typically damaged due to large
et al. 1998). wood accumulation and clogging (Furniss et al.
1998) (Figure 5-8).
A geomorphic-engineering design approach
(Figure 5-7) (Schiff et al. 2014) is recommended Proper structure design must consider the
to optimize structure size and type so that the floodplain setting (i.e., does the channel have
river channel form and processes can play out in broad floodplains or narrow benches?). The
a more natural way. Structures that are sized at floodplain width and frequency of inundation are
the bankfull channel width or larger are important to fine tune the structure width to
achieve an acceptable flow width during floods.
 Able to convey more water, sediment, large
Overflow structures should be considered in
wood, and ice.
broad floodplain settings although structures
 Less prone to clogging. placed away from the channel and at higher
elevations than the banks can be prone to
 Less prone to bridge scour.
clogging due to slower flow velocities than in the
 More compatible with a stable channel. channel.
 Able to pass fish and wildlife.

A central theme of the geomorphic-engineering


design approach is that bridges and culverts
should be sized to operate with a headwater–to-
depth ratio less than 1 (i.e., part full) to leave
space for sediment and wood. Even with a large
volume of historical information indicating that
wood and sediment are the primary mechanisms
of structure failure, the majority of structure
design is still currently done primarily with
consideration of only clear-flow hydraulics,
leading to recurring public safety, financial, and
environmental impacts.
One example of geomorphically based stream
crossing design is the USFS stream simulation
approach (USFS 2008). Another is presented in
the box below.
The potential input of large wood during a flood
needs to be evaluated along the channel reach
and watershed to know if a structure is prone to
clogging. In northern climates, ice can also clog

Large Wood National Manual July 2015


5-26
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

GUIDANCE

Geomorphic Engineering Design Recommendations Required by the


Vermont Stream Alteration General Permit (2013) (Schiff et al. 2014; VTANR 2014).
 Wstructure = 1.0 x Wbankfull channel.
 Hopening = 4 x Dbankfull channel.
 Dembed = 30% Hopening or D84 for boulder bed, whichever is larger.
 Match channel profile and create uniform longitudinal transitions at inlet and outlet.
 Structure shall not obstruct aquatic organism passage.
 Evaluate structure for clear-flow hydraulics and perform checks for large wood and sediment deposition and
scour.
 Where physical constraints preclude achievement of the 4.0X opening height standard and any potential
increase in flooding hazard associated with a reduced opening height will be offset by other factors such as a
lower roadway fill height, the minimum opening height shall be > 3.0X the mean bankfull channel depth, as
approved by the Secretary of the Vermont Agency of Natural Resources (VTANR), and as specified in the most
current version of the Vermont Agency of Transportation (VTrans) Hydraulics Manual (2001).
 Where more capacity is needed based on flow, material deposition, or scour, structure width shall be 1.2 x
bankfull width or larger (e.g., floodprone width).
 Where channel gradient is 0.5% or less or the structure is under outlet control, depth of embeddedness may
be reduced, as approved by the Secretary of the VTANR.
 Retain sediment throughout structure and maintain natural sediment transport.
 Avoid backwatering at inlet and naturalize the movement of large wood and ice.
 Design Q and Hw/Hopening from state hydraulic standards (VTrans 2001).
 Match channel hydraulic conditions for design flood, fish passage, and low flows.
 Align structure parallel to flow in channel.
 Maximize fish and wildlife passage.

Large Wood National Manual July 2015


5-27
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Figure 5-7. Geomorphic Engineering Structure Sizing Method

Source: Schiff et al. (2014)

Large Wood National Manual July 2015


5-28
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Figure 5-8. Schematic of Culvert Performance at Varying Stages and Alignments

Source: Furniss et al. (1998)

Large Wood National Manual July 2015


5-29
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

5.7.4 Additional Bridge and Pier Locations Out of Main Flow – If piers are
needed, they should not be placed at the channel
Culvert Design thalweg in the main flow location of the channel
Considerations or on the outside of meander bends where
debris tends to preferentially accumulate.
Traditional debris jam mitigation strategies
focus on bridge design, debris source control, Ample Vertical Clearance – Bridges and
and debris shedding/passage at structures. Many culverts over rivers with significant wood loads
debris problems at bridges and culverts can be should have ample vertical clearance above the
reduced through proper design of new design water profile to pass the material without
structures. Specific techniques for new bridges it hitting the top of the bridge or culvert.
include the following considerations.
No Pier Skew – Bridges that have an alignment
Wide Channel Span – Bridge and culverts skewed to the river should have piers aligned
should span the entire active bankfull channel parallel to the flow.
and minimize hydraulic dead zones behind
Scour Protection – Scour protection should be
embankments. Bridge abutments should not be
provided at piers and abutments where local
located in the channel. For small channels, single
scour is anticipated due to the anticipated wood
spans of up to 24 meters (80 feet) will eliminate
load and size of the structure opening.
the need for mid-channel piers. Where piers are
Replacement structures in high risk areas should
necessary due to wide channels, each span
have deep footings, piles, or armor to reduce the
between piers and between the end abutments
risk of structure damage.
and piers should exceed the length of expected
logs if possible. Debris deflection structures may be used, but
they require maintenance to keep clean and in
Channel Design – A compound channel cross-
functioning order. Debris deflectors are
section with two or three benches is common for
structures installed upstream of the bridge or
concentrating low and moderate flows to
culvert to redirect wood and ice away from
promote wood transport through the structure
bridges and culverts. Debris fins are one of the
while maintaining a wide floodplain to allow
more common devices located on the upstream
flow to spread out during large floods and
extension of bridge piers. They consist of a
deposit wood across the floodplain. A vegetated
sloping wall or series of piles that are at the
floodplain will trap wood away from hydraulic
leading edge of piers and are parallel to the flow.
openings at structures.
Freestanding piles or fenders are also used to
Overflow Structures – Wet and dry side redirect large wood. Debris racks can be used to
channels in floodplains should have structures in block wood accumulation from a structure.
addition to the main channel that pass under the Racks are porous steel bar or concrete structures
road embankment. Although prone to clogging that are inclined or skewed to deflect debris.
unless regular maintenance is performed, Proper structure design is preferred over debris
overflow structures minimize lateral and trapping structures.
converging flows and can reduce the wood load
to the primary hydraulic opening. Floodplain 5.8 Floods, Recovery,
equalization culverts can be used to convey
floodplain flows away from secondary and and Large Wood
tertiary channels, whether wet or dry, and can
Some of the information in this section has been
provide passage for terrestrial, amphibious, and
adapted from the Vermont Standard River
aquatic species at minimal risk to travelers.
Management Protocols (Schiff et al. 2014; see
http://www.anr.state.vt.us/dec/waterq/rivers).

Large Wood National Manual July 2015


5-30
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

5.8.1 Large Wood removal of large elements that provide hydraulic


friction, stabilizing bed features and dimensional
Assessment morphology. Large wood is essential to stabilize
After a flood, an evaluation of the load of large both the channel bottom and banks in most
wood must be completed within the context of channel types.
the watershed wood budget of source, transport, Removal of large wood also leads to loss of
and retention. A channel survey or flight to habitat, so limiting wood removal is
document the size and distribution of channel- recommended. Many channels with watersheds
spanning jams and smaller deposits should be having young forests lack regular inputs of large
performed. The following information is typically wood; thus, pulses delivered through material
collected to describe post-flood large wood contribution zones (Smith et al. 2008) or mass
accumulations: wasting events during floods are critical to long-
 Location in or near structure, channel, or term habitat formation and maintenance.
floodplain (GPS if possible).
 Photo-documentation. RISK
 Number of pieces and lengths.
Large wood accumulations should be assessed to see
 Dimensions, area, and volume (small, if they pose high, moderate, or low risk during future
medium, or large) (Bradley et al. 2005). floods (Homer et al. 2004). Do not default to
 Embeddedness. removing all of the large wood from post-flood
channels and floodplains as some likely does not pose
 Prediction of stability. a risk during future floods (Table 5-3). If the transport
 Distance and proximity to nearby property potential of wood is small, the material will remain in
and infrastructure. place over time and can have a positive effect on
habitat and stability with relatively low risk to
 The remaining load from accumulated or downstream infrastructure. If large wood is highly
sources or recent mass failures. mobile, it can lead to high risks downstream and
A stream walk and sketch with notes is helpful to should be removed. Tyler (2011) provides a flow chart
document the post-flood large wood load. A tally for predicting the likelihood of wood transport. When
form can be used to count and size large wood evaluating risk, consider that large wood mobilized in
pieces and jams (e.g., Schiff et al. 2008a). a flood can commonly be deposited on the nearest
downstream bar (Bertoldi et al. 2013). Cut large wood
on small bars that is likely to be highly mobile into
5.8.2 Large Wood small pieces so it may pass through downstream
Alternatives Analysis structures and remain in the ecosystem.

A primary objective of the post-flood assessment


is to develop a list of wood retention/removal Large jams that could lead to avulsion, bank
alternatives to avoid the unwarranted wholesale erosion, or clogging of downstream structures
removal of large wood that is often performed due to sudden release should be removed. Large
out of habit, fear of future damages, and ease of trees should be stockpiled for future habitat
construction. Large wood removal often leads to restoration work if storage space is available
long-term channel destabilization due to (NRCS 2007e).

Large Wood National Manual July 2015


5-31
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Table 5-3. Large Wood Removal Recommendations

Risk Level Risk Description Recommendation


HIGH Channel-spanning large wood Remove large wood jam.
jams with altered flow path and
high risk of avulsion.
Remobilization of large amounts
of wood and downstream
structure clogging likely. Structure
completely or mostly clogged.
HIGH TO MODERATE Large mid channel or bank Remove large wood.
accumulations of large wood. Flow
path may be altered, but risk of
avulsion is low. Remobilization of
a large amount of large wood and
downstream structure clogging
likely. Structure partially clogged.
MODERATE Large mid channel or bank Leave large wood in place.
accumulations of large wood. Flow
path may be altered, but risk of
avulsion is low. Remobilization of
a large amount of wood is not
likely.
LOW Bank accumulations of large wood Leave large wood in place.
jam or individual embedded
pieces of wood in channel. Flow
path may be altered, but risk of
avulsion is low. Remobilization of
large wood not likely.
Source: Adapted from Homer et al. (2004) and Schiff et al. (2014).

The combined understanding of the distribution  Selective removal of large wood jams that
of large wood (e.g., Magilligan et al. 2007), the are not embedded and likely to mobilize and
wood budget, the potential for additional create downstream flood and erosion risks.
transport to an area, the stability and roughness  Cutting of larger pieces to allow for safe
benefits offered by wood in channels and future passage through structures in the
floodplains, and the concurrent potential risks event of transport.
posed by the accumulations will allow for a
proper alternatives analysis. A range of large  Removal of all wood from the channel, but all
wood retention and removal alternatives should wood left on the floodplain to slow flood
be considered for design and implementation, flows, catch sediment, and create seed
and assessment should be in concert with propagule areas.
consideration of stream crossing retrofits and  Wholesale removal of wood from the
replacements to improve wood conveyance. channel and floodplain in areas with
 All wood left in place if little or no risk to abundant risk to bridges and culverts, or in
property or infrastructure exists. channels prone to avulsion into areas with
improved property.
 Selective removal of wood upstream of
bridges and culverts.

Large Wood National Manual July 2015


5-32
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

5.8.3 Large Wood Flood Standing and rooted trees in the river corridor
should not be removed following a flood. A
Recovery Design common misperception exists that these pose
uniform and hazardous threats during floods.
GUIDANCE Standing trees reduce bank erosion and typically
decrease flood risks by slowing flow velocity and
Goals of Large Wood Retention or Removal reducing erosion. They also catch sediment and
 Reduce flood and erosion risks downed wood and hold it on the floodplain
instead of allowing all material to deposit in the
 Improved long-term channel stability
channel.
 Maintain or improved instream habitat
Rootwads and tangles of large trees remaining
 Protect water quality on the banks should not be excavated. It is
Design Considerations for Large Wood Removal usually preferable to cut trees that must be
removed 2 to 3 meters (6 to 10 feet) above the
 Retain standing trees with intact roots on the
base of the trunks to remove only the upper
banks and floodplain.
sections. The remaining roots will hold the bank
 Minimize large wood removal to limit channel together. Minimize the use of large machinery in
and ecosystem impacts. Individual pieces, side the channel and the number of access points to
bar accumulations, or mid-channel accumulations control impacts.
of wood can remain that will not dictate
hydraulics or clog structures. Large wood should Large wood removal has historically been
be retained in the post-flood bankfull channel performed without proper design or
where possible. consideration of resultant flood and erosion
 Remove large or channel-spanning wood jams risks. There is a high degree of experience with
that alter flow path and increase the chance of removal of large wood, but the assessment,
avulsions or future clogging of the downstream alternatives analysis, and design approach
channel or structures. identified here will be new to most designers
 Retain large wood jams with limited risk of full and construction crews. A detailed plan review
movement such as those wedged on the during a preconstruction meeting and frequent
upstream end of islands or in large stable jams construction oversight at the beginning of the
(Ravazzolo et al. 2015). Wood on smaller bars is project are essential for proper implementation
likely to move in more frequent floods. Jams far of large wood removal, retention, or
from infrastructure can be left in place to decay re-utilization.
and disperse material downstream.
 Stockpile cleared trees with diameter >30
centimeters ( 12 inches ) for future use in channel
or floodplain stabilization or restoration. Intact
root balls are preferred.

Large Wood National Manual July 2015


5-33
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

GUIDANCE

New York Large Wood Removal Guidelines

The New York Department of Environmental Conservation (2014) provides recommendations for post-flood wood
management in rivers. They recognize that woody debris such as trees and branches are an important part of
healthy stream systems, providing habitat, roughness, energy dissipation, and slowing floodwater. Wood should be
left in place unless it endangers infrastructure.

The guidelines state that woody debris and trash can be removed from a stream without an Article 15 Protection
of Waters Permit due to lower risks of impacts under the following conditions.
 Fallen trees and debris may be pulled from a stream by vehicles and motorized equipment operating from the
top of streambanks using winches, chains, or cables.
 Handheld tools such as chainsaws, axes, hand saws, etc. may be used to cut debris into smaller pieces.
 Downed trees still attached to streambanks should be cut off near their stumps. Do not grub (pull out) tree
stumps from banks. Stumps keep streambanks from eroding.
 All trees, brush, and trash removed from a channel should be removed from the floodplain as well. Trash
should be properly disposed at a waste management facility. Trees and brush can be used as firewood. To
prevent the spread of invasive species such as the emerald ash borer, do not move firewood more than
80 kilometers (50 miles) from its point of origin. DEC has additional information on invasive insects.

Projects likely to disturb a streambed or banks and using motorized vehicular heavy equipment in the stream
channel or anywhere below the top of banks require either a Protection of Waters Permit or an Excavation or Fill in
Navigable Waters Permit.

distribution, and abundance of individual aquatic


5.9 Climate Change species. Average temperatures influence habitat
suitability for riverine species, and temperature
5.9.1 Climate-Driven extremes may limit species’ ranges due to
Processes Related to physical tolerances. Species distribution and
abundance also affects biotic interactions,
Large Wood including predator-prey and competition;
In riverine ecosystems, a variety of ecological disruption of these interactions can alter the
processes are influenced by climate. The climate ecosystem function and services.
can have direct temperature or precipitation
Flow regime in a river controls or influences
impacts, as well as indirect impacts through
many biological processes, such as reproduction
effects on terrestrial inputs (e.g., large wood,
and migration, as well as the physical
nutrients, and sediment), structure and species
environment (Wenger et al. 2011). Streamflow
composition, and disturbance regimes (see
also influences water temperature, nutrient
Figure 5-9). Different aspects of climate, such as
concentrations, and sediment. Streamflow
climate averages, extremes, or variability, may
quantity impacts the extent of available aquatic
be relevant to particular ecological processes.
habitat, while variability in streamflow regulates
Atmospheric temperature influences water species abundance and persistence.
temperature, which has direct effects on aquatic
species. Temperature regulates metabolism,

Large Wood National Manual July 2015


5-34
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Figure 5-9. Pathways by Which Climate Change which leads to large wood inputs to riverine
May Alter Stream Ecosystem Structure and systems. Extreme precipitation events can
Function trigger landslides, especially when following a
wildfire (Cannon and DeGraff 2009). Change in
average temperature and increases in minimum
temperatures are known to control insect
outbreaks, such as with bark beetles in boreal
forests (Bentz et al. 2010), which also influence
large wood inputs. Extreme winds are also
known to contribute large wood to riverine
systems (CCSP 2008b). Nutrients carried by
runoff following average and extreme
precipitation events impact riverine systems,
and may lead to an oversupply of nutrients that
leads to spread of harmful invasive species or
decreased habitat quality. Sediment inputs from
surface runoff and high-flow events also impact
Climate change will have direct effects on riverine systems, altering water quality.
temperature and precipitation regimes, which will
The structure and species composition in the
influence indirect effects of terrestrial disturbance
that will alter terrestrial ecosystem structure. terrestrial environment is influenced by climate
Changes in terrestrial ecosystem structure will in and is important to the potential for large wood
turn alter terrestrial inputs to streams. Source: inputs into riverine systems. Structure of
Davis et al. (2013). watershed vegetation and species composition
Precipitation is central to flow regime, directly also affect the infiltration of water, soil moisture,
through surface runoff and rainfall inputs, as groundwater recharge, and erosion patterns
well as through soil moisture and groundwater (Davis et al. 2013). Average temperature,
recharge in a watershed. Extreme high average precipitation, seasonality and timing of
precipitation is directly related to flooding and seasons, and disturbance regimes can all
high flow rates. Extreme low precipitation can influence the terrestrial structure and species
also be a primary driver of drought, leading to composition in a watershed.
extreme low flows in aquatic systems, stressing
organisms. Extreme precipitation can be defined 5.9.2 Recent and Future
many ways, such as a total threshold amount in Climate Change
an event (i.e., 50 millimeters [2 inches]) or rate
(i.e. 10 millimeters/hour [0.4 inches/hour]) or Climate change refers to any significant change
relative to historical averages (i.e., greater than in the measures of climate lasting for an
90% or less than 10% of all other events in the extended period of time (usually decades or
historical record). longer), including major changes in temperature,
precipitation, or wind patterns, among other
Terrestrial disturbances in a watershed are effects. Projected changes in such climatic
connected to the processes in the aquatic variables will potentially impact ecological
environment and affect terrestrial inputs. processes related to large wood. Changes may be
Climate plays an important role in terrestrial related to long-term averages, variability, and
disturbance, including wildfires, landslides, extreme events. The characteristics of extreme
drought, and insect outbreaks. Extreme high events vary from place to place, but are generally
temperatures and extreme low precipitation defined as rare occurrences within a statistical
contribute to wildfire frequency and intensity, reference distribution (e.g., rarer than the 10th

Large Wood National Manual July 2015


5-35
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

or 90th percentile) at a particular place (IPCC applied (planners and decision-makers get
2007). While a single extreme event may not be information) and available (information
directly attributed to climate change, the long- providers collect and create useful results).
term trend of more frequent and severe weather
While many of the connections between
events is projected under climate change for the
ecosystems and climate are well understood,
United States (USGCRP 2009), and
there is uncertainty in projecting future climate
decisionmakers should consider this trend to
conditions and ecosystem response. Sources of
adequately prepare for future risks to
uncertainty include climate model uncertainty,
ecosystems and infrastructure from large wood
the future greenhouse gas concentration
in streams.
pathway, climate sensitivity, novel ecosystem
A variety of climate information may be relevant conditions or thresholds, and biotic interactions
to decisions regarding large wood in restoration and feedbacks. Each of these sources of
projects. This information spans a range of time uncertainty is an active area of research.
scales and different geographies, and can include Communication with information providers and
historical and current observations, and modeled other stakeholders can improve the
projections of possible future conditions for understanding of these uncertainties and how to
variables such as temperature and precipitation. address them, if necessary. This dialogue can
Projections of future climate over the next also help build capacity to update and adapt
century are not precise forecasts; considering plans over time as conditions change or new
output from multiple models, over multiple information becomes available.
scenarios and time periods provides insight into
Recent changes in temperature have been
the range of possible future conditions (see
observed in many parts of the United States, with
Knutti et al. 2010).
water temperature increasing in some rivers
Data on possible future climate conditions are (Kaushal et al. 2010). Average temperatures are
maintained and collected by numerous expected to increase across the United States by
institutions including universities, federal and the later part of the century (2071–2099)
state climate agencies, and other organizations relative to historical (1970–1999) averages (see
working on climate change. Building a formal or Figure 5-10). While temperature increases have
informal network that includes decision-makers been observed in all parts of the country,
and information providers can help improve the regional variations in the magnitude of warming
ability to plan for changes in climate and are projected (Melillo et al. 2014).
extreme events. While many different types of
Changes in climate are already altering the water
climate information exist, restoration planners
cycle in multiple ways over different geographic
and decision-makers should focus on the specific
areas and time scales (Mellilo et al. 2014). By the
decision of interest to guide their use of this
middle of the century (2040–2070), changes in
information. For example, in a remote area of
runoff and related river-flow are projected in
low-value habitat where a rapid assessment is
many parts of the United States relative to the
needed, knowing if the area is likely to become
historical patterns (1971–2000). Spring runoff is
wetter or drier in the future may be sufficient to
projected to decline in the southwest and
inform planning. In other cases, such as where
southern Rockies (e.g., the Rio Grande and
there are species of concern and critical human
Colorado River basins) and southeast (Figure
infrastructure, more detailed information on
5-11; U.S. Department of Interior—Bureau of
precipitation, temperature, and extreme flow
Reclamation 2011). In many cases, the projected
conditions and trends may be necessary. A two-
changes in streamflow are outside the range of
way dialogue with information providers can
historical variability (Melillo et al. 2014). In the
help ensure that the best available information is
northwest to north-central United States, basins

Large Wood National Manual July 2015


5-36
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

like the Columbia River and Missouri River are end of the century. Soil moisture is also
projected to see little change by the middle of projected to decrease across the southwest,
this century, with some potential increase by the which would impact flow regime.

Figure 5-10. Projected Temperature Change by 2071–2099

The largest uncertainty in projecting climate change beyond the next few decades is the level of heat-trapping
gas emissions. Results are shown for four Representative Concentration Pathways (RCP): a low scenario that
assumes rapid reductions in emissions (RCP 2.6); a high scenario that assumes continued increases in
emissions (RCP 8.5) and the corresponding greater amount of warming; an intermediate scenario RCP 4.5;
and RCP 6.0. Projections show change in average temperature in the later part of this century (2071–2099)
relative to the late part of last century (1970–1999). Source: Melillo et al. (2014).

Large Wood National Manual July 2015


5-37
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Figure 5-11. Streamflow Projections for River Basins in the Western United States

Annual and seasonal streamflow projections are based on possible climate scenarios for eight river basins in
the western United States The panels show percent change in average runoff. Projections are for annual, cool,
and warm seasons, for three future decades (2020s 2050s, and 2070s) relative to the 1990s. Source: Melillo
et al. (2014) after U.S. Department of the Interior – Bureau of Reclamation (2011).

In mountain watersheds, warming under some driven by precipitation, as pre-existing soil


scenarios is projected to directly affect flooding moisture conditions, topography, and other
even in places that are not projected to have an factors, including human-caused changes,
overall increase in precipitation, due to more influence flooding in a basin-specific manner
precipitation falling as rain rather than snow, or (e.g., Poff et al. 2006).
more rain falling on the snowpack (Knowles et
The extent of area burned in wildfires in the
al. 2006). Rainfall on snowpack can also affect
continental United States has significantly
the timing of peak runoff. In much of the western
increased from 1987–2003 compared to the
United States, earlier peak river levels have
period from 1970–1986 (Westerling et al. 2006),
already been observed in snowmelt-fed rivers
due to multiple factors, including management
(see Figure 5-12), which are related to rain on
practices that allowed fuel build-up and changes
snow, as well as other influences like dust and
in temperature and precipitation in some
soot on snowpack and natural variability
regions (Melillo et al. 2014). Projections of future
(Creamean et al. 2013). River flooding is not only
wildfires are difficult due to regional differences

Large Wood National Manual July 2015


5-38
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

in seasonality of fire activity. There is a direct Drought is expected to intensify in many parts of
connection between land surface drying due to the United States due to longer periods of dry
increases in temperature and increases in the weather and more extreme heat (Melillo et al.
size and intensity of wildfires (Westerling et al. 2014). More intense drought would lead to more
2003). moisture loss from plants, potentially affecting
the risk of wildfire and large wood inputs to
Increases in wildfire activity, particularly in the
riverine systems. Long-term drought conditions
western United States, are correlated with
(multi-season) are projected to increase in part
earlier snowmelt, longer growing seasons, and
of the southeast United States (Melillo et al.
higher summer temperatures (Westerling et al.
2014).
2006).

Figure 5-12. Changes in Timing of Streamflow 5.9.3 Potential Climate


from Snowmelt Change Impacts on the
Riverine Environment
and Built Infrastructure
The water cycle is dynamic, and riverine
ecosystems are able to maintain a healthy and
self-sustaining condition in the face of large year-
to-year variation in temperature conditions.
However, in many places the range of projected
changes in climate may lead to impacts that
exceed the natural resilience of these ecosystems
(CCSP 2008b). If the rate of climate change
outpaces the ability of plant and animal species
to adjust to temperature changes, population
loss or extinction may result (Loarie et al. 2009).

Habitat fragmentation, pollution, increased


urbanization, and other stressors will interact
with climate change, exacerbating the level of
impact (CCSP 2008b). In the western United
States, this includes increasing vulnerability due
to projected increases in drought and water
shortages (Falke et al. 2011). In southern states,
Red dots indicate stream gauge locations where projections are less certain but include
half of the annual flow is arriving from 5–20 days potentially drier conditions, and the impacts may
earlier (2001–2010 compared to 1951–2000 be greater due to interactions with projected
average). Blue dots indicate locations where annual increases in water withdrawals (Melillo et al.
flow is arriving later (2001–2010 compared to
2014). Irruptions of forest pests have increased
1951–2000 average). Crosses are locations where
observed changes are not significantly different tree mortality in many regions and may affect
from the past century baseline (90% confidence wood loading rates significantly.
level); dots and diamonds indicate where timing is
different (95% and 90% confidence level, The combined impacts of projected climate
respectively). Source: Melillo et al. (2014, page 768, change and water withdrawals can lead to
Appendix 3). habitat loss and local extinctions of fish and
aquatic species (Spooner et al. 2011). Climate

Large Wood National Manual July 2015


5-39
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

impacts that result in changes to vegetation or biotic interactions in changes in species


stream hydrology will likely also affect instream distribution in response to changes in climate
wood, and the potential for small streams to (Wenger et al. 2011).
form dynamic aquatic habitat (Hough-Snee et al.
The projected possible changes in timing of peak
2014). For example, trout habitat in the interior
flows and runoff conditions can impact riverine
western United States is projected to decrease by
ecosystems, as well as the operation,
2080 under several different possible future
maintenance, and service of water management
climate scenarios (Wegner et al. 2011; see Figure
infrastructure. Water management plans and
5-13). The projected impact on four trout species
policies designed to provide adequate service
is driven not only by changes in water
and meet regulation for human use and
temperature beyond physiological optima, but
environmental flows may not be adapted to
also potential shifts in flow regime and biotic
possible future conditions (EPA 2013). In
interactions, which are projected to change with
addition, ecosystems provide important services
climate (Wegner et al. 2011).
for improving water quality and regulating water
It is likely that temperature, flow regime, and flows, which are projected to be impacted by
biotic interactions will have a strong influence projected changes in climate (Melillo et al. 2014).
on changes in species distribution in response to
climate change (Wenger et al. 2011). However,
additional research is needed on the role of

Figure 5-13. Occurrence Probability of Trout Species as a Function of Air Temperature and Winter High
Flow Frequency

Projected loss of suitable habitat is driven by possible changes in climate that impact temperature and winter
flooding (caused by warmer, rainier winters). Green indicates cutthroat trout; blue indicates brook trout; red
indicates rainbow trout; and brown indicates brown trout. Source: Wegner et al. (2011).

Large Wood National Manual July 2015


5-40
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Climate change, in combination with existing Insect outbreaks, driven at least in part by
environmental stressors, is overwhelming the changes in climate, are already having a
capacity of some ecosystems to recover from significant impact on forests in the United States.
impacts from major disturbances, such as In particular, bark beetles have damaged boreal
wildfires, floods, and storms (Melillo et al. 2014). conifer forests in the western United States, with
Projected changes in climate may alter higher temperatures allowing more beetles to
disturbance regimes, such as fire, landslides, and survive the winter and to extend their range to
insect outbreaks (CCSP 2008b), which can alter higher elevations and more northern latitudes
the terrestrial inputs into riverine ecosystems. (Raffa et al. 2008), such as new areas in the
Reductions in forest cover or leaf area due to Greater Yellowstone Ecosystem (Logan et al.
disturbances will likely alter the hydrology in a 2010). The damage to forest areas can alter fire
watershed (CCSP 2008b), which will impact regimes and terrestrial inputs of large wood into
riverine ecosystems. riverine systems. Insect outbreaks can also
increase base flows and advance the timing of
Potential increases in flood magnitude or
peak runoff, resulting in impacts on riverine
frequency could lead to impacts on terrestrial
ecosystems (CCSP 2008b).
inputs to riverine systems. For example, more
intense overbank flooding may change patterns The increased intensity in individual
of sediment erosion and deposition, resulting in precipitation events will likely affect
transitions in riparian vegetation from large transportation and stormwater infrastructure.
long-lived conifer trees to early-successional Bridges, culverts, and other stormwater
shrubs that do not contribute large wood to the infrastructure will be vulnerable to the impacts
riverine ecosystem (Hough-Snee et al. 2014). of precipitation and flooding from higher water
The flood regime also alters the stream power levels, increased flows, scour, sedimentation, etc.
and channel geometry, and potential changes to (CCSP 2008a). Runoff resulting from such events
the hydrologic regime would therefore affect could lead to increased peak streamflow, which
wood mobility (Hough-Snee et al. 2014). could affect the sizing requirement for bridges
and culverts (CCSP 2008a). Historically, bridges
Possible changes in fire regime may lead to
and culverts have not been designed well enough
transition in forest vegetation toward early-seral
to convey sediment and large wood, much less
species, altering the contribution of large wood
deal with increased flood peaks (see Figure
to channels (Hough-Snee et al. 2014). Intense
5-14). Both disaster planning and restoration
forest fires can also increase sediment
efforts should consider replacing inadequately
production and water yield as much as 10 to
sized stream crossings and restoring riparian
1,000 times (CCSP 2008b), impacting the
forests and stable instream large wood to
riverine environment.
attenuate flood peaks.
Changing climatic conditions, along with other
The accumulation of large wood and
drivers of change, can impact distribution and
transportation of material downstream can pose
success of invasive species in a watershed (CCSP
risks to infrastructure. Large wood is a concern
2008b). The ability to outcompete in novel
for highway engineering planning because it can
climate conditions will lead to altered forest
accumulate at and obstruct the waterway
stand composition. The species composition
entrance of culverts or bridges, adversely
controls aspects of terrestrial inputs, including
affecting the operation of the structure or
large wood, to riverine systems. Invasive species
causing failure of the structure.
may also alter watershed erosion regimes due to
shallow root systems, altering sediment inputs,
as well.

Large Wood National Manual July 2015


5-41
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Figure 5-14. Wood Inhibiting the Flow of Water in sediment load due to indirect effects of climate
through a Culvert under Highway 4 Following the change can be modulated by large wood, which
Las Conchas, New Mexico Fire (2011) is known to effectively trap sediment. The
increase in connectivity between rivers and
groundwater (i.e., hyporheic exchange) due to
large wood may provide a buffer to riverine
habitats against nutrient loading and thermal
impacts from climate change (Sawyer et al.
2011).

Large wood also influences the biological


condition of riverine ecosystems, which can
reduce the vulnerability to climate change. Large
wood placements can improve biological
structure and ecosystem productivity, improving
resilience of these systems to indirect impacts
from changes in terrestrial disturbance. Complex
Source: Jake Quintana, USFS. cover provided for aquatic organisms and
improved water quality also can improve habitat
5.9.4 Large Wood quality, reducing some of the impact from
climate change, such as higher temperatures and
Contribution to changes in species composition.
Reducing Climate
Restoring stream corridors with mature timber
Vulnerabilities in helps attenuate the effects of fires and debris
Riverine Ecosystems flows by trapping sediment and debris before it
reaches areas with infrastructure. Trees in
and Built Infrastructure riparian areas tend to be more resistant to fires
Many of the important roles that large wood because they have higher soil moisture; in many
debris plays in riverine ecosystems also can recent western fires, riparian corridors acted as
reduce the vulnerability of riverine ecosystems critical fire breaks. Fires tend to burn tree
to impacts from climate change. Large wood canopies and leave most of the trunk, so if trees
influences the physical condition of riverine are mature they are more likely to provide stable
ecosystems, including the temperature, wood to the channel.
hydrology, and sediment load. Potential
Heavy precipitation events can increase the flow
increases in water temperatures with future
velocity and flow depth of a stream or river,
climate change can be moderated by large wood
which can affect local scour depth. During flood
placements. The effects of large wood in raising
conditions, if the stream elevation reaches the
local water elevations, scouring bed, and creating
low cord bridge elevation, the local scour depths
low-velocity refugia can provide protection to
could be increased by 200 to 300%. Using stable
aquatic species and habitat against the possible
large wood placements will help to restore
increase in duration and intensity of drought in
streams, increase flow resistance, partition shear
some regions, like the southwestern United
stress, and slow flood peaks (Fischenich and
States Increases in floodplain connectivity due to
Morrow 2000; Anderson 2006; Abbe and Brooks
the presence of large wood in rivers may also
2011), protecting downstream bridges.
increase resilience of aquatic species to direct
impacts of climate change on habitat quality Large wood in channels can be maintained by
within certain portions of a watershed. Increases enlarging infrastructure (e.g., culverts and

Large Wood National Manual July 2015


5-42
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

bridges) to allow large wood passage


downstream and reduce risks to downstream
5.10 Conclusion
infrastructure. Designing bridges and culverts to Common themes in this chapter include
withstand more frequent and severe storm uncertainty, change, variation within basins and
events (e.g., 500-year events rather than 50-year between basins, variation within
events) can allow for enhanced passage of large hydrophysiographic region and between regions,
wood under and through this infrastructure. The and other caveats and frustratingly
American Association of State Highway and unpredictable dynamics and interactions.
Transportation Officials Load and Resistance Nonetheless, certain patterns and challenges
Factor Design specifications require that scour at emerge and become clear. Returning integrity
bridge foundations be designed for the 100-year and resilience to our aquatic and riparian
flood, while some bridges should be designed to ecosystems requires consideration of all
withstand the 500-year flood (the “super flood”) regimes, including wood. Restoration must
(Ghosn et al. n.d.). develop further expertise in assessing the
efficacy of large scale, long term, passive
In addition to changing the design of
processes and integrating them where necessary
infrastructure, measures can also be taken to
with more expensive and failure-prone
reduce the impact of large wood on bridges and
engineered approaches. The limits of society’s
culverts. Both structural and nonstructural
limited perspective, and the mixed performance
measures can be used to mitigate the effects of
of our decision-making and management
large wood and protect infrastructure (see Table
institutions in protecting residents,
5-2). Structural measures either prevent debris
infrastructure, economies, and ecosystems from
from entering or blocking passageways, or assist
the ravages of large floods are well recognized,
in the passage of debris through the
but agencies, stakeholders, and the restoration
passageways. Nonstructural measures include
community can and must do better in the
regular maintenance and clearing of debris
contexts of increasingly developed river
during extreme events (FHWA 2005). Many of
corridors, the various predicted effects of climate
these measures are used to protect
change, and increasing recognition of the
infrastructure from current conditions. These
ecosystem values and services of watersheds.
types of measures may be increasingly used as
Managers, restoration practitioners, and
conditions change; therefore, decision-makers
stakeholders must plan for future scenarios and
should consider climate change and extreme
not historical norms. This chapter offers some
events in future applications of these measures.
powerful initial principles, concepts, and tools to
achieve these goals.

Large Wood National Manual July 2015


5-43
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

5.11 Uncertainties and Research Needs


1. The subjects addressed in this chapter, by their large scale and long term nature, all represent
domains of uncertainty.
2. The transport dynamics of pulsed wood inputs from stochastic events are variable and largely
unstudied.
3. The roles of thresholds and multiple or alternative stable states, in hydrologic, geomorphic, and
ecological terms, remain largely unstudied.
4. Future peak flow hydrology as the driving variable for multiple processes and concerns,
particularly in the context of climate change, is uncertain in terms of precise quantification but
poses significant concerns.
5. Secondary effects of climate change, such as vegetative stress induced by base flow alteration as
well as forest disease or insect issues, are largely unpredictable but have already induced
dramatic impacts in some settings.

5.12 Key Points


1. The capacity of the watershed to produce a large wood supply of appropriate volume and size
range and deliver it to the channel network is ultimately more significant and cost-effective than
engineered wood features.
2. The trapping and transport roles played by the largest wood pieces (relative to channel
geometry) are pivotal.
3. Similarly, the ability of the system to convey mobile wood elements is critical to mid-basin,
lower basin, and terminus supply. This includes natural supply and recruitment as well as large
wood management at dams and channel crossings that recognizes and addresses the
importance of wood to downstream reaches.
4. The risks of flooding and structural damage at stream and river crossings are most effectively
addressed by crossing retrofits and redesigns, offering long-term economic, public safety, and
ecological benefits—particularly in view of peak flow increases predicted by climate change
scientists.
5. Although pulsed wood inputs and jams created by stochastic floods must sometimes be
managed through removal, this should not be the default response. Large wood offers important
stability and habitat values. If removal is required, it should be retained for channel and
floodplain restoration use elsewhere.
6. The patterns and trends of climate change remain unpredictable at quantifiable and local scales,
along with associated hydrologic and ecological responses (e.g., disease and insect outbreaks, or
the effects of altered base flows on vegetative stress), but significant impacts have been
observed in multiple locations. The irruption of mountain pine bark beetle in many locations in
the American west is one example.

Large Wood National Manual July 2015


5-44
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

5.13 References
Abbe, T. B., and A. P. Brooks. 2011. Geomorphic, Engineering, and Ecological Considerations when
Using Wood in River Restoration. Pages 419–451 in A. Simon, S. J. Bennett, and J. M. Castro
(eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools.
Geophysical Monograph Series 194. Washington, D.C.: American Geophysical Union.

Agrawal, A., M. A. Khan, and Z. Yi. 2007. Handbook of Scour Countermeasures and Design. FHWA-NJ-
2005-027. New Jersey Department of Transportation and Federal Highway Administration,
Washington, DC.

Allan, J. D. 1995. Stream Ecology: Structure and Function of Running Waters. Boston, MA: Kluwer
Academic Publishers.

Anderson, D. B. 2006. Quantifying the Interaction between Riparian Vegetation and Flooding: from
Cross-Section to Catchment Scale. University of Melbourne.

Angradi, T. R., E. W. Schweiger, D. W. Bolgrien, P. Ismert, and T. Selle. 2004. Bank Stabilization,
Riparian Land Use and the Distribution of Large Woody Debris in a Regulated Reach of the
Upper Missouri River, North Dakota, USA. River Research and Applications 20:829–846.

Arnáez, J., V. Larrea, and L. Ortigosa. 2004. Surface Runoff and Soil Erosion on Unpaved Forest Roads
from Rainfall Simulation Tests in Northeastern Spain. Catena 57:1–14.

Arneson, L. A., L. W. Zevenbergen, P. F. Lagasse, and P. E. Clopper. 2012. Evaluating Scour at Bridges.
Hydraulic Engineering Circular 18, FHWA-HIF-12-003, National Highway Institute, Federal
Highway Administration, Arlington, VA.

Beckman, N. D., and E. Wohl. 2014. Carbon Storage in Mountainous Headwater Streams: The Role of
Old-Growth Forest and Logjams. Water Resources Research 50:2376–2393.

Beechie, T. J., G. Pess, P. Kennard, R. E. Bilby, and S. Bolton. 2000. Modeling Recovery Rates and
Pathways for Woody Debris Recruitment in Northwestern Washington Streams. North American
Journal of Fisheries Management 20:436–452.

Benda, L. E., and J. C. Sias. 2003. A Quantitative Framework for Evaluating the Mass Balance of In-
Stream Organic Debris. Forest Ecology and Management 172:1–16.

Benda, L., Miller, D., Bigelow, P., Andras, K. 2003a. Effects of Post-Wildfire Erosion on Channel
Environments, Boise River, Idaho. Forest Ecology and Management 178:105–119.

Benda, L., D. Miller, J. Sias, D. Martin, R. Bilby, C. Veldhuisen, and T. Dunne. 2003b. Wood Recruitment
Processes and Wood Budgeting. American Fisheries Society Symposium 37:49–73.

Benke, A. C. 2001. Importance of Flood Regime to Invertebrate Habitat in an Unregulated River-


Floodplain Ecosystem. Journal of the North American Benthological Society 20:225–240.

Benke, A. C., and J. B. Wallace. 1990. Wood Dynamics in Coastal Plain Blackwater Streams. Canadian
Journal of Fisheries and Aquatic Sciences 47:92–99.

Large Wood National Manual July 2015


5-45
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Bentz, B. J., J. Régnière, C. J. Fettig, M. Hansen, J. L. Hayes, J. A. Hicke, R. G. Kelsey, J. F. Negrón, and S. J.
SeyboLd. 2010. Climate Change and Bark Beetles of the Western United States and Canada:
Direct and Indirect Effects. BioScience 60:602–613.

Bertoldi, W., A. M. Gurnell, and M. Welber. 2013. Wood Recruitment and Retention: The Fate of
Eroded Trees on a Braided River Explored Using a Combination of Field and Remotely-Sensed
Data Sources. Geomorphology 180–181(0):146–155.

Bilby, R. E., and G. E. Likens. 1980. Importance of Debris Dams in the Structure and Function of
Stream Ecosystems. Ecology 61:1107–1113.

Bilby, R. E. and J. W. Ward. 1991. Characteristics and Function of Large Woody Debris in Streams
Draining Old-Growth, Clear-Cut, and Second-Growth Forests in Southwestern Washington.
Canadian Journal of Fisheries and Aquatic Sciences 48:2499–2508.

Bocchiola, D., M. C. Rulli, and R. Rosso. 2006. Transport of Large Woody Debris in the Presence of
Obstacles. Geomorphology 76(1):166–178.

Bocchiola, D., M. C. Rulli, and R. Rosso. 2008. A Flume Experiment on the Formation of Wood Jams in
Rivers. Water Resources Research 44: W02408, doi:10.1029/2006WR005846.

Bradley, J., D. Richards, and C. Bahner 2005. Debris Control Structures – Evaluation and
Countermeasures. Hydraulic Engineering Circular No. 9. FHWA-IF-04-016. U.S. Department of
Transportation, Federal Highway Administration., Salem, OR.

Bragg, D. C. 2000. Simulating Catastrophic and Individualistic Large Woody Debris Recruitment for a
Small Riparian Ecosystem. Ecology 81:1383–1394.

Braudrick, C. A., and G. E. Grant. 2000. When Do Logs Move in Rivers? Water Resources Research
36(2):571–583.

Braudrick, C. A., and G. E. Grant. 2001. Transport and Deposition of Large Woody Debris in Streams:
A Flume Experiment. Geomorphology 41:263–283.

Braudrick, C. A., G. E. Grant, Y. Ishikawa, and H. Ikeda. 1997. Dynamics of Wood Transport in
Streams: A Flume Experiment. Earth Surface Processes and Landforms 22:669–683.

Brenkman, S., J. Duda, C. E> Torgersen, E. Welty, G. R. Pess, R. Peters, and M. L. McHenry. 2012. A
Riverscape Perspective of Pacific Salmonids and Aquatic Habitats Prior to Large-Scale Dam
Removal in the Elwha River, Washington, USA. Fisheries Management and Ecology 19:36–53.

Brooks, A. P., G. J. Brierly, and R. G. Millar. 2003. The Long-Term Control of Vegetation and Woody
Debris on Channel and Flood-Plain Evolution: Insights from a Paired Catchment Study in
Southeastern Australia. Geomorphology 51:7–30.

Brooks, A. P., T. Abbe, T. Cohen, N. Marsh, S. Mika, A. Boulton, T. Broderick, D. Borg, and I. Rutherfurd
2006b. Design Guidelines for the Reintroduction of Wood into Australian Streams. Land & Water
Australia, Canberra, Australia.

Buckley B. M., and F. J. Triska 1978. Presence and Ecological Role of Nitrogen-Fixing Bacteria
Associated with Wood Decay in Streams. Internationale Vereinigung für Theoretische und
Angewandte Limnologie Verhandlungen 20:1333–1339.

Large Wood National Manual July 2015


5-46
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Buffington, J. M., T. E. Lisle, R. D. Woodsmith, and S. Hilton. 2002. Controls on the Size and
Occurrence of Pools in Coarse-Grained Forest Rivers. River Research and Applications 18:507–
531.

Cannon, S. H., and J. DeGraff. 2009. The Increasing Wildfire and Post-Fire Debris-Flow Threat in
Western USA, and Implications for Consequences of Climate Change. Pages 177–190 in K. Sassa
and P. Canuti (eds.), Landslides—Disaster Risk Reduction. Berlin Heidelberg: Springer-Verlag.
Available: <http://landslides.usgs.gov/docs/cannon/Cannon_Degraff_2008_Springer.pdf>.

Carlson, J. Y., C. W. Andrus, and H. A. Froehlich. 1990. Woody Debris, Channel Features, and
Macroinvertebrates of Streams with Logged and Undisturbed Riparian Timber in Northeastern
Oregon, USA. Canadian Journal of Fisheries and Aquatic Sciences 47:1103–1111.

Collins, B. D., and D. R. Montgomery. 2002. Forest Development, Wood Jams, and Restoration of
Floodplain Rivers in the Puget Lowland, Washington. Restoration Ecology 10:237–247.

Collins, B. D., D. R. Montgomery, and A. D. Haas. 2002. Historical Changes in the Distribution and
Functions of Large Wood in Puget Lowland Rivers. Canadian Journal of Fisheries and Aquatic
Sciences 59:66–76.

Collins, B. D., D. R. Montgomery, K. L. Fetherston, and T. B. Abbe. 2012. The Floodplain Large-Wood
Cycle Hypothesis: A Mechanism for the Physical and Biotic Structuring of Temperate Forested
Alluvial Valleys in the North Pacific Coastal Ecoregion. Geomorphology 139/140:460–470.

Comiti, F., A. Andreoli, M. A. Lenzi, and L. Mao. 2006. Spatial Density and Characteristics of Woody
Debris in Five Mountain Rivers of the Dolomites (Italian Alps). Geomorphology 78:44–63.

Cook, W. J. 2014. Bridge Failure Rates, Consequences, and Predictive Trends. PhD Dissertation. Utah
State University, Logan, UT.

Creamean, J. M., K. J. Suski, D. Rosenfeld, A. Cazorla, P. J. DeMott, R. C. Sullivan, A. B. White, F. M.


Ralph, P. Minnis, J. M. Comstock, J. M. Tomlinson, and K. A. Prather. 2013. Dust and Biological
Aerosols from the Sahara and Asia Influence Precipitation in the Western U.S. Science 339:
1572–1578, doi:10.1126/ science.1227279.

Curran, J. H., and E. E. Wohl 2003. Large Woody Debris and Flow Resistance in Step-Pool Channels,
Cascade Range, Washington. Geomorphology 51:141–157.

Daniels, M. D., and B. L. Rhoads. 2004. Effect of Large Woody Debris Configuration on Three-
Dimensional Flow Structure in Two Low-Energy Meander Bends at Varying Stages. Water
Resources Research 40: W11302, doi:10.1029/2004WR003181.

Daniels, M. D., and B. Rhoads. 2007. Influence of Experimental Removal of Large Woody Debris on
Spatial Patterns of Three-Dimensional Flow in a Meander Bend. Earth Surface Processes and
Landforms 32:460–474.

Davis, J. M., C. V. Baxter, E. J. Rosi-Marshall, J. L. Pierce, and B. T. Crosby. 2013. Anticipating Stream
Ecosystem Responses to Climate Change: Toward Predictions that Incorporate Effects via Land–
Water Linkages. Ecosystems 16:909–922. DOI:10.1007/s10021-013-9653-4.

Dean, D. J., M. L. Scott, P. B. Shafroth, and J. C. Schmidt. 2011. Stratigraphic, Sedimentologic, and
Dendrogeomorphic Analyses of Rapid Floodplain Formation Along the Rio Grande in Big Bend
National Park, Texas. Geological Society of America Bulletin 123:1908–1925.

Large Wood National Manual July 2015


5-47
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Diehl, T. H. 1997. Potential Drift Accumulation at Bridges. Publication FHWA-RD-97-028. U.S.


Department of Transportation, McLean, VA.

Downs, P.W., and A. Simon. 2001. Fluvial Geomorphological Analysis of the Recruitment of Large
Woody Debris in the Yalobusha River Network, Central Mississippi, USA. Geomorphology 37: 65-
91.

Doyle, M. W., E. H. Stanley, J. M. Harbor, and G. S. Grant, G.S. 2003. Dam Removal in the United States:
Emerging Needs for Science and Policy. Eos, Transactions American Geophysical Union 84:29.

Draut, A. E., J. B. Logan, and M. C. Mastin, M.C. 2011. Channel Evolution on the Dammed Elwha River,
Washington, USA. Geomorphology 127:71–87.

Dunkerley, D. 2014. Nature and Hydro-Geomorphic Roles of Trees and Woody Debris in a Dryland
Ephemeral Stream: Fowlers Creek, Arid Western New South Wales, Australia. Journal of Arid
Environments 102:40-49.

Dwire, K. A., and J. B. Kauffman. 2003. Fire and Riparian Ecosystems in Landscapes of the Western
USA. Forest Ecology and Management 178:61–74.

East, A. E., G. R. Pess, J. A. Bountry, C. S. Magirl, A. C. Ritchie, J. B. Logan, T. J. Randle, M. C. Mastin, J. T.


Minear, J. J. Duda, M. C. Liermann, M. L. McHenry, T. J. Beechie, and P. B. Shafroth. 2014. Large-
Scale Dam Removal on the Elwha River, Washington, USA: River Channel and Floodplain
Geomorphic Change. Geomorphology 228:765–786.

Eaton, B. C., M. A. Hassan, and S. L. Davidson. 2012. Modeling Wood Dynamics, Jam Formation, and
Sediment Storage in a Gravel-Bed Stream. Journal of Geophysical Research 117:F00A05,
doi:10.1029/2012JF002385.

Falke, J. A., K. D. Fausch, R. Magelky, A. Aldred, D. S. Durnford, L. K. Riley, and R. Oad. 2011. The Role
of Groundwater Pumping and Drought in Shaping Ecological Futures for Stream Fishes in a
Dryland River Basin of the Western Great Plains, USA. Ecohydrology 4L682–697. doi:10.1002/
eco.158. Available: http://onlinelibrary.wiley.com/doi/10.1002/eco.158/pdf.

Faustini, J. M., and J. A. Jones. 2003. Influence of Large Woody Debris on Channel Morphology and
Dynamics in Steep, Boulder-Rich Mountain Streams, Western Cascades, Oregon. Geomorphology
51:187–205.

Federal Emergency Management Agency (FEMA). 1999. Riverine Erosion Hazard Areas; Mapping
Feasibility Study. FEMA Technical Services Division, Hazard Study Branch.

Federal Highway Administration (FHWA) 1985b. Hydraulic Design of Highway Culverts. FHWA-IP-
58-15. Available: http://www.fhwa.dot.gov/engineering/hydraulics/pubs/12026/hif12026.pdf.

Federal Highway Administration (FHWA). 2005. Debris Control Structures Evaluation and
Countermeasures. Hydraulic Engineering Circular No. 9. Publication No. FHWA-IF-04-016.
Available: <http://www.fhwa.dot.gov/engineering/hydraulics/pubs/04016/>.

FHWA 2012b. Stream Stability at Highway Structures. Hydraulic Engineering Circular No. 20.
Publication No. FHWA-HIF-12-004. Federal Highway Administration, U.S. Department of
Transportation, Washington, DC.

Large Wood National Manual July 2015


5-48
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Fischenich, C., and J.V. Morrow, Jr. 2000. Streambank Habitat Enhancement with Large Woody Debris.
Publication No. ERDC TN-EMRRP-SR-13. U.S. Army Engineer Research and Development Center.
Available: <http://el.erdc.usace.army.mil/elpubs/pdf/sr13.pdf>.

Fox, M. J. and S. Bolton. 2007. A Regional and Geomorphic Reference for Quantities and Volumes of
Instream Wood in Unmanaged Forested Basins of Washington State. North American Journal of
Fisheries Management 27:342–359.

Fremier, A. K., J. I. Seo, and F. Nakamura. 2010. Watershed Controls on the Export of Large Wood
from Stream Corridors. Geomorphology 117:33–43.

Friedman, J. M., G. T. Auble, P. B. Shafroth, M. L. Scott, M. F. Merigliano, M. D. Preehling, and E.


K.Griffin. 2005. Dominance of Non-Native Riparian Trees in Western USA. Biological Invasions
7:747–751.

Furniss, M., T. Ledwith, M. Love, B. McFadin, and S. Flanagan 1998. Response of Road-Stream
Crossings to Large Flood Events in Washington, Oregon, and Northern California. USDA-Forest
Service, Technology & Development Program, Corvallis OR.

Ghosn, M., F. Moses, and J. Wang. Undated. Design of Highway Bridges for Extreme Events.

Gillespie, N., A. Unthank, L. Campbell, P. Anderson, R. Gubernick, M. Weinhold, D. Cenderelli, B.


Austin, D. McKinley, S. Wells, J. Rowan, C. Orvis, M. Hudy, A. Bowden, A. Singler, E. Fretz, J. Levine,
and R. Kirn 2014. Flood Effects on Road-Stream Crossing Infrastructure: Economic and
Ecological Benefits of Stream Simulation Designs. Fisheries 39(2):62–76.

Goode, J. R., C. H. Luce, and J. M. Buffington. 2012. Enhanced Sediment Delivery in a Changing Climate
in Semi-Arid Mountain Basins: Implications for Water Resource Management and Aquatic
Habitat in the Northern Rocky Mountains. Geomorphology 139-140:1–15.

Gosnell, H., and E. Kelly. 2010. Peace on the River? Social-Ecological Restoration and Large Dam
Removal in the Klamath basin, USA. Water Alternatives 3:362–383.

Gowan, C., and K. D. Fausch. 1996. Long-Term Demographic Responses of Trout Populations to
Habitat Manipulation in Six Colorado Streams. Ecological Applications 6(3):931–946.

Graf, W. L. 1999. Dam Nation: A Geographic Census of American Dams and Their Large-Scale
Hydrologic Impacts. Water Resources Research 35:1305–1311.

Grant, G., J. Schmidt, J., and S. Lewis. 2003. A Geological Framework for Interpreting Downstream
Effects of Dams on Rivers. Page 209–226 in A Unique River, Water Science Application, Volume 7.
American Geophysical Union.

Gregory, K. J., R. J. Davis, and S. Tooth. 1993. Spatial Distribution of Coarse Woody Debris Dams in
the Lymington Basin, Hampshire, UK. Geomorphology 6:207–224.

Gregory, S.V., M.A. Meleason, and D.J. Sobota. 2003b. Modeling the Dynamics of Wood in Streams and
Rivers. American Fisheries Society Symposium 37:315–335.

Gurnell, A. M., H. Piegay, F. J. Swanson, and S. V. Gregory. 2002. Large Wood and Fluvial Processes.
Freshwater Biology 47(4):601–619.

Large Wood National Manual July 2015


5-49
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Gurnell, A. J., W. Bertoldi, and D. Corenblit. 2012. Changing River Channels: The Roles of Hydrological
Processes, Plants and Pioneer Fluvial Landforms in Humid Temperate, Mixed Load, Gravel Bed
Rivers. Earth-Science Reviews 111:129–141.

Guyette, R. P., D. C. Dey, and M. C. Stambaugh 2008. The Temporal Distribution and Carbon Storage
of Large Oak Wood in Streams and Floodplain Deposits. Ecosystems 11:643–653.

Haga, H., T. Kumagai, K. Otsuki, and S. Ogawa. 2002. Transport and Retention of Coarse Woody
Debris in Mountain Streams: An In Situ Field Experiment of Log Transport and a Field Survey of
Coarse Woody Debris Distribution. Water Resources Research 38:1126,
doi:10.1029/2001WR001123.

Hamill, L. 1999. Bridge Hydraulics. New York, NY: Routledge.

Harmon, M. E., J. F. Franklin, F. J. Swanson, P. Sollins, S. V. Gregory, J. D. Lattin, N. H. Anderson, S. P.


Cline, N. G. Aumen, J. R. Sedell, G. W. Lienkaemper, K. Cromack, Jr., and K. W. Cummins. 1986.
Ecology of Coarse Woody Debris in Temperate Ecosystems. Advances in Ecological Research
15:133–302.

Hart, E. A. 2002. Effects of Woody Debris on Channel Morphology and Sediment Storage in
Headwater Streams in the Great Smoky Mountains, Tennessee-North Carolina. Physical
Geography 23:492–510.

Harwood, K., and A. G. Brown. 1993. Fluvial Processes in a Forested Anastomosing River: Flood
Partitioning and Changing Flow Patterns. Earth Surface Processes and Landforms 18:741–748.

Hester, E. T., and M. W. Doyle. 2008. In-Stream Geomorphic Structures as Drivers of Hyporheic
Change. Water Resources Research 44:W03427.

Homer, C. C., L. Huang, B. W. Yang, and M. Coan. 2004. Development of a 2001 National Landcover
Database for the United States. Photogrammetric Engineering and Remote Sensing 70(7):829–
840.

Hough-Snee, N., A. Kasprak, B. B. Roper, and C. S. Meredith. 2014. Direct and Indirect Drivers of
Instream Wood in the Interior Pacific Northwest, USA: Decoupling Climate, Vegetation,
Disturbance, and Geomorphic Setting. Riparian Ecology and Conservation 2:14–34.

Huryn, A. D., and J. B. Wallace. 1987. Local Geomorphology as a Determinant of Macrofaunal


Production in a Mountain Stream. Ecology 68:1932–1942.

Intergovernmental Panel on Climate Change (IPCC). 2007. Climate Change 2007: Impacts,
Adaptation and Vulnerability. Contribution of Working Group II to the Fourth Assessment
Report of the Intergovernmental Panel on Climate Change. M. L. Parry, O. F. Canziani, J. P.
Palutikof, P. J. van der Linden, and C. E. Hanson (eds.). Cambridge, UK, and New York, NY:
Cambridge University Press.

Jeffries, R., S. E. Darby, and D. A. Sear. 2003. The Influence of Vegetation and Organic Debris on
Flood-Plain Sediment Dynamics: Case Study of a Low-Order Stream in the New Forest, England.
Geomorphology 51:61–80.

Johnson, S. L., F. J. Swanson, G. E. Grant, and S. M. Wondzell. 2000. Riparian Forest Disturbances by a
Mountain Flood—The Influence of Floated Wood. Hydrological Processes 14:3031–3050.

Large Wood National Manual July 2015


5-50
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Kasprak, A., F. J. Magilligan, K. H. Nislow, and N. P. Snyder, N.P. 2012. A Lidar-Derived Evaluation of
Watershed-Scale Large Woody Debris Sources and Recruitment Mechanisms: Coastal Maine,
USA. River Research Applications 28:1462–1476.

Kaushal, S. S., G. E. Likens, N. A. Jaworski, M. L. Pace, A. M. Sides, D. Seekell, K. T. Belt, D. H. Secor, and
R. L. Wingate. 2010: Rising Stream and River Temperatures in the United States. Frontiers in
Ecology and the Environment 8:461–466. doi:10.1890/090037.

Kloehn, K., T. Beechie, S. Morley, H. Coe, and J. Duda, J. 2008. Influence of Dams on River-Floodplain
Dynamics in the Elwha River, Washington. Northwest Science 82:224–235.

Knowles, N., M. D. Dettinger, and D. R. Cayan. 2006. Trends in Snowfall versus Rainfall in the
Western United States. Journal of Climate 19:4545–4559. doi:10.1175/JCLI3850.1. Available:
http://journals.ametsoc.org/doi/pdf/10.1175/JCLI3850.1.

Knutti, R., G. Abramowitz, M. Collins, V. Eyring, P. J. Gleckler, B. Hewitson, and L. Mearns. 2010. Good
Practice Guidance Paper on Assessing and Combining Multi Model Climate Projections. In T. F.
Stocker, D. Qin, G.-K. Plattner, M. Tignor, and P.M. Midgley (eds.), Meeting Report of the
Intergovernmental Panel on Climate Change Expert Meeting on Assessing and Combining Multi
Model Climate Projections. IPCC Working Group I Technical Support Unit, University of Bern,
Bern, Switzerland.

Kraft, C. E., and D. R. Warren. 2003. Development of Spatial Pattern in Large Woody Debris and
Debris Dams in Streams. Geomorphology 51:127–139.

Kramer, N., and E. Wohl. 2014. Estimating Fluvial Wood Discharge using Time-Lapse Photography
with Varying Sampling Intervals. Earth Surface Processes and Landforms 39:844–852.

Kreutzweiser, D.P., K. P. Good, and T. M. Sutton. 2005. Large Woody Debris Characteristics and
Contributions to Pool Formation in Forest Streams of the Boreal Shield. Canadian Journal of
Forest Research 35:1213–1223.

Kukulak, J., A. Pazdur, and T. Kuc. 2002. Radiocarbon Dated Wood Debris in Floodplain Deposits of
the San River in the Bieszczady Mountains. Geochronometria 21:129–136.

Lagasse, P. F., P. Clopper, L. Zevenbergen, W. Spitz, and L. G. Girard.2010. Effects of Debris on Bridge
Pier Scour. Federal Highway Administration. Washington, D.C.

Lagasse, P. F., L. W. Zevenbergen, W. J. Spitz, and L. A. Arneson 2012. Stream Stability at Highway
Structures, Fourth Edition. Hydraulic Engineering Circular No. 20. Publication No. FHWA-HR-12-
004. Office of Technology, Federal Highway Administration, Washington, DC.

Lassettre, N. S. and R. R. Harris 2001. The Geomorphic and Ecological Influence of Large Woody
Debris in Streams and Rivers. University of California, Berkeley, CA.

Lassettre, N. S., and G. M. Kondolf. 2012. Large Woody Debris in Urban Stream Channels: Redefining
the Problem. River Research and Applications 28(9):1477–1487.

Lassettre, N., and H. Piégay. 2008. Decadal Changes in Distribution and Frequency of Wood in a Free
Meandering River, the Ain River, France. Earth Surface Processes and Landforms 1112:1098–
1112.

Large Wood National Manual July 2015


5-51
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Lautz, L. K., and R. M. Fanelli. 2008. Biogeochemical Hotspots in the Streambed around Restoration
Structures. Biogeochemistry 91:85–104.

Lautz, L. K., D. I. Siegel, and R. L. Bauer. 2006. Impact of Debris Dams on Hyporheic Interaction along
a Semi-Arid Stream. Hydrological Processes 20:183–196.

Lisle, T. E., Y. Cui, G. Parker, J. E. Pizzuto, and A. M. Dodd. 2001. The Dominance of Dispersion in the
Evolution of Bed Material Waves in Gravel-Bed Rivers. Earth Surface Processes and Landforms
26:1409–1420.

Loarie, S. R., P. B. Duffy, H. Hamilton, G. P. Asner, C. B. Field, and D. D. Ackerly. 2009. The Velocity of
Climate Change. Nature 462:1052–1055. doi:10.1038/nature08649.

Logan, J. A., W. W. Macfarlane, and L. Willcox. 2010. White-Bark Pine Vulnerability to Climate Change
Induced Mountain Pine Beetle Disturbance in the Greater Yellowstone Ecosystem. Ecological
Application 20:895–902. doi:10.1890/09- 0655.1. Available:
http://www.esajournals.org/doi/pdf/10.1890/09-0655.1.

MacFarlane, W. A., and E. Wohl. 2003. Influence of Step Composition on Step Geometry and Flow
Resistance in Step-Pool Streams of the Washington Cascades. Water Resources Research
39:1037. doi:10.1029/2001WR001238.

MacVicar, B., and H. Piégay. 2012. Implementation and Validation of Video Monitoring for Wood
Budgeting in a Wandering Piedmont River, the Ain River (France). Earth Surface Processes and
Landforms 37:1272–1289.

Magilligan, F. J., and K. H. Nislow. 2005. Changes in Hydrologic Regime by Dams. Geomorphology
71:61–78.

Magilligan, F. J., K. H. Nislov, G. B. Fisher, J. Wright, G. Mackey, and M. Laser 2007. The Geomorphic
Function and Characteristics of Large Woody Debris in Low Gradient Rivers, Coastal Maine, USA.
Geomorphology 97:467–482.

Major, J., J. O’Connor, C. Podolak, M. K. Keith, G. E. Grant, K. Spicer, S. Pittman, H. M. Bragg, J. R.


Wallick, D. Q. Tanner, A. Rhode, and P. Wilcock. 2012. Geomorphic Response of the Sandy River,
Oregon, to Removal of Marmot Dam. U.S. Geological Survey Professional Paper 1792.

Manga, M., and J. W. Kirchner. 2000. Stress Partitioning in Streams by Large Woody Debris. Water
Resources Research 36:2373–2379.

Manners, R. W., M. W. Doyle, and M. J. Small. 2007. Structure and Hydraulics of Natural Woody
Debris Jams. Water Resources Research 43, doi:10.1029/2006WR004910.

Manners, R. B., J. C. Schmidt, and M. L. Scott, M.L. 2014. Mechanisms of Vegetation-Induced Channel
Narrowing on an Unregulated Canyon Bound River: Results from a Natural Field-Scale
Experiment. Geomorphology 211:100–115.

Mao, L., and F. Comiti. 2010. The Effects of Large Wood Elements During an Extreme Flood in a Small
Tropical Basin of Costa Rica. WIT Transactions on Engineering Sciences 67:225–236.

Marcus, W. A., R. A. Marston, C. R. Colvard, and R. D. Gray. 2002. Mapping the Spatial and Temporal
Distributions of Woody Debris in Streams of the Greater Yellowstone Ecosystem, USA.
Geomorphology 44:323–335.

Large Wood National Manual July 2015


5-52
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Marcus, W. A., J. Rasmussen, and M. A. Fonstad. 2011. Response of the Fluvial Wood System to Fire
and Floods in Northern Yellowstone. Annals of the Association of American Geographers 101:21–
44.

Martin, D. J., and L. E. Benda. 2001. Patterns of Instream Wood Recruitment and Transport at the
Watershed Scale. Transactions of the American Fisheries Society 130:940–958.

Massachusetts Department of Transportation (MassDOT). 2010. Design of Bridges and Culverts for
Wildlife Passage at Freshwater Streams. Highway Division, Environmental, Bridge, Construction,
and Hydraulics Sections, Boston, MA.

Matheussen, B., R. L. Kirschbaum, I. A. Goodman, G. M. O’Donnell, and D. P. Lettenmaier. 2000. Effects


of Land Cover Change on Streamflow in the Interior Columbia River Basin (USA and Canada).
Hydrological Processes 14:867–885.

May, C. L., and R. E. Gresswell. 2003a. Large Wood Recruitment and Redistribution in Headwater
Streams in the Southern Oregon Coast Range, USA. Canadian Journal of Forest Research
33:1353–1362.

May, C. L., and R. E. Gresswell. 2003b. Processes and Rates of Sediment and Wood Accumulation in
Headwater Streams of the Oregon Coast Range, USA. Earth Surface Processes and Landforms
28:409–424.

May, C. L., E. B. Welch, R. R. Horner, J. R. Karr, and B. W. Mar, B.W. 1997. Quality Indices for
Urbanization Effects on Puget Sound Lowland Streams. Water Resource Series Tech Report 154.
Seattle, Washington.

Meleason, M. A., R. J. Davies-Colley, and G. M. J. Hall. 2007. Characterizing the Variability of Wood in
Streams: Simulation Modelling Compared with Multiple-Reach Surveys. Earth Surface Processes
and Landforms 32:1164–1173.

Melillo, J. M., T. C. Richmond, and G. W. Yohe (eds.). 2014. Climate Change Impacts in the United
States: The Third National Climate Assessment. U.S. Global Change Research Program, 841 pp.
doi:10.7930/J0Z31WJ2.

Meredith, C., B. Roper, and E. Archer. 2014. Reductions in Instream Wood in Streams near Roads in
the Interior Columbia River Basin. North American Journal of Fisheries Management 34(3):493–
506.

Merritt, D. M., N. L. R. Poff. 2010. Shifting Dominance of Riparian Populus and Tamarix Along
Gradients of Flow Alteration in Western North American Rivers. Ecological Applications 20:135–
152.

Merten, E., J. Finlay, L. Johnson, R. Newman, R., H. Stefan, and B. Vondracek. 2010. Factors
Influencing Wood Mobilization in Stream. Water Resources Research 46:W10514.

Merten, E. C., P. G. Vaz, J. A. Decker-Fritz, J. C. Finlay, and H. G. Stefan. 2013. Relative Importance of
Breakage and Decay as Processes Depleting Large Wood from Streams. Geomorphology 190:40–
47.

Mikuś, P., B. Wyżga, R. J. Kaczka, E. Walusiak, and J. Zawiejska. 2013. Islands in a European Mountain
River: Linkages with Large Wood Deposition, Flood Flow and Plant Diversity. Geomorphology
202:115–127.

Large Wood National Manual July 2015


5-53
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Montgomery, D. R. 1999. Process Domains and the River Continuum. Journal of the American Water
Resources Association 35:397–410.

Montgomery, D. R., and T. B. Abbe. 2006. Influence of Logjam-Formed Hard Points on the Formation
of Valley-Bottom Landforms in an Old-Growth Forest Valley, Queets River, Washington, USA.
Quaternary Research 65:147–155.

Montgomery, D. R. and H. Piégay 2003. Wood in Rivers: Interactions with Channel Morphology and
Processes. Geomorphology 51:1–5.

Montgomery, D. R., T. B. Abbe, J. M. Buffington, N. P. Peterson, K. M. Schmidt, and J. D. Stock. 1995a.


Distribution of Bedrock and Alluvial Channels in Forested Mountain Drainage Basins. Nature
381:587–589.

Morris, A. E. L., P. C. Goebel, and B. J. Palik. 2010. Spatial Distribution of Large Wood Jams in Streams
Related to Stream-Valley Geomorphology and Forest Age in Northern Michigan. River Research
and Applications 26:835–847.

Moulin, B., E. R. Schenk, and C. R. Hupp. 2011. Distribution and Characterization of In-Channel Large
Wood in Relation to Geomorphic Patterns on a Low-Gradient River. Earth Surface Processes and
Landforms 36:1137–1151.

Munn, N. L., and J. L. Meyer. 1990. Habitat-Specific Solute Retention in Two Small Streams: An
Intersite Comparison. Ecology 71:2069–2082.

Mutz, M. 2003. Hydraulic Effects of Wood in Streams. American Fisheries Society Symposium 37:93–
107.

Nagayama, S., F. Nakamura, Y. Kawaguchi, and D. Nakano. 2012. Effects of Configuration of Instream
Wood on Autumn and Winter Habitat Use by Fish in a Large Remeandering Reach. Hydrobiologia
680:159–170.

Naiman, R. J., R. E. Bilby, and P. Bisson. 2000. Riparian Ecology and Management in the Pacific
Coastal Rain Forest. Bioscience 50:996–1011.

Naiman, R. J., J. S. Becthold, T. J. Beechie, J. J. Laterell, and R. Van Pelt. 2010. A Process-Based View of
Floodplain Forest Patterns in Coastal River Valleys of the Pacific Northwest. Ecosystems 13:1–31.

Nakamura, F., and F. J. Swanson. 1993. Effects of Coarse Woody Debris on Morphology and Sediment
Storage of a Mountain Stream System in Western Oregon. Earth Surface Processes and Landforms
18:43–61.

National Resources Conservation Service (NRCS) 2007e. Use of Large Woody Material for Habitat
and Bank Protection - Technical Supplement 14j of Part 654. The National Engineering
Handbook. 210–VI–NEH, August 2007. U.S. Department of Agriculture, Washington, DC.

New York State Department of Environmental Conservation (NYSDEC). 2014. Removal of Woody
Debris and Trash from Rivers and Streams. In Post-Flood Stream Reconstruction: Guidelines and
Best Practices. Albany, NY.

Nilsson, C., and K. Berggren. 2000. Alterations of Riparian Ecosystems Caused by River Regulation.
Bioscience 50:783–792.

Large Wood National Manual July 2015


5-54
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Oswald, E. B., and E. Wohl. 2008. Wood-Mediated Geomorphic Effects of a Jökuhlaup in the Wind
River Mountains, Wyoming. Geomorphology 100:549–562.

Pealer, S. 2012. Lessons from Irene – Building Resiliency as We Rebuild. Vermont Agency of Natural
Resources Climate Change Team, Montpelier, VT. Available:
http://www.anr.state.vt.us/anr/climatechange/Pubs/Irene_Facts.pdf.

Pettit, N. E., and R. J. Naiman. 2006. Flood-Deposited Wood Creates Regeneration Niches for Riparian
Vegetation on a Semi-Arid South African River. Journal of Vegetation Science 17:615–624.

Pettit, N. E., R. J. Naiman, K. H. Rogers, and J. E. Little. 2005. Post-Flooding Distribution and
Characteristics of Large Woody Debris Piles Along the Semi-Arid Sabie River, South Africa. River
Research and Applications 21:27–38.

Phillips, J. D. 2012. Log-Jams and Avulsions in the San Antonio River Delta, Texas. Earth Surface
Processes and Landforms 37:936–950.

Piégay, H. 1993. Nature, Mass and Preferential Sites of Coarse Woody Debris Deposits in the Lower
Ain Valley (Mollon Reach), France. Regulated Rivers: Research and Management 8:359–372.

Poff, N. L., and H. K. H. Zimmerman. 2010. Ecological Responses to Altered Flow Regimes: A
Literature Review to Inform the Science and Management of Environmental Flows. Freshwater
Biology 55:194–205.

Poff, N. L., B. P. Bledsoe, and C. O. Cuhaciyan. 2006. Hydrologic Variation with Land Use across the
Contiguous United States: Geomorphic and Ecological Consequences for Stream Ecosystems.
Geomorphology 79:264–285. doi:10.1016/ j.geomorph.2006.06.032.

Pohl, M. M. 2002. Bringing Down Our Dams: Trends in American Dam Removal Rationales. Journal of
the American Water Resources Association 38:1511–1519.

Raffa, K. F., B. H. Aukema, B. J. Bentz, A. L. Carroll, J. A. Hicke, M. G. Turner, and W. H. Romme. 2008.
Cross-Scale Drivers of Natural Disturbances Prone to Anthropogenic Amplification: The
Dynamics of Bark Beetle Eruptions. Bio-Science 58:501–517. doi:10.1641/b580607. Available:
http://www.jstor.org/stable/pdfplus/10.1641/B580607. pdf.

Raikow, D. F., S. A. Grubbs, and K. W. Cummins. 1995. Debris Dam Dynamics and Coarse Particulate
Organic Matter Retention in an Appalachian Mountain Stream. Journal of the North American
Benthological Society 14:535–546.

Ravazzolo, D., L. Mao, L. Picco, and M. A. Lenzi 2015. Tracking Log Displacement During Floods in the
Tagliamento River Using RFID and GPS Tracker Devices. Geomorphology 228:226-233.

Richmond, A. D., and K. D. Fausch. 1995. Characteristics and Function of Large Woody Debris in
Subalpine Rocky Mountain Streams in Northern Colorado. Canadian Journal of Fisheries and
Aquatic Sciences 52:1789–1802.

Rigon, E., F. Comiti, and M. A. Lenzi. 2012. Large Wood Storage in Streams of the Eastern Italian Alps
and the Relevance of Hillslope Processes. Water Resources Research 48:W01518,
doi:10.1029/2010WR009854 18 p.

Robison, E. G., and R. L. Beschta. 1990. Identifying Trees in Riparian Areas that can Provide Coarse
Woody Debris to Streams. Forest Science 36:790–801.

Large Wood National Manual July 2015


5-55
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Rood, S. B., and J. M. Mahoney. 1990. Collapse of Riparian Poplar Forests Downstream from Dams in
Western Prairies: Probable Causes and Prospects for Mitigation. Environmental Management
14:451–464.

Rood, S. B., C. R. Gourley, E. M. Ammon, L. G. Heki, J. R. Klotz, M. L. Morrison, D. Mosley, G. G.


Scoppettone, S. Swanson, and P. L. Wagner. 2003. Flows for Floodplain Forests: A Successful
Riparian Restoration. Bioscience 53:647–656.

Rosgen, D., and H. L. Silvey 1996. Applied River Morphology. Pagosa Springs, CO: Wildland Hydrology.

Ryan, S. E., E. L. Bishop, and J. M. Daniels. 2014. Influence of Large Wood on Channel Morphology and
Sediment Storage in Headwater Mountain Streams, Fraser Experimental Forest, Colorado.
Geomorphology 217:73–88.

Sawyer, A. H., M. B. Cardenas, and J. Buttles. 2011. Hyporheic Exchange due to Channel-Spanning
Logs. Water Resources Research 47(8):W08502.

Schenk, E. R., J. W. McCargo, B. Moulin, C. R. Hupp, and J. M. Richter. 2014a. The Influence of Logjams
on Largemouth Bass (Micropterus salmoides) Concentrations on the lower Roanoke River, a
Large Sand-Bed River. River Research and Applications 2014(DOI: 10.1002/rra.2779).

Schenk, E. R., B. Moulin, C. R. Hupp, and J. M. Richter. 2014b. Large Wood Budget and Transport
Dynamics on a Large River Using Radio Telemetry. Earth Surface Processes and Landforms
39:487–498.

Schiff, R., J. S. Clark, G. Alexander, and M. Kline 2008a. The Vermont Agency of Natural Resources
Reach Habitat Assessment (RHA). Prepared by Milone & MacBroom, Inc. with the Vermont
Agency of Natural Resources, Departments of Environmental Conservation and Fish and
Wildlife, Waterbury, VT.

Schiff, R., J. S. Clark, and S. Jaquith 2008b. The Vermont Culvert Geomorphic Compatibility Screening
Tool. Prepared by Milone & MacBroom, Inc. with the VT DEC River Management Program,
Waterbury, VT.

Schiff, R., E. Fitzgerald, J. MacBroom, M. Kline, and S. Jaquith 2014. The Vermont Standard River
Management Principles and Practices (Vermont SRMPP): Guidance for Managing Vermont's
Rivers Based on Channel and Floodplain Function. Prepared by Milone & MacBroom and
Fitzgerald Environmental Associates for and in collaboration with the Vermont Rivers Program,
Montplelier, VT.

Schmidt, J. C., and P. R. Wilcock. 2008. Metrics for Assessing the Downstream Effects of Dams. Water
Resources Research 44:W04404. doi:10.1029/2006WR005092.

Scott, M. L., J. M. Friedman, G. T. Auble. 1996. Fluvial Process and the Establishment of Bottomland
Trees. Geomorphology 14:327–339.

Sear, D. A., C. E. Millington, D. R. Kitts, and R. Jeffries. 2010. Logjam Controls on Channel:Floodplain
Interactions in Wooded Catchments and their Role in the Formation of Multi-Channel Patterns.
Geomorphology 116:305–319.

Senter, A. E., and G. B. Pasternack. 2010. Large Wood Aids Spawning Chinook Salmon (Oncorhynchus
Tshawytscha) in Marginal Habitat on a Regulated River in California. River Research and
Applications 27:550–565.]

Large Wood National Manual July 2015


5-56
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Shields, F. D., and R. H. Smith. 1992. Effects of Large Woody Debris Removal on Physical
Characteristics of a Sand-Bed River. Aquatic Conservation: Marine and Freshwater Ecosystems
2:145–163.

Sklar, L. S., J. Fadde, J. G. Venditti, P. Nelson, M. A. Wydzga, Y. Cui, and W. E. Dietrich. 2009.
Translation and Dispersion of Sediment Pulses in Flume Experiments Simulating Gravel
Augmentation Below Dams. Water Resources Research 45:W08439,
doi:10.1029/2008WR007346.

Sleeter, B., T. Wilson, W. Acevedo, W. 2012. Status and Trends of Land Change in the Western United
States—1973–2000. U.S. Geological Survey Professional Paper 1794-A.

Smith, M. P., R. Schiff, A. Olivero, and J. G. MacBroom 2008. The Active River Area: A Conservation
Framework to Protect Rivers and Streams. Boston, MA: The Nature Conservancy.

Smith, R. D., R. C. Sidle, and P. E. Porter. 1993. Effects on Bedload Transport of Experimental
Removal of Woody Debris from a Forest Gravel-Bed Stream. Earth Surface Processes and
Landforms 18:455–468.

Spanhoff, B., and E. I. Meyer. 2004. Breakdown Rates of Wood in Streams. Journal of the North
American Benthological Society 23(2):189–197.

Spooner, D. E., M. A. Xenopoulos, C. Schneider, and D. A. Woolnough. 2011. Coextirpation of Host-


Affiliate Relationships in Rivers: The role of Climate Change, Water Withdrawal, and Host-
Specificity. Global Change Biology 17:1720–1732. doi:10.1111/j.1365-2486.2010.02372.x.

Thompson, D. M. 1995. The Effects of Large Organic Debris on Sediment Processes and Stream
Morphology in Vermont. Geomorphology 11(3):235–244.

Tyler, R. N. 2011. River Debris: Causes, Impacts, and Mitigation Techniques. Prepared for Ocean
Renewable Power Company by the Alaska Center for Energy and Power, Fairbanks, Alaska.

Umazano, A.M., R.N. Melchor, E. Bedatou, E.S. Bellosi, and J.M. Krause. 2014. Fluvial Response to
Sudden Input of Pyroclastic Sediments During the 2008–2009 Eruption of the Chaitén Volcano
(Chile): The Role of Logjams. Journal of South American Earth Sciences 54:140–157.

U.S. Environmental Protection Agency, Office of Water. 2013. Climate Change Adaptation
Implementation Plan. Available: http://epa.gov/climatechange/Downloads/impacts-
adaptation/office-of-water-plan.pdf.

U.S. Climate Change Science Program (CCSP). 2008a. Preliminary Review of Adaptation Options for
Climate-Sensitive Ecosystems and Resources. A Report by the U.S. Climate Change Science
Program and the Subcommittee on Global Change Research. (S. H. Julius and J.M. West [eds.], J. S.
Baron, B. Griffith, L. A. Joyce, P. Kareiva, B. D. Keller, M. A. Palmer, C. H. Peterson, and J. M. Scott
[Authors]). U.S. Environmental Protection Agency. Washington, D.C. 873 pp.

U.S. Climate Change Science Program (CCSP ). 2008b. The Effects of Climate Change on Agriculture,
Land Resources, Water Resources, and Biodiversity in the United States. A Report by the U.S.
Climate Change Science Program and the Subcommittee on Global Change Research (P.
Backlund, A. Janetos, D. Schimel, J. Hatfield, K. Boote, P. Fay, L. Hahn, C. Izaurralde, B.A. Kimball,
T. Mader, J. Morgan, D. Ort, W. Polley, A. Thomson, D. Wolfe, M. G. Ryan, S. R. Archer, R. Birdsey,
C. Dahm, L. Heath, J. Hicke, D. Hollinger, T. Huxman, G. Okin, R. Oren, J. Randerson, W.

Large Wood National Manual July 2015


5-57
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Schlesinger, D. Lettenmaier, D. Major, L. Poff, S. Running, L. Hansen, D. Inouye, B. P. Kelly, L.


Meyerson, B. Peterson, and R. Shaw). U.S. Department of Agriculture. Washington, D.C. 362 pp.

U.S. Climate Change Science Program (CCSP). 2008c. Impacts of Climate Change and Variability on
Transportation Systems and Infrastructure: Gulf Coast Study, Phase I. A Report by the U.S. Climate
Change Science Program and the Subcommittee on Global Change Research (M. J. Savonis, V. R.
Burkett, and J. R. Potter [eds.]). U.S. Department of Transportation. Washington, D.C.

U.S. Department of the Interior – Bureau of Reclamation. 2011. West-Wide Climate Risk Assessments:
Bias-Corrected and Spatially Downscaled Surface Water Projections. Technical Memorandum No.
86-68210–2011-01

U.S. Environmental Protection Agency, Office of Water. 2013. Climate Change Adaptation
Implementation Plan. Available: http://epa.gov/climatechange/Downloads/impacts-
adaptation/office-of-water-plan.pdf.

U.S. Forest Service (USFS). 2008. Stream Simulation: An Ecological Approach to Providing Passage for
Aquatic Organisms at Road-Stream Crossings. Forest Service Stream-Simulation Working Group.
San Dimas, CA. May. Available:
http://www.stream.fs.fed.us/fishxing/publications/PDFs/AOP_PDFs/
08771801.pdf. Accessed: February 27, 2015.

U.S. Global Change Research Program (USGCRP). 2009. Global Climate Change Impacts in the United
States. Edited by T. R. Karl, J. M. Melillo, and T. C. Peterson. Cambridge, MA: Cambridge
University Press.

University of New Hampshire (UNH). 2009. New Hampshire Stream Crossing Guidelines. University of
New Hampshire, Durham, NH.

Vermont Agency of Natural Resources (VTANR). 2014. Vermont Stream Alteration General Permit.
Department of Environmental Conservation, Montpelier, VT.

Vermont Agency of Transportation (Vtrans). 2001. Hydraulics Manual. Montpelier, VT.

Walsh, C., and A. Roy. 2005. The Urban Stream Syndrome: Current Knowledge and the Search for a
Cure. Journal of the North American Benthological Society 24:706–723.

Walter, R.C. and D.J. Merritts. 2008. Natural Streams and the Legacy of Water-Powered Mills. Science
319(5861):299-304.

Warren, D. R. and C.E. Kraft. 2008. Dynamics of large wood in an eastern U.S. mountain stream.
Forest Ecology and Management 256(4):808-814.

Warren, D. R., C. E. Kraft, W. S. Keeton, J. S. Nunery, and G. E. Likens. 2009. Dynamics of Wood
Recruitment in Streams of the Northeastern US. Forest Ecology and Management 258:804–813.

Watts, R. J., B. D. Richter, J. J. Opperman, and K. H. Bowmer. 2011. Dam Reoperation in an Era of
Climate Change. Marine and Freshwater Research 62:321–327.

Webb, A. A., and W. D. Erskine. 2003. Distribution, Recruitment, and Geomorphic Significance of
Large Woody Debris in an Alluvial Forest Stream: Tonghi Creek, Southeastern Australia.
Geomorphology 51:109–126.

Large Wood National Manual July 2015


5-58
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Welber, M., W. Bertoldi, and M. Tubino. 2013. Wood Dispersal in Braided Streams: Results from
Physical Modeling. Water Resources Research 49:7388–7400.

Welty, J. J., T. Beechie, K. Sullivan, D. M. Hyink, R. E. Bilby, C. Andrus, and G. Pess. 2002. Riparian
Aquatic Interaction Simulator (RAIS): A Model of Riparian Forest Dynamics for the Generation of
Large Woody Debris and Shade. Forest Ecology and Management 162:299–318.

Wemple, B. C., and J. A. Jones 2003. Runoff Production on Forest Roads in a Steep, Mountain
Catchment. Water Resources Research 39(8). doi:10.1029/2002WR001744

Wenger, S. J., D. J. Isaak, C. H. Luce, H. M. Neville, K. D. Fausch, J. B. Dunham, D. C. Dauwalter, M. K.


Young, M. M. Elsner, B. E. Rieman, A. F. Hamlet, and J. E. Williams. 2011. Flow Regime,
Temperature, and Biotic Interactions Drive Differential Declines of Trout Species under Climate
Change. Proceedings of the National Academy of Sciences 108:14175–14180.
doi:10.1073/pnas.1103097108. Available: http://www.pnas.org/content/108/34/14175.full.
pdf+html.

Westerling, A. L., A. Gershunov, T. J. Brown, D. R. Cayan, and M. D. Dettinger. 2003. Climate and
Wildfire in the Western United States. Bulletin of the American Meteorological Society 84:595–
604. doi:10.1175/BAMS-84-5-595. Available:
http://journals.ametsoc.org/doi/pdf/10.1175/BAMS-84-5-595.

Westerling, A. L., H. G. Hidalgo, D. R. Cayan, and T. W. Swetnam. 2006. Warming and Earlier Spring
Increase Western U.S. Forest Wildfire Activity. Science 313: 940–943. doi:10.1126/science.
1128834.

White, P. S., and S. T. A. Pickett. 1985. Natural Disturbance and Patch Dynamics: An Introduction.
Pages 3–9 in S. T. A. Pickett and P. S. White (eds.), The Ecology of Natural Disturbance and Patch
Dynamics. San Diego, CA:Academic Press.

Whiting, P. J. 2002. Streamflow Necessary for Environmental Maintenance. Annual Review of Earth
and Planetary Sciences. 30:181–206.

Williams, G. P. 1986. River Meanders and Channel Size. Journal of Hydrology 88(1-2):147–164.

Williams, G.P., and M. G. Wolman. 1984. Downstream Effects of Dams on Alluvial Rivers. USG S
Professional Paper 1286.

Wipf, T. J., B. M. Phares, and J. Dahlberg 2012. Debris Mitigation Methods for Bridge Piers. Iowa State
University, Ames, IA.

Wohl, E. 2011a. Seeing the Forest and the Trees: Wood in Stream Restoration in the Colorado Front
Range, United States. Pages 399–418 in A. Simon, S. J. Bennett, and J. Castro (eds.), Stream
Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools. Washington,
D.C.: American Geophysical Union Press.

Wohl, E. 2011b. What Should these Rivers Look Like? Historical Range of Variability and Human
Impacts in the Colorado Front Range, USA. Earth Surface Processes and Landforms 36:1378–
1390.

Wohl, E. 2011c. Threshold-Induced Complex Behavior of Wood in Mountain Streams. Geology


39:587–590.

Large Wood National Manual July 2015


5-59
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Watershed-Scale and Long-Term Considerations

Wohl, E. 2013a. Redistribution of Forest Carbon Caused by Patch Blowdowns in Subalpine Forests of
the Southern Rocky Mountains, USA. Global Biogeochemical Cycles 27:1205-1213.

Wohl, E. 2014. A Legacy of Absence: Wood Removal in US Rivers. Progress in Physical Geography
38:637–663.

Wohl, E., and N. Beckman. 2014a. Controls on the Longitudinal Distribution of Channel-Spanning
Logjams in the Colorado Front Range, USA. River Research and Applications 30:112–131.

Wohl, E., and N. D. Beckman. 2014b. Leaky rivers: Implications for the loss of longitudinal fluvial
disconnectivity in headwater streams. Geomorphology 205:27–35.

Wohl, E., and D. Cadol. 2011. Neighborhood Matters: Patterns and Controls on Wood Distribution in
Old-Growth Forest Streams of the Colorado Front Range, USA. Geomorphology 125:132–146.

Wohl, E., and J. R. Goode. 2008. Wood Dynamics in Headwater Streams of the Colorado Rocky
Mountains. Water Resources Research 44:W09429.

Wohl, E., and K. Jaeger. 2009. A Conceptual Model for the Longitudinal Distribution of Wood in
Mountain Streams. Earth Surface Processes and Landforms 34:329–344.

Wohl, E. and D. M. Merritt. 2008. Reach-Scale Channel Geometry of Mountain Streams.


Geomorphology 93(3-4):168–185.

Wohl, E., and F. L. Ogden. 2013. Organic Carbon Export in the Form of Wood During an Extreme
Tropical Storm, Upper Rio Chagres, Panama. Earth Surface Processes and Landforms 38:1407–
1416.

Wohl, E., F. L. Ogden, and J. Goode. 2009. Episodic Wood Loading in a Mountainous Neotropical
Watershed. Geomorphology 111:149–159.

Wohl, E., L. E. Polvi, and D. Cadol. 2011. Wood Distribution Along Streams Draining Old-Growth
Forests in Congaree National Park, South Carolina, USA. Geomorphology 126:108–120.

Wohl, E., S. Bolton, D. Cado, F. Comiti, J. R. Goode, and L. Mao. 2012. A Two End-Member Model of
Wood Dynamics in Headwater Neotropical Rivers. Journal of Hydrology 462-463:67–76.

Wondzell, S. M., J. LaNier, R. Haggerty, R. D. Woodsmith, and R. T. Edwards. 2009. Changes in


Hyporheic Flow Following Experimental Removal of a Small, Low-Gradient Stream. Water
Resources Research 45:W05406, 13 pp.

Wyżga, B., and J. Zawiejska. 2005. Wood Storage in a Wide Mountain River: Case Study of the Czarny
Dunajec, Polish Carpathians. Earth Surface Processes and Landforms 30:1475–1494.

Young, M. K., E. A. Mace, E. T. Ziegler, and E. K. Sutherland. 2006. Characterizing and Contrasting
Instream and Riparian Coarse Wood in Western Montana Basins. Forest Ecology and
Management 226:26–40.

Large Wood National Manual July 2015


5-60
Chapter 6
ENGINEERING CONSIDERATIONS

Complex timber revetment designed to protect bank by partitioning shear stress while also
creating cover and hydraulic refugia for salmonids,
South Fork Nooksack River, Northwest Washington (Tim Abbe 2012)

AUTHORS

Doug Shields (Shields Engineering, LLC)


Tim Abbe (NSD)
Mike Hrachovec (NSD)
Leif Embertson (ICF International)
Carl Jensen (ICF International)
This page intentionally left blank.
developing a set of bid documents (plans and
6.1 Overview specifications) that can be constructed within
This chapter provides an introduction to the the allocated budget. Preparation of and format
engineering design of large wood placements in of plans, specifications, and estimates typically
streams. There are many factors to consider in follows standards established by local, state, or
any stream engineering endeavor and many federal agencies, which are familiar to
assumptions to be made to assess a design contractors. Because stream restoration typically
quantitatively. Project documentation should involves unique circumstances and structures,
include a basis of design describing the special provisions are often required, which
methodology and assumptions used to develop underscores the need for experience and
the design, no matter how simple or intricate the expertise working in fluvial systems.
project. As with other aspects of large wood
The information and guidelines presented here
placement projects, the effort devoted to
provide an introduction to the use of wood in
engineering should be commensurate with
restoring fluvial systems. Designs should always
project risk and scale (Figure 6-1). Higher risk
be led and reviewed by professionals with
projects should receive more intensive
expertise and experience in fluvial systems and
engineering (e.g., higher resolution pre-project
restoration. This chapter assumes that objectives
surveys, multidimensional computer models,
for a given project have been set prior to the
collection of calibration data, establishment of
design process using information from within
ecological baselines) to reduce uncertainty
this manual. Typically, these objectives will
surrounding technical issues. Even so, much
involve either habitat rehabilitation, channel
uncertainty will remain after state-of-the art
stabilization, or both. Clear, written objectives
engineering is employed.
are needed to drive and justify design decisions.

CROSS-REFERENCE
6.2 Introduction
Chapter 7, Risk Considerations, provides detailed Naturally occurring large wood influences or
guidance on overall project risk assessment and governs hyporheic exchange, habitat complexity,
management. hydraulics, sediment storage and transport, and
reach-scale geomorphology.

As described previously, fluvial systems are


directly influenced by biological (e.g., plants), CROSS-REFERENCE
physical (e.g., hydrologic and geomorphic), and
social (e.g., recreation and flood control) factors, Chapter 1, Large Wood Introduction, Chapter 3,
all of which interact and change over time. This Ecological and Biological Considerations, and Chapter
underlies the need for interdisciplinary design. 4, Geomorphology and Hydrology Considerations,
The design team’s responsibility is to develop a provide much more information about the functional
role of large wood in riverine ecosystems.
conceptual design that achieves the project
objectives within the site constraints. The role of
professional geologists is to ensure the design
achieves the desired geomorphic conditions and Much evidence attests to the fact that North
can be constructed given surface and subsurface American rivers had much higher rates of large
conditions, and to predict how the design will wood loading (both on the floodplain and
perform through time. The role of professional instream) prior to European settlement.
engineers is to take this design to reality by

Large Wood National Manual July 2015


6-1
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Figure 6-1. Impact of Spatial Scale and Relative Risk on Engineering Aspects of a Large Wood Project

km2 = square kilometers; m3/s = cubic meters per second.


Appropriate commitment of project resources to design, breadth of stakeholder engagement, and
multidisciplinary involvement in the design process should reflect both spatial scale and relative risk. For
example, low-risk, small scale projects might be designed using qualitative impressions from visual
reconnaissance, professional judgment and manual computations. Intermediate level projects would rely on
aerial photos, survey data, gage records, and spreadsheet analysis. Higher risk projects, particularly those in
larger streams, require high-resolution hydrologic, sediment, and topographic and bathymetric data to
construct, calibrate, and verify numerical (computer) models. Some small scale projects are high risk due to
land use context, geomorphology, or hydrology; and some large scale projects are low risk due to similar
factors.

Currently, large wood density and stability have Large wood and structures comprising large
been drastically compromised by the combined wood have been used for river training and
influence of large wood removal, beaver decline, stabilization for centuries (Figure 6-2). In the
riparian and watershed deforestation, “splash latter half of the twentieth century timber was
damming” to transport logs to mill, largely replaced by rock, concrete, and steel in
channelization, dam construction, and channel channel stabilization. Wooden structures
enlargement due to incision. Accordingly, stream intended to improve fish habitat have been
restoration efforts often include replacement of described in literature from the nineteenth and
stable large wood by constructing instream large twentieth centuries (Thompson and Stull 2002).
wood structures, supplying loose large wood to Entering the twenty-first century there has been
the channel (replenishment of supply), or an increase in timber use driven by
trapping mobile wood. environmental concerns.

Large Wood National Manual July 2015


6-2
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Figure 6-2. Examples of Wood Placements Used to Stabilize River Banks

Photo a. Bundles of small wood have been used for Photo b. Placement of cedar brush mattress along
several thousand years in China to stabilize banks toe of Puyallup River North Levee, May 24, 1916
and levees (Glenn Wilson). (photo courtesy of Pierce County, Washington).

Photo c. Large timber cribs constructed in 1930s to Photo d. Complex timber revetment with internal
deflect flows on the Eel River, Northern California, rock collar ballast, 2010 (Tim Abbe).
circa 1960).

Photo e. Series of ELJ flow deflectors constructed in Photo f. Detail of engineered logjam, Cispus River,
1999, Cispus River, Washington (Tim Abbe). Washington, 2004 (Tim Abbe).

Large Wood National Manual July 2015


6-3
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Photo g. Wood structure placed to accelerate flows Photo h. Large wood with complex fine branches
and flush fine sediment deposited on gravel bed of placed along outside of bend to trigger sediment
Fawn River, Ohio, following rapid drawdown of an deposition along steep, eroding bank of a Georgia
upstream impoundment. (Photo courtesy of Fawn stream.
River Restoration and Conservation Charitable
Trust.)

Photo i. Headwater or small perennial stream Photo j. The wood is not visible in the post-project
construction can use wood scaled appropriately to photo, taken 10 years later, but will continue to
provide overhead cover and bank stability that can provide habitat value in this stream for many
last 100 years or more(Photo by Inter-Fluve). decades (Photo by Inter-Fluve).

The new generation of timber structures builds emulated the form and function of naturally
upon basic principles of earlier structures such occurring, stable accumulations of wood,
as crib walls and deflectors, but they represent particularly in rivers of the Pacific Northwest
a major change to more physical complexity (Abbe et al. 1997; Hilderbrand et al. 1998).
that better emulates natural conditions (Abbe et Additional research and successful installations
al. 1997, 2003b, 2003c; Abbe and Brooks 2011). have been carried out in Australia since about
2000 (Brooks 2006; Brooks et al. 2006; Simon
Many of the earliest river training structures
et al. 2012).
built on large rivers in the United States
included willow mattresses, brush mattresses, Although success rates for Australian projects
or wooden pilings driven into the bed (Vanoni have been relatively high, the results of large
1975; Keown et al. 1977). During the 1990s, wood installations for ecological restoration in
increasing appreciation of the importance of the United States have varied widely (Roni et al.
large wood in natural riverine ecosystems 2008). A 1986 evaluation of 137 log habitat
triggered efforts to design structures that structures in the Northwest revealed high rates

Large Wood National Manual July 2015


6-4
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

of damage and failure (Frissell and Nawa 1992). designed to resist movement up to a specified
Of 72 large wood structures placed within a discharge.
short reach of a small stream in the Southeast,
51 were damaged or destroyed within 3 years
(Shields et al. 2008). Nevertheless, careful CAVEAT
planning and design can reduce the risk
associated with large wood projects (see Designing for Dynamic Process not Static Structure
Chapter 7, Risk Considerations). When engineers design structures for the river
environment, normally great care is taken to ensure
The planning process should include
that the structures will retain a constant position in
establishment of measurable objectives. space despite fluctuating flow and sediment load.
Biological objectives should be based on Although stream beds may fill scour and forces
assessment of current and desired habitat imposed by flow vary widely and hydrographs rise
quality and quantity. and fall, we expect well-designed revetments,
training structures, bridges, dams, or gates to stay in
place so that they will fulfill their intended function.
CROSS-REFERENCE However, designing large wood additions is often a
different proposition. Instead of static structure, we
are striving for more or less static function. If a large
More quantitative analyses include assessment of
wood structure is intended to create and maintain
limiting factors for populations of target species
pool habitat or cover, it may do this even if the
according to principles found in Chapter 3, Ecological
individual wood members in the structure shift,
and Biological Considerations.
rotate, or are replaced by fluvially transported wood.
Wood structures shrink and subside as wood decays
and grow as floating wood is racked up, sediments
The outcome of such analyses provides a
deposit, and, in some cases, as trees colonize the
rationale for selecting the numbers and types of structures and associated sediment bars. If a wood
large wood structures to be added to the project project is intended to shift the channel morphology
reach. of a reach, say from a braided condition to an
anastomosed channel or from a channel evolution
This chapter focuses on the design of large
model type IV to type V or VI, the original placed
wood structures, which is a bit paradoxical. On wood may be buried or otherwise “lost” as the
the one hand, large wood reintroduction is a channel shifts to the desired state.
step toward a more natural fluvial system in
which large wood is both plentiful and mobile. Therefore, temporally dynamic wood structures do
not represent failure. Wood can provide habitat
Conversely, in almost all large wood structure
benefits or temporary channel stabilization benefits
projects, the designer’s intention is for the large
even if large wood structures lose their integrity
wood to be stationary for years if not decades. when placed wood is completely washed away and is
In many cases placed large wood structures will not replaced by other wood, when undesirable scour
accumulate additional large wood that is or deposition occurs, or when the expected habitat
naturally transported from upstream. Natural benefits are not realized, project outcomes are not
large wood residence times vary widely from deemed successful.
hours to centuries. However, unless a piece of
large wood is much longer than the channel As discussed in Chapter 3, Ecological and
width and has a large enough wood volume Biological Considerations, riparian revegetation
relative to the channel cross-section to act as a is a key component of large wood addition.
key member or is deeply buried in the bed or Vegetation growing on sediments deposited in
floodplain, it eventually moves downstream. On or adjacent to large wood can anchor and
the other hand, most large wood structures are restrain the wood, serve many of the same
functions as nonliving wood, and, over the

Large Wood National Manual July 2015


6-5
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

longer term, supply additional large wood to With these data, the project team should be able
the fluvial system as in lightly degraded stream to ascertain the trajectory of ongoing channel
corridors. Large wood is not added to be a evolution (e.g., incision, aggradation, widening,
permanent feature of the river system, but to narrowing, braiding, avulsion.). Large wood
assist the natural fluvial system in recovering a structures may not be successful if applied in an
cycle that involves wood addition, riparian zone effort to force a fluvial system to reverse the
regeneration, natural vegetative succession, and overall course of geomorphic evolution acting at
more wood addition. the watershed scale (Shields et al. 2008),
although reach-scale transformations have been
initiated by some projects in the Pacific
CROSS-REFERENCE Northwest. Clearly, natural large wood
accumulations have exerted major landscape
The content of this chapter presupposes completion impacts (Montgomery et al. 1995a, 1995b; Abbe
of a geomorphic assessment (Chapter 4,
2000; Abbe and Montgomery 2003;
Geomorphology and Hydrology Considerations) and
biological evaluation (Chapter 3, Ecological and Montgomery et al. 2003; Montgomery and Abbe
Biological Considerations) of the project site. 2006; Collins et al. 2012; Wohl 2013). The most
ecologically beneficial large wood projects have
floodplain-scale effects. There are still many
The geomorphic assessment should include a rural areas where this scale is possible, and the
description of the regions upstream and projects can deliver important benefits to
downstream as well as the project reach. At a downstream human communities by trapping
minimum, the assessment should include these mobile debris and attenuating flood peaks. But
features. most sites constrain project scale due to
 Characterization of geometry (thalweg floodplain development. While there are natural
profile, bed slope, cross section circumstances such as confined bedrock
characteristics). canyons where natural large wood frequencies
and densities are quite low, wood can be
 Historic changes in geometry.
effectively applied in a wide range of site
 Sediments (size, cohesion). conditions, including urban streams. Ecological
evaluation is needed to determine if existing
 Banks (erosion rates, locations, and
large wood loading and stability is lower than a
processes).
reference or other desirable state, or if positive
 Riparian vegetation. biotic response is likely to follow stable large
 Wood loading. wood reintroduction. In cases where there is a
desire to reintroduce mobile large wood, such
 Hydrology (frequency of overflow, as mitigating the impacts of dams, careful
magnitude of floods, and duration of analysis should be done regarding the
droughts). downstream fate of the large wood. If the large
 A disturbance history (dams, dam removals, wood simply flushes through the system it will
channelization, instream mining, fires, not provide desired benefits. Therefore if it is
floods, logging, farming, etc.). assumed the wood will be retained to enhance
habitat, the project proponents should describe
 Major sediment fluxes associated with these where and how large wood will be trapped and
disturbances. the function it will provide, and ensure
 Assessment of dynamic stakeholders it will not be a threat to
equilibrium/disequilbrium. infrastructure, which might entail improving

Large Wood National Manual July 2015


6-6
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

infrastructure or creating stable structures to be required that present difficulties in terms of


trap large wood in desirable locations. flow conveyance or cost. Because channel
incision is typically a progressive condition with
negative implications for both habitat and
CROSS-REFERENCE infrastructure, robust large wood placement
may be a priority for checking or reversing
Effects of dams, wood sources, floodplain incision and attendant lateral instability
interactions and other topics relating to large spatial
(widening and narrowing). In general, large
scales are discussed in Chapter 5, Watershed-Scale
and Long-Term Considerations. wood structures applied to channels that are
not actively incising and have cobble or finer
bed material may be designed with lower safety
factors because they incur less risk (Table 6-1).
6.3 Area of Applicability
Because most tree species produce wood that
The site context of large wood structure decays within a few years unless it is
projects is an essential aspect of success or continuously submerged, large wood structures
failure. First, candidate sites should display low are generally not suited for long-term
large wood loading relative to reference states stabilization unless the wood is preserved by
or sites, but large wood should be a natural continuous submersion. Section 6.4, Design Life
component of the geomorphic landscape. Reach of Places Wood, discusses decay rates and
context plays an important role in large wood design life considerations. Large wood projects
risk profile and overall feasibility. If local land are best viewed as measures that will require
use and infrastructure permits significant periodic maintenance or as bridges toward a
channel change and increased flooding, great target state that facilitates riparian forest
ecological benefit may be derived from large regrowth, large wood recruitment, and ongoing
wood projects. Large wood projects can be geomorphic evolution (Abbe and Brooks 2011).
extensive enough to raise flood stages, increase The latter approach necessarily involves
floodplain connection, develop or accentuate regeneration of forested riparian zones and
side channels, trigger migration or channel floodplains that serve as large wood sources
avulsions that create habitat, and reinitiate the (Erskine et al. 2012). The target state for the
cycle of tree growth and large wood supply fluvial system features higher levels of naturally
from the floodplain (Collins et al. 2012). recruited stable large wood, channel stability,
Because much of the local habitat value or habitat quality (Kail et al. 2007). Large wood
associated with large wood is due to scour and loadings featuring extremely large trees
deposition, streams with nonerodible produce resilient fluvial systems characterized
boundaries that transport little sediment are by high levels of biogeomorphic complexity
poor candidates to improve local habitat. (Collins et al. 2012). So, in most cases, placed
Deeply incised channels1 or channels with very large wood will not persist in its constructed
narrow or absent floodplains present form. Wood will decay, be re-arranged by river
difficulties due to the relatively frequent deep flows, or accumulate and trap drifting wood.
submergence of the large wood, with attendant Sustainable benefits of wood placement are
higher shear stresses and buoyant forces. In usually not due to the immobility or
such channels, larger, more robust designs may permanence of the placed members but due to
rejuvenation of constructed large wood features
with colonizing woody vegetation or recruited
1 Defined here as channels with average depths more large wood material and sediment.
than three times the average depth of nonincised
reaches or channels with flow capacity greater than
the 10-year return interval discharge.

Large Wood National Manual July 2015


6-7
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Table 6-1. Limitations on the Applicability of Large Wood Structures

Variable Considerations
Habitat Provides physical diversity, cover, velocity shelter, substrate sorting, pool development,
requirements undercut banks, and sites for terrestrial plant colonization using natural materials.
Existing large Absent or depressed relative to similar nearby reaches that are lightly degraded.
wood density
Sediment load Generally best for gravel bed systems, but have been applied to sand and cobble systems.
Resultant habitat value diminished when placed in streams with very low sediment loads.
Large wood structures may be rapidly buried in high sediment load reaches, diminishing
their aquatic habitat value, but accelerating recovery of terrestrial riparian habitats.
Bed material Anchoring will be difficult in hard beds such as cobble, boulder, or bedrock. Structures
placed in cobble bed rivers are often held in place with bed material used as ballast.
Bed stability Not suitable for rapidly avulsing, degrading or incising channels unless riparian
infrastructure and land use can tolerate large-scale channel movement. The best
situations include areas of general or local sediment deposition along reaches that are
stable or gradually aggrading. Deposition induced by large wood structures may be
stabilized by planted or volunteer woody vegetation, fully rehabilitating a naturally stable
bank by the time the placed woody materials decay. Unlike some of the other structure
types, rootwads often create scour zones, not deposition.
Bank material Large wood structures placed adjacent to erodible banks are subject to flanking, with
special care needed for structures on sandy banks.
Bank erosion Not recommended where the mechanism of failure is mass failure, subsurface
processes entrainment, or channel avulsion. Best when toe erosion is the primary process.
Flow velocity Well-anchored structures have been successfully applied to situations with estimated
or shear velocities ~2.5 meters/second (D’Aoust and Millar 2000). Rootwad installations have
stress withstood velocities of 2.7 to 3.7 meters/second (Allen and Leech 1997). ELJ-type
structures withstood 1.2 meters/second in a sand-bed stream (Shields et al. 2004) and
flows that produced estimated mean boundary shear stresses of 50 to 170 N/m2 1.0 to
3.5 lbs/sq ft (Abbe and Brooks 2011).
Site access Heavy equipment access to bring in and place large trees with rootwads is needed for all
but the smallest project.
Conveyance Large wood structures can increase flow resistance if they occupy significant parts of the
channel prism (Shields and Gippel 1995).
Navigation Design should minimize potential hazards to commercial or recreational navigation.
and Potential hazards are greatest for structures that span the channel.
recreation
Raw Suitable sources of adequately sized logs needed within economically feasible haul
materials distance.
Risk Situations where failure would endanger human life or critical infrastructure call for
rigorous risk analysis and higher safety factors.
Source: Fischenich and Morrow 2000.

Large Wood National Manual July 2015


6-8
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

GUIDANCE 6.4 Design Life of


Regional Considerations in Large Wood Design
Placed Wood
Large wood naturally decays and breaks into
The basic physics of water, sediment, and wood are
universal. However, wood decay rates are only one of smaller pieces. Large logs that are subjected to
the important regional differences facing large wood wetting and drying cycles may last only a couple
project planners and designers. Workers in other of years in hot, humid climates but decades or
regions should be alert to differences between their even longer in cooler regions like the Pacific
project context and constraints and those found in the Northwest. Accordingly, large wood placement
Northwest. Among these are hydrology,
should be viewed as a “transitional
geomorphology, tree size, riparian land use and land
ownership, aquatic ecology, endangered species, and rehabilitation technique” (Lester and Boulton
patterns of recreational and navigational use of 2008), and should not be attempted if reach-
waterways. As noted by Shields et al. (2004): and watershed-scale geomorphology are not
conducive to sediment retention, woody plant
Design of large wood structures…has been described
for gravel-bed rivers … in the Pacific Northwest. colonization, and stabilization of sediments
Placing structures in incised, sand-bed channels of retained by wood structures (Shields et al.
smaller streams typical of the Midwestern and 2008). A long-term goal of a large wood project
southeastern U.S. presents a different set of is to replace the natural wood source—the
challenges. In addition to basic differences in ecology, riparian forest—and associated processes that
available wood tends to be smaller, material coarser
will naturally replenish instream large wood
than fine gravel for ballast is unavailable, and
channel erosion rates (relative to channel width) are and on the floodplain (Abbe and Brooks 2011).
higher. Channel width–depth ratios are an order of
Decay rates for logs that are periodically wetted
magnitude smaller ~typically, 10), so storm flows
tend to be deep, and structures are more frequently and dried vary radically with tree species
submerged. Bed slopes and current velocities are (Table 6-2) and with the wetting frequency,
typically lower in sand-bed systems. ambient temperatures, and humidity due to the
requirements of the fungi responsible for
Additional concerns attend the presence of ice loading
and ice flows in northern states and the much wider aerobic decomposition of wood (Harmon et al.
difference between base flows and high flows for much 1986). These decay rates are accelerated with
of the rest of the country than those found in the increasing temperature and precipitation.
Northwest. Regional differences exist in the likelihood Scheffer (1971) developed the following index
of major wood loading associated with hurricanes, for comparing potential decay rates of above-
tornadoes, or avalanches.
ground wood structures in different climatic
Because much of the research on and implementation regions of the United States.
of large wood projects to date have been in the Pacific
Northwest, there is a strong regional bias in the Equation 6-1:
literature. Sources of regional design information
include the following:
 Pacific Northwest: Knutson and Fealko (2014)
 Lower reaches of the Sacramento and San Joaquin
Rivers and their tributaries: ICF International
(2010) where T is the mean monthly temperature (°F),
 Southeastern Coastal Plain: Shields et al. (2004 and
Dp is the mean number of days in the month
2008)
 Appalachians: Hilderbrand et al. (1998) with 0.03 centimeter (0.01 inch) or more of
 Australia: Brooks (2006) precipitation, and the summation represents
the sum of products for all of the months of the
year. The sum is divided by 30 to make the

Large Wood National Manual July 2015


6-9
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

index fall between 0 and 100 for most of the is accelerated by periodic wetting and drying.
United States. Shields et al. (2008) reported breakup and
decay of large wood structures comprising
For example, Scheffer (1971) computed values
primarily deciduous species in a flashy,
of 82.5, 44.8, and 22.0 for Atlanta, Georgia, Des
sand-bed Mississippi stream within 3 years.
Moines, Iowa, and Casper, Wyoming,
Cederholm et al. (1997a, 1997b, 1997c)
respectively (Figure 6-3). This implies a wood
reported significant degradation of partially
structure would last about four times longer in
submerged red alder logs in a western
a climate typical of Wyoming than one typical of
Washington stream after 3 years, but little
Georgia, all other factors being equal.
degradation of conifer logs in an adjacent reach
Hardwood species decay very slowly if of the same stream.
continuously wet (Bilby et al. 1999) while decay

Figure 6-3. Climate Index for Wood Decay Hazard

Source: Forest Products Laboratory 2010

Large Wood National Manual July 2015


6-10
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Table 6-2. Comparison of Desirability of Various Tree Species for Stream Structures
Durability
Species (assuming wetting and drying) Source1
Cottonwood (Populus spp.) Poor Johnson and Stypula (1993)
Alder (Alnus spp.) Poor Johnson and Stypula (1993)
Cederholm et al. (1997a,
1997b, 1997c)
Maple (Acer spp.) Fair (will survive 5 to 10 years) Johnson and Stypula (1993)
Hemlock (Tsuga spp.) Least of conifers Johnson and Stypula (1993)
Sitka spruce (Picea sitchensis) Excellent Johnson and Stypula (1993)
Douglas-fir (Pseudotsuga spp.) Excellent, will survive 25–50 Johnson and Stypula (1993)
years
32–56 Harmon et al. (1986)
Western red cedar (Thuja plicata) Most desirable, will survive 50 to Johnson and Stypula (1993)
100 years
Yellow-poplar (Liriodendron tulipifera) 0.4 year Harmon et al. (1986)
Aspen (P. tremuloides) 5 years Harmon et al. (1986)
White fir (A. concolor) 4 years Harmon et al. (1986)
Norway spruce (Picea abies) ~30 years Kruys et al. (2002)
Conifers (P. sitchensis, T. heterophylla, P. Half-life of ~20 years Hyatt and Naiman (2001)
menziesii, T. plicata)
Black locust (Robinia pseudoacacia), red Exceptionally high heartwood Simpson and TenWolde
mulberry (Morus rubra), Osage orange decay resistance (1999)
(Maclura pomifera), Pacific yew (Taxus
brevifolia)
Old growth bald cypress (Taxodium Resistant or very resistant to Simpson and TenWolde
distichum), catalpa (Catalpa spp.), cedars heartwood decay (1999)
(Cedrus), black cherry (Prunus serotine),
chestnut (Castanea spp.), Arizona cypress
(Cupressus arizonica), junipers (Juniperus
spp.), honey locust (Gleditsia triacanthos),
mesquite (Prosopis glandulosa), old
growth redwood (Sequoia sempervirens),
sassafras (Sassafras albidum), black
walnut(Juglans nigra)
Young growth bald cypress (Taxodium Moderately resistant to Simpson and TenWolde
distichum), western larch (Larix heartwood decay (1999)
occidentalis), longleaf old growth pine
(Pinus palustris), old growth slash pine
(Pinus elliottii), young growth redwood
(Sequoia sempervirens), tamarack (Larix
laricina), old growth eastern white pine
(Pinus strobus)

Large Wood National Manual July 2015


6-11
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Durability
Species (assuming wetting and drying) Source1
Red alder (Alnus rubra), ashes (Fraxinus Slightly or nonresistant to Simpson and TenWolde
spp.), aspens (Populus tremuloides), beech heartwood decay (1999)
(Fagus spp.), birches (Betula spp.),
buckeye (Aesculus glabra), butternut
(Juglans cinerea), cottonwood (Populus
spp.), elms (Ulmus spp.), basswood (Tilia
Americana), true firs (Abies spp.),
hackberry (Celtis occidentalis), hemlocks
(Tsuga spp.), hickories (Carya spp.),
magnolia (Magnolia grandiflora), maples
(Acer spp.), pines (Pinus spp.), spruces
(Picea spp.), sweetgum (Liquidambar
styraciflua), sycamore (Platanus
occidentalis), tanoak (Notholithocarpus
densiflorus), willows (Salix spp.), yellow-
poplar (Liriodendron tulipifera)
1 Information from Johnson and Stypula (1993) is qualitative and unsubstantiated. Evidently these comments pertain to
the region of King County, Washington. Harmon et al. (1986) provide a review of scientific literature dealing with
decomposition rates of snags and logs in forest ecosystems. The times from Harmon et al. (1986) represent the time
required for 20% decomposition (mineralization) of a log based on exponential decay constants obtained from the
literature. Fragmentation of logs in streams due to mechanical abrasion would accelerate the decay process, as would
more frequent wetting and drying. Kruys et al. (2002) provide data on decay of fallen and standing dead trees in a forest
in mid-northern Sweden. Hyatt and Naiman (2001) provide data on residence time of large wood in Queets River,
Washington. Simpson and TenWolde (1999) provide data for evaluating wood products, not whole trees. Additional data
on lumber (not large wood) are available from Forest Products Laboratory (2010).

Effects of the red alder logs on habitat were effective design life of “50 years or more.” Wohl
projected to disappear within 5 years of (2013) reviewed available literature and found
placement, but the conifer structures had a that the decay rates of logs on a forest floor are
design life of 25 years. Hertzberg (1954) as follows:
reported that fence-type wooden revetments
made of wood impregnated with preservative Time for Full Decay
(creosote) had a design life of 20 years along Climate (years)
the Lower Mississippi River, while structures Cold Boreal/Subarctic >100
made of cylindrical bundles of fresh willow Dry 50–100
brush exhibited “rapid deterioration.” Humid-Temperate 10–100
In sharp contrast to the above, certain natural Tropical <10
log accumulations have been shown to be stable
for centuries in the Pacific Northwest and in
Wohl (2013) further noted that although the
Australia (Abbe and Montgomery 1996; Nanson
rates of decay for waterlogged instream wood
et al. 1995). Conifers in the Pacific Northwest
may be slower due to anaerobic conditions, the
and eucalypt species in Australia are decay-
relative rates of decay between regions based
resistant relative to other species. In general,
on forest-floor decay rates likely hold for
decay rates are lowest for species with high-
instream wood.
density wood. Writing about decay-resistant
wood placed in Australian rivers, Brooks (2006) Clearly, large wood species, climate, local
suggests that a “well designed structure in the hydraulics, and reach hydrology all play a role
right conditions” may be expected to have an in the stability of natural wood and large wood

Large Wood National Manual July 2015


6-12
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

structures. It is very important to note that


decaying wood loses mass and volume at an
6.5 Level of Design
exponential rate (Harmon et al. 1986): Effort
Equation 6-2: Fluvial systems and their interactions with
riparian vegetation and large wood are
exceedingly complex. The engineer must devote
sufficient effort to planning and design of large
where wood structure projects to achieve an
Y0 = initial quantity (density, mass, or acceptable standard of care and rate of success
volume) of wood without wasting resources on excessively
elaborate analyses. As described in Table 6-3,
Yt = amount left at time t (year), and the levels of design effort used for instream
large wood structures are used to determine
k is a decay-rate constant (year-1)
the context and expertise needed to complete
Harmon et al. (1986) reported species-specific an analysis. In some cases it may be wise to
values of k for logs lying on a forest floor range confine feasibility studies to Levels I and II, with
from 0.004 to 0.52 year-1, which implies that Level III used in preliminary and final designs.
logs of these species will lose 1–63% of their Pencil and paper analyses are rapid and
density within only 2 years. A more recent relatively cheap while spreadsheets are
study of six species in Sweden documented k dynamic and allow for sensitivity analyses in
values ranging from 0.039 to 0.102 year-1 the computation of applied and resisting forces
(Freschet et al. 2012).Therefore, using or costs when key parameter values are
appropriate k values, Equation 6-2 may be used uncertain and only a range may be specified.
to predict the relative density or mass of large Only numerical simulation allows the engineer
wood after a given period of time (Abbe 2000; to approach understanding the impacts of
Abbe et al. 2003b; Abbe and Brooks 2011). structures on flows and sediment movements at
the reach scale. Accordingly, modeling allows
In some cases, the design life of components
the designer to view the impacts (e.g., on
other than wood should be considered when
conveyance) of changing the numbers,
estimating the design life of a large wood
locations, or sizes of structures placed in the
installation. For example, fiber rope may decay
study reach. Additional modeling may allow
faster than wood, or metal hardware may
quantification of habitat benefits, although
corrode and weaken.
biotic responses may not follow habitat
Living plant materials such as willow cuttings improvements.
may be used in a large wood project to
promptly establish vegetation on sediment
deposits induced by large wood structures or to 6.6 Design Decisions
revegetate riparian areas disturbed during and Data
construction or pre-project erosion. Guidance
such as that of the Natural Resources Requirements
Conservation Service (NRCS) (2007a) can be Large wood structure design may be viewed as
quite helpful regarding use of such measures. a series of data gathering and analysis exercises,
each of which leads to a decision point.
Feedback is required as certain design decisions
affect previous design decisions. Key design
issues are outlined in Table 6-4.

Large Wood National Manual July 2015


6-13
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 5. Engineering Considerations

Table 6-3. Levels of Design Effort for Instream Large Wood Structures
Level Description and Example Context Relative Risk Level
I Pencil and paper Small stream, experienced designer, overall Low
risk to infrastructure or human life small,
small equipment or hand tools, construction
done under arrangement (i.e., hourly hire)
that freely allows adjustment of design in the
field.
II Spreadsheet, Bank Small to medium stream, overall risk to Medium
Stability and Toe Erosion infrastructure or human life small, heavy
Model (BSTEM) equipment, construction done under
contract that requires development of plans
and specifications.
III 1D and 2D numerical Medium to large stream, design by team of High
simulation including specialists across range of disciplines,
geomorphic response significant risk, construction under contract
streambank stability and that requires development of plans and
other types of specifications. See below for additional
geotechnical engineering discussion on numerical model selection.

GUIDANCE

Minimum Data Requirements for Successful Designs (Brooks 2006)

1. Cross-section surveys including representative sections spaced at no more than one channel width, ideally at
each structure location, with a minimum of 10 per reach to try to capture more than one complete riffle-pool
sequence (if they exist).

2. Thalweg profile survey (at least three riffle-pool sequences or 15 to 20 channel widths long). This information
will help determine the reach bed slope (i.e., as a regression line passing through crossing or riffle crests).

3. One bed material sample from the center of each cross section.

4. Streamflow data or regional regression relations for flow frequency.

5. Desired wood loading.

6. Wood dry density and approximate sizes of available wood.

Large Wood National Manual March 2015


6-14
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Table 6-4. Key Engineering Issues for Instream Large Wood Structure Placement

Category Decisions
Hydrology What is the design event? How will the structures affect/interact with smaller and
larger flows? Should ice be considered in the design?
Reach layout How many structures will be placed and where?

Materials What types and sizes of logs and other materials will be used? Sources?

Structure dimensions What type/shape of structures will be employed? What will their dimensions be?
and details
Hydraulics How will the project affect habitat quality and high flow stages?

Sediment What effect will the project have on local scour and deposition, bank erosion, reach
scale morphology, channel response, habitat value, and terrestrial plant
colonization?
Vegetation How much effort should be devoted to planting vegetation? Should effects of
vegetation on structural stability (surcharge, sediment cohesion), erosion, and
bank stability be included in analysis, and, if so, how?
Anchoring What is the magnitude of forces that the structures must be designed to resist?
Will anchoring involve passive or active restraints? What factors of safety will be
used?
Construction What construction methods will be utilized? Will channel be de-watered or large
wood placed in the “wet”? What adverse impacts will be created by construction,
and how can they be controlled? What time windows (seasons) will be used for
construction?
Economics Can the project be delivered within budget? How can value be increased?

larger events will impose larger shear forces


6.6.1 Hydrology (drag, lift, buoyancy) on the structures.
However, peak velocities may be associated
6.6.1.1 Determination of Design with more frequent (e.g., bankfull) events. In
Discharge some systems, the design condition may
correspond to prolonged high flows, while in
Features within stream corridors are normally other cases driving forces will be greatest
designed to withstand loadings imposed by during a sudden rise in stage after a long
discharges with a certain frequency or return drought that desiccated the wood (Abbe and
interval such as the 100-year, 10-year, or 2-year Brooks 2011). In such a case, the buoyant forces
event. Computation of hydraulic loadings and wave drag would reach maxima when the
imposed by a design event requires knowledge structure is initially overtopped (Shields and
of the discharge associated with the design Alonso 2012). Design discharge selection
event frequency. Design discharges may be should be consistent with the risk analysis and
determined by statistical analyses of data from be completed prior to design (see Chapter 7,
nearby gages or using other techniques Risk Considerations). The expected frequency of
described below. the design event should be compatible with the
design life of the structures and the planned
The design event return interval may be based
intensity of maintenance.
on the design life of the structure (Knutson and
Fealko 2014); the tacit assumption here is that

Large Wood National Manual July 2015


6-15
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

The ratio of more frequent to less frequent Turnipseed (2010) provide details on stage and
discharges may be considered. For example, in discharge measurement at gaging stations.
some systems the difference between the Q2
(2-year return interval discharge) and Q25 is
small, while in others it is an order of GUIDANCE
magnitude or more. So in some cases the
difference between designing for the smaller
StreamStats
and the larger event is small. In some cases,
large wood design for Q100 may not be The StreamStats software is a web-based GIS tool
economically or technically feasible. In the (http://water.usgs.gov/osw/streamstats/) that can
Pacific Northwest it is not unusual for designs be used to obtain streamflow statistics, basin
to be required by regulators to be stable at or characteristics, and other information for
above the 100-year flood level. Streams with user-selected sites on streams to aid in regional
regression analysis. Specific capabilities of the
large variability usually feature flashy
National Streamflow Statistics Program include:
hydrology, and large wood structures are
repeatedly wet and dried, accelerating decay  Estimate rural and urban flood-frequency
and deterioration and shortening design life. discharges for ungaged streams by use of
regression equations, or for six states, by
Hydrologic Data from Gaged Sites region-of-influence analysis.

Ideally, a gaging station is located near the  Estimate a wide range of low-flow duration and
project site with an established gaging record of frequency discharges for ungaged streams.
mean daily flow (typically an average of all the  Estimate discharges for natural streams. The
15-minute interval flows reported in a day). The program does not account for the effects of
U.S. Geological Survey (USGS) is the primary water diversions, dams, flood-detention
source of mean daily flow data in the United structures, and other human-made works.
States, with gaging location information and
 Statistically weight estimated peak discharges
data available online
for ungaged sites with drainage basins that span
(http://water.usgs.gov/osw/). Data from gages
multiple hydrologic regions using the
operated by the state, local municipalities, percentage of drainage area in each region
water districts, and hydropower companies within a given state.
may also be available.
 Statistically weight estimated and observed
Collecting Data from Ungaged Sites peak discharges for stream gaging stations using
the equivalent years of record of the regression
Depending on the scale of the project and time estimate and the number of years of observed
available to collect new data, it may be practical record as the weighting factors.
to collect data at least for a few high-frequency
 Statistically weight estimated peak discharges
events. The basic approach to developing a new
for ungaged sites obtained from regression
stream gaging record is to correlate equations and from the flow per unit area for an
observations of flow stage with discharge upstream or downstream gaging station.
measurements made at the same time to
develop a stage-discharge relationship. Multiple  Plot hydrographs of flood and low flows.
flow measurements are required to create  Generate frequency graphs for both high- and
enough data points spanning low to high flows low-flow frequency analyses.
through which a stage-discharge curve could be
drawn. Stream gaging is described in textbooks
and other restoration handbooks and will not
be described further here. Sauer and

Large Wood National Manual July 2015


6-16
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Flow Frequency Analysis site. If drainage areas for the two gages have
similar terrain and land cover, discharges from
Once a design discharge frequency is selected, the gaged site may be transferred to the
standard analyses may be used to derive the ungaged site using a ratio of drainage areas or
discharge magnitude using gage data. The more sophisticated approaches as described by
designer should check on the adequacy of NRCS (2007f) and Saur (1974).
available data. A minimum of 10 years of peak
discharge data is required under Bulletin 17B Also for ungaged sites, discharges of a given
guidelines (IACWD 1982), and 30 or more years frequency may be estimated using region-
of data are preferred. Estimating flood specific regression formulas that use watershed
frequencies for recurrence intervals more than characteristics as dependent variables (National
twice the annual series record length is Streamflow Statistics Program2). Many of these
cautioned against. Longer periods of record formulas are included in StreamStats. 3 In
reduce the need for extrapolation to determine addition to application of the USGS tools,
infrequent return period discharges, and thus regional regression equations applicable to the
increase the certainty in the estimate. However, project may have been developed by others.
the designer should evaluate long periods of The designer is cautioned that the empirical
record for time-homogeneity (i.e., stationary) to regression equations are based on statistical
establish that the causative hydrological models, and must be applied within the limits of
processes remain consistent over the annual the data used to develop the equations with the
series and that two events of the same acknowledgement of associated scatter in the
magnitude in the annual series are likely to data (Gotvald et al. 2012). StreamStats provides
occur at any time in the series (Subramanya confidence limits for peak discharge estimates.
2008). Altered hydrology arising from changed For example, standard error values from the
land use conditions and actions such as dam Sierra Nevada hydrologic region were reported
construction, changes in reservoir operations, by the USGS as ranging from 51.5% for a 4%
or water diversions could systematically change exceedance probability to 74.4% for a 50%
peak flow values, which must be accounted for exceedance probability (Gotvald et al. 2012).
in the frequency analysis by only including the Discharge values from regression formulas may
most recent continuous homogeneous portion be checked against uniform flow computations
of the annual series. Likewise, long-term (e.g., Manning formula) based on survey data
changes to peak flows due to climate change and appropriate resistance coefficients.
and extreme events that could occur must be Rainfall-Runoff-Routing Modeling
considered in the implementation and
interpretation of the flood frequency analysis. Rainfall-runoff-routing models can be used
where gaging data is not available or reliance
Procedures for frequency analysis of discharge
on historical data or regional regression
data are outlined in textbooks (e.g., Subramanya
equations alone is not sufficient. Results from
2008; Eslamian 2014) and other river
hydrologic models can be compared with
restoration handbooks (e.g., NRCS 2007f) and
separate flood frequency estimates, and offer
will not be described further here.
the benefit of evaluating hydrologic conditions
If gaging data are unavailable or the period of for future conditions where climate and land
record is insufficient, then alternative use may vary appreciably from the historic
techniques are required. If a stream gage is conditions of the gaging record. Numerous
located upstream or downstream of the project rainfall-runoff-routing models are available,
reach with 10 or more years of record, it may be
2 Available at http://water.usgs.gov/software/NSS/
possible to transfer information from the gaged
3 Available at http://streamstats.usgs.gov/.

Large Wood National Manual July 2015


6-17
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

including the relatively easy to use Soil base flow is readily observed as the slowly
Conservation Service’s curve number based decreasing flow on the receding portion of the
WinTR-55 (for areas <65 square kilometers hydrograph that reaches a minimum prior to
[25 square miles]) and without snowmelt the addition of direct runoff from the next
capability), USACE’s Hydrologic Modeling rainfall or snowmelt event. The designer should
System, and the U.S. Environmental Protection evaluate how base flow conditions may change
Agency’s (EPA’s)/USGS’ Hydrological between wet and dry water year types to
Simulation Program-Fortran model. These ensure that elevations of the structure intended
models require information on watershed land to be continually submerged will in fact be
use, topography, soil characteristics and submerged.
infiltration, storage, and other variables to
transform design precipitation events into Exceedance or Flow Duration Curve
runoff hydrographs that are routed through Analysis
channel networks. The designer must use
intensity-duration-frequency data to determine Exceedance analysis is performed to determine
the precipitation hyetograph of a specified the probability that a flow of a particular
magnitude is equaled or exceeded based on
design storm frequency and duration and use
statistical analysis of the flow record. A plot of
the rainfall-runoff-routing model to calculate
volume, stage, and peak flow. discharge versus the percentage of time the
discharge is equaled or exceeded at a given site
The frequency of the design storm event is is called a flow duration curve. The analysis is
commonly assumed to be approximately equal typically performed on the mean daily flow
to the frequency of the design flood event. record available from a stream gage or
However, this is not always true due to determined from a rainfall-runoff simulation
watershed complexities and variations in storm model or other method. For flashy streams
intensity and duration. Instead of calculating exhibiting rapid changes in flow magnitude
peak flows from all different types of 100-year over the course of several hours as opposed to
storms, it is common practice that for smaller days, the 15-minute flow record can be used in
watersheds the storm duration is set to equal the analysis instead of daily flows.
the time of concentration, whereas in larger
watersheds with times of concentration over Procedures for developing flow duration curves
are presented in standard texts (e.g.,
1 hour it is common to only evaluate the 6-hour
and 24-hour storm (Viessman and Lewis 2003). Subramanya 2008) and handbooks (e.g., NRCS
2007f) and will not be described here. The
entire measured or simulated flow record can
6.6.1.2 Hydrologic Design for be used in the analysis; however, the same data
Habitat considerations discussed above for peak flow
analysis also apply, namely analysis of the data
Base Flow to establish that the causative hydrological
Base flow is typically the minimum flow in the processes remain consistent and time-
stream supplied by groundwater and release of homogeneity is achieved. Factors that have
water stored in the channel’s banks. In affected the long-term flow record, such as flow
regulated systems base flow conditions can be regulation, land use change, or climate change,
altered by controlled reservoir flow releases or should be accounted for in the period of record
water diversions or inputs. Various techniques selected for the analysis. Furthermore, the
are available for determining base flow designer should evaluate the length of the flow
conditions from hydrographs (Subramanya record and determine if it contains a
2008). In many perennial stream hydrographs, representative sample of wet and dry water

Large Wood National Manual July 2015


6-18
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

years, and consider performing separate departures from the data trend indicative of a
analyses for different water year types to better non-stationary flow record.
understand flow variability between water
Performing an exceedance analysis of the flow
years. For cases where a gaging record has
record on a monthly (or other desired time-
incomplete data for a particular water year, it is
step) basis is helpful for assessing the habitat
common to exclude all of the data for the year
value of a wood structure during various
to prevent skewing of the probability analysis.
seasons. Ecological events or seasons tied to life
Separate exceedance analyses can be performed
cycles for species of interest may be displayed
on different periods of time in the flow record
on an annual hydrograph showing various flow
(e.g., 10- to 20-year increments) to test for
exceedance levels and key project attributes to
assess habitat performance (Figure 6-4).

Figure 6-4. Graphical Output of Mean Daily Flow Monthly Exceedance Analysis and Project-Specific
Salmonid Life Stages

costs, it is important to strategically size and


6.6.2 Reach Layout place wood structures. Initial plans may be
Reach layout refers to the arrangement of large based on study of target reaches. A target reach
wood within the channel. Approximate is defined here as a lightly impacted reach with
locations of proposed large wood placements similar width, depth, slope and grain size to the
should be recorded on hard copy or digital project site. There are no true target reaches
maps of the project reach. Because the number due to the ubiquity of human influences,
of structures will directly influence project although our understanding of pre-European

Large Wood National Manual July 2015


6-19
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

settlement conditions indicates that rivers of an island, in mid channel (to foster
North America had levels of wood loading much development of a bar or island), along the bank
greater than even contemporary heavily loaded on the outside of a bend, at the upstream
systems (see Chapter 1, Large Wood entrance to a side channel, fully spanning the
Introduction, and Chapter 4, Geomorphology and channel, or secured on floodplain surfaces (to
Hydrology Considerations). A target reach must reduce potential for channel avulsions by
be selected with reference to the project goals increasing flow resistance or to serve as
(e.g., habitat rehabilitation, erosion control). instream wood after future channel shifting
Target reach levels and types of wood loading occurs) (Cramer 2012).
may be used as design analogs for the amount
Although the specific types of large wood
and distribution of wood in the design reach.
structures should be selected in the next design
Reach layout should tie in heavily with project step, a general determination must be made for
geomorphic objectives: are wood structures reach layout. Large wood placement usually has
intended to facilitate or mitigate channel a primary goal of improving habitat quality by
avulsion and braiding; engage side channel adding woody substrate, cover, scour pools, and
development; and control bank erosion, store physical heterogeneity. Large wood structures
sediments, or trap wood? How much may be intended to address either vertical
aggradation or incision is currently occurring, (bed) or horizontal (bank) erosion processes.
and how will the large wood interact with that Cramer (2012) noted the variation of large
process? As for naturally occurring large wood, wood configuration with stream size and the
structures may be placed at the head of a bar or associated function and risk (Table 6-5).

Table 6-5. Recommendations for Placement of Large Wood in Streams for Aquatic Habitat Benefits

Stream Width Large Wood Structure Natural Large Wood


Size (meters) Functions and Risks Configurations
Small <10 Single or multiple pieces of wood can be effectively Logs most often lie
used to create habitat, stabilize the channel, dissipate perpendicular or are angled
energy, and store sediment. Logs in small streams downstream to flow, but
may be used to create step pools (i.e., plunge pools). any orientation is feasible.
Because small streams generally have less energy to They may span the channel
move large wood, a greater variety of large wood or intrude partway into the
locations and orientations can be employed without channel.
excess risk.
Medium 10–20 Channel-spanning wood structures may be applicable, Wood tends to accumulate
but the results are less predictable than for small in jams, but single pieces
streams and their vulnerability to flood damage is and small complexes also
relatively high. occur. The outside of bends
and the head of natural
gravel bars tend to be
relatively stable locations
for wood jams.

Large Wood National Manual July 2015


6-20
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Stream Width Large Wood Structure Natural Large Wood


Size (meters) Functions and Risks Configurations
Large >20 Stabilizing woody debris becomes a significant Lateral jams, as opposed to
concern on larger streams. Wood placement in the full-spanning jams, are a
main stem of the channel is only recommended in the common feature. As with
form of anchored structures (i.e., log jams, large wood medium-sized streams,
complexes, and wood trapping structures), unless locations at the outside of
transport can be tolerated. Key pieces and log bends and the head of
complexes can be effectively used in side channels and natural gravel bars tend to
floodplain habitats. be relatively stable.
Source: Saldi-Caromile et al. (2004); Cramer (2012).

6.6.2.1 Bed Control spur-type structures (Figure 6-6). Spacing for


intermittent structures is normally expressed as
Bed control structures (Figure 6-5) should be a multiple of the length of the structure from
spaced so that backwater from one structure bank to riverward tip, measured perpendicular
reaches the next structure upstream during to the approach flow (this distance is called the
channel-forming flows (about Q2 for many projected crest length or effective length or
channels). Bank erosion control structures may structure protrusion width) (Table 6-6).
be continuous blankets or intermittent,

Figure 6-5. Large Wood Bed-Control Structures

Photo a. Natural channel-spanning large wood, Trail Photo b. Constructed bed control large wood structure in
Creek in the Snowy Range of the Medicine Bow Australia (Andrew Brooks).
Mountains, Wyoming (Claire Ruffing).

Large Wood National Manual July 2015


6-21
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Figure 6-6. Continuous and Intermittent, Spur-Type Large Wood Structures

Photo a. Continuous blanket type structure. Cabled spruce Photo b. Intermittent large wood structures; Little
trees and brush layering immediately after installation, Topashaw Creek, Mississippi, showing sediment deposition
Ciechanski Recreation Site, Kenai River. Alaska at toe of eroding bank induced by structures.
Department of Fish and Game.

Table 6-6. Criteria for Spacing Intermittent Large Wood Structures along the Outside of Meander
Bendsa

Channel Planform Large Wood Structure Spacing Source


Rc/W > 3 3 to 5 x projected crest length Sylte and Fischenich (2000)
Rc/W < 2.5 Spacing goes to zero—use
continuous type structure
Tight bends 3 x projected crest length Drury et al. (1999), Brooks (2006)
Straight reaches 5 x projected crest length
All 1.5 to 2.0 x crest length Petersen (1986) in Shields et al. (2004)
All 2/3 to 2.5 times the length of the Pokrefke (2013)
upstream structure
Rc/W = bend radius of curvature divided by channel top width
a These design criteria are extracted from works guiding placement of river training and bank protection structures

designed to produce channel stability. If higher levels of dynamism are desired or tolerable, spacing should be increased.
Erosion between widely spaced structures may lead to flanking (river avulsion around the land side of structure).

6.6.2.2 Bank Protection in Lagasse et al. 2009). Although large wood


may be effectively employed to control bank
For erosion control objectives, reach layouts erosion and protect infrastructure (Figures 6-7
may be designed using guidelines for traditional and 6-8) (Abbe et al. 1997, 2003c; Shields et al.
spur dikes and groins (Ahmad 1951; Copeland 2004), well-designed placements also offer
1983; Klingeman et al. 1984; Design Guideline 2 aquatic habitat benefits (Shields et al. 2006).

Large Wood National Manual July 2015


6-22
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Figure 6-7. ELJ Spacing to Protect Road and Enhance Habitat Along the Cispus River

Gifford Pinchot National Forest, Washington—project was constructed in 1999 and had been subjected to a
25-year and two 10-year flood events by 2014. Structures successfully established a forest buffer between
highway and river.

Large Wood National Manual July 2015


6-23
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

may be used for an initial estimate of the


Figure 6-8. Example of the Use of Two Sets of
required downstream extent of bank protection.
ELJs in Bank Protection Along Hoh River
The initial extent of bank protection determined
from Figure 6-9 should be adjusted according to
field observations of active scour, channel
surveys at low flow, and aerial photography and
field investigations at high flow. Investigators of
field installations of bank protection have found
that protection commonly extends farther
upstream than necessary and not far enough
downstream.
Figure 6-9. Recommended Extent of Riprap
Revetment for 110o Bend

Source: USACE (1981) in Lagasse et al. (2009).

More aggressive designs have used large wood


structures to split a main channel into smaller
channels prior to reaching the eroding bank
where additional large wood structures are
positioned (Figure 6-8). For rivers that have a
sufficient floodplain corridor this approach is a
sustainable means of balancing restoration and
flood protection. An array of ELJs that increases
in density with distance from large migrating
rivers will split the channel into smaller
The Hoh River is a large gravel bedded river draining
the Olympic Mountains of Northwest Washington—the
anabranches and diminish erosive energy to
first set is laid out upstream of eroding bank to deflect effectively limit the extent of channel migration,
flow into chute channels across point bar. The second is restore floodplain forests, and protect areas
placed along outer bend similar to series of spur dikes. that might otherwise be at risk of eroding
Continuous or intermittent structures should (Figure 6-10) (QIN 2008). This strategy
cover the entire zone of potential erosion. Many diminishes the stream power reaching the site
streambank projects fail because protection of concern in a way that restores the channel
was not extended far enough upstream, and floodplain complexity the river once had
downstream, or into the bed. Protection should (Figures 6-10 and 6-11) (QIN 2008).
be extended well past the anticipated zone of
current attack during design events. Figure 6-9

Large Wood National Manual July 2015


6-24
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Figure 6-10. Valley Scale Restoration Approach to Limiting Bank Erosion Along Valley Margins

Density of roughness elements (i.e., ELJs) increases toward the margin of the valley. This type of layout
breaks up the channel into smaller and smaller channels or anabranches with distance from the main channel
(from QIN 2008)

Large Wood National Manual July 2015


6-25
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Figure 6-11. Upper Quinault River Valley Floodplain and Side Channel Restoration

Array of ELJs constructed on large point bar in 2008 (depicted in LiDAR DEM at top). Main channel is in upper
right and flowing to right to left (west) By 2012 the river migrated into ELJ 08-10, forming a new pool, added
more than 100 feet of racked wood, initiated new side channels, and two new logjams formed. By 2013 the
side channels are more pronounced and ELJs are forming forested islands.

Large Wood National Manual July 2015


6-26
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

6.6.3 Select Types of GUIDANCE

Structures
Key Considerations for Selecting the Types of Large
Reach layout and selection of large wood Wood Structures for a Given Reach
structure type are closely related, and some
 The configuration should address the dominant
cycling back and forth between these two fluvial (erosion, deposition, etc.) processes
decision steps is usually required. More than operating on the site.
one type of structure should be used in a reach,
and structure type should be matched to the  Key habitat deficiencies (e.g., lack of pools, lack
local morphology and desired functions. of cover, lack of woody substrate) should be
addressed. These should have been established
Structures that protrude into the flow like in accordance with principles described in
logjams, weirs, or spurs tend to create greater Chapters 3 and 4 of this manual.
habitat diversity than those that parallel banks  The project should be in harmony with the
like revetments. When selecting structures that anticipated future geomorphic and riparian
fully, span the channel, avoid using single-log response of the reach.
weirs, they are subject to undercutting and have
no redundancy should the log fail. The more  Economic, political, institutional, social, and
construction access issues should be considered.
logs used, the stronger the structure and
greater the factor of safety. Whether using a  Suitable materials must be available at a
step-pool or reinforced riffle design, it is reasonable cost. “Key” logs of adequate size to
important to minimize the magnitude of be naturally stable without anchoring may not
individual drops and thus create broad-crested be available, and designs must be modified
structures (Table 6-7; Figure 6-8). This typically accordingly.
increases the cost, but greatly increases  Safety issues for recreational use of the
structure stability and enhances fish passage. In completed project reach should be addressed, if
steep step-pool or cascade channels this may appropriate (Chapter 7, Risk Considerations).
entail placing wood throughout the length of
 The most desirable types of structure emulate
the stream. Considerable research has been
naturally occurring large wood formations.
recently completed on step-pool streams (e.g.,
Permanently fixed structures placed at regular
Comiti and Mao 2012). intervals for erosion control are often necessary
but do not replicate features typical of natural
Two schools of thought exist on large wood
settings. When possible (i.e., when dynamic,
structure design and typology: one relies on
mobile boundaries and wood are acceptable),
emulation of natural large wood formations structures should look and behave like stable
(jams) observed in the Pacific Northwest (Abbe wood jams.
et al. 2003b), while the other is loosely based on
more traditional river training structures
(Shields and Wood 2007). A combination of
CROSS-REFERENCE
these two schemes is presented in Table 6-7.
Uniform spacing and structure configuration
should be avoided in favor of variation in the See Chapter 1, Large Wood Introduction, and
frequency, size, and type of structure applied Chapter 4, Geomorphology and Hydrology
Considerations, for images and descriptions of
(Erskine et al. 2012).
natural large wood formations.

Large Wood National Manual July 2015


6-27
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Table 6-7. Classification of Large Wood Instream Structures Based on Architecture4

Functional Role and


Configuration Sketch Description Strengths and Weaknesses References
ELJs or flow Intermittent structures built into Emulates natural formations Abbe et al. 1997;
deflection jams eroding banks by stacking whole if dimensions and spacing Drury et al. 1999;
trees and logs with rootwads in vary. Creates diverse Drury 1999;
crisscross arrangements. Often physical conditions, traps Shields et al. 2004;
filled with gravel or cobble as additional debris. Suitable Brooks 2006;
ballast. Large quantities of smaller for banks subject to mass Brooks et al. 2006
wood (racked debris) may be failure.
added to upstream face.
Log vanes/step jams Single logs or small bundles of Low-cost, minimally Derrick 1997;
logs secured to bed. Also called intrusive. Generally limited ODFW 2010
log bendway weirs (if partially to channels with low banks provides nine
spanning channel and angling not subject to mass failure. configurations.
upstream) or log steps (if fully May be used to retard bed
spanning channel, and usually erosion (fully spanning logs)
placed perpendicular to channel). or divert flow away from
Ends of logs held in place by concave bank (log bendway
burying in sediment or in bank, or weirs). High failure rates due
secured against trees, boulders, or to undermining by
bedrock. downstream scour hole.
Log weirs/valley jams Weir-like accumulations built Creates pool habitat. Prone D’Aoust and Millar
around one or more large logs to failure by flanking or 2000
(key members). undermining.

4Many variations on these basic configurations have been used. Note that “strengths and weaknesses” are subject to the project goals and objectives. In some
settings, erosion control and channel stabilization are desirable while other projects are intended to increase fluvial dynamism.

Large Wood National Manual July 2015


6-28
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Functional Role and


Configuration Sketch Description Strengths and Weaknesses References
Rootwads or Logs buried in bank with Protects low banks by Wood and Jarrett
meander jams rootwads protruding into channel. reducing shear stress acting 2004
Usually placed on outside of on bank toe, provides scour
bends. pools with woody cover
Accumulates fluvially
transported wood. Does not
emulate natural features.
Tree revetments or Whole trees placed along bank Deflects high flows and Cramer et al. 2002
roughness logs or parallel to current. Trees are shear from outer banks; may
bench jams overlapped (shingled) and induce sediment deposition
securely anchored or lodged into and halt erosion. Provides
bedrock outcrops, boulders, or complex cover until smaller
other obstructions. branches decay or break
away.
Toe logs One or two rows of logs or whole Temporary toe protection. Cramer et al. 2002;
trees running parallel to current Generally only for low banks Brooks 2006
and secured to bank toe. Gravel fill because banks above toe
may be placed immediately remain unprotected and
behind logs. therefore allow toe logs to
be flanked if banks are high
and erodible.
Bar apex jam Wood structure composed of 10– Readily accumulates fluvially Abbe and
30 logs placed in the middle of the transported wood. Designed Montgomery 1996;
channel to initiate bar formation to emulate a commonly Cramer 2012;
or placed on the upstream end of occurring natural large Brooks 2006
an existing bar or island. wood formation.

Large Wood National Manual July 2015


6-29
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

spurs oriented in an upstream direction cause


6.6.4 Determine Dimensions greater scour than if oriented normal to the
bank, and spurs oriented in a downstream
The next step is to determine the constructed direction cause less scour. Wood members
dimensions of the structures. Following embedded in the bank so that their butts or
construction, dimensions will increase slightly rootwads are pointing upstream may gain
as additional large wood is trapped by (or stability as drag forces tend to push them into
“racked up on”) the structures, and some the bank.
lowering will occur due to settlement and
decay. However, as-built dimensions are useful For structures that will not overtop at flows less
for estimating required wood quantities and for than bankfull stage, an orientation
hydraulic analyses. perpendicular to flow is generally considered to
be the most effective (e.g., Klingeman et al.
The geometry of intermittent structures (ELJs 1984). Rapidly changing, dynamic sites should
or flow-deflection jams) may be specified by six be evaluated for likely changes in the effect of
parameters: crest angle, crest length, channel migration on orientation angle with
embedment length, crest elevation, structure incident flow.
length, and spacing (Figure 6-12). Guidelines for
selecting these parameters for wood structures, Figure 6-12. Definition Sketch for Large Wood
as for stone river training structures, are Geometric Variables
primarily based on long-running experience
(e.g., Design Guideline 2 in Lagasse et al. 2009; crest angle

Pokrefke 2013). However, a few systematic structure length

investigations have been conducted using flow

physical models in flumes (e.g., Kuhnle et al. A spacing

1999, 2002; Thompson 2005; Svoboda and effective crest length

Russell 2011), which focus on effects of top bank

structure design on local scour. Additionally, a point bar


crest length
A’

few workers have conducted investigations embedment length

using numerical models (e.g., Jia et al. 2009). crest elevation


For high-risk projects, it may be helpful to top bank
construct site-specific numerical or physical section A-A’
models and analyze effects of design
parameters (Table 6-3). Below we encapsulate Structures such as bendway weirs, which are
guidance that may be used to generate trial designed to overtop at relatively low flows, are
dimensional characteristics that should be oriented upstream to set up a hydraulic
further refined and modified through iterative gradient directed away from the bank.
hydraulic analysis or modeling to fully complete Accordingly, the crest angle for structures that
design. are overtopped frequently may be set at
15° upstream from a line drawn perpendicular
6.6.4.1 Crest Angle to flow to promote deflection of overtopping
flow away from eroding banks. However, for
The crest angle is defined as the angle between sites with heavy drift wood loading, it is best to
a line normal to the approach flow vector and assume the structure will end up with an
the weir crest. A crest angle of 90° has the effect upstream face oriented perpendicular to the
of forcing the main flow current and channel flow, or possibly with a downstream
thalweg farther from the concave bank than orientation. As wood accumulates upstream of a
upstream or downstream orientations. Stone structure the pile of racked material tends to

Large Wood National Manual July 2015


6-30
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

taper, giving a structure a blunt arrowhead structures must be high enough so that the
form pointed upstream. sediment berms that form over the structures
stabilize the existing near-vertical banks. Stable
6.6.4.2 Crest Length bank heights and angles may be based on
geotechnical analyses (e.g., the Bank Stability
The crest length for structures that do not span and Toe Erosion Model [BSTEM]) (Simon et al.
the channel may be based on a projected value 2000, 2014, USDA 2013) or empirical criteria
for the equilibrium width of the channel. Crest based on regional data sets.
length will then be the difference between the
existing channel width and the equilibrium 6.6.4.5 Structure Streamwise
width times the cosine of the crest angle.
Alternatively, crest length may be based on a Length
target flow conveyance for the design
Structure length is dependent upon the upper
cross-section. In any event, crest length should
limit length of available logs for simple
be small enough that blockage is less than
structures and often is 1 to 3 times the crest
one-third the channel width (Johnson et al. length. Length may be adjusted to achieve
2001). Flume experiments by Thompson (2002)
specific geomorphic or ecological objectives.
showed that high (overtopped by the 0.27-year
event) deflectors that projected only 25% of the
way across the flume produced more scour
6.6.4.6 Spacing
during high flows than lower ones (overtopped Spacing is set as a preliminary or trial value in
at 9% bankfull discharge) that projected 75% of initial reach layout (see Section 6.6.2, Reach
the way across the channel. Layout), but refined as dimensions of individual
structures are selected. Spacing between
6.6.4.3 Embedment Length intermittent wood structures is measured crest
to crest. Spacing should be great enough to
Embedment length is critical for structural provide segments of unprotected bankline
stability. The approach outlined below under between structures to reduce cost and to create
“Geotechnical Forces” may be used to compute
physical habitat diversity (Shields et al. 1995),
embedment length, but a rule of thumb is to
but prevent flanking and structural failure. See
embed at least two-thirds of the log or structure Table 6-6 above for standard guidelines. Similar
length (Oregon Department of Transportation
guidance is provided by Shields et al. (2004)
2011).
and by Lagasse et al. (2009). A rule of thumb is
that the maximum downstream influence of a
6.6.4.4 Crest Elevation structure will be less than 7 times the effective
crest length when Rc/W <3; spacing should
Abbe et al. (1997) and Castro and Sampson
always be less than this (Drury et al. 1999).
(2001) suggest crest elevation be set equal to
Lagasse et al. (2009:2.12–2.15) provide a
that of the channel-forming flow stage. Still
procedure for locating and spacing structures in
other practitioners suggest that to achieve
a given meander bend. Considerable research
effective flow deflection the general rule of
has been conducted on step-pool channels (e.g.,
thumb is that the height of the structure
Comiti and Mao 2012).
(distance from channel bed to crest) should be
0.5 times the “channel-forming flow” depth
(Klingeman et al. 1984; Drury 1999). All other
factors being equal, local scour depths tend to
be greater for higher structures. In incised
channels crest elevations for ELJ-type

Large Wood National Manual July 2015


6-31
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Species that are decay resistant are preferred,


6.6.5 Select Wood Materials such as eastern red cedar (Juniperus virginiana),
Minimum dimensions, species, and sources for western red cedar (Thuja plicata), coastal
woody materials should be specified during redwood (Sequoia sempervirens), Douglas-fir
design. Cramer et al. (2002) suggest the (Pseudotsuga spp.), or bald cypress (Taxodium
following guidelines for roughness trees. distichum). Rapidly decaying species such as
cottonwood (Populus spp.), pines native to the
Southeast (e.g., Pinus echinata and Pinus taeda),
and alder (Alnus spp.) should be avoided.
Dimension Minimum Size However, as noted above, use of freshly cut or
Rootwad Bankfull discharge depth grubbed willow or cottonwood trees may be
Diameter desirable for quick revegetation in structures
Trunk 0.5 x bankfull discharge depth that are partially buried. Additional information
Diameter about desirability of various species is found in
Tree 0.25 x bankfull discharge width Table 6-2 above.
Length
In some cases, material other than wood may be
more desirable for process restoration. If
Clearly, wood materials this large are not another material can be used that provides the
always available on site (some scientists have same function and better meets stability and
suggested that the near universal mobility of longevity requirements at a lower cost, it is
large wood in present-day rivers [e.g., Curran certainly worthwhile, especially if it helps to
2010] is due to the scarcity or absence of retain natural wood moving through the
extremely large trees that were present prior to system. For example, large concrete jacks
European settlement). Although using onsite (dolosse), together with large quantities of
wood is preferred given the extreme cost of native wood, have offered an economical and
bringing in large materials, importation may be long-term alternative to simulate the function of
necessary to obtain large enough logs and avoid logjams in restoring habitat and protecting
detrimental environmental impacts. Benefits to roads in the Pacific Northwest (Abbe and
the habitat and stream ecosystem must be Brooks 2011). Synthetic large wood material for
weighed against the impacts of clearing and stream work is available commercially (Bolton
grubbing on existing riparian and floodplain et al. 1998). Synthetic wood products are
habitat. engineered to compare favorably with natural
materials in terms of durability or habitat value.
Complex woody material structures that feature
However, they may be less effective in terms of
numerous branches and thus high stem density
habitat creation or more costly than natural
locally depress velocity, inducing sediment
materials. Cost comparisons should consider
deposition. Accordingly, materials should be
full project life cycles.
selected that have numerous branches, and
breakage and removal of branches should be
avoided during construction. Clearing within 6.6.6 Hydraulic Analysis
the stream corridor should be avoided, but bar
scalping to provide temporary relief of outer Well-designed large wood projects have
bank erosion in a sharp bend may be advisable reportedly withstood flows in channels with
in certain cases, and resulting woody materials average bed shear stresses of 50–170 Pa and
(e.g., willow rootwads and stems) may be used estimated velocities of 3 to 4 m s-1 (Allen and
in structures to trigger rapid revegetation. Leech 1997, Plate 13, and Table 1 in Abbe and
Brooks 2011). However, design of large wood
for a specific site should not be based on the

Large Wood National Manual July 2015


6-32
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

average shear stress or velocity but on a dimensions). Examples are provided by Abbe in
detailed force balance because turbulent flows Brooks (2006), He et al. (2009), and Smith et al.
in the immediate vicinity of the boundary are (2011). Calibration and validation of hydraulic
quite complex and loading on a given structure analyses and hydrodynamic models for projects
is poorly represented by cross-sectional mean not yet constructed are problematic. In years to
velocities or stresses. come, numerical modeling capabilities should
allow detailed models of water and sediment
Except for extremely simple projects, a
movement in reaches with a range of large
hydraulic analysis is strongly recommended.
wood structure sizes and frequencies.
Such an analysis should include assessment of
the flow conveyance, sediment transport
capacity, and velocity and shear stress at design 6.6.7 Scour Analysis
discharge for the existing channel and for the
Channel bed scour or degradation is often a
channel after large wood structure construction
primary causal factor in large wood structural
(Cramer 2012: Appendix E). Rough analysis
failures (Shields et al. 2004, 2006; Herrera
may be based on pencil and paper or worksheet
Environmental Consultants 2006). Scour pools
computations using uniform flow formulas
provide important aquatic habitat, but scour
(Gippel et al. 1996) and simple sediment
that undercuts an instream structure can pose a
transport relations, but a collection of cross-
significant threat to the structural integrity.
section and thalweg profile surveys allows 1D
Undercutting occurs when the depth of bed
modeling with tools such as the Hydrologic
scour exceeds the depth of the structure. Scour
Engineering Centers River Analysis System
estimates are needed to design the portion of
(HEC-RAS). Large wood structures may be
the structure that will be placed below bed
simulated by modifying cross sections, adding
level. A first order approximation of scour
blocked obstructions, or increasing roughness
depths in gravel bed channels may be obtained
(Manning) coefficients (Valverde 2013). HEC-
using a regression line fitted to the thalweg
RAS has limited sediment transport capability
profile. The difference between the maximum
and can simulate unsteady flows but will not
positive residual and the maximum negative
simulate dynamic boundaries (i.e., bed or bank
residual (riffle/pool amplitude) provides an
scour) during a given hydrologic event.
estimate of the scour potential within the reach
Therefore, HEC-RAS will likely over-predict
(Brooks 2006).
peak flood stages, particularly for channels with
sand or fine-gravel beds that readily scour More detailed analyses include estimates of
during the rising limb of hydrographs (Brooks different types of scour. Total scour estimates
2006). Flow depths and velocities may be used are the sum of general scour, contraction scour,
as input to scour analysis, as described below. and local scour. Local scour for large wood
structures placed on or beside banks may be
Higher risk projects may call for two-
estimated using approaches used for bridge
dimensional simulations, which are far superior
abutments, while local scour at mid-channel
to 1D models in examining large wood effects.
structures (e.g., bar apex jams) may be
These tools allow more detailed analysis of the
estimated using bridge pier scour equations.
local impacts of structures on flow stages,
Local scour associated with large wood may be
velocities, shear stresses, bed scour, and habitat
quite dynamic in sand-bed channels with
characteristics as well as reach-scale effects.
considerable scour and fill occurring during
Such efforts are more resource intensive, but
flow events (Borg et al. 2007).
these models do allow some estimation of the
morphologic response to a given large wood Detailed guidance for scour analysis is not
design (placement of structures and their provided here; designers must consult the

Large Wood National Manual July 2015


6-33
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

references below. A range of empirical formulas a model implemented in Microsoft Excel to


is available to generate scour estimates, as estimate bank stability, failure modes, methods,
described by Cramer (2012: Appendix E), and distances. BSTEM can be used to test the
Arneson et al. (2012), and Shields (2007); and effects of hydraulic scour, water table height,
examples of scour analyses applied to large vegetation, and stage on stability; used
wood structure projects are provided by Brooks iteratively with knowledge of the flow regime to
(2006) and Abbe and Brooks (2011). Drury predict widening rates; and used to test various
(1999) provides example scour computations mitigation strategies to control undercutting
for engineered log jams placed in the North and mass failure (Simon et al. 2014). BSTEM
Fork Stillaguamish River, Washington. requires significant amounts of site-specific
input data.
Scour depths are sensitive to structure
dimensions. Structures may become smaller 6.6.9 Force and Moment
due to decay or loss of members, but they may
become much larger if they trap and retain Analysis
floating wood. If significant amounts of large
wood are being transported into the project Forces that should be considered for large wood
reach and wood trapping is likely, approaches design include net buoyancy, friction between
described by Lagasse et al. (2010) or by Elliot the wood structure and the bed, fluid drag and
et al. (2012) for estimating the amount of large lift, and geotechnical forces on buried members.
wood trapped on bridge piers and the For very large, critical projects with asymmetric
associated scour, modified by the user for large structures, a moment analysis of each large
wood structures, may be useful. Knutson and wood structure is recommended, and a separate
Fealko (2014) offer guidance for estimating factor of safety with respect to moments should
wood trapping on large wood structures in the be computed as shown by Shields and Wood
Pacific Northwest. (2007) and Knutson and Fealko (2014). Below
is an example of a simpler approach similar to
the one developed by Drury (1999) that
6.6.8 Bank Erosion considers vertical and horizontal forces
Bank erosion analyses are less straightforward separately. The large wood structure is treated
than bed scour analyses and subject to more as a unit; in other words, it is assumed that
site-specific factors. Large wood placement may large wood members (logs) are secured to one
deflect flows toward banks and locally another by hardware (chain or cable5) or by
accelerate bank erosion; this may not be interlocking construction and ballast. A free
undesirable as it may scour zones that provide body diagram is useful in ensuring inclusion of
pool habitat or cover at base flow. Approaches all forces in the analysis (Figure 6-13). If the
for assessing bank erosion potential include a structure cannot be assumed to act as an
general assessment of erosion rates through the integrated unit, a more complex force and
project reach using historical aerial photos, moment analysis for each structural component
surveys, landowner interviews, and geomorphic is required.
assessments. Assessments specific to structure
locations may be based on professional
judgment or computation of peak shear stresses
on banks using numerical (2D) models and
5 As used herein, the term “cable” refers to cable
comparing those values with critical values for
the sediments on the boundary. Additional larger than about 10 millimeters in diameter. Such
cable is often called “wire rope” because it consists of
analysis may be performed using BSTEM
several strands of metal wire laid (or “twisted”) into
(Simon et al. 2000, 2014; USDA 2013), which is a helix. Steel is the main material used for wire ropes.

Large Wood National Manual July 2015


6-34
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

6.6.9.1 Vertical Forces Equation 6-3:

Buoyant Force and Gravity


Vertical forces often result in lower factors of
where  represents specific weight, and V
safety than horizontal forces and thus control
represents volume with subscripts d and w
design (Drury 1999; Shields et al. 2004). The
referring to wood and water, respectively. It is
buoyant force is equal to the weight of the
important to note that Vw represents displaced
displaced water volume. The net buoyant
water volume. If the design water surface
force, 𝐹⃗𝑏 , is the only vertical component of the
elevation is high enough to inundate the
driving force and is equal to the difference
structure, then Vw = Vd. If the structure is not
between the weight of the structure and the
fully submerged, Vw is only the displaced water
weight of displaced water:
volume, and is equal to the volume of
submerged wood, not the total wood volume for
the entire structure.

Figure 6-13. Typical Free Body Diagram for a Large Wood Structure

Fgh

Flow Flow
FL Fb Ff Fd
Ff Fd

Wbl Fah
Fav Fgv

Profile Plan

Forces may be determined as follows. Fav = restraining force due to anchors or other restraints in vertical direction, Wbl =
weight of ballast, Fgv = geotechnical forces in vertical direction, Ff = force of friction between LW and stream boundary,
Fd = drag force, FL = lift force, Fb = buoyant force, Fgh = geotechnical force in horizontal direction, Fah = force due to anchors
or other restraints in horizontal direction. Points of application for force vectors shown are arbitrary.

Wood structures may have complex geometries, treating rootwads and boles as separate
which makes determination of volume difficult, elements (Braudrick and Grant 2000; Shields et
particularly for partially submerged structures. al. 2004; Abbe and Brooks 2011). For example,
Shields et al. (2004: Appendix) provide an for a structure with n logs, we may approximate
example computation of Vwater as a function of the boles as cylinders and the rootwads as
flow depth. Computations may be simplified by cones:
assuming that logs are cylinders or cones,
adopting advantageous coordinate systems, and

Large Wood National Manual July 2015


6-35
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Equation 6-4: forces on simple logs with few branches showed


lift can be quite large in cases where significant
flow occurs above and below trunks and
branches (Shields and Alonso 2012). However,
lift coefficients ranged from 0.8 to 1.7 for large
where lk and rk represent the length (exclusive wood with less interstitial space (a rootwad)
of rootwad) and DBH radius of the kth log, that was suspended above the channel bed. In
respectively, and tk and wk represent the the absence of better information, CL may be
thickness (measured in direction parallel to assumed = 1.0 for complex large wood
trunk) and radius of the kth rootwad. Brooks structures that are submerged. Lift may be
(2006) suggests that tk and wk may be assumed = 0 for large wood in contact with the
approximated by 4rk and 2.5rk, respectively. bed that is not fully submerged. For purposes of
computing A, a large wood structure may be
The specific weight of wood, d, should be treated as a single body, rather than as
assumed to represent worst-case or driest and individual cylinders if the upstream face of the
partially decayed conditions. Unless more structure is only slightly porous due to ballast,
specific data are available (Shields 2004; racked debris, or trash (Gippel et al. 1996). For
Brooks 2006; Shields and Wood 2007; Ruiz- structures located on the outside of bends, the
Villanueva et al. 2014; Miles and Smith 2009; approach flow velocity may be assumed equal
Forest Products Laboratory 2010), a value of to 1.5 times the cross-sectional mean velocity
3,900 N/m3 (corresponding to a specific gravity (U.S. Army Corps of Engineers 1994).
of 0.40) should be used. If rootwads are
approximated by a cone (as in the above Ballast
equation), the resulting volume should be If ballast is provided by gravel or cobble fill, the
multiplied by an appropriate void ratio to allow maximum ballast volume is the volume
for the empty spaces between roots after soil is computed using external structure dimensions
removed. less the volume of wood. If the structure is
Lift approximated by a rectangular prism, then the
weight of the ballast, Wbl, is computed using the
Fluid lift is generated by flow acceleration following equation.
above and below a solid object. Drag on a large
Equation 6-6:
wood structure may be computed using the
following equation.
Equation 6-5:
where bl = specific weight of ballast (typically
14,000 N/m3 < bl < 19,000 N/m3 for gravel or
cobble), and Ws, Hs, and Ls are the width, height,
and length of the large wood structure,
where 𝐹⃗𝐿 = lift force, CL = drag coefficient, A = respectively. If boulders are used for ballast,
area of structure projected in the plane their volume may be approximated by that of an
perpendicular to flow, and Uo = approach flow equal number of spheres, and
velocity in the absence of the structure. Many
designers neglect lift when analyzing forces on Equation 6-7:
large wood structures (Shields 2004; Abbe and
𝜋𝑑𝑏6
Brooks 2011), as lift forces are usually small ⃗⃗⃗⃗𝑏𝑙 = 𝑛𝛾𝑏𝑙 (
𝑊 )
2
relative to buoyant forces (Merten et al. 2010;
Knutson and Fealko 2014). Measurements of lift

Large Wood National Manual July 2015


6-36
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

where n is the number of boulders, bl is the the submerged ballast and the weight of the
displaced water:
specific weight of the boulder (not bulk weight),
usually 25,000–27,000 N/m3, and db is the Equation 6-9:
diameter of a sphere with volume equal to that
of a representative boulder. Alternatively, a
more conservative approach would be to apply
Equation 6-6 above using the bulk density of and

boulders for bl. Equation 6-10:

Conservative designers may wish to reduce 𝑊 ⃗⃗⃗⃗𝑏𝑙


in subsequent computations to allow for loss of
ballast due to erosion at structure margins. If Drag
velocities within the structure exceed critical
Fluid drag is a driving force in the horizontal
velocities for erosion of the ballast particles,
direction. Drag on a large wood structure may
rapid loss of ballast may occur, destroying the
be computed using the equation
structure (Shields et al. 2004).
Equation 6-11:
6.6.9.2 Horizontal Forces
The free body diagram (Figure 6-13) is also
useful in considering and analyzing forces in the where 𝐹⃗𝑑 = drag force, CD = drag coefficient, A =
horizontal plane. area of structure projected in the plane
Friction perpendicular to flow, and Uo = approach flow
velocity in the absence of the structure. Drag
The movement of large wood structures by coefficients vary greatly from one large wood
sliding along the bed will be resisted by a formation to another due to differences in the
frictional force 𝐹⃗𝑓 with magnitude equal to the way the members engage the flow. However,
normal force times the coefficient of friction CD tends to decline to values typical of cylinders
between the woody material and the bed. If 𝐹⃗𝑛 > (0.5–1.0) when large wood becomes so complex
0, that interstitial flow is nil. CD values reach their
maximum (~1.5) when large wood structures
Equation 6-8: are just barely overtopped due to the additional
 drag incurred due to formation of standing
F f  bed Fn waves (Shields and Alonso 2012). In the
In the absence of measured data, it may be absence of better information, CD may be
assumed that bed = tan , where is the friction assumed = 0.9 for fully submerged conditions
angle for the bed sediments (Braudrick and and = 1.5 for conditions where the water
Grant 2000; D’Aoust and Millar 2000; Castro surface is within one (typical) log diameter of
and Sampson 2001). If vertical restraint is the top of the structure (Shields and Gippel
provided by ballast, the normal force will be 1995; Brooks 2006; Shields and Alonso 2012).
equal to the weight of the wood and ballast Drag forces rapidly diminish with time during
above the waterline plus the submerged weight the first few high flow events as patterns of
of the wood and ballast below the waterline. If scour and deposition reshape the local
the structure is fully submerged, then the topography (Wallerstein et al. 2001). As for lift
submerged weight of the ballast, Wbl(sub), will be computations, when computing A, a large wood
equal to the difference between the weight of structure may be treated as a single body,
rather than as individual cylinders if the

Large Wood National Manual July 2015


6-37
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

upstream face of the structure is only slightly Figure 6-14. Entanglement of Logs in Riparian
porous due to ballast, racked debris, or trash Stumps and Boles for Passive Restraint,
(Gippel et al. 1996). For structures located on Hylebos Creek, Milton, Washington
the outside of bends, the approach flow velocity
may be assumed equal to 1.5 times the cross-
sectional mean velocity (U.S. Army Corps of
Engineers 1994).

6.6.9.3 Determining How


Structural Elements Will Be
Restrained
Large wood added to a stream may be
unrestrained, ballasted, or anchored. Many
natural large wood accumulations are stable for
decades or centuries, particularly when the (Photo by Mike Hrachovec)
length of the large wood is large relative to
Active restraining approaches include placing
channel width. Natural large wood stability is
ballast (soil, cobbles, boulders) on or within the
improved by the presence of rootwads, shallow
structure; embedding part or all of the large
flow depth, partial burial, high length relative to
wood in the bank or in a stone structure such as
channel width, and bracing against trees, rocks,
a revetment or spur dike; and using cable, rope,
banks, etc. (Merten et al. 2010). In some cases
or chain to secure the structure to boulders,
(such as undeveloped watersheds), transport of
mechanical anchors placed in soil, rock or
wood downstream may not be objectionable.
concrete, stumps, trees, deadmen, or pilings
However, in most cases, large wood structures
(Fischenich and Morrow 2000; Shields and
are intended to remain in place over their
Wood 2007; Knutson and Fealko 2014).
design life in order to generate the intended
benefits and to avoid hazards to downstream
bridges and other infrastructure, so large wood
installation design must include passive or CROSS-REFERENCE
active restraint.
Additional information on options for large wood
Passive anchoring restraint refers to a design restraints is provided in Table 8-3 of Chapter 8,
approach in which the shape, weight, ballast Regulatory Compliance, Public Involvement, and
and placement of large wood structures are Implementation.
adequate to resist movement in events up to the
design flow. Passive anchoring along smaller
streams includes entanglement of logs within A wide range of mechanical anchors for soil and
boles of trees adjacent to the stream (Figure 6- rock are available; vendors can supply
14). Logs within a passively anchored structure information about each type. Detailed guidance
may be attached to one another, but not to for large wood structure active restraint
external anchors. Passive anchoring is not suitable for smaller structures is provided by
recommended for high hazard situations, for the Natural Resources Conservation Service
sites with vulnerable infrastructure (2007b) and by Cramer (2012: Appendix G).
downstream, or for sites where structures will
be frequently overtopped. When boulders are used for ballast, the
buoyant, drag, and lift forces on the ballast rock

Large Wood National Manual July 2015


6-38
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

must be considered in the force balance Equation 6-13:


(D’Aoust and Millar 2000). Logs in complex
structures may be attached to one another or to
boulders by drilling holes through them and
pinning them together with rebar. Alternatively,
logs may be fastened with chains, which are less where Fsv and Fsh are safety factors with respect
likely to fail from repeated bending stress. to vertical (floating) and horizontal (sliding)
Epoxy adhesive has been used for attaching movement, 𝐹⃗𝑔𝑣 and 𝐹⃗𝑔ℎ are vertical and
chains, cables, or metal rods to holes drilled in horizontal restraint forces, respectively,
boulders or bedrock. More details about the provided by geotechnical processes (embedded
methods to epoxy cable into rocks can be found
logs or piles), and 𝐹⃗𝑎𝑣 and 𝐹⃗𝑎ℎ are vertical and
in published documents such as the California
horizontal restraint forces, respectively,
Salmonid Stream Habitat Restoration Manual
provided by anchors. 𝑊 ⃗⃗⃗⃗𝑏𝑙 is the vertical force
(Flosi et al. 1998). Cable (wire rope) should be
due to ballast as defined by Equation 6-6.
galvanized or stainless steel and sized to
withstand loads greater than those computed.
Hardware used to affix cable to itself or to other 6.6.9.5 Geotechnical Forces
components should be carefully specified. See
Horizontal Large Wood Embedded in
Cramer (2012: Appendix G) for
recommendations. Bank
Members of a large wood structure may provide
6.6.9.4 Safety Factors for Designing significant restraining forces if they are
Restraints embedded in banks by excavating trenches and
burying them (“keying in”). However, because
Safety factors are ratios of resisting to driving of the disturbance required, this approach may
forces. Restraint systems should be designed to not be practical for extremely high, steep banks
achieve safety factors that are scaled to the risk or banks providing sensitive habitats. The
profile of each large wood placement (see embedment length or bank key-in distance for
Chapter 7, Risk Considerations). Resisting forces structures that are partially buried in the bank
include the weight of the structure plus ballast, should vary with bank height, soil type, and
friction with the bed, and forces due to anchors. stream size. As a rule of thumb, a log will be
Driving forces include buoyancy and fluid lift stable if two-thirds of the log is buried in the
and drag. Additional driving forces may arise bank (Oregon Department of Transportation
due to waves, ice action, or collisions from 2011). The key-in should be sufficient to
floating debris,6 but computation of these forces maintain the position of the rest of the structure
is beyond the scope of this guide. Put simply, throughout its design life, and should be greater
restraining systems should be designed to meet: for frequently overtopped and highly erodible
Equation 6-12: banks (Sylte and Fischenich 2000). Site-specific
computations are suggested (D’Aoust and Millar
2000). A simplified analysis is presented below;
a more detailed treatment that includes sloping
banks and a nonhorizontal water table is
presented by Wood and Jarrett (2004). Because
and
these geotechnical analyses involve
considerable professional judgment, an
6 Knutson and Fealko (2014) present suggestions on experienced geotechnical engineer should
consideration of ice and debris loading in large wood perform or review this part of the design.
structure design.

Large Wood National Manual July 2015


6-39
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

RISKS

Safety Factors

Safety factors are ratios of resisting to driving forces. Engineers often compensate for uncertainty in design
computations by modifying designs in order to increase the safety factor. Safety factors recommended for design
of concrete gravity structures like dams, pumping stations, and floodwalls range from 1.1 to 3.0 based on the
anticipated loading and the quality of site information (U.S. Army Corps of Engineers 2005). Factors for bearing
capacity of soils range from 2 to 4 (U.S. Army Corps of Engineers 1992). Knutson and Fealko (2014) recommend
different safety factors for large wood structures based on risk profile and failure mode as shown below.

Public Safety Property Stability Design FOSrotation


Risk Damage Risk Flow Criteria FOSsliding FOSbouyancy FOSoverturning
High High 100-year 1.75 2.0 1.75
High Moderate 50-year 1.5 1.75 1.5
High Low 25-year 1.5 1.75 1.5
Low High 100-year 1.75 2.0 1.75
Low Moderate 25-year 1.5 1.75 1.5
Low Low 10=year 1.25 1.5 1.25

Designers often deal with uncertainty by making conservative assumptions when computing the components that
constitute the right-hand sides of Equations 6-12 and 6-13. These assumptions can result in “implicit” factors of
safety that should be considered when assessing the overall factor of safety for the design. If the large wood
placement is intended to survive several years, increased factors of safety may be needed to allow for the
possibility that:

1. Wood is partially decayed and thus less dense,

2. The current angle of attack has shifted, with flow impinging directly on the structure,

3. Branches and twigs have been removed, simplifying the wood and increasing drag coefficients, or

4. Ballast has been eroded or moved.

For example, if a designer decides to use 300 N/m3 for the specific weight of wood rather than a more realistic
⃗⃗𝑳 is
value of 500 N/m3 in order to be conservative, there is an implicit 𝑭𝒔𝒗 = 1.66 (assuming the lift force 𝑭
negligible). In other cases, elevated safety factors are used to compensate for uncertainties in the underlying data
and assumptions used in the design analysis.

Embedment depths must be increased if bank bank is horizontal; the bank is composed of
erosion is likely so that depths are adequate homogeneous, isotropic soil with bulk specific
even after erosion has occurred. weight s, effective friction angle 'and
The resistive forces due to passive soil pressure effective cohesion c’; and the groundwater table
acting on buried portions of logs are direct elevation in the bank is approximately equal to
reactions to fluid forces. The following the stream surface elevation, which is high
equations assume that the log is embedded enough to fully submerge the log (Figure 6-15),
horizontally in the streambank; the top of the and is a distance Dw below top bank elevation. In

Large Wood National Manual July 2015


6-40
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

addition, the bank slope is assumed to be near Equation 6-15:


vertical, and the log is assumed to be
frictionless. The log has a length L, a diameter d,
and is buried a distance D below the top bank
and a horizontal depth Lem (embedment length).
where Bd is the width of the ditch and Cw is a
The soil passive resistance distribution is
coefficient that captures the interaction
assumed to be triangular with its maximum
between the ditch walls and the fill. Cw is given
value at the bank face and decreasing linearly to
by
zero at the embedded tip of the log. This implies
that the resultant passive resistance force acts Equation 6-16:
on the log a distance of two-thirds Lem from the
embedded tip. The active earth pressure force is Cw 
1  e 0.38D / Bd

assumed to be small relative to the passive 0.38
force. 𝐷
for < 2 and
𝐵𝑑
Figure 6-15. Definition Sketch for Derivation of
Geotechnical Forces on a Horizontally 𝐷
𝐶𝑤 =
Embedded Log 𝐵𝑑
𝐷
for ≥ 2. Here, e is the base of natural logs.
𝐵𝑑
The two approaches for computing Fsoil
converge for ditches with widths just slightly
greater than the log diameter. Clearly, greater
restraint can be provided at shallower depth
with denser soil or by using gravel or rock
instead of soil due to higher γsoil in Equation
6-14.

Assuming friction between the soil and log is


negligible, the passive soil pressure force, 𝐹⃗𝑝 , is
Source: Shields and Wood (2007) given by
Equation 6-17:
The vertical loading on the log due to the weight
of the soil above it will be given by
𝐹⃗𝑝 = 0.5𝜎𝑝 𝐿𝑒𝑚 𝑑
Equation 6-14:
where p, the passive soil pressure, is given by
Equation 6-18:

𝜎𝑝 = 𝜎𝑣′ 𝐾𝑝 + 2𝑐′√𝐾𝑝
where 𝛾𝑠 is the bulk (or moist) unit weight of
the soil above the log. Alternatively, 𝐹⃗𝑠𝑜𝑖𝑙 may be where Kp, the Rankine coefficient of passive
computed using equations developed to earth pressure, is given by
compute soil loading on conduits buried in
ditches. When the ditch width is no greater than
three times the log diameter

Large Wood National Manual July 2015


6-41
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Equation 6-19: Therefore, scour analysis must precede


geotechnical analysis for vertical piles.
∅′
𝑲𝒑 = 𝒕𝒂𝒏𝟐 (𝟒𝟓 + )
𝟐 Applicable methods selected under advisement
from a geotechnical engineer should be used to
where ∅′ is the effective angle of internal determine the number of pilings, piling depth.
friction and c’ is the effective soil cohesion. If and other properties needed. Knutson and
unknown, effective soil cohesion may Fealko (2014) suggest use of Brom’s equation7
conservatively be assumed to equal 0. for computing lateral resisting forces due to a
group of vertical piles.
6.6.9.6 Moments, Horizontal Large
Equation 6-22:
Wood Embedded in Bank
1 3
𝐿𝑒𝑚 𝑑𝑝 𝐾𝑝 (𝛾𝑠 − 𝛾𝑤 )
The driving moment about the buried tip of the 𝐹𝑔ℎ(𝑝𝑖𝑙𝑒𝑠) = 𝑁 (2 )
embedded log will be given by the vector sum (ℎ𝑙𝑜𝑎𝑑 + 𝐿𝑒𝑚 )

Equation 6-20: where N is the number of piles, Lem is the length


⃗⃗⃗𝑑 = [(𝐹⃗𝑑 + 𝐹⃗𝐿 )(𝐿𝑒𝑚 + 𝐿𝑒𝑥 /2) + 𝐹⃗𝑏 (𝐿/2)] × 𝑙⃗ of the pile buried below the bed (allowing for
𝑀
scour), dp is the pile diameter, hload is the
where 𝑙⃗ is the unit vector along the axis of the distance above the scoured bed that the load is
buried log and positive in the direction away applied, and all other variables are as
⃗⃗⃗𝑟 previously defined. A trial and error approach
from the buried tip. The resisting moment 𝑀
may be used to adjust piling parameters to
will act opposite the driving moment and will
be given by the vector sum obtain the value of 𝐹⃗𝑔ℎ that will produce 𝐹⃗𝑠ℎ ≥ 2
in Equation 6-12 above. It is important to look
Equation 6-21: at the ultimate pile strength (shear and
moment) versus the applied loads to ensure
that the pile material is structurally sound and
⃗𝑴
⃗⃗⃗𝒓 = [(𝑭
⃗⃗𝒔𝒐𝒊𝒍 (𝟏/𝟐)𝑳𝒆𝒎 + ⃗𝑭⃗𝒑 (𝟐/𝟑)(𝑳𝒆𝒎 + ⃗𝑭⃗𝒂𝒗 𝑳𝒄 ]
will not snap or shear off during the design
× 𝒍⃗ event. Piling strength is often the limiting factor,
and consideration of this factor can result in a
where 𝐹⃗ av is the restraining force due to higher number of required piles.
anchors (if there are any) as defined above, and
Lc is the appropriate moment arm about the
buried tip of the embedded log. 6.6.10 Planting Vegetation
Vertical Large Wood Placed as Piling Because instream wood is tightly linked to the
riparian forest (see Chapter 4, Geomorphology
Logs or metal beams may be driven vertically and Hydrology Considerations), ecosystem
through the large wood structure as it is built so recovery is often aided by planting woody
that they act as pilings to resist horizontal vegetation concurrent with wood addition
driving forces. If a log with a rootwad is buried projects (sometimes referred to as
vertically with the rootwad at the bottom, the
pullout resistance will be increased several-fold 7 Brom’s equation assumes a maximum allowable
relative to a log without a rootwad (Abbe and deflection of the pile at the ground of 0.002 to 0.006
Brooks 2011). Pilings must be placed deeply radians. Assuming rotation about the buried pile tip,
enough to provide adequate safety factors even this assumption implies that a pile embedded 3
meters (10 feet) below the channel bed would have a
after burial depth has been reduced by scour.
theoretical maximum displacement of 1.8
centimeters (0.7 inches) at the channel bed.

Large Wood National Manual July 2015


6-42
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

“revegetation”). Plantings are usually


flood-tolerant, pioneering species that may be
6.6.11 Constructability
propagated by large (~1- to 10-centimeter [0.4- Assessment
to 4-inch] diameter) cuttings. In some cases,
small living trees may be transplanted by Preliminary designs should be subjected to a
excavating root balls and incorporating the constructability assessment (also see Chapter 8
living material and associated soil in the placed for guidance on constructability). Key issues for
wood or adjacent to it. Cuttings may be placed such an assessment include the compatibility of
in banks adjacent to wood placements, on the design with site geology, hydrology,
nearby bars, or planted in sediments that hydraulics, and biota. For example, stream
deposit on or within large structures after the channel substrate must allow for pile driving or
first few high flows. A wide variety of planting excavation if required for construction of
techniques and practices are available; full restraint systems. Posts may be used for
treatment is beyond the scope of this manual. restraints if site substrate allows excavation of
Additional guidance is provided in Chapter 8, post holes, but bedrock is an obvious limitation.
Regulatory Compliance, Public Involvement, and
In deeply incised channels, consideration must
Implementation, and by the Federal Interagency
be given to the challenge of access to the
Stream Restoration Working Group (FISRWG
channel bed from high banks. Certain types and
1998), NRCS (2007f), and Fischenich (2001).
sizes of equipment may be able to reach the
channel from top banks, but it will often be
CAVEAT necessary for equipment to operate within the
channel during low flow periods. Temporary
ramps may be needed to allow movement of
Key Engineering Considerations for Plantings
(NRCS 2007f) equipment and materials from the top
bank/floodplain down into the incised channel
 Protection of plantings from erosion during (Figure 6-16).
period of establishment.
Figure 6-16. Construction of Temporary Ramp
 Ability of established plant materials to for Access to Channel for Large Wood
withstand hydraulic loadings imposed by high
flows.
Structure Construction in Little Topashaw
Creek, Mississippi
 Protection of plantings on banks from
undercutting.

 Effects of slope instability on plantings on high,


steep banks.

 Effects of mature vegetation on reach hydraulics


and sediment transport.

 Site conditions (soil fertility; shade; frequency,


timing, and duration of inundation).

 Potential damage to plants by beaver, muskrat,


deer, or other herbivores.

 Problems due to invasive exotic species. One of the key decisions that affects both design
and construction is whether to construct in the
 Availability of plant materials from local sources. dry using some type of berm or coffer dam to

Large Wood National Manual July 2015


6-43
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

allow site dewatering or with ordinary water rapidly delivers water to streams, peak flows
levels, otherwise referred to as “in the wet.” are elevated in urban areas. Designers are
challenged to produce structures with the
Project economics are heavily influenced by
capacity to withstand high flows while still
haul distances for wood, ballast rock, and other
providing habitat enhancement during lower
materials, so sources and haul routes must be
flows. Design is further complicated by the
carefully planned, and the types of equipment
tendency of urban channels to be disconnected
to be used must be selected and specified.
from adjacent floodplains due to floodplain fill
Limitations on heavy equipment use are or channel entrenchment. In such channels it is
provided by site hydrology and hydraulics, typical for the water width-to-depth ratio to
environmental restrictions (e.g., water quality decrease with increasing flows, resulting in
considerations, migratory or other seasonal increasing velocities and shear stresses as the
windows), and potential conflicts with flows increase up to and beyond Q100. Hydraulic
recreation and navigation. modeling is particularly critical to quantify the
forces acting on the large wood structures and
the stream bed and bank materials. As
6.7 Special described above, assessing the response of
added large wood and channel boundaries to
Considerations for hydraulic loading is a key task within the design
Urban Streams process.

There are special considerations for the If the spatial limits of the project can be
evaluation and design of large wood structures expanded, it may be possible to maintain
in an urban setting due to the constrained instream flow conveyance even with large wood
nature of many of such sites; the extreme added or to restore floodplain function beside
modifications to water, sediment, and wood or within the incised channel to regain lost
loading; and potential impacts on public floodwater storage. Floodplain restoration can
infrastructure and safety. Crossing structures produce a more natural aesthetic through
(bridges, culverts, and pipelines) are more floodplain vegetation and use of softer
prominent in urban settings, and while newly bioengineering techniques such as soil wraps,
constructed crossings should be designed with coir logs and vegetated flood swales.
consideration of passage of large wood Where floodplain restoration cannot be
(Lassettre and Kondolf 2012), existing crossings achieved and forces remain high within the
are likely to retain large wood, and risks should project reach, more non-deformable bank
be carefully assessed. Figure 6-17 illustrates stabilization may be required around large
built projects in urban settings. The following wood. This may include boulder placement,
are some key parameters to consider in the nondegradable geotextiles for bioengineering,
design of large wood structures in the urban and/or cobble backfill within the wood
environment. structures to resist erosion of the native soils
where burial has been used as the primary
anchoring technique. Structure heights may be
6.7.12 Design Discharge up to or above the level of Q100 where bank
It is common for Q100 to be the design flow in stabilization is being installed at high-risk,
urban settings due to the need for stability constrained sites.
within tightly constrained, high-risk project
settings. Due to the prevalence of impervious
surfaces (40% or more), the prevalence of turf
grass, and storm drainage infrastructure that

Large Wood National Manual July 2015


6-44
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Figure 6-17. Examples of Large Wood Projects in Urban Settings of the Pacific Northwest
Before After Description
Reestablishing pool-riffle
morphology using hand-
placed logs and gravel
addition in entrenched
channel.

Bed and bank stabilization


and instream habitat
enhancement with large
wood installed as part of
fish passage culvert
replacement with adjacent
houses and utilities in
entrenched channel.

Large wood installation as


part of channel
realignment and
floodplain/wetland
restoration for flood relief
in former agricultural site.

Installation of large wood


in plane-bed gravel
channel to restore pool
habitat through
constriction scour.

Installation of large wood


to raise bed elevation to
reengage floodplain,
reestablish channel
Before photo not available. sinuosity and provide
instream habitat.

Large Wood National Manual 6-45 July 2015


Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Before After Description


Installation of whole tree
spanning ¼ of river
channel to promote
deposition of sediment
Before photo not available. along channel edge and
increase center-channel
scour

(Photos by Mike Hrachovec)

Though many cities are implementing Exceptions to this policy include an exception
stormwater management techniques intended issued by Region X of FEMA that allows the “no-
to reduce both the flood peaks and the amount rise analysis” to be replaced by the judgment of
of fine sediments delivered to urban channels, a qualified professional such as staff of the
such as Low Impact Development (LID) Rural Conservation and Development or the
techniques, it will be many years before these NRCS. “The qualified professional should, at a
projects are prevalent enough within most minimum, provide a feasibility analysis and
watersheds to result in measureable reductions certification that the project was designed to
in flood peaks in urban channels. keep any rise in 100-year flood levels as close to
zero as practically possible and that no
6.7.13 Floodplain Regulation structures would be impacted by a potential
rise.” Additional provisions of the policy include
Many urban settings are under constraints maintenance considerations and further
imposed by FEMA floodplain regulations, which analysis to address river dynamics (FEMA
require a “no-rise” condition within the channel. 2009).
In these cases the project must be designed so
that hydraulic analysis of existing and proposed 6.7.14 Existing Utilities
conditions demonstrate that the proposed large
wood installation does not increase flood stages It is critical to fully understand the constraints
within the project reach or upstream. In these placed on the project by the presence of existing
situations excavation of channel or floodplain utilities (power, sewer, gas, storm drainage,
cross-section may be required to provide flood water) by obtaining the horizontal and vertical
storage equal to or greater than that portion of location of all utilities at the start of design. In
the cross-section occupied by the large wood, many cases these utilities will provide a limit of
incorporating into the proposed model excavation that will influence where large wood
conditions the additional channel roughness can be installed as well as appropriate
imposed by the large wood. These constraints, anchoring methods. In some cases, anchoring
when imposed, will often drive the with bole burial may not be feasible due to
development of project alternatives and the location of utilities within the bank. The
final design. An LOMR or Conditional Letter of relocation of utilities to allow installation of
Map Revision (CLOMR) prepared by a Certified large wood requires involvement of a civil
Floodplain Manager to demonstrate the project engineer as well as review and concurrence by
is in compliance with FEMA regulations may be the utility owner.
required.

Large Wood National Manual July 2015


6-46
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

6.7.14.1 Water and Sewer Lines 6.7.14.3 Power Lines


A general rule and utility requirement is to limit The main constraint with power lines is the
excavation to areas at least 3 meters vertically limitation of overhead height, which will affect
and horizontally away from major sewer and the type of equipment used to move the large
water lines. In some circumstances exceptions wood into position and install it. Presence of
to this rule may occur, but such cases may power poles within the proposed project’s
require costly temporary shoring and channel or floodplain corridor is a common
stabilization and additional geotechnical constraint, though it may be feasible to relocate
analyses. It is very common for sewer and the power poles to a different location to permit
water lines to cross through and under urban installation of the large wood.
stream corridors. Channel incision driven by
urbanization frequently results in exposure and 6.7.15 Sediment and Debris
breakage of water and sewer lines buried under
the channel. Restoration of utility service will In urban settings the incoming sediment supply
often require mitigation in the form of may be much lower in quantity and finer than
placement of large wood within the channel. In under predevelopment conditions. In urban
these circumstances, placement of large wood settings, sediment sources may be cut off by
must be designed in a way such that erosion or revetments or retaining walls and impervious
deposition triggered by the large wood does not surfaces, with a shift from gravel substrate to a
threaten the restored utility crossing. Minor sand/silt-dominated substrate associated with
water lines (less than 20 centimeters [8 inches] road runoff including street sanding and soils
in diameter) can occasionally be routed around from residential yards. In some cases a bimodal
the project area, depending on the number of sediment size distribution develops, with higher
lateral service connections affected. The re- proportional quantities of both fine sediments
routing of larger (more than 31 centimeters (sand/silt) and larger materials (bricks, broken
[12 inches] in diameter) sewer lines is typically concrete) with little gravel or cobble present.
expensive due to gravity flow constraints and Large wood installations must be designed to
multiple lateral connections, but it may be accommodate these materials; for example,
feasible to lower elevations of smaller sewer anticipating infill within the voids between logs
lines to allow installation of large wood where by fine materials. Additionally, the designer
sufficient gradient exists to connect to the should recognize that conditions conducive to
receiving sewer trunk line. An additional bed degradation may exist, and gravel backfill
exceptional situation is revision of the base must be sized for stability under design
flood elevation to reflect levee setbacks or other conditions.
physical changes that have occurred
subsequent to the current base flood elevation Incoming debris from storm conditions can
determination. Base flood elevation revision is consist of trash, human-made objects of all
typically a costly process. types, yard waste, log rounds, tires, and
construction debris, as well as branches, leaves,
6.7.14.2 Gas Lines logs, and even whole trees. In tightly
constrained settings, it is prudent to avoid
Often it is feasible to relocate minor (less than structures that fully span the channel, as
10 centimeters [4 inches] in diameter) gas lines incoming debris can accumulate and exacerbate
to accommodate large wood placement, flooding. A general rule for reach layout is to
typically with directional drilling of the lines position structures on alternating banks at
around the zones where large wood is to be longitudinal intervals of at least one to two
installed. channel widths to avoid impairing flow

Large Wood National Manual July 2015


6-47
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

conveyance. When analyzing conveyance projects that meet ecological objectives and that
impacts, it is prudent to consider large wood can also be safely enjoyed by the public. Large
structures as impermeable and not allocate wood structures in these multiuse landscapes
channel capacity for flow that may flow through must be designed properly to benefit aquatic
or under the large wood structure. organisms and at the same time provide an
opportunity for both active and passive human
6.7.16 Existing and Historic interaction.

Structures 6.8.1 Landscape Integration


Existing and historic structures will impose Integrating large wood into larger habitat
lateral limits to installation. Excavation limits restoration projects takes additional planning
must be designed to avoid destabilization of and presents unique design and construction
structure foundations. In many cases the soils challenges. The horizontal and vertical elements
between the stream channel and historic of large wood structures sometimes require
structures may be poorly consolidated fill, and excavation with heavy machinery.
it may be difficult to obtain representative
borings. In constrained sites, soil retention
measures such as sheet piling may be required
GUIDANCE
to allow installation of large wood. High
groundwater tables and saturated soils create
the hazard of soil slumping during excavation to Depending on the timing of installation, several post-
install large wood; in these situations, installation actions may need to take place to ensure
other elements of the restoration project or
involvement of a geotechnical engineer is
surrounding environment are not adversely affected.
prudent. Where there is any likelihood of past
human habitation within the project area,  If the structures are installed in a newly graded
archaeological investigations should be landscape once installation is complete, the soil
scheduled early in the project to identify should be returned to the design finish grade
sensitive areas and features and required and compaction specification to ensure proper
drainage and sufficient soil cover over elements
avoidance or mitigation of impacts.
of the structure anchored in the bank. This will
help ensure there is sufficient rooting zone for
6.8 Integrating plantings and cover for irrigation lines that may
cross the buried portions of the structures.
Landscape  Any areas or soil around the work area that
Architecture were not excavated and backfilled should be
uncompacted as necessary and stabilized to
Landscape architecture is the profession that reduce erosion.
applies artistic and scientific principles to the
 If large wood structures are installed in areas
research, planning, design, and management of around existing vegetation or other sensitive
both the natural and built environments resources, precautions should be taken to avoid
(American Society of Landscape Architects adverse effects on these areas. This could
2013). In these two environments, the include exclusion fencing, onsite monitors, and
functional outcomes achieved from this process working within prescribed seasonal work
are often distinct and disparate but will need to windows.
overlap and coexist in multiuse landscapes.
Accommodating these interactions during the
planning and design process will help create

Large Wood National Manual July 2015


6-48
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

should be placed near the installations,


6.8.2 Public Use indicating that the slippery and uneven surfaces
Considerations as well as the unpredictable hydrodynamics on
and around the structures persist and that
Large wood structures often represent the focal climbing or standing on them should be
point of stream restoration and enhancement avoided. Depending on seasonal water
projects. Aesthetically they represent the elevations, submerged or partially submerged
seldom-seen macro-scale successional structures can also present a hazard to boaters.
processes that occur naturally in forests Local boat rental companies should be notified
adjacent to riverine systems. Wood structures about these hazards so they can inform
represent a link between terrestrial and aquatic customers as appropriate. Signage or other
ecosystems, and the microclimates they create notifications should be posted at put-in
are frequently heavily utilized by aquatic locations and launch ramps to inform boaters
organisms for refugia and foraging. Placed large about the structure locations and the potential
wood can provide opportunities for the public hazard they represent to watercraft.
to see these functional outcomes occurring.
Because large wood habitats are often attractive 6.8.3 Graphic Standards
to fish, they provide opportunities for fishing or
snorkeling. Figure 6-18 illustrates some common symbols
that can be used in restoration construction
To maximize the educational and passive
drawings to illustrate the size, extent, and
recreation opportunities presented by large
location of large wood structures. Regardless of
wood structures, trails and interpretive exhibits
the symbol type, large wood pieces and
should be sited close when possible.
structures should always be drawn to scale to
Interpretive exhibits can explain the ecological
convey the size and extent of each placement to
functions and benefits offered by large wood
the design team, community stakeholders, and
structures, and situating the exhibits next to
construction contractors. Construction details
trails will maximize their exposure to the
for individual structures should be developed
public. In areas where the public will interact
on a project-by-project basis so that structures
with large wood structures, adequate
are engineered properly for existing site
precautions should be taken to ensure the
conditions and prevailing hydrology.
public can safely do so and reduce the potential
liability for the landowner. Warning signs

Large Wood National Manual July 2015


6-49
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Figure 6-18. Large Wood Structure Graphic Standards

Large Wood National Manual July 2015


6-50
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

6.9 Uncertainties and Research Needs


1. Much of the existing knowledge base is derived from the Pacific Northwest. Regional differences
in climate, hydrology, geomorphology, large wood loading processes, wood species, wood sizes,
and aquatic ecology are significant; therefore, it is likely that some of the principles developed to
date are not universally applicable. There is a need for development of regional data and
expertise.
2. Ideally, large wood placements foster conditions where wood contributions from riparian zones
and floodplains sustain target-loading levels without further intervention. A basis is needed to
estimate the time required for natural regrowth of riparian and floodplain forests to sustain
instream wood levels, particularly for warm, humid regions where wood decay rates are most
rapid.
3. More information is needed about the role of vertical and inclined timber piling in stabilizing
engineered logjams and collecting natural wood, including guidelines for pile sizing,
embedment, species, and condition (e.g., using trees that are not certified as pile quality).
4. More modeling and data is needed on the role of “racking wood” on the stability and
performance of engineered structures.
5. More modeling and data and guidance is needed regarding the spatial layout and sizing of
engineered logjams within migrating channels
6. More information is needed on use of live trees as instream large wood.
7. Wood placements/structures have been designed to produce certain hydraulic effects related to
habitat character or erosion control. Accordingly, considerable research has targeted hydraulic
effects of wood, but little is known regarding the properties of wood placement that control
trapping efficiency of fluvially transported wood.
8. Effects of ice and ice-related events on instream wood, wood contributions to channels, and the
stability of placed wood is poorly understood.
9. Approaches for computing forces on wood structures due to ice, floating debris collisions,
bedload collisions, waves, and mud glows are needed.
10. Additional information is needed on the interaction between ice, instream wood, and channel
erosion and sedimentation.
11. Existing technology for computing forces due to wind waves and waves produced by boat wakes
should be adapted for large wood design.
12. Continued development of multidimensional hydrodynamic and sediment transport models is
needed to facilitate reach-scale design and comparison of alternatives. User-friendly interfaces
and utilities to facilitate selection of input values for coefficients, for example, are needed to
facilitate adoption and use of these tools by practitioners.

Large Wood National Manual July 2015


6-51
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

6.10 Key Points


1. Use of large wood in fluvial systems is not a new practice, but is becoming more common with
more of a focus on restoring physical and temporal complexity.
2. Large wood is generally used with a goal of assisting recovery of a degraded system to a state
where it naturally sustains levels of large wood loading commensurate with the target
ecosystem requirements.
3. Wood is not intended to be static like most engineering structures placed in channels. It decays,
shifts, accretes sediment and other wood and organic matter, and may be colonized by plants. It
is not considered a failure if it produces the desired channel and habitat characteristics even if is
not stationary.
4. Engineered (introduced, placed. or otherwise managed) wood is most common in the Pacific
Northwest. Wood is a major component in lotic ecosystems nationwide, but higher wood decay
rates, more extreme hydrologic variation, absence of boulders and cobbles, presence of ice, and
human encroachments into stream corridors may make instream wood engineering much more
complex in other regions.
5. Large wood introduction is not feasible in all settings. Applicability is sharply limited by site
geomorphology, availability of wood, and adjacent infrastructure.
6. Levels of effort appropriate for engineering wood projects vary from desktop, pencil and paper
exercises to multidimensional hydrodynamic and geomorphic modeling. Level of effort should
be proportional to the stream and project size and relative risk associated with failure.
7. Hydrologic considerations include selection of a design event that will produce the greatest
forces on the wood structure and assessment of wood influences on physical aquatic habitat
across the range of flows and seasonal requirements for species/life stages of interest.
8. Design of wood projects is a multi-disciplinary endeavor. This is reflected in the stamping of
plans by not just the professional engineer, but others with key design input such as the
professional geologist (with geomorphology expertise).
9. Design includes reach layout, or determining locations and configurations for placed wood.
Reach layout may be driven by habitat or erosion control requirements. Habitat may be viewed
at a local scale (increasing pool habitat or woody substrate) or at much larger scales (sediment
retention, island formation, or inducing channel planform evolution). Erosion control includes
both bed and bank erosion.
10. Following reach layout, designers should determine the dimensions of each structure. Usually,
this process follows experience and rules-of-thumb, but multidimensional computer models and
physical models may be justified for higher risk projects.
11. Design includes specifying the size, species, and other characteristics of wood materials. In
general, larger wood is in short supply and is often too costly to import. Decay-resistant species
are preferred. Environmental impacts of wood harvest should be considered.
12. Except for extremely simple projects, hydraulic analysis should include assessment of the flow
conveyance, sediment transport capacity, and velocity and shear stress at design discharge for
the existing channel and for the channel after large wood structure construction.

Large Wood National Manual July 2015


6-52
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

13. Flow forces on large wood include buoyancy, lift, and drag. A free body diagram may be used to
visualize the interactions among these forces and restraining forces of friction, gravity, and
restraints such as ballast, buried members, or anchors.
14. Restraints should be designed so that safety factors exceed a predetermined minimum.
15. Plans and designs may include provisions for planting vegetation that provides additional
stability and accelerates ecological recovery.
16. Design should include an assessment of constructability that deals with issues of access,
availability of materials, construction techniques and equipment, safety concerns, and
environmental restrictions.
17. Projects in urban areas face additional constraints due to utilities and other infrastructure in the
stream corridor, perturbed hydrology, historic structures, and floodplain regulation.

Large Wood National Manual July 2015


6-53
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

6.11 References
Abbe, T. B. 2000. Patterns, Mechanics, and Geomorphic Effects of Wood Debris Accumulations in a
Forest River System. Ph.D. dissertation. University of Washington, Seattle, WA. 222 pp.

Abbe, T. B., and A. P. Brooks. 2011. Geomorphic, Engineering, and Ecological Considerations when
Using Wood in River Restoration. Pages 419–451 in A. Simon, S. J. Bennett, and J. M. Castro
(eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools.
Geophysical Monograph Series 194. Washington, D.C.: American Geophysical Union.

Abbe, T. B., and D. R. Montgomery. 1996. Large Woody Debris Jams, Channel Hydraulics and Habitat
Formation in Large Rivers. Regulated Rivers: Research and Management 12:201–221.

Abbe, T. B., and D. R. Montgomery. 2003. Patterns and Processes of Wood Debris Accumulation in
the Queets River Basin, Washington. Geomorphology 51:81–107.

Abbe, T. B., D. R. Montgomery, and C. Petroff. 1997. Design of Stable In-Channel Wood Debris
Structures for Bank Protection and Habitat Restoration: An Example from the Cowlitz River, WA.
Pages 809–816 in S. S. Y. Wang, E. J. Langendoen, and F. D. Shields, F.D. (eds.), Proceedings of the
Conference on Management of Landscapes Disturbed by Channel Incision. University of
Mississippi, Oxford, MS.

Abbe, T. B, A. P. Brooks, and D. R. Montgomery. 2003b. Wood in River Rehabilitation and


Management. Pages 367–389 in S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology
and Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society.

Abbe, T. B., G. Pess, D. R. Montgomery, and K. L. Fetherston. 2003c. Integrating Engineered Log Jam
Technology into River Rehabilitation. In D. R. Montgomery, S. Bolton, D. Booth, and L. Wall
(eds.), Restoration of Puget Sound Rivers. Center for Water and Watershed Studies, University of
Washington, Seattle.

Ahmad, M. 1951. Spacing and Projection of Spurs for Bank Protection. Civil Engineering and Public
Works Review. March:172–174; April:256–258.

Allen, H. H., and J. R. Leech. 1997. Bioengineering for Streambank Erosion Control. Technical Report E
97-8, U.S. Army Corps of Engineers Waterways Experiment Station, Vicksburg, MS.

American Society of Landscape Architects (ASLA). 2013. American Society of Landscape Architects
(ASLA). Available: http://www.asla.org. Accessed: August 27, 2013.

Arneson, L. A., L. W. Zevenbergen, P. F. Lagasse, and P. E. Clopper. 2012. Evaluating Scour at Bridges.
Hydraulic Engineering Circular 18, FHWA-HIF-12-003, National Highway Institute, Federal
Highway Administration, Arlington, VA.

Bilby, R. E., J. T. Heffner, B. R. Fransen, F. W. Ward, and P. A. Bisson. 1999. Effects of Immersion in
Water on Deterioration of Wood from Five Species of Trees Used for Habitat Enhancement
Projects. North American Journal of Fisheries Management 19(3):687–695.

Bolton, S., A. Watts, T. Sibley, and J. Dooley. 1998. A Pilot Study Examining the Effectiveness of
Engineered Large Woody Debris (ELWD™) as an Interim Solution to Lack of LWD in Streams.
EOS, Transactions of the American Geophysical Union 79(45):F346.

Large Wood National Manual July 2015


6-54
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Borg, D., I. Rutherford, and M. Stewardson. 2007. The Geomorphic and Ecological Effectiveness of
Habitat Rehabilitation Works: Continuous Measurement of Scour and Fill around Large Logs in
Sand-Bed Streams. Geomorphology 89(1/2):205–216.

Braudrick, C. A., and G. E. Grant. 2000. When do Logs Move in Rivers? Water Resources Research
36(2):571–583.

Brooks, A. P. 2006. Design Guidelines for the Reintroduction of Wood into Australian Streams. Land &
Water Australia, Canberra.

Brooks, A. P., T. Howell, T. B. Abbe, and A. H. Arthington. 2006. Confronting Hysteresis: Wood Basin
River Rehabilitation in Highly Altered Riverine Landscapes of South-Eastern Australia.
Geomorphology 79(3/4):395–422.

Castro, J., and R. Sampson. 2001. Incorporation of Large Wood into Engineering Structures. Natural
Resource Conservation Service Engineering Technical Note Number 15. U.S. Department of
Agriculture. Boise, ID.

Cederholm, C. J., R. E. Bilby, P. A. Bisson, T. W. Bumstead, B. R. Fransen, W. J. Scarlett, and J. W. Ward.


1997a. Response of Juvenile Coho Salmon and Steelhead to Placement of Large Woody Debris in
a Coastal Washington Stream. North American Journal of Fisheries Management 17:947–963.

Cederholm, C. J., R. E. Bilby, P. A. Bisson, T. W. Bumstead, B. R. Fransen, W. J. Scarlett, and J. W. Ward.


1997b. Response of Juvenile Coho Salmon and Steelhead to the Placement of Large Woody
Debris in a Coastal Washington Stream. Transactions of the American Fisheries Society. 118:368–
378.

Cederholm, C. J., L. G. Dominguez, and T. W. Bumstead. 1997c. Rehabilitating Stream Channels and
Fish Habitat Using Large Woody Debris. In P. A. Slaney and D. Zaldokas (eds.), Fish Habitat
Procedures. Watershed Restoration Program, Ministry of Environment, Lands and Parks,
Vancouver, British Columbia.

Collins, B. D., D. R. Montgomery, K. L. Fetherston, and T. B. Abbe. 2012. The Floodplain Large-Wood
Cycle Hypothesis: A Mechanism for the Physical and Biotic Structuring of Temperate Forested
Alluvial Valleys in the North Pacific Coastal Ecoregion. Geomorphology 139/140:460–470.

Comiti, F. and Mao, L. 2012. Recent Advances in the Dynamics of Steep Channels. Gravel-bed Rivers:
Processes, Tools, Environments, pages 351–377.

Copeland, R. R. 1983. Bank Protection Techniques Using Spur Dikes. Paper No. HL-83-1. Hydraulics
Laboratory. U.S. Army Waterways Experiment Station. Vicksburg, MS.

Cramer, M. L. (ed.). 2012. Stream Habitat Restoration Guidelines. Copublished by the Washington
Departments of Fish and Wildlife, Natural Resources, Transportation and Ecology, Washington
State Recreation and Conservation Office, Puget Sound Partnership, and the U.S. Fish and
Wildlife Service. Olympia, WA.

Cramer, M., K. Bates, D. Miller, K. Boyd, L. Fotherby, P. Skidmore, T. Hoitsma, B. Heiner, K. Buchanan,
P. Powers, G. Birkeland, M. Rotar, and D. White. 2002. Integrated Streambank Protection
Guidelines. Washington State Aquatic Habitat Guidelines Program, Washington State Department
of Fish and Wildlife. Olympia, WA.

Large Wood National Manual July 2015


6-55
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Curran, J. C. 2010. Mobility of Large Woody Debris (LWD) Jams in a Low Gradient Channel.
Geomorphology 116:320–329.

D'Aoust, S. G. and R. G. Millar. 2000. Stability of Ballasted Woody Debris Habitat Structures. Journal
of Hydraulic Engineering 126(11):810–817.

Derrick, D. L. 1997. Twelve low-Cost, Innovative, Landowner Financed, Streambank Protection


Demonstration Projects. Pages 446–451 in Management of Landscapes Disturbed by Channel
Incision, Stabilization, Rehabilitation, and Restoration. Center for Computational Hydroscience
and Engineering, University of Mississippi.

Drury, T. A. 1999. Stability and Pool Scour of Engineered Log Jams in the North Fork Stillaguamish
River, Washington. Thesis, Master of Science in Civil Engineering. University of Washington,
Seattle.

Elliot, R., D. Froehlich, and R. MacArthur. 2012. Calculating the Potential Effects of Large Woody
Debris Accumulations on Backwater, Scour, and Hydrodynamic Loads. Pages 1213–1222 in
Proceedings of the World Environmental and Water Resources Congress 2012. Reston, VA:
American Society of Civil Engineers.

Erskine, W. D., M. J. Saynor, A. C. Chalmers, and S. J. Riley. J. 2012. Water, Wind, Wood, and Trees:
Interactions, Spatial Variations, Temporal Dynamics, and their Potential Role in River
Rehabilitation. Journal of Geographical Research 50(1):60–74.

Eslamian, S. 2014. Handbook of Engineering Hydrology. Boca Raton, FL: CRC Press.

Federal Emergency Management Agency (FEMA). 2009. NFIP Floodplain Management Guidebook: A
Local Administrator’s Guide to Floodplain Management and the National Flood Insurance
Program. Fifth Edition, Federal Emergency Management Agency Region 10. Bothell, WA.

Federal Interagency Stream Restoration Working Group (FISRWG). 1998. Restoration


Implementation, Monitoring, and Management. Chapter 9 in Stream Corridor Restoration:
Principles, Processes and Practices. National Technical Information Service, U. S. Department of
Commerce, Springfield, VA. Also published as NRCS, U.S. Department of Agriculture (1998)
National Engineering Handbook (NEH), Part 653. Washington, D.C.

Fischenich, C. 2001. Stability Thresholds for Stream Restoration Materials, Publication No. ERDC
TNEMRRP-SR-29. U.S. Army Engineer Research and Development Center, Vicksburg, MS.

Fischenich, C., and J.V. Morrow, Jr. 2000. Streambank Habitat Enhancement with Large Woody Debris.
Publication No. ERDC TN-EMRRP-SR-13. U.S. Army Engineer Research and Development Center.

Flosi, G., S. Downie, J. Hopelain, M. Bird, R. Coey, and B. Collins (eds.). 1998. California Salmonid
Stream Habitat Restoration Manual. 3rd ed. California: California Department of Fish and Game,
Inland Fisheries Division, Sacramento.

Forest Products Laboratory. 2010. Wood Handbook—Wood as an Engineering Material. General


Technical Report FPL-GTR-190. U.S. Department of Agriculture, Forest Service, Forest Products
Laboratory. Madison, WI.

Freschet, G. T., J. T. Weedon, R. Aerts, J. R. van Hal, and J. H. Cornelissen. 2012. Interspecific
Differences in Wood Decay Rates: Insights from a New Short-Term Method to Study Long-Term
Wood Decomposition. Journal of Ecology 100:161–170. doi: 10.1111/j.1365-2745.2011.01896.x.

Large Wood National Manual July 2015


6-56
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Frissell, C. A., and R. K. Nawa. 1992. Incidence and Causes of Physical Failure of Artificial Habitat
Structures in Streams of Western Oregon and Washington. North American Journal of Fisheries
Management 12 182–197.

Gippel, C. J., I. C. O’Neill, and B. L. Finlayson. 1996. Distribution and Hydraulic Significance of Large
Woody Debris in a Lowland Australian River. Hydrobiologia 318:179–194.

Gotvald, A. J., N. A. Barth, A. G. Veilleux, and C. Parrett. 2012. Methods for Determining Magnitude and
Frequency of Floods in California, Based on Data Through Water Year 2006. U.S. Geological Survey
Scientific Investigations Report 2012–5113. Available: http://pubs.usgs.gov/sir/2012/5113/.

Gurnell, A. M., H. Piegay, F. J. Swanson, F. J. and S. V. Gregorys. 2002. Large Wood and Fluvial
Processes. Freshwater Biology 47(4):601–619.

Harmon, M. E., J. F. Franklin, F. J. Swanson, P. Sollins, S. V. Gregory, J. D. Lattin, N. H. Anderson, S. P.


Cline, N. G. Aumen, J. R. Sedell, G. W. Lienkaemper, K. Cromack, Jr., and K. W. Cummins. 1986.
Ecology of Coarse Woody Debris in Temperate Ecosystems. Advances in Ecological Research
15:133–302.

He, Z., W. Wu, and F. D. Shields, Jr. 2009. Numerical Analysis of Effects of Large Wood Structures on
Channel Morphology and Fish Habitat Suitability in a Southern U.S. Sandy Creek. Ecohydrology 2
(3):370–380. doi: 10.1002/eco.60.

Herrera Environmental Consultants, Inc. 2006. Conceptual Design Guidelines: Application of


Engineered Logjams. Prepared for Scottish Environmental Protection Agency, Galashiels, United
Kingdom.

Hertzberg, R. 1954. Wave-Wash Control on Mississippi River Levees. Transactions of the ASCE
119(2688):628–638.

Hilderbrand, R. H., A. D. Lemly, C. A. Dolloff, and K. L. Harpster. 1998. Design Considerations for
Large Woody Debris Placement in Stream Enhancement Projects. North American Journal of
Fisheries Management 18:161–167.

Hyatt, T. L., and R. J. Naiman. 2001. The Residence Time of Large Woody Debris in the Queets River,
Washington, USA. Ecological Applications 11(1):191–202.

ICF International. 2010. Instream Woody Material Installation and Monitoring Guidance Manual.
Sacramento Area Flood Control Agency, Sacramento, California.

Interagency Advisory Committee on Water Data (IACWD). 1982. Guidelines for Determining Flood
Flow Frequency. Bulletin 17B of the Hydrology Subcommittee, Office of Water Data Coordination,
U.S. Geological Survey, Reston, Virginia. 183 p.

Jia, Y., S. Scott, Y. Xu, and S. S. Y. Wang. 2009. Numerical Study of Flow Affected by Bendway Weirs in
Victoria Bendway, the Mississippi River. Journal of Hydraulic Engineering 135(11):902–916.

Johnson, A. W., and J. M. Stypula (eds.). 1993. Guidelines for Bank Stabilization Projects in the Riverine
Environments of King County. King County Department of Public Works, Surface Water
Management Division. Seattle, WA.

Johnson, P. A., R. D. Hey, M. Tessier, and D. L. Rosgen. 2001. Use of Vanes for Control of Scour at
Vertical Wall Abutments. Journal of Hydraulic Engineering 127(9):772–778.

Large Wood National Manual July 2015


6-57
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Kail, J., D. Hering, S. Muhar, J. Gerhard, and S. Preis. 2007. The Use of Large Wood in Stream
Restoration: Experiences from 50 Projects in Germany and Austria. Journal of Applied Ecology
44:1145–1155.

Keown, M. P., N. R. Oswalt, E. B. Perry, and E. A. Dardeau Jr. 1977. Literature Survey and Preliminary
Evaluation of Streambank Protection Methods. Technical Report No. WES-TR-H-77-9, U.S. Army
Engineer Waterways Experiment Station, Vicksburg, MS.

Klingeman, P. C., S. M. Kehe, and Y. A. Owusu. 1984. Streambank Erosion Protection and Channel
Scour Manipulation Using Rockfill Dikes and Gabions. Water Resources Research Institute,
Oregon State University. Salem, OR.

Knutson, M., and P. Fealko. 2014. Pacific Northwest Region Resource and Technical Services—Large
Woody Material Risk Based Design Guidelines. U.S. Department of the Interior, Bureau of
Reclamation, Pacific Northwest Region, Boise, Idaho.

Kruys, N., B. G. Jonsson, and G. Stahl. 2002. A Stage-Based Matrix Model for Decay-Class Dynamics of
Woody Debris. Ecological Applications 12(3):773–781.

Kuhnle, R. A., C. V. Alonso, and F. D. Shields Jr. 1999.Volume of Scour Holes Associated with 90-
degree Spur Dikes. Journal of Hydraulic Engineering 125(9):972–978.

Kuhnle, R. A., C. V. Alonso, and F. D. Shields Jr. 2002. Local Scour Associated with Angled Spur Dikes.
Journal of Hydraulic Engineering 128(12):1087–1093.

Lagasse, P. F., P. E. Clopper, J. E. Ortiz-Page, L. W. Zevenbergen, L. A. Ameson, J. D. Schall, and L. G.


Girard. 2009. Bridge Scour and Stream Instability Countermeasures Experience, Selection and
Design Guidance Volumes 1 and 2. HEC-23, Third Edition. Federal Highway Administration.
Washington, D.C.

Lagasse, P. F., P. Clopper, L. Zevenbergen, W. Spitz, and L. G. Girard. 2010. Effects of Debris on Bridge
Pier Scour. Federal Highway Administration. Washington, D.C.

Lassettre, N. S., and G. M. Kondolf. 2012. Large Woody Debris in Urban Stream Channels: Redefining
the Problem. River Research and Applications 28.9 (2012):1477–1487.

Lester, R. E. and A. J. Boulton. 2008. Rehabilitating Agricultural Streams in Australia with Wood: A
Review. Environmental Management 42(2):310–326.

Merten, E., J. Finlay, L. Johnson, R. Newman, R., H. Stefan, and B. Vondracek. 2010. Factors
Influencing Wood Mobilization in Stream. Water Resources Research 46:W10514.

Miles, P. D., and W. B. Smith. 2009. Specific Gravity and Other Properties of Wood and Bark for 156
Tree Species Found in North America. U.S. Forest Service, Newtown Square, PA.

Montgomery, D. R., and T. B. Abbe. 2006. Influence of Logjam-Formed Hard Points on the Formation
of Valley-Bottom Landforms in an Old-Growth Forest Valley, Queets River, Washington, USA.
Quaternary Research 65:147–155.

Montgomery, D. R., T. B. Abbe, J. M. Buffington, N. P. Peterson, K. M. Schmidt, and J. D. Stock. 1995a.


Distribution of Bedrock and Alluvial Channels in Forested Mountain Drainage Basins. Nature
381:587–589.

Large Wood National Manual July 2015


6-58
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Montgomery, D. R., J. M. Buffington, R. D. Smith, K. M. Schmidt, and G. Pess. 1995b. Pool Spacing in
Forest Channels. Water Resources Research 31:1097–1105.

Montgomery, D. R., B. D. Collins, J. M. Buffington, and T. B. Abbe. 2003. Geomorphic Effects of Wood
in Rivers. Pages 21–47 in S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology and
Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society.

Nanson, G. C., M. Barbetti, and G. Taylor. 1995. River Stabilisation due to Changing Climate and
Vegetation During the late Quaternary in Western Tasmania, Australia. Geomorphology
13.1(1995):145–158.

Natural Resources Conservation Service (NRCS). 2007a. Streambank Soil Bioengineering. Technical
Supplement TS 14I in Stream Restoration Design, National Engineering Handbook Part 654.
USDA-NRCS, Washington, D.C. CD-ROM.

Natural Resources Conservation Service (NRCS). 2007b. Use and Design of Soil Anchors. Technical
Supplement TS 41E in Stream Restoration Design, National Engineering Handbook Part 654.
USDA-NRCS, Washington, D.C. CD-ROM.

Natural Resources Conservation Service (NRCS). 2007f. Stream Hydrology. Chapter 5 in Stream
Restoration Design, National Engineering Handbook Part 654. USDA-NRCS, Washington, D.C. CD-
ROM.

Oregon Department of Fish and Wildlife (ODFW). 2010. Guide to Placement of Wood, Boulders, and
Gravel for Habitat Restoration. Oregon Departments of Forestry and Fish and Wildlife. Salem, OR.

Oregon Department of Transportation (ODOT). 2011. Hydraulics Manual, Engineering and Asset
Management. Unit Geo-Environmental Section. Salem, OR.

Petersen, M. S. 1986. River Engineering. Englewood Cliffs, NJ: Prentice-Hall.

Piégay, H., and J. P. Bravard. 1997. Response of a Mediterranean Riparian Forest to a 1 in 400 Year
Flood, Ouveze River, Drome-Vaucluse, France. Earth Surface Processes and Landforms 22(1):31–
43.

Pokrefke, T. J. (ed.) 2013. Inland Navigation: Channel Training Works. ASCE Manual of Practice 124.
American Society of Civil Engineers. Reston, VA.

Quinault Indian Nation (QIN). 2008. Salmon Habitat Restoration Plan for the Upper Quinault River.
Quinault Indian Nation Department of Fisheries. Taholah, Washington. Prepared by T. Abbe and
others.

Roni, P., K. Hanson, and T. Beechie. 2008. Global Review of the Physical and Biological Effectiveness
of Stream Habitat Rehabilitation Techniques. North American Journal of Fisheries Management
28(3):856–890.

Ruiz-Villanueva, V., M. Stoffel, H. Piégay, V. Gaertner, and F. Perret. 2014. Wood Density Assessment
to Improve Understanding of Large Wood Buoyancy in Rivers. Pages 2503–2508 in A. Schleiss,
G. De Cesare, M. Franca, and M. Pfister (eds.), River Flow. London, England: Taylor and Francis.

Saldi-Caromile, K., K. Bates, P. Skidmore, J. Barenti, and D. Pineo. 2004. Stream Habitat Restoration
Guidelines: Final Draft. Co-published by the Washington Departments of Fish and Wildlife and
Ecology and the U.S. Fish and Wildlife Service. Olympia, Washington.

Large Wood National Manual July 2015


6-59
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Sauer, V. B. 1974. Flood Characteristics of Oklahoma Streams Techniques for Calculating Magnitude
and Frequency of Floods in Oklahoma, with Compilations of Flood Data Through 1971. U.S.
Geological Survey Water-Resources Investigations Report 73–52. 307 p.

Scheffer, T. C. 1971. A Climate Index for Estimating Potential for Decay in Wood Structures Above
Ground. Forest Products Journal 21(10):25–31.

Shields, F. D., Jr. 2007. Scour Calculations. Technical Supplement 14B in Stream Restoration Design.
National Engineering Handbook Part 654. USDA-NRCS. Washington, D.C. CD-ROM.

Shields, F. D., Jr., and C. V. Alonso. 2012. Assessment of Flow Forces on Large Wood in Rivers. Water
Resources Research 48(4):W04156.

Shields, F. D., Jr., and C. J. Gippel. 1995. Prediction of Effects of Woody Debris Removal on Flow
Resistance. Journal of Hydraulic Engineering 121 (4):341–354.

Shields, F. D., Jr., and A. D. Wood. 2007. The Use of Large Woody Material for Habitat and Bank
Protection. Technical Supplement 14J in Stream Restoration Design, National Engineering
Handbook Part 654. USDA-NRCS Washington, D.C. CD-ROM.

Shields, F. D., Jr., A. J. Bowie, and C. M. Cooper. 1995. Control of Streambank Erosion due to Bed
Degradation with Vegetation and Structure. Water Resources Bulletin 31(3):475–489.

Shields, F. D., Jr., N. Morin, and C. M. Cooper. 2004. Large Woody Debris Structures for Sand-Bed
Channels. Journal of Hydraulic Engineering 130(3):208–217.

Shields, F. D. Jr., S. S. Knight, and J. M. Stofleth. 2006. Large Wood Addition for Aquatic Habitat
Rehabilitation in an Incised, Sand-Bed Stream, Little Topashaw Creek, Mississippi. River
Research and Applications 22:803–817.

Shields, F. D., Jr., S. R. Pezeshki, G. V. Wilson, W. Wu, and S. M. Dabney. 2008. Rehabilitation of an
Incised Stream with Plant Materials: The Dominance of Geomorphic Processes. Ecology and
Society 13 (2):54.

Simon, A., A. Curini, S. E. Darby, and E. J. Langendoen. 2000. Bank and Near-Bank Processes in an
Incised Channel. Geomorphology 35(3):193–217.

Simon, A., A. Brooks, and N. Bankhead. 2012. Effectiveness of Engineered Log Jams in Reducing
Streambank Erosion to the Great Barrier Reef: The O’Connell River, Queensland, Australia. Pages
2570–2577 in World Environmental and Water Resources Congress 2012: Crossing Boundaries.
Reston, VA: ASCE.

Simon, A., R. Thomas, A. Curini, and N. Bankhead. 2014. Development of the Bank-Stability and Toe-
Erosion Model (BSTEM version 5.4). Available: www.kwo.org/reports_publications/
Presentations/pp_Development_of_BSTEM_012811_sm.pdf.

Simpson, W. and A. TenWolde. 1999. Physical Properties and Moisture Relations of Wood. Chapter 3
in Wood Handbook: Wood as an Engineering Material. Report FPL-GTR-113. U.S. Department of
Agriculture Forest Service. Forest Products Laboratory. Madison, WI.

Smith, D. L., J. B. Allen, O. Eslinger, M. Valenciano, J. Nestler, and R. A. Goodwin. 2011. Hydraulic
Modeling of Large Roughness Elements with Computational Fluid Dynamics for Improved
Realism in Stream Restoration Planning. Geophysical Monograph Series 194:115–122.

Large Wood National Manual July 2015


6-60
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

Subramanya, K., 2008. Engineering Hydrology. New York: McGraw-Hill. 434 pp.

Svoboda, C. D. and K. Russell, K. 2011. Flume Analysis of Engineered Large Wood Structures for
Scour Development and Habitat. Pages 2572–2581 in Proceedings, World Environmental and
Water Resources Congress, ASCE, Reston, VA.

Sylte, T., and C. Fischenich. 2000. Rootwad Composites for Streambank Erosion Control and Fish
Habitat Enhancement. U.S. Army Corps of Engineers. Vicksburg, MS.

Thompson, D. M. 2002. Channel-bed Scour with High Versus Low Deflectors. Journal of Hydraulic
Engineering 128(6):640–643.

Thompson, D. M., and Stull, G. N. 2002. The Development and Historic Use of Habitat Structures in
Channel Restoration in the United States: The Grand Experiment in Fisheries Management.
Géographie physique et Quaternaire 56(1):45–60.

Turnipseed, D. P., and V. B. Sauer. 2010. Discharge Measurements at Gaging Stations: U.S. Geological
Survey Techniques and Methods Book 3, Chapter A8, U.S. Geological Survey.

U.S. Army Corps of Engineers, 1981. The Streambank Erosion Control Evaluation and Demonstration
Act of 1974. Final Report to Congress, Main Report. Washington, D.C.

U. S. Army Corps of Engineers. 1992. Engineering and Design: Bearing Capacity of Soils. EM 1110-1-
1905. Department of the Army, U.S. Army Corps of Engineers. Washington, D.C.

U. S. Army Corps of Engineers. 1994. Engineering and Design: Hydraulic Design of Flood Control
Channels. EM 1110-2-1601. Department of the Army, U.S. Army Corps of Engineers. Washington,
D.C.

U. S. Army Corps of Engineers. 2005. Engineering and Design: Stability Analysis of Concrete
Structures. EM 1110-2-2100. Department of the Army, U.S. Army Corps of Engineers.
Washington, D.C.

U.S. Department of Agriculture (USDA), Agricultural Research Service. 2013. Bank Stability and
Erosion Model. Available: http://www.ars.usda.gov/Research/docs.htm?docid=5044&page=1.

Valverde, R. S. 2013. Roughness and Geometry Effects of Engineered Log Jams on 1-D Flow
Characteristics. M. S. Thesis, Civil Engineering, Oregon State University, Corvallis.

Vanoni, V. 1975. Sedimentation Engineering, ASCE Manuals and Reports on Engineering Practice—No.
54. American Society of Civil Engineers, New York, NY, pp. 531–538.

Viessman, W. J., and G. L. Lewis. 2003. Introduction to Hydrology. Prentice Hall. 612 pp.

Wallerstein, N. P., C. V. Alonso, S. J. Bennett, and C. R. Thorne. 2001. Distorted Froude-Scaled Flume
Analysis of Large Woody Debris. Earth Surface Processes and Landforms 26:1265–1283.

Wohl, Ellen. 2013. Floodplains and Wood. Earth-Science Reviews 123:194–212.

Wood, A. D., and A. R. Jarrett. 2004. Design Tool for Rootwads in Streambank Restoration. Paper
042047, Annual International Meeting, Ottawa. American Society of Agricultural Engineers. St.
Joseph, MI.

Large Wood National Manual July 2015


6-61
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 6. Engineering Considerations

This page intentionally left blank.

Large Wood National Manual July 2015


6-62
Chapter 7
RISK CONSIDERATIONS

Channel spanning logjam in the Deschutes River of central Oregon providing a complex range
of habitat conditions and cover (Tim Abbe, March 2013).

AUTHORS

Tim Abbe (NSD)


Leif Embertson (NSD)
This page left intentionally blank.
large wood to geomorphic processes, habitat
7.1 Purpose complexity, and ultimately, the health of aquatic
habitat has become better understood. This has
This chapter provides an overview of how to led to federal, state, local, tribal, and private
assess risk when introducing and managing citizens placing large wood in streams in the
wood in stream restoration and management. form of ELJs and large wood placements. These
Risk is a major concern with many stakeholders endeavors are an effort to restore channel
based on real and perceived threats that wood processes, create habitat for the purposes of
may pose or because infrastructure was never enhancing geomorphic and habitat processes,
designed for wood conveyance. This is not and restoring flood-protection measures that
surprising given over a century of channel will enhance fluvial processes while also
clearing led by local, state, and federal protecting infrastructure and property.
government agencies. Inappropriate
placements and poorly designed structures can While the placement of large wood is an
introduce unacceptable risks. Risk also applies important component of river restoration
to ecological effects caused by design strategies, many of the rivers and streams have
alternatives that do more harm than good. significant constraints such as civil and private
Every project should consider a range of infrastructure, private property ownership, and
alternatives that always includes a no-action a host of recreational activities within the river
scenario. An objective risk assessment not only corridors. These constraints are commonly
provides insight to how and where things can damaged by natural fluvial processes and are at
go wrong, but provides justification on why risk from flooding and erosion, resulting in
restoration is needed. significant public and private investments in
river-training structures and flood-control
projects.
7.2 Introduction
Current river and floodplain management
The design and placement of large wood practices acknowledge that there are inherent
structures and riparian reforestation has been risks associated with any river that could
recognized as a beneficial element of stream negatively affect recreation opportunities,
and river restoration strategies (Roni et al. commerce, infrastructure, and existing
2014a). Historically, large wood was naturally buildings in a flood or channel migration zone.
abundant and strongly influenced fluvial Since the European settlement of North
morphology in virtually every river and stream America started, large financial investments
network where riparian forests were present. have been made in the United States to reduce
In the past 100 to 200 years, large wood has flood risks and improve navigation through the
been actively removed from streams for a removal of large wood, dam building, channel
variety of reasons, such as: to increase channel confinement, channel training, and flood-
flood capacity, improve channel navigability, control projects, which have singly and in
and improve safety to the general public. combination greatly simplified aquatic habitat
Although most of the initial wood and disrupted natural fluvial processes. It is
reintroduction projects in North America were important to recognize that extensive natural
well intentioned, some projects had limited wood accumulations can coexist with existing
success due to the insufficient understanding of infrastructure. For example, large
the fluvial processes, and the forces to which accumulations of wood in Long Tom Creek in
the structures would be subjected, how the Venata, Oregon, occur adjacent to the annual
project would influence these processes, and Oregon Country Fair (Figure 7-1). The wood
how changes in those processes would affect probably increases the frequency of floodplain
habitat. In the past 30 years, the importance of
Large Wood National Manual July 2015
7-1
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

inundation during the fall and winter, but incision (i.e., the natural or anthropogenic
holding the Country Fair during the summer downcutting of a river that occurs in the long-
has avoided this potential issue. In the Upper term erosion of a landscape). In many
Yakima River near the town of Easton, watersheds the clearing of wood and channel
Washington, a channel-spanning logjam just straightening has increased the speed at which
100 meters (328 feet) from Interstate 90 has floods move downstream and the stream’s
obstructed flows for decades without erosive power to create incised channels that
threatening the highway, due to natural are disconnected from their floodplains, further
development of new side channels (Abbe et al. exaggerating downstream flooding. Channel
2003a). incision can threaten infrastructure buried
Figure 7-1. Natural Logjam on Long Tom Creek under, going over, or near existing channels. It
near Venata, Oregon is imperative to consider that the failure to
correctly restore wood in channels can also
pose significant risks. Too often risk
assessments are one-sided, simply focusing on
traditional definitions of risk and the historic
perceptions of wood. Restoration of fluvial
systems using large wood should always
consider the risk of not restoring the system
correctly, and, in most cases, it can be
demonstrated that restoration provides the
greatest long-term benefits and reductions in
risk. Much of this chapter focuses on factors to
consider in assessing risks of reintroducing
wood to streams from the perspective of
flooding, erosion, and public safety. Evaluating
risk in river restoration has received increased
The simplification of rivers and streams has attention, and some sort of risk assessment is
also led to public misconceptions of riparian typically included in many restoration projects.
systems and their natural fluvial processes. As Thorne et al. (2014b) describe the project risk
addressed in previous chapters, a healthy and screening matrix, including the RiverRAT
productive river is complex on multiple levels. guidelines published by the National Oceanic
The “mess” many people perceive when seeing and Atmospheric Administration
wood accumulating in a river channel is quite (http://restorationreview.com/).
the opposite from an ecological perspective.
With any fluvial project, managing risk must The initial and essential step in conducting the
also include gaining an understanding of and risk assessment is to document the existing,
managing people’s perceptions. By helping the inherent, or background risks found within the
public understand fluvial science and the stream or river in question. This is particularly
historic context, project sponsors and important given that the assessment is intended
practitioners can expect greater public support to show whether or not the addition of large
for future projects. For instance, wood wood will introduce risks not found in the
accumulations, once common in streams, were system or will increase existing risks. Natural
very effective at dissipating energy, slowing wood can pose direct and indirect risks. Direct
down flows, trapping sediment, engaging risks are those that create a direct impact, such
floodplains, and creating vast wetlands. These as a person who is entangled in wood situated
naturally occurring events moderated in flowing water. Indirect risks are those where
downstream flooding and reduced channel wood contributes to a problem, such as wood

Large Wood National Manual July 2015


7-2
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

blocking a culvert, which in turn leads to educating communities about the importance of
flooding. Both of these risks can exist where large wood placements and the conditions they
wood is naturally entering streams. For create, a safer environment can be attained.
example, natural logjams still occur throughout This can be partially achieved by providing local
North America and pose a significant hazard to user groups with interpretative kiosks at entry
unprepared boaters (e.g., kayakers, canoeists, points and posting warning signage where
rafters, and fishermen) and have been instream structures are located. Regardless of
contributing factors in fatalities, injuries, and these educational efforts, there may continue to
close calls. In all such instances, the people be concerns about placing wood in channels
involved assumed personal responsibility for even though project sponsors and designers
the risks associated with entering the river. The have done their due diligence and have
persistence of the traditional view that “clean” conducted public meetings, performed other
rivers are safer contradicts the reality that in means of outreach, and included signage as part
some circumstances wood provides benefits of the project. To address these concerns and
that slow flows down, limit channel incision, also meet the professional engineering design
and create safer conditions for the public on the standards, some level of risk assessment should
whole. The widespread perception that wood is be incorporated into the standard of practice
a hazard will have to be addressed by many for wood design and managing streams. It
restoration projects, increasing the time and should also be recognized that restoring rivers
cost of implementation. At the same time, many to their natural state will change the way they
user groups are strong advocates for restoring look to the public. After decades of clearing
river corridors to the benefit of aquatic species wood from streams it can be expected that
and allowing the natural channel-forming restoration professionals will need to work
processes to continue. Additionally, user safety with local communities and recreational users.
is a goal for these advocates, where large wood
Accurately predicting the geomorphic response
is seen as a hazard, suggesting many current
of restoration projects or changes in stream
river users and advocates are unaware of the
management can influence relationships with
influence large wood historically had on
stakeholders and local communities which
properly functioning fluvial processes and
affect future projects and underscore the
aquatic habitat conditions.
importance of having well qualified
It is also important to consider that the river geomorphologists involved with design. In
environment is inherently dangerous because many cases, large wood placements have been
of large wood delivery and the dynamic designed to achieve the maximum geomorphic
behavior of channels as they continually adjust and habitat benefit, which has in some instances
their form, alignment, and character through resulted in conflicts with the general public and
changes in water and sediment delivery. recreational users. In 2010, public concerns
over large wood placements in the Entiat River
Wood poses little threat when boaters are
watershed in Washington State led to
careful to inspect the channel prior to floating it
construction delays and added coordination
to identify potential obstructions and where to
costs. Ultimately, maintaining and improving
line or portage boats around a hazard. Where
the safety around large wood placements
wood is placed in recreational rivers, warning
requires careful planning, public education, and
and educational signage is sometimes required
outreach to reduce conflicts. Urban rivers and
by local authorities. Education, such as signage
streams may pose extra consideration because
at boat ramps and information supplied to user
of infrastructure, private property, upstream
groups, leads to less risk, and, conversely, little
effects, downstream effects, and recreational
or no information leads to greater risks. By
uses or aesthetic considerations. The most

Large Wood National Manual July 2015


7-3
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

common failure of large wood placements the design life of large wood placements may be
during early attempts has been the wood simply subjective, based on typical design life of
washing away and the failure to achieve the existing infrastructure, or subject to such
intended effects at the reach scale (Frissell and stochastic events as major ice dam breaks.
Nawa 1992). But advancements in the Figure 7-2. Scour Undermining Downstream
understanding of wood stability and Corner of an ELJ on Upper Quinault, Washington
engineering instream structures have
dramatically increased the performance of
wood placements (Nooksack Tribe 2013).

The design life of wood structures continues to


be a common question that risk assessments
address, particularly the mechanical failure of a
structure during a particular design flow. Force
balance calculations (see Chapter 6, Engineering
Considerations) should be done to ensure any
structure has the desired stability. These
calculations should not only account for
buoyancy and drag, but how the structure is
designed to deal with bed scour (Figure 7-2).

Scour is one of the most common failure


mechanisms of instream structures, both in To summarize, managing all streams involves
traditional river engineering and restoration. some level of risk, including leaving the system
Wood structure design should also assess and current processes in their current
design life with respect to decay. Wood that condition. Placing anything, including wood,
remains submerged can last indefinitely, so into a stream immediately incurs risk as the
wood placed below base flow will have a much object could be washed away, which may negate
longer design life than wood subjected to the purpose for which it was intended and pose
wetting and drying. Some types of wood such as downstream risks. Stable large wood structures
cedar are naturally more resistant to decay, so induce changes that can be beneficial or create
that should be taken into account in estimating undesired results. It is important that large
design life. Some permitting agencies may not wood designs carefully consider project goals,
require projects to clearly describe the define acceptable risk, determine the critical
measures to retain wood placements, but may factors contributing to unacceptable risks, and
require an assessment of the fate of wood develop designs that achieve project goals
should it be washed downstream (e.g., will it within acceptable risks. Therefore, the simple
threaten downstream infrastructure?). act of conducting a risk assessment helps to
reduce the chances of an adverse effect.
In the restoration context, large wood
placements should have a design life sufficient
to restore habitat conditions, achieve the
project objectives, and sustain long-term
recovery of the system. This is typically the time
needed to reestablish riparian trees large
enough to create functional (stable) instream
wood and natural processes to deliver the wood
into the river (e.g., bank erosion, wind throw
associated with severe weather). In other cases
Large Wood National Manual July 2015
7-4
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

stability, hydraulics, channel response, human


RISKS
behavior, and monetary costs. Any risk
assessment is inherently subjective given the
Using Wood in Stream Restoration many factors that influence streams, so a great
deal depends on professionals familiar with
 Loss or washout of wood placement that results
fluvial processes and historical knowledge of
in failure to achieve restoration goals.
regional streams in unmanaged forest settings
 Washout of wood placement that would and in disturbed settings. An assessment can be
threaten downstream infrastructure: as simple as using empirical guidelines to size
o Bridge pier scour and then place functional wood into a remote
stream. This scenario assumes that there would
o Bridge or culvert blockages causing flooding be no adverse consequences because the stream
 Large wood structures triggering unintended
is located in a protected watershed with little to
geomorphic changes in a river corridor which no development or recreation. Conversely, a
damage adjacent infrastructure, property, or sophisticated assessment can include detailed
habitat (e.g., structure intended to protect a engineering, flood scenarios, wood transport,
bank could cause unintended erosion of an wood decay rates, and other analyses, to clearly
adjacent or opposite bank). evaluate all reasonable scenarios that could
occur. Potential changes in both the natural and
 Large wood structures raise water elevations
above existing regulatory mandates (e.g., the built environment, such as climate change and
FEMA 100-year flood). increased development, introduce uncertainty
and may require more detailed analysis and
 Large wood structures collecting sufficient predictions to adequately assess risk. Rivers
debris to increase flooding or create hazardous themselves have inherent dangers where the
conditions.
variability in river conditions are constantly
 Large wood structures altering sediment changing and evolving. As such, risks for any
transport characteristics within the reach, given project are situation-specific and should
resulting in local aggradation or scour affecting be evaluated relative to the watershed and user
flooding or infrastructure failure. groups associated with the project, stream type,
 The presence of large wood structures project context, and project components.
encouraging beaver activity in the area leading Risk assessments are increasingly being
to formation of beaver dams, with subsequent
incorporated into stream restoration and river
flooding impacts on adjacent properties or
creation of fish passage barriers. management to better ensure that projects have
considered potential adverse consequences.
 Large wood used in bank stabilization rotting Standard engineering practices typically include
out over time, leaving a “soft spot” in the road a risk assessment involving a structure’s
prism or hillside that would then be exposed to stability and safety. In stream restoration
future slope instability.
practice, large wood placements have been a
primary driver for completing risk assessments
because of the concern about public safety
7.3 Defining and Assessing associated with instream structures that many
people are not familiar with, do not understand,
Risk or see as a potential threat.
Assessing risk can be relatively simple when Risk is ubiquitous within the river environment
there is little or no consequence to failure, or it given channel responses to changes related to
can entail complex quantitative analyses of water, sediment, and the delivery of natural

Large Wood National Manual July 2015


7-5
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

large wood. The purpose of any risk assessment areas that will have economic damages
is not to eliminate risk but to objectively (consequences). In the case of large wood
evaluate the potential risk elements and assess placement, if it is known that a large wood
how a particular large wood placement or structure would fail during a 100-year flood
project can be designed and installed to address event, then consequences must be assigned to
and alleviate those risks. It is also important to the structure’s failure. This could be as simple
note that there can often be a significant risk to as the economic loss associated with how much
continued geomorphic and habitat degradation it cost to build the wood structure or additional
if large wood is not reintroduced to a stream or factors such as the wood accumulating on a
river, and this should be considered in every bridge pier. Assuming a large wood structure
risk assessment. This highlights the importance has a 50% probability of surviving the 1%
of having a professional geologist with expertise probability flood event, then the large wood
in fluvial geomorphology involved with design structure has a 0.5% (0.5 x 0.01 = 0.005)
and risk assessments, including approving plan probability of failing in any given year. If the
sets and reports. A primary purpose of a risk structure does fail, the consequences also may
assessment is to assure the design team, have a particular probability. For instance, if the
stakeholders, and local community that the structure fails, there will be some probability
short- and long-term effects of the project have from 0 to 1 of wood accumulating on a bridge
been considered, and the expected benefits of pier that would pose a problem. If that
the project outweigh the potential probability was 1%, then the actual probability
consequences. of wood causing a problem at the bridge would
be 0.005% (0.005 x 0.1 = 0.00005) in any given
Risk is most commonly defined as the product
year. If the consequences are $100,000 in
of the probability of a certain event occurring
emergency maintenance then the total risk
with the consequences of that event. This is
would be $5 per year summed over 50 years,
expressed as the following equation.
which would be $250. But if wood on the bridge
Risk  Ph x  C  resulted in a failure requiring a $5,000,000
bridge replacement, then the cumulative risk
Where: would be $500 per year, or $25,000 over
50 years.
P(h) = Probability of a specific event or
combination of events occurring. Risk also must be computed for the no-action
alternative and can often result in identification
∑(C) = Summation of the consequences of of greater risk to both ecological and
event occurring, typically presented socioeconomic conditions than restoration
as a monetary cost. involving wood placement. This is especially
true in the case of habitat restoration when the
If there are no negative consequences of a
no-action approach results in further
particular event occurring, then there is no risk.
degradation of habitat that leads to higher costs
If the consequences are very severe, then even
to restore in the future. A no-action alternative
an event with low probability of occurrence
can also put infrastructure at risk, particularly
may pose more risk than is acceptable. Critical
in cases of channel incision that undermines
to evaluating risk is how events and
bridge foundations, buried pipelines, or road
consequences are defined. For instance, the
embankments. Channel incision has been
100-year recurrence flood has a 1% probability
shown to result in the loss of floodplain
of occurring in any given year. The 100-year
connectivity and associated side channels and
flood is then associated with particular
wetlands. Geomorphic assessments can
consequences to have meaning, such as flooding
quantitatively define incision and the impact on

Large Wood National Manual July 2015


7-6
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

both habitat and infrastructure, and provide they will create potential hazards, so even
input on how wood can be used to treat the managing for passive restoration may need to
problem. address the benefits and risks of natural snags
and logjams to prevent their removal. Historic
For example, assume a geomorphic assessment
management incurred costs to remove snags.
found that there is a 90% probability that
Currently many parts of the country remove
100 acres of floodplain wetland will be
large quantities of wood after major storms and
disconnected and lost in the next 20 years if
floods, much of which could be left to provide
incision is allowed to proceed. In 20 years the
substantial ecological benefits with little risk.
incision will also expose a pipeline that would
Current management leaving wood may have
then need to be lowered at an estimated cost of
no costs associated with removal, but may
$1,500,000. Stabilizing the stream channel is
entail costs for public outreach and education.
estimated to cost $2,000,000 if done today. If
done in 20 years, the project will be more Not all wood placements are equal. Properly
challenging because of the deeper channel, engineered wood placements should not pose a
which together with predicted cost inflation is risk to downstream infrastructure because they
estimated to be at least $3,500,000, and not will be stable and act to trap mobile wood that
recover the wetlands. The estimated may otherwise put infrastructure at risk. Poorly
replacement or mitigation cost of the wetlands designed wood placements can pose a major
is $10,000/acre, adding an additional risk if the material were to plug a culvert and
$1,000,000 cost if the wetlands are not trigger a road washout. Conversely, properly
protected. The total cost of a no-action scenario engineered wood placements could lower risk
is $4,500,000—and $6,000,000 if the pipeline to the culvert by capturing mobile wood and
has to be lowered. This simple example clearly sediment prior to reaching the culvert. The
shows the value of stopping the incision as soon engineered wood placements will be more
as possible. expensive, but will lower risk. Because risk
increases at sites upstream of inadequately
In evaluating restoration risks, it is important to
designed infrastructure, so will project costs.
understand the ecological trajectory of the
Most culverts were never designed considering
project site. In some situations where riparian
sediment transport, much less wood transport.
forests are protected, a stream may gradually
Restoration should always consider upgrading
be restored under the no-action scenario. In
infrastructure as a critical element to achieving
these cases of “passive restoration,” the
restoration goals.
question to ask is how long will recovery take
and is that acceptable with regard to goals such
as restoring habitat for endangered fish. 7.3.1 Quantitative and
Recovery metrics must be defined, such as the Qualitative Risk
stream reaching a specified wood loading, a
volume per channel length, or number of
Assessment
functional pieces per channel length (e.g., Fox A risk assessment should be completed for
and Bolton 2007). Risk is evaluated by taking every large wood placement and restoration
the probability of not achieving the desired goal project, regardless of the size and scope. Risk
within a specified time, multiplied by the cost of assessments can either be quantitative or
placing that wood at that time. Like other risk qualitative depending on the level of
assessments, this requires estimating future background risk, acceptable limits of risk,
costs. available data, and project resources. Given the
subjective nature associated with key risk
With regard to human safety, even passive
elements, most often risk assessments begin
restoration has risks. As trees fall into rivers
Large Wood National Manual July 2015
7-7
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

qualitatively and proceed to a quantitative stage evaluated then multiplied by the consequences
if initial findings warrant further detailed of those events (impact of wood placement
assessments. A thorough risk assessment may moving), and summed for each alternative. The
even provide an economic study to assess the most common method used is to equate a
positive or negative monetary effects from a monetary value or loss to risk elements of
large wood project. By using a project-screening interest events, and then sum all risk elements
matrix developed by RiverRAT (Skidmore et al. for each alternative considered (e.g.,
2011), practitioners can evaluate the relative replacement cost of a bridge failing, property
level of thoroughness needed for a specific value loss due to an eroding bank, economic
project (i.e., a high response stream with a high impacts of the loss of a commercial fishery).
impact potential likely warrants a more Assessing the value or loss of any risk elements
thorough risk assessment than a low response can be very subjective, and a certain degree of
stream with a low-impact potential; see Figure objectivity should be used to provide
7-3). reasonable assessments. Niezgoda and Johnson
(2007) and Jones and Johnson (2015) provide
To complete a quantitative risk assessment, the
examples of cost-based risk assessments.
probability of certain events (e.g., a flood of
particular discharge, depth, and velocity
capable of moving wood placement) are

Figure 7-3. RiverRAT Screening Matrix

Source: Thorne et al. (2014b).

Large Wood National Manual July 2015


7-8
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

7.3.2 Elements of Risk expertise in fluvial geomorphology, bank and


hillslope stability, and hydrogeology; a biologist
Assessment with expertise in aquatic habitat and species;
Defining risk and identifying the appropriate and civil engineering professionals with
risk elements to consider is specific to each expertise in hydraulics, structural stability, and
project. However, the processes of assessing construction management. Because all of these
risk should not significantly vary between fields have a wide range of specific skills, each
projects. professional is mandated to practice within
their area of expertise or under the guidance of
someone who has that experience.
GUIDANCE
Consequently, anyone performing large wood
design should have education, knowledge, and
Main Components of a Risk Assessment experience in all the relevant fields.
1. Description of key risk concerns: (A) human
There are situations where wood is not
development (e.g., infrastructure, recreation,
and flooding) and (B) ecology (e.g., preventing appropriate; natural wood rarely occurs within
further habitat degradation, failing to achieve confined bedrock channels (canyons) because
restoration goals) wood is more easily transported to such
locations due to high flow depths and velocities,
2. Description of channel morphology and stability and channels with little resistance. Additionally,
(e.g., is channel incising or aggrading?)
because there is no bank erosion in bedrock
3. Description of existing or current fluvial and channels, there is little to no wood recruitment.
habitat conditions, including presence and type Where historic incision (due to human
of wood currently in the system disturbance) has transformed alluvial channels
4. Description of historic changes to the system to bedrock, wood may play an important role in
and how they influenced risk concerns (e.g., restoring an alluvial channel. Such sites will
channel clearing, recreational boating) involve greater risk and more intensive
engineering, but also represent important
5. Description of risks associated with the no- restoration opportunities. Other channel types
action alternative
where wood may not be appropriate are highly
6. Description of existing infrastructure such as confined urban channels with little or no
downstream bridges or culverts and how they tolerance for increasing water levels.
could be impacted by wood Understanding linkages between channel
7. Description for how the proposed large wood evolution and wood accumulations is a key
placements will affect these processes and key element of any risk assessment involving wood
risk elements placement or removal.

It is important to understand channel stability


A risk assessment should begin with a and the role wood will have on channel forming
description of the existing project site, key processes. Wood can be used to stabilize or
watershed processes, and adjacent fluvial destabilize banks. Restoration projects will fall
morphology (Table 7-1). Typically, this portion into three different categories regarding
of the assessment forms the basis for the no- channel stability.
action alternative. Common risk elements for
1. Dynamic Channel Corridors:
each component are shown in Table 7-2.
Elements of risk assessments should be a. Project sites where channel migration is
completed by licensed professionals in each a natural process, which wood
relevant science—typically a geologist with

Large Wood National Manual July 2015


7-9
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

placements are intended to The Trinity River in northwestern California is


accommodate. an example of a restoration program where
water withdrawals have dramatically reduced
b. Project sites where natural channel
peak flows and sediment supply. The altered
migration has been halted:
flow and sediment regime has resulted in an
1) Sites where channel migration does unnaturally stable channel devoid of the
not occur due to regulated flow and physical and temporal complexity that once
sediment regimes where wood is characterized the river. One goal of the Trinity
intended to help restore channel Restoration Program (2015) is to rehabilitate
migration. channel forming processes such as bar
2) Channelized reaches where development and bank erosion. Restoration
restoration includes setting back actions include using engineered logjams to
levees and revetments and wood is deflect flow and induce turbulence that will
intended to restore channel help trigger the morphologic response the river
migration. once had, where wood is being used to help
trigger channel migration in reaches where it
3) Sites where wood is intended to can be accommodated. Some reaches of the
locally stabilize banks to protect river underwent significant development
infrastructure or property within a because flow regulation so the restoration
dynamic reach. program must deal with different risk issues for
2. Restoration of Disturbed Channels that each project reach. Some sites must
were stable under their natural condition— demonstrate they will not adversely affect
many restoration projects have a goal of existing development, with regard to either
restoring straightened channels to their bank erosion or flooding. The Trinity River also
natural meandering planform. Wood is has recreational and commercial rafting and
often used to stabilize banks until mature fishing. Through public meetings with fishing
riparian vegetation is established that guides and rafting representatives and designs
naturally stabilizes the channel banks. that maintained navigable pathways, the
restoration program has received widespread
3. Stable or Constrained Channel Corridors— public support, and the ELJs have created
sites where wood placements are intended popular new fishing holes.
to stabilize banks and not trigger bank
erosion elsewhere.

Large Wood National Manual July 2015


7-10
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Table 7-1. Important Project Characteristics Defining Existing Conditions and Geomorphic Setting

Element Considerations
Project Goals Clearly state the project goals and the role wood placements have in achieving
those goals; use metrics to quantify goals if possible.
Project Site Where is the project area located? What are the main features of the project
area? What and where are constraints (e.g., levees, bridges, buried pipelines)?
Project Reach How do upstream and downstream reaches influence the project site?
Watershed Water, Are peak flows increasing? Is sediment supply expected to be relatively constant
Sediment, Wood (e.g., upstream dam or landsliding)? Development, logging, agriculture, and
Loading climate change can significantly alter these conditions and influence a project.
Stable Wood Are there stable wood accumulations in the project reach? If so, what are their
characteristics (size and shape of key pieces)? How much wood is enough; how
much is too much?
Mobile Wood How much wood is moving through the project reach, and how will it influence
the project? What will consequences be of wood accumulations within the
treatment reach?
Geology What are the characteristics of the riverbed and river banks? What is depth of
alluvium? What is bedrock material made of (e.g., glacial clay or hard rock?)
Habitat What are current habitat conditions? Is there high-value habitat in the project
reach that could be affected?
Channel Migration Is the channel actively migrating? If so, what are the average rates of migration?
Are there avulsion risks?
Channel Confinement Is the channel confined by levees, revetments, or incision?
Existing Large Wood What is the frequency and function of existing large wood in the project area?
Will the project significantly increase the frequency of large wood in the project
reach?
Floodplain Do flows frequently access the floodplain?
Connectivity
Riparian Condition What are the size, species, and distribution of trees in the project reach and
channel migration zone, and are they available for potential recruitment?
Historical Context Was large wood historically present? What were its likely effects?
Future Context Will large wood loading remain constant? How will climate change effect fluvial
habitat processes?
Channel Bed Material What is the size and gradation of channel material? Has an armor layer formed?
Channel Bank Material What is bank stability and resistance to erosion? Will wood placements trigger
bank erosion on either side of channel? (Refer to Simon and Collison 2002;
Simon et al. 2000.)

Large Wood National Manual July 2015


7-11
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Table 7-2. Important Elements for Consideration in Risk Assessment

Element Considerations
Infrastructure Are there bridges or culverts downstream of the project area? Do they have in-
channel piers? What is their ability to convey large wood?
Are there levees or revetments adjacent to or downstream of the project area?
What is their condition? Were they designed to withstand extreme floods?
Are there buried utilities in the project area? How deep are they buried? If the
channel avulses or migrates are they likely to be exposed?
Are there public or private roads within the adjacent floodplain? If so, are they
overtopped frequently and by how much flow depth?
Property Is the adjacent floodplain public or private property? How will large wood
placements affect flood depths on adjacent properties?
Where is the project area located in relation to property boundaries?
What structures (houses, outbuildings, recreational facilities) exist within or
downstream of the project area?
Is the channel actively eroding or migrating? How will large wood placements
affect erosion and migration rates? Would channel migration into adjacent
properties be perceived negatively?
Are there avulsion pathways through adjacent properties? How will large wood
placements affect the likelihood of a major channel avulsion? Would a major
avulsion through adjacent properties be perceived negatively?
Habitat What will happen if no project is completed? Will habitat conditions for the
species of interest improve or decline?
How will large wood placements affect habitat conditions in the short (1 to
5 years) to long term (5 to 50 years)?
Will there be temporary impacts during the construction process? Will those
create any permitting issues?
How will large wood placements affect future large wood recruitment?
Public Safety Would failure of infrastructure (described above) cause a threat to human safety
or welfare?
Would erosion, channel migration, or avulsion (described above) cause a threat
to human safety or welfare?
Does the reach experience recreational use? If so, what is the experience level of
the normal user? Are most users accustomed to large wood hazards?
Construction How does the local regulatory environment view large wood installations? Will
local policies and/or viewpoints affect how the large wood placements are
located and constructed?
How will the large wood placements be constructed? How will sediment and
turbidity be minimized?
Will de-watering be required? If so, is a de-watering plan feasible? What are the
contingencies if the plan’s de-watering method proves to be infeasible?
When will the large wood placements be constructed? Is there a risk of high
flows during the construction window? If so, what would the consequences be?
Can a flood event (e.g., summer rainstorm) pose a threat to construction? What
is the probability and how can risk be minimized?

Large Wood National Manual July 2015


7-12
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Element Considerations
Is there a regulatory “fish-window” or timeframe the project will need to be
constructed within? If so, is that timeframe sufficient to complete construction
for all elements?
Will the construction methods generate significant noise that will affect nesting
birds or wildlife, particularly threatened or endangered species?
Is buried wood expected within the excavation area during pile driving? If so,
what is the plan or contingencies for how to handle?
Is bedrock expected in the excavation area during pile driving? If so, what is the
plan or contingencies for how to handle this?
What level of design is being developed for the large wood placement? Has the
contractor built large wood placements? How will change-orders be handled
during the construction process?
How will the contractor access the site and are there constraints on that access
posed by landowners, length of access route, traffic control, wetlands, stream
channels, or soft soils?

7.3.3 Professional Liability


In many regions where large wood placements GUIDANCE
are installed, property owners, designers,
contractors, sponsors, counties, and regulatory
Key Tenets of Recent Washington Legislation
agencies are professionally liable for damages
caused by these placements (Andrus and 1. The project was designed by licensed
Gessford 2007). To minimize liability associated professional engineers and geologists
with damages related to large wood experienced in riverine restoration.
placements, practitioners are advised to 2. The project was designed to withstand the 100-
conduct rigorous and defensible analyses of the year flood.
risks associated with each project (Tonglao and
Eckberg 2012). This analysis should include 3. The project is not located within 0.40 kilometer
hydraulic and geomorphic information that will (0.25 mile) of an established boat launch.
evaluate how the large wood placement would 4. The project is designed to allow adequate
affect flow patterns, and verify that the response time for recreational users to safely
predicted changes are not likely to result in evade large wood placements.
significant damages to adjacent property
5. Large wood placements larger than 3 meters (10
owners (Tonglao and Eckberg 2012). In feet) long and 0.3 meter (1 foot) in diameter
practice, liability is assumed by the design include tagging of individual pieces that will last
professionals while sponsors and property for at least 3 years.
owners often pursue indemnification
agreements to protect themselves. Washington
While these requirements represent specific
and Oregon have passed legislation granting
guidelines for the State of Washington, it is
immunity to private property owners to remove
reasonable to expect similar legislation in other
barriers to aquatic habitat restoration.
states as the design and implementation of
wood-based restoration activities becomes
more frequent.

Large Wood National Manual July 2015


7-13
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Locating and designing large wood placements machines, equipment, processes, works, or
is a multidisciplinary exercise that requires projects.
involvement of trained professionals that Other professionals, such as geologists,
include professional engineers, licensed maintain similar definitions and guidelines. In
geologists, fisheries scientists, and wetland/ Washington, any analysis report describing
riparian scientists to ensure long-term success surface and subsurface water flow and earth
(Tonglao and Eckberg 2012). Project sponsors materials is supposed to be stamped by a
and regulatory agencies are encouraged to professional licensed geologist. Geology
require stamped and signed plans from every specialty licenses include engineering geology
(e.g., rock and soil mechanics, hillslope stability,
key discipline involved as part of the review
and stabilization of excavated areas) and
process. hydrogeology (e.g., ground and surface water
The design of large wood placements should, at modeling, solute transport, water quality).
a minimum, include licensed professionals and
scientists with river and wood expertise. The 7.3.4 Defining Risk on Your
final design package (plans, specifications, and Project
estimates) should be stamped by a professional
geologist and engineer. The geologist ensures Each risk assessment will be unique to each
that designs have taken into account an project given the historical context, restoration
understanding of site conditions, geomorphic goals, site constraints, recreational use, and
processes, and responses. The engineer certifies public concerns. However, as noted above, the
that designs have the desired stability, are assessment process should not vary
buildable, provide sufficient detail for the significantly between projects, though the scale
contractor, and include inspection criteria for of the efforts may be quite different. A risk
ensuring the project is constructed per the assessment is only finalized once a project is
design. Due to the charge of civil engineers to completed and deemed to be functioning
“to use their knowledge and skill for the properly. The critical stages for the risk
enhancement of human welfare and the evaluation are completed at critical stages
environment” and “engineers shall hold during the lifecycle for the project. An early
paramount the safety, health and welfare of the evaluation is completed during the project
public…in the performance of their professional concept phase with details on a variety of
duties” (Tonglao and Eckberg 2012) they are analyses performed during the project
often, and in some circumstances, required development phase, while a final evaluation is
(SRFB 2013) to be responsible for the design of completed during the post-project review and
large wood placements. Furthermore the monitoring phase. Each phase of the evaluation
development of large wood placement designs process is described below.
generally falls into the standard definition of the
“practice of engineering” as follows. 7.3.4.1 Project Conception
Practice of engineering means any professional This phase of risk assessment begins with
service of creative work requiring engineering engaging local and regional stakeholders at the
education, training, and experience and the onset of the project to provide an opportunity
application of special knowledge of
for input. Initiating the process with
mathematical, physical, and engineering
services to such professional services or stakeholder engagement reduces the chances
creative work as consultation, investigation, for costly changes near the end of the design
evaluation, planning, design, and supervision of process, and it also engages the community in a
construction for the assuring compliance with way that engenders support for large wood
specifications, in connection with any public or projects. Following public input and developing
private utilities, structures, buildings,

Large Wood National Manual July 2015


7-14
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

extensive site knowledge, project sponsors and punch-list is simply a checklist of important
designers should develop a list of key risk actions that can be clearly accounted for with
elements (a partial list is included in Table 7-2) regard to when they were completed, by whom,
to consider during the project development and and as intended. During the development of the
post-project phases. punch-list items, each large wood placement
should be inspected and evaluated for
7.3.4.2 Project Development compliance with the final plans and
specifications, while any deviations that create
Risk elements identified during the project a high-risk situation should be addressed before
conception phase should be considered in the the contractor demobilizes from the site.
project development phase. Alternatives should
Following the post-construction walk through,
be developed that consider the geomorphic and large wood placements should be inspected on
habitat restoration goals, and how each an annual, multiyear timeframe to ensure the
alternative could affect the identified risk
structures are performing as intended and a
elements. Effort should also be made to
high-risk situation has not developed. If during
minimize risk, while maintaining intended periodic monitoring a high-risk situation is
geomorphic and habitat benefits to the observed, sponsors should consider an adaptive
maximum extent possible. Following
management protocol to reduce risk and
development of alternatives, an analysis of improve public safety. Situations that could
hydraulic, hydrologic, scour, and stability warrant high risk could include natural large
factors can be performed to evaluate the effects
wood that has racked on large wood placements
on geomorphic and habitat processes, flooding, creating a strainer condition or a channel
erosion, and sediment transport. This spanning logjam that increases unacceptable
information will aid in evaluating the relative
flooding or erosion.
risks and benefits of each alternative. At the end
of the project development phase, the results of
the risk assessment should be documented and CROSS-REFERENCE
presented to the local and regional
stakeholders. Chapter 9, Assessing Ecological Performance,
describes the adaptive management process in
7.3.4.3 Post-Project Monitoring detail.

During the development of a large wood


project, there is potential to encounter
difficulties requiring adjustments to the design
7.3.4.4 Special Considerations for
elements, schedule, and funding requirements. Recreational Users
During the construction phase most challenges Safety attributes of ELJ and large wood
arise due to unforeseen site conditions, short placements specific to recreational users can be
construction timeframes, materials that do not divided into two categories: reach and
meet specifications, and inexperienced structure-specific assessments. Reach
contractors. The purpose of this phase of the categories include definition of the recreational
risk assessment is to ensure that the key use, access, and reach-scale geomorphic factors.
assumptions and design elements were Structure-specific categories include structure
completed and the large wood placements are location, structure type and characteristics, and
performing as intended. The majority of the risk avoidance potential of each specific structure
assessment for this phase can be completed (i.e., line of sight distance, path around
during the post-construction punch-list with the structure, including portage, and response
contractor and project sponsor. A time). These categories, as they relate to public

Large Wood National Manual July 2015


7-15
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

safety and the engineered placement of large of large wood or ELJ placements, it is important
wood, are further explained in the following to note a low-risk structure can be placed in a
section. high-risk environment creating a hazardous
scenario, and a high-risk structure can be
The American Whitewater Association (2012)
placed in a low-risk environment and not
suggests assessing how individual wood
significantly affect the safety of recreational
accumulations function both ecologically and as
users.
a hazard to recreational boaters, suggesting that
hazardous wood is typically just a fraction of
the total wood loading (Figure 7-4). 7.3.5 Reach Factors
Figure 7-4. Relative Quantity of Wood Within a 7.3.5.1 Recreational Use
Reach, the Subset with High Geomorphic and
Habitat Benefits, and the Subset that Causes An important consideration when assessing
Public Safety Concerns public safety impacts on recreational users
associated with large wood placements is
determining the various recreational uses, the
primary use period, frequency of use, and the
general skill level of the primary user group.
Most recreational rivers in North America
experience seasonal use based on weather and
flows. Summer is typically the highest use time
and may correspond to relatively high flows
(particularly in regulated rivers with irrigation
flow releases) or low flows. Expert whitewater
enthusiasts can be an exception; their most
A. All wood in stream reach. intensive use is typically during periods of high
B. Portion of wood causing high-risk flow, such as fall and winter in the Pacific
recreational hazards in stream reach. Northwest. While recreational use in some form
C. Portion of wood that is both ecologically most
is possible on most rivers in North America, not
functional and causes high-risk recreational
hazards. all rivers experience a high frequency of use.
D. Portion of ecologically most functional pieces The flow range occurring during the majority of
of wood in stream reach. that use period is also important and is defined
in the assessment as the recreational flow range.
Source: American Whitewater Association (2012).
When considering recreational use categories,
there are often outliers or extremes to many of
A more detailed risk matrix was developed in the categories described. When performing a
Washington State that considers both structure recreational risk assessment it is recommended
and reach characteristics (Figure 7-5). Similar to focus on the majority or typical value for the
risk matrices could be created by expanding or specific category and omit outliers or extremes.
considering different risk elements specific to a
project. Risk matrices are not recommended to
evaluate a precise risk level but for evaluating
the general effects relative to the no-action and
project alternatives. When evaluating the safety

Large Wood National Manual July 2015


7-16
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Figure 7-5. Risk Assessment Chart

Source: Embertson and Monahan (2011).

Large Wood National Manual July 2015


7-17
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

frequency of use in the immediate area. When


The skill level of recreational users is an
determining normal access points it is
important consideration when completing a
important to note that the type of access varies
reach assessment. Large wood in the river
for different types of recreational use. More
environment is very common in certain
sophisticated, experienced user groups can
physiographic regions (albeit significantly more
often use road and highway pullouts with steep
rare than historic norms), and avid recreational
wooded banks and relatively little to no calm
users may be accustomed to dealing with
water along the channel bank. Less-experienced
hazards associated with large wood (Figure 7-
recreationalists require larger access points and
6). For instance, expert and advanced user
boat ramps to the water edge to aid in carrying
groups will generally not be challenged
heavy equipment and loads. Consulting with
navigating safely around large wood given their
known private or commercial users groups is
experience with naturally occurring wood.
often the easiest way to locate common
However, safely avoiding ELJs or large wood
recreational access points and to assess the
may be more difficult for beginner-to-
experience level and frequency of access by
intermediate user groups. Therefore, skill level
different user groups.
and frequency of use are important factors to
consider because structures placed in reaches
frequented by beginner to intermediate users
7.3.5.3 Geomorphic Factors
will pose a greater risk to those users than Reach-scale characteristics described in a
structures placed in reaches frequented by geomorphic risk assessment can also aid in
expert users. assessing recreation-based public safety. Wood
Figure 7-6. Natural Wood Accumulation in Idaho naturally plays a role in many channel types
throughout a drainage network (e.g., Keller and
Swanson 1979; Hickin 1984; Triska 1984; Abbe
and Montgomery 2003; Montgomery et al 2003;
Wallerstein and Thorne 2004). A fluvial
geomorphic analysis should compile a spatial
and temporal database that includes: valley
morphology, channel planform and
confinement, bed (surface and subsurface) and
bank materials, sediment supply, channel
gradient, riparian conditions, artificial
structures (e.g., levees, revetments, diversions,
weirs, dams, bridges), pool frequency, rates of
Source: American Whitewater Association (2012). channel migration, evidence of channel incision
or aggradation, flood frequency, flow regime,
7.3.5.2 Access flow depths and velocities, and estimates of
natural wood quantities (refer to Chapter 4,
The ability to access a given reach can
Geomorphology and Hydrology Considerations).
significantly influence many of the recreational
factors discussed above. Reaches with poor The valley type within a reach can help identify
access will generally have a low frequency of safety issues that could arise associated with
use and are well suited as locations for the instream wood. For instance, large wood in a
placement of ELJs and large wood to maximize confined bedrock canyon would pose a greater
habitat enhancement. Locating large wood and risk to recreational users than placements in a
ELJ structures near known access points can broad alluvial valley where a user would likely
also be considered a higher risk due to the

Large Wood National Manual July 2015


7-18
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

be able to portage (get out and walk around) The channel stability of a given reach is also an
wood that posed a possible hazard. important geomorphic reach characteristic.
Most structures placed in the river environment
Channel types can be used to evaluate potential
are located such that they do not pose a
risk by providing information on hydraulics,
significant safety hazard following construction.
bed material, and the influence of wood.
However, if the river channel migrates or
Montgomery and Buffington (1997) provide a
dramatically changes position, a significant
process-based classification of channel types
safety hazard could result due to changes in the
found within a channel drainage network
channel location, flow direction, and potential
primarily based on gradient and confinement.
accumulation of large wood on the ELJ or other
Both of these factors can help identify the
structure. The likelihood of this occurring in a
degree to which a recreational user might be
dynamic and active reach is higher than in a less
challenged to navigate safely through a given
dynamic, slow-reacting system. ELJs or large
reach and the hazards of wood accumulations.
wood placements in a dynamic geomorphic
Large wood placements in a reach with a
reach should be considered higher risk than
confined bedrock, cascade, or step pool
those located in a slow-reactive system.
morphology (Montgomery and Buffington
1997) should be considered higher risk because Large wood is very common in river
these channel types are inherently more environments within certain physiographic
difficult and dangerous than other channel regions in North America (Figure 7-4). Almost
types due to the flow velocities, transitions any stream flowing within forested banks will
between subcritical and supercritical flow, and have wood inputs. Because bank erosion is a
channel confinement that makes it difficult to major recruitment mechanism, forest channels
reach safe ground. In contrast, placements in a with alluvial bank heights greater than root
reach with pool-riffle or plane-bed morphology depth of trees typically have large quantities of
are generally lower risk because these channel wood. In rivers and streams where there are
types are easier for recreational users to existing accumulations of large wood,
navigate or portage in order to avoid large recreational users are generally aware of the
wood placements. Any wood placements in inherent risk in that area. The addition of large
rivers with recreational boaters should wood or ELJ placements as part of habitat
maintain a sufficient portion of the channel for improvement projects should clearly
safe navigation. demonstrate the scientific justification of
placements and how they could affect
The average channel gradient within a reach
recreational users. Education and signage can
can both help identify the inherent difficulty for
help to mitigate the recreational effects of the
a recreational user and estimate the relative
placement of wood. In some cases it may be
speed with which a recreational user
wise to close the river to recreational use. In
approaches the large wood or ELJ placement. A
cases where recreational areas will be
steep-gradient reach should be considered
maintained, wood placement should both mimic
higher risk than a low-gradient reach for similar
natural wood accumulations and maintain a
reasons as the channel type described above. A
navigable path for boaters. Placing structures in
steep-gradient reach generally has a high
sites downstream of natural large wood reaches
approach velocity, reducing the reaction time of
should be considered lower risk if boaters have
a recreational user to large wood or ELJ
to navigate those reaches before encountering a
placements. Consequently, high-gradient
restoration reach.
reaches should be considered higher risk than
lower-gradient reaches. To estimate natural wood loading, it is common
to use reference conditions determined by Fox

Large Wood National Manual July 2015


7-19
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

and Bolton (2007). As part of a study of over As flow moves through a channel bend, floating
150 stream segments, unmanaged basins were objects will tend to move toward the outside of
surveyed for wood quantities and volumes in a bend. Recreational users navigating through a
Washington State. The results were segregated channel bend have a harder time avoiding
into bankfull width classes, forest zones, and structures placed along the outside of a sharp
percentile of classes listed. The most important channel bend than structures placed along the
thing to consider for wood and boaters is inside of a broad channel meander. But if there
whether flow goes around the wood is sufficient roughness (e.g., wood) along the
(“deflector”) or through it (“strainer”). Wood outside of a bend it will introduce strong
design in recreational rivers should focus on secondary vortices along the bank that not only
deflector structures, which are easy to navigate dramatically reduce near-bank velocities, but
because flow goes around them. They also tend effectively push the thalweg away from the
to create downstream eddies that provide a safe bank (Blanckaert et al. 2010; Konsoer 2013).
refuge or pull-out point for boaters. Structures placed along the outside of a channel
bend should be considered higher risk to
7.3.6 Large Wood Structure recreational users than structures placed in a
linear reach or on the inside of a channel bend.
Factors The degree to which a recreational user is
influenced by local hydraulic patterns is a
7.3.6.1 Structure Location function of the user’s maneuverability. For
The location of large wood and ELJ structures in instance, a recreational “tuber” has a low
a stream channel and floodplain is an important maneuverability and, therefore, would be more
consideration when assessing the recreational influenced by hydraulic patterns (and at a
safety of a particular structure. The primary higher risk) than an intermediate to advanced
consideration related to the location of whitewater kayaker who is more
structures is the amount of engagement of the maneuverable.
structure with the wetted channel during the
expected recreational flow range, and whether 7.3.6.2 Structure Characteristics
the structure is located along the outside of a The characteristics of different large wood and
channel bend. ELJ structures have varying degrees of risk to
The more a structure is engaged in the wetted recreational users (Table 7-3). Structure
channel, the more likely it is that the structure characteristics of most concern are those that
poses a risk to the safety of recreational users. create a “strainer” condition that could trap a
Structures that are not engaged in the wetted person or boat. A strainer condition occurs
channel during the expected recreational flow when a piece or pieces of large wood in a
range pose a much lower risk to the safety of structure allow water to pass under, over, or
recreational users. For instance, many large ELJ through the piece or pieces. The force of the
structures are often constructed on dry gravel moving water through the strainer can trap or
bars and out of the low-flow channel due to pin a person or their recreational craft against
permitting and constructability constraints and the large wood and create a dangerous scenario
may not significantly engage with the channel (i.e., potential drowning). The most common
during the recreational flow range. These strainer condition is a single piece of large
structures pose a much lower risk to wood that extends out perpendicular to the
recreational users than structures constructed channel bank and direction of flow, at or below
in the wetted channel that are fully engaged the water surface.
with the low-flow channel.

Large Wood National Manual July 2015


7-20
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Large wood placed in a rootwad bank location in the channel, approach line of sight
protection method can commonly form a distance, flow velocity, angle of flow deflection,
strainer condition if scour and channel and proximity to calm water (e.g., eddies) to
migration is not considered during the design rest or pull out. Ratings are influenced by
and placement process. A strainer condition is location along the outside of a channel bend
not as common for ELJs but can occur if the (higher risk), sight distance (often poor), and
structure is not backfilled with alluvium to tendency to create a sweeper/ strainer
prevent flow through the structure, if individual condition (higher risk).
log pieces extend out beyond the general limits
of the structure, or if the structure shifts and
Avoidance Potential
unravels over time. While a strainer can create If recreational users can safely avoid large wood
a dangerous condition for recreational users, or ELJ structures by either portaging around the
strainers can also increase channel complexity, structure or paddling well away from the
cover, and habitat variability all of which are structure, the relative risk of that structure is
beneficial for many types of aquatic. lower than if portaging or paddling away from
the structure is difficult (Figure 7-7). Key
Large wood and ELJ placements designed to
factors when considering avoidance potential
emulate natural wood assemblages create flow
are egress potential, sight distance, approach
hydraulics that are more familiar to
velocity, and the combined values of depth and
recreational users and pose a lower risk than
velocity at the approach to the structure (depth
nonnatural structure types (e.g., log crib wall,
and velocity product).
tethered log structures). Abbe et al. (2003a)
classified instream woody debris accumulations Figure 7-7. Egress and Portage
observed on the Queets River in three distinct
types: grade control, revetment, and flow
deflection. A summary of the different types,
brief descriptions, and relative recreational
risks is provided in Table 7-3.

For the purpose of this assessment, a subjective


relative risk rating is provided for each
structure type, based on the intended function
of the structure (Table 7-3). In this assessment,
the only structure qualifying for a low rating
Source: American Whitewater (2012).
was a step-type structure, because its design
standard requires a high level of embeddedness The egress (exit) potential of a structure can be
in the channel bed and also provides the low- defined as the ability of a recreational user to
risk flow profile over the structure. Valley-type exit the channel upstream of the structure in
structures receive a high-risk rating due to their order to walk around the structure. An egress
size, the chaotic assemblage of woody material point is a specific location where a recreational
within each structure, and the presence of flow user could exit the channel upstream of the
through the structure. Bar apex structures may large wood or ELJ. Conversely, an ingress point
be assigned a low to moderate rating in cases is a location where a recreational user could
where there is a clear navigational path around enter or re-enter the channel. Steep bedrock
them and a higher risk if navigation is unclear canyons or an incised channel with steep banks
or obstructed. Variability in rating should focus generally have poor egress potential. Broad
on ease of navigation around the structure, and alluvial valleys with frequent gravel mid-
take into account such things as structure channel and point bars generally have good
egress potential.
Large Wood National Manual July 2015
7-21
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Table 7-3. Relative Risk of Instream Wood to Recreational River Users


Type Description Relative Recreation Risk
Step Single log structure spanning the channel Low to moderate in small
width and forming a scour/plunge pool streams where boating is
immediately downstream. Flow generally uncommon or there is
proceeds over the structure. adequate submergence to
eliminate vertical drop and
recirculating plunge pool. A
large drop and standing
wave increases risk.
Valley Multiple log structure with a width High
greater than the bankfull width and
accompanying a significant portion of the
valley width. Flow through and over the
structure.
Bankfull Bench Multiple log structure located along the Moderate
outside of channel bend, with a width less
than the bankfull width, and creating a
bench surface. Flow generally proceeds
along the structure.
Flow Deflection Multiple log structure located along the Low to high, depending on
outside of channel bend, with a width less how much of the channel
than the bankfull width that accumulates width is obstructed and
wood over time. Flow generally response time (line of sight
approaches normal to the structure and is and velocity).
then deflected away at a moderate to
severe angle via parallel log members.
Bar Apex Multiple log structure located at the head Low to moderate where the
of mid-channel bar, with a width less than navigational path is around
the bankfull width, creating a stable the structure.
depositional zone downstream. Flow Moderate to high where the
generally approaches normal to the navigational path around the
structure and is then deflected away at a structure is unclear or
small to moderate angle. obstructed.
Meander Multiple log structure located along the Moderate to high
outside of channel bend, with a width less
than the bankfull width, and creating a
bench surface. Flow generally proceeds
along the structure.

Large Wood National Manual July 2015


7-22
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

The sight distance of a structure can be defined where:


as the maximum distance from which a
V = Flow Velocity (meters/second)
recreational user will be able to see the
structure when approaching along the thalweg D = Flow Depth (meters)
of the channel. The lower the sight distance, the
less time a recreational user will have to Note that the drag force acting on a wader will
develop a plan for how to avoid the structure be proportional to the wader’s submerged area
and react appropriately. Structures with more normal to flow.
sight distance are safer than structures with
less sight distance. Because many ELJ structures 7.4 Bridges and Culverts
are constructed to be equal to or above the peak
flood water surface, the effects of the channel Mobile wood moving down streams can be an
profile can generally be ignored. However, in issue with human-made crossings that constrict
certain circumstances the effect of the channel the channel, such as small bridge spans or
profile may be an important consideration culverts. A risk assessment should always
when determining the available sight distance. include a description of downstream crossings
and whether the project would increase risk
The mathematical product of flow depth, D, and exposure. If the wood placements are stable and
velocity, V, referred to as the wading safety designed to catch mobile wood debris, then the
factor, is used to evaluate the potential for a restoration will reduce risk to downstream
recreational user to walk away or “self-save” crossings. Efforts to restore wood by putting in
within a stream. Researchers at Colorado State mobile pieces should be carefully thought
University conducted flume tests to identify the through because this process could elevate
depth and velocity of flow at which a person downstream risks. Well-designed large wood
could safely maneuver and stand in moving placements can help downstream infrastructure
water. Subjects participating in the study (Abbe et al. 2003a; Abbe and Brooks 2011).
became unstable at product numbers that Whenever possible, channel crossings (culverts
ranged from 8–23 square meters per second. or bridges) should be used to accommodate the
Results were found to be dependent on body transport of wood material (e.g., Cafferata et al.
stance, position, and type (Abt et al. 1989). 2004; Flanagan 2004, 2005). There has been
Given the broad range of results, the study extensive research regarding the risk wood
concluded safe wading conditions should not material poses at bridge crossings, with specific
exceed a product of 10. To evaluate the wading concern for conveyance and bed scour (e.g.,
safety factor specific to large wood or ELJ Diehl 1997; Lagasse et al. 2010). Recent flume
structures, the expected flow depth and velocity research found that scour risk is reduced if the
upstream from field observations or hydraulic wood accumulation on a bridge pier extends the
modeling is commonly utilized. Flow depth and full depth of flow, while it increases if there is
velocity should be determined upstream of the flow beneath the wood (Lagasse et al. 2010).
structure (about 8–15 meters [26–49 feet]; Abt
et al. 1989) and outside the expected scour hole
CROSS-REFERENCE
influence. The wading safety factor (depth and
velocity product) is defined as follows.
Additional information on the effects associated
𝑊𝑎𝑑𝑖𝑛𝑔𝑆𝑎𝑓𝑒𝑡𝑦𝐹𝑎𝑐𝑡𝑜𝑟 = 𝑉 ∗ 𝐷 with bridges and climate change can be found in
Chapter 5, Watershed-Scale and Long-Term
Considerations.

Large Wood National Manual July 2015


7-23
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

7.5 Uncertainties and Research Needs


1. Regional data is needed regarding the ecosystem role of wood and impacts of wood removal.
2. Regional data on existing wood loading, specifically the location, size, and mobility of large wood
pieces, is needed.
3. Predictive models need to be designed to identify wood transport and deposition reaches in
stream systems.
4. Outreach should be conducted to educate residents, recreational users, flood districts, and
public works departments on the value of instream wood and best practices for wood placement
and stream management. Awareness and education on wood is one of most important means of
reducing risks over time.
5. Guidelines should be developed for culvert design with respect to sediment and wood
conveyance.
6. Guidelines should be developed for wood management (what to leave or remove) following
major storms or floods.
7. Regional legal guidelines are needed with respect to the liability of wood placement in, or its
removal from, streams.
8. Hydraulic models need to be developed that show the influence of instream wood loading on
flood stage and discharge throughout a channel network.
9. Predictive models should be developed to evaluate how a wood placement may influence the
accumulation of wood and its impacts.
10. Regional data should be compiled on channel incision attributed to removal of instream wood
removal.
11. Regional and species data need to be compiled on wood longevity (decay) and its influence on
wood stability and function.
12. Regional models are needed that predict the time wood projects must last to establish mature
riparian forests and self-sustaining conditions.
13. Guidelines should be developed for establishing geomorphic corridors needed to sustain
physical and ecological processes and minimize risks to human development.

7.6 Key Points


1. Risk is the product of the probability of an event occurring times the consequences the event
will have with regards to impacts on habitat, public safety or property.
2. Snags and logjams are natural elements of streams throughout the United States and thus
represent an inherent risk that recreational users accept when entering the system.
3. For over a century, local municipalities and the federal government have cleared streams of
wood and riparian trees to foster navigation and flood conveyance. Although channel clearing
has been scaled back in many systems, it has left many people with a simplified perception of

Large Wood National Manual July 2015


7-24
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

streams that is not consistent with natural conditions and an expectation of channel
maintenance that is unrealistic both economically and environmentally.
4. The loss of instream wood has led to major geomorphic changes and severe ecological impacts
in most streams throughout the country. Failure to rehabilitate wood loading and the functions
it provided poses a serious risk to further ecosystem degradation.
5. Clearing wood from streams can result in channel incision that increases the risk to
infrastructure such as buried pipeline crossing, bridge piers and abutments, road embankments,
and water intakes.
6. Instream wood and riparian vegetation diffuse flood peaks and lower flood peaks downstream;
therefore, large-scale stream clearing increases the risk of downstream flood discharge and
staging.
7. When municipalities take direct actions to clear wood from streams, their liability may be
increased if the practice is not sustained.
8. Wood accumulations also can raise water elevations, which can increase the frequency and
magnitude of overbank inundation. This provides very beneficial ecosystem services but can be
problematic in areas where development has encroached into flood-prone areas.
9. Riparian forests are integral to restoring and sustaining wood to stream ecosystems, yet
estimates of human impacts on riparian areas range from over 50% (Swift 1984) to 95%
(Brinson et al. 1981). This further underscores the risk posed by no-action alternatives.
10. Most stream crossings (i.e., culverts and bridges) have not been designed to accommodate the
passage of wood material. Actions that increase wood flux into inadequate crossings will
increase the risk of blockages that could compromise the facilities or increase upstream
flooding. Risk can be addressed by upgrading infrastructure to accommodate wood material.
11. In streams with high recreational boating usage care should be taken to ensure engineered
wood placements do not create strainers and that there is sufficient line of sight and response
time to provide navigable passage.
12. Wood placement projects in rivers with recreational users should include public engagement
and education, particularly with local communities, emergency service providers (e.g., fire and
law enforcement departments with search and rescue teams), and river user groups (e.g., rafting
companies, fishing guides).

7.7 References
Abbe, T. B., and A. P. Brooks. 2011. Geomorphic, Engineering, and Ecological Considerations when
Using Wood in River Restoration. Pages 419–451 in A. Simon, S. J. Bennett, and J. M. Castro
(eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools.
Geophysical Monograph Series 194. Washington, D.C.: American Geophysical Union.

Abbe, T. B., and D. R. Montgomery. 2003. Patterns and Processes of Wood Debris Accumulation in
the Queets River Basin, Washington. Geomorphology 51:81–107.

Abbe, T. B., D. R. Montgomery, K. Fetherston, and E. M. McClure. 1993. A Process-Based Classification


of Woody Debris in a Fluvial Network: Preliminary Analysis of the Queets River, Washington.
EOS, American Geophysical Union Transactions 73(43):296.

Large Wood National Manual July 2015


7-25
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Abbe, T. B., J. Carrasquero, M. McBride, A. Ritchie, M. McHenry, and K. Dublanica. 2003a.


Rehabilitating River Valley Ecosystems: Examples of Public, Private, and First Nation Cooperation
in Western Washington. Proceedings of the Georgia Basin/Puget Sound 2003 Research
Conference, Vancouver, B.C., March 31–April 1, 2003, T. Droscher (ed.). Puget Sound Action
Team, Olympia, WA.

Abt, S. R., R. J. Wittler, A. Taylor, and D. J. Love. 1989. Human Stability in a High Flood Hazard Zone.
American Water Resources Association. Water Resources Bulletin 25(4):881–889.

American Whitewater Association. 2012. Integrating Recreational Boating Considerations into


Stream Channel Modificatin & Design Projects. Written by Kevin Colburn, National Stewardship
Director. Illustrations by Chad Lewis. Figure 8.4, page 13. Available:
http://www.americanwhitewater.org/content/Document/fetch/documentid/1006/.raw.

Anderson, D. B. 2006. Quantifying the Interaction between Riparian Vegetation and Flooding: from
Cross-Section to Catchment Scale. University of Melbourne.

Andrus, B., and J. Gessford. 2007. Understanding the Legal Risks Associated with the Design and
Construction of Engineered Logjams. Skellenger Bender Attorneys. Seattle, WA.

Beechie, T. J., H. Imaki, J. Greene, A. Wade, H. Wu, G. Pess, P. Roni, J. Kimball, J. Stanford, P. Kiffney,
and N Mantua. 2012. Restoring Salmon Habitat for a Changing Climate. River Research and
Applications 29:939–960.

Blanckaert, K. A. Duarte, and A. J. Schleiss. 2010. Influence of Shallowness, Bank Inclination and Bank
Roughness on the Variability of Flow Patterns and Boundary Shear Stress due to Secondary
Currents in Straight Open-Channels. Advances in Water Resources 33(9):1062–1074.

Booth, D. 1991. Urbanization and the Natural Drainage System: Impacts, Solutions, and Prognoses.
The Northwest Environmental Journal 7, 93-118.

Brinson, M. M., B. L. Swift, R. C. Plantico, and J. S. Barclay. 1981. Riparian ecosystems: Their Ecology
and Status. FWS/OBS-81/17. Office of Biological Services. United States Department of the
Interior Fish and Wildlife Service., Washington D.C.

Cafferata, P., T. Spittler, M. Wopat, G. Bundros, and S. Flanagan. 2004. Designing Watercourse
Crossings for Passage of 100-Year Flood Flows, Wood, and Sediment. California Forestry Report
No.1. California Department of Forestry and Fire Protection. Available:
http://www.fire.ca.gov/php/rsrc-mgt_forestpractice_pubsmemo.php.

Cannon, S. H., and J. DeGraff. 2009. The Increasing Wildfire and Post-Fire Debris-Flow Threat in
Western USA, and Implications for Consequences of Climate Change. Pages 177–190 in K. Sassa
and P. Canuti (eds.), Landslides—Disaster Risk Reduction. Berlin Heidelberg: Springer-Verlag.
Available: <http://landslides.usgs.gov/docs/cannon/Cannon_Degraff_2008_ Springer.pdf>.

Cannon, S. H., J. E. Gartner, M. G. Rupert, J. A. Michael, A. H. Rea, and C. Parrett. 2010. Predicting the
Probability and Volume of Postwildfire Debris Flows in the Intermountain Western United
States. Geological Society of America Bulletin 122(1-2):127–144. doi:10.1130/B26459.1.

Colburn, K. 2011. Integrating Recreational Boating Considerations into Stream Channel Modification
and Design Projects. American Whitewater (2011).

Large Wood National Manual July 2015


7-26
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Diehl, T. H. 1997. Potential Drift Accumulation at Bridges. Publication FHWA-RD-97-028. U.S.


Department of Transportation, McLean, VA.

Embertson, L, and J. Monahan. 2011. Public Safety Assessment of Habitat Enhancement Projects Fobes
and Skookum Reach Restoration Projects South Fork Nooksack River. GeoEngineers, Bellingham
Washinghton. March 1, 2011.

Fox, M. J. and S. Bolton. 2007. A Regional and Geomorphic Reference for Quantities and Volumes of
Instream Wood in Unmanaged Forested Basins of Washington State. North American Journal of
Fisheries Management 27:342–359.

Federal Highway Administration (FHWA). 2001. Evaluating Scour at Bridges, Fourth Edition.
Hydraulic Engineering Circular No. 18. Publication No. FHWA NHI 01-001. Available:
http://www.stream.fs.fed.us/fishxing/fplibrary/FHWA_2001_Evaluating_Scour_at_Bridges.pdf.

Federal Highway Administration (FHWA). 2005. Debris Control Structures Evaluation and
Countermeasures. Hydraulic Engineering Circular No. 9. Publication No. FHWA-IF-04-016.
Available: <http://www.fhwa.dot.gov/engineering/hydraulics/pubs/04016/>.

Federal Highway Administration (FHWA). 2012. Climate Change & Extreme Weather Vulnerability
Assessment Framework. FHWA Publication No: FHWA-HEP-13-005.

Fischenich, C., and J.V. Morrow, Jr. 2000. Streambank Habitat Enhancement with Large Woody Debris.
Publication No. ERDC TN-EMRRP-SR-13. U.S. Army Engineer Research and Development Center.
Available: <http://el.erdc.usace.army.mil/elpubs/pdf/sr13.pdf>.

Flanagan, S. A. 2004. Woody Debris Transport Through Low-Order Stream Channels of Northwest
California – Implications for Road-Stream Crossing Failure. M.S. Thesis. Humboldt State
University, Arcata, CA. Available:
http://www.bof.fire.ca.gov/board/msg_supportedreports.html.

Flanagan, S. A. 2005. Woody Debris Transport at Road-Stream Crossings. Stream Notes. Rocky
Mountain Research Station. U.S. Forest Service. Fort Collins, CO. October 2005. Available:
http://www.stream.fs.fed.us/news/streamnt/pdf/SNOct_05.pdf.

Fox, M. J. and S. Bolton. 2007. A Regional and Geomorphic Reference for Quantities and Volumes of
Instream Wood in Unmanaged Forested Basins of Washington State. North American Journal of
Fisheries Management 27:342–359.

Frissell, C. A., and R. K. Nawa. 1992. Incidence and Causes of Physical Failure of Artificial Habitat
Structures in Streams of Western Oregon and Washington. North American Journal of Fisheries
Management 12:182–197.

Ghosn, M., F. Moses, and J. Wang. 2003. Design of Highway Bridges for Extreme Events. NCHRP
(National Cooperative Highway Research Program) Report 489. National Transportation Board.
Washington D.C. Available: http://www.national-academies.org/trb/bookstore.

Hamlet, A. F., M. M. Elsner, G. S. Mauger, S.-Y. Lee, I. Tohver, and R. A. Norheim. 2013. An Overview of
the Columbia Basin Climate Change Scenarios Project: Approach, Methods, and Summary of Key
Results. Atmosphere-Ocean, 51(4):392–415.

Hammer, T. R. 1972. Stream Channel Enlargement due to Urbanization. Water Resources Research
8:1530–1540.

Large Wood National Manual July 2015


7-27
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Hickin, E. J. 1984. Vegetation and River Channel Dynamics. Canadian Geographer 28(2):111–126.

Hollis, G. E. 1975. The Effects of Urbanization on Floods of Different Recurrence Intervals. Water
Resources Research 11:431–435.

Intergovernmental Panel on Climate Change (IPCC). 2007. Climate Change 2007: Impacts, Adaptation
and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change. M. L. Parry, O. F. Canziani, J. P. Palutikof, P. J. van der
Linden, and C. E. Hanson (eds.). Cambridge, UK, and New York, NY: Cambridge University Press.

Jones, C. and P. Johnson. 2015. Risk Assessment for Stream Modification Projects in Urban Ssettings.
ASCE-ASME Journal of Risk and Uncertainty in Engineering Systems, Part A: Civil Engineering.
10.1061/AJRUA6.0000815, 04015001.

Keller, E. A., and F. J. Swanson. 1979. Effects of Large Organic Material on Channel Form and Fluvial
Processes. Earth Surface Processes 4:361–380.

Knox, J. C. 1993. Large Increases in Flood Magnitude in Response to Modest Changes in Climate.
Nature, 361(6411):430–432.

Knox, J. C. 2000. Sensitivity of Modern and Holocene Floods to Climate Change. Quaternary Science
Reviews, 19(1):439–457.

Konsoer, K. M., J. A. Zinger, and G. Parker, G., 2013. Bankfull Hydraulic Geometry of Submarine
Channels Created by Turbidity Currents: Relations Between Bankfull Channel Characteristics
and Formative Flow Discharge. Journal of Geophysical Research – Earth Surface 118:1–13. doi:
10.1029/2012JF00242.

Lassettre, N.S., and G.M. Kondolf. 2003. Process Based Management of Large Woody Debris at the
Basin Scale, Soquel Creek, California. Report Presented to California Department of Forestry and
Fire Protection and Soquel Demonstration State Forest. Available:
http://www.fire.ca.gov/resource_mgt/downloads/reports/LWDinSoquelCreek.pdf.

Lagasse, P. F., P. Clopper, L. Zevenbergen, W. Spitz, and L. G. Girard.2010. Effects of Debris on Bridge
Pier Scour. Federal Highway Administration. Washington, D.C.

Montgomery, D. R., and J. M. Buffington. 1997. Channel-Reach Morphology in Mountain Drainage


Basins. Geological Society of America Bulletin 109:596–611.

Montgomery, D. R., B. D. Collins, J. M. Buffington, and T. B. Abbe. 2003. Geomorphic Effects of Wood
in Rivers. Pages 21–47 in S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology and
Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society.

Niezgoda, S., and P. Johnson. 2007. Case Study in Cost-Based Risk Assessment for Selecting a Stream
Restoration Design Method for a Channel Relocation Project. Journal of Hydraulic Engineering
133(5):468–481

Nooksack Tribe. 2013. ELJ Assessment. Deming, WA.

Roni, P., T. J. Beechie, G. R. Pess, and K. M. Hanson. 2014a. Wood Placement in River Restoration:
Fact, Fiction and Future Direction. Canadian Journal of Fisheries and Aquatic Sciences.

Rose, S. and N. E. Peters. 2001. Effects of Urbanization on Streamflow in the Atlanta Area (Georgia,
USA): A Comparative Hydrological Approach. Hydrological Processes 15:1441–1457.
Large Wood National Manual July 2015
7-28
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Salmon Recovery Funding Board (SRFB). 2013. Manual 18 Salmon Recovery Grants. Washington
State Recreation and Conservation Office. Salmon Recovery Funding Board. January.

Simon, A., and A. J. C. Collison. 2002. Quantifying the Mechanical and Hydrological Effects of Riparian
Vegetation on Stream-Bank Stability. Earth Surface Processes and Landforms 27(5):527–546.

Simon, A., A. Curini, S. E. Darby, and E. J. Langendoen. 2000. Bank and Near-Bank Processes in an
Incised Channel. Geomorphology 35(3):193–217.

Skidmore, P. B., C. R. Thorne, B. L. Cluer, G. R. Pess, J. M. Castro, T. J. Beechie, and C. C. Shea. 2011.
Science Base and Tools for Evaluating Stream Engineering, Management, and Restoration
Proposals. U.S. Department of Commerce. NOAA Tech. Memo. NMFS-NWFSC-112.

Swift, B. L. 1984. Status of Riparian Ecosystems in the United States. Water Resources Bulletin
20:223–228.

Thorne, C., J. Castro, B. Cluer, P. Skidmore, and C. Shea. 2014b. Project Risk Screening Matrix for
River Management and Rrestoration. River Research and Applications, April 2014, DOI:
10.1002/rra.2753.

Tonglao, P., and D. Eckberg. 2012. FAQ’s about Wood Placements in Rivers. March Bulletin, Skellenger
Bender Attorneys, Seattle, Washington. Available:
http://www.hallandcompany.com/php_uploads/resources/library/2012%20March%20Bulleti
n%20-
%20FAQ%27s%20About%20Wood%20Placements%20in%20Rivers%20%283%29%202012.
pdf.

Trinity River Restoration Program. 2015. Main Web Page. Available: http://www.trrp.net/.
Accessed: February 28, 2015.

Triska, F. J. 1984. Role of Large Wood in Modifying Channel Morphology and Riparian Areas of a
Large Lowland River under Pristine Conditions: A Historical Case Study. Verhandlungen-
InternationaleVereinigung für Theorelifche und Angewandte Limnologie 22:1876–1892.

U.S. Climate Change Science Program (CCSP). 2008a. Preliminary Review of Adaptation Options for
Climate-Sensitive Ecosystems and Resources. A Report by the U.S. Climate Change Science
Program and the Subcommittee on Global Change Research. (S. H. Julius and J.M. West [eds.], J. S.
Baron, B. Griffith, L. A. Joyce, P. Kareiva, B. D. Keller, M. A. Palmer, C. H. Peterson, and J. M. Scott
[Authors]). U.S. Environmental Protection Agency. Washington, D.C. 873 pp.

U.S. Climate Change Science Program (CCSP ). 2008b. The Effects of Climate Change on Agriculture,
Land Resources, Water Resources, and Biodiversity in the United States. A Report by the U.S.
Climate Change Science Program and the Subcommittee on Global Change Research (P.
Backlund, A. Janetos, D. Schimel, J. Hatfield, K. Boote, P. Fay, L. Hahn, C. Izaurralde, B.A. Kimball,
T. Mader, J. Morgan, D. Ort, W. Polley, A. Thomson, D. Wolfe, M. G. Ryan, S. R. Archer, R. Birdsey,
C. Dahm, L. Heath, J. Hicke, D. Hollinger, T. Huxman, G. Okin, R. Oren, J. Randerson, W.
Schlesinger, D. Lettenmaier, D. Major, L. Poff, S. Running, L. Hansen, D. Inouye, B. P. Kelly, L.
Meyerson, B. Peterson, and R. Shaw). U.S. Department of Agriculture. Washington, D.C. 362 pp.

U.S. Climate Change Science Program (CCSP). 2008c. Impacts of Climate Change and Variability on
Transportation Systems and Infrastructure: Gulf Coast Study, Phase I. A Report by the U.S. Climate

Large Wood National Manual July 2015


7-29
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 7. Risk Considerations

Change Science Program and the Subcommittee on Global Change Research (M. J. Savonis, V. R.
Burkett, and J. R. Potter [eds.]). U.S. Department of Transportation. Washington, D.C.

U.S. Environmental Protection Agency (EPA). 2014. Green Infrastructure. Available:


http://water.epa.gov/infrastructure/greeninfrastructure/index.cfm.

U.S. Global Change Research Program (USGCRP). 2009. Global Climate Change Impacts in the United
States. Edited by T. R. Karl, J. M. Melillo, and T. C. Peterson. Cambridge, MA: Cambridge
University Press.

Wallerstein, N. P., and C. R. Thorne. 2004. Influence of Large Woody Debris on Morphological
Evolution of Incised, Sand-Bed Channels. Geomorphology 57:53–73.

Warner, M. D., C. F. Mass, E. P. Salathé Jr. 2012. Wintertime Extreme Precipitation Events along the
Pacific Northwest Coast: Climatology and Synoptic Evolution. Monthly Weather Review,
140(7):2021–2043.

Washington Department of Fish and Wildlife (WDFW). 2012. Stream Habitat Restoration Guidelines.
Washington Department of Fish and Wildlife, Olympia, Washington, 2012.

Large Wood National Manual July 2015


7-30
Chapter 8
REGULATORY COMPLIANCE, PUBLIC INVOLVEMENT,
AND IMPLEMENTATION

Photo credit: Ken DeCamp

AUTHORS

Doug Shields (Shields Engineering, LLC)


Gregg Ellis (ICF International)
Leo D. Lentsch (ICF International)
David Bandrowski (U.S. Bureau of Reclamation)
Mike Hrachovec (NSD)
Rocco Fiori (Fiori Geosciences)
This page intentionally left blank.
The first step in implementation is synonymous
8.1 Introduction with the last step in design: a constructability
This chapter discusses the way federal, state, and assessment.
local regulations influence the placement,
operation, and long-term operation of large
CROSS-REFERENCE
wood. It describes the regulatory background,
offers potential scenarios under which the
For an overview of all of the steps involved in a large
regulations may apply, and provides potential
wood project, see Chapter 2, Large Wood and the
best management practices designers and Fluvial Ecosystem Restoration Process. For more on
installers should consider. constructability assessments, see Chapter 6,
Engineering Considerations.
Secondly, the chapter focuses on public outreach
during a large wood project, describing how to
best inform the public of project goals, design, Two overriding considerations in
and construction and maintenance to ensure constructability assessment are cost efficiency
their understanding and support. and minimization of collateral environmental
damage. Although some disturbance to the
Finally, the bulk of the chapter describes the stream corridor is necessary for large wood
implementation process itself. Large wood projects, implementation should be conducted to
project implementation is crucial because even ensure that project benefits are not outweighed
the best planning and design can be entirely by impacts. Those who implement restoration
negated by shoddy construction work. If a projects would do well to adopt an ethic that
project is poorly implemented so that the design requires them to avoid projects that risk doing
intent or intended outcomes are compromised, more harm than good. Examples of
actual or perceived project failure may make it constructability issues include the following
difficult to pursue future large wood projects. (NRCS 2007c; Cramer 2012):
Skills for implementing large wood projects
draw from typical logging work, heavy  Safety for construction personnel.
construction, and, more specifically, stream  Permits and agreements.
restoration methods, but also include a suite of
 Constraints due to existing infrastructure
practices and techniques not required for other
and utilities.
types of projects. Accordingly, contractors
working with instream or floodplain wood for  Material availability.
the first time will encounter a learning curve that  Equipment availability and capability.
will tax the vigilance and communication skills of
inspectors and designers. For example,  Site and staging area access for various types
contractors may not appreciate the importance and sizes of equipment.
of retaining tree crowns and rootwads on boles  Labor requirements.
for certain types of wood construction. Others
may have limited or no experience in  Dewatering requirements and trafficability
transporting large wood, earthmoving in a for equipment.
stream corridor, dewatering, or affixing cables to  Sequencing and seasonal restrictions that
large wood. Handling and successful installation include, but are not limited to, these:
of plant materials requires specialized expertise.
 Optimal periods for material sourcing
Safety is of paramount importance during
and vegetation establishment.
implementation, and the project team must
comply with applicable laws and regulations as  Trafficability properties associated with
well as ordinary common sense. wet or frozen soils.

Large Wood National Manual July 2015


8-1
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

 Migratory/breeding period constraints. and construction approach. This will


increase the odds of receiving qualified bids
 Turbidity control and in-water work
and reduce the potential for quality control
windows.
issues and change orders.
 Flooding and site dewatering issues.
 Contract type. Using the right contracting
 Measures to protect habitat and water mechanism for the project is an important
quality, such as these: element during the planning process. There
 Limiting impact zones by working from are considerable differences between
banks, platforms, or gravel bars. standard fixed-price and time and material
type contracts. It is difficult to define exact
 Using special equipment. scopes of work for large wood projects and
 Avoiding sensitive soils and vegetation. therefore flexibility is required for
construction. See Appendix A-1 and NRCS
After the constructability assessment, initial (2007c) for additional detail on types of
steps in project implementation include the federal contracts available for large wood
following (Federal Interagency Stream projects.
Restoration Working Group [FISRWG] 1998):
 Scheduling implementation events and Usually designers will prepare plans (drawings)
obtaining required permits. and specifications that govern construction
activities. The level of detail provided in plans
CROSS-REFERENCE and specifications depends on the type of
contract, as explained in Appendix A-1. Typical
specifications will contain sections dealing with
For guidance on permits and regulations, see the following:
Chapter 2, Large Wood and the Fluvial Restoration
Process, and sections below dealing with Regulatory  Mobilization and demobilization
Compliance and Public Involvement.
 Pollution and erosion control
 Removal of structures (if necessary)
 Informing landowners and other
stakeholders.  Site preparation (haul roads/staging areas)

 Securing site access and easements.  Large wood sources and harvesting

 Locating utilities.  Large wood transport

 Confirming sources and quality of large  Excavation


wood, plant materials, and other materials.  Fill
 Prebid meeting (making the prebid meeting  Quality control
attendance required for all bidders increases
 Large wood placement
the odds of receiving qualified bids and
reduces the potential for change orders).  Site closure and cleanup
 Contractor selection. Due to the inherent  Revegetation
complexity and uncertainty of site conditions
and wood materials in many restoration
projects, it is highly advised that contractor
selection not be based solely on low bids, but
also on contractor experience, qualifications,

Large Wood National Manual July 2015


8-2
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

placement or detailed for each site), details for


GUIDANCE
haul roads and staging areas, and locations and
details for revegetation and plantings.
Regional Considerations in
Large Wood Project Implementation In addition to typical safety concerns for a
construction project within a stream corridor,
Each of the key constructability issues listed below large wood projects incur other hazards. Logs
displays variation across regions. People implementing
may shift in unforeseen and potentially lethal
large wood projects may turn to case studies and
personal communications with experts from other ways. Wood breakage may result in rapid
regions, but should remain aware of regional movement of heavy objects or ejection of
differences, including the following: dangerous fragments. Cable breakage or slippage
 Total funding available. Costly approaches such as can create extreme danger to personnel due to
helicopter transport of wood or importation of rapid rebound. Additional discussion of safety
logs with rootwads from a distant harvest site are issues is provided below and guidance of a
not feasible in many regions with less funding general nature is provided by USACE (2008).
support for restoration work.
 Size and type of wood available. Large, decay-
resistant logs are simply not found or are not GUIDANCE
economically accessible in some parts of the
Midwest and Southeast. Types of Contracts for Large Wood Construction
 Flow seasonality. Regions with snowmelt
hydrology and less intense storms tend to present Fixed-Price Contracts—Place the maximum risk and full
responsibility on the contractor for all costs and resulting
different risk profile with respect to high flows
profit or loss associated with the work. This type of
during construction.
contract provides the maximum incentive for the
 Environmental protection. The rigor required of contractor to control costs and perform effectively and
contractors working adjacent to and within imposes a minimum administrative burden on project
streams in terms of sediment and turbidity control, sponsors. A fixed-price contract requires the contractor to
fish exclusion, temporary crossings, and understand, in detail, what is to be constructed before
dewatering protocols varies widely from state to bidding to do the work. This requires a design that includes
state and even from stream to stream. detailed drawings, specifications, and a bid schedule
containing a bid item for each major item of work. The
 Wood restraint. Acceptability of the use of natural designer must provide a cost estimate by bid item so that
ballast and wood entanglement rather than pilings, the cost of the work can be estimated and the contracting
wood burial, and attachment of wood to fixed officer can assess the reasonableness of the bids. Most
points using cables, chain, or hardware varies. fixed-price contracts are awarded after contractors have
 Revegetation. The impact of exotic plant species, submitted a sealed bid in response to an Invitation for Bids
ice, and herbivores on planted vegetation can be (IFB). The IFB includes the drawings and specifications for
extreme in some areas. the work and specific contract requirements. The design
effort and level of detail may be the same for simplified
 Contractor expertise. The availability of capable, fixed-price contracts as it is for formal fixed-price
experienced contractors is much greater in regions contracts.
with strong restoration funding and a history of
large wood projects. Cost-Reimbursement Contracts—Suitable for use when the
cost of the work cannot be estimated with sufficient
accuracy to use a fixed-price contract. The cost of the work
is estimated for the purpose of obligating funds; however,
Methods for measuring quantities for payment
a detailed cost analysis is not required. The contractor
are often specified if required by the type of must have an accounting system adequate for determining
contract selected. incurred costs that are reimbursable. This type of contract
requires significantly more oversight during the
Typical drawings contain sheets showing the construction phase to document that efficient
project location (maps), plan and profile construction methods and efficient cost controls are being
drawings of large wood placements and used. It provides little incentive for the contractor to
control costs and perform effectively and imposes a much
structures (either typical for each type of

Large Wood National Manual July 2015


8-3
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

larger administrative burden on the contractor and project requirements or planning principles, is
sponsors. discussed.
Incentive Contracts—Link the contractor’s profit to
performance by establishing reasonable and attainable
There are many federal, state, and local
targets that are clearly communicated to the contractor. regulations that could apply to the installation of
These contracts are designed to motivate the contractor to large wood within a stream or river. Although
achieve certain goals such as completion by a target date. federal regulations apply throughout the United
Incentive contracts discourage inefficiency and waste.
States, others obviously vary among the states
They can be fixed-price incentive contracts or cost-
reimbursable incentive contracts. These types of contracts and local jurisdictions. Information provided
are normally used for performance-based service about state and local regulations is intended to
contracts and rarely for construction work. serve as examples of what may be applicable and
Time-and-Materials Contracts—Used to procure supplies should be considered individually for each
or services on the basis of direct labor and materials costs. project. The primary regulations relevant to the
Time-and-materials contracts should be used only when it installation of large wood are discussed below.
is not possible to accurately estimate the extent or
duration of work or to anticipate costs with any degree of
confidence. With this type of contract, there is no 8.2.1 Federal Regulations
incentive to the contractor to control costs, significant
sponsor oversight is required, and a much larger Actions involving modification of channel
administrative burden is imposed on the sponsor. structure and instream habitats (e.g., placement
Labor-Hour Contracts—A variation of the time-and- of large wood) will most likely involve activities
materials contract, differing only in that materials are not in navigable waters or waters of the United
supplied by the contractor. States and involve the discharge of fill material,
Equipment Rental Contracts—Used in instances when it is triggering the need for compliance with Section
not feasible or desirable to prepare detailed drawings and 404 of the Clean Water Act and Section 10 of the
specifications. Require substantial construction oversight Rivers and Harbors Act, administered by USACE.
and impose an additional administrative burden on the The extent of the action will dictate if a
sponsor.
Nationwide Permit can be utilized or if an
individual permit is required.

8.2 Regulatory In some cases, the installation of large wood may


be considered “vegetation” and, if within the
Compliance and footprint of federal flood control levees, is
Public Involvement subject to USACE Engineering Technical Letter
1110-2-583, Guidelines for Landscape Planting
Most restoration activity decisions must address and Vegetation Management at Floodwalls,
environmental policy. This section describes the Levees, Embankment Dams, and Appurtenant
types of federal, state, and local regulations that Structures. This may limit the locations in which
control or may influence the initial placement large wood can be installed.
and long-term operation and maintenance of
large wood. It also describes some hypothetical Any alterations or modifications to existing
scenarios under which the various regulations USACE projects, such as federal flood control
may apply and strategies for compliance. levees, must request and be granted permission
Potential safety issues associated with placing from USACE pursuant to Section 14 of the Rivers
large wood in areas where recreational activities and Harbors Act of 1899 (Title 33 of the United
are common are also discussed, along with a States Code [USC], Section 408)—hereinafter
recommended process for addressing these referred to as Section 408—for the alteration of
issues. Finally, the need for outreach to the a federal work (e.g., levee).
public, whether driven by regulatory

Large Wood National Manual July 2015


8-4
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Because actions that include installation of large require compliance with state Waste Discharge
wood will most likely occur in areas where Requirements.
species that are federally listed as threatened or
Over 90% of states have some form of
endangered or that are candidates for listing
endangered species act. They vary widely with
may be present, compliance with the federal
some just prohibiting either the “taking” of or
Endangered Species Act (ESA) (in coordination
trafficking in an endangered species to more
with the U.S. Fish and Wildlife Service [USFWS]
comprehensive processes for species listing,
and the National Marine Fisheries Service
management, protection, and recovery. It is
[NMFS]) may be required. A current list of
essential to know the details of the applicable
species protected under the ESA should be
state program in order to ensure proper
obtained from the regulating agencies in order to
compliance. Alabama, North Dakota, West
begin the compliance process.
Virginia, and Wyoming are the states that
Separate from federal agency involvement in currently do not have their own state-level
proposing the project (e.g., placement of large endangered species acts. Because these actions
wood), which would itself require National may occur in areas where species that are state-
Environmental Protection Act (NEPA) listed as threatened or endangered or that are
compliance, involvement of USACE and/or candidates for state listing may be present,
USFWS and NMFS may trigger the need for NEPA compliance with any applicable state
compliance, and may trigger the need for endangered species acts may be required.
compliance with the Fish and Wildlife
These actions are also likely to involve changing
Coordination Act when actions involve the
a streambed or altering streambed material,
modification of surface water. Federal agency
triggering the need for compliance with various
involvement may trigger the need for
state fish and wildlife code and/or regulation.
compliance with Section 106 of the National
For example, in California a Section 1600
Historic Preservation Act if the action would
Streambed Alteration Agreement with the
occur in an area where properties are listed, or
California Department of Fish and Wildlife is
are eligible for listing, on the National Register of
required. Many other states have similar
Historic Places.
requirements.
Actions in this category will most likely occur in
Compliance with state and local flood
or affect wetlands, triggering the need for
management agencies and reclamation districts
compliance with Executive Order 11990
may also be required, especially for actions in
(protection of wetlands). They may also be
state-designated floodways, floodplains, and
located within a floodplain and require
shorelines. If both a state and local agency exist,
compliance with Executive Order 11988
they are often used to working in tandem, and
(floodplain management).
early coordination with both agencies is
recommended. The local reclamation district
8.2.2 State and Local (sometimes called a levee maintenance district)
Regulations is often responsible for the maintenance of local
waterways and will have a keen interest in large
Because actions would involve activities with the wood projects and how they may influence their
potential to mobilize contaminants in surface ability to conduct maintenance in the long term.
waters and require compliance with Section 404,
state certification under Section 401 of the Clean Requirements for state and local authorizations
Water Act will usually be required. Because such will trigger the need for state-level
actions could result in the temporary discharge environmental compliance in those states with
of waste affecting surface water, many may applicable laws (e.g., the California

Large Wood National Manual July 2015


8-5
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Environmental Quality Act [CEQA]). The of these permits and requirements. On the other
following states have some form of hand, these projects can be complicated and
environmental impact assessment laws: subject to more than a dozen regulatory
California, Connecticut, Delaware, Georgia, compliance processes. Therefore, the timeframe
Hawaii, Indiana, Maryland, Massachusetts, for receiving agency authorizations can vary
Michigan, Minnesota, Montana, New Jersey, New greatly and should be initiated as soon as
York, North Carolina, South Dakota, Virginia, adequate information about the project has been
Washington, and Wisconsin. developed.

In many cases the environmental effects Table 8-1 elaborates further on the preceding
associated with the placement of large wood are narrative by posing questions that serve as a
expected to be minimal and the environmental basic regulatory compliance decision analysis
benefits are expected to be high. In some cases, tool.
this may help speed the processing and issuance

Table 8-1. Large Wood Regulatory Compliance Decision Analysis

Do the Following Apply? If Yes, Compliance is Required With:


Federal Compliance
The action is considered a federal agency proposal. National Environmental Policy Act (NEPA)
The action is located in waters of the United States, Section 404 of the Clean Water Act and Section 10 of
including wetlands, and/or the action is located in the Rivers and Harbors Act
navigable waters of the United States; and the
action is considered a discharge of dredged or fill
material; or the action would affect facilities
designed, built, or managed by USACE.
The action would alter a federal project, such as a Section 408 (33 USC 408)
federal flood control levee.
The action is considered a major construction Section 7 or 10 of the ESA
activity, and species listed as threatened or
endangered under the federal ESA may be found in
the project area; the action may affect the listed
species (Section 7).
The action may result in the “take” of a species
listed as threatened or endangered under the ESA
(Section 7 or 10).
The action is considered a federal agency proposal Fish and Wildlife Coordination Act
and proposes to control or modify surface water.
The action is considered a federal agency proposal National Wild and Scenic Rivers Act
and affects a river within the National Wild and
Scenic Rivers system.
The action is considered a federal agency proposal Executive Order 11988 – Floodplain Management
and is located within or may affect a floodplain.
The action is considered a federal agency proposal Executive Order 11990 – Protection of Wetlands
and is located within or may affect wetlands.
The action is considered a federal agency proposal Executive Order 12898 – Environmental Justice
and may affect minority or low-income populations.

Large Wood National Manual July 2015


8-6
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Do the Following Apply? If Yes, Compliance is Required With:


The action is considered a federal agency proposal American Indian Religious Freedom Act of 1978
and may affect Native American religious practices.
The action may affect Indian Trust Assets. Indian Trust Assets
State or Local Compliance
The action involves a state or local agency action State Environmental Impact Assessment Laws (e.g.,
and is considered a project for such purposes. CEQA, Washington’s State Environmental Policy Act
[SEPA])
The action involves a federal license or permit that Section 401 of the Clean Water Act
may affect state water quality, and the action would
result in a discharge of a pollutant into waters of the
United States.
A species listed as candidate, threatened, or Fish and Game Code – California Endangered
endangered under the California Endangered Species Act
Species Act may be present in the project area or
the action may result in the “take” of a state listed
species.
The action involves any activity that will divert or State Streambed Alteration Agreement (e.g., Section
obstruct the natural flow or change the bed, 1600 of California’s Fish and Game Code)
channel, or bank of any river, stream, or lake; the
action involves the use or alteration of any
streambed material; the action occurs within the
annual high-water mark of a wash, stream, or lake.
The action occurs in tideland; submerged land; the State agency overseeing sovereign lands of the state
bed of a navigable river, stream, lake, bay, estuary, (e.g., California State Lands Commission requires a
inlet, or strait; swamp land, or overflowed land; the land use lease)
action would affect water-related commerce,
navigation, fisheries, recreation, open space, or
other public trust uses.
The action would affect existing state flood control State or local agency overseeing any state/local plan
project facilities, including levees, dams, reservoirs, of flood control (e.g., California’s Central Valley
and floodways and flood control plans. Flood Protection Board, local reclamation and/or
levee districts)
The action would involve grading, building or City or county approvals and entitlements
modifying structures, special or conditional uses,
modification or approval of general or specific plans
(local or regional), and/or zoning ordinance
amendment.

Large Wood National Manual July 2015


8-7
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

between outreach, engineering, and science


8.3 Public Involvement ensures a strongly integrated project.
and Input As appropriate and tailored to meet the scope
Public outreach during a large wood project and scale of the project, coordination with a
may occur for several reasons. In general, variety of stakeholders—local fire/sheriff/
outreach will be associated with public noticing rescue, local flood control entities, rafting
required by regulations, public outreach to companies, fishing guides—will lead to project
solicit design input and to build project support, support. Throughout the development of the
and outreach to inform river users about the preliminary design, the project proponent
presence of large wood to help ensure their should engage these stakeholders through
long-term safety. project briefings, one-on-one meetings,
presentations, and social media and email
Public noticing is required by several communication—once again, as appropriate
regulations, including federal and state depending on the size and location of the
environmental impact assessment laws (e.g., project. These stakeholders should be engaged
NEPA, CEQA, SEPA, Clean Water Act Section early and often in order to provide feedback on
404). Each of these regulations and the the preliminary designs and to express their
regulating agencies have very specific guidance preferences or concerns to the design team. It is
on the type and content of noticing materials. In important to ensure that the right stakeholders
addition to meeting the regulatory are sitting at the table, that the outreach
requirements, this type of outreach can be an process is tailored to respect the contribution
effective method of reaching a wide variety of and time the stakeholders will give to project
important stakeholders in order to make them review and development, and that these
aware of the proposed project and solicit their stakeholders are provided with regular and
input on a number of matters. meaningful opportunities for engagement in the
One type of feedback that can result from planning process. Key stakeholders such as
effective outreach is input on the design of a those listed above have the potential to be
project. While engineers and aquatic scientists project champions/ambassadors of the project
are typically key participants in a large wood as it becomes reality.
design team, input from stakeholders can also Once a project is constructed, it is important to
be very important. People tend to support what continue the outreach to inform river users on
they help build, and early involvement of an ongoing basis about the presence of large
stakeholders can lead to the most successful wood to help ensure their long-term safety.
buy-in and project support. One of the greatest There are always new river users, so providing
elements of a successful public outreach permanent, educational sources of information
program is the development and maintenance is essential. Appropriate methods include, but
of strong relationships with key stakeholders are not limited to, signage, websites, and
and members of the community, built on and interpretive displays.
sustained by trust. Opportunities for effective
two-way communication are vital to building
trust and support.

All aspects of public outreach need to be


tailored to engage the community and be the
appropriate level of effort for the project—no
two projects are alike. Developing a synergy

Large Wood National Manual July 2015


8-8
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Environmental Impact Report is required to


8.4 Regulatory satisfy NEPA and CEQA; Biological Assessments
Compliance are needed to comply with the federal and state
endangered species acts; and compliance with
Approaches Clean Water Act Sections 404 and 401 is
needed.
8.4.1 Scenario 1: Project
While both of these projects involve the
Site with an installation of large wood, the differences in the
Endangered Species specific circumstances of each result in
substantially different regulatory compliance
Consider a large wood project on a stream that requirements. It is very important to consider,
supports federally endangered salmon. The understand, and plan for these types of
project involves a relatively small area and is situations and develop your approach
intended to enhance conditions for rearing and accordingly. A key strategy for successful
migrating salmon by providing instream regulatory compliance is to begin project
structure in an area where it does not exist. Due coordination early and make use of the
to the relatively small size of the installation information gleaned through coordination with
and its purpose (enhancing aquatic habitat), a regulators, resource specialists, river users, and
Clean Water Act Section 404 Nationwide Permit other relevant stakeholders.
(NWP) is utilized (NWP 27: Aquatic Habitat
Restoration, Establishment, and Enhancement
Activities). NEPA compliance is addressed 8.5 Construction
through the NWP. The project description
clarifies that the project will be Key elements of construction include:
constructed/installed when the endangered  Construction oversight
salmon are not present and NMFS issues a
Letter of No Effect under the ESA. Due to the  Dewatering and diversion
scope and location of this project, no other  Excavation
regulatory approvals are required.
 Wood placement

8.4.2 Scenario 2: Erosion  Securing wood

Control Project  Finish work

Another scenario involves a much larger and


complex undertaking. Consider a project
8.5.1 Construction
intended to control erosion that is threatening a Oversight
flood control levee while at the same time
would provide instream structure for 8.5.1.1 Risk Management
endangered salmon through the use of large
wood. The river is in California and has been
CROSS-REFERENCE
designated as Wild and Scenic under both
federal and state acts, and it supports a wide
variety of recreational uses and a high number Chapter 7, Risk Considerations, provides detailed
guidance on overall project risk assessment and
of users. Due to the nature of the project,
management.
compliance with a majority of the regulations
described in Table 8-1 is required. For example,
an Environmental Impact Statement/

Large Wood National Manual July 2015


8-9
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Construction projects often result in litigation Small large wood projects may proceed with
due to disputes regarding liability for accidental minimal plans and specifications. Minimal
injuries, cost overruns, project failure, and the drawings and specifications may be sufficient if
like. The distribution of legal liability among a time-and-material or a labor-hour contract is
owners, designers, and contractors is a highly employed, experienced inspection personnel
technical subject, and personnel involved in familiar with the design will be on site, and the
large wood projects should obtain legal counsel contractor is experienced and reliable. Simple
for review of contractual arrangements. projects in smaller streams may benefit from
reliance on typical sketches of certain types of
8.5.1.2 Contract Types and Risk large wood configurations rather than detailed
drawings (e.g., McMillen 2014). Exact locations
A brief review of some of the types of federal for large wood may be revised just prior to
contracts available for large wood projects is construction due to events such as channel
provided in Appendix A-1 and NRCS (2007c).
migration or tree wind throw, or due to
Fundamental differences in liability
constructability logistics; as a result, the
apportionment exist between arrangements original site-specific drawings are often
where designers act as advisors to the project obsolete by the time staking occurs
owner or sponsor and those in which the
immediately prior to construction.
designer is contracted by the owner or sponsor
to perform construction. In the former, the Although time-and-materials or labor-hour
designer creates the design, and prepares plans contracts or arrangements are sometimes
and specifications, cost estimates, and perhaps employed in part or in whole for larger scale
schedules and constructability reviews. The projects, the formality of project management
designer may also provide inspection services should increase with project scale. The number
or high level advice during construction, but and experience of construction inspectors and
does not award contracts for construction, the involvement of designers is key in larger
materials, or labor. In the latter arrangement scale, high-risk projects. Formal submission of
(sometimes called design-build), the design- operational, safety and health, pollution control,
constructor performs the functions listed above and other plans by the contractor and
but also subcontracts for actual construction. In documentation of reviews, approvals, and
this case, the design-constructor typically denials by the project sponsor must be more
assumes the same risks and responsibilities as meticulous. Accounting for construction
the general contractors, including safety. activities, delivery, and disposition of supplies
and materials, and hydrologic and geotechnical
8.5.1.3 Risk and Project Scale conditions encountered should also increase
with project scale.
In theory, risk and project scale are
independent as even very small projects in
8.5.1.4 Risk and Project
critical locations may incur large risk, while
larger projects in remote locations may incur
Management
moderate risk. However, because project cost is Key construction oversight tasks involve clearly
often a function of scale and projects in remote delineating work zones, access routes, haul
locations that encounter less potential risk to roads, and staging areas on the ground and on
adjacent infrastructure are likely to affect more project documents. Material quality and
valuable habitats, most projects exhibit a tight quantity should be assessed and recorded, and
linkage between risk and scale (geographic and measurements of quantities of large wood
economic). material, excavation, fill, plant materials, etc.
should be conducted as specified in contract

Large Wood National Manual July 2015


8-10
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

documents. The order of operations should be Construction Equipment, for a discussion on the
controlled in a manner that allows the appropriate equipment for transporting wood.
contractor leeway when possible but avoids Large wood stockpiles may require security
undesirable impacts. Considerable ingenuity measures because wood may be vandalized for
may be required to avoid sensitive habitats, firewood sourcing.
cultural resources sites, highly erodible soils,
Effort should be applied to minimize collateral
and soils too wet or soft to support vehicles and
environmental impacts associated with harvest,
equipment.
transport, and stockpiling of large wood. Use of
Daily records should be kept in a log file by the locally derived materials, when available, tends
sponsor and/or designated field to reduce costs and overall impacts (Figure
engineer/scientist when on site to document 8-1). However, strict limits should be placed on
field conditions, construction progress, the acceptable size, species, condition, and
compliance with design plans, and distance from the stream for local large wood.
conversations with the contractor. Daily logs For example, the contract could stipulate “Only
should include photos from fixed points and live trees more than 8 meters (25 feet) away
plan mark-ups documenting whether the from the channel top bank may be used,” and
project is being constructed as designed or if “Only downed wood more than 0.3 meter
changes are needed. These logs can serve as key (1 foot) vertically above specified elevation may
information to the sponsor and/or designer be used.” Trees with obvious cavities used for
during any disputes and can limit the potential nesting may be excluded.
for unnecessary change orders that increase
costs. Figure 8-1. Use of Locally Sourced Large Wood

8.5.1.5 Material Sourcing


Often large wood construction projects are
awarded with little consideration regarding the
timing needed for proper wood sourcing. A
wood sourcing plan should be developed during
the design phase of the project. The wood
sourcing plan should specify harvest locations
and equipment, hauling and loading equipment,
and stockpile or staging locations near
placement sites.

Wood sourcing should be a separate phase of


Little Topashaw Creek, Mississippi. Haul Distance
the construction contract or secured through a
from fencerow harvest zone <0.8 kilometer
different contract or agreement (see Appendix (0.5 mile).
A-3 for sample large wood harvest and hauling
contract language). Some wood material may be Alternatively, living and downed wood that is
harvested on site as a consequence of the suitable for use may be flagged by construction
excavation, but this rarely produces the inspectors working with personnel qualified to
required quantity for the project. Offsite assess the environmental significance of the
sourcing is challenging and may require a materials, and the contractor may be prohibited
timber harvest plan or environmental from using materials that are not flagged.
compliance documents for harvesting in
Because forest products used for restoration
sensitive areas. Hauling can also be a major
projects include soil and biomass, and these
constraint; see Section 8.5.7, Typical

Large Wood National Manual July 2015


8-11
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

materials will likely be placed in flowing water, component in large wood construction. Slash
the potential to spread disease and unwanted material supplies should allow for significant
species is possible unless proper precautions compaction, and specified volume of slash may
are taken. Wood materials should be sourced be increased by 50 to 75%.
and stockpiled in areas free of disease and
invasive species (e.g., root-rot, rust, sudden oak Table 8-2. Size Categories for Large Wood
death syndrome, insects, ivy, and pampas
Nominal DBH
grass). In some regions of the United States, length Range
inspections by the county agricultural (feet) Morphology (inches)
department need to occur and the disease 32 Log with rootwad 6–12
status of forest products signed-off before
32 Log with rootwad 12–24
transporting materials from the harvest
32 Log with rootwad 24–36
location.
20 Butt Log–Racking material 4–6
and logs sharpened at one
8.5.1.6 Material Types end to be used as pins
Designers should specify quantities and sizes 32 Horizontal Logs and logs 6–12
(diameter and length) of different materials sharpened at one end to
including rootwads, butt logs, slash, tree tops, be used as structural piles
etc. Most large wood structures contain a few 32 Butt log 12–24
very large key pieces, usually with rootwads 32 Butt log 24–36
attached. The numbers of slightly smaller logs
Large wood delivered to the site should be
and boles required to construct the body of the
inspected to ensure it meets the species, size,
large wood structure around the key members
and quality specifications provided in the bid
is usually much greater (~10x) than for the key
documents. Any pieces not meeting the
members. Even more numerous smaller limbs
specifications should be tagged and removed
and slash are used to fill gaps and interstices
from the site immediately.
within the structure. For example, oblique logs
are small (15- to 30-centimeter [6- to 12–inch]) Similar concerns attend procurement of
diameter logs that are wedged, at off-vertical boulders and other coarse sediments for
angles, into the gaps of a logjam. Material types anchorage and ballasting. Importation of these
should be specified by DBH, length, with or materials is likely to be costly, and removal
without a rootwad, and wood species. Material from the base flow channel may create
specifications affect costs. For example, a unacceptable impacts.
60-centimeter (24-inch) DBH tree with a
rootwad can be an order of magnitude more 8.5.1.7 Wood Transport
costly than a 30-centimeter (12-inch) DBH butt
log. Harvest plans should allow for procurement The cost to harvest and deliver logs with
of about 25% more large wood than is indicated attached rootwads is generally three to four
by design drawings. Large wood material can be times greater than the cost for similar size logs
placed into the general size categories as shown without rootwads. Rootwad complexity directly
in Table 8-2. benefits fish habitat quality, geomorphic
function, and stability of a constructed logjam.
In addition to the log types listed in Table 8-2, Therefore, the importance of retaining as much
material plans should include quantities for of the root system as possible cannot be
whole tree tops and smaller slash material. overstated and should be emphasized at each
Slash is generally too small to be included as step of the log-handling process (i.e., harvest,
large wood material but is an important loading, transport, stockpiling, and placement).

Large Wood National Manual July 2015


8-12
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Furthermore, rough handling can split and/or Figure 8-2. Equipment Useful for Handling Large
gouge the bole, which weakens the structural Wood
integrity of the log and makes it more
susceptible to decay. Several measures can be
employed to avoid degrading the habitat quality
and premature failure of a logjam due to rough
log handling practices.

An excavator equipped with a bucket and


hydraulic thumb is typically used as the prime
mover of logs, slash, and earth materials for
large wood projects. Because a bucket and
thumb has limited capabilities to rotate and
position logs, contractors can opt to use an
excavator with a bucket and thumb for earth-
moving tasks and an excavator with a rotating
grapple or log shovel for most of the log
handling tasks (Figure 8-2a and 8-2b.).
Although having both types of equipment is
reasonable for large-scale projects, this can be
cost-prohibitive for smaller scale projects.
Alternatively, contractors can equip an
excavator with a detachable heel rack or
rotating grapple, or simply attach a set of log
tongs, via heavy chain, to the back of the
excavator bucket (Figure 8-2c). Any of these
options will allow the operator to more
efficiently position logs, while minimizing
damage to the rootwad, bole, and surrounding
environment. However, even when these
attachments are available on a job site, most
operators need instruction to minimize
handling the log by the rootwad and to avoid
damaging the material.

8.5.1.8 Change Orders


Change orders during construction can be
costly. Some amount of change from the design
plans during the construction process should be
expected on every large wood project for the
reasons previously mentioned.
(a) Use of excavator with bucket and hydraulic
thumb to position log with rootwad. (b) Use of
rotating grapple to move multiple small logs.
(c) Log tongs used to position rootwad log in tight
quarters.

Large Wood National Manual July 2015


8-13
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Projects where a certain amount of uncertainty  Blocking flows upstream and bypassing
is anticipated should build that expectation into water through pipes using pumps (Figure
the design plans, contract documents, and 8-3) or a gravity system.
discussions with contractors to reduce the  Isolating a portion of the site using
potential for costly change orders during the cofferdams or push-up dikes.
construction process. Common strategies to
build flexibility into the contract documents Figure 8-3. Diversion of a Small Stream Around a
include bidding items lump sum and creating Construction Zone Through Plastic Pipe
force account items for miscellaneous items
(e.g., setting up bid item for contractor to lock in
cost of machine and operator time).

8.5.2 Water Management


Large wood placement methods vary greatly
from project to project due to site conditions,
with economic and regulatory constraints
dictating the project approach. Water
management is a key component for managing
collateral environmental impacts. Construction
plans should include stream crossing plans and Eel River Headwaters, Massachusetts. Photo
prescribe wood placement methods. Available courtesy SumCo Eco-Contracting.
approaches may be broadly classified as:
 Working in dry conditions, 8.5.2.2 Wet but Controlled
Conditions
 Working in wet, but controlled conditions,
and In many cases, working in completely dry
conditions is not possible. Full dewatering is not
 Working in existing (wet) conditions, not
always needed or desirable, but isolating a
controlled.
work zone is often required to control
downstream turbidity. Working in more
8.5.2.1 Dewatering and Diversion controlled conditions where the project area is
Working in dry conditions is typically still wet creates a functional approach to
completed in smaller channels, in perennial complete the activity (Figure 8-4). This scenario
streams where complete diversion is easily is common in tidally influenced areas, large
achieved, or on gravel bars in large rivers. It is river systems, estuaries, shorelines, and ponds.
especially advantageous to work in dry While complete diversion is typically not
conditions when placing bed control structures, achievable, a variety of methodologies to divert
placing vertical posts or pilings, excavating to flow and partially or fully isolate the work site
scour depth, or having to build forms for cast- from flowing water are listed in the Guidance
in-place concrete. When water diversion is box below. Dewatering or isolation must be
needed to ensure dry conditions, available complete prior to excavation or large wood
methods include the following: placement.
 Rerouting of the stream through a bypass
channel.

Large Wood National Manual July 2015


8-14
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Figure 8-4. Water Management Techniques for duration, excavation below the water line,
Large Wood Projects operation of equipment in flowing water, and
other actions that mobilize suspended
sediments.

GUIDANCE

Dewatering and Turbidity Control Techniques

Construction of Push-Up Berms using onsite material is


relatively inexpensive and easy. If native material proves
to be too porous, has a high clay content, or if there is not
enough suitable material on site, a specified mix can be
imported and used and exported at the end of the project
or repurposed on site. The width and length of the berm
will depend on the volume of water that needs to be
diverted. Typically, the wider and longer the berm, the
more the berm will be more resistant to the forces of the
water.

Sand Bagging is the filling of individual bags of woven


geotechnical fabric, burlap, or plastic with sand or rounded
gravel. The bags come in several colors including white,
green, and orange; and, dependent on the situation, can
be left in place to degrade or removed at the end of the
project. The bags can also vary in size from commonly
available small bags (approximately 1-cubic-foot volume)
(a) Wet but controlled conditions used to isolate to large “bulk” bags (approximately 1-cubic-yard volume).
turbidity impacts in work zone from main Typically, the bags are sacked no higher than 1.2 meters (4
channel. (b) Highway traffic barriers being used to feet) or 2–3 courses high because of their non-structural
construct a temporary isolation cell prior to wood material contents making them susceptible to falling over
easily.
placement. Photos by Ken DeCamp.
Water Filled Bladders or “Aqua Barriers” are portable
8.5.2.3 Uncontrolled Wet water-inflated temporary dams and are designed for
construction worksite dewatering and/or water diversion
Conditions work. They are manufactured using high quality industrial-
grade vinyl, which provides an economical, effective, and
While working in dry conditions or semi-dry
safe alternative to conventional dam methods like
conditions is desirable, it is sometimes not sandbags and push-up berms.
feasible. Examples where wet construction is
warranted include emergency work, placing Highway Traffic Barriers (precast concrete) normally used
as separators for highway construction may be used in
materials in areas where water is too deep or combination with turbidity curtains or heavy plastic to
swift for dewatering or diversion, and work isolate work zones. Barriers are available in 81–107
limited to sites along the shoreline (gravel centimeter (32–42 inch) heights and can be butted
bars). Wet construction also includes situations together to form the necessary isolation length. Plywood
can be anchored against the barriers to make them higher.
where equipment is driven across a flowing Barriers may be placed and moved using common
channel for accessing dry areas on the other excavators and loader equipment and chains.
side. When working in such conditions,
Turbidity Curtains are flexible, impermeable barriers used
regulatory authorities may require impacts on
to trap sediment in water bodies. Curtains are generally
habitat and water quality be avoided or weighted at the bottom to ensure that sediment does not
minimized by fish exclusion (see Section 8.7.2, travel underneath, and are supported at the top by
Fish Exclusion), minimizing construction flotation units that are integrated within the curtain.

Large Wood National Manual July 2015


8-15
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

barrier. Careful monitoring of turbidity


Pumps are the most frequently used piece of equipment
to re-route water from the construction area. There are a
downstream of the work zone is important to
variety of pumps that a contractor may use, dependent on ensure environmental regulations are not
the amount of water that is required/needs to be violated. Because rewatering must be done
removed, frequency, volume, regulatory requirements, slowly, time requirements must be considered
etc. Pumps can either be run using diesel or regular gas
in the sequencing schedule of the contract.
and may be submersible or placed on dry land.

Gravity Bypass may be performed using a combination of 8.5.2.5 Stream Crossings


some types of cofferdam at the upstream end of the
project site, where the cofferdam has an outlet into a In many regulatory environments, construction
flume, pipe, trench, or temporary channel to allow the
operations must be planned to avoid impacts
water to bypass the site using free-flowing conditions.
associated with fording channels with
equipment. An overall stream crossing plan
8.5.2.4 Rewatering should be prepared within the water
management plan to make use of opportunities
Some projects may require care when provided by existing crossings, fording
rewatering portions of the channel that have locations with hard bottoms, and strategically
been dewatered or placed in slackwater status located temporary bridges. Temporary bridge
for large wood construction. When streamflow construction should be considered to limit
is returned to these areas, a pulse of turbidity impacts when multiple stream crossings are
may occur. If such a release of sediment and required for equipment access and material
turbidity violates the applicable permit(s), site staging (Figure 8-5). Temporary bridges may be
rewatering should be done very slowly to constructed using log stringers, steel plates, old
minimize scour and resuspension of sediments. railroad cars, or large concrete blocks and span
The dewatered zone can be flooded and allowed from 6 to 30 meters (20 to 100 feet) (combining
to stand under quiescent conditions for a period multiple smaller spans). Costs are dependent on
of time prior to full restoration of streamflow. both the length of the required span and design
Filtration or sediment retention ponds may be load. If contractors design temporary bridges,
used for mitigating rewatering impacts. they may be required to submit design
Infiltration ponds are simply excavated holes in drawings prior to construction.
the floodplain where the substrate is sand or
gravel material. A filtration pond can be built so 8.5.2.6 Sequencing Plan
it is hydraulically connected or disconnected to A sequencing plan detailing water management
the isolated work zone. The turbid water from activities should be developed early in the
an isolated work zone is pumped into the construction process to facilitate
filtration pond to filter out the sediment-laden communications among the project
water by infiltrating into the sediments that stakeholders: designers, sponsors, inspectors,
make up the pond boundary and discharging by contractors, and regulatory agencies. Water
gravity into an off-channel area (backchannel, management details should be integrated
wetland, slough area). throughout excavation and large wood
Another technique is allowing controlled entry placement plans with emphasis on the initial
of upstream clean water into the isolated, and final stages of those activities. Sequencing
turbid work zone and metering out the turbid plans may address activities such as
water at the downstream end. Controlled  Barrier installation
opening up of the upstream/downstream
isolation area can be done by strategically  Fish removal
removing sand-bags or opening up a corner of a  Excavation

Large Wood National Manual July 2015


8-16
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

 Large wood placement or holes excavated below projected maximum


scour elevation and then partially secured by
 Sediment and turbidity control
backfilling around the large wood with ballast.
 Rewatering Overlying large wood is secured to the key
 Re-engaging isolated work zones and members, and additional ballast is often placed
putting the design feature “on-line” with the within the constructed matrix of large wood.
active river Excavated sediments may or may not be
suitable for ballast, and, if not, a specified type
Figure 8-5. Examples of Temporary Bridges of backfill material may need to be imported to
Constructed for Large Wood Projects provide structural stability.

Figure 8-6. Minimal Excavation for Placing First


Layer of Large Wood

Little Topashaw Creek, Mississippi

Excavation can present some of the biggest


implementation challenges. Excavation
approaches vary according to site conditions
such as bed material, vegetation, foundation
rock, hydrology, and the presence of utilities,
infrastructure, and cultural resources.

The use of appropriate equipment and skilled


Photos courtesy of Tracy Drury, Anchor QEA, LLC. personnel is critical. Using inadequately sized
equipment, the wrong type of equipment, or
8.5.3 Excavation unskilled operation is hazardous to human
safety and environmental resources. Additional
Some, but not all, large wood designs require considerations regarding excavation are listed
excavation of the channel bed or banks to in the Guidance box that follows.
secure placed wood against fluid forces and to
prevent undermining by local scour. Excavation
may be limited to minor grading to smooth the
bed prior to wood placement, excavation of key
trenches into the bank for burying boles of key
members (Figure 8-6), or excavation of deep
pits for burial of key members. In the latter
case, key large wood members are placed in pits

Large Wood National Manual July 2015


8-17
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

access path to the site.


GUIDANCE
Depth to Water—Excavation below stream surface
or water table level poses challenges in release of
Considerations for Excavations for turbidity as well as placement of large wood. When
Instream Large Wood Placements placing wood below the water line, equipment must
hold the wood in place to resist buoyant forces while
Depth of Excavation—Excavation deeper than additional topping logs, ballast, or backfill is placed
1 meter typically requires laying back side slopes to on the lower log layers.
1H:1V or less, or use of trench boxes when
personnel must enter the pit. Soil stability is critical
for deeper excavation due to the risk of sidewall
collapse. Excavation depth should be specified
8.5.4 Wood Placement
during the design phase based on projected scour
depth and anchorage requirements. See Chapter 6, 8.5.4.1 Large Wood Configurations
Engineering Considerations. Configuration of emplaced large wood is usually
Site Hydrology—Dry sites are optimal for excavation. selected in the design phase.
Excavations that intercept the groundwater table are
at greater risk for sudden side slope collapse and
development of boils. Site dewatering prior to CROSS-REFERENCE
excavation is recommended in high groundwater
conditions. Chapter 6, Engineering Considerations, provides
Existing Soils—Coarse aggregate soils, if dry and detailed information on available options for
uncompacted, will require laying back of excavation configuring large wood emplacements.
side slopes to prevent side wall collapse. Stiff, highly
compacted sands or sand/gravel mixtures can
Large wood configurations in current use vary
typically retain steep faces. Geotechnical exploration
using borings or trenches are recommended for
in complexity from single pieces to complex
excavations over 1.5 meters (5 feet) deep. structures with dozens of large logs, rootwads,
and hundreds of smaller pieces (Table 8-3).
Backfill—Clay soils are difficult to compact when Pieces with rootwads tend to be more stable
used as backfill. Sand is highly erodible until (Braudrick and Grant 2001), but are much
stabilized with vegetation. When native soils are not
harder to transport and manipulate. The
suitable, larger sized materials may be imported.
Sometimes existing soils can be blended with some
stability of individual pieces is related to the
amount of imported aggregate to improve stability ratio of their length to channel width,1 and
for backfilling and compaction. some scientists have found lengths as great as
2.5 channel widths are needed for long-term
Location of Equipment—Equipment should not be stability. Others have found values of this ratio
located at tops of slopes, which are potentially
closer to 1.0 for geometrically complex channels
subject to failure unless slope stabilization measures
have been implemented, such as use of trench
with higher rates of natural large wood loading
boxes. Equipment should have sufficient boom (e.g., Bocchiola et al. 2006).
length to reach at least the bottom of the excavation
pit when located safely back from the edge of
excavation.
1 Stability of individual wood pieces is related to
Existing Vegetation—Excavation location should, many factors in addition to the length/channel width
where possible, be selected to preserve existing ratio, and many of these are region- or site-specific.
mature vegetation, with the edge of excavation Among these are the presence of rootwads, branches,
remaining outside of the drip line of trees. When the ratio of log diameter to depth, channel sinuosity,
designing the project, identify existing mature cross-sectional shape, riparian vegetation, and many
vegetation both at the excavation site and in the other factors. However, relations between mobility
and wood length consistently arise in all regions.

Large Wood National Manual July 2015


8-18
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Table 8-3. Configurations for Instream Large Wood Placement

Type of
configuration Advantages Disadvantages Illustration
Individual large Less expensive, Less stable, less hydraulic
wood piece easier to transport and habitat complexity
without rootwad1 and for hand crews
to place, more
maneuverable

Individual large More stable, more More expensive, harder to


wood piece with hydraulic and place with hand crews,
rootwad habitat complexity. usually requires heavy
Soil in rootwad equipment to place and
provides ballast operator skill

Grouping Added stability by Added cost, generally


interlocking, added requires some excavation or
complexity, anchoring, increased
increased potential for
likelihood of adverse/unanticipated
engagement channel response

Log matrix High surface Aesthetics, limited benefit as


complexity and bank armoring
roughness, simple
to construct

without the use of large machinery. The wood


8.5.4.2 Small Wood for these installations can be harvested locally
and should be staged no farther than 9 meters
Small wood placement is typically installed in (30 feet) from the construction site. Depending
smaller systems and provides complexity, on the high flows of the system, the wood can
diversity, and overhead cover. In these smaller
be generally placed on the surface of the
systems, the introduction of a small amount of channel and intertwined with other wood
wood can provide a significant amount of adjacent to or in the channel to help maintain
habitat value. Large wood in this context is
stability.
considered to be 10–30 centimeters
(4–12 inches) in diameter and no longer than
6 meters (20 feet). The installation of wood is
fairly easy and can be completed by hand crews

Large Wood National Manual June 2015


8-19
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

8.5.4.3 Oblique Logs


Oblique logs are small (15- to 30-centimeter
GUIDANCE
[6- to 12-inch) diameter logs that are wedged,
at off-vertical (oblique) angles into the gaps of a Uses for Slash in Instream Large Wood Projects
logjam and used to interlock other members
and reduce vibration during high flows. Oblique Base Layer—When placed prior to wood, slash
provides prime juvenile fish habitat if submerged
logs are placed in a manner to pin underlying
(below base flow water surface elevation).
key logs against vertical pilings and under the
ballasted weight of overlying key logs. Backfill Retention—A layer of slash underlying
Placement of oblique logs is a strategic part of backfill helps retain overlying alluvium from
the construction process and done to bleeding out if undercut by scour.
complement and not interfere with placement Soil Amendment—Alternating layers of soil and
of successive layers of logs, slash, and ballast. slash during backfill provides an organic soil
amendment that will help retain soil moisture.
8.5.4.4 Slash After slash decays, it provides nutrients but will be
a nitrogen sink until it does. In order to place slash
Wood slash refers to branches, limbs, twigs, and layers, place a lift of loose slash at least 61–91
other residue left over after a tree is felled centimeters (24–36 inches) thick and then track or
(Figure 8-7). Slash may be used for erosion wheel compact with heavy machinery so that the
control and as a component of large wood layer is 8–10 centimeters (3–4 inches) thick.
structures, as described in the Guidance box Alternate with layers of soil that are 46–60
that follows. Slash used in large wood projects centimeters (18–24 inches) soil thick.
should be that generated from large wood Final Ground Surface Layer—This layer protects soil
harvest and transport to the placement site; no surface from sunlight (lowering soil temperatures),
extra slash should be extracted. increases surface roughness if subjected to
overland flow, retains moisture, and limits wind-
Figure 8-7. Pile of Slash Available for Use in Large blown dust.
Wood Project
In Large Wood Structures—Use slash to fill large
voids and supplement racked logs. Incorporating
slash into the face of the wood structure creates a
more hydraulically diverse condition and provides
refuge for juvenile fish.

Special Cases—When using dolosse (concrete


jacks), steel piles, or other artificial materials, loose
slash offers an excellent way to improve aesthetics
by covering over exposed artificial elements.

Large Wood National Manual July 2015


8-20
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

8.5.5 Securing Wood


Properly securing the wood installation is CROSS-REFERENCE
probably the most important task in an
instream large wood construction project, and a
Chapter 6, Engineering Considerations, provides
wide range of approaches are possible (Table
detailed information on selecting and designing
8-4). In most cases, the methods to be used are
restraints for placed wood.
selected and details specified in the design
phase of the project.
By properly installing and securing the wood,
this will prevent future property damage, loss
of life and possible litigation.

Table 8-4. Comparison of Methods for Securing Instream Large Wood

Method Technique Advantages Disadvantages Illustration


Burial Trenching/ Precise Cost of excavation
backfill. See placement. and challenges
discussion of Embedment in working below
ballast highly cohesive waterline
requirements (e.g., lacustrine
below. clay) material
has been
effective
(Southerland
and Reckendorf
2010)
Pinning Rebar/dowels Inexpensive way Holes from rebar
to use small create a weak point
wood to create a that tends to rot out.
larger structure; Leaves behind relict
wooden dowels steel. Wooden dowels
are have little strength
biodegradable against rotational
forces

Lashing Manila rope, Bundle If structures wash


cable, chain small/mid-size away, they may not
together to act break up and
as large therefore more
members. Quick readily get tangled up
to install. Manila in downstream
degrades in few structures. Cable and
years chain leave relict
steel. Loose cable can
be safety hazard to
boaters.

Large Wood National Manual July 2015


8-21
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Method Technique Advantages Disadvantages Illustration


Tethering Cable or chain Simple and low Limited habitat
(Nichols and cost benefit as single logs;
Sprague also, tethered logs
2003) often come to rest
above the baseflow
level and provide no
aquatic habitat at
lower stages; added
safety risk due to
tether, no
redundancy for
stability
Mechanical Helical, rotating Large holding Leaves behind relict
anchors plate force with small steel; difficult holding
(Shields et al. anchor in some alluvial soils;
2008) time consuming to
install

Pile- Driven or placed Smaller Subsurface


supported in excavated excavation obstruction; piles
structures1 holes and refilled footprint; quick must be driven deep
(Abbe and (the latter installation, enough to avoid
Brooks 2011) required for relatively low scour
placing piles cost; high
with rootwads); stability, adds
sharpen piles for redundancy
quicker driving when
incorporated in
larger structures
Entanglement On-bank trees No additional Dictated by existing
on bank anchoring trees; requires large
trees2 required wood that is longer
than ~2.5 times the
channel width for
permanent stability

Gravity Structure (wood No additional Structure height must


anchorage + ballast) is anchoring or be great enough that
heavy enough to manufactured it is not submerged
resist imposed materials during design event;
forces during required; not feasible at many
design flows natural sites
appearance

Large Wood National Manual July 2015


8-22
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Method Technique Advantages Disadvantages Illustration


Woven Hybrid of pile- No additional Structure height must
supported and anchoring or be great enough that
gravity; manufactured it is not submerged
horizontal logs materials during design event;
are entangled required; not feasible at many
with vertical natural sites
piling logs; appearance.
vertical piles
used to
counteract Photo by Ken DeCamp
horizontal forces
and ballast to
counteract
vertical forces
Unanchored Placing wood No anchoring Safety concerns, may
directly in required dislodge in
system unexpected flows;
requires large wood
that is longer than
~2.5 times the
channel width for
permanent stability

1 Photo used by permission of Office of Response and Restoration, National Ocean Service, National Oceanic
and Atmospheric Administration.
2 Photo used by permission of Long Tom Watershed Council, Eugene, Oregon.

applications of cable, chain, or rope should be


8.5.5.1 Cable, Chain, and Rope carefully considered to ensure that these
materials do not create public safety hazards.
Restraining devices such as steel cable, chain,
and rope are commonly used for either
8.5.5.2 Mechanical Anchors
temporary or permanent large wood restraint
(Table 8-4). As temporary features, these Earth anchors in unconsolidated material such
materials are used during the construction as alluvial streambed or banks rely on skin
process to secure wood to heavy equipment for friction and surcharge on the cable and passive
transport and installation and to secure wood in earth pressure acting on the anchor—all of
a particular location or logjam until additional which depend on the material remaining static.
ballast or wood members are placed. All
If the streambed or bank erodes, an earth
temporary restraining devices should be
anchor will fail. If wood attached to a cable
viewed with caution due to corrosion and wear
moves, so will the cable and the alluvium
and should be inspected by the contractor prior
around the cable, destabilizing the earth anchor
to construction commencing. As permanent
and potentially causing erosion of the
features, these materials are used to connect
streambed or bank. Because of these potential
large wood pieces and to connect large wood to
issues, using earth anchors is strongly
anchors, boulders, and trees. Permanent
discouraged, and if they are used, it is

Large Wood National Manual July 2015


8-23
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

recommended to use three or more at each some examples of typical ballasts used in wood
attachment to minimize strain and installations and their advantages and
displacement, similar to cable stays on a tower. disadvantages. Each technique will vary given
specific site conditions and the availability of
8.5.5.3 Ballast material. Interstices in large ballast can be filled
with finer-grained alluvium washed in using
Large wood structures are often stabilized by water jets supplied by pumps and hoses. Jetting
filling the spaces within the wood members
in the backfill improves consolidation, increases
with alluvium. Because native alluvium (e.g., ballast weight, and provides better media to
that removed for excavation in preparation for support vegetation planted within the structure.
large wood placement) tends to be mobile at
high flows, ballast must often be imported,
which can increase costs. Table 8-5 presents

Table 8-5. Ballast Materials for Instream Large Wood Structures

Ballast Advantages Disadvantages Illustration


Native alluvium1 Reuse onsite material Can wash out of
lower costs structures (Shields et al.
2004; Southerland and
Reckendorf 2010)

Imported rock2 Select material Additional cost to


appropriate for job3 purchase and haul in;
nonnative to site

Rock collars Quick placement and Poor quality rock can


flexible to shatter when drilled;
accommodate multiple match cable strength to
log configurations ultimate loading
Dolosse4 Complex shape helpful Artificial material;
to entrap wood; shape aesthetics poor; limited
of unit inherently manufacturing
stable; can be colored locations and sources
and textured to imitate
wood

1 Arrow in photograph indicates alluvium placed as ballast.


2 Illustration shows Sulphur Creek, Redding, California. Used by permission of John McCullah, Salix Applied
Earthcare, Redding, California.
3 Sizing and selection of imported rock is beyond the scope of this document. Guidance for use of quarry

stone riprap for erosion control in stream channel is provided by many authorities including Brown and
Clyde (1989), U.S. Army Corps of Engineers (1994), and Escarameia (1998). Natural rounded stone is
aesthetically superior to angular riprap for many applications, but tends to be less stable because it does
not interlock. Use of rounded stone may call for larger stone or thicker layers of stone.
4 Illustration used by permission of Pierce County Public Works and Utilities, Surface Water Division,

Washington.

Large Wood National Manual July 2015


8-24
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

8.5.6 Finish Work implementation of the blankets, the blankets


must be securely held down with no bare soil
An essential element in the process of installing showing. Additionally, a variety of methods to
wood is cleaning up the site and completing the securely hold down the blankets are available,
finish work and preparing the site for plant such as using staples and stakes. Stakes and
material. An important factor when staples are also made using a variety of
demobilizing from the site is to try and not materials such as wood, metal, and
leave the soil compacted in the construction biodegradable material and should be selected
area itself or the access route to the site. Wet based on the site constraints and long-term
soils and seep areas should be avoided when maintenance considerations.
operating vehicles, and when these areas must
be affected, they may be protected with slash, 8.5.6.2 Revegetation
matting, or logs.
Compacted soils are difficult for vegetation to Natural large wood accumulations produce
penetrate the compacted soil and become gradual changes in stream corridors by
established. If compaction does occur, it is providing sites for terrestrial plant colonization.
important to scarify the soil with the equipment Heterogeneous floodplains develop as woody
on site or return to the site and un-compact the plants sprout and mature on sediments trapped
soil with a different approach. on or within large wood formations, trapping
and recruiting more large wood and sediment.
The weight of trees growing on large wood adds
8.5.6.1 Soil Amendments and
ballast to counteract buoyancy, drag, and lift.
Erosion Control Large wood projects often attempt to emulate
To ensure successful plant propagation, adding these natural processes by including provisions
soil amendments can help in the long-term for planting fast-growing riparian species in
stability for the site. Ideally, the contractor exposed soils and sediments adjacent or within
should blend in approximately 46 centimeters newly constructed large wood structures.
of organic material with the native soil. Once
Revegetation may proceed by inserting or
the final grade has been established and the site
burying dormant cuttings of adventitious
has been planted, the application of arborist
species such as willow (Salix spp.) (Figure 8-8);
chips will aid in the retention of moisture in the
using nursery-grown bare root, potted, or
soil and can potentially slow erosion on the
burlap-wrapped specimens; or seeding. In some
disturbed area.
cases plant materials may be harvested or
Erosion control for the site is a construction salvaged from stands on the project site, and
activity that requires diligent implementation some projects have successfully transplanted
and constant maintenance to ensure willow tree root balls. Selection of plant species
compliance with a variety of regulators. One of and propagule should include consideration of
the more common applications is the use of site hydrology, soil conditions, sun exposure,
erosion control blankets, which will provide existing plant material, and long-term
2–5 years of erosion control before the blankets maintenance and management goals. Plants
biodegrade. Depending of the degree of erosion should be robust under anticipated hydraulic
probability, the blankets come in a variety of loads; deeper-rooting species may be needed to
sizes and weights to mitigate the degradation of withstand higher velocities. Soil tests should
the site. Typically, the variety of the blankets is precede planting, and amendments and
contingent on the frequency and size of the mulches used as necessary. Plants are sensitive
woven matrix. To ensure successful to soil density and drainage, so overcompaction

Large Wood National Manual July 2015


8-25
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

and grading that leave plants flooded should be roots are exposed to direct sun. Soaking rooted
avoided. Moist, well-drained soils are optimal. stock is not recommended, but 1 to 2 hours of
immersion immediately prior to planting is a
Figure 8-8. Planting Willow Cuttings in Recent common practice. Bare-rooted or burlap-
Sediment Deposits Adjacent to Placed Large wrapped stock should be heeled into damp
Wood Using Water Jetting
ground or mulch while awaiting final
installation. Cuttings must be planted with the
same vertical orientation they have grown in,
and bundling and marking of cuttings should
proceed accordingly.

During the period of establishment, irrigation


may be needed. Also, fencing to deter herbivory
by wildlife (e.g., beaver) or livestock is often
required, although some evidence indicates
moderate beaver herbivory is not deleterious to
willows (Li et al. 2005; NRCS 2007d). Vandalism
is also a potential problem in populated areas
(FISRWG 1998).
Little Topashaw Creek, Mississippi

Soaking cuttings in well-oxygenated water for 8.5.7 Typical Construction


up to 21 days prior to planting may be helpful Equipment
(Schaff et al. 2002; Martin et al. 2004). Some
evidence indicates that survival rates may be While the use of mechanized equipment allows
higher for larger-diameter cuttings (Greer et al. restoration professionals to complete site work
2006). More complete treatment of plant faster and on a larger scale, the use of “light”
material installation is provided by FISRWG construction techniques is often required to
(1998), NRCS (2007d), and Fischenich (2001). complete a variety of wood installments. Some
Goldsmith et al. (2014) present a series of installations require access to remote sites that
useful case studies. Key principles include are not accessible by vehicles or are located in
reliance on local or regional expertise and environmentally sensitive areas where access is
native species. Control of exotic species may be limited and alternative construction methods
included in the project or may be required to are required.
eliminate competition for plantings.
8.5.7.1 Light Construction
Plant materials impose severe constraints on
implementation scheduling (FISRWG 1998). Hand Placement
Cuttings should be dormant when harvested
and planted, and they should be planted within Human-powered labor is a useful tool when
a day of harvest or completion of the soaking access is extremely limited or site conditions do
period. Ideally, planting schedules should allow not allow vehicles or construction machinery
time for establishment and rooting between (Figure 8-9). Typical applications are along
planting and high-flow season, but this may not small, sensitive streams. Of course, hand
be possible. placement is not adequate for dealing with large
quantities of excavation or fill, pieces of wood
Plant materials should be kept moist between or rock that are heavier than a few hundred
harvest and planting. Rooted stock is prone to pounds, or working in water more than 0.6 to
drying, particularly if pots or burlap-wrapped 1.0 meters (2 to 3 feet) deep.

Large Wood National Manual July 2015


8-26
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Figure 8-9. Manual Labor Team Stockpiling Large Figure 8-10. Belgian Draft Horses Moving Large
Wood Prior to Stream Installation Wood for Instream Placement

Some workers report good results using draft


animals (e.g., Belgian draft horses) to enhance
hand labor teams (Figure 8-10). However, use Photo courtesy of the British Columbia Ministry of
of draft animals requires skilled handlers and the Environment.
access for horse trailers. When the local Figure 8-11. Cable Yarding Large Wood for
regulatory agency permits, onsite thinning of Transport to Channel
wood along the stream banks can be used to
drop whole trees directly into the channel. Riley
(1998) provides detailed instructions for
implementing stream stabilization measures,
many of which include wood and plant cuttings
with hand labor.

Cable Yarding
Mechanical systems such as overhead cables,
grip-hoists, or pulley systems can be useful for
transporting wood from staging sites into the
stream channel. This practice is referred to as Cable system feasibility depends on site
“cable yarding,” and may be used to transport topography, and requires anchor trees or
wood overhead or drag it along the ground towers at both ends to which the cable tight line
(Figure 8-11). Overhead cable systems can can be anchored. Typically, cable yarding is a
typically be used for lengths of up to 91 meters manual process, but using small machines can
(300 feet) each, and log weights up to speed up the process and facilitate work with
1,360 kilograms (3,000 pounds) may be larger wood. It is important to note that cable
accommodated. trading can be extremely dangerous and
consulting with an expert in the profession
prior to starting work is recommended.

Walking Excavators
Walking excavators or “spiders” provide
unequaled performance in difficult terrain such

Large Wood National Manual July 2015


8-27
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

as steep slopes or in river beds. The four larger than this, a larger excavator will typically
hydraulically adjustable legs allow for the be required.
excavator to position the machine in a variety of
ways, unlike other excavators. Some other 8.5.7.2 Heavy Equipment
benefits of walking excavators include very
small turning radius, large lifting forces, and Characteristics of typical equipment used in
compact loading dimensions. Conversely, large wood work in the Northwest are provided
in Table 8-6 below.
walking excavators are slow to move, have
difficulty moving through soft soil, and are not
ideal for transporting materials.
Trucks
Dump trucks are used to transport material to
Small Excavators and around the site. Wood may be hauled in
semi-end dump trucks on highways and in off-
The use of small excavators or “minis” is
road articulated dump trucks elsewhere. Dump
particularly useful when the site will not
trucks come in a variety of sizes and are
accommodate larger equipment or the scope of
the project does not warrant the expense. Minis typically identified by volumetric capacity: five-
are typically rubber-tracked and can access yard or ten-yard dump trucks are common.
When used in large wood placement, a ten-yard
remote sites and environmentally sensitive
areas with little disturbance. Minis are limited dump truck can transport logs shorter than
to excavating in soils (not rock) and cannot drag 6 meters (20 feet). Highway hauling restrictions
typically limit log length to 10 meters (32 feet).
weights greater than about 1360 kilograms
(3,000 pounds). They typically are useful for Trash haulers and trucks with flatbed trailers
lifting and relocating materials weighing up to may be used for logs longer than 6 meters
about 544 kilograms (1,200 pound)s; this is (20 feet) and for logs with large rootwads. Logs
equivalent to a 6-meter (20-foot) long log with are also often delivered via self-loading log
an average diameter of 46 centimeters trucks. Logs with rootwads do not fit well in
(19 inches) or boulders with a 76-centimeter logging trucks.
(30-inch) average diameter. If materials are

Large Wood National Manual July 2015


8-28
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Table 8-6. Examples of Heavy Equipment Used in Large Wood Installation Including Machine and Lift
Weights as Appropriate

Typical
Approximate Approximate Lift
Machine Weight Capacity
Machine Examples (1,000 lbs) (1,000 lbs) Illustration
Tracked CAT 330 75 8
excavator with Kobelco SK300
thumb

CAT 350 105


Kobelco SK400

Rubber tired CAT 930K 30


wheel loader
with forks1

Rotating grapple
sometimes used
on excavator
arm instead of
bucket with
thumb

Mini excavator

Tracked-type CAT D8R 83


tractor
(bulldozer)

Walking or Schaeff HS 41 M 18 ~11


“spider”
excavator

Large Wood National Manual July 2015


8-29
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Typical
Approximate Approximate Lift
Machine Weight Capacity
Machine Examples (1,000 lbs) (1,000 lbs) Illustration
Log skidder CAT 525 28 6

Tracked log Thunderbird 100 5


loader2 840

Low-boy truck, A variety of


logging truck truck types are
utilized.

Crane National 456A 20

Photo by Ken DeCamp


Pile driver Bermingham Attachment to
Foundation crane
Solutions B21
Diesel Hammer

Excavator-
mounted
vibrator

Photo by Ken DeCamp


1 Permission pending for use of image. Obtained from: http://www.cat.com/en_US/products/new/
equipment/wheel-loaders/small-wheel-loaders/18262632.html.
2 Illustration shows log loader with tires, not tracks.

Large Wood National Manual July 2015


8-30
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

an access road, thus reducing cost and site


Tracked vs. Rubber Tire Machines disturbance.
Most types of heavy machinery are available in
Bulldozers
either tracked or rubber-tired versions. Rubber-
tired vehicles are best suited for compact Bulldozers are useful for moving large amounts
driving surfaces such as well-established access of soil. Dependent of the size of the bulldozer,
roads. Rubber-tired machines tend to have less they can be utilized in large wood placement to
effect on native soils and vegetation but can be excavate the area where the wood jam will be
limited in off-road capabilities. Tracked placed, temporarily stage excavated material,
machinery is best suited for sites with and complete the finish grading for many
challenging access conditions. Due to the construction sites.
surface area of tracks (relative to rubber tires),
tracked machinery also exerts less pressure on Cranes
native soils.
Large cranes can be used to move single pieces
or bundles of large wood. Cranes are often used
Excavators
to move other large objects such as culverts,
Excavators are the most often-used type of pre-cast control structures, machinery, and
equipment in large wood placement. Excavators construction materials. Construction cranes can
come in a variety of sizes, and selection for a be mobile or stationary. Crane mobility is
given project depends on large wood size. The related to size and lift capacity with large cranes
smaller excavators are more nimble, while requiring firm soil conditions and having
larger machines can handle heavier material limited off-road capabilities. Crane rental can be
but are somewhat cumbersome on the site. very expensive, and their utilization should be
understood early in the design process and built
Most excavators are tracked, allowing them to
into the construction budget. Stationary cranes
travel well on variable terrain. Ground pressure
typically remain on site for the duration of the
for tracked vehicles is lower than equivalently
project.
sized equipment with tires, reducing site
damage and soil compaction. For large wood Pile Drivers
projects, an excavator may be fitted with a
bucket with a “thumb,” or device that allows the Pile drivers are usually attachments for cranes
machine to hold onto materials while being that include a heavy weight placed between
placed. Instead of a bucket with a thumb, some guides so that it is able to freely slide up and
machines use a rotating grapple that allows for down in a single line. The weight is repeatedly
more precise placement of material. dropped on the head of a pile to drive it into the
ground. Pile drivers are useful for placing
Log Loaders vertical piles to anchor large wood structures
and formations. Piles may be driven into the
Log loaders are an efficient means to transport
bed prior to placement of horizontal large wood
logs from source or staging areas to
members or driven through an existing large
construction zones in or near the channel. Most
wood jam or structure. The feasibility of pile
log loaders have a straight arm assembly that is
driving is strongly tied to subsurface conditions
optimized for log handling whereas excavators
because buried large wood, bedrock, or
have a curved arm assembly that is optimized
boulders can severely hamper pile driving.
for digging. Tracked and wheeled loaders are
available, and both types can traverse rough Pile driving requires less dewatering and
terrain with little disturbance. Neither require disturbance than excavation for large wood

Large Wood National Manual July 2015


8-31
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

structure foundation member burial. Noise (hammer) into the timber pile. The
levels are similar to that for other types of recommended diesel hammer for large wood
construction equipment operation when timber projects is the DELMAG (D-12) hammer due to
piles are used. Piling materials include steel its size, reliability, and availability. Use of larger
beams, sheet pilings, pipes, or concrete. Only hammers may result in impact cracking or
timber piles are recommended for river shattering.
restoration applications. Piles should be
Hydraulic-powered hammers operate similar
untreated, green-harvested large logs (30–60
to diesel hammers but have cylinders stocked
centimeters [12–24 inches] in diameter) and
with hydraulic fluid rather than a compressed
often have one end sharpened to a point.
air-fuel combination. Hydraulic hammers are
Leads can be one of two types: hanging or fixed. not as efficient, cost effective, or available as
Detailed information on lead systems can be diesel hammers. However, hydraulic hammers
found at: http://www.delmag.com/lead- can often be less noisy and decrease concerns of
systems.html. air/water pollution.

Hanging leads typically are more versatile and Vibratory impact drivers (“Vibros”) operate
do not need a level surface. Swinging leads using counter-rotating weights that are
provide more flexibility, but need a spotter on powered by hydraulic motors. Although
the ground and a lot of head room for the crane vibratory hammers are often used for driving
to position them. A typical set of swinging leads hollow piles or sheet piles through fine
is approximately 18 meters (60 feet) long and sediments, they are not recommended for large
60 centimeters (24 inches) square. For driving wood projects.
timber piles for large wood projects, it is
Alternative equipment includes devices
recommended to use a hanging lead system for
designed for mounting on excavators. Use of
flexibility and to help provide bracing and
excavators removes the need for crane
support for the timber pile itself during driving
mobilization. Excavator-mounted vibratory
operations.
drivers can work well in finer sediments up to
Fixed leads are typically only used in situations gravel size, but may not function well in cobble.
where increased precision control is needed for Other alternatives have also been used to drive
detailed positioning by the operator or to timber piling, but are not recommended. Among
reduce ground support of spotters. The fixed these are conversion of excavator-mounted soil
leads are usually mounted to an excavator compaction vibrators and direct use of
boom or other rigid controlled machine. excavator boom to push down timber piles.

Pile driving equipment can be placed into four


categories (Table 8-7): (1) diesel hammers,
(2) hydraulic hammers; 3) vibratory hammers;
4) alternative equipment.

Diesel-powered hammers are the most


common, and are very efficient and effective at
installing timber piles below the river bed, even
in difficult conditions like coarse cobble
substrate or in flowing water. Diesel hammers
operate with a piston-cylinder apparatus using
an air-fuel compression-impact-combustion
energy combination to drive the piston

Large Wood National Manual July 2015


8-32
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Table 8-7. Comparison of Pile-Driving Methods

Hammer Advantage Disadvantage Remarks More information


Diesel Commonly Potential Recommended http://www.delmag.com/
hammer available, low cost, environmental issues option for driving diesel-pile-hammers.html
high strength to for air quality, timber piles on delmag.com/technical-
weight ratio, can be increased noise level large wood projects data.html
used in cobble
substrate
Hydraulic Decreased noise Decreased strength Reasonable http://www.apevibro.com/
hammer level, decreased and efficiency, higher alternative to diesel ver2/products/hih/default.
risk to air quality cost hammers asp
Vibratory Very little noise Difficult to install per This is not http://www.apevibro.com/
impact level pile—mounting recommended ver2/products/vibro/defau
hammer required lt.asp
Excavator- Removes need to May not work for http://www.movax.com
mounted mobilize crane cobble beds
vibrators

CASE STUDY

Driving Piles to Secure Large Wood Structures

On the Trinity River in Northern California, restoration practitioners found that the diesel hammer was the most effective
at driving timber piles into coarse sediments in moving water. The large wood design called for foundation piles installed
approximately 3 meters (10 feet) below grade to allow for scour. Over ten piles needed to be installed as the vertical
members to the large wood structure. Excavation and backfilling were not feasible due to regulatory requirements for
maintaining navigability and holding the turbidity below 20 NTU at 152 meters (500 feet) downstream.

The construction sequence is illustrated in Figure 8-12.

1. Clean gravel fill was placed in the large wood footprint approximately 0.3 meter (1 foot) above the water level to
serve as a staging pad.

2. The planned location for each pile was surveyed and marked on the gravel pad.

3. An excavator was used to dig a pilot hole for each pile through the gravel pad and 0.3–0.6 meter (1–2 feet) into the
bottom of the river bed.

4. Piles were inserted in the holes and backfill was placed to hold them vertical.

5. A crane and DELMAG D-12 diesel pile-driver hammer were mobilized to the project site. The crane was positioned
strategically to be able to logistically reach each of the piles from one central location. The diesel hammer and fixed
leads were lowered onto each timber pile and were driven according to conventional pile-driving protocols, except
that the excavator boom was used to stabilize the fixed leads to ensure vertical placement during initial blows for
each pile.

The diesel hammer was able to drive the timber piles 3 meters (10 feet) into the river bed through coarse cobble in
approximately 30 minutes or less per pile. Positioning the piles prior to crane mobilization allowed the crane and diesel
hammer to mobilize, install the piles, and demobilize in one full day. The approximate cost per day for the crane, diesel
hammer (D-12), and crew was around $7,500 (2014).

Large Wood National Manual July 2015


8-33
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Figure 8-12. Sequence for Constructing Large Wood Structure With Vertically Driven Piles Used to Secure
the Structure

a. b.

c. d.
Trinity River, California. (a) Gravel pad placed over river bed at site for large wood structure. Holes excavated
in pad with track hoe, logs (piles) inserted, and holes backfilled to stabilize piles. (b) Crane for driving piles
mobilized to site. (c) Crane driving piles with diesel hammer. (d) Gravel pad removed to allow flow around
placed wood and attendant vegetated bar. Completed large wood formation functioning as bar apex jam. All
photos by Ken DeCamp.

access and in-water placement. On the other


8.5.7.3 Helicopter Construction hand, they are relatively high cost, have limited
ability to work with extremely large loads, and
Helicopters (“aerial cranes” or “skycranes”) are require increased planning and coordination as
used to lift heavy loads with long cables or well as special safety expertise for the ground
slings and have been used in the logging crew. Helicopter routes must avoid active
industry for decades. They are useful for
roadways and residential areas when
delivering imported wood and other materials transporting material.
and maneuvering wood for final placement
(Figure 8-13). Helicopters offer advantages of
low site impact and rapid construction. They
are especially useful for sites with difficult

Large Wood National Manual July 2015


8-34
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Figure 8-13. Use of Helicopter to Transport Large


Wood to Remote Project Site GUIDANCE

Improving the Efficiency and Cost-Effectiveness of


Helicopter Operations

 Provide a design booklet for the pilot and


ground crew. The booklet should be 5.5 by
8.5 inches (or similar) and show basic
schematics for each site design and be arranged
to show the order of operations for the project.

 Pre-arrange in-flight design terminology.

 Have the wood numbered, laid out and grouped


to coincide with the construction sequencing.

 Review the project construction sequencing and


safety hazards at daily pre-flight safety
meetings.

Requirements for Project Design


Drawings
Construction drawings for large wood projects
involving helicopters must address several
issues that are normally not taken into account.
For example, to eliminate lost air time, plans
may specify staging areas and refueling stations
and contain lists of exact coordinates for the
delivery or placement of materials. Additionally,
the construction sequence should be defined in
detail. This will allow for efficient staging and
streamline helicopter operations.

Ground Crew
Ground crews for helicopter placements are an
important component. Ground crew personnel
are generally provided by the helicopter
contractor and are trained in operating and
(a) Helicopter delivering wood. (b) Use of hand safety procedures specific to this machinery.
signals by ground crew to coordinate with pilot. Crews composed of two to four persons will be
stationed at both the loading and unloading
Timing and Costs areas and responsible for communicating with
Using helicopters for implementing large wood the pilots, connecting cable chokers to the
projects can be highly efficient and cost controller yoke, helping position large wood,
effective. However, few firms have the required and collecting cables at the unloading area. For
expertise, and they are often unavailable during projects with in-water placements, care should
forest fire season (summer). be taken not to exceed depth and velocity
criteria for safe wading.

Large Wood National Manual July 2015


8-35
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Most helicopter companies working in North a 2010 fatality rate of 73.7 deaths per
America provide the option of moving wood 100,000 workers, or about 21 times higher than
using either chokers or various types of the overall population fatality rate There is a
specialized grapples. Grapples are more large body of regulations and supporting
efficient at moving one or two large logs at a documents dealing with safety in the logging
time, whereas chokers can move several and construction industries, and no effort will
smaller logs but require ground crews to hook be made here to reproduce all of it. Highlights
and unhook the logs. Grapples allow the pilot to from key topics will be introduced. The single
select, arrange, and place logs with greater most comprehensive document in this topic
precision compared to chokers, and are safer area is Engineer Manual 385-1-1 (USACE 2008),
for ground crews because there is no hooking which governs activities by USACE personnel
involved, and the crew can maintain a greater and contractors and is a valuable information
distance from the load and flying debris. resource for others. Safety guidance for logging
operations is provided by the U.S. Occupational
See the next section for other safety
Safety and Health Administration (OSHA
considerations associated with helicopter
undated a), the National Institute of
operations.
Occupational Safety and Health (NIOSH 2012),
The large wood project designer (or their the State of Idaho (undated), the Washington
representative[s]) should be on the ground, State legislature (undated), the Southwide
working either in direct communication with Safety Committee (2010), and USACE (2008).
the pilots, or through the helicopter company’s
It is advisable to include provisions in large
ground crew, to direct final placements. This
wood placement contracts that require Job
role will depend on the designer’s experience
Hazard Analyses (JHA) for each distinct phase
and physical ability to traverse the terrain, and
of work. The JHA should be prepared and
the helicopter company’s safety policies.
submitted by the contractor and approved by
the government prior to beginning work on the
8.6 Safety relevant phase. Furthermore, the contractor
may be required to develop a site-specific safety
Successful implementation of safety plans is a and health plan prior to starting work. The plan
hallmark of effective projects and contractors. should cover all aspects of on- and offsite
Personnel implementing instream large wood operations and activities associated with the
projects incur hazards associated with logging, contract, and include noise monitoring and
construction, and amphibious operations. Such material safety data sheets for activities
operations often occur in non-ideal weather. requiring hazardous materials. Generic safety
The synergy of these hazards heightens the plans do not fulfill these requirements. The
importance of safety issues in implementation. project safety and health plan should provide a
Standard practices such as furnishing first aid list of the JHA anticipated throughout the
kits and training, holding regular safety project and a statement that additional JHA will
meetings, and complying with applicable local, be provided as required as the project
state, and federal laws and regulations should progresses.
be followed and will not be detailed here. A
safety and health section from an actual large 8.6.1 Potential Safety Issues
wood placement contract is provided as
Appendix A-4. Public agencies and other responsible and
interested parties may be concerned about both
Logging has consistently been one of the most public safety and assumption of liability that
hazardous industries in the United States, with may be associated with large wood projects.

Large Wood National Manual July 2015


8-36
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

The level of concern, and, as a result, the level of Although recreational activities vary greatly
effort to address potential issues, needs to be among different streams, in general, several
driven by the actual level of risk. The number types of recreation may occur, such as fishing,
and types of users on any given stream or river swimming, wading, rafting, kayaking, and inner-
can vary dramatically when compared to other tubing. In addition, water skiing, personal
rivers. Some rivers have millions of users watercraft use, and recreational power boating
during a single year while others may have are popular activities on some larger rivers,
close to none. In either case, public safety especially during the summer and fall months
should be a strong driver for project design. when water levels are generally at their lowest
However, the level of analysis and method of and large wood structures are most exposed.
accommodation should reflect the level and Incorrect placement of large wood can increase
types of use. One way to address the public potential safety hazards to recreationists,
safety concern is to document the pre- and post- especially for swimmers, waders, water skiers,
project conditions and use the design process to and personal watercraft users.
identify all potential hazards, systematically
evaluate them, and ensure that the design has 8.6.2 Potential Best
minimized the level of risk to an acceptable
level.
Management Practices
Public safety in the broader sense, and Designers and installers should consider best
specifically recreational safety, is a primary management practices when installing large
consideration during the design and wood to minimize the potential for
construction phases of instream projects that compromising recreational safety. Each project
include the installation of large wood. It is site is different, and the site-specific details
important to consider recreational safety need to greatly influence if and how these
throughout the project development process to recommendations are incorporated.
ensure that public safety is maintained over the
life of the project (MTZ Associates 2000).

Large Wood National Manual July 2015


8-37
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

GUIDANCE

Best Management Practices

1. Avoid installing large wood in such a way that large, single branches project far out into the river channel
where they can create a hazard to boaters; placement of large, rigid woody structures in strong currents has
the greatest potential to present a hazardous condition.

2. Where applicable, install trimmed large wood in such a way that finer-textured material projects above the
water surface at low flows; this provides recreationists with visual cues of the presence of shallowly
submerged structures.

3. Do not install large wood where there is limited opportunity for river users to recognize and avoid submerged
structures (i.e., hazards are greatest when there is limited approach visibility).

4. Orient large wood downward toward the water and in a downstream direction (i.e., between 45 and 90
degrees relative to a line running perpendicular to the channel) to minimize hazards to swimmers and
waders.

5. Do not place large wood near bridge piers and crossings of other infrastructure because complex velocity
patterns are often associated with these structures, and large wood at these sites can create special hazards
to recreationists.

6. If materials such as cable or chain are used to secure large wood (e.g., into rock revetment), minimize the
length of cable or chain that is exposed above the rock revetment to avoid creating a tripping hazard.

7. Secure cable ends under rock revetment or near trunk sections to minimize exposure of the public to sharp
objects.

8. Approach visibility is a critical indicator for river users and should be considered in all aspects of project
design, including selection of the location of large wood placement, approach velocities under a variety of
flow scenarios, and signage at entry points and sufficiently upstream to warn oncoming river users.

9. In various locations around each large wood structure (e.g., entry points, upstream of the structure), install
warning and/or interpretive signage panels to advise the public of presence. Warning signs should be very
specific about the risks and strategies for avoidance. Interpretive sign panels should describe the functions of
a large wood structure, native fish and fauna that utilize wood structures, and precautions boaters and
recreationists should take when near a large wood structure.

Large Wood National Manual July 2015


8-38
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

8.6.3 Personal Protective maintained condition (particularly brakes and


throttles), using equipment that provides
Equipment structural barriers such as bulkheads to protect
Personnel involved in large wood construction operators, and staging operations to provide
activities should be protected “from head to visibility to operators. For example, information
toe.” Clothing and protective equipment such as the following should be provided:
(personal protective equipment [PPE]) is no “Truck drivers shall be in the clear and in view
substitute for hazard control at the source. For of the log unloader operator before forks are
example, hearing protection is not to be moved into the load or against it, before a lift is
substituted for functional mufflers on made. All persons are prohibited from standing
machinery. Employees should wear clothing under, or near, the ends of logs being lifted or
suitable for the expected weather conditions moved,” and “All workers shall be in the clear
and work conditions. At a minimum, clothing and in view of the machine operator before a lift
should include short- or long- sleeved shirts, is made.”
long pants, leather boots, and hard hats.
Employees should be trained in the use and 8.6.5 Excavation and Earth
adjustment of protective equipment, and Moving
nonfunctional or damaged equipment and attire
should be destroyed or discarded. Basic Alluvial soils and sediments are quite
requirements are described by USACE (2008) heterogeneous. Bearing capacity or slope
and are summarized in Table 8-8. stability may change drastically between two
points separated by only a few feet, leading to
8.6.4 Log Handling hazards for equipment or even personnel on
foot (Figure 8-14). Geotechnical stability is also
Handling large wood is extremely dangerous. a consideration for excavation, as saturated
Very large wood pieces are often required as alluvial soils tend to be weak and prone to slope
key members, and use of large wood with failure. Safety standards for construction
rootwads is attractive for functionality in many excavation and trenching have been
applications. However, the asymmetrical mass promulgated by OSHA (undated b). Excavation
distribution of these elements leads to deeper than 1 meter typically requires laying
unpredictable behavior when they are being back side slopes to 1H:1V or less, or use of
moved about. Clearances between personnel trench boxes when personnel must enter the
and large wood under transport should be pit. All trenches 1.5 meters deep or greater
conservatively large to allow for these shifts require a protective system unless the
and movements. excavation is made entirely in stable rock
(OSHA undated b).
Additional detailed guidelines for log handling
are provided within the logging safety sources
cited above. In general, these resources
describe the necessity for proficient equipment
operators, keeping equipment in well-

Large Wood National Manual July 2015


8-39
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Table 8-8. Personal Protective Equipment and Attire for Large Wood Project Implementation

Category of Gear Examples Appropriate Uses and Remarks


Head protection Hard hats, helmets Hard hats needed whenever heavy
equipment is present
Eye and face protection Safety glasses, goggles, face shields Glasses with added shields for side
protection; face shields needed for
chainsaw operation
Hearing protection Disposable, preformed, or custom- When noise exposure exceeds 85–90
molded ear inserts, ear plugs, ear A-weighted decibels (dBA) for a
muffs time-weighted 8-hour exposure; see
USACE (2008) for exposure limit
details and computation
High visibility apparel Apparel meeting American National Workers in proximity to heavy
Standards Institute/International equipment or vehicles
Safety Equipment Association
(ANSI/ISEA) 107-2004 Performance
Class 2 requirements
Chaps Protective leg chaps meeting Chainsaw operators; workers
American Society for Testing and potentially exposed to poisonous
Materials (ASTM) Standard F1897 snakes can be protected with snake
chaps or knee-high snake boots
Gloves Hand gloves designed to protect from Important when working with metal
cold, poisonous plants, cuts, abrasions, cables and chainsaws
punctures, burns, and chemical
irritants
Personal flotation devices Inherently buoyant Type III, Type V Whenever working on floating plant
work vests, or better U.S. Coast or over or adjacent to water such that
Guard–approved personal flotation a drowning hazard exists; see USACE
devices that are international orange (2008) regarding the use of auto-
(or orange/red) or ANSI 107 yellow- inflatable devices
green in color
Protective footwear Safety-toed boots meeting ASTM Whenever on a work site
Standards F2412 and F2413

Large Wood National Manual July 2015


8-40
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Figure 8-14. Log Skidder Mired in an Isolated under hovering aircraft except while hooking or
Deposit of Highly Plastic Clay in a Stream Bed unhooking loads. Communications and signals
between helicopter crews and ground
personnel must be clear, continuous, and
unambiguous.

Figure 8-15. Construction Laborers Work to


Secure Fabrics Around Large Wood Toe
Placements on the Outside of a Meander Bend in
a Shallow Channel

Little Topashaw Creek, Mississippi

Great care is needed to ensure safety of


personnel working near heavy equipment. High
visibility apparel and PPE, scrupulous attention
to communication, and maintaining clear lines
of sight for operators are essential. For all but
the smallest projects, it is advisable to prepare
an internal traffic control plan (Roadway Safety
Alliance [undated]). Key safety principles for Note the spotter (orange hard hat) in visual
construction site management include the contact with both the equipment operator and
laborers. Source: Inter-Fluve.
following:
 Separate on-foot workers from equipment 8.6.7 Chainsaw Operation
as much as possible.
Chainsaws are efficient, but extremely
 Design the work space and operations to dangerous, tools. All types of power saws
eliminate/minimize backing and blind should be kept sharp and in good repair at all
spots. times. All exhaust parts on power chainsaws
 Train workers and equipment operators on should be constructed and maintained so the
communication methods. operator is exposed to a minimum amount of
fumes and noise. Chainsaws should not be
 When necessary, use a spotter so the
operated from unstable water craft or floating
vehicles do not run over workers or back
plant or while standing in water. Guidelines for
into other vehicles (Figure 8-15).
the safe use of chainsaws are widely available
and include the following (USACE 2008):
8.6.6 Helicopters
 Chainsaws shall have an automatic chain
Helicopter operations generate safety hazards brake or kickback device.
in addition to those associated with lifting
 The idle speed shall be adjusted so that the
equipment due to the danger of moving
chain does not move when the engine is
propellers, the effects of propeller wash, and
idling.
the great heights of helicopter lifts. Propeller
wash can dislodge treetops, tree limbs, and  Operators will wear proper PPE. Eye, ear,
other objects high above and endanger hand, foot (safety shoes), and leg protection
personnel below. Employees should not work are required as a minimum.

Large Wood National Manual July 2015


8-41
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

 Chainsaws will not be fueled while running, sediment-related water quality impacts.
while hot, or near an open flame. Saws will Guidelines for construction best management
not be started within 3 meters of a fuel practices are widely available (e.g., Fifield
container. 2011), and a good basic summary from a
European perspective is provided by Scottish
 The operator will hold the saw with both
EPA (2009).
hands during all cutting operations.
 A chainsaw must never be used to cut above
the operator’s shoulder height. CROSS-REFERENCE

 Chainsaws shall have sprockets and drive


Section 8.2, Regulatory Compliance and Public
end of the bar adequately guarded. Idler Involvement, provides more information on required
ends, when used as two-man saw, shall also water quality protection measures in large wood
be guarded. projects.
 Combustion engine-driven power saws
shall be equipped with a clutch. Saws with Permit regulations for most wood installations
faulty clutches shall not be used. require special care be taken to prevent harmful
 Combustion engine-driven power saws chemical spills from occurring during
shall be equipped with an automatic construction. Common requirements include
throttle, which will return the motor to replacement of hydraulic fluids in heavy
idling speed upon release of the throttle. machinery with food-grade vegetable oil,
pressure washing machinery prior to arriving
 Power saw motors shall be stopped while on site to remove debris and chemicals, and
being fueled. maintaining spill kits kept onsite during
construction. In addition, either the contractor
or owner is often required to submit a spill
8.7 Managing prevention plan for approval by the regulatory
Environmental agency prior to construction.

Impacts 8.7.2 Fish Exclusion


Large wood projects are intended to
rehabilitate environmental resources Fish exclusion refers the removal of fish from
the work area to allow for continued survival.
associated with stream corridors. It is therefore
incumbent upon those who implement these Detailed guidelines are provided by the
projects to do so in ways that minimize Washington Department of Transportation
(2012). Some regulatory authorities require
collateral environmental damages. Sections
above may be consulted for methods for exclusion of live fish from work areas prior to
procuring, transporting, and placing large wood instream large wood construction. This
in ways that reduce impacts. During requirement is most common in streams with
construction, actions described below may be anadromous fishes in the Pacific Northwest, and
used to minimize impacts on water quality and fish exclusion is virtually unheard of in many
ecological resources. other places. Excluded fish are typically
relocated in reaches adjacent to the project
area.
8.7.1 Water Quality
Fish exclusion is scheduled so that fish are
Construction in stream corridors require removed prior to complete dewatering or
permits that specify erosion controls to limit initiation of construction below the water line.

Large Wood National Manual July 2015


8-42
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Prior to removal of fish, block nets are placed


GUIDANCE
around the construction area to keep removed
fish from re-entering the work area. The basic
Fish Capture for Exclusion
idea for fish capture is to concentrate in areas
where they can be easily seined and netted. Seining–Most common method used. Large nets are swept
Complex cover structures such as long culverts across the bed by teams of people holding each end.
Seines are pursed by drawing the ends together and then
can present a challenge to fish exclusion
retained partially in the water while fish are removed with
operations. For example, it may be appropriate dip nets. Seines with a “bag” minimize handling stress and
to place block nets at the ends of culverts. Once are preferred. Small mesh sizes are more effective across
most or all of the fish have been removed from the full range of fish size (and age class), but also increase
other parts of the work area, block nets may be resistance and can make deployment/ retrieval more
difficult in flowing waters. Seines with a small mesh size in
removed to encourage volitional downstream the bag (or body) and a larger, less resistant mesh size in
movement of fish. the wings offer a compromise.

To be most effective and to minimize stress and Baited Minnow Traps—Typically used before seining. Traps
risk of injury to fish (including stranding), in the should be inspected at least four times daily to remove
captured fish and minimize predation within the trap.
Pacific Northwest, regulatory agency personnel Predation risk to juvenile salmonids is greater at night
coordinate fish exclusion operations with plans from large sculpin.
for dewatering or flow diversion. Plans for
Dip Netting—Commonly used in conjunction with seining;
dewatering and/or flow diversion proceed at a nets are particularly effective during gradual dewatering or
measured pace (within constraints), to flow diversion. Once netted, fish should remain in water
encourage the volitional downstream until transferred to a bucket, cooler, or holding tank. Dip
movement of fish, and reduce the risk of nets that retain a volume of water (“sanctuary nets”) are
preferred method to transfer fish but may be ineffective
stranding. The directing biologist monitors the
unless flow velocity is low. When water depths are very
dewatering process to ensure that water is shallow or fish are concentrated in very small receding
removed slowly to allow for fish capture and to pools or coarse substrate, “aquarium” nets may be a
preclude stranding. Dewatering or flow better, more effective choice. Use of dip nets in
diversion should not proceed unless there are conjunction with snorkeling, flushing of the cover, or
around the hours of dawn or dusk (i.e., during low-light
sufficient staff and materials on site to capture conditions) can be effective for capturing fish sheltered
and safely remove fish in a timely manner. below cover.
Generally this will require a minimum of two
Connecting Rod Snakes—Connecting rod snakes are
persons (three if electrofishing), but large or composed of wood sections approximately 1 meter
complicated sites may require higher levels of (3.3 feet) in length. They may be used to flush fish out of
effort. stream crossing structures (i.e., culverts).

Electrofishing—Electrofishing or electroshocking is
8.7.3 Cultural Resources commonly performed only when other methods are
impracticable or ineffective. In shallow (wadeable) water,
Because stream corridors have long attracted electrofishing may be performed using hand crews and
human use and activity, they are often rich in backpack-mounted equipment. In deeper water, boat-
mounted electro-shock equipment is used, and boat crews
historical and archeologically significant remove fish with long-pole dip nets. Larger fish (i.e., adult
resources. Assessing potential impacts on and sub-adult fish with comparatively longer spine
cultural resources and avoidance or mitigation lengths) are more susceptible to electrofishing injury than
should be similar to practices for any water smaller fish. As a general rule, electrofishing is not
conducted under conditions that offer poor visibility (i.e.,
resources project. Federal and state regulations
visibility of less than 0.5 meter) due to the potential for
govern these assessments. increased fish mortality.

Large Wood National Manual July 2015


8-43
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Section 106 of the National Historic archaeological site may involve long-term site
Preservation Act requires that federal agencies protection, monitoring, or site excavation and
consult with state and local groups before non- data recovery.
renewable cultural resources, such as
It is important to note that these efforts are
archaeological sites and historic structures, are
usually completed well before implementation,
affected. The Advisory Council on Historic
during the planning phase of the project.
Preservation (ACHP) is responsible for
However, if significant cultural resources or any
developing regulations to enforce Section 106
type of human remains are discovered during
compliance (ACHP 2010). Basically, the law
construction, federal and many state
requires federal agencies to initiate a review of
jurisdictions require immediate cessation of
applicable actions, identify potential impacts on
activities that affect the remains and
significant resources, and explore alternatives
notification of authorities.
for avoiding or mitigating impacts. These
alternatives include site preservation,
monitoring, data recovery, and other actions. 8.7.4 Noise
Approvals must be obtained from the State
Noise is any sound that has the potential to
Historic Preservation Officer or native
annoy or disturb humans, or cause an adverse
American tribe or ACHP, depending on resource
psychological or physiological effect on humans.
details.
Sound becomes noise when it is too loud,
Archaeological sites may be directly affected by unexpected, or perceived as uncontrollable.
construction traffic and excavation or indirectly Most sounds that humans are capable of
affected by soil compaction and erosion. hearing have a decibel (dB) range of 0 to 140.
Diversion of flows may erode banks containing A whisper is about 30 dB, conversational speech
artifacts, remains, or other resources. Routinely, 60 dB, and 130 dB is the threshold of physical
a three-staged approach is followed to comply pain. Human exposure limits are based on
with Section 106: duration, with 90 dB a typical upper limit for an
8-hour exposure. Construction activities
1. Identification of the resources present in
involving heavy machinery generate noise that
the project area through background
may adversely affect workers, nearby residents,
research and a field survey.
or wildlife. Many states and federal agencies
2. If resources are present, evaluation of their have promulgated guidelines and regulations
significance. for construction-related noise management.
These policies and global guidance relevant to
3. Mitigation of impacts on the significant
transportation construction projects are
resources.
provided by the Federal Highway
Mitigation means to alleviate any destructive Administration (FHWA 2006).
impacts the project may have on the cultural
Noise generation on most construction projects
resource. ACHP regulations describe mitigation
is the result of equipment operation, principally
as a consultative process that allows for leeway
diesel engines. In assessing noise generation,
in the actual details. If the affected resource
construction equipment can be grouped into
comprises standing structures, mitigation may
two categories, stationary and mobile.
consist in having them properly recorded to the
Equipment noise can also be categorized as
Historic American Building Survey or Historic
being either continuous or impulse in nature.
American Engineering Record standards (i.e.,
Stationary equipment is considered to operate
architectural drawings or large format
in one location for one or more days at a time;
photographs) or some other standard before
pumps, generators, compressors, and screens
moving or demolishing them. Mitigation for an

Large Wood National Manual July 2015


8-44
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

are typical examples of stationary equipment. In observed to decrease by 65% as it aged


addition, pile drivers are sometimes categorized (Rheinhardt et al. 2012). Dynamic fluvial
as stationary equipment. Mobile equipment systems exhibit a complex pattern of erosion,
includes machinery that performs cyclic deposition, and avulsion in response to large
processes such as bulldozers, scrapers, loaders, wood addition, particularly during periods of
and haul trucks. Newer equipment tends to be higher flows.
much quieter than older equipment due to
design features to intentionally reduce noise. It is important to note that maintenance
Noise mitigation measures include specifying requirements for large wood projects should be
the types of equipment that may be used, assessed differently than for more orthodox
scheduling, limiting travel routes and work river works. Displacement of large wood,
zones, and erecting noise barriers trapping additional wood and sediment,
(FHWA 2006). unanticipated scour, or even avulsion may not
necessitate remediation and may, in fact,
indicate that the large wood is functioning as
8.8 Maintenance and intended by restoring natural fluvial and biotic
processes. Maintenance needs should be
Adjustments assessed based on the functional performance
Ideally, instream and floodplain large wood of the large wood, not its appearance. Some
projects create conditions that foster natural features noted on inspection (e.g., loss of wood
recruitment and retention of appropriate and or ballast) may legitimately motivate
desirable levels of large wood loading and maintenance even though the large wood
channel dynamism, making them self-sustaining remains functional because they indicate trends
and maintenance-free over the long term. In that will lead to project failure if not addressed
reality, such ideal conditions rarely occur, and (Thorne et al. 2014).
the effects of a large wood project will often be
temporary without some level of adjustment, 8.8.1 Three Types of
adaptive management, or maintenance. Wood
that is alternatively wet and dry, especially in
Maintenance
regions that are relatively warm and humid, Guidance for river restoration project planning
decays rapidly (e.g., Shields et al. 2008). and design identifies three types of
maintenance (FISRWG 1998).
CROSS-REFERENCE  Remedial maintenance is triggered by
results of routine inspection. Inspections
should identify and prioritize maintenance
Wood decay rates are discussed in Section 6.4,
needs that are not emergencies but that are
Design Life of Placed Wood, and in Chapter 3,
Ecological and Biological Considerations, Chapter 4, unlikely to be addressed through
Geomorphic and Hydrologic Considerations, Chapter maintenance actions that are already
5, Watershed-Scale and Long-Term Considerations, planned or routine.
and Chapter 7, Risk Considerations.
 Scheduled maintenance refers to activities
that are planned during project planning or
However, even in the Pacific Northwest wood design and for which funds are budgeted.
becomes lighter and more brittle within 3 Scheduled maintenance is typically rare for
large wood addition projects, but can
years of placement (e.g., Thorne et al. 2014).
include (for example) replenishment of
Dry density of large wood in riparian zones of
smaller (“racked”) wood; control of exotic
the North Carolina Coastal Plain was

Large Wood National Manual July 2015


8-45
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

vegetation; planting cuttings of pioneer of anchoring hardware such as cables to


species on recent sediment deposits; and maintain tension as wood decays and shrinks.
reconfiguring wood, stone, or sediment to
Vegetative components often require intensive
redirect impinging flows.
maintenance over the short- to intermediate
 Emergency maintenance requires term to replace dead plantings, provide
immediate mobilization to repair, limit, or irrigation during dry seasons, and combat
prevent damage. It may include measures damage due to herbivory or vandalism as noted
such as replacement of plants that fail to in Section 8.5.6.2, Revegetation.
establish, or repair or replacement of wood
intended to provide bank protection or 8.8.3 Adjustments Based on
channel stabilization. Sources of funding,
labor, and materials for emergency Monitoring and
maintenance should be identified prior to Adaptive
project implementation as part of the
contingency planning process. Plans should
Management
include a strategy for allowing rapid
response to any emergency. Plans also CROSS-REFERENCE
should include a process for obtaining
required permits, access routes for
Detailed guidance for preparing large wood project
emergency construction, and coordination
monitoring and adaptive management plans is
with agencies and utilities that are provided in Chapter 9, Assessing Ecological
responsible for riparian roads, transmission Performance.
lines, and utility crossings.

Once environmental documentation is


8.8.2 Maintenance approved, permits are received, and
Activities construction is completed, the monitoring and
adaptive management phase of the project
Typical maintenance activities include begins. Although completion of the
removing or replacing large wood in environmental documentation and permitting
constructed groupings or structures to maintain process may introduce new requirements that
structural stability and habitat benefits or to require modification of the monitoring and
avoid undesirable local scour of banks or bed. adaptive management plan, by this stage in the
Natural large wood formations often experience project, all of the key elements of the plan
cyclical replenishment of wood from upstream should be approved and in place. The
sources so that the appearance of the jam seems monitoring and adaptive management plan will
more or less static even as most individual have clear criteria stating which elements will
members change. It may be necessary to adjust be monitored, the frequency of monitoring, and
the frequency or volume of supplemental whether performance standards have been met.
upstream wood inputs to achieve similar effects
in regulated systems. Use of balloons (Russell and Holburn 2009) or
drones to obtain images that may be analyzed
Restraining elements such as pilings, anchors, using photogrammetric techniques may be
or ballast should be carefully inspected and efficient for physical monitoring. Additional
replaced or adjusted as needed. Cables can guidance pertaining especially to projects in
become safety hazards and warrant special remote locations is provided by Davis et al.
attention in reaches subject to recreational use. (2001). An example of a thorough post-project
Some project plans call for periodic tightening appraisal is provided by Thorne et al. (2014).

Large Wood National Manual July 2015


8-46
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

As the monitoring is implemented and reports projects based on budget constraints, but the
are written, the stakeholders will use the longer the monitoring periods, the greater the
approved adaptive management plan to probability the project will achieve its
determine any remedial work that must be objectives.
performed. The transfer of “bottom line” type
summaries from monitoring to those
responsible for maintenance and adaptive 8.9 Acknowledgments
management is a key link. Adaptive
Dave Porter (BCI Contracting, Inc.), Travis
management actions should be based on
Sumner (SumCo Eco-Contracting), Dave Lyste
monitoring information. Monitoring may also
(Rachel Contracting), Kim Erion (LKE
lead to modification of maintenance plans and
Corporation), Jon Fripp (USDA NRCS), Tracy
schedules. Few large wood projects will trigger
Drury (Anchor QEA, LLC), and Will Harman
natural processes and plant succession rapidly
(Stream Mechanics) shared their views and
enough to eliminate all maintenance
expertise with the authors of this chapter.
requirements. The length of monitoring and
adaptive management will vary between

8.10 Uncertainties and Research Needs


1. Development of approaches for inducing formation of large wood accumulations rather than
constructing them.
2. Enhanced techniques for rapid revegetation of riparian zones and floodplains.
3. Equitable distribution of risk among project designers, constructors, sponsors, and other
stakeholders.
4. Improved techniques for rapidly and economically driving or inserting piles in streambeds and
banks.
5. Guidance for using drones and webcams to monitor implementation.
6. Development of general principles (“rules of thumb”) for deciding what proportion of project
resources should be reserved for adaptive management activities.

8.11 Key Points


1. Even the best planning and design can be entirely negated by shoddy construction work.
2. A variety of contractual arrangements are available for procuring implementation services. See
descriptions applicable to federal agencies in Appendix A-1. Because construction contracts
often result in litigation and distribution of risk varies with the type of contract, personnel
involved in implementation should seek legal counsel and review of contracting arrangements.
3. Maintaining a daily log is an important part of implementation project management. The log
should include photos from fixed points and notes regarding materials, equipment, personnel,
and conversations with contractors.
4. A wood sourcing plan should be developed during project design to allow adequate lead time.

Large Wood National Manual July 2015


8-47
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

5. Water management is a key issue in implementation. Available approaches may be categorized


as working in the dry conditions, working in wet but controlled conditions, and working in
uncontrolled wet conditions. Working in the dry requires dewatering or diversion.
6. Substantial regional variations exist in requirements for mitigating project impacts on water
quality (turbidity), aquatic habitat, and fish or other organisms. Temporary crossing structures
are sometimes used to reduce heavy equipment impacts on streambeds.
7. Key construction steps include excavation, wood placement, and securing wood. Risks may be
reduced by careful selection of appropriate equipment and methods for each step.
8. Large wood projects are complex from a safety management standpoint because they combine
potential hazards incurred by logging, construction, amphibious operations, and non-ideal
weather conditions. It is advisable to require Job Hazard Analyses for each phase of work and a
site-specific safety plan.
9. The effects of a large wood project will often be temporary without some level of adjustment,
adaptive management, or maintenance.

8.12 References
Abbe, T. B., and A. P. Brooks. 2011. Geomorphic, Engineering, and Ecological Considerations when
Using Wood in River Restoration. Pages 419–451 in A. Simon, S. J. Bennett, and J. M. Castro
(eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools.
Geophysical Monograph Series 194. Washington, D.C.: American Geophysical Union.

Advisory Council on Historic Preservation (ACHP). 2010. Protecting Historic Properties: A Citizen’s
Guide to Section 106 Review. Advisory Council on Historic Preservation. Washington, D.C.

Bocchiola, D., M. C. Rulli, and R. Rosso. 2006. Transport of Large Woody Debris in the Presence of
Obstacles. Geomorphology 76(1):166–178.

Braudrick, C. A., and G. E. Grant. 2000. When do Logs Move in Rivers? Water Resources Research
36(2):571–583.

Brown, S. A. and E. S. Clyde. 1989. Design of Riprap Revetment. Hydraulic Engineering Circular 11,
Publication No. FHWA-IP-89-016, Federal Highway Administration, US Department of
Transportation, Washington, DC.

Cramer, M. L. (ed.). 2012. Stream Habitat Restoration Guidelines. Copublished by the Washington
Departments of Fish and Wildlife, Natural Resources, Transportation and Ecology, Washington
State Recreation and Conservation Office, Puget Sound Partnership, and the U.S. Fish and
Wildlife Service. Olympia, WA.

Davis, J. C., G. Minshall, W. Robinson, T. Christopher, and P. Landres. 2001. Monitoring Wilderness
Stream Ecosystems. Gen. Tech. Rep. RMRS-GTR-70. Ogden, UT: U.S. Department of Agriculture,
Forest Service, Rocky Mountain Research Station.

Escarameia, M. (1998). River and Channel Revetments: A Design Manual. London: Thomas Telford.

Federal Highway Administration (FHWA). 2006. Construction Noise Handbook. Report numbers
FHWA-HEP-06-015, DOT-VNTSC-FHWA-06-02, NTIS No. PB2006-109102. U.S. Department of

Large Wood National Manual July 2015


8-48
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Transportation, Research and Innovative Technology Administration, Cambridge, MA. Available:


http://www.fhwa.dot.gov/environment/noise/construction_noise/handbook/. Accessed: July
10, 2014.

Federal Interagency Stream Restoration Working Group (FISRWG). 1998. Restoration


Implementation, Monitoring, and Management. Chapter 9 in Stream Corridor Restoration:
Principles, Processes and Practices. National Technical Information Service, U. S. Department of
Commerce, Springfield, VA. Also published as NRCS, U.S. Department of Agriculture (1998)
National Engineering Handbook (NEH), Part 653. Washington, D.C.

Fifield, J. S. 2011. Designing and Reviewing Effective Sediment and Erosion Control Plans. Third
Edition, Santa Barbara, CA: Forester Press.

Fischenich, J. C. 2001. Plant Material Selection and Acquisition. EMRRP Technical Notes Collection
(ERDC TNEMRRP-SR-33), U.S. Army Engineer Research and Development Center, Vicksburg, MS.

Goldsmith, W., D. H. Gray, and J. McCullah. 2014. Bioengineering Case Studies. New York Springer.

Greer, E. S., S. R. Pezeshki, and F. D. Shields, Jr. 2006. Root Elongation of Black Willow Stakes in
Response to Cutting Size and Soil Moisture Regime (Tennessee). Ecological Restoration
24(3):195–197.

Idaho Office of the Administrative Rules Coordinator. Undated. Idaho Minimum Standards for
Logging—Sections 17.08.01 through 17.08.16. Available:
http://adminrules.idaho.gov/rules/current/17/index.html. Accessed: July 10, 2014. Industrial
Commission. Idaho Office of the Administrative Rules Coordinator, Boise, Idaho.

Li, S., L. T. Martin, S. R. Pezeshki, and F. D. Shields Jr. 2005. Responses of Black Willow (Salix nigra)
Cuttings to Herbivory and Flooding. Acta Oecologica 28(2):173–180.

Martin, L. T., S. R. Pezeshki, and F. D. Shields Jr. 2004. High Oxygen Level in a Soaking Treatment
Improves Early Root and Shoot Development of Black Willow Cuttings. The Scientific World
Journal 4:899–907.

McMillan, LLC. 2014. Fourth of July Creek Stream Restoration Draft Design Report. Prepared for Pend
Oreille County Public Utility District. February 14, 2014.

MTZ Associates. 2000. River Corridor Recreation Safety Study: Sacramento River Bank Protection
Project. Prepared for the U.S. Army Corps of Engineers, Contract 42E, Sacramento, CA.

National Institute for Occupational Safety and Health. 2012. Logging Safety. Centers for Disease
Control and Prevention. Atlanta, GA. Available: http://www.cdc.gov/niosh/topics/logging/.
Accessed: July 10, 2014.

National Resources Conservation Service (NRCS). 2007c. Chapter 15—Project Implementation. In


Part 654, Stream Restoration Design National Engineering Handbook. U.S. Department of
Agriculture, Washington, D.C.

National Resources Conservation Service (NRCS). 2007d. Technical Supplement 14I—Streambank


Soil Bioengineering. In Part 654, Stream Restoration Design National Engineering Handbook. U.S.
Department of Agriculture, Washington, D.C.

Large Wood National Manual July 2015


8-49
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

Nichols, R. A. and S. G. Sprague. 2003. The Use of Long-Line Cabled Logs for stream Bank
Rehabilitation. Pages 422–442 in D. R. Montgomery, S. M. Bolton, D. B. Booth, and L. Wall (eds.),
Restoration of Puget Sound Rivers. University of Washington Press: Seattle.

Occupational Safety and Health Administration (OSHA). Undated a. Logging eTool. Occupational
Safety and Health Administration. Washington, D.C. Available:
https://www.osha.gov/SLTC/etools/logging/index.html. Accessed: July 10, 2014.

Occupational Safety and Health Administration (OSHA). Undated b. Trenching and Excavation Safety.
Fact Sheet. Occupational Safety and Health Administration. Washington, D.C. Available:
https://www.osha.gov/SLTC/trenchingexcavation/construction.html. Accessed: July 10, 2014.

Rheinhardt, R., M. Brinson, G. Meyer, and K. Miller. 2012. Integrating Forest Biomass and Distance
from Channel to Develop an Indicator of Riparian Condition. Ecological Indicators 23:46–55.

Riley, A. L. 1998. Restoring Streams in Cities: A Guide for Planners, Policymakers, and Citizens.
Washington, D.C. Island Press.

Roadway Safety Alliance. Undated. Internal Traffic Control Plans. Developed under a contract with
the Centers for Disease Control and Prevention contract No. 212-2003-M-02677, Laborers’
Health and Safety Fund of North America, Washington, D.C.

Russell, K., and E. Holburn. 2009. Field Evaluation of Engineered Large Woody Debris for Structure
Performance and Habitat Value. Pages 3234–3243 in World Environmental and Water Resources
Congress 2009. American Society of Civil Engineers.

Schaff, S. D., S. R. Pezeshki, and F. D. Shields Jr. 2002.The Effect of Pre-Planting Soaking on Growth
and Survival of Black Willow (Salix nigra) Cuttings. Restoration Ecology 10(2):267–274.

Scottish Environment Protection Agency. 2009. Engineering in Water Environment Good Practice
Guide: Temporary Construction Methods. First edition.

Shields, F. D., Jr., N. Morin, and C. M. Cooper. 2004. Large Woody Debris Structures for Sand-Bed
Channels. Journal of Hydraulic Engineering 130(3):208–217.

Shields, F. D., Jr., S. R. Pezeshki, G. V. Wilson, W. Wu, and S. M. Dabney. 2008. Rehabilitation of an
Incised Stream with Plant Materials: The Dominance of Geomorphic Processes. Ecology and
Society 13 (2):54. Available: http://www.ecologyandsociety.org/vol13/iss2/art54/.

Southerland, B. S., and F. Reckendorf. 2010. Performance of Engineered Log Jams in Washington
State—Post Project Appraisal. In Joint Federal Interagency Conferences 2010: Book of Abstracts,
446 pp., ISBN 097790027X, Joint Fed. Interagency Conf., 2010, Conference on Sedimentation and
Hydrologic Modeling, June 27–July 1, Las Vegas, Nev., Government Printing Office, Washington,
D. C.

Southwide Safety Committee. 2010. Timber Harvesting Safety Manual. Rockville, MD. National
Timber Harvesting and Transportation Safety Foundation. Available:
http://loggingsafety.com/content/timber-harvesting-safety-manual. Accessed: July 10, 2014.

Thorne, C., J. Townsend, and T. Ashley. 2014. Geomorphic and Ecological Assessment and Evaluation
of Grade Building Structures on the SRS Sediment Plain, North Fork Toutle River Final Report.
Performed for the U.S. Army Corps of Engineers, Portland District, OR.

Large Wood National Manual July 2015


8-50
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

U. S. Army Corps of Engineers. 1994. Engineering and Design: Hydraulic Design of Flood Control
Channels. EM 1110-2-1601. Department of the Army, U.S. Army Corps of Engineers. Washington,
D.C.

U.S. Army Corps of Engineers. 2008. Safety and Health Requirements. Engineer Manual 385-1-1. U.S.
Army Corps of Engineers Headquarters, Washington, D.C.

Washington Department of Transportation. 2012. WSDOT Fish Exclusion Protocols and Standards.
Available: http://www.wsdot.wa.gov/Environment/Biology/BA/BAtemplates.htm>.
Washington DOT, Olympia.

Washington State Legislature. Undated. Safety Standards—Logging Operations. Chapter 296-54


WAC. Available: http://apps.leg.wa.gov/WAC/default.aspx?cite=296-54. Accessed: July 10,
2014. Washington State Legislature. Olympia, Washington.

Large Wood National Manual July 2015


8-51
Bureau of Reclamation and Chapter 8. Regulatory Compliance, Public
U.S. Army Corps of Engineers Involvement, and Implementation

This page intentionally left blank.

Large Wood National Manual July 2015


8-52
Chapter 9
ASSESSING ECOLOGICAL PERFORMANCE

AUTHORS

Leo D. Lentsch (ICF International)


C. Anna Toline (National Park Service)
Willis McConnaha (ICF International)
This page intentionally left blank.
provide fish and other organisms with
9.1 Introduction sustainable wood habitats at the watershed scale
The bulk of evidence presented in the previous over the long term.” In other words, large wood
chapters supports the notion that the addition of enhancement should be viewed as an interim
large wood and large wood structures in restoration measure until natural processes of
streams, as a restoration action, can enhance wood recruitment recover to natural levels.
ecological functions, and generally results in Within this context, assessing the value of
greater abundance and/or biomass of fish and placing wood in a stream channel at a specific
other aquatic species. However, a considerable site while determining the performance of
amount of uncertainty remains associated with restoration actions at reestablishing natural
the use of large wood for restoring function in ecosystem processes and functions will likely
aquatic ecosystems. require different assessment perspectives.

As highlighted in the previous chapters,


CROSS-REFERENCE ecosystems are highly variable and have
inherent uncertainties. As such, resource
See the Uncertainties and Research Needs sections managers often need to accept the reality of
of Chapters 3, Ecological and Biological uncertainty and address it through a structured
Considerations, 4, Geomorphology and Hydrology evaluation process (e.g., adaptive management),
Considerations, 5, Watershed-Scale and Long-Term while minimizing management risks associated
Considerations, 6, Engineering Considerations, with proposed activities (Keith et al. 2011). In
7, Risk Considerations, and 8, Regulatory
other words, while measures can be put in place
Compliance, Public Involvement, and
that help to reduce the uncertainty of
Implementation, for detailed lists of uncertainties
that remain for each area of concern. management decisions, uncertainty and its
associated risks will always be a component of
ecological systems and restoration actions.
The observed response of aquatic ecosystems to
Successful resource managers must be both
wood enhancement can reflect a suite of
flexible, to accommodate uncertainty in future
watershed-level conditions that can obscure the
events, and be able to respond to scientific
effects of site-specific wood restoration.
paradigm shifts associated with new
Engineering solutions that do not account for
information. In this light, each large wood
species habitat needs, stream dynamics,
placement project should be viewed, to some
disturbance regimes, and watershed
degree, as experimental, with a minimum level of
characteristics are often unsuccessful (Beschta
scientific effort dedicated to addressing
1997). Nagayama and Nakamura (2010) found
uncertainties. That is, large wood placements
ample examples of restoration projects that
that produce some ecological benefits but do not
failed, concluding that, “restoration projects
provide some level of learning for future efforts
should be aimed at restoring natural processes
are not successful overall.
of wood recruitment and routing, which can
It is alarming, however, to note the rate at which Bernhardt et al. (2005), who reported that only
evaluation approaches are left out of restoration It is alarming, however, to note the rate at which
projects. For 42% of 50 European large wood evaluation approaches are left out of restoration
projects reviewed, no monitoring occurred (Kail projects. For 42% of 50 European large wood
et al. 2007). Those results were similar to the projects reviewed, no monitoring occurred (Kail
findings of Bash & Ryan (2002), who reported et al. 2007). Those results were similar to the
the lack of evaluation of 47% of restoration findings of Bash & Ryan (2002), who reported
projects in the state of Washington. Most the lack of evaluation of 47% of restoration
notable, however, were the findings by projects in the state of Washington. Most

Large Wood National Manual July 2015


9-1
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

notable, however, were the findings by 9.2.1 Using Best Available


Bernhardt et al. (2005), who reported that only
10% of more than 37,000 projects evaluated Knowledge
incorporated any form of project monitoring, Several federal and state mandates or directives
and little if any of this information was either offer insights to the application of best available
appropriate or available for assessing the science. A number of authors have addressed
ecological effectiveness of restoration activities. this issue (Doremus 2004; Murphy and Weiland
This chapter emphasizes the use of carefully 2010). For example, Murphy and Weiland (2010)
designed approaches and/or experiments that reviewed the incorporation of best available
address key uncertainties associated with large science into the ESA compliance process. They
wood restoration actions, monitoring the effects noted that the ESA, along with the Marine
of the restoration actions, and subsequently Mammal Protection Act and the Magnuson-
directing necessary adjustments. It provides the Stevens Fishery Conservation and Management
foundation for a discussion of monitoring and Act, require federal agencies implementing
research activities necessary to answer actions to use the best available scientific and
questions associated with the placement of large commercial data when making decisions. Under
wood in streams and provides practitioners with ESA, USFWS and NMFS must follow the best
information needed to help guide them to make available scientific data mandate when making
informed resource management decisions. The listing decisions, designating critical habitat, and
information includes: completing the consultation process on
proposed federal actions. USFWS and NMFS have
1. Best Science Practices not issued regulations that interpret the
2. Measurable Outcomes and Performance requirement to use the best scientific and
Indicators commercial data available. However, they issued
a policy statement on information standards
3. Monitoring Approaches
under the ESA in 1994.
4. Research Approaches
Two additional federal statutes provide guidance
5. Decision Making and Choices on the use of best available science. The
Administrative Procedure Act of 1946 provides
parties affected by final agency actions with a
9.2 Incorporating Best means to seek judicial review of those actions. In
Science Practices addition, it requires that a reviewing court set
aside an agency action that is ‘‘arbitrary,
Science plays an increasingly important role in capricious, an abuse of discretion, or otherwise
contributing to how people perceive and not in accordance with law.’’ Under the
respond to restoration of ecosystem processes. Information Quality Act of 2001, the Office of
The current understanding of ecosystem Management and Budget issued guidance to
processes is quite different from that of a few federal agencies to ensure the ‘‘quality,
decades ago. Additionally, constant changes in objectivity, utility, and integrity’’ of information
population growth, land subsidence, catastrophic disseminated by those agencies to the public.
events, and climate change ensure that the future Additionally, the standards in these statutes
will be very different from today. As such, emphasize the importance of transparent
incorporating best science practices is an ever- decision making to allow affected individuals
changing and critical component of any and reviewing courts to determine that federal
restoration project. agencies have considered the full record before
them and have made agency determinations

Large Wood National Manual July 2015


9-2
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

based upon the data, analyses, and findings in processes. Instead, they illustrate a logic path
that record. that links ecosystem components to indicators of
desired species and ecosystem conditions. They
While evaluating the use of large wood in
are useful for management because they can
50 restoration projects, researchers came to
help identify which factors may be important in a
similar conclusions (Kail et al. 2007). They
system, which of these factors may be influenced
emphasized: (1) that it is not possible to predict
by management, and hence which attribute
precisely the effect of restoration measures (Kail
(component or condition) of the system should
et al. 2007; Kondolf 1998), and, therefore,
be assessed.
information from surrogate metrics such as
monitoring stream morphology and biota should Conceptual models can inform the research
be used to obtain information to help make program in several important ways, by:
corrections (Bryant 1995; Kail et al. 2007); and providing a basis from which to test assumptions
(2) monitoring results may provide valuable about the relative importance of certain
information for the improvement of future processes, helping to identify threats or
project designs (Bryant 1995; Kondolf 1995, stressors, identifying species or other attributes
1996, 1998; Bash and Ryan 2002; Downs and that function as ecosystem indicators, and
Kondolf 2002; Bisson et al. 2003; Reich et al. serving as a repository of our changing
2003; Kail et al. 2007). To this end, restoration understanding of the system as more data
projects can be successful in providing valuable become available. Conceptual models can also be
information for the design of future projects, used to communicate the understanding of the
even if the projects fail to achieve some of the system to other scientists and the public and to
performance objectives (Kail et al. 2007: Kondolf facilitate review. For a multi-species, habitat-
1995). It is also important to emphasize the based conservation plan, these types of models
importance of incorporating learning objectives provide a useful framework to help us
into restoration projects in addition to understand how species react to management
performance objectives (Downs & Kondolf 2002; actions. These models must be complex enough
Kail et al. 2007). to capture the relationships that drive the
system and translate these relationships to
9.2.2 Using Conceptual covered species, but must be streamlined enough
to be useful as management and monitoring
Models tools.
Conceptual models describe our current As ecological conceptual models are refined with
understanding of a functioning ecosystem. They data from monitoring and research, the effects of
provide a framework for learning about a system conservation measures and associated
and help formulate hypotheses about cause-and- management actions on ecological parameters
effect relationships. (as identified in monitoring actions) can be more
readily anticipated. The anticipated effects can
CROSS-REFERENCE ultimately be stated as hypotheses and tested
with data from targeted studies and research. In
A detailed detailed discussion of the use of this manner, effects can be systematically
ecological models is included in Chapter 3, Ecological analyzed. From this approach we can increase
and Biological Considerations. our understanding of the system and potential
effects of conservation measures.
Conceptual models differ from quantitative
models in that they do not posit any
mathematical relationship between factors or

Large Wood National Manual July 2015


9-3
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

1. Define the question. Research strategies


GUIDANCE
should be designed to address specific
hypotheses. Conceptual, statistical, or
Restoration actions, intended to offset the effects of
spatially explicit models will define those
anthropogenic activities, can affect species habitat
hypotheses.
through two major pathways:
2. Determine what to measure. Establish the
1. Some habitat restoration approaches focus on
restoring natural processes (e.g., road removal,
attributes or variables that the research will
riparian replanting) and thus affect ecosystem measure to answer the question defined
functions by influencing the underlying watershed above.
processes (e.g., sediment supply, delivery of 3. Develop data collection protocols.
organic material).
Questions to be answered by the research
2. Other techniques focus on manipulating or program can be at the species, community,
enhancing habitats for organisms at specific sites or landscape level. Research protocols will
(e.g., wood placement for cover). Restoration vary with scale and with the target of the
actions should be at a scale commensurate with question.
environmental problems. (Roni 2005)
4. Use indicator species, if appropriate. In
some cases groups of species or indicator
species will streamline data collection.
Indicators are selected because they are easy
9.2.3 Following Scientific to survey and provide usable information on
Principles and the species or system in question.
Guidelines 5. Consider sampling design. Sampling design
needs to be a consideration prior to
Evaluation of restoration actions should be
initiating the experiment. The experimental-
based on scientific principles that guide
management approach requires that
continual refinement of restoration efforts to
questions of site selection, pseudo-
effectively implement a restoration strategy. In
replication, power, and significance be
this way, the adaptive management process is
incorporated, as much as possible.
likely to lead to the development of alternative
management strategies and ultimately the
testing of the effectiveness of those strategies. 9.2.4 Existing Protocols and
Because of this, there is a continuum of Indices
management actions that incorporate scientific
principles to varying degrees. The most basic 9.2.4.1 Protocols
monitoring involves simply assessing effects
once a management action has occurred without When available, scientifically accepted
any replication, controls, or comparison of monitoring protocols that are compatible with
management treatments. At the other end of the measuring the success of a restoration project
spectrum are directed studies that test a should be adopted to facilitate data comparison
hypothesis in a manner that can be validated with other studies. For example, in addition to
through statistical inference. Even simple others, the National Park Service’s Inventory and
experimental methods will yield important Monitoring Program guidelines for monitoring
results to help guide and improve management. protocols (Oakley et al. 2003) or the Bureau of
Land Management’s guidelines (Elzinga et al.
In addition to these scientific guidelines, the 1998) can be used as references for developing
following steps should precede experimental research and monitoring protocols. To be
design.

Large Wood National Manual July 2015


9-4
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

successful, these protocols need to be index has been used throughout the United
appropriate to the task, implemented precisely, States and many countries internationally, and
and as cost-effective as possible. Research and has proven to be a reliable means of assessing
monitoring protocols should be standardized the effect of human disturbance on streams and
(implemented consistently) as much as possible watersheds. As such, it has application for
across restoration projects. Ongoing training assessing the ecological value of restoration
may be necessary to ensure there is consistency activities.
in protocol implementation.
IBI was first developed by Dr. James Karr to help
resource managers’ sample, evaluate, and
GUIDANCE describe the condition of small warm water
streams in central Illinois and Indiana (Karr
 Monitoring and research activities should 1981). The original version had 12 metrics that
incorporate scientific principles of replication, reflected fish species’ richness and composition,
control, and pre- and post-treatment monitoring number and abundance of species, trophic
when feasible. organization and function, reproductive
 Monitoring and research actions should be linked behavior, fish abundance, and condition of
to hypotheses about species’ ecological individual fish. In 1993, Karr developed a
relationships and responses to management Benthic-Index of Biotic Integrity (B-IBI),
actions, when possible. modeled after the fish IBI. The B-IBI included 13
metrics based on benthic macroinvertebrate
 When feasible, research should include an
data collected from rivers in the Tennessee
experimental design with appropriate
Valley (Kerans and Karr 1994).
significance levels (alpha level) as well as
sufficient power to detect effects (beta level). The phrase “biological integrity” comes from the
1972 Clean Water Act, which established
“restoration and maintenance of the chemical,
Research and monitoring protocols can be at a
physical, and biological integrity of the Nation’s
landscape, community, or species scale. The level
waters.” Integrity implies an unimpaired
of detail of data collected will depend on the
condition or quality or state of being complete.
scale and also on the available opportunities for
“Biotic integrity” is based on the premise that the
detecting monitored variables. For example,
status of living organisms provides the most
monitoring protocols will vary by covered
direct and effective measure of the “integrity of
species. For species that are difficult to detect in
water.” As such, IBI provides managers with a
the project area, monitoring may be limited to
technique for evaluating the biological condition
determining whether the species persists from
of the water resource management activities
sample period to sample period, what features
(Teels and Danielson 2001; Karr et al. 1986;
define its habitat, and what threats it faces.
Simon and Lyons. 1995).
Surveys for species that are more readily
detectible may indicate whether the species’
occurrence locations are increasing or 9.3 Measurable Outcomes
decreasing
and Performance
9.2.4.2 Indices Indicators
The Index of Biotic Integrity (IBI) is an example
A key component of a restoration action
of a well-known indexing procedure commonly
evaluation framework is defining measurable
used by academia, agencies, and resource
outcomes and associated performance metrics
managers to assess watershed condition. This
that are directly related to the project objectives.

Large Wood National Manual July 2015


9-5
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

Measurable outcomes can be predicted using solids), temperature, and ionic strength (e.g.,
quantitative models. Each outcome should have conductivity). Chemical parameters include the
at least one associated performance indicator concentrations of dissolved gases, major cations,
(e.g., Carignan and Villard 2002), a target for anions, and nutrients (i.e., nitrogen,
successful achievement of that outcome, a phosphorus).
monitoring program designed to identify
progress toward that target, and decision points
for amending actions if acceptable progress is GUIDANCE
not being made. For the purposes of this manual,
performance indicators are biotic and abiotic Characteristics of Effective Performance Indicators
variables that are selected to facilitate  Relevant to project goals and objectives and can
monitoring of systems or species that are be used to assess the project performance at
otherwise difficult to examine. the appropriate spatial and temporal levels.

Ecological indicators can be used in many ways:  Sensitive to changes in the ecosystem, providing
to predict species richness (MacNally and early warning of response to environmental or
Fleishman 2004), to estimate biodiversity (Kati management impacts.
et al. 2004), to assess levels of disturbance, or to  Indicate the cause of change, not just the
provide targeted information on a system or existence of change.
species (Caro and O’Doherty 1999; Carignan and
Villard 2004). In general, ecological indicators  Provide a continuum of responses to a range of
stressors such that the indicator will not quickly
demonstrate changes or trends that are
reach a minimum or maximum threshold.
quantifiable. Indicators may include a variety of
measures or a single indicator species. An  Have known statistical properties, with baseline
indicator species is an organism whose data, references, or benchmarks available.
characteristics are used as an index of attributes
 Are technically feasible, easily understood, and
too difficult, inconvenient, or expensive to
cost effective to measure by all personnel
measure that relate to other species or involved in the monitoring.
environmental conditions of interest (Landres et
al. 1988).  Can be measured with an adequate level of
precision and accuracy. (Carignan and Villard
Ecological variables or structure-based 2002).
characteristics, such as water inundation depth
and duration are also used as indicators of
performance Some examples of potential Information from these analyses is used to
ecological indicators, in the riverine evaluate a stream’s condition with respect to
environment, include those discussed in the stressors such as acidic deposition, nutrient
following sections. enrichment, and other inorganic contaminants.
In addition, streams can be classified with
respect to water chemistry type, water clarity,
9.3.1 Water Quality mass balance budgets of constituents,
Physiochemical water quality characteristics temperature regime, and the presence of anoxic
affect the ability of species to persist in a given conditions. Examples of relationships between
lotic (flowing water) habitat. Water quality data stream chemistry and watershed-level land use
are collected to determine the acid-base status, data are described in Herlihy et al. (1998).
trophic condition (nutrient enrichment), and
chemical stressors. Physical parameters include
light penetration (e.g., turbidity, suspended

Large Wood National Manual July 2015


9-6
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

9.3.2 Periphyton 9.3.3 Aquatic


Periphyton are algae, fungi, bacteria, protozoa, Macroinvertebrates
and organic matter associated with channel Aquatic macroinvertebrates play important
substrates. They are useful indicators of functional roles in lotic ecosystems and are good
environmental conditions because they respond indicators of stream quality. Aquatic
rapidly and are sensitive to a number of macroinvertebrates represent a fundamental
anthropogenic disturbances, including habitat link in the food web between organic matter
degradation, and contamination by nutrients, resources (e.g., leaf litter, periphyton, detritus)
metals, herbicides, hydrocarbons, and and fishes. Within specific biogeographical
acidification (Banta et al. 2000). regions, aquatic macroinvertebrate assemblages
Periphyton exhibit high diversity and are a major respond in predictable ways to changes in
component in energy flow and nutrient cycling in stream environmental variables. Because many
aquatic ecosystems. Many characteristics of aquatic macroinvertebrates have limited
periphyton community structure and function migration patterns or a sessile mode of life, they
can be used to develop indicators of ecological are particularly well suited for assessing site-
conditions in streams. Periphyton are sensitive specific effects.
to many environmental conditions, which can be Benthic macroinvertebrates inhabit the
detected by changes in species composition, cell sediment or live on the bottom substrates of
density, ash free dry mass (AFDM), chlorophyll, streams. Macroinvertebrate assemblages in
and enzyme activity (e.g., alkaline and acid streams reflect the overall biological integrity of
phosphatase). Each of these characteristics may the benthic community, and monitoring these
be used, singly or in concert, to assess condition assemblages is useful in assessing the status of
with respect to societal values such as biological the water body and discerning trends. Benthic
integrity and trophic condition. communities respond differently to a wide array
A hierarchical framework can be used in the of stressors. As a result of this, it is often possible
development of the periphyton indices of stream to determine the type of stress that has affected a
condition. The framework involves the benthic macroinvertebrate community (Barbour
calculation of composite indices for biotic et al. 1999; Kerans and Karr 1994). Additionally,
integrity, ecological sustainability, and trophic macroinvertebrate community structure can
condition. The composite indices will be sometimes be used as an indicator of past
calculated from measured or derived first-order conditions.
and second-order indices. The first-order indices There are generally two different approaches
include species composition (richness, diversity), being used for developing ecological indicators
cell density, AFDM, chlorophyll, and enzyme based on benthic invertebrate assemblages. The
activity, which individually are indicators of first is a multimetric approach, where different
ecological condition in streams. Second-order structural and functional attributes of the
indices will be calculated from periphyton assemblage are characterized as "metrics."
characteristics, such as the autotrophic index Individual metrics that respond to different
(Lakowicz and Weber 1973), community types of stressors are scored against
similarity compared to reference sites, and expectations under conditions of minimal human
autecological indices (e.g., Lowe 1974; Lange- disturbance. The individual metric scores are
Bertalot 1979; Dixit et al. 1992). Banta et al. then summed into an overall index value that is
(2000) describe the development of a used to judge the overall level of impairment of
multimetric index based on periphyton an individual stream reach. Examples of
assemblages in wadable streams. multimetric indices based on benthic

Large Wood National Manual July 2015


9-7
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

invertebrate assemblages include Fore et al. Amphibians also comprise a substantial portion
(1996), Barbour et al. (1996), and Resh et al. of vertebrate biomass in streams throughout the
(1995). United States (Hairston 1987; Bury and Corn
1991). Reports of dramatic declines in
The second approach is to develop indicators of
amphibian biodiversity (e.g., Blaustein and Wake
conditions based on multivariate analysis of
1990) have increased the level of interest in
benthic assemblages and associated abiotic
monitoring these assemblages. Amphibians may
variables. Examples of this type of approach as
also provide more information about ecosystem
applied to benthic invertebrate assemblages
conditions in headwater or intermittent streams
include the IBI discussed above (Kerans and Karr
in certain areas of the country than other
1994), the River Invertebrate Prediction and
biological response indicators (Hughes 1993).
Classification System (RIVPACS; Wright 1995),
and the Benthic Assessment of Sediment Overall, field sampling is used to collect a
(BEAST; Reynoldson et al. 1995). Rosenberg and representative sample of the aquatic vertebrate
Resh (1993) present several approaches to assemblage by methods designed to (1) collect
biological monitoring using benthic all except very rare species in the assemblage
invertebrates, and Norris (1995) briefly and (2) provide a measure of the abundance of
summarizes approaches to analyzing benthic species in the assemblages (McCormick 1993).
macroinvertebrate community data.

9.3.4 Fish and Aquatic 9.4 Monitoring


Vertebrate Assemblage Roni et al. (2003) emphasize the importance of
monitoring activities associated with ecosystem
Fish and other aquatic vertebrates can indicate restoration projects while providing excellent
stream and riparian quality. Extensive life guidelines on developing and implementing
history information is available for many species, monitoring programs. A well-designed
and because many are high order consumers, monitoring plan includes well-developed
they often reflect the responses of the entire testable hypotheses, data collection and data
trophic structure to environmental stress. Also, management to answer questions, and clear
fish provide a more publicly understandable communication of the results and conclusions.
indicator of environmental degradation. Fish
generally have long life histories and integrate
pollution effects over longer time periods and
9.4.1 Compliance Monitoring
large spatial scales. The purpose of compliance monitoring is to
(1) track progress of project implementation in
The fish assemblage represents a critical
accordance with established timetables, and
component of biological integrity from both an
(2) ensure compliance with terms and conditions
ecosystem function and a public interest
of the project permits. Compliance monitoring is
perspective. Historically, fish assemblages have
required to ensure that avoidance and
been used for biological monitoring in streams
minimization measures are properly carried out
more often than in lakes (e.g., Karr and Kerans
where specific sensitive occurrences of covered
1991). Fish assemblages can serve as good
species (e.g., an active nesting site for a covered
indicators of ecological conditions because they
bird species or a population of a highly restricted
are long-lived and mobile, forage at different
covered plant species) or other risks (e.g.,
trophic levels, integrate effects of lower trophic
sedimentation of a wetland) have been identified
levels, and reasonably easy to identify.
at or adjacent to a construction site.

Large Wood National Manual July 2015


9-8
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

monitoring designs should be based on where


GUIDANCE and when effects are expected to occur (both
spatially and temporally), what organisms are
Monitoring Types
expected to be affected (fish, wildlife, plants,
Compliance Monitoring aquatic invertebrates, etc.), what the expected
benefits are (magnitude, duration), potential
Required by permits; focuses on whether the
mitigating factors (including distribution and
restoration activities are being implemented as
exposure), and how various factors may alter
designed.
exposure and effect.
Effectiveness Monitoring, including
It is anticipated that the extent to which
 Performance monitoring. Identifies whether effectiveness monitoring would occur can be
conservation measures are achieving the reduced over time as causal relationships
expected outcomes or targets. between the implementation of restoration
 Mechanistic monitoring. Demonstrates whether actions and the responses of species and
the mechanisms thought to link a restoration ecosystems to those measures are better
action to the desired outcomes are working as understood (as a result of knowledge gained
predicted. under the monitoring and research activities).
For example, if relationships between the
 System-level monitoring. Identifies the degree of
the program’s success relative to the desired placement of a large wood structure and macro
outcome. This requires a sustained, long-term invertebrate production are established through
commitment to monitoring critical features of monitoring and research on initially restored
the system. channels, then effectiveness monitoring for
assessing the production of macro invertebrates
Long-Term Status and Trend Monitoring
associated with subsequent restoration of a
Used to determine the status and trends of stream channel may be reduced or no longer
ecosystems, natural communities, and species. required. Effectiveness monitoring can also be
spatially stratified and occur long enough to
establish the effectiveness of the restoration
9.4.2 Effectiveness activities in ecologically relevant portions of the
planning area.
Monitoring
As described above, research and monitoring
Effectiveness monitoring assesses ecosystem-,
plans associated with specific restoration actions
natural community–, and covered species–scale
should be considered as part of the
responses to the implementation of restoration
implementation process. These plans should be
actions and monitors progress made toward
reviewed on a regular basis and adjustments
achieving biological goals and objectives.
made in response to new information and/or
Effectiveness monitoring will be closely
identified research needs. Plan implementation,
coordinated with research actions to support
monitoring, analysis, and research are all part of
adaptive management. Evaluating clearly
an overall adaptive management process. This is
discernible change in environmental conditions
not intended to be a stand-alone process, but
is often difficult, due to the multitude of
rather one that integrates information and
interacting factors. For example, it is often not
learning to facilitate decision making, including
clear which environmental component will be
decisions to adjust the design and
affected by a stressor manipulation and what
implementation of restoration actions, and the
type of change will occur. A changing
type and extent of monitoring associated with
environment is natural, and variation due to
those activities.
natural effects may be great. To account for this,

Large Wood National Manual July 2015


9-9
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

9.4.3 Long-Term Status and involved persons. A database and a clear


reporting procedure is also required for permit
Trend Monitoring compliance.
Long-term monitoring is used to determine the
status and trends of ecosystems, natural GUIDANCE
communities, and species. Long-term monitoring
should use the framework developed during the Issues to Consider During Data Analysis
baseline studies to carry out effectiveness
monitoring and to implement adaptive  Availability of sites on which treatments can
be applied.
management.
 Availability of reference sites.

 The site-selection process (i.e., is it random,


GUIDANCE
stratified random, non-random).

Tasks During Long-Term Monitoring  Systematic versus opportunistic sampling.


 Assess status and trends at the landscape and
 Detection probability of the protocol.
natural community levels.
 Monitor species response to enhancement,  Replication versus pseudo-replication
restoration, and habitat creation. (Hurlbert 1984).

 Monitor restoration sites for success.  The clarity of hypotheses.

 Sufficient statistical power (1-≤≥) or


significance level (≤≥).(Scheiner and
9.4.4 Collect, Analyze, Gurevitch 2001)

Synthesize, and
Evaluate Data 9.5 Research and
Collection, analysis, synthesis, and evaluation of
project actions and follow up monitoring are
Experimentation
crucial to improve our current understanding of Adjustments to natural resource management
the use of large wood. Analysis and synthesis actions might entail more than minor corrective
should incorporate how conditions have actions. This may require the need for a
changed, expectedly and unexpectedly, because commitment, most often driven by quantitative
of project actions. Evaluation should address models, for identifying and experimentally
whether the objectives have been met and why. evaluating alternative hypotheses about
In addition to ecological information, the right responses to resource management actions
data can provide valuable information about (Briceño-Linares et al. 2011; Kingsford et al.
non-ecological factors such as project costs and 2011; Van Wilgen and Biggs 2011; Walters
compliance, and efficiency 2002).
Proper data management, analysis, and Management programs associated with
reporting are critical to the success of an ecological restoration have an experimental
adaptive management program. All data and component aimed at improving the performance
metadata related to monitoring methods, results, of restoration actions (Keith et al. 2011).
and analysis must be managed, stored, and made Well-defined experiments, supplemented by
available to Implementation Office staff, expert knowledge, are often applied to evaluate
decision-makers, scientific advisors, and other the assumptions underlying resource

Large Wood National Manual July 2015


9-10
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

management strategies (Rumpff et al. 2011). about key factors for the landscape, natural
Simple experimental designs can go a long way community, and/or species for which the
toward separating resource management action restoration action is applied. Adaptive
effects from other causes of ecological change management actions and monitoring should be
(Mackenzie and Keith 2009). In some cases, low directed toward confirming or disproving those
numbers, small areas, and urgent time frames hypotheses. In this light, targeted studies should
place severe constraints on experimental design. be conducted using an experimental design that
will yield statistically valid results to address
critical uncertainties (see Section 9.4,
GUIDANCE Monitoring).
For species that are sufficiently detectable to obtain
estimates of population size or probability of 9.5.1 Research
detection, monitoring a randomly selected subset of
Research may be conducted to resolve specific
the population to make statistical inference to the
whole population can be achieved through the questions related to the following.
principles listed below:  Key ecological processes.
 Develop and state the assumptions in the  Technologies and methods for effectively
hypotheses and models before collecting implementing and measuring the outcome of
monitoring data or conducting manipulations conservation measures.
such as experiments and adaptive management.
 Development of new and more sensitive
 When designing an experiment or using adaptive indicators and metrics.
management, select the number and location of
sampling units so as to apply sufficient scientific  Increasing understanding of the ecological
rigor for evaluating the hypothesis being requirements of covered species for effective
advanced (although flexibility is needed because implementation of conservation measures.
the number of units required to arrive at a
statistically valid result may depend on the  Modeling and assessing responses of
variability of the characteristic being measured). covered species to conservation measures.

 Use spatial and temporal survey site replicates  Determining causal relationships between
for population estimates and/or those receiving a ecological stressors and drivers and changes
management action/treatment. Use controls in natural communities and covered species.
when appropriate.
 Identifying and evaluating trade-offs among
 Measure the sensitivity of variables to reflect restoration options.
true changes in the resource being sampled.
When appropriate, adjust counts, measures of Research results should be sufficient to help
species richness, and determinations of patch direct and prioritize subsequent implementation
occupancy (i.e., presence/absence) with an of restoration projects though the adaptive
estimate of detection probability as described by management process. Ideally, directed research
Yoccoz et al. (2001) and Pollock et al. (2002). can detect both false negatives and false
positives, yielding statistically valid results. This
type of research should answer specific
In these situations, a succession of trial-and- restoration-related questions that arise from
error evaluations may offer the only practical monitoring results and should address data gaps
insights into how to adjust management and provide information necessary to
strategies (Briceño-Lenares et al. 2011). The successfully implement restoration actions. The
design of targeted studies that address key design of experimental research should be
uncertainties should be driven by hypotheses driven by hypotheses about key factors in the

Large Wood National Manual July 2015


9-11
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

natural community in which management is that an effect of the perturbation has been
applied. Management actions and monitoring detected.
should be directed towards confirming or
disproving those hypotheses. For key
management questions, directed research should GUIDANCE
be tested on a small scale using an experimental
design that will yield statistically valid results. Tasks for Documenting Baseline Conditions

 Inventory and document resources and improve


9.5.2 Before-and-After mapping.

Studies  Use the results of land acquisition assessments


as the first source of baseline data.
Studies intended to evaluate restoration projects
often have failed to include the collection of  Standardize data-collection methodologies and
baseline data prior to the restoration (Anderson nomenclature to facilitate information sharing.
and Dugger 1998; Wissmar and Beschta 1998).  Conduct baseline surveys for plants in areas
Without pre-restoration data as a benchmark for where covered activities may impact plant
comparison with post-restoration data, however, occurrences.
it is not possible to document what changes have
 Research and document historical data and
occurred. Conclusions about achievement of the
trends, as appropriate.
restoration project’s goals, and the success or
failure of the project, are enhanced by an ability  Use baseline data to validate and refine species-
to prove quantitatively and statistically that the habitat models as lands are surveyed and
restored system has changed. Change detection acquired (species models will be updated
relative to a baseline condition is therefore an annually as new, relevant information becomes
important aspect of restoration evaluation. available).

 Conduct post-acquisition biological inventories.


Baseline conditions need to be defined to serve
Additional surveys will be needed to
as a comparison point for future monitoring
supplement data gathered in pre-acquisition
actions. Accordingly, resources of interest that assessments.
occur on a site need to be documented, mapped,
and inventoried.  Conduct post-construction surveys for covered
plants in areas where covered activities may
Before-and-after studies involve measurement of have impacted occurrences of covered plants.
a variable prior to and following a perturbation
 Use aerial photos and ground surveys, as
both at a location that will be affected by the
needed, to assess quality and location of local
perturbation (impact) and in an area that will and regional landscape linkages between
not be affected (control) (Stewart-Oaten et al. unprotected natural areas and adjacent
1992). This approach is analogous to an protected lands.
experimental design in which some subjects
receive a treatment and others do not, although  Collect additional baseline data needed to
refine conceptual models.
true replication in the experimental sense may
not be possible. One classic approach to analysis
proposed by Stewart-Oaten et al. (1992) is to
compare the mean difference between the 9.5.3 Pilot Projects
control and impact area in the before period
Pilot projects can be used to ascertain which
with the mean control-impact difference in the
management actions may ultimately yield the
after period. A significant difference suggests
desired restoration gains prior to initiating a
long-term project. Pilot projects are also a cost-

Large Wood National Manual July 2015


9-12
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

effective way to test restoration actions that can Pilot projects may also be short-term
and should be used during the early phases of experiments or observations that give
project implementation to field-test different information on long-term effects. For example,
management actions. Pilot projects are designed Opperman and Merenlander (2004) evaluated
to evaluate alternative monitoring protocols and the effectiveness of methods to restore riparian
sampling designs and to select the best vegetation along stream corridors. The study
technique for obtaining information. For examined the long-term effects of grazing within
example, if the objective is to quantify wildlife the riparian corridor by comparing historically
use of a corridor crossing, a pilot project may grazed stream reaches to ungrazed reaches.
test the effectiveness of different tracking Although the study was short-term (<1 year), it
methodologies (i.e., comparisons between using provided information on long-term effects of
tracking plates, bait stations, and trail cameras). grazing and led to recommendations on riparian
The results of the projects would then be used to corridor management.
develop long-term monitoring protocols.

Different management techniques should be 9.6 Making Decisions and


implemented and evaluated experimentally. In
some cases, restoration, enhancement, and Choices
monitoring methods are not known or have not
As described in Chapter 2, Large Wood and the
been successfully reproduced on a large scale by
Fluvial Ecosystem Restoration Process, this
land managers or the scientific community.
manual recommends that practitioners adopt a
Before restoration or enhancement through
structured process to design, implement, and
management can occur successfully, these
evaluate restoration projects. As such,
methodologies need to be tested on a smaller
practitioners should consider integrating
scale.
adaptive management into their restoration
Pilot projects designed to test the effectiveness projects. Adaptive management is systematic
of restoration and enhancement can be long- and designed to address uncertainty predicated
term (i.e., 5- to 15-year) endeavors. They can on principles of experimental design. Adequate
inform long-term management and can be data are gathered and statistically analyzed to
included as part of a long-term restoration identify effective alterations to a restoration
program. Results from these types of projects program or project. Even if quantitative data
can guide future restoration efforts. This acquisition is limited, a record of qualitative
feedback can increase the efficiency with which observations can produce information valuable
restoration projects can be managed and the for advancing the state of the art.
overall success rate of the actions. Similar pilot
Adaptive management is a structured approach
projects can be developed as research studies
to addressing uncertainty about the potential
when multiple techniques are intended to
environmental and biological response to
achieve a desired outcome and are appropriate
management actions. The process promotes
for monitoring habitat function within a broader
flexible decision making that can be adjusted
study area. Testing the use of indicators for
based on outcomes of management actions and
ecosystem function or covered species; refining
changes in ecological processes (Holling 1978;
monitoring protocols; establishing control plots
Walters 2002). It requires well-articulated
for long-term management; and reviewing the
management objectives and explicit assumptions
literature for guidance on sampling,
about expected outcomes to compare against
experimental design, and management will all be
actual outcomes (Williams et al. 2009).
a part of research.
Importantly, adaptive management requires
explicit recognition of uncertainties and how

Large Wood National Manual July 2015


9-13
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

these may affect achievement of management of adaptive management to aquatic ecosystem


goals. Also, applying adaptive management restoration.
principles requires scientific rigor, including
Adaptive management relies on existing
using models to develop hypotheses about
information to develop and implement
potential resource responses to management
ecosystem restoration actions in such a manner
actions, maintain flexibility in management, and
that new information can be gained and utilized
committing to carry out monitoring and re-
in the process. The process is designed to use
evaluating management goals over time. With
new information to inform a systematic and
these qualities, adaptive management programs
integrated critical review at regular intervals. As
can reduce uncertainty and associated
restoration actions are implemented, the
management risks by improving our
knowledge base is expanded, biological
understanding through monitoring and
assumptions are revised, and changes may be
researching the outcomes of restoration actions.
made to the restoration project objectives and
The challenge in using an adaptive management
associated hypotheses, metrics, targets, and
approach lies in finding the correct balance
monitoring metrics (Figure 9-1).
between gaining knowledge to improve
management in the future and achieving the best
Figure 9-1. Key Components of an Adaptive
short-term outcome based on current knowledge Management Framework
(Allan and Stankey 2009).

One of the earliest applications of the adaptive


management concept for use with natural
resource decisions involved management of
commercial fishing in 1957 (Beverton and Holt
1957). However, it took until the 1970s for the
traditional concept and application of adaptive
management as a natural resource management
tool to be improved and evolve (Holling 1978;
Walters 2002; Pahl-Wostl 1995; Lee 1999;
Oglethorpe 2002). Since that time, it has been
applied to a wide range of resource management
approaches (Walters 2002; Christensen et al.
1996; Stanford and Poole 1996; Oglethorpe
2002; Habron 2003; Kaplan 2008; Lyons et al. Modified from USFWS 2014.
2008; Williams et al. 2009). Many of these
involve water supply management and Decision-makers should reexamine the steps of
ecosystem restoration activities (Poff et al. the adaptive management framework and make
2003), such as the Glen Canyon Dam and the revisions when needed. This may include
Colorado River ecosystem (National Research modifying the goals and objectives, modifying
Council 1999); the Missouri River ecosystem the metrics, applying new and modified
(Prato 2003); USACE water resource project analytical tools and models, modifying
planning (National Research Council 2004); conservation measures, and implementing new
Columbia River system (Vail and Skaggs 2002; or modified monitoring. As described above, the
Volkman and McConnaha 1993); and the targets and criteria used to define a restoration
Everglades ecosystem (Gunderson and Light project should reflect judgments based on the
2006). Lessons learned from these applications, best available science. As the project is
as well as advances in other scientific disciplines, implemented, however, new information may
have greatly improved the utility and application indicate that some of these targets or criteria are

Large Wood National Manual July 2015


9-14
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

less effective at producing desired outcomes. To need to be implemented, such as refining the
allow for flexible and responsive implementation objectives, models, and conservation measures,
of the project, it is important that the plan and potentially selecting an alternative action.
identify decision points that establish the
Fir an adaptive management plan to succeed,
parameters that will be used to improve the
technical staff and decision-makers must be
effectiveness of the project, respond to changing
regularly involved in the exchange of
biological conditions, and/or respond to social
information as data are analyzed and
and economic directives.
synthesized. The information should be provided
Communicating the current understanding of the to those directly involved in the adaptive
results of the restoration action is an important management process as well as all those
step for informing and equipping policy makers, interested in the outcome. The communication
managers, stakeholders, and the public. The should be ongoing and occur at appropriate
information communicated should be technically decision points.
sound, well synthesized, and translated into
formats conducive to informing nontechnical
audiences. If necessary, the communication
should include any potential adjustments that

Large Wood National Manual July 2015


9-15
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

Guidance

Important Considerations
 Does the project meet the expected benefits?
 Is the project consistent with and contribute to fulfilling the restoration plan and objectives?
 Is the project compliant with federal and state law?
 Is the project being implemented within a reasonable timeframe?
 Do the restoration activities have clear, measurable, and achievable end points?
 Does the monitoring plan enable evaluation of the project’s progress and ultimate success?
 Is the cost to carry out and monitor the project reasonable relative to benefits and available funds?
 Are the project’s potential harmful effects on natural resources and ecological services deemed acceptable?
 Is the project resulting in a net benefit or improvement for the environment?
 Have any adverse impacts resulting from the project been fully mitigated by restoring, replacing,
rehabilitating, or acquiring the equivalent of the same or similar resources harmed by the project?
 Has the project benefited multiple species or resources?
 Has the project contributed to an ecologically balanced and integrated approach to restoration?
 Has the project benefited any of the following economic sectors: tourism, fisheries, maritime, and recreation?
 Has the project built community resiliency and benefited communities vulnerable to disasters?
 Has the project addressed underlying sources of environmental stress and provided a long-term approach
and/or solution to restoring natural processes rather than addressing the symptoms of environmental
degradation through short-term fixes?
 Has the project provided long-term ecological benefits commensurate with the investment?
 Has the project enhanced resilience and adaptation of river and stream environments and species with
respect to climate change impacts?
 Does the project represent a restoration approach for which the public has expressed support or would likely
provide support based on previous public comment or input?
 Does the project contain a public education component such as onsite interpretation, signage. or some other
means to inform the public about the project’s importance and results.

Large Wood National Manual July 2015


9-16
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

9.7 References
Allan, C., and G. H. Stankey. 2009. Adaptive Environmental Management. Volume 351. Springer.

Anderson, D. H., and B. D. Dugger. 1998. A Conceptual Basis for Evaluating Restoration Success.
Transactions of the North American Wildlife and Natural Resources Conference Volume 63.
Wildlife Management Institute.

Banta, E. R., M. C. Hill, and M. G. McDonald. 2000. MODFLOW-2000, the US Geological Survey
Modular Ground-Water Model: User Guide to Modularization Concepts and the Ground-Water
Flow Process. Reston, VA, USA: US Geological Survey.

Barbour, M. T., J. Gerritsen, G. E. Griffith, R. Frydenborg, E. McCarron, J. S. White, and M. L. Bastian.


1996. A Framework for Biological Criteria for Florida Streams Using Benthic
Macroinvertebrates. Journal of the North American Benthological Society (1996):185–211.
Available:
http://www.jstor.org/discover/10.2307/1467948?uid=3739936&uid=2&uid=4&uid=3739256
&sid=21104863995063.

Barbour, M. T., J. Gerritsen, B. D. Snyder, and J. B. Stribling. 1999. Rapid Bioassessment Protocols for
use in Streams and Wadeable Rivers. USEPA, Washington. Available:
http://zoology.okstate.edu/zoo_lrc/biol1114/study_guides/labs/lab12/gen_usepa_barbouretal
_1999_rba.pdf.

Bash, J. S. and C. M. Ryan. 2002. Stream Restoration and Enhancement Projects: Is Anyone
Monitoring? Environmental Management 29(6):877¬885.

Bernhardt, E. S., and 24 others. 2005. Synthesizing U.S. River Restoration Efforts. Science.
308(5722):636–637

Beschta, R. L. 1997. Restoration of Riparian and Aquatic Systems for Improved Aquatic Habitats in
the Upper Columbia River Bbasin. Pages 475-491 in D. J. Stouder and P. A. Bisson (eds.). Pacific
Salmon and Their Ecosystems: Status and Future Options. New York: Chapman Hall.

Beverton, R. J. H., and S. J. Holt. 1957. On the Dynamics of Exploited Fish Populations. Fisheries
Investigation Series 2 (19). Fisheries and Food. London: Ministry of Agriculture.

Bisson, P. A., B. E. Rieman, C. Luce, P. F. Hessburg, D. C. Lee, J. L. Kershner, G. H. Reeves, and R. E.


Gresswell. 2003. Fire and Aquatic Ecosystems of the Western USA: Current Knowledge and Key
Questions. Forest Ecology and Management 178:213–229.

Blaustein, A. R., and D. B. Wake. 1990. Declining Amphibian Populations: A Global Phenomenon?
Trends in Ecology & Evolution 5 (7):203–204.

Briceño-Linares, J. M., J. P. Rodríguez, K. M. Rodríguez-Clark, F. Rojas-Suárez, P. A. Millán, E. G. Vitton,


and M. Carrasco-Muñoz. 2011. Adapting to changing poaching intensity of yellow-shouldered
parrot (Amazona barbadensis) nestlings in Margarita Island, Venezuela. Biological Conservation
144 (4):1188–1193.

Bury, R. B., and P. S. Corn. 1991. Sampling Methods for Amphibians in Streams in the Pacific
Northwest. Available: http://www.treesearch.fs.fed.us/pubs/3069.

Large Wood National Manual July 2015


9-17
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

Carignan, V., and M.-A. Villard. 2002. Selecting Indicator Species to Monitor Ecological Integrity: A
Review. Environmental Monitoring and Assessment 78(1):45–61.

Carignan, V., and M. A. Villard. 2004. Biological Indicators in Environmental Monitoring Programs:
Can We Increase Their Effectiveness. Environmental Monitoring:567–582.

Caro, T. Mi, and G. O'Doherty. 1999. On the Use of Surrogate Species in Conservation Biology.
Conservation Biology 13(4):805–814.

Christensen, N. L., and 12 others. 1996. The Report of the Ecological Society of America Committee
on the Scientific Basis for Ecosystem Management. Ecological Applications 6(3):665–691.

Dixit, S. S., J. P. Smol, J. C. Kingston, and D. F. Charles. 1992. Diatoms: Powerful Indicators of
Environmental Change. Environmental Science & Technology 26 (1):22–33.

Doremus, H. 2001. Adaptive Management, the Endangered Species Act, and the Institutional
Challenges of “New Age” Environmental Protection. Washburn Law Journal 41:50–89.

Doremus, H. 2004. The Purposes, Effects, and Future of the Endangered Species Act's Best Available
Science Mandate. Environmental Law 34:397.

Downs, P. W., and G. M. Kondolf. 2002. Post-project Appraisals in Adaptive Management of River
Channel Restoration. Environmental Management 29(4):477–496.

Elzinga, C. L., D. W. Salzer, and J. W. Willoughby. 1998. Measuring & Monitoring Plant Populations.
Denver, CO: Bureau of Land Management Technical Reference 1730-1.

Fore, L. S., J. R. Karr, and R. W. Wisseman. 1996. Assessing Invertebrate Responses to Human
Activities: Evaluating Alternative Approaches. Journal of the North American Benthological
Society 15(2):212–231.

Gunderson, L., and S. S. Light. 2006. Adaptive Management and Adaptive Governance in the
Everglades Ecosystem. Policy Sciences 39(4):323–334.

Habron, G. 2003. Role of Adaptive Management for Watershed Councils. Environmental Management
31(1):0029–0041.

Hairston, N. G. 1987. Community Ecology and Salamander Guilds. Cambridge University Press, 1987

Herlihy, A. T., J. L. Stoddard, and C. Burch Johnson. 1998. The Relationship between Stream
Chemistry and Watershed Land Cover Data in the Mid-Atlantic Region, U.S. Pages 377–386 in
Biogeochemical Investigations at Watershed, Landscape, and Regional Scales. Springer
Netherlands.

Holling, C. S. 1978. Adaptive Environmental Assessment and Management. Chichester, UK: Wiley
Interscience.

Hudson, P. L., R. W. Griffiths, and T. J. Wheaton. 1992. Review of Habitat Classification Schemes
Appropriate to Streams, Rivers, and Connecting Channels in the Great Lakes Drainage Basin.
Pages 73–107 in W. D. N. Busch and P. G. Sly (eds.), The Development of an Aquatic Habitat
Classification System for Lakes. Ann Arbor, MI: CRC Press.

Hughes, R. M. (ed.). 1993. Stream Indicators and Design Workshop. EPA/600/R-93/138. U.S.
Environmental Protection Agency, Corvallis, OR.

Large Wood National Manual July 2015


9-18
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

Hurlbert, S. H. 1984. Pseudoreplication and the Design of Ecological Field Experiments. Ecological
Monographs 54 (2:187–211.

Kail, J., D. Hering, S. Muhar, J. Gerhard, and S. Preis. 2007. The Use of Large Wood in Stream
Restoration: Experiences from 50 Projects in Germany and Austria. Journal of Applied Ecology
44:1145–1155.

Kaplan, R. S. 2008. Conceptual Foundations of the Balanced Scorecard. Handbook of Management


Accounting Research 3:1253–1269.

Karr, J. R. 1981. Assessment of Biotic Integrity Using Fish Communities. Fisheries 6(6):21–27.

Karr, J. R., and B. L. Kerans. 1991. Components of Biological Integrity: Their Definition and Use in
Development of an Invertebrate IBI. Midwest Pollution Control Biologists Meeting, Chicago, Ill.,
1991, Proceedings: US Environmental Protection Agency, Region V, EPA-905/R-92-003.

Karr, J. R., K. D. Fausch, P. L. Angermeier, P. R. Yant, and I. J. Schlosser. 1986. Assessing Biological
Integrity in Running Waters: A Method and its Rationale. Illinois Natural History Survey Special
Publication 5, Urbana, IL.

Kati, V., P. Devillers, M. Dufrene, A. Legakis, D. Vokou, and P. Lebrun. 2004. Testing the Value of Six
Taxonomic Groups as Biodiversity Indicators at a Local Scale. Conservation Biology 18(3):667–
675.

Kaufmann, P. R., P. Levine, E. G. Robison, C. Seeliger, and D. V. Peck. 1999. Quantifying Physical
Habitat in Streams. U.S. Environmental Protection Agency, Washington, D.C. Available:
http://www.epa.gov/emap2/html/pubs/docs/groupdocs/surfwatr/field/phyhab.pdf.

Keith, D. A., T. G. Martin, E. McDonald-Madden, and C. Walters. 2011. Uncertainty and Adaptive
Management for Biodiversity Conservation. Biological Conservation 144(4):1175–1178.

Kerans, B. L., and J. R. Karr. 1994. A Benthic Index of Biotic Integrity (B-IBI) for Rivers of the
Tennessee Valley. Ecological Applications. 4(4):768–785.

Kingsford, R. T., K. F. Walker, R. E. Lester, W. J. Young, P. G. Fairweather, J. Sammut, and M. C. Geddes.


2011. A Ramsar Wetland in Crisis—The Coorong, Lower Lakes and Murray Mouth, Australia.
Marine and Freshwater Research 62(3):255–265.

Kondolf, G. M. 1995. Five Elements for Effective Stream Restoration. Restoration Ecology 3:133–136.

Kondolf, G. M. 1996. A Cross Section of Stream Channel Restoration. Journal of Soil and Water
Conservation 51(2):119–125.

Kondolf, G. M. 1998. Lessons Learned from River Restoration Projects in California. Aquatic
Conservation: Marine and Freshwater Ecosystems 8(1):39–52.

Lakowicz, J. R., and G. Weber. 1973. Quenching of Fluorescence by Oxygen. Probe for Structural
Fluctuations in Macromolecules. Biochemistry 12(21): 4161–4170.

Landres, P. B., J. Verner, and J. W. Thomas. 1988. Ecological Uses of Vertebrate Indicator Species: A
Critique. Conservation Biology 2(4):316–328.

Lange-Bertalot, H. 1979. Pollution Tolerance of Diatoms as a Criterion for Water Quality Estimation.
Nova Hedwigia, Beih. 64:285–304.

Large Wood National Manual July 2015


9-19
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

Lee, K. N. 1999. Appraising Adaptive Management. Conservation Ecology 3(2):3.

Lowe, R. L. 1974. Environmental Requirements and Pollution Tolerance of Freshwater Diatoms.


National Environmental Research Center, Office of Research and Development, U.S.
Environmental Protection Agency.

Lyons, J. E., M. C. Runge, H. P. Laskowski, and W. L. Kendall. 2008. Monitoring in the Context of
Structured Decision-Making and Adaptive Management. The Journal of Wildlife Management
72(8):1683–1692.

Mackenzie, B. D. E, and D. A. Keith. 2009. Adaptive Management in Practice: Conservation of a


Threatened Plant Population. Ecological Management & Restoration 10(s1):S129–S135.

MacNally, R., and E, Fleishman. 2004. A Successful Predictive Model of Species Richness Based on
Indicator Species. Conservation Biology 18(3):646–654.

McCormick, F. H. 1993. Fish Communities as Indicators of Stream Condition. In R. M. Hughes (ed.),


EMAP Streams Bioassessment Workshop. Report of the Proceedings. EPA/600/R-93/138. U.S.
Environmental Protection Agency, Corvallis, OR.

Murphy, D. D., and P. S. Weiland. 2010. The Route to Best Science in Implementation of the
Endangered Species Act’s Consultation Mandate: The Benefits of Structured Effects Analysis.
Environmental Management 47(2):161–172.

Nagayama, S., and F. Nakamura. 2010. Fish Habitat Rehabilitation Using Wood in the World.
Landscape and Ecological Engineering 6(2):289–305.

National Research Council (US). 1999. Committee on Health Risks of Exposure to Radon. Health
Effects of Exposure to Radon. Vol. 6. National Academies Press.

National Research Council. 2004. Adaptive Management for Water Resources Planning, The National
Academies Press. Washington, D.C.

Norris, R. H. 1995. Biological Monitoring: The Dilemma of Data Analysis." Journal of the North
American Benthological Society:440–450.

Oakley, K. L., L. P. Thomas, and S. G. Fancy. 2003. Guidelines for Long-Term Monitoring Protocols.
Wildlife Society Bulletin 31(4):1000–1003.

Oglethorpe, J. 2002. Adaptive Management: From Theory to Practice. The World Conservation Union
(IUCN).

Opperman, J. J., and A. M. Merenlender. 2004. The Effectiveness of Riparian Restoration for
Improving Instream Fish Habitat in Four Hardwood-Dominated California Streams. North
American Journal of Fisheries Management 24(3):822–834.

Pahl-Wostl, C. 1995. The Dynamic Nature of Ecosystems: Chaos and Order Entwined. Chichester:
Wiley.

Pastorok, R. A., A. MacDonald, J. R. Sampson, P. Wilber, D. J Yozzo, and J. P Titre. 1997. An Ecological
Decision Framework for Environmental Restoration Projects. Ecological Engineering 9 (1–
2):89–107.

Large Wood National Manual July 2015


9-20
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

Poff, N. L., J. D. Allan, M. A. Palmer, D. D. Hart, B. D. Richter, A. H. Arthington, K. H. Rogers, J. L. Meyer,


and J. A. Stanford. 2003. River Flows and Water Wars: Emerging Science for Environmental
Decision Making. Frontiers in Ecology and the Environment 1:298–306.

Pollock, K. H., J. D. Nichols, T. R. Simons, G. L. Farnsworth, L. L. Bailey, and J. R. Sauer. 2002. Large
Scale Wildlife Monitoring Studies: Statistical Methods for Design and Analysis. Environmetrics
13:105–119.

Prato, T. 2003. Adaptive Management of Large Rivers with Special Reference to the Missouri River.
Journal of the American Water Resources Association 39(4):935–946.

Reich, P. B., I. Wright, J. Cavender-Bares, J. Craine, J. Oleksyn, M. Westoby, and M. B. Walters. 2003.
The Evolution of Plant Functional Variation: Traits, Spectra, and Strategies. International Journal
of Plant Sciences 164:s143–s164.

Resh, V. H., R. H. Norris, and M. T. Barbour. 1995. Design and Implementation of Rapid Assessment
Approaches for Water Resource Monitoring Using Benthic Macroinvertebrates. Australian
Journal of Ecology 20(1):108–121.

Reynoldson, T. B., R. C. Bailey, K. E. Day, and R. J. Norris. 1995. Biological Guidelines for Freshwater
Sediment Based on BEnthic Assessment of SedimenT (the BEAST) Using a Multivariate
Approach for Predicting Biological State. Australian Journal of Ecology 20(1):198–219.

Roni, P. (ed.). 2005. Monitoring Stream and Watershed Restoration. American Fisheries Society,
Bethesda, MD.

Roni, P., M. Liermann, and A. Steel. 2003. Monitoring and Evaluating Fish Response to Instream
Restoration. In D. Montgomery, S. Bolton, D. Booth, and L. Wall (eds.), Restoration of Puget Sound
Rivers. Center for Water and Watershed Studies. University of Washington Press: Seattle.

Rosenberg, D. M., and V. H. Resh (eds.). 1993. Freshwater Biomonitoring and Benthic
Macroinvertebrates. Springer.

Rumpff, L., Duncan, D. H., P. A. Vesk, D. A. Keith, and B. A. Wintle. 2011. State-and-Transition
Modelling for Adaptive Management of Native Woodlands. Biological Conservation 144(4):1224–
1236.

Scheiner S. M., and J. Gurevitch (eds.). 2001. Design and Analysis of Ecological Experiments. Oxford
University Press.

Simon, T.P., and J. Lyons. 1995. Application of the Index of Biotic Integrity to Evaluate Water
Resource Integrity in Freshwater Ecosystems. Chapter 16 in W. S. Davis and T. P. Simon.
Bioassessment and Criteria: Tools for Water Resources Planning and Decision Making. CRC Press.

Stanford, J. A., and G. C. Poole. 1996. A Protocol for Ecosystem Management. Ecological
Applications:741–744.

Stewart-Oaten, A., J. R. Bence, and C. W. Osenberg. 1992. Assessing Effects of Unreplicated


Perturbations: No Simple Solutions. Ecology:1396–1404.

Teels, B. M., and T. Danielson. 2001. Using a Regional IBI to Characterize Condition of Northern
Virginia Streams, with Emphasis on the Occoquan Watershed. USDANRCS. Technical Note 190-
13-1. December 2001

Large Wood National Manual July 2015


9-21
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 9. Assessing Ecological Performance

U.S. Environmental Protection Agency (EPA). 1995. A Decision-Making Guide for Restoration in
Ecological Restoration. EPA 841-F-95-007 (November)

Vail, L. W., and R. L. Skaggs. 2002. Adaptive Management Platform for Natural Resources in the
Columbia River Basin. Pacific Northwest National Laboratory. Available:
http://www.pnl.gov/main/publications/external/technical_reports/PNNL-13875.pdf.

Van Wilgen, B. W., and H. C. Biggs. 2011. A Critical Assessment of Adaptive Ecosystem Management
in a Large Savanna Protected Area in South Africa. Biological Conservation 144(4):1179–1187.

Volkman, J. M., and W. E. McConnaha. 1993. Through a Glass, Darkly: Columbia River Salmon, the
Endangered Species Act, and Adaptive Management. Environmental Law 23:1249–1272.

Walters, C. 2002. Adaptive Management of Renewable Resources. The Blackburn Press.

Williams, B. K., R. C. Szaro, and C. D. Shapiro. 2009. Adaptive Management: The U.S. Department of the
Interior Technical Guide. Adaptive Management Working Group, U.S. Department of the Interior,
Washington, D.C.

Wissmar, R. C., and R. L. Beschta. 1998. Restoration and Management of Riparian Ecosystems: A
Catchment Perspective. Freshwater Biology 40(3):571–585.

Wright, L. D. (ed.). 1995. Morphodynamics of Inner Continental Shelves. CRC Press.

Yoccoz, N. G., J. D. Nichols, and T. Boulinier. 2001. Monitoring of Biological Diversity in Space and
Time. Trends in Ecology & Evolution 16(8):446–453.

Large Wood National Manual July 2015


9-22
Chapter 10
LARGE WOOD BIBLIOGRAPHY
This page intentionally left blank.
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Abbe, T. B. 2000. Patterns, Mechanics, and Geomorphic Effects of Wood Debris Accumulations in
a Forest River System. Ph.D. dissertation. University of Washington, Seattle, WA.

Abbe, T. 2008. Alluvial Landscape Response to Climate Change in Glacial Rivers and the Implications to
River Restoration. Presentation at River Restoration Northwest. Available:
http://archive.rrnw.org/docs/2008/zSession%202/4-Abbe%20RRNW%202008%20final.pdf.

Abbe, T. B., and A. P. Brooks. 2011. Geomorphic, Engineering, and Ecological Considerations when
Using Wood in River Restoration. Pages 419–451 in A. Simon, S. J. Bennett, and J. M. Castro
(eds.), Stream Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools.
Geophysical Monograph Series 194. Washington, D.C.: American Geophysical Union.

Abbe, T. B., and D. R. Montgomery. 1996. Large Woody Debris Jams, Channel Hydraulics and Habitat
Formation in Large Rivers. Regulated Rivers: Research and Management 12:201–221.

Abbe, T. B., and D. R. Montgomery. 2003. Patterns and Processes of Wood Debris Accumulation in
the Queets River Basin, Washington. Geomorphology 51:81–107.

Abbe, T. B., D. R. Montgomery, K. Fetherston, and E. M. McClure. 1993. A Process-Based Classification


of Woody Debris in a Fluvial Network: Preliminary Analysis of the Queets River, Washington.
EOS, American Geophysical Union Transactions 73(43):296.

Abbe, T. B., D. R. Montgomery, and C. Petroff. 1997. Design of Stable In-Channel Wood Debris
Structures for Bank Protection and Habitat Restoration: An Example from the Cowlitz River, WA.
Pages 809–816 in S. S. Y. Wang, E. J. Langendoen, and F. D. Shields, F.D. (eds.), Proceedings of the
Conference on Management of Landscapes Disturbed by Channel Incision. University of
Mississippi, Oxford, MS.

Abbe, T. B., J. Carrasquero, M. McBride, A. Ritchie, M. McHenry, and K. Dublanica. 2003a.


Rehabilitating River Valley Ecosystems: Examples of Public, Private, and First Nation Cooperation
in Western Washington. Proceedings of the Georgia Basin/Puget Sound 2003 Research
Conference, Vancouver, B.C., March 31–April 1, 2003, T. Droscher (ed.). Puget Sound Action
Team, Olympia, WA.

Abbe, T. B, A. P. Brooks, and D. R. Montgomery. 2003b. Wood in River Rehabilitation and


Management. Pages 367–389 in S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology
and Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society.

Abbe, T. B., G. Pess, D. R. Montgomery, and K. L. Fetherston. 2003c. Integrating Engineered Log Jam
Technology into River Rehabilitation. In D. R. Montgomery, S. Bolton, D. Booth, and L. Wall
(eds.), Restoration of Puget Sound Rivers. Center for Water and Watershed Studies, University of
Washington, Seattle.

Abbe, T. B., C. Miller, and A. Michael. 2009. Self-Mitigating Protection for Pipeline Crossings in
Degraded Streams: A Case Study from Woodward Creek, Washington. 9th International Right of
Way Symposium. 2009. Portland, OR.

Abbe. T., J. Bjork, A. Zehni, T. Nelson, and J. Park. 2011. New Innovative, Habitat-Creating Bank
Protection Method. Pages 2011–2021 in Proceedings of ASCE World Environmental and Water
Resources Congress.

Large Wood National Manual July 2015


10-1
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Abbe. T., M. Ericsson, and L. Embertson, L. 2013. Geomorphic Assessment of the Larson Reach of the
South Fork Nooksack River, NW Washington. Report submitted to Lummi Indian Nation.

Abt, S. R., R. J. Wittler, A. Taylor, and D. J. Love. 1989. Human Stability in a High Flood Hazard Zone.
American Water Resources Association. Water Resources Bulletin 25(4):881–889.

Advisory Council on Historic Preservation (ACHP). 2010. Protecting Historic Properties: A Citizen’s
Guide to Section 106 Review. Advisory Council on Historic Preservation. Washington, D.C.

Agee, J. K. 1990. The Historical Role of Fire in Pacific Northwest Forests. Pages 25–38 in J. Walstad,
S. R. Radosevich, and D. V. Sandberg (eds.), Natural and Prescribed Fire in Pacific Northwest
Forests. Corvallis: Oregon State University Press.

Agee, J. K. 1992. The Historical Role of Fire in Pacific Northwest Forests. Pages 25–38 in J. Walstad,
S. R. Radosevich, and D. V. Sandberg (eds.), Natural and Prescribed Fire in Pacific Northwest
Forests. Corvallis: Oregon State University Press.

Agee, J. K. 1993. Fire Ecology of Pacific Northwest Forests. Washington, DC: Island Press.

Agrawal, A., M. A. Khan, and Z. Yi. 2007. Handbook of Scour Countermeasures and Design. FHWA-NJ-
2005-027. New Jersey Department of Transportation and Federal Highway Administration,
Washington, DC.

Ahmad, M. 1951. Spacing and Projection of Spurs for Bank Protection. Civil Engineering and Public
Works Review. March:172–174; April:256–258.

Allan, C., and G. H. Stankey. 2009. Adaptive Environmental Management. Volume 351. Springer.

Allan, J. D. 1995. Stream Ecology: Structure and Function of Running Waters. Boston, MA: Kluwer
Academic Publishers.

Allan, J. D., M. S. Wipfli, J. P. Caouette, A. Prussian, and J. Rodgers. 2003. Influence of Streamside
Vegetation on Inputs of Terrestrial Invertebrates to salmonid Food Webs. Canadian Journal of
Fisheries and Aquatic Sciences 60(3):309–320.

Allen, H. H., and J. R. Leech. 1997. Bioengineering for Streambank Erosion Control. Technical Report
E 97-8, U.S. Army Corps of Engineers Waterways Experiment Station, Vicksburg, MS.

Ambrose, H. E., M. A. Wilzbach, and K. W. Cummins. 2004. Periphyton Response to Increased Light
and Salmon Carcass Introduction in Northern California Streams. Journal of the North American
Benthologocial Society 23(4):701–712.

American Fisheries Society. 1983. Stream Obstruction Removal Guidelines. Stream Renovation
Guidelines Committee. The Wildlife Society and American Fisheries Society. Published by AFS,
Washington D.C.

American Society of Landscape Architects (ASLA). 2013. American Society of Landscape Architects
(ASLA). Available: http://www.asla.org. Accessed: August 27, 2013.

American Whitewater Association. 2012. Integrating Recreational Boating Considerations into


Stream Channel Modificatin & Design Projects. Written by Kevin Colburn, National Stewardship
Director. Illustrations by Chad Lewis. Figure 8.4, page 13. Available:
http://www.americanwhitewater.org/content/Document/fetch/documentid/1006/.raw.

Large Wood National Manual July 2015


10-2
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Anderson, B. 2005. Will Replanting Vegetation along River Banks Make Floods Worse?
8th International River Symposium, Brisbane, Australia.

Anderson, D. B. 2006. Quantifying the Interaction between Riparian Vegetation and Flooding: from
Cross-Section to Catchment Scale. University of Melbourne.

Anderson, D. H., and B. D. Dugger. 1998. A Conceptual Basis for Evaluating Restoration Success.
Transactions of the North American Wildlife and Natural Resources Conference Volume 63.
Wildlife Management Institute.

Anderson, N. H., R. J. Steedman, and T. Dudley. 1984. Patterns of Exploitation by Stream


Invertebrates of Wood Debris (Xylophagy). Verhandlungen der Internationalen Vereinigung für
theoretische und angewandte Limnologie 22:1847-1852.

Anderson, P. D., D. J. Larson, and S. S. Chan. 2007. Riparian Buffer and Density Management
Influences on Microclimate of Young Headwater Forests of Western Oregon. Forest Science
53(2):254–269.

Andrus, B., and J. Gessford. 2007. Understanding the Legal Risks Associated with the Design and
Construction of Engineered Logjams. Skellenger Bender Attorneys, Seattle, WA.

Andrus, C. W., B. A. Long, and H. A. Froehlich. 1988. Woody Debris and its Contribution to Pool
Formation in a Coastal Stream 50 Years after Logging: Canadian Journal of Fish and Aquatic
Science 45:2080–2086.

Angradi, T. R., E. W. Schweiger, D. W. Bolgrien, P. Ismert, and T. Selle. 2004. Bank Stabilization,
Riparian Land Use and the Distribution of Large Woody Debris in a Regulated Reach of the
Upper Missouri River, North Dakota, USA. River Research and Applications 20:829–846.

Armstrong J. D., and K. H. Nislow. 2006. Critical Habitat During the Transition from Maternal
Provisioning in Freshwater Fish, with Emphasis on Atlantic Salmon (Salmo salar) and Brown
Trout (Salmo trutta). Journal of Zoology 269,403–413.

Arnáez, J., V. Larrea, and L. Ortigosa. 2004. Surface Runoff and Soil Erosion on Unpaved Forest Roads
from Rainfall Simulation Tests in Northeastern Spain. Catena 57:1–14.

Arneson, L. A., L. W. Zevenbergen, P. F. Lagasse, and P. E. Clopper. 2012. Evaluating Scour at Bridges.
Hydraulic Engineering Circular 18, FHWA-HIF-12-003, National Highway Institute, Federal
Highway Administration, Arlington, VA.

Arno, S. F., and R. J. Hoff. 1989. Silvics of Whitebark Pine (Pinus albicaulis). USDA For. Serv. Gen Tech.
Rep. INT-253.

Babakaiff, S., D. Hay, and C. Fromuth. 1997. Rehabilitating Stream Banks. In P. A. Slaney and
D. Zaldokas (eds.), Fish Habitat Rehabilitation Procedures, Watershed Restoration Program.
Ministry of Environment, Lands and Parks, Vancouver, BC.

Bailey, R. G. 1995. Description of Ecoregions of the United States, 2nd Edition, USDA, US Forest Service.
Washington, D.C. Miscellaneous Publication No. 1391.

Bailey, R. G. 2009. Ecosystem Geography – From Ecoregions to Sites. New York, NY: Springer Science
and Business Media.

Large Wood National Manual July 2015


10-3
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Baillie, B. R., and T. R. Davies. 2002. Influence of Large Woody Debris On Channel Morphology in
Native Forest and Pine Plantation Streams in the Nelson Region, New Zealand. New Zealand
Journal of Marine and Freshwater Research. 36:763–774.

Baillie, B. R., L. G. Garret, and A. W. Evanson. 2008. Spatial Distribution Influence of LWD in an
Old-growth Forest River System, New Zealand. Forest Ecology and Management 256:20–27.

Bain, M. B., and N. J. Stevenson (eds.). 1999. Aquatic Habitat Assessment: Common Methods. Bethesda,
MD: American Fisheries Society.

Baker, C. O. 1979. The Impacts of Logjam Removal on Fish Populations and Stream Habitat in Western
Oregon. M.S. Thesis, Oregon State University, Corvallis, OR.

Banta, E. R., M. C. Hill, and M. G. McDonald. 2000. MODFLOW-2000, the US Geological Survey Modular
Ground-Water Model: User Guide to Modularization Concepts and the Ground-Water Flow Process.
Reston, VA, USA: US Geological Survey.

Barbour, M. T., J. Gerritsen, G. E. Griffith, R. Frydenborg, E. McCarron, J. S. White, and M. L. Bastian.


1996. A Framework for Biological Criteria for Florida Streams Using Benthic
Macroinvertebrates. Journal of the North American Benthological Society (1996):185–211.
Available: http://www.jstor.org/discover/10.2307/1467948?uid=3739936&uid=
2&uid=4&uid=3739256&sid=21104863995063.

Barbour, M. T., J. Gerritsen, B. D. Snyder, and J. B. Stribling. 1999. Rapid Bioassessment Protocols for
use in Streams and Wadeable Rivers. USEPA, Washington. Available:
http://zoology.okstate.edu/zoo_lrc/biol1114/study_guides/labs/lab12/gen_usepa_barbouretal
_1999_rba.pdf.

Bardini, L., F. Boano, M. B. Cardenas, A. H. Sawyer, R. Revelli and L. Ridolfi. 2013. Small-Scale
Permeability Heterogeneity has Negligible Effects on Nutrient Cycling in Streambeds.
Geophysical Research Letters 40:1118–1122.

Barker, B. L., R. D. Nelson, and M. S. Wigmosta. 1991. Performance of Detention Ponds Designed
According to Current Standards. Puget Sound Water Quality Authority, Puget Sound Research '91:
Conference Proceedings. Seattle, Washington.

Barrett, M. L. 1996. Environmental Reconstruction of a 19th Century Red River Raft Lake: Caddo
Lake, Louisiana and Texas. Gulf Coast Association of Geological Societies Transactions 46:471–
471.

Bartholow, J. 1988. Stream Segment Shade Model (SSSHADE) Version 1.4. Temperature Model
Technical Note #3. U.S. Fish and Wildlife Service, Fort Collins, CO.

Bartz, K. K., K. M. Lagueux, M. D. Scheuerell, T. Beechie, A. D. Haas, and M. H. Ruckelshaus. 2006.


Translating Restoration Scenarios into Habitat Conditions: An Initial Step in Evaluating
Recovery Strategies for Chinook Salmon (Oncorhynchus tshawytscha). Canadian Journal of
Fisheries and Aquatic Sciences 63(7):1578–1595.

Bash, J. S. and C. M. Ryan. 2002. Stream Restoration and Enhancement Projects: Is Anyone
Monitoring? Environmental Management 29(6):877¬885.

Baxter, C. V., K. D. Fausch, and W. Carl Saunders. 2005. Tangled Webs: Reciprocal Flows of
Invertebrate Prey Link Streams and Riparian Zones. Freshwater Biology 50(2):201–220.

Large Wood National Manual July 2015


10-4
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Beckman, N. D., and E. Wohl. 2014. Carbon Storage in Mountainous Headwater Streams: The Role of
Old-Growth Forest and Logjams. Water Resources Research 50:2376–2393.

Beamer, E. M., and R. A. Henderson. 1998. Juvenile Salmonid use of Natural and Hydromodified
Stream Bank Habitat in the Mainstem Skagit River, Northwest Washington. Miscellaneous Report.
Skagit System Cooperative. La Connor, WA.

Beecher, H. A., B. A. Caldwell, S. B. DeMond, D. Seiler, D., and S. N. Boessow. 2010. An Empirical
Assessment of PHABSIM Using Long-Term Monitoring of Coho Salmon Smolt Production in
Bingham Creek, Washington. North American Journal of Fisheries Management 30(6):1529–
1543. doi:10.1577/M10.

Beechie, T. J., and S. Bolton. 1999. An Approach to Restoring Salmonid Habitat-Forming Processes in
Pacific Northwest Watersheds. Fisheries 24:6–15.

Beechie, T. J., and K. Wyman. 1992. Stream Habitat Conditions, Unstable Slopes and Status of Roads in
Four Small Watersheds of the Skagit River. Skagit System Cooperative, Fisheries services for the
Swinomish Tribal Community, Upper Skagit and Sauk-Suiattle Indian Tribes.

Beechie, T. J., G. Pess, P. Kennard, R. E. Bilby, and S. Bolton. 2000. Modeling Recovery Rates and
Pathways for Woody Debris Recruitment in Northwestern Washington Streams. North American
Journal of Fisheries Management 20:436–452.

Beechie, T. J., D. A. Sear, J. D. Olden, G. R. Pess, J. M. Buffington, H. Moir, P. Roni, and M. M. Pollock.
2010. Process-Based Principles for Restoring River Ecosystems. Bioscience 60:209–222.

Beechie, T. J., H. Imaki, J. Greene, A. Wade, H. Wu, G. Pess, P. Roni, J. Kimball, J. Stanford, P. Kiffney,
and N Mantua. 2012. Restoring Salmon Habitat for a Changing Climate. River Research and
Applications 29:939–960.

Beesley, L. 1996. The Ecological Importance of Large Woody Debris in the Sandy River Systems of the
Swan Coastal Plain (Perth, Western Australia). Honors thesis, University of Western Australia,
Perth.

Beets, P. N., I. A. Hood, M. O. Kimberley, G. R. Oliver, S. H. Pearce, and J. F. Gardner. 2008. Coarse
Woody Debris Decay Rates for Seven Indigenous Tree Species in the Central North Island of New
Zealand. Forest Ecology and Management 256:548–557.

Begon, M., J. L. Harper, and C. R. Townsend. 1990. Ecology: Individuals, Populations, Communities,
2nd Edition. Blackwell Scientific Publications: Boston.

Beltaos, S. 1983. River Ice Jams: Theory, Case Studies, and Applications. Journal of Hydraulic
Engineering 109(10):1338–1359.

Bencala, K. E. 2005. Hyporheic Exchange Flows. Encyclopedia of Hydrological Sciences,


M. G. Anderson and J. J. McDonnell (eds.). Wiley-Blackwell.

Bencala, K. E., M. N. Gooseff, and B. A. Kimball. 2011. Rethinking Hyporheic Flow and Transient
Storage to Advance Understanding of Stream-Catchment Connections. Water Resources Research
47(3):W00h03.

Benda, L. and T. W. Cundy. 1990. Predicting Deposition of Debris Flow in Mountain Channels.
Canadian Geotechnical Journal 27:409–417.

Large Wood National Manual July 2015


10-5
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Benda, L. E., and J. C. Sias. 2003. A Quantitative Framework for Evaluating the Mass Balance of
In-Stream Organic Debris. Forest Ecology and Management 172:1–16.

Benda, L., Miller, D., Bigelow, P., Andras, K. 2003a. Effects of Post-Wildfire Erosion on Channel
Environments, Boise River, Idaho. Forest Ecology and Management 178:105–119.

Benda, L., D. Miller, J. Sias, D. Martin, R. Bilby, C. Veldhuisen, and T. Dunne. 2003b. Wood Recruitment
Processes and Wood Budgeting. American Fisheries Society Symposium 37:49–73.

Benda, L. E., D. Miller, K. Andras, P. E. Bigelow, G. H. Reeves, and D. Michael. 2007. NetMap: A New
Tool in Support of Watershed Science and Resource Management. Forest Science 53(2):206–219.

Bender, L. C., G. J. Roloff, and J. B. Haufler. 1996. Evaluating Confidence Intervals for Habitat
Suitability Models. Wildlife Society Bulletin 24(2):347–352.

Benke, A. C. 2001. Importance of Flood Regime to Invertebrate Habitat in an Unregulated


River-Floodplain Ecosystem. Journal of the North American Benthological Society 20:225–240.

Benke, A. C., and J. B. Wallace. 1990. Wood Dynamics in Coastal Plain Blackwater Streams. Canadian
Journal of Fisheries and Aquatic Sciences 47:92–99.

Benke, A. C., and J. B. Wallace. 2010. Influence of Wood on Invertebrate Communities in Streams and
Rivers. American Fisheries Society Symposium 37:149–177.

Benke, A. C., R. L. Henry III, D. M. Gillespie, and R. J. Hunter. 1985. Importance of Snag Habitat for
Animal Production in Southeastern Streams. Fisheries 10:8–12.

Bentz, B. J., J. Régnière, C. J. Fettig, M. Hansen, J. L. Hayes, J. A. Hicke, R. G. Kelsey, J. F. Negrón, and
S. J. SeyboLd. 2010. Climate Change and Bark Beetles of the Western United States and Canada:
Direct and Indirect Effects. BioScience 60:602–613.

Berg, N. A., A. Carlson, and D. Azuma. 1998. Function and Dynamics of Woody Debris in Stream
Reaches in the Central Sierra Nevada, California. Canadian Journal of Fisheries and Aquatic
Sciences 55:1807–1820.

Bernhardt, E. S., and 24 others. 2005. Synthesizing U.S. River Restoration Efforts. Science.
308(5722):636–637

Bernhardt, E. S., E. B. Sudduth, M. A. Palmer, J. D. Allan, J. L. Meyer, G. Alexander, J. Follastad-Shah,


B. A. Hassett, R. Jenkinson, R. Lave, J. Rumps, and L. Pagano. 2007. Restoring Rivers One Reach at
a Time: Results from a Survey of U.S. River Restoration Practitioners. Restoration Ecology
15(3):482–493.

Bertoldi, W., A. M. Gurnell, and M. Welber 2013. Wood Recruitment and Retention: The Fate of
Eroded Trees on a Braided River Explored Using a Combination of Field and Remotely-Sensed
Data Sources. Geomorphology 180–181(0):146–155.

Beschta, R. L. 1997. Restoration of Riparian and Aquatic Systems for Improved Aquatic Habitats in
the Upper Columbia River Bbasin. Pages 475-491 in D. J. Stouder and P. A. Bisson (eds.). Pacific
Salmon and Their Ecosystems: Status and Future Options. New York: Chapman Hall.

Beschta, R. L., R. E. Bilby, L. B. Brown, L. B. Holtby, and T. D. Hofstra. 1987. Stream Temperature and
Aquatic Habitat: Fisheries and Forestry Interactions. Pages 191-232 in E. O. Salo and

Large Wood National Manual July 2015


10-6
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

T. W. Cundy (eds.), Streamside Management: Forestry and Fishery Interactions. College of Forest
Resources, University of Washington, Seattle, WA. 471p.

Beverton, R. J. H., and S. J. Holt. 1957. On the Dynamics of Exploited Fish Populations. U.K. Ministry of
Agriculture, Fisheries Investigation Service 2(19):553.

Bilby, R. E. 1981. Role of Organic Debris Dams in Regulating the Export of Dissolved and Particulate
Matter from a Forested Watershed. Ecology 62(5):1234–1243.

Bilby, R. E. 1984. Removal of Woody Debris May Affect Stream Channel Stability. Journal of Forestry,
609–613. October.

Bilby, R. E. 2003. Decomposition and Nutrient Dynamics of Wood in Streams and Rivers. The Ecology
and Management of Wood in World Rivers, American Fisheries Society Symposium 37:135–147.

Bilby, R. E., and P. A. Bisson. 1998. Function and Distribution of Large Woody Debris. Pages 324–346
in R. J. Naiman and R. E. Bilby (eds.), River Ecology and Management: Lessons from the Pacific
Coast Ecoregion. New York, NY: Springer-Verlag.

Bilby, R. E., and G. E. Likens. 1980. Importance of Debris Dams in the Structure and Function of
Stream Ecosystems. Ecology 61:1107–1113.

Bilby, R. E. and J. W. Ward. 1991. Characteristics and Function of Large Woody Debris in Streams
Draining Old-Growth, Clear-Cut, and Second-Growth Forests in Southwestern Washington.
Canadian Journal of Fisheries and Aquatic Sciences 48:2499–2508.

Bilby, R. E., and J. W. Ward. 1989. Changes in Characteristics and Function of Woody Debris With
Increasing Size of Streams in Western Washington. Transactions of the American Fisheries
Society 118:368–378.

Bilby, R. E., and L. J. Wasserman. 1989. Forest Practices and Riparian Management in Washington
State: Data Based Regulation Development. In R. E. Gresswell, B. A. Barton, and J. L. Kershner
(eds.), Practical Approaches to Riparian Management. U.S. Bureau of Land Management, BLM MT
PT 89 001 4351, Billings, Montana.

Bilby, R. E., B. R. Fransen, P. A. Bisson, and J. K. Walter. 1998. Response of Juvenile Coho Salmon
(Oncorhynchus kisutch) and Steelhead (Oncorhynchus mykiss) to the Addition of Salmon
Carcasses to Two Streams in Southwestern Washington, U.S.A. Canadian Journal of Fisheries and
Aquatic Science 55:1909–1918.

Bilby, R. E., J. T. Heffner, B. R. Fransen, F. W. Ward, and P. A. Bisson. 1999. Effects of Immersion in
Water on Deterioration of Wood from Five Species of Trees Used for Habitat Enhancement
Projects. North American Journal of Fisheries Management 19(3):687–695.

Binns, N. A., and F. M. Eiserman. 1979. Quantification of Fluvial Trout Habitat in Wyoming.
Transactions of the American Fisheries Society 109(3):215–228.

Bisson, P. A., J. L. Nielsen, and R. A. Palmason. 1982. A System of Naming Habitat Types in Small
Steams, with Examples of Habitat Utilization by Salmonids during Low Steamflow. Pages 62–73
in N. B. Armantrout (ed.), Acquisition and Utilization of Aquatic Habitat Inventory Information.
Bethesda, MD: American Fisheries Society.

Large Wood National Manual July 2015


10-7
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Bisson, P. A., R. E. Bilby, M. D. Bryant, C. A. Dolloff, G. B. Grette, R. A. House, M. L. Murphy, K. V. Koski,


and J. R. Sedell. 1987. Large Woody Debris in Forested Streams in the Pacific Northwest: Past,
Present, and Future. Pages 143–190, in E. O. Salo and T. W. Cundy (eds.), Streamside
Management: Forestry and Fishery Interactions. College of Forest Resources, University of
Washington, Seattle, Washington.

Bisson, P. A., B. E. Rieman, C. Luce, P. F. Hessburg, D. C. Lee, J. L. Kershner, G. H. Reeves, and


R. E. Gresswell. 2003. Fire and Aquatic Ecosystems of the Western USA: Current Knowledge and
Key Questions. Forest Ecology and Management 178:213–229.

Bjornn, T. C., and D. W. Reiser. 1991. Habitat Requirements of Salmonids in Streams. Pages 83-138 in
W. R. Meehan (ed.), Influences of Forest and Rangeland Management on Salmonid Fishes and their
Habitats. Bethesda, MD: American Fisheries Society.

Blair, G. R., L. C. Lestelle, and L. E. Mobrand. 2009. The Ecosystem Diagnosis and treatment Model: A
Tool for Assessing Salmonid Performance Potential Based on Habitat Conditions. Pages 289–309
in E. E. Knudsen and J. J. Michael, Jr., Pacific Salmon Environment and Life History Models:
Advancing Science for Sustainable Salmon in the Future. Bethesda, MD: American Fisheries
Society.

Blanckaert, K. A. Duarte, and A. J. Schleiss. 2010. Influence of Shallowness, Bank Inclination and Bank
Roughness on the Variability of Flow Patterns and Boundary Shear Stress due to Secondary
Currents in Straight Open-Channels. Advances in Water Resources 33(9):1062–1074.

Blaustein, Andrew R., and David B. Wake. 1990. Declining Amphibian Populations: A Global
Phenomenon? Trends in Ecology & Evolution 5 (7):203–204.

Boano, F., A. Demaria, R. Revelli, and L. Ridolfi. 2010. Biogeochemical Zonation due to Meander
Hyporheic Flow. Water Resources Research 46:W20511.

Bocchiola, D., M. C. Rulli, and R. Rosso. 2006. Transport of Large Woody Debris in the Presence of
Obstacles. Geomorphology 76(1):166–178.

Bocchiola, D., M. C. Rulli, and R. Rosso. 2008. A Flume Experiment on the Formation of Wood Jams in
Rivers. Water Resources Research 44: W02408, doi:10.1029/2006WR005846.

Boisclair, D. 2001. Fish Habitat Modeling: From Conceptual Framework to Functional Tools.
Canadian Journal of Fisheries and Aquatic Sciences 58:1–9.

Bolton, S., A. Watts, T. Sibley, and J. Dooley. 1998. A Pilot Study Examining the Effectiveness of
Engineered Large Woody Debris (ELWD™) as an Interim Solution to Lack of LWD in Streams.
EOS, Transactions of the American Geophysical Union 79(45):F346.

Boose, E. R., K. E. Chamberlin, and D. R. Foster. 2001. Landscape and Regional Impacts of Hurricanes
in New England. Ecological Monographs 71:27–48.

Booth, D. B. 1990. Stream-Channel Incision Following Drainage-Basin Urbanization. Water Resources


Bulletin 26:407–417.

Booth, D. 1991. Urbanization and the Natural Drainage System: Impacts, Solutions, and Prognoses.
The Northwest Environmental Journal 7, 93-118.

Large Wood National Manual July 2015


10-8
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Booth, D., and L. E. Reinelt. 1993. Consequences of Urbanization on Aquatic Systems—Measured


Effects, Degradation Thresholds, and Corrective Strategies. Watershed ’93, American Water
Resources Association, pages 545–550.

Borg, D., I. Rutherford, and M. Stewardson. 2007. The Geomorphic and Ecological Effectiveness of
Habitat Rehabilitation Works: Continuous Measurement of Scour and Fill around Large Logs in
Sand-Bed Streams. Geomorphology 89(1/2):205–216.

Boulton, A. J. 1989. Over-Summering Refuges of Aquatic Macroinvertebrates in Two Intermittent


Streams in Central Victoria. Transactions of the Royal Society of South Australia 113:23–34.

Boulton, A. J., T. Datry, T. Kasahara, M. Mutz, and J.A. Stanford. 2010. Ecology and Management of the
Hyporheic Zone – Groundwater Interactions of Running Waters and Their Floodplains. Journal
of the North American Benthological Society 29:26–40.

Bouwes, N., J. Moberg, N. Weber, B. Bouwes, S. Bennett, C. Beasley, C. E. Jordan, P. Nelle, M. Polino, S.
Rentmeester, B. Semmens, C. Volk, M. B. Ward, and J. White. 2011. Scientific Protocol for
Salmonid Habitat Surveys within the Columbia Habitat Monitoring Program. Prepared by
Terraqua, Inc., Wauconda, WA, for Integrated Status and Effectiveness Monitoring Program,

Bovee, K. D., T. J. Newcomb, and T. G. Coon. 1994. Relations Between Habitat Variability and
Population Dynamics of Bass in the Huron River, Michigan. Biological Report 21. National
Biological Survey, U.S. Department of the Interior. Washington D.C.

Bovee, K. D., B. L. Lamb, J. M. Bartholow, C. B. Stalnaker, J. Taylor, and J. Henriksen. 1998. Stream
Habitat Analysis using the Instream Flow Incremental Methodology. Washington, DC: U.S.
Geological Survey.

Boyce, M. S., and L. L. McDonald. 1999. Relating Populations to Habitats Using Resource Selection
Functions. Trends in Ecology and Evolution 14(7):268–272.

Bradford, M. J., J. Korman, and P. S. Higgins. 2005. Using Confidence Intervals to Estimate the
Response of Salmon Populations (Oncorhynchus spp.) to Experimental Habitat Alterations.
Canadian Journal of Fisheries and Aquatic Sciences 62(12):2716–2726.

Bradley, J., D. Richards, and C. Bahner 2005. Debris Control Structures – Evaluation and
Countermeasures. Hydraulic Engineering Circular No. 9. FHWA-IF-04-016. U.S. Department of
Transportation, Federal Highway Administration., Salem, OR.

Bragg, D. C. 2000. Simulating Catastrophic and Individualistic Large Woody Debris Recruitment for
a Small Riparian Ecosystem. Ecology 81:1383–1394.

Braudrick, C. A., and G. E. Grant. 2000. When do Logs Move in Rivers? Water Resources Research
36(2):571–583.

Braudrick, C. A., and G. E. Grant. 2001. Transport and Deposition of Large Woody Debris in Streams:
A Flume Experiment. Geomorphology 41:263–283.

Braudrick, C. A., G. E. Grant, Y. Ishikawa, and H. Ikeda. 1997. Dynamics of Wood Transport in
Streams: A Flume Experiment. Earth Surface Processes and Landforms 22:669–683.

Braun, A., K. Auerswald, and J. Geist. 2012. Drivers and Spatio-Temporal Extent of Hyporheic Patch
Variation: Implications for Sampling. PLOS One 7:e42046.

Large Wood National Manual July 2015


10-9
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Brenkman, S., J. Duda, C. E. Torgersen, E. Welty, G. R. Pess, R. Peters, and M. L. McHenry. 2012.
A Riverscape Perspective of Pacific Salmonids and Aquatic Habitats Prior to Large-Scale Dam
Removal in the Elwha River, Washington, USA. Fisheries Management and Ecology 19:36–53.

Bretherton, W. D., J. S. Kominoski, D. G. Fischer, and C. J. LeRoy. 2011. Salmon Carcasses Alter Leaf
Litter Species Diversity Effects on In-stream Decomposition. Canadian Journal of Fisheries and
Aquatic Sciences 68(8):1495–1506.

Briceño-Linares, J. M., J. P. Rodríguez, K. M. Rodríguez-Clark, F. Rojas-Suárez, P. A. Millán, E. G. Vitton,


and M. Carrasco-Muñoz. 2011. Adapting to changing poaching intensity of yellow-shouldered
parrot (Amazona barbadensis) nestlings in Margarita Island, Venezuela. Biological Conservation
144 (4):1188–1193.

Brigham, M. E., D. A. Wentz, G. R. Aiken, and D. P. Krabbenhoft. 2009. Mercury Cycling in Stream
Ecosystems. 1. Water Column Chemistry and Transport. Environmental Science and Technology
43:2720–2725.

Brooks, A. P. 2006. Design Guidelines for the Reintroduction of Wood into Australian Streams. Land &
Water Australia, Canberra.

Brooks, A. P., and G. J. Brierly. 2002. Mediated Equilibrium: The Influence of Riparian Vegetation and
Wood on the Long-Term Evolution and Behavior of a Near-Pristine River. Earth Surface
Processes and Landforms 27:343–367.

Brooks, A., and G. J. Brierly. 2004. Framing Realistic River Rehabilitation Targets in Light of Altered
Sediment Supply and Transport Relationships: Lessons from East Gippsland, Australia.
Geomorphology 58:107–123.

Brooks, A. P., G. J. Brierly, and R. G. Millar. 2003. The Long-Term Control of Vegetation and Woody
Debris on Channel and Flood-Plain Evolution: Insights from a Paired Catchment Study in
Southeastern Australia. Geomorphology 51:7–30.

Brooks, A. P., P. Gehrke, J. D. Jansen, and T. B. Abbe. 2004. Experimental Reintroduction of Woody
Debris on the Williams River, NSW: Geomorphic and Ecological Responses. River Research and
Applications 20:513–536.

Brooks, A. P., T. Howell, T. B. Abbe, and A. H. Arthington. 2006a. Confronting Hysteresis: Wood Basin
River Rehabilitation in Highly Altered Riverine Landscapes of South-Eastern Australia.
Geomorphology 79(3/4):395–422.

Brooks, A. P., T. Abbe, T. Cohen, N. Marsh, S. Mika, A. Boulton, T. Broderick, D. Borg, and I. Rutherfurd
2006b. Design Guidelines for the Reintroduction of Wood into Australian Streams. Land & Water
Australia, Canberra, Australia.

Brosofske, K. D., J. Chen, R. J. Naiman, and J. F. Franklin. 1997. Harvesting Effects on Microclimatic
Gradients from Small Streams to Uplands in Western Washington. Ecological Applications
7(4):1188–1200.

Brown, S. A. and E. S. Clyde. (1989). Design of Riprap Revetment. Hydraulic Engineering Circular 11,
Publication No. FHWA-IP-89-016, Federal Highway Administration, U.S. Department of
Transportation, Washington, D.C.

Large Wood National Manual July 2015


10-10
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Brummer, C., T. B. Abbe, J. R. Sampson, and D. R. Montgomery. 2006. Influence of Vertical Channel
Change Associated with Wood Accumulations on Delineating Channel Migration Zones,
Washington State, USA. Geomorphology 80:295–309.

Bryant, M. D. 1980. Evolution of large, Organic Debris after Timber Harvest: Maybeso Creek, 1949 to
1978. USDA Forest Service, General Technical Report, PNW-101.

Bryant, M. D. 1983. The Role and Management of Woody Debris in West Coast Salmonid Nursery
Stream. North American Journal of Fisheries Management 3(3):322–330.

Bryant, M. D., and J. R. Sedell. 1995. Riparian Forests, Wood in the Water, and Fish Habitat
Complexity. Pages 202–224 in N. B. Armantrout and R. J. Wolotira, Jr. (eds.), Conditions of the
World's Aquatic Habitats. Proceedings of the World Fisheries Congress Theme 1. Oxford and IBH
Publishing Co. Pvt. Ltd., New Delhi.

Buckley B. M., and F. J. Triska 1978. Presence and Ecological Role of Nitrogen-Fixing Bacteria
Associated with Wood Decay in Streams. Internationale Vereinigung für Theoretische und
Angewandte Limnologie Verhandlungen 20:1333–1339.

Buffington, J. M., and D. R. Montgomery. 1999a. A Procedure for Classifying Textural Facies in
Gravel-bed Rivers. Water Resources Research 35(6):1903-1914.

Buffington, J. M., and D. R. Montgomery. 1999b. Effects of Hydraulic Roughness on Surface Textures
of Gravel-Bed Rivers. Water Resources Research 35(11):3507–3521.

Buffington, J. M., and D. R. Montgomery. 1999c. Effects of Sediment Supply on Surface Textures of
Gravel-Bed Rivers. Water Resources Research 35(11):3523–3530.

Buffington, J. M. and D. Tonina. 2009. Hyporheic Exchange in Mountain Rivers II: Effects of Channel
Morphology on Mechanics, Scales, and Rates of Exchange. Geography Compass 3:1038–1062.

Buffington, J. M., T. E. Lisle, R. D. Woodsmith, and S. Hilton. 2002. Controls on the Size and
Occurrence of Pools in Coarse-Grained Forest Rivers. River Research and Applications 18:507–
531.

Burchsted, D., M. Daniels, R. Thorson, and J. Vokoun. 2010. The River Discontinuum: Beaver
Modifications to Baseline Conditions for Restoration of Forested Headwaters. BioScience
60(11):908–922.

Burkholder, B. K., G. E. Grant, R. Haggerty, T. Khangaonkar, and P. J. Wampler. 2008. Influence of


Hyporheic Flow and Geomorphology on Temperature of a Large, Gravel-Bed River, Clackamas
River, Oregon, USA. Hydrological Processes 22:941–953.

Bury, R. B., and P. Stephen Corn. 1991. Sampling Methods for Amphibians in Streams in the Pacific
Northwest. Available: http://www.treesearch.fs.fed.us/pubs/3069.

Cafferata, P., T. Spittler, M. Wopat, G. Bundros, and S. Flanagan. 2004. Designing Watercourse
Crossings for Passage of 100-Year Flood Flows, Wood, and Sediment. California Forestry Report
No.1. California Department of Forestry and Fire Protection. Available:
http://www.fire.ca.gov/php/rsrc-mgt_forestpractice_pubsmemo.php.

Camp, A., C. Oliver, P. Hessburg, and R. Everett. 1996. Predicting Late-Successional Fire Refugia
Pre-Dating European Settlement in the Wenatchee Mountains. USDA PNW, Wenatchee For. Sci.

Large Wood National Manual July 2015


10-11
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Lab., Univ. of Washington, Seattle. Elsevier Science Publishers B.V. Forest Ecology and
Management 95:63–77.

Cannon, S. H., and J. DeGraff. 2009. The Increasing Wildfire and Post-Fire Debris-Flow Threat in
Western USA, and Implications for Consequences of Climate Change. Pages 177–190 in K. Sassa
and P. Canuti (eds.), Landslides—Disaster Risk Reduction. Berlin Heidelberg: Springer-Verlag.
Available: http://landslides.usgs.gov/docs/cannon/Cannon_Degraff_2008_Springer.pdf.

Cannon, S. H., J. E. Gartner, M. G. Rupert, J. A. Michael, A. H. Rea, and C. Parrett. 2010. Predicting the
Probability and Volume of Postwildfire Debris Flows in the Intermountain Western United
States. Geological Society of America Bulletin 122(1-2):127–144. doi:10.1130/B26459.1.

Carignan, V., and M.-A. Villard. 2002. Selecting Indicator Species to Monitor Ecological Integrity:
A Review. Environmental Monitoring and Assessment 78(1):45–61.

Carignan, V., and M. A. Villard. 2004. Biological Indicators in Environmental Monitoring Programs:
Can We Increase Their Effectiveness. Environmental Monitoring:567–582.

Carlson, J. Y., C. W. Andrus, and H. A. Froehlich. 1990. Woody Debris, Channel Features, and
Macroinvertebrates of Streams with Logged and Undisturbed Riparian Timber in Northeastern
Oregon, USA. Canadian Journal of Fisheries and Aquatic Sciences 47:1103–1111.

Caro, T. Mi, and G. O'Doherty. 1999. On the Use of Surrogate Species in Conservation Biology.
Conservation Biology 13(4):805–814.

Castelle, A. J., A. W. Johnson, and C. Conolly. 1994. Wetland and Stream Buffer Size Requirements—
A Review. Journal of Environmental Quality 23(5):878–882.

Castro, J., and R. Sampson. 2001. Incorporation of Large Wood into Engineering Structures. Natural
Resource Conservation Service Engineering Technical Note Number 15. U.S. Department of
Agriculture. Boise, ID.

Cederholm, C. J., W. J. Scarlett, N. P. and Peterson. 1988. Low-Cost Enhancement Technique for
Winter Habitat of Juvenile Coho Salmon. North American Journal of Fisheries Management
8:438–441.

Cederholm, C. J., R. E. Bilby, P. A. Bisson, T. W. Bumstead, B .R. Fransen, W. J. Scarlett, and J. W. Ward.
1997a. Response of Juvenile Coho Salmon and Steelhead to Placement of Large Woody Debris in
a Coastal Washington Stream. North American Journal of Fisheries Management 17:947–963.

Cederholm, C. J., R. E. Bilby, P. A. Bisson, T. W. Bumstead, B. R. Fransen, W. J. Scarlett, and J. W. Ward.


1997b. Response of Juvenile Coho Salmon and Steelhead to the Placement of Large Woody
Debris in a Coastal Washington Stream. Transactions of the American Fisheries Society. 118:368–
378.

Cederholm, C. J., L. G. Dominguez, and T. W. Bumstead. 1997c. Rehabilitating Stream Channels and
Fish Habitat Using Large Woody Debris. In P. A. Slaney and D. Zaldokas (eds.), Fish Habitat
Procedures. Watershed Restoration Program, Ministry of Environment, Lands and Parks,
Vancouver, British Columbia.

Cederholm, C. J., M. D. Kunze, T. Murota, and A. Sibatani. 1999. Pacific salmon Carcasses: Essential
Contributions of Nutrients and Energy for Aquatic and Terrestrial Ecosystems. Fisheries
24(10):6–15.

Large Wood National Manual July 2015


10-12
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Chambers, J. Q., J. I. Fisher, H. Zeng, E. L. Chapman, D. B. Baker, and G. C. Hurtt. 2007. Hurricane
Katrina’s Carbon Footprint on U.S. Gulf Coast Forests. Science 318 (5853):1107.

Chapman, D. W. 1966. Food and Space as Regulators of Salmonid Populations in Streams. American
Naturalist 100:345–357.

Chen, J., J. F. Franklin, and T. A. Spies. 1995. Growing-Season Microclimatic Gradients from Clearcut
Edges into Old-Growth Douglas-Fir Forests. Ecological Applications 5(1):74–86.

Chesney, C. 2000. Functions of Wood in Small, Steep Streams in Eastern Washington: Summary of
Results for Project Activity in the Ahtanum, Cowiche, and Tieton Basins. Washington Department
of Natural Resources. Prepared for the Timber/Fish/Wildlife Monitoring Advisory Group and
the Northwest Indian Fisheries Commission. TFW Effectiveness Monitoring Report: TFW-MAGl-
00-002.

Chin, A., M. D. Daniels, M. A. Urban, H. Piegay, K. J. Gregory, W. Bigler, A. Z. Butt, J. L. Grable, S. V.


Gregory, M. Lafrenz, L. R. Laurencio, and E. Wohl. 2008. Perceptions of Wood in Rivers and
Challenges for Stream Restoration in the United States. Environmental Management 41:893–903.

Christensen, N. L., and 12 others. 1996. The Report of the Ecological Society of America Committee
on the Scientific Basis for Ecosystem Management. Ecological Applications 6(3):665–691.

Claussens, L., C. L. Tague, P. M. Groffman, and J. M. Melack. 2010. Longitudinal and Seasonal Variation
of N Uptake in an Urbanizing Watershed: Effect of organic Matter, Stream Size, Transient Storage
and Debris Dams. Biogeochemistry 98:45–62.

Clay, C. 1949. The Colorado River Raft. The Southwestern Historical Quarterly 102 (4):400–426.

Clemen, R. T., and T. Reilly. 2001. Making Hard Decisions with Decision Tools. South-Western Cengage
Learning.

Coe, H. J., P. M. Kiffney, G. R. Press, K. K. Kloehn, and M. L. McHenry. 2009. Periphyton and
Invertebrate Response to Wood Placement in Large Pacific Coastal Rivers. River Research and
Applications 25(8):1025–1035.

Coho, C., and S. J. Burges. 1993. Dam-Break Floods in Low Order Mountain Channels of the PNW.
Water Resources Series Tech Rep no. 138. Dept. Civil Engineering, Univ. of Washington, Seattle.

Coho, C., and S. J. Burges. 1994. Dam Break Floods in Low Order Mountain Channels of the Pacific
Northwest. TFW SH9 93 001. Timber Fish and Wildlife, Department of Natural Resources,
Olympia.

Colburn, K. 2011. Integrating Recreational Boating Considerations into Stream Channel Modification
and Design Projects. American Whitewater (2011).

Collins, B. D., and A. J. Sheikh. 2005. Historical Reconstruction, Classification, and Change Analysis of
Puget Sound Tidal Marshes. University of Washington (Seattle, WA) and the Nearshore Habitat
Program, Washington State Dept. of Natural Resources, Olympia, WA. See more at:
http://www.eopugetsound.org/science-review/3-tidal-wetlands#sthash. T4OyhfFd.dpuf

Collins, B. D., and D. R. Montgomery. 2002. Forest Development, Wood Jams, and Restoration of
Floodplain Rivers in the Puget Lowland, Washington. Restoration Ecology 10:237–247.

Large Wood National Manual July 2015


10-13
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Collins, B. D., D. R. Montgomery, and A. D. Haas. 2002. Historical Changes in the Distribution and
Functions of Large Wood in Puget Lowland Rivers. Canadian Journal of Fisheries and Aquatic
Sciences 59:66–76.

Collins, B. D., D. R. Montgomery, and A. J. Sheikh. 2003. Reconstructing the Historical Riverine
Landscape of the Puget Lowland. Pages 79–128 in D. R. Montgomery, S. M. Bolton, D. B. Booth,
and L. Wall (eds.), Restoration of Puget Sound Rivers. University of Washington Press, Seattle.

Collins, B. D., D. R. Montgomery, K. L. Fetherston, and T. B. Abbe. 2012. The Floodplain Large-Wood
Cycle Hypothesis: A Mechanism for the Physical and Biotic Structuring of Temperate Forested
Alluvial Valleys in the North Pacific Coastal Ecoregion. Geomorphology 139/140:460–470.

Comiti, F. and Mao, L. 2012. Recent Advances in the Dynamics of Steep Channels. Gravel-bed Rivers:
Processes, Tools, Environments, pages 351–377.

Comiti, F., A. Andreoli, M. A. Lenzi, and L. Mao. 2006. Spatial Density and Characteristics of Woody
Debris in Five Mountain Rivers of the Dolomites (Italian Alps). Geomorphology 78:44–63.

Comiti, F., A. Andreoli, L. Mao, and M. A. Lenzi. 2008. Wood Storage in Three Mountain Streams of the
Southern Andes and its Hydro-Morphological Effects. Earth Surface Processes and Landforms
33:244–262.

Compton, J. E., M. R. Church, S. T. Larned, and W. E. Hogsett. 2003. Nitrogen Export from Forested
Watersheds in the Oregon Coast Range: The Role of N2-Fixing Red Alder. Ecosystems 6(8):773–
785.

Conquest, L. L., and S. C. Ralph. 1998. Statistical Design and Analysis Considerations for Monitoring
and Assessment. Pages 455-475 in R. J. Naiman and R. E. Bilby (eds.), River Ecology and
Management: Lessons from the Pacific Coastal Ecoregion. New York: Springer.

Convertino, M., K. M. Baker, J. T. Vogel, C. Lu, B. Suedel, and I. Linkov. 2013. Multi-Criteria Decision
Analysis to Select Metrics for Design and Monitoring of Sustainable Ecosystem Restorations. U.S.
Army Research. Paper 190. Available: http://digitalcommons.unl.edu/usarmyresearch/1905.

Cook, W. J. 2014. Bridge Failure Rates, Consequences, and Predictive Trends. PhD Dissertation. Utah
State University, Logan, UT.

Copeland, R. R. 1983. Bank Protection Techniques Using Spur Dikes. Paper No. HL-83-1. Hydraulics
Laboratory. U.S. Army Waterways Experiment Station. Vicksburg, MS.

Cordova, J. M., E. J. Marshall-Rosi, A. M. Yamamoto, and G. A. Lamberti. 2007. Quantity, Controls, and
Functions of Large Woody Debris in Midwestern USA Streams. River Research and Applications
23(1):21–33.

Costa, J. E. 1984. Physical Geometry of Debris Flows. Pages 268–317 in J. E. Costa and P. J. Fleisher
(eds.), Developments and Applications of Geomorphology. Springer-Verlag, Berlin.

Cramer, M. L. (ed). 2012. Stream Habitat Restoration Guidelines. Copublished by the Washington
Departments of Fish and Wildlife, Natural Resources, Transportation and Ecology, Washington
State Recreation and Conservation Office, Puget Sound Partnership, and the U.S. Fish and
Wildlife Service. Olympia, WA.

Large Wood National Manual July 2015


10-14
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Cramer, S. P., and N. K. Ackerman. 2009. Prediction of Stream Carrying Capacity for Steelhead; The
Unit Characterization Method. Pages 255–288 in Pacific Salmon Environment and Life History
Models. Bethesda, MD: American Fisheries Society.

Cramer, M., K. Bates, D. Miller, K. Boyd, L. Fotherby, P. Skidmore, T. Hoitsma, B. Heiner, K. Buchanan,
P. Powers, G. Birkeland, M. Rotar, and D. White. 2002. Integrated Streambank Protection
Guidelines. Washington State Aquatic Habitat Guidelines Program, Washington State Department
of Fish and Wildlife. Olympia, WA.

Creamean, J. M., K. J. Suski, D. Rosenfeld, A. Cazorla, P. J. DeMott, R. C. Sullivan, A. B. White, F. M.


Ralph, P. Minnis, J. M. Comstock, J. M. Tomlinson, and K. A. Prather. 2013. Dust and Biological
Aerosols from the Sahara and Asia Influence Precipitation in the Western U.S. Science 339:
1572–1578, doi:10.1126/ science.1227279.

Crispell, J. K., and T. A. Endreny. 2009. Hyporheic Exchange Flow around Constructed In-Channel
Structures and Implications for Restoration Design. Hydrological Processes 23:1158–1168.

Crook, D., and A. Robertson. 1999. Relationships between Riverine Fish and Woody Debris:
Implications for Lowland Rivers. Marine and Freshwater Research 50:941–953.

Cummins, K. W. 1974. Structure and Function of Stream Ecosystems. BioScience 24(11):631–641.

Cummins, K. W., G. W. Minshall, J. R. Sedell, C. E. Cushing, and R. C. Petersen. 1984. Stream Ecosystem
Theory. Internationale Vereinigung fur theoretische und angewandte Limnologie, Verhandlungen
22:1818–1827.

Curran, J. C. 2010. Mobility of Large Woody Debris (LWD) Jams in a Low Gradient Channel.
Geomorphology 116:320–329.

Curran, J. H., and E. E. Wohl 2003. Large Woody Debris and Flow Resistance in Step-Pool Channels,
Cascade Range, Washington. Geomorphology 51:141–157.

Cushman, M. J. 1981. The Influence of Recurrent Snow Avalanches on Vegetation Patterns in the
Washington Cascades. Ph.D. dissertation. University of Washington, Seattle, Washington.

D'Aoust, S. G. and R. G. Millar. 2000. Stability of Ballasted Woody Debris Habitat Structures. Journal
of Hydraulic Engineering 126(11):810–817.

Dacy, G. H. 1921. Pulling the Mississippi’s Teeth. Scientific American 75(4):60, 70.

Daley, J. S. 2012. Taming the Hungry Beast: The Effectiveness of Engineered Log Jams in an Incising
Gravel Bedded River. B.S. Honors Thesis, School of Earth and Environmental Science, University
of Wollongong, Australia. Available: http://ro.uow.edu.au/thsci/38.

Danehy, R. J., and B. J. Kirpes. 2000. Relative Humidity Gradients across Riparian Areas in Eastern
Oregon and Washington Forests. Northwest Science 74(3):224–233.

Daniels, M. D., and Rhoads, B. L. 2004. Effect of Large Woody Debris Configuration on
Three-Dimensional Flow Structure in Two Low-Energy Meander Bends at Varying Stages. Water
Resources Research (40):W11302.

Daniels, M. D., and B. L. Rhoads. 2004. Spatial Pattern of Turbulence Kinetic Energy and Shear Stress
in a Meander Bend with Large Woody Debris. Pages 87–97 in S. J. Bennett and A. Simon (eds.),

Large Wood National Manual July 2015


10-15
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Riparian Vegetation and Fluvial Geomorphology, Water Science and Application 8. American
Geophysical Union, Washington D.C.

Daniels, M. D., and B. Rhoads. 2007. Influence of Experimental Removal of Large Woody Debris on
Spatial Patterns of Three-Dimensional Flow in a Meander Bend. Earth Surface Processes and
Landforms 32:460–474.

Davis, J. C., G. Minshall, W. Robinson, T. Christopher, and P. Landres. 2001. Monitoring Wilderness
Stream Ecosystems. Gen. Tech. Rep. RMRS-GTR-70. Ogden, UT: U.S. Department of Agriculture,
Forest Service, Rocky Mountain Research Station.

Davis, J. M., C. V. Baxter, E. J. Rosi-Marshall, J. L. Pierce, and B. T. Crosby. 2013. Anticipating Stream
Ecosystem Responses to Climate Change: Toward Predictions that Incorporate Effects via Land–
Water Linkages. Ecosystems 16:909–922. DOI:10.1007/s10021-013-9653-4.

Davis, W. M. 1901. Physical Geography. Boston, MA: Ginn and Company.

Dean, D. J., M. L. Scott, P. B. Shafroth, and J. C. Schmidt. 2011. Stratigraphic, Sedimentologic, and
Dendrogeomorphic Analyses of Rapid Floodplain Formation Along the Rio Grande in Big Bend
National Park, Texas. Geological Society of America Bulletin 123:1908–1925.

DeBano, L. F., S. J. DeBano, D. E. Wooster, and M. B. Baker, Jr. 2004. Linkages between Riparian
Corridors and Surrounding Watersheds. Pages 77–97 in M. B. Baker, Jr., P. E. Ffolliott, L. F.
DeBano, and D. G. Neary (eds.), Riparian Areas of the Southwestern United States. Lewis
Publishers.

Derrick, D. L. 1997. Twelve low-Cost, Innovative, Landowner Financed, Streambank Protection


Demonstration Projects. Pages 446–451 in Management of Landscapes Disturbed by Channel
Incision, Stabilization, Rehabilitation, and Restoration. Center for Computational Hydroscience
and Engineering, University of Mississippi.

Dickman, A., and S. Cook. 1989. Fire and Fungus in a Mountain Hemlock Forest. Canadian Journal of
Botany 67:2005–2016.

Diehl, T. H. 1997. Potential Drift Accumulation at Bridges. Publication FHWA-RD-97-028.


U.S. Department of Transportation, McLean, VA.

Dixit, S. S., J. P. Smol, J. C. Kingston, and D. F. Charles. 1992. Diatoms: Powerful Indicators of
Environmental Change. Environmental Science & Technology 26 (1):22–33.

Doloff, C. A., and M. L. Warren, Jr. 2003. Fish Relationships With Large Wood in Small Streams.
American Fisheries Symposium 37:179–193.

Dominguez, L. G., and C. J. Cederholm. 2000. Rehabilitating Stream Channels Using Large Woody
Debris with Considerations for Salmonid Life History and Fluvial Geomorphic Processes. Pages
545–563 in E. E. Knudsen. C. R. Steward, D. D. MacDonald, J. E. Williams, and D. W. Reiser (eds.),
Sustainable Fisheries Management: Pacific Salmon. Lewis Publishers, New York.

Doremus, H. 2001. Adaptive Management, the Endangered Species Act, and the Institutional
Challenges of “New Age” Environmental Protection. Washburn Law Journal 41:50–89.

Doremus, H. 2004. The Purposes, Effects, and Future of the Endangered Species Act's Best Available
Science Mandate. Environmental Law 34:397.

Large Wood National Manual July 2015


10-16
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Downs, P. W., and G. M. Kondolf. 2002. Post-project Appraisals in Adaptive Management of River
Channel Restoration. Environmental Management 29(4):477–496.

Doyle, M. W., and F. D. Shields, Jr. 2000. Incorporation of Bed Texture into a Channel Evolution
Model. Geomorphology 34:291–309.

Doyle, M. W., E. H. Stanley, J. M. Harbor, and G. S. Grant, G.S. 2003. Dam Removal in the United States:
Emerging Needs for Science and Policy. Eos, Transactions American Geophysical Union 84:29.

Draut, A. E., J. B. Logan, and M. C. Mastin, M.C. 2011. Channel Evolution on the Dammed Elwha River,
Washington, USA. Geomorphology 127:71–87.

Drury, T. A. 1999. Stability and Pool Scour of Engineered Log Jams in the North Fork Stillaguamish
River, Washington. Thesis, Master of Science in Civil Engineering. University of Washington,
Seattle.

Drury, T. A., C. Petroff, T. B. Abbe, D. R. Montgomery, and G. R. Pess. 1999. Evaluation of Engineered
Log Jams as a Soft Bank Stabilization Technique: North Fork Stillaguamish River, Washington.
Proceedings of Annual Conference, Water Resources Engineering. American Society of Civil
Engineers. Reston, VA.

Dunkerley, D. 2014. Nature and Hydro-Geomorphic Roles of Trees and Woody Debris in a Dryland
Ephemeral Stream: Fowlers Creek, Arid Western New South Wales, Australia. Journal of Arid
Environments 102:40-49.

Dunne, T., and L. B. Leopold. 1978. Water in Environmental Planning. New York, NY: W.H. Freeman
& Co.

Dwire, K. A., and J. B. Kauffman. 2003. Fire and Riparian Ecosystems in Landscapes of the Western
USA. Forest Ecology and Management 178:61–74.

East, A. E., G. R. Pess, J. A. Bountry, C. S. Magirl, A. C. Ritchie, J. B. Logan, T. J. Randle, M. C. Mastin,


J. T. Minear, J. J. Duda, M. C. Liermann, M. L. McHenry, T. J. Beechie, and P. B. Shafroth. 2014.
Large-Scale Dam Removal on the Elwha River, Washington, USA: River Channel and Floodplain
Geomorphic Change. Geomorphology 228:765–786.

Eaton, B. C. 2006. Bank Stability Analysis for Regime Models of Vegetated Gravel Bed Rivers. Earth
Surface Processes and Landforms 31:1438–1444.

Eaton, B. C., M. Chuch, and R. G. Millar. 2004. Rational Regime Model of Alluvial Channel Morphology
and Response. Earth Surface Processes and Landforms 29:511–529.

Eaton, B.C., R. C. Millar, and S. Davidson. 2010. Channel Patterns: Braided, Anabranching and Single
Thread. Geomorphology 120:353–364.

Eaton, J. M., and D. Lawrence. 2006. Woody Debris Stocks and Fluxes During Succession. Forest
7Ecology and Management 232(1-3):46–55.

Eaton, B. C., M. A. Hassan, and S. L. Davidson. 2012. Modeling Wood Dynamics, Jam Formation, and
Sediment Storage in a Gravel-Bed Stream. Journal of Geophysical Research 117:F00A05,
doi:10.1029/2012JF002385.

Large Wood National Manual July 2015


10-17
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Edmonds, R. L., T. B. Thomas, and K. P. Maybury. 1993. Tree Population Dynamics, Growth, and
Mortality in old-Growth Forests in the Western Olympic Mountains, Washington. Canadian
Journal of Forest Research 23:512–519.

Eggert, S. L., and J. B. Wallace. 2007. Wood Biofilm as a Food Resource for Stream Detritivores.
Limnology and Oceanography 52(3):1239–1245.

Ehrman, T. P., and G. A. Lamberti. 1992. Hydraulic and Particulate Matter Retention in a 3rd-Order
Indiana Stream. Journal of the North American Benthological Society. 11:341–349.

Elliot, R., D. Froehlich, and R. MacArthur. 2012. Calculating the Potential Effects of Large Woody
Debris Accumulations on Backwater, Scour, and Hydrodynamic Loads. Pages 1213–1222 in
Proceedings of the World Environmental and Water Resources Congress 2012. Reston, VA:
American Society of Civil Engineers.

Elmore, W., and R. L. Beschta. 1988. The Fallacy of Structures and the Fortitude of Vegetation.
Proceedings of California Riparian Systems Conference. Davis, Calif.

Elosegi, A., J. Diez, and J. Pozo. 2007. Contribution of Dead Wood to the Carbon Flux in Forested
Streams. Earth Surface Processes & Landforms 32(8):1219–1228.

Elzinga, C. L., D. W. Salzer, and J. W. Willoughby. 1998. Measuring & Monitoring Plant Populations.
Denver, CO: Bureau of Land Management Technical Reference 1730-1.

Embertson, L, and J. Monahan. 2011. Public Safety Assessment of Habitat Enhancement Projects Fobes
and Skookum Reach Restoration Projects South Fork Nooksack River. GeoEngineers, Bellingham
Washinghton. March 1, 2011.

Endreny, T., L. Lautz, and D. I. Siegel. 2011. Hyporheic Flow Path Response to Hydraulic Jumps at
River Steps: Flume and Hydrodynamic Models. Water Resources Research 47:W02517.

Ensign, S. H., and M. W. Doyle. 2005. In-channel Transient Storage and Associated Nutrient
Retention: Evidence from Experimental Manipulations. Limnology and Oceanography
50(6):1740–1751.

Environmental Agency. 2009. The Hyporheic Handbook. a Handbook of the Groundwater-Surface


Water Interface and Hyporheic Zone for Environmental Managers. Science Report SC050070.
Available: http://www.hyporheic.net/SCHO1009BRDX-e-e.pdf. Accessed: June 13, 2014.

Erskine, W. D., and A. A. Webb (2003). Desnagging to Resnagging: New Directions in River
Rehabilitation in Southeastern Australia. River Research and Applications. 19:233–249.

Erskine, W. D., M. J. Saynor, A. C. Chalmers, and S. J. Riley. J. 2012. Water, Wind, Wood, and Trees:
Interactions, Spatial Variations, Temporal Dynamics, and their Potential Role in River
Rehabilitation. Journal of Geographical Research 50(1):60–74.

Eslamian, S. 2014. Handbook of Engineering Hydrology. Boca Raton, FL: CRC Press.

Evergreen Funding Consultants (EFC). 2003. A Primer on Habitat Project Costs. Available:
http://www.evergreenfc.com/section_services/resources/primer.pdf.

Fahnestock, G. R. 1976. Fires, Fuel, and Flora as Factors in Wilderness Management: The Pasayten
Case. Tall Timbers Fire Ecology Conf. 15:33–70.

Large Wood National Manual July 2015


10-18
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Fahnestock, G. R., and J. K. Agee. 1983. Biomass Consumption and Smoke Production by Prehistoric
and Modern Forest Fires in Western Washington. Journal of Forestry 81:653–657.

Falke, J. A., K. D. Fausch, R. Magelky, A. Aldred, D. S. Durnford, L. K. Riley, and R. Oad. 2011. The Role
of Groundwater Pumping and Drought in Shaping Ecological Futures for Stream Fishes in
a Dryland River Basin of the Western Great Plains, USA. Ecohydrology 4L682–697. doi:10.1002/
eco.158. Available: http://onlinelibrary.wiley.com/doi/10.1002/eco.158/pdf.

Fausch, K. D., C. E. Torgersen, C. V. Baxter, and H. W. Li. 2002. Landscapes to Riverscapes: Bridging
the Gap Between Research and Conservation of Stream Fishes. BioScience 52(6):483–498.

Faustini, J. M., and J. A. Jones. 2003. Influence of Large Woody Debris on Channel Morphology and
Dynamics in Steep, Boulder-Rich Mountain Streams, Western Cascades, Oregon. Geomorphology
51:187–205.

Federal Emergency Management Agency (FEMA). 1999. Riverine Erosion Hazard Areas; Mapping
Feasibility Study. FEMA Technical Services Division, Hazard Study Branch.

Federal Emergency Management Agency (FEMA). 2009. NFIP Floodplain Management Guidebook:
A Local Administrator’s Guide to Floodplain Management and the National Flood Insurance
Program. Fifth Edition, Federal Emergency Management Agency Region 10. Bothell, WA.

Federal Highway Administration (FHWA). 1985a. Design of Spur-Type Streambank Stabilization


Structures. Federal Highway Administration Report No. FHWA/RD-84/101. U.S. Department of
Transportation. Washington D.C.

Federal Highway Administration (FHWA) 1985b. Hydraulic Design of Highway Culverts. FHWA-IP-
58-15. Available: http://www.fhwa.dot.gov/engineering/hydraulics/pubs/12026/hif12026.pdf.

Federal Highway Administration (FHWA). 2001. Evaluating Scour at Bridges, Fourth Edition.
Hydraulic Engineering Circular No. 18. Publication No. FHWA NHI 01-001. Available:
http://www.stream.fs.fed.us/fishxing/fplibrary/FHWA_2001_Evaluating_Scour_at_Bridges.pdf.

Federal Highway Administration (FHWA). 2005. Debris Control Structures Evaluation and
Countermeasures. Hydraulic Engineering Circular No. 9. Publication No. FHWA-IF-04-016.
Available: http://www.fhwa.dot.gov/engineering/hydraulics/pubs/04016/.

Federal Highway Administration (FHWA). 2006. Construction Noise Handbook. Report numbers
FHWA-HEP-06-015, DOT-VNTSC-FHWA-06-02, NTIS No. PB2006-109102. U.S. Department of
Transportation, Research and Innovative Technology Administration, Cambridge, MA. Available:
http://www.fhwa.dot.gov/environment/noise/construction_noise/handbook/. Accessed: July
10, 2014.

Federal Highway Administration (FHWA). 2012a. Climate Change & Extreme Weather Vulnerability
Assessment Framework. FHWA Publication No: FHWA-HEP-13-005.

Federal Highway Administration (FHWA). 2012b. Stream Stability at Highway Structures. Hydraulic
Engineering Circular No. 20. Publication No. FHWA-HIF-12-004. Federal Highway
Administration, U.S. Department of Transportation, Washington, DC.

Federal Interagency Stream Restoration Working Group (FISRWG). 1998. Restoration


Implementation, Monitoring, and Management. Chapter 9 in Stream Corridor Restoration:
Principles, Processes and Practices. National Technical Information Service, U. S. Department of

Large Wood National Manual July 2015


10-19
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Commerce, Springfield, VA. Also published as NRCS, U.S. Department of Agriculture (1998)
National Engineering Handbook (NEH), Part 653. Washington, D.C.

Federal Writers’ Project. 1952. West Virginia A Guide to the Mountain State. U.S. History Publishers.

Fetherston, K. L., R. J. Naiman, and R. E. Bilby. 1995. Large Woody Debris, Physical Process, and
Riparian Forest Development in Montane River Networks of the Pacific Northwest.
Geomorphology 13:133–144. Elsevier Science B.V.

Fifield, J. S. 2011. Designing and Reviewing Effective Sediment and Erosion Control Plans. Third
Edition., Santa Barbara, CA: Forester Press.

Findlay, S., J. Tank, S. Dye, H. M. Valett, P. J. Mulholland, W. H. McDowell, S. L. Johnson, S. K. Hamilton,


J. Edmonds, W. K. Dodds, and W. B. Bowden. 2002. A Cross System Comparison of Bacterial and
Fungal Biomass in Detritus Pools of Headwater Streams. Microbial Ecology 43(1):55–66.

Findlay, S. E. G., R. L. Sinsabaugh, W. V. Sobczak, and M. Hoostal. 2003. Metabolic and Structural
Response of Hyporheic Microbial Communities to Variations In Supply of Dissolved Organic
Matter. Limnology and Oceanography 48:1608–1617.

Fischenich, C. 2001. Stability Thresholds for Stream Restoration Materials, Publication No. ERDC
TNEMRRP-SR-29. U.S. Army Engineer Research and Development Center, Vicksburg, MS.

Fischenich, C., and J.V. Morrow, Jr. 2000. Streambank Habitat Enhancement with Large Woody Debris.
Publication No. ERDC TN-EMRRP-SR-13. U.S. Army Engineer Research and Development Center.

Fischenich, J. C. 2001. Plant Material Selection and Acquisition. EMRRP Technical Notes Collection
(ERDC TNEMRRP-SR-33), U.S. Army Engineer Research and Development Center, Vicksburg, MS.

Fisher, G. B., F. J. Magilligan, J. M. Kaste, and K. H. Nislow. 2010. Constraining the Timescales of
Sediment Sequestration Associated with Large Woody Debris using Cosmogenic 7Be. Journal of
Geophysical Research 115 (F3), DOI: 10.1029/2009JF001352.

Flanagan, S. A. 2004. Woody Debris Transport Through Low-Order Stream Channels of Northwest
California – Implications for Road-Stream Crossing Failure. M.S. Thesis. Humboldt State
University, Arcata, CA. Available: http://www.bof.fire.ca.gov/board/
msg_supportedreports.html.

Flanagan, S. A. 2005. Woody Debris Transport at Road-Stream Crossings. Stream Notes. Rocky
Mountain Research Station. U.S. Forest Service. Fort Collins, CO. October 2005. Available:
http://www.stream.fs.fed.us/news/streamnt/pdf/SNOct_05.pdf.

Flebbe, P. A. 1999. Trout Use of Wood Debris and Habitat in Wine Spring Creek, North Carolina.
Forest Ecology and Management 114:367–376.

Flores, L., A. Larranaga, J. Diez, and A. Elosegi. 2011. Experimental Wood Addition in Streams: Effects
on Organic Matter Storage and Breakdown. Freshwater Biology 56(10):2156–2167.

Flosi, G., S. Downie, J. Hopelain, M. Bird, R. Coey, and B. Collins (eds.). 1998. California Salmonid
Stream Habitat Restoration Manual. 3rd ed. California: California Department of Fish and Game,
Inland Fisheries Division, Sacramento.

Large Wood National Manual July 2015


10-20
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Flynn, K.M., Kirby, W.H., and Hummel, P.R. 2006. User’s Manual for Program PeakFQ Annual
Flood-Frequency Analysis Using Bulletin 17B Guidelines: U.S. Geological Survey, Techniques and
Methods. Book 4, Chapter B4.

Fore, L. S., J. R. Karr, and R. W. Wisseman. 1996. Assessing Invertebrate Responses to Human
Activities: Evaluating Alternative Approaches. Journal of the North American Benthological
Society 15(2):212–231.

Forest Ecosystem Management Assessment Team (FEMAT). 1993. Forest Ecosystem Management:
An Ecological, Economic, and Social Assessment. Report of the Forest Ecosystem Management
Assessment Team. July.

Forest Products Laboratory. 2010. Wood Handbook—Wood as an Engineering Material. General


Technical Report FPL-GTR-190. U.S. Department of Agriculture, Forest Service, Forest Products
Laboratory. Madison, WI.

Foster, D. R., and E. R. Boose. 1992. Patterns of Forest Damage Resulting from Catastrophic Wind in
Central New England, USA. Journal of Ecology 80:79–98.

Fox, M. J. 2001. A New Look at the Quantities and Volumes of Instream Wood in Forested Basins within
Washington State. Master of Science thesis. College of Forest Resources, University of
Washington.

Fox, M. J. 2003. Spatial Organization, Position, and Source Characteristics of Large Woody Debris in
Natural Systems. Ph.D. dissertation. College of Forest Resources, University of Washington.
Seattle, Washington.

Fox, M. J. and S. Bolton. 2007. A Regional and Geomorphic Reference for Quantities and Volumes of
Instream Wood in Unmanaged Forested Basins of Washington State. North American Journal of
Fisheries Management 27:342–359.

Frangi, J. L., and A. E. Lugo. 1991. Hurricane Damage to a Flood Plain Forest in the Luquillo
Mountains of Puerto Rico. Biotropica 23(4a):324–335.

Franklin, J. F., and C. T. Dyrness. 1973. Natural Vegetation of Oregon and Washington. USDA Forest
Service. Gen. Tech. Rep. PNW-8.

Fremier, A. K., J. I. Seo, and F. Nakamura. 2010. Watershed Controls on the Export of Large Wood
from Stream Corridors. Geomorphology 117:33–43.

Freschet, G. T., J. T. Weedon, R. Aerts, J. R. van Hal, and J. H. Cornelissen. 2012. Interspecific
Differences in Wood Decay Rates: Insights from a New Short-Term Method to Study Long-Term
Wood Decomposition. Journal of Ecology 100:161–170. doi: 10.1111/j.1365-2745.2011.01896.x.

Friedman, J. M., G. T. Auble, P. B. Shafroth, M. L. Scott, M. F. Merigliano, M. D. Preehling, and


E. K.Griffin. 2005. Dominance of Non-Native Riparian Trees in Western USA. Biological Invasions
7:747–751.

Frissell, C. A., and R. K. Nawa. 1992. Incidence and Causes of Physical Failure of Artificial Habitat
Structures in Streams of Western Oregon and Washington. North American Journal of Fisheries
Management 12 182–197.

Large Wood National Manual July 2015


10-21
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Frissell, W. J., W. J. Liss, C. E. Warren, and M. D. Hurley. 1986. A Hierarchical Framework for Stream
Habitat Classification: Viewing Streams in a Watershed Context. Environmental Management
10(2):199–214.

Furniss, M., T. Ledwith, M. Love, B. McFadin, and S. Flanagan 1998. Response of Road-Stream
Crossings to Large Flood Events in Washington, Oregon, and Northern California. USDA-Forest
Service, Technology & Development Program, Corvallis OR.

Gandy, C. J., and A. P. Jarvis. 2006. Attenuation of Nine Pollutants in the Hyporheic Zone. Environment
Agency, Bristol, England, June.

Gandy, C. J., J. W. N. Smith, and A. P. Jarvis. 2007. Attenuation of Mining-Derived Pollutants in the
Hyporheic Zone: A Review. Science of the Total Environment 373:435–446.

Garshelis, D. L. 2000. Delusions in Habitat Evaluation: Measuring Use, Selection, and Importance.
Pages 11–165 in L. Boitani and T. K. Fuller (eds.), Research Techniques in Animal Ecology. New
York: Columbia University Press.

Gastaldo, R. A., and C. W. Degges. 2007. Sedimentology and Paleontology of a Carboniferous Log Jam.
International Journal of Coal Geology 69:103–113.

Gastaldo, R. A., and T. M. Demko. 2011. The Relationship Between Continental Landscape Evolution
and the plant-Fossil Record: Long Term Hydrologic Controls on Preservation. Pages 249–285 in
P. A. Allison and D. J. Bottjer (eds.), Taphonomy: Process and Bias Through Time. Aims & Scope
Topics in Geobiology Voume 32. Springer Netherlands.

Gerecht, K. E., M. B. Cardenas, A. J Guswa, A. H. Sawyer, J. D. Nowinski, and T. E. Swanson. 2011.


Dynamic Hyporheic Flow and Heat Transport across a Bed-to-Bank Continuum in a Large
Regulated River. Water Resources Research 47:W03524.

Ghosn, M., F. Moses, and J. Wang. 2003. Design of Highway Bridges for Extreme Events. NCHRP
(National Cooperative Highway Research Program) Report 489. National Transportation Board.
Washington D.C. Available: http://www.national-academies.org/trb/bookstore.

Gibling, M. R., and N. S. Davies. 2012. Palaeozoic Landscapes Shaped by Plant Evolution. Nature
Geoscience 5(2):99–105.

Gibling, M. R., A. R. Bashforth, H. J. Falcon-Lang, J. P. Allen, and C. R. Fielding. 2010. Log Jams and
Flood Sediment Buildup Caused Channel Abandonment and Avulsion in the Pennsylvanian of
Atlantic Canada. Journal of Sedimentary Research 80:268–287.

Gillespie, Major G. L. 1881. Report of the Chief of Engineers, U.S. Army. Appendix OO 10, 2603–2605.

Gillespie, N., A. Unthank, L. Campbell, P. Anderson, R. Gubernick, M. Weinhold, D. Cenderelli,


B. Austin, D. McKinley, S. Wells, J. Rowan, C. Orvis, M. Hudy, A. Bowden, A. Singler, E. Fretz, J.
Levine, and R. Kirn 2014. Flood Effects on Road-Stream Crossing Infrastructure: Economic and
Ecological Benefits of Stream Simulation Designs. Fisheries 39(2):62–76.

Gippel, C. J. 1995. Environmental Hydraulics of Large Woody Debris in Streams and Rivers. Journal of
Environmental Engineering 121:388–395.

Gippel, C. J., I. C. O’Neill, and B. L. Finlayson. 1992. The Hydraulic Basis of Snag Management. Center
for Environmental Applied Hydrology, University of Melbourne, Melbourne, Victoria, Australia.

Large Wood National Manual July 2015


10-22
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Gippel, C. J., I. C. O’Neill, B. L. Finlayson, and I. Schnatz, I. 1994. Hydraulic Guidelines for
Reintroduction and Management of Large Woody Debris in Degraded Lowland Rivers. Pages 225–
239 in Proceedings of the Conference on Habitat Hydraulics. International Association for
Hydraulic Research.

Gippel, C. J., I. C. O’Neill, and B. L. Finlayson. 1996. Distribution and Hydraulic Significance of Large
Woody Debris in a Lowland Australian River. Hydrobiologia 318:179–194.

Goldsmith, W., D. H. Gray, and J. McCullah. 2014. Bioengineering Case Studies. New York Springer.

Goode, J. R., C. H. Luce, and J. M. Buffington. 2012. Enhanced Sediment Delivery in a Changing Climate
in Semi-Arid Mountain Basins: Implications for Water Resource Management and Aquatic
Habitat in the Northern Rocky Mountains. Geomorphology 139-140:1–15.

Gordon, R. P., L. K. Lautz, and T. L. Daniluk. 2013. Spatial Patterns of Hyporheic Exchange and
Biogeochemical Cycling around Cross-Vane Restoration Structures: Implications for Stream
Restoration Design. Water Resources Research 49:20185.

Gosnell, H., and E. Kelly. 2010. Peace on the River? Social-Ecological Restoration and Large Dam
Removal in the Klamath basin, USA. Water Alternatives 3:362–383.

Gotvald, A. J., N. A. Barth, A. G. Veilleux, and C. Parrett. 2012. Methods for Determining Magnitude and
Frequency of Floods in California, Based on Data Through Water Year 2006. U.S. Geological Survey
Scientific Investigations Report 2012–5113. Available: http://pubs.usgs.gov/sir/2012/5113/.

Gowan, C., and K. D. Fausch. 1996. Long-Term Demographic Responses of Trout Populations to
Habitat Manipulation in Six Colorado Streams. Ecological Applications 6(3):931–946.

Graf, W. L. 1975. The impact of Suburbanization on Fluvial Morphology. Water Resources Research
11:690–692.

Graf, W. L. 1999. Dam Nation: A Geographic Census of American Dams and Their Large-Scale
Hydrologic Impacts. Water Resources Research 35:1305–1311.

Graham, R., and K. Cromack. 1982. Mass, Nutrient Content, and Decay Rate of Dead Boles in Rain
Forests of Olympic National Park. Canadian Journal of Forest Research 12(3):511–521.

Grant, G. E., and F. J. Swanson. 1995. Morphology and Processes of Valley Floors in Mountain
Streams, western Cascades, Oregon. Pages 83–101 in J. D. Costa, A. J. Miller, K. W. Potter, and
P. R. Wilcock (eds.). Natural and Anthropogenic Influences in Fluvial Geomorphology. Geophysical
Monograph 89. American Geophysical Union, Washington DC.

Grant, G. E., M. J. Crozier, and F. J. Swanson. 1984. An Approach to Evaluating Off-Site Effects of
Timber Harvest Activities on Channel Morphology. Proceedings of the Symposium on the Effects
of Forest and Land Use on Erosion and Slope Stability. Environment and Policy Institute, E-West
Center, University of Hawaii, Honolulu 177–186.

Grant, G., J. Schmidt, J., and S. Lewis. 2003. A Geological Framework for Interpreting Downstream
Effects of Dams on Rivers. Page 209–226 in A Unique River, Water Science Application, Volume 7.
American Geophysical Union.

Gray, D. H. 1974. Reinforcement and Stabilization of Soil by Vegetation. Journal of Geotechnical


Engineering, 100(GT6):695–699.

Large Wood National Manual July 2015


10-23
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Gray, D. H. and D. Barker. 2004. Root-soil Mechanics and Interactions. Pages 113–123 in S. J. Bennett
and A. Simon (eds.). Riparian Vegetation and Fluvial Geomorphology, Water Science and
Application 8, American Geophysical Union, Washington D.C.

Greer, E. S., S. R. Pezeshki, and F. D. Shields, Jr. 2006. Root Elongation of Black Willow Stakes in
Response to Cutting Size and Soil Moisture Regime (Tennessee). Ecological Restoration
24(3):195–197.

Gregory, K. J., R. J. Davis, and S. Tooth. 1993. Spatial Distribution of Coarse Woody Debris Dams in
the Lymington Basin, Hampshire, UK. Geomorphology 6:207–224.

Gregory, R., L. Failing, M. Harstone, G. Long, T. MacDaniels, and D. Ohlson. 2012. Structured Decision
Making A Practical Guide to Environmental Choices. Wiley-Blackwell.

Gregory, S. V., F. J. Swanson, W. A. McKee, and K. W. Cummins. 1991. An Ecosystem Perspective of


Riparian Zones. BioScience 41(8):540–551.

Gregory, S. V., K. L. Boyer, and A. M. Gurnell (eds.). 2003a. The Ecology and Management of Wood in
World Rivers. Bethesda, MD: American Fisheries Society.

Gregory, S.V., M.A. Meleason, and D.J. Sobota. 2003b. Modeling the Dynamics of Wood in Streams and
Rivers. American Fisheries Society Symposium 37:315–335.

Grette, G. B. 1985. The role of Large Organic Debris in Juvenile Salmonid Rearing Habitat in Small
Streams. MS thesis, University of Washington, Seattle, WA.

Grimm, W. C. 1967. Familiar Trees of America. New York: Harper & Row.

Grizzel, J. D., and N. Wolff. 1998. Occurrence of Windthrow in Forest Buffer Strips and its Effect on
Small Streams in Northwest Washington. Northwest Science 72:214–223.

Grizzel, J., M. McGowan, D. Smith, and T. Beechie. 2000. Streamside Buffers and Large Woody Debris
Recruitment: Evaluating the Effectiveness of Watershed Analysis Prescriptions in the North
Cascades Region. TFW-MAGI-00-003. Washington State Timber, Fish & Wildlife.

Groffman, P. M., and eight others. 2009. Challenges to Incorporating Spatially and Temporally
Explicit Phenomena (Hotspots and Hot Moments) in Denitrification Models. Biogeochemistry
DOI 10.1007/s10533-008-9277-5.

Groot, C., and L. Margulis (eds.). 1991. Pacific Salmon Life Histories. Vancouver, BC: University of
British Columbia Press.

Guardia, J. E. 1933. Some Results of the Log Jams in the Red River. The Bulletin of the Geographical
Society of Philadelphia 31(3):103–114.

Gunderson, L., and S. S. Light. 2006. Adaptive Management and Adaptive Governance in the
Everglades Ecosystem. Policy Sciences 39(4):323–334.

Gurnell, A. M., G. E. Petts, N. Harris, J. V. Ward, K. Tockner, P. J. Edwards, and J. Kollman. 2000. Large
Wood Retention in River Channels: The Case of the Fiume Tagliamento, Italy. Earth Surface
Processes and Landforms 25:255–275.

Gurnell, A. M., H. Piegay, F. J. Swanson, and S. V. Gregory. 2002. Large Wood and Fluvial Processes.
Freshwater Biology 47(4):601–619.

Large Wood National Manual July 2015


10-24
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Gurnell, A. J., W. Bertoldi, and D. Corenblit. 2012. Changing River Channels: The Roles of Hydrological
Processes, Plants and Pioneer Fluvial Landforms in Humid Temperate, Mixed Load, Gravel Bed
Rivers. Earth-Science Reviews 111:129–141.

Guyette, R. P., and M. Stambaugh. 2003. The Age and Density of Ancient and Modern Oak Wood in
Streams and Sediments. The International Association of Wood Anatomists (IAWA) Journal
24:345–353.

Guyette, R. P., D. C. Dey, and M. C. Stambaugh 2008. The Temporal Distribution and Carbon Storage
of Large Oak Wood in Streams and Floodplain Deposits. Ecosystems 11:643–653.

Habron, G. 2003. Role of Adaptive Management for Watershed Councils. Environmental Management
31(1):0029–0041.

Hafs, A. W., L. R. Harrison, R. M. Utz, and T. Dunne. 2014. Quantifying the Role of Woody Debris in
Providing Bioenergetically Favorable Habitats for Juvenile Salmon. Ecological Modelling 285:
30–38.

Haga, H., T. Kumagai, K. Otsuki, and S. Ogawa. 2002. Transport and Retention of Coarse Woody
Debris in Mountain Streams: An In Situ Field Experiment of Log Transport and a Field Survey of
Coarse Woody Debris Distribution. Water Resources Research 38:1126,
doi:10.1029/2001WR001123.

Hairston, N. G. 1987. Community Ecology and Salamander Guilds. Cambridge University Press, 1987.

Hall, L. S., P. R. Krausman, and M. L. Morrison. 1997. The Habitat Concept and a Plea for Standard
Terminology. Wildlife Society Bulletin 25(1):173–182.

Hall, R. O., and J. L. Meyer. 1998. The Trophic Significance of Bacteria in a Detritus-Based Stream
Food Web. Ecology 79:1995–2012.

Hall, R. O., J. B. Wallace, and S. L. Eggert. 2000. Organic Matter Flow in Stream Food Webs with
Reduced Detrital Resource Base. Ecology 81(12):3445–3463.

Hamill, L. 1999. Bridge Hydraulics. New York, NY: Routledge.

Hamlet, A. F., M. M. Elsner, G. S. Mauger, S.-Y. Lee, I. Tohver, and R. A. Norheim. 2013. An Overview of
the Columbia Basin Climate Change Scenarios Project: Approach, Methods, and Summary of Key
Results. Atmosphere-Ocean, 51(4):392–415.

Hammer, T. R. 1972. Stream Channel Enlargement due to Urbanization. Water Resources Research
8:1530–1540.

Harmon, M. E., J. F. Franklin, F. J. Swanson, P. Sollins, S. V. Gregory, J. D. Lattin, N. H. Anderson,


S. P. Cline, N. G. Aumen, J. R. Sedell, G. W. Lienkaemper, K. Cromack, Jr., and K. W. Cummins. 1986.
Ecology of Coarse Woody Debris in Temperate Ecosystems. Advances in Ecological Research
15:133–302.

Harrod, R. 2000. Ecologist with the Wenatchee National Forest Service, Wenatchee, WA. Personal
Communication.

Hart, E. A. 2002. Effects of Woody Debris on Channel Morphology and Sediment Storage in
Headwater Streams in the Great Smoky Mountains, Tennessee-North Carolina. Physical
Geography 23:492–510.

Large Wood National Manual July 2015


10-25
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Hartopo, 1991. The Effect of Raft Removal and Dam Construction on the Lower Colorado River, Texas.
Unpublished M.S. Thesis, Texas A & M University.

Harvey, J. W., and ten others. 2012. Hydrogeomorphology of the Hyporheic Zone: Stream Solute and
Fine Particle Interactions with a Dynamic Streambed. Journal of Geophysical Research
117:G00N11.

Harvey, M. D, D. S. Biedenharn, and P. Combs. 1988. Adjustments of Red River Following Removal of
the Great Raft in 1873 [abs.]. Eos, Transactions of the American Geophysical Union 69(18):567.

Harwood, K., and A. G. Brown. 1993. Fluvial Processes in a Forested Anastomosing River: Flood
Partitioning and Changing Flow Patterns. Earth Surface Processes and Landforms 18:741–748.

Hauer, F. R. 1989. Organic Matter Transport and Retention in a Blackwater Stream Recovering from
Flow Augmentation and Thermal Discharge. Regulated Rivers: Research and Management 4:371–
380.

Hawkins, C. P., J. L. Kershner, P. A. Bisson, M. D. Bryant, L. M. Decker, S. V. Gregory, D. A. McCullough,


C. K. Overton, G. H. Reeves, R. J. Steedman, and M. K. Young. 1993. A Hierarchical Approach to
Classifying Stream Habitat Features. Fisheries 18(6):3–12.

Hax, C. L., and S. W. Golladay. 1993. Macroinvertebrate Colonization and Biofilm Development on
Leaves and Wood in a Boreal River. Freshwater Biology 29(1):79–87.

Hayes, D. B., C. P. Ferreri, and W. W. Taylor. 1996. Linking Fish Habitat to Their Population
Dynamics. Canadian Journal of Fisheries and Aquatic Sciences 53(S1):383–390.

Hazard, J. T. 1948. Our Living Forests, the Story of Their Preservation and Multiple Use. Seattle,
WA: Superior Publishing.

He, Z., W. Wu, and F. D. Shields, Jr. 2009. Numerical Analysis of Effects of Large Wood Structures on
Channel Morphology and Fish Habitat Suitability in a Southern U.S. Sandy Creek. Ecohydrology
2 (3):370–380. doi: 10.1002/eco.60.

Hedman, C. W., D. H. Van Lear, and W. T. Swank. 1996. In-Stream Large Woody Debris Loading and
Riparian Forest Seral Stage Associations in the Southern Appalachian Mountains. Canadian
Journal of Forest Research 26:1218–1227.

Heede, B. H. 1972. Influences of a Forest on the Hydraulic Geometry of Two Mountain Streams.
Water Resources Bulletin 8:523–530.

Heede, B. H. and J. N. Rinne, 1990. Hydrodynamic and Fluvial Morphological Processes: Implications
for Fisheries Management and Research. North American Journal of Fisheries Management,
10:249–268.

Helmus, M. R. and G. G. Sass. 2008. The Rapid Effects of a Whole-Lake Reduction of Coarse Woody
Debris on Fish and Benthic Macroinvertebrates. Freshwater Biology 53:1423–1433.

Helton, A. M., and 22 others. 2011. Thinking outside the Channel: Modeling Nitrogen Cycling in
Networked River Ecosystems. Frontiers in Ecology and the Environment 9:229–238.

Henderson, J. 1996. Unpublished Data Regarding Tree Height vs. Age for Two Common Plant
Association Groups. USDA Forest Service, Pacific Northwest Region, Mount Lake Terrace, WA.

Large Wood National Manual July 2015


10-26
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Henderson, J. A., R. D. Lesher, D. H. Peter and D. C. Shaw. 1992. Field Guide to the Forested Plant
Associations of the Mt. Baker-Snoqualmie National Forest. USDA Forest Service, Pacific Northwest
Region. Tech paper R6 ECOL TP 028-91.

Herlihy, A. T., J. L. Stoddard, and C. Burch Johnson. 1998. The Relationship between Stream
Chemistry and Watershed Land Cover Data in the Mid-Atlantic Region, U.S. Pages 377–386 in
Biogeochemical Investigations at Watershed, Landscape, and Regional Scales. Springer
Netherlands.

Herrera Environmental Consultants, Inc. 2006. Conceptual Design Guidelines: Application of


Engineered Logjams. Prepared for Scottish Environmental Protection Agency, Galashiels, United
Kingdom.

Hershey, K. 1995. Characteristics of Forests at Spotted Owl Nest Sites in the Pacific Northwest.
M.S. thesis, Oregon State University, Corvallis.

Hertzberg, R. 1954. Wave-Wash Control on Mississippi River Levees. Transactions of the ASCE
119(2688):628–638.

Hester, E. T., and M. W. Doyle. 2008. In-Stream Geomorphic Structures as Drivers of Hyporheic
Change. Water Resources Research 44:W03427.

Hester, E. T., and M. N. Gooseff. 2010. Moving Beyond the Banks: Hyporheic Restoration is
Fundamental to Restoring Ecological Services and Functions of Streams. Environmental Science
and Technology 44:1521–1525.

Hester, E. T., M. W. Doyle, and G. C. Poole. 2009. The Influence of in-Stream Structures on Summer
Water Temperatures via Induced Hyporheic Exchange. Limnology and Oceanography 54:355–
367.

Hewitt, E. R. 1934. Hewitt’s Handbook of Stream Improvement. New York: The Marchbanks Press.

Hickin E. J. 1984. Vegetation and River Channel Dynamics. Canadian Geographer 28(2):111–126.

Hilborn, R., and M. Mangel. 1997. The Ecological Detective. Princeton, NJ: Princeton University Press.

Hilborn, R., and C. J. Walters. 1992. Quantitative Fish Stock Assessment. London: Chapman and Hall.

Hilderbrand, R. H., A. D. Lemly, C. A. Dollof, and K. L. Harpster. 1997. Effects of Large Woody Debris
Placement on Stream Channels and Benthic Macroinvertebrates. Canadian Journal of Fisheries
and Aquatic Sciences 54:931–939.

Hilderbrand, R. H., A. D. Lemly, C. A. Dolloff, and K. L. Harpster. 1998. Design Considerations for
Large Woody Debris Placement in Stream Enhancement Projects. North American Journal of
Fisheries Management 18:161–167.

Hinkle, S. R., J. H. Duff, F. J. Triska, A. Laenen, E. B. Gates, K. E. Bencala, D. A. Wentz, and S. R. Silva.
2001. Linking Hyporheic Flow and Nitrogen Cycling near the Willamette River – A Large River in
Oregon, USA. Journal of Hydrology 244:157–180.

Hinkle, S. R., D. S. Morgan, L. L. Orzol, and D. J. Polette. 2007. Ground Water Redox Zonation near
La Pine, Oregon: Relation to River Position within the Aquifer-Riparian Zone Continuum.
U.S. Geological Survey Scientific Investigations Report 2007–5239.

Large Wood National Manual July 2015


10-27
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Hinkle, S. R., K. E. Bencala, D. A. Wentz, and D. P. Krabbenhoft. 2013. Mercury and Methylmercury
Dynamics in the Hyporheic Zone of an Oregon Stream. Water Air and Soil Pollution 225, article
1694.

Hoellein, T. J., J. L. Tank, E. J. Rosi-Marshall, S. A. Entrekin, and G. A. Lamberti. 2007. Controls on


Spatial and Temporal Variation of Nutrient Uptake in Three Michigan Headwater Streams.
Limnol. Oceanogr.52:1964–1977.

Hogan, D. L. 1987. The influence of large organic debris on channel recovery in the Queen Charlotte
Islands, British Columbia, Canada. Pages 343–353 in R. L. Beschta, T. Blinn, G. E. Grant,
F. J. Swanson, and G. G. Ice (eds.), Erosion and Sedimentation in the Pacific Rim. IAHS Publication
No.165.

Holling, C. S. 1978. Adaptive Environmental Assessment and Management. Chichester, UK: Wiley
Interscience.

Hollis, G. E. 1975. The Effects of Urbanization on Floods of Different Recurrence Intervals. Water
Resources Research 11:431–435.

Holstine, C. 1992. An Historical Overview of the Wenatchee National Forest, Washington. Rep. 100-80.
Archaeological and historical Services. Eastern Washington University, Cheney.

Homer, C. C., L. Huang, B. W. Yang, and M. Coan. 2004. Development of a 2001 National Landcover
Database for the United States. Photogrammetric Engineering and Remote Sensing 70(7):829–
840.

Hough-Snee, N., A. Kasprak, B. B. Roper, and C. S. Meredith. 2014. Direct and Indirect Drivers of
Instream Wood in the Interior Pacific Northwest, USA: Decoupling Climate, Vegetation,
Disturbance, and Geomorphic Setting. Riparian Ecology and Conservation 2:14–34.

House, R. A., and P. L. Boehne. 1985. Evaluation of Instream Enhancement Structures for Salmonid
Spawning and Rearing in a Coastal Oregon Stream. North American Journal of Fish Management
5:283–295.

House, R. A., and P. L. Boehne. 1986. Effects of Instream Structures on Salmonid Habitat and
Populations in Tobe Creek, Oregon. North American Journal of Fisheries Management 6:283–295.

Howey, C. A. F., and S. A. Dinkelacker. 2009. Habitat Selection of the Alligator Snapping Turtle
(Macrochelys temminckii) in Arkansas. Journal of Herpetology 43(4):589–596.

Hudson, P. L., R. W. Griffiths, and T. J. Wheaton. 1992. Review of Habitat Classification Schemes
Appropriate to Streams, Rivers, and Connecting Channels in the Great Lakes Drainage Basin.
Pages 73–107 in W. D. N. Busch and P. G. Sly (eds.), The Development of an Aquatic Habitat
Classification System for Lakes. Ann Arbor, MI: CRC Press.

Hughes, R. M. (ed.). 1993. Stream Indicators and Design Workshop. EPA/600/R-93/138.


U.S. Environmental Protection Agency, Corvallis, OR.

Hughes, T. J. R., G. R. Feijoo, L. Mazzei, and J.-B. Quincy. 1998. The Variational Multiscale Method—
A Paradigm for Computational Mechanics. Computer Methods in Applied Mechanics and
Engineering 166 (1-2):3–24.

Large Wood National Manual July 2015


10-28
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Hurlbert, S. H. 1984. Pseudoreplication and the Design of Ecological Field Experiments. Ecological
Monographs 54 (2:187–211.

Huryn, A. D., and J. B. Wallace. 1987. Local Geomorphology as a Determinant of Macrofaunal


Production in a Mountain Stream. Ecology 68:1932–1942.

Hutchens, J. J., E. F. Benfield, and J. R. Webster. 1997. Diet and Growth of a Leaf-Shredding Caddisfly
in Southern Appalachian Streams of Contrasting Disturbance History. Hydrobiologia 346:193–
201.

Hutchinson, G. E. 1957. Concluding Remarks. Cold Springs Harbor Symposium on Quantitative Biology
22:415–427.

Hyatt, T. L., and R. J. Naiman. 2001. The Residence Time of Large Woody Debris in the Queets River,
Washington, USA. Ecological Applications 11(1):191–202.

Hygelund, B., and M. Manga. 2003. Field Measurements of Drag Coefficients for Model Large Woody
Debris. Geomorphology 51:175–185.

Ibrahim, T. D. Lerner, and S. Thornton. 2011. Accounting for the Groundwater-Surface Water
Interface in Contaminated Land Assessments. CL:AIRE Technical Bulletin 15.

ICF International. 2010. Instream Woody Material Installation and Monitoring Guidance Manual.
Sacramento Area Flood Control Agency, Sacramento, California.

Idaho Office of the Administrative Rules Coordinator. Undated. Idaho Minimum Standards for
Logging—Sections 17.08.01 through 17.08.16. Available:
http://adminrules.idaho.gov/rules/current/17/index.html. Accessed: July 10, 2014. Industrial
Commission. Idaho Office of the Administrative Rules Coordinator, Boise, Idaho.

Ikeya, H. 1981. A Method for Designation Forested Areas in Danger of Debris Flows. In Erosion and
Sediment Transport in Pacific Rim Steeplands. Edited by T. R. H. Davies and A. J. Pearce.
International Association of Hydrological Sciences, Publication 132:576–588.

Interagency Advisory Committee on Water Data (IACWD). 1982. Guidelines for Determining Flood
Flow Frequency. Bulletin 17B of the Hydrology Subcommittee, Office of Water Data Coordination,
U.S. Geological Survey, Reston, Virginia. 183 p.

Intergovernmental Panel on Climate Change (IPCC). 2007. Climate Change 2007: Impacts, Adaptation
and Vulnerability. Contribution of Working Group II to the Fourth Assessment Report of the
Intergovernmental Panel on Climate Change. M. L. Parry, O. F. Canziani, J. P. Palutikof, P. J. van der
Linden, and C. E. Hanson (eds.). Cambridge, UK, and New York, NY: Cambridge University Press.

International Union for Conservation of Nature (IUCN). 2014. IUCN Releases an Economic Framework
for Analyzing Forest Landscape Restoration Decisions. Available:
http://cmsdata.iucn.org/downloads/flr_economic_analysis_tutorial___july_2014_1.pdf.

Jackson, T. R., R. Haggerty, and S. V. Apte. 2013. A Fluid-Mechanic Classification Scheme for Surface
Transient Storage in Riverine Sediments: Quantitatively Separating Surface from Hyporheic
Storage. Hydrology and Earth System Sciences 17:2747–2779.

Large Wood National Manual July 2015


10-29
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Jacobson, P. J., K. M. Jacobson, P. L. Angermeier, and D. S. Cherry. 1999. Transport, Retention, and
Ecological Significance of Woody Debris within a Large Ephemeral River. Journal of the North
American Benthological Society 18:429–444.

James, L. D. 1965. Using a Digital Computer to Estimate the Effects of Urban Development on flood
Peaks. Water Resources Research 1:223–234.

Jeffries, R., S. E. Darby, and D. A. Sear. 2003. The Influence of Vegetation and Organic Debris on
Flood-Plain Sediment Dynamics: Case Study of a Low-Order Stream in the New Forest, England.
Geomorphology 51:61–80.

Jenkins, M. J., and E. G. Hebertson. 1998. Using Vegetative Analysis to Determine the Extent and
Frequency of Avalanches in Little Cottonwood Canyon, Utah. Department of Forest Resources,
Utah State University. WestWide Avalanche Network, UT.

Jia, Y., S. Scott, Y. Xu, and S. S. Y. Wang. 2009. Numerical Study of Flow Affected by Bendway Weirs in
Victoria Bendway, the Mississippi River. Journal of Hydraulic Engineering 135(11):902–916.

Johnson, A. W., and J. M. Stypula (eds.). 1993. Guidelines for Bank Stabilization Projects in the Riverine
Environments of King County. King County Department of Public Works, Surface Water
Management Division. Seattle, WA.

Johnson, L. B., D. H. Breneman, and C. Richards. 2003. Macroinvertebrate Community Structure and
Function Associated with Large Wood in Low Gradient Streams. River Research and Applications
19:199–218.

Johnson, P. A., R. D. Hey, M. Tessier, and D. L. Rosgen. 2001. Use of Vanes for Control of Scour at
Vertical Wall Abutments. Journal of Hydraulic Engineering 127(9):772–778.

Johnson, S. L., F. J. Swanson, G. E. Grant, and S. M. Wondzell. 2000. Riparian Forest Disturbances by a
Mountain Flood—The Influence of Floated Wood. Hydrological Processes 14:3031–3050.

Johnston, N. T., E. A. MacIsaac, P. J. Tschaplinski, and K. J. Hall. 2004. Effects of the Abundance of
Spawning Sockeye Salmon (Oncorhynchus nerka) on Nutrients and Algal Biomass in Forested
Streams. Canadian Journal of Fisheries and Aquatic Sciences 61:384–403.

Johnston, N. T., S. A. Bird, D. L. Hogan, and E. A. MacIsaac. 2011. Mechanisms and Source Distances
for the Input of Large Woody Debris to Forested Streams in British Columbia, Canada Canadian
Journal of Forest Research. 41(11):2231–2246.

Jones, C. and P. Johnson. 2015. Risk Assessment for Stream Modification Projects in Urban Settings.
ASCE-ASME Journal of Risk and Uncertainty in Engineering Systems, Part A: Civil Engineering.
10.1061/AJRUA6.0000815, 04015001.

Jonsson, B., and N. Jonsson. 2003. Migratory Atlantic Salmon as Vectors for the Transfer of Energy
and Nutrients Between Freshwater and Marine Environments. Freshwater Biology 48:21–27.

Jowett, I. G., J. W. Hayes, and M. J. Duncan. 2008. A Guide to Instream Habitat Survey Methods and
Analysis. NIWA Science and Technology Series No. 54. Available:
http://www.niwa.co.nz/sites/niwa.co.nz/files/a_guide_to_instream_habitat_survey_methods_an
d_analysis.pdf.

Large Wood National Manual July 2015


10-30
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Junk, W. J., P. B. Bayley, and R. E. Sparks. 1989. The Flood Pulse Concept in River Floodplain Systems.
Pages 110-127 in D. P. Dodge (ed.), Proceedings of the International Large River Symposium.

Kail, J., D. Hering, S. Muhar, J. Gerhard, and S. Preis. 2007. The Use of Large Wood in Stream
Restoration: Experiences from 50 Projects in Germany and Austria. Journal of Applied Ecology
44:1145–1155.

Kanes, W. H. 1970. Facies and Development of the Colorado River Delta in Texas. Pages 78–106 in
J. P. Morgan and R. H. Shaver (eds.), Deltaic Sedimentation Modern and Ancient. Special
Publication No.15. Society of Economic Paleontologists and Mineralogists. Tulsa, Oklahoma.

Kaplan, R. S. 2008. Conceptual Foundations of the Balanced Scorecard. Handbook of Management


Accounting Research 3:1253–1269.

Karr, J. R. 1981. Assessment of Biotic Integrity Using Fish Communities. Fisheries 6(6):21–27.

Karr, J. R., and B. L. Kerans. 1991. Components of Biological Integrity: Their Definition and Use in
Development of an Invertebrate IBI. Midwest Pollution Control Biologists Meeting, Chicago, IL,
1991, Proceedings: US Environmental Protection Agency, Region V, EPA-905/R-92-003.

Karr, J. R., K. D. Fausch, P. L. Angermeier, P. R. Yant, and I. J. Schlosser. 1986. Assessing Biological
Integrity in Running Waters: A Method and its Rationale. Illinois Natural History Survey Special
Publication 5, Urbana, IL.

Kasprak, A., F. J. Magilligan, K. H. Nislow, and N. P. Snyder, N.P. 2012. A Lidar-Derived Evaluation of
Watershed-Scale Large Woody Debris Sources and Recruitment Mechanisms: Coastal Maine,
USA. River Research Applications 28:1462–1476.

Kati, V., P. Devillers, M. Dufrene, A. Legakis, D. Vokou, and P. Lebrun. 2004. Testing the Value of Six
Taxonomic Groups as Biodiversity Indicators at a Local Scale. Conservation Biology 18(3):667–
675.

Kauffman, J. B., R. L. Beschta, N. Otting, and D. Lytjen. 1997. An Ecological Perspective of Riparian
and Stream Restoration in the Western United States. Fisheries (Bethesda) 22:12–24.

Kaufmann, P. R., P. Levine, E. G. Robison, C. Seeliger, and D. V. Peck. 1999. Quantifying Physical
Habitat in Streams. U.S. Environmental Protection Agency, Washington, D.C. Available:
http://www.epa.gov/emap2/html/pubs/docs/groupdocs/surfwatr/field/phyhab.pdf.

Kaushal, S. S., G. E. Likens, N. A. Jaworski, M. L. Pace, A. M. Sides, D. Seekell, K. T. Belt, D. H. Secor, and
R. L. Wingate. 2010: Rising Stream and River Temperatures in the United States. Frontiers in
Ecology and the Environment 8:461–466. doi:10.1890/090037.

Keeney, R. L., and H. Raiffa. 1993. Decisions with Multiple Objectives Preferences and Value Tradeoffs.
Cambridge University Press.

Keeton, W. S., C. E. Kraft, and D. R. Warren. 2007. Mature and Old-Growth Riparian Forests:
Structure, Dynamics and Effects on Adirondack Stream Habitats. Ecological Applications 17:852–
868.

Keith, D. A., T. G. Martin, E. McDonald-Madden, and C. Walters. 2011. Uncertainty and Adaptive
Management for Biodiversity Conservation. Biological Conservation 144(4):1175–1178.

Large Wood National Manual July 2015


10-31
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Keller, E. A. and A. MacDonald. 1995. River Channel Change: The Role of Large Woody Debris. Pages
217–236 in A. Gurnell and G. Petts (eds.), Changing River Channels. John Wiley and Sons,
Chichester. 217-235.

Keller, E. A., and F. J. Swanson. 1979. Effects of Large Organic Material on Channel Form and Fluvial
Processes. Earth Surface Processes 4:361–380.

Keller, E. A., and T. Tally. 1979. Effects of Large Organic Debris on Channel Form and Fluvial
Processes in the Coastal Redwood Environment. Pages 169–197 in D. D. Rhodes and G. P.
Williams (eds.), Adjustments of the Fluvial System. Proceedings of the 10th Annual Binghamton
Geomorphology Symposium. Kendal-Hunt. Dubuque, IA.

Kellndorfer, J., W. Walker, E. LaPoint, J. Bishop, T. Cormier, G. Fiske, M. Hoppus, K. Kirsch, and
J. Westfall. 2012. NACP Aboveground Biomass and Carbon Baseline Data (NBCD 2000), U.S.A.
2000. Data set. Available: http://daac.ornl.gov from ORNL DAAC, Oak Ridge, Tennessee, U.S.A.
http://dx.doi.org/10.3334/ORNLDAAC/1081. See more at: http://www.whrc.org/mapping/
nbcd/#sthash.7OB99UEf.dpuf.

Kennard, P., G. Pess, T. Beechie, B. Bilby, and D. Berg. 1998. Riparian-in-a-Box: A Manager’s Tool to
Predict the Impacts of Riparian Management on Fish Habitat. Pages 483-490. in M. K. Brewin
and D. M. A. Monita (eds.), Forest-Fish Conference: Land Management Practices Affecting Aquatic
Ecosystems. Proceedings of Forest-fish conference, May 1-4, 1996, Calgary, Alberta. Natural
Resources Canada. North For. Cent., Edmonton, Alberta Inf. Rep. NOR-X-356.

Kennedy, B. P., K. H. Nislow, and C. L. Folt. 2008. Habitat-Mediated Foraging Limitations Drive
Survival Bottlenecks for Juvenile Salmon. Ecology 89(9):2529–2541.

Keown, M. P., N. R. Oswalt, E. B. Perry, and E. A. Dardeau Jr. 1977. Literature Survey and Preliminary
Evaluation of Streambank Protection Methods. Technical Report No. WES-TR-H-77-9, U.S. Army
Engineer Waterways Experiment Station, Vicksburg, MS.

Kerans, B. L., and J. R. Karr. 1994. A Benthic Index of Biotic Integrity (B-IBI) for Rivers of the
Tennessee Valley. Ecological Applications. 4(4):768–785.

Kingsford, R. T., K. F. Walker, R. E. Lester, W. J. Young, P. G. Fairweather, J. Sammut, and M. C. Geddes.


2011. A Ramsar Wetland in Crisis—The Coorong, Lower Lakes and Murray Mouth, Australia.
Marine and Freshwater Research 62(3):255–265.

Kirkwood, C. W. 1997. Strategic Decision Making. Multi-Objective Decision Analysis with Spreadsheets.
Wadsworth.

Klingeman, P. C., S. M. Kehe, and Y. A. Owusu. 1984. Streambank Erosion Protection and Channel
Scour Manipulation Using Rockfill Dikes and Gabions. Water Resources Research Institute,
Oregon State University. Salem, OR.

Kloehn, K., T. Beechie, S. Morley, H. Coe, and J. Duda, J. 2008. Influence of Dams on River-Floodplain
Dynamics in the Elwha River, Washington. Northwest Science 82:224–235.

Knowles, N., M. D. Dettinger, and D. R. Cayan. 2006. Trends in Snowfall versus Rainfall in the
Western United States. Journal of Climate 19:4545–4559. doi:10.1175/JCLI3850.1. Available:
http://journals.ametsoc.org/doi/pdf/10.1175/JCLI3850.1.

Large Wood National Manual July 2015


10-32
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Knox, J. C. 1993. Large Increases in Flood Magnitude in Response to Modest Changes in Climate.
Nature, 361(6411):430–432.

Knox, J. C. 2000. Sensitivity of Modern and Holocene Floods to Climate Change. Quaternary Science
Reviews, 19(1):439–457.

Knutson, M., and P. Fealko. 2014. Pacific Northwest Region Resource and Technical Services—Large
Woody Material Risk Based Design Guidelines. U.S. Department of the Interior, Bureau of
Reclamation, Pacific Northwest Region, Boise, Idaho.

Knutti, R., G. Abramowitz, M. Collins, V. Eyring, P. J. Gleckler, B. Hewitson, and L. Mearns. 2010. Good
Practice Guidance Paper on Assessing and Combining Multi Model Climate Projections. In T. F.
Stocker, D. Qin, G.-K. Plattner, M. Tignor, and P.M. Midgley (eds.), Meeting Report of the
Intergovernmental Panel on Climate Change Expert Meeting on Assessing and Combining Multi
Model Climate Projections. IPCC Working Group I Technical Support Unit, University of Bern,
Bern, Switzerland.

Koehn, J. D., and S. J. Nicol. 2014. Comparative Habitat Use by Large Riverine Fishes. Marine and
Freshwater Research 65(2):164–174.

Koehn, J. D., W. G. O’Connor, P. D. Jackson. 1994. Seasonal and size-Related Variation in Microhabitat
Use of a Small Victorian Stream Fish Assemblage. Australian Journal of Marine and Freshwater
Research 45:1353–1366.

Kondolf, G. M. 1995. Five Elements for Effective Stream Restoration. Restoration Ecology 3:133–136.

Kondolf, G. M. 1996. A Cross Section of Stream Channel Restoration. Journal of Soil and Water
Conservation 51(2):119–125.

Kondolf, G. M. 1998. Lessons Learned from River Restoration Projects in California. Aquatic
Conservation: Marine and Freshwater Ecosystems 8(1):39–52.

Kondolf, G. M. 2000. Some Suggested Guidelines for Geomorphic Aspects of Anadromous Salmonid
Habitat Restoration Proposals. Restoration Ecology 8:48–55.

Konsoer, K. M. 2014. Influence of Riparian Vegetation on Near-Bank Flow Structure and Rates of
Erosion on a Large Meandering River. Ph.D. Dissertation, University of Illinois, Urbana-
Champaign. 236 p.

Konsoer, K. M., J. A. Zinger, and G. Parker, G., 2013. Bankfull Hydraulic Geometry of Submarine
Channels Created by Turbidity Currents: Relations Between Bankfull Channel Characteristics
and Formative Flow Discharge. Journal of Geophysical Research – Earth Surface 118:1–13. doi:
10.1029/2012JF00242.

Kraft, C. E., and D. R. Warren. 2003. Development of Spatial Pattern in Large Woody Debris and
Debris Dams in Streams. Geomorphology 51:127–139.

Kramer, N., and E. Wohl. 2014. Estimating Fluvial Wood Discharge using Time-Lapse Photography
with Varying Sampling Intervals. Earth Surface Processes and Landforms 39:844–852.

Kratzer, J. F., and D. R. Warren. 2013. Factors Limiting Brook Trout Biomass in Northeastern
Vermont Streams. North American Journal of Fisheries Management 33(1):130–139.

Large Wood National Manual July 2015


10-33
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Krause, C., and C. Roghair. 2014. Inventory of Large Wood in the Upper Chattooga River Watershed,
2007–2013. U.S. Forest Service Southern Research Station, Center for Aquatic Technology
Transfer. Blacksburg, VA.

Krause, S., M. J. Klaar, D. M. Hannah, J. Mant, J. Bridgeman, M. Trimmer, and S. Manning-Jones. 2014.
The Potential of Large Woody Debris to Alter Biogeochemical Processes and Ecosystem Services
in Lowland Rivers. Wiley Interdisciplinary Reviews (WIREs): Water 1:263–275.

Kreutzweiser, D.P., K. P. Good, and T. M. Sutton. 2005. Large Woody Debris Characteristics and
Contributions to Pool Formation in Forest Streams of the Boreal Shield. Canadian Journal of
Forest Research 35:1213–1223.

Kruys, N., B. G. Jonsson, and G. Stahl. 2002. A Stage-Based Matrix Model for Decay-Class Dynamics of
Woody Debris. Ecological Applications 12(3):773–781.

Kukulak, J., A. Pazdur, and T. Kuc. 2002. Radiocarbon Dated Wood Debris in Floodplain Deposits of
the San River in the Bieszczady Mountains. Geochronometria 21:129–136.

Kuhnle, R. A., C. V. Alonso, and F. D. Shields Jr. 1999.Volume of Scour Holes Associated with
90-degree Spur Dikes. Journal of Hydraulic Engineering 125(9):972–978.

Kuhnle, R. A., C. V. Alonso, and F. D. Shields Jr. 2002. Local Scour Associated with Angled Spur Dikes.
Journal of Hydraulic Engineering 128(12):1087–1093.

Lagasse, P. F., P. E. Clopper, J. E. Ortiz-Page, L. W. Zevenbergen, L. A. Ameson, J. D. Schall, and


L. G. Girard. 2009. Bridge Scour and Stream Instability Countermeasures Experience, Selection and
Design Guidance Volumes 1 and 2. HEC-23, Third Edition. Federal Highway Administration.
Washington, D.C.

Lagasse, P. F., P. Clopper, L. Zevenbergen, W. Spitz, and L. G. Girard.2010. Effects of Debris on Bridge
Pier Scour. Federal Highway Administration. Washington, D.C.

Lagasse, P. F., L. W. Zevenbergen, W. J. Spitz, and L. A. Arneson 2012. Stream Stability at Highway
Structures, Fourth Edition. Hydraulic Engineering Circular No. 20. Publication No. FHWA-HR-12-
004. Office of Technology, Federal Highway Administration, Washington, DC.

Lakowicz, J. R., and G. Weber. 1973. Quenching of Fluorescence by Oxygen. Probe for Structural
Fluctuations in Macromolecules. Biochemistry 12(21): 4161–4170.

Lampert, W. 1978. Release of Dissolved Organic-Carbon by Grazing Zooplankton. Limnology and


Oceanography 23(4):831–834.

Lancaster, S. T., S. K. Hayes, and G. E. Grant. 2001. Modeling Sediment and Wood Storage and
Dynamics in Small Mountainous Watersheds. Geomorphic Processes and Riverine Habitat, Water
Science and Application Volume 4:85–102. American Geophysical Union.

Landres, P. B., J. Verner, and J. W. Thomas. 1988. Ecological Uses of Vertebrate Indicator Species:
A Critique. Conservation Biology 2(4):316–328.

Langbien, W. B., and S. B. Schumm. 1958. Yield of Sediment in Relation to Mean Annual Precipitation.
American Geophysical Union Transactions 39:1076–1084.

Lange-Bertalot, H. 1979. Pollution Tolerance of Diatoms as a Criterion for Water Quality Estimation.
Nova Hedwigia, Beih. 64:285–304.

Large Wood National Manual July 2015


10-34
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Lansdown, K., and seven others. 2012. Characterization of Key Pathways of Dissimitory Nitrate
Reduction and Their Response to Complex Organic Substrates in Hyporheic Sediments.
Limnology and Oceanography 57:387–400.

Lassettre, N. S. and R. R. Harris 2001. The Geomorphic and Ecological Influence of Large Woody Debris
in Streams and Rivers. University of California, Berkeley, CA.

Lassettre, N.S., and G.M. Kondolf. 2003. Process Based Management of Large Woody Debris at the
Basin Scale, Soquel Creek, California. Report Presented to California Department of Forestry and
Fire Protection and Soquel Demonstration State Forest. Available:
http://www.fire.ca.gov/resource_mgt/downloads/reports/LWDinSoquelCreek.pdf.

Lassettre, N. S., and G. M. Kondolf. 2012. Large Woody Debris in Urban Stream Channels: Redefining
the Problem. River Research and Applications 28(9):1477–1487.

Lassettre, N., and H. Piégay. 2008. Decadal Changes in Distribution and Frequency of Wood in a Free
Meandering River, the Ain River, France. Earth Surface Processes and Landforms 1112:1098–
1112.

Latterell, J. J., and R. J. Naiman. 2007. Sources and Dynamics of Large Logs In a Temperate Floodplain
River. Ecological Applications 17:1127–1141.

Latterell, J. J., J. S. Bechtold, T. C. O'Keefe, R. Van Pelt, and R. J. Naiman. 2006. Dynamic Patch Mosaics
and Channel Movement in an Unconfined River Valley of the Olympic Mountains Freshwater
Biology 51(3):523–544.

Lautz, L. K., and R. M. Fanelli. 2008. Biogeochemical Hotspots in the Streambed around Restoration
Structures. Biogeochemistry 91:85–104.

Lautz, L. K., D. I. Siegel, and R. L. Bauer. 2006. Impact of Debris Dams on Hyporheic Interaction along
a Semi-Arid Stream. Hydrological Processes 20:183–196.

Lautz, L. K., R. M. Fanelli, N. Kranes, and D. I. Siegel. 2007. Abstract. Sediment distribution around
Debris Dams: Impacts on Streambed Hydrology, Biogeochemistry and Temperature Dynamics in
Small Streams. Geological Society of America Annual Meeting (28–31 October 2007), Paper No.
180-4.

Lautz, L. K., N. T. Kranes, and D. I. Siegel. 2010. Heat Tracing of Heterogeneous Exchange Adjacent to
In-Stream Geomorphic Features. Hydrological Processes 24:3074–3086.

Ledger, M. E. and M. J. Winterbourn. 2000. Growth of New Zealand Stream Insect Larvae in Relation
to Food Type. Archiv fur Hydrobiologie 149:353–364.

Lee, K. N. 1999. Appraising Adaptive Management. Conservation Ecology 3(2):3.

Lee, P. C., C. Smyth, and S. Boutin. 2004. Quantitative Review of Riparian Buffer Width Guidelines
from Canada and the United States. Journal of Environmental Management 70:165–189.

Lehane, B. M., P. S. Giller, J. O’Halloran, C. Smith, J. Murphy. 2002. Experimental Provision of Large
Woody Debris in Streams as a Trout Management Technique. Aquatic Conservation-Marine and
Freshwater Ecosystems 12:289–311.

Lemly, A. D., and R. H. Hilderbrand. 2000. Influence of Large Woody Debris on Stream Insect
Communities and Benthic Detritus. Hydrobiologia 421:179–185.

Large Wood National Manual July 2015


10-35
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Leopold, L. B. 1973. River Channel Change with Time: An Example. Geological Society of America
Bulletin 84:1845–1860.

Lester, R. E. and A. J. Boulton. 2008. Rehabilitating Agricultural Streams in Australia with Wood:
A Review. Environmental Management 42(2):310–326.

Lester, R. E., and W. Wright. 2009. Reintroducing Wood to Streams in Agricultural Landscapes:
Changes in Velocity Profile, Stage and Erosion Rates. River Research and Applications 25(4):276–
392.

Levi, P. S., J. L. Tank, S. D. Tiegs, J. Rüegg, D. T. Chaloner, and G. A. Lamberti. 2011. Does Timber
Harvest Influence the Dynamics of Marine-Derived Nutrients in Southeast Alaska Streams?
Canadian Journal of Fisheries and Aquatic Sciences 68(8):1316–1329.

Li, R., and H. W. Shen. 1973. Effect of Tall Vegetation on Flow and Sediment. Journal of
the Hydraulic Division, ASCE 99(5):793–814.

Li, S., L. T. Martin, S. R. Pezeshki, and F. D. Shields Jr. 2005. Responses of Black Willow (Salix nigra)
Cuttings to Herbivory and Flooding. Acta Oecologica 28(2):173–180.

Lichatowich, J. A., L. E. Mobrand, L. Lestelle, and T. Vogel. 1995. An Approach to the Diagnosis and
Treatment of Depleted Pacific Salmon Populations in Freshwater Ecosystems. Fisheries
20(1):10–18.

Lienkaemper, G. W., and F. J. Swanson. 1987. Dynamics of Large Woody Debris in Streams in
Old-Growth Douglas-Fir Forests. Canadian Journal of Forest Research 17:150–156.

Linkov, I., and E. Moberg. 2011. Multi-Criteria Decision Analysis: Environmental Applications and Case
Studies. Boca Raton, FL: CRC Press.

Linkov. I., and E. Moberg. 2012. Multi-Criteria Decision Analysis: Environmental Applications and Case
Studies. CRC Press. ISBN: 978-1-4398-5318-4.

Linkov, I., A. Varghese, S. Jamil, T. P. Seager, G. Kiker, and T. Bridges. 2005. Multi-Criteria Decision
Analysis: A Framework for Structuring Remedial Decisions at Contaminated Sites. In I. Linkov
and A. Bakr Ramadan (eds.), Comparative Risk Assessment and Environmental Decision Making.
NATO Science Series Volume 38 2005ISBN: 978-1-4020-1895-4 (Print) 978-1-4020-2243-2
(Online).

Lisle, T. 1995. Effects of Coarse Woody Debris and its Removal on a Channel Affected by the 1980
Eruption of Mount St. Helens, Washington. Water Resources Research 31:1797–1808.

Lisle, T. E., Y. Cui, G. Parker, J. E. Pizzuto, and A. M. Dodd. 2001. The Dominance of Dispersion in the
Evolution of Bed Material Waves in Gravel-Bed Rivers. Earth Surface Processes and Landforms
26:1409–1420.

Lister, D. B., and H. S. Genoe. 1970. Stream Habitat Utilization by Cohabiting Underyearlings of
Chinook (Oncorhynchus tshawytscha) and Coho (O. kisutch) Salmon in the Big Qualicum River,
British Columbia. Journal of the Fisheries Research Board of Canada 27:1215–1224.

Loarie, S. R., P. B. Duffy, H. Hamilton, G. P. Asner, C. B. Field, and D. D. Ackerly. 2009. The Velocity of
Climate Change. Nature 462:1052–1055. doi:10.1038/nature08649.

Large Wood National Manual July 2015


10-36
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Lockaby, B. G., J. A. Stanturf, and M. G. Messina. 1997. Effects of Silvicultural Activity on Ecological
Processes in the Floodplain Forests of the Southern United States: A Review of Existing Reports.
Forest Ecology and Management 90:93–100.

Logan, J. A., W. W. Macfarlane, and L. Willcox. 2010. White-Bark Pine Vulnerability to Climate Change
Induced Mountain Pine Beetle Disturbance in the Greater Yellowstone Ecosystem. Ecological
Application 20:895–902. doi:10.1890/09- 0655.1. Available: http://www.esajournals.org/
doi/pdf/10.1890/09-0655.1.

Long, S. L., and C. R. Jackson. 2014. Variation of Stream Temperatures Among Mesoscale Habitats
within Stream Reaches: Southern Appalachians. Hydrological Processes 28:3041–3052.

Lowe, R. L. 1974. Environmental Requirements and Pollution Tolerance of Freshwater Diatoms.


National Environmental Research Center, Office of Research and Development,
U.S. Environmental Protection Agency.

Lunetta, R. S., B. L. Cosentino, D. R. Montgomery, E. M. Beamer, and T. J. Beechie. 1997. GIS-Based


Evaluation of Salmon Habitat in the Pacific Northwest. Photogrammetric Engineering & Remote
Sensing 63(10):1219–1229.

Lyell, C. 1830. Principles of Geology, Volume I. London, UK: John Murray. Published in 1990 by
University of Chicago Press. Chicago, IL.

Lyons, J. E., M. C. Runge, H. P. Laskowski, and W. L. Kendall. 2008. Monitoring in the Context of
Structured Decision-Making and Adaptive Management. The Journal of Wildlife Management
72(8):1683–1692.

MacFarlane, W. A., and E. Wohl. 2003. Influence of Step Composition on Step Geometry and Flow
Resistance in Step-Pool Streams of the Washington Cascades. Water Resources Research
39:1037. doi:10.1029/2001WR001238.

Macka, Z., K. Lukas, B. Louckova, and P. Lucie. 2011. A Critical Review of Field Techniques Employed
in the Survey of Large Woody Debris in River Corridors: A Central European Perspective.
Environmental Monitoring and Assessment 181(1–4):291–316.

Mackenzie, B. D. E, and D. A. Keith. 2009. Adaptive Management in Practice: Conservation of


a Threatened Plant Population. Ecological Management & Restoration 10(s1):S129–S135.

MacNally, R., and E, Fleishman. 2004. A Successful Predictive Model of Species Richness Based on
Indicator Species. Conservation Biology 18(3):646–654.

MacVicar, B., and H. Piégay. 2012. Implementation and Validation of Video Monitoring for Wood
Budgeting in a Wandering Piedmont River, the Ain River (France). Earth Surface Processes and
Landforms 37:1272–1289.

Magilligan, F. J., and K. H. Nislow. 2005. Changes in Hydrologic Regime by Dams. Geomorphology
71:61–78.

Magilligan, F. J., K. H. Nislov, G. B. Fisher, J. Wright, G. Mackey, and M. Laser 2008. The Geomorphic
Function and Characteristics of Large Woody Debris in Low Gradient Rivers, Coastal Maine, USA.
Geomorphology 97:467–482.

Large Wood National Manual July 2015


10-37
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Major, J., J. O’Connor, C. Podolak, M. K. Keith, G. E. Grant, K. Spicer, S. Pittman, H. M. Bragg,


J. R. Wallick, D. Q. Tanner, A. Rhode, and P. Wilcock. 2012. Geomorphic Response of the Sandy
River, Oregon, to Removal of Marmot Dam. U.S. Geological Survey Professional Paper 1792.

Makaske, B., D. G. Smith, and H. J. Berendsen. 2002. Avulsions, Channel Evolution and Floodplain
Sedimentation Rates of the Anastomosing Upper Columbia River, British Columbia, Canada.
Sedimentology 49(5):1049–1071.

Manga, M., and J. W. Kirchner. 2000. Stress Partitioning in Streams by Large Woody Debris. Water
Resources Research 36:2373–2379.

Manly, B. F. 2002. Resource Selection by Animals. Springer, New York.

Manners, R. B. and Doyle, M. W. 2008. A Mechanistic Model of Woody Debris Jam Evolution and its
Application to Wood-based Restoration and Management. River Research and Applications
24:1104-1123.

Manners, R. W., M. W. Doyle, and M. J. Small. 2007. Structure and Hydraulics of Natural Woody
Debris Jams. Water Resources Research 43, doi:10.1029/2006WR004910.

Manners, R. B., J. C. Schmidt, and M. L. Scott, M.L. 2014. Mechanisms of Vegetation-Induced Channel
Narrowing on an Unregulated Canyon Bound River: Results from a Natural Field-Scale
Experiment. Geomorphology 211:100–115.

Mao, L., and F. Comiti. 2010. The Effects of Large Wood Elements During an Extreme Flood in a Small
Tropical Basin of Costa Rica. WIT Transactions on Engineering Sciences 67:225–236.

Marcus, W. A., R. A. Marston, C. R. Colvard, and R. D. Gray. 2002. Mapping the Spatial and Temporal
Distributions of Woody Debris in Streams of the Greater Yellowstone Ecosystem, USA.
Geomorphology 44:323–335.

Marcus, W. A., J. Rasmussen, and M. A. Fonstad. 2011. Response of the Fluvial Wood System to Fire
and Floods in Northern Yellowstone. Annals of the Association of American Geographers 101:21–
44.

Marschall, E. A., T. P. Quinn, D. A. Roff, J. A. Hutchings, N. B. Metcalfe, T. A. Bakke, R. L. Saunders, and


N. L. Poff. 1998. A Framework for Understanding Atlantic Salmon (Salmo salar) Life History.
Canadian Journal of Fisheries and Aquatic Sciences 55(S1):48–58.

Marston, R. A. 1982. The Geomorphic Significance of Log Steps in Forested Streams. Annals of the
Association of American Geographers 72:99–108.

Martin, D. J., and L. E. Benda. 2001. Patterns of Instream Wood Recruitment and Transport at the
Watershed Scale. Transactions of the American Fisheries Society 130:940–958.

Martin, L. T., S. R. Pezeshki, and F. D. Shields Jr. 2004. High Oxygen Level in a Soaking Treatment
Improves Early Root and Shoot Development of Black Willow Cuttings. The Scientific World
Journal 4:899–907.

Maser, C. and J. R. Sedell, J.R. 1994. From the Forest to the Sea: The Ecology of Wood in Streams, Rivers,
Estuaries, and Oceans. St. Lucie Press.

Maser, C., and J. R. Sedell. 1988. From the Forest to the Sea: the Ecology of Wood in Streams, Rivers,
Estuaries and Oceans. Delray Beach, FL: St. Lucie Press.

Large Wood National Manual July 2015


10-38
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Maser, C., and J. M. Trappe (eds.). 1984. The Seen and Unseen World of the Fallen Tree. Gen. Tech. Rep.
PNW-164. Portland, OR: U.S. Forest Service, Pacific Northwest Forest and Range Experiment
Station.

Maser, C., R. F. Tarrant, J. M. Trappe, and J. F. Franklin (eds.). 1988. From the Forest to the Sea: A Story
of Fallen Trees. General Tech. Report PNW-GTR-229. USFS.

Massachusetts Department of Transportation (MassDOT). 2010. Design of Bridges and Culverts for
Wildlife Passage at Freshwater Streams. Highway Division, Environmental, Bridge, Construction,
and Hydraulics Sections, Boston, MA.

Masterman, R., and C. R. Thorne. 1992. Predicting Influence of Bank Vegetation on Channel Capacity.
Journal of Hydraulic Engineering 118:1052–1058.

Matheussen, B., R. L. Kirschbaum, I. A. Goodman, G. M. O’Donnell, and D. P. Lettenmaier. 2000. Effects


of Land Cover Change on Streamflow in the Interior Columbia River Basin (USA and Canada).
Hydrological Processes 14:867–885.

Mathur, D., W. H. Bason, E. J. Purdy, and C. A. Silver. 1985. A Critique of Instream Flow Incremental
Methodology. Canadian Journal of Fisheries and Aquatic Science 42(4):825–831.

May, C. L., and R. E. Gresswell. 2003a. Large Wood Recruitment and Redistribution in Headwater
Streams in the Southern Oregon Coast Range, USA. Canadian Journal of Forest Research
33:1353–1362.

May, C. L., and R. E. Gresswell. 2003b. Processes and Rates of Sediment and Wood Accumulation in
Headwater Streams of the Oregon Coast Range, USA. Earth Surface Processes and Landforms
28:409–424.

May, C. L., E. B. Welch, R. R. Horner, J. R. Karr, and B. W. Mar, B.W. 1997. Quality Indices for
Urbanization Effects on Puget Sound Lowland Streams. Water Resource Series Tech Report 154.
Seattle, Washington.

McCall, E. 1984. Conquering the Rivers. Louisiana State University Press. Baton Rouge, LA.

McClain, M. E., and 11 others. 2003. Biogeochemical Hot Spots and Hot Moments at the Interface of
Terrestrial and Aquatic Ecosystems. Ecoystems 6:301-312.

McCormick, F. H. 1993. Fish Communities as Indicators of Stream Condition. In R. M. Hughes (ed.),


EMAP Streams Bioassessment Workshop. Report of the Proceedings. EPA/600/R-93/138. U.S.
Environmental Protection Agency, Corvallis, OR.

McDade, M. H., F. J. Swanson, W. A. McKee, J. F. Franklin, and J. Van Sickle. 1990. Source Distances for
Coarse Woody Debris Entering Small Streams in Western Oregon and Washington. Canadian
Journal of Forest Research 20:326–330.

McElhany, P., M. H. Ruckelshaus, M. J. Ford, T. C. Wainwright, and E. P. Bjorkstedt. 2000. Viable


Salmonid Populations and the Recovery of Evolutionary Significant Units. U.S. Department of
Commerce, Seattle, WA. NOAA Tech. Memo NMFS-NWFSC-42.

McHenry, M. L., E. Shott, R. H. Conrad, and G. B. Grette. 1998. Changes in the Quantity and
Characteristics of LWD in Streams of the Olympic Peninsula, Washington, USA (1982-1993).
Canadian Journal of Fisheries and Aquatic Sciences 55(6):1395–1407.

Large Wood National Manual July 2015


10-39
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

McMahon, T. E. 1983. Habitat Suitability Index Models: Coho Salmon. U.S. Fish and Wildlife Service.

McMillan, LLC. 2014. Fourth of July Creek Stream Restoration Draft Design Report. Prepared for Pend
Oreille County Public Utility District. February 14, 2014.

McNeely, C., and M. E. Power. 2007. Spatial Variation in Caddisfly Grazing Regimes Within
a Northern California Watershed. Ecology 88(10):2609–2619.

Means, J. E., K. Cromack Jr., and P. C. MacMillan, 1986, Comparison of Decomposition Models Using
Wood Density of Douglas-Fir Logs. Canadian Journal of Forestry Research 15:1092–1098.

Meile, T., J. Boillat, and A. Schleiss. 2011. Flow Resistance Caused by Large-Scale Bank Roughness in
a Channel. Journal of Hydraulic Engineering 137(12):1588–1597.

Meleason, M. A., R. J. Davies-Colley, and G. M. J. Hall. 2007. Characterizing the Variability of Wood in
Streams: Simulation Modelling Compared with Multiple-Reach Surveys. Earth Surface Processes
and Landforms 32:1164–1173.

Melillo, J. M., R. J. Naiman, J. D. Aber, and K. N. Eshleman. 1983. The Influence of Substrate Quality
and Stream Size on Wood Decomposition Dynamics. Oecolgia (Berlin) 58:281–285.

Melillo, J. M., T. C. Richmond, and G. W. Yohe (eds.). 2014. Climate Change Impacts in the United
States: The Third National Climate Assessment. U.S. Global Change Research Program.
doi:10.7930/J0Z31WJ2.

Mellina, E. and S. G. Hinch. 2009. Influences of Riparian Logging and in-Stream Large Wood Removal
on Pool Habitat and Salmonid Density and Biomass: A Meta-Analysis. Canadian Journal of Forest
Research 39:1280–1301.

Meredith, C., B. Roper, and E. Archer. 2014. Reductions in Instream Wood in Streams near Roads in
the Interior Columbia River Basin. North American Journal of Fisheries Management 34(3):493–
506.

Merten, E., J. Finlay, L. Johnson, R. Newman, R., H. Stefan, and B. Vondracek. 2010. Factors
Influencing Wood Mobilization in Stream. Water Resources Research 46:W10514.

Merritt, D. M., N. L. R. Poff. 2010. Shifting Dominance of Riparian Populus and Tamarix Along
Gradients of Flow Alteration in Western North American Rivers. Ecological Applications 20:135–
152.

Merten, E. C., P. G. Vaz, J. A. Decker-Fritz, J. C. Finlay, and H. G. Stefan. 2013. Relative Importance of
Breakage and Decay as Processes Depleting Large Wood from Streams. Geomorphology 190:40–
47.

Merz, J., and P. B. Moyle. 2006. Salmon, Wildlife and Wine: Marine-Derived Nutrients in
Human-Dominated Ecosystems of Central California. Ecological Applications 13(3):999–1009.

Meyer, J. L., J. B. Wallace, and S. L. Eggert. 1998. Leaf Litter as a Source of Dissolved Organic Carbon
in Streams. Ecosystems 1:240–249.

Micheli, E. R., J. W. Kirchner, and E. W. Larsen. 2003. Quantifying the Effect of Riparian Forest Versus
Agricultural Vegetation on River Meander Migrations Rates, Central Sacramento River,
California, USA. River Research and Applications 19:1–12.

Large Wood National Manual July 2015


10-40
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Mikuś, P., B. Wyżga, R. J. Kaczka, E. Walusiak, and J. Zawiejska. 2013. Islands in a European Mountain
River: Linkages with Large Wood Deposition, Flood Flow and Plant Diversity. Geomorphology
202:115–127.

Miles, P. D., and W. B. Smith. 2009. Specific Gravity and Other Properties of Wood and Bark for
156 Tree Species Found in North America. U.S. Forest Service, Newtown Square, PA.

Miller, D., C. Luce, and L. Benda. 2003. Time, Space, and Episodicity of Physical Disturbance in
Streams. Forest Ecology and Management 178(1):121–140.

Miller, S. W., P. Budy, and J. C. Schmidt. 2010. Quantifying Macroinvertebrate Responses to In-Stream
Habitat Restoration. Applications of Restoration Ecology 18:8–19.

Millward, A. A., C. E. Kraft, and D. R. Warren. 2010. Ice Storm Damage Greater Along the
Terrestrial-Aquatic Interface in Forested Landscapes. Ecosystems 13:249–260.

Minakawa, N., and R. I. Gara. 2005. Spatial and Temporal Distribution of Coho Salmon Carcasses in a
Stream in the Pacific Northwest, USA. Hydrobiologia 539:163–166.

Minore, D. 1979. The Wild Huckleberries of Oregon and Washington: A Dwindling Resource. USDA
Forest Service Research Paper 143.

Montgomery, D. R. 1999. Process Domains and the River Continuum. Journal of the American Water
Resources Association 35:397–410.

Montgomery, D. R., and T. B. Abbe. 2006. Influence of Logjam-Formed Hard Points on the Formation
of Valley-Bottom Landforms in an Old-Growth Forest Valley, Queets River, Washington, USA.
Quaternary Research 65:147–155.

Montgomery, D. R., and J. M. Buffington. 1993. Channel Classification, Prediction of Channel Response,
and Assessment of Channel Condition. TFW-SH10-93-002. Washington State Timber, Fish &
Wildlife.

Montgomery, D. R., and J. M. Buffington. 1997. Channel-Reach Morphology in Mountain Drainage


Basins. Geological Society of America Bulletin 109:596–611.

Montgomery, D. R., and J. M. Buffington. 1998. Channel Processes, Classification and Response. Pages
13–42 in R. J. Naiman and R. E. Bilby (eds.), River Ecology and Management: Lessons from the
Pacific Coastal Ecoregion. New York: Springer.

Montgomery, D. R. and H. Piégay 2003. Wood in Rivers: Interactions with Channel Morphology and
Processes. Geomorphology 51:1–5.

Montgomery, D. R., T. B. Abbe, J. M. Buffington, N. P. Peterson, K. M. Schmidt, and J. D. Stock. 1995a.


Distribution of Bedrock and Alluvial Channels in Forested Mountain Drainage Basins. Nature
381:587–589.

Montgomery, D. R., J. M. Buffington, R. D. Smith, K. M. Schmidt, and G. Pess. 1995b. Pool Spacing in
Forest Channels. Water Resources Research 31:1097–1105.

Montgomery, D. R., J. M. Buffington, N. P. Peterson, D. Schuett-Hames, and T. P. Quinn. 1996a.


Streambed Scour, Egg Burial Depths and the Influence of Salmonid Spawning on Bed Surface
Mobility and Embryo Survival. Canadian Journal of fisheries and Aquatic Sciences 53:1061–1070.

Large Wood National Manual July 2015


10-41
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Montgomery, D. R., T. Abbe, N. P. Peterson, J. M. Buffington, K. M. Schmidt, and J. D. Stock 1996b.


Distribution of Bedrock and Alluvial Channels in Forested Mountain Drainage Basins. Nature
381:587–589.

Montgomery, D. R., B. D. Collins, J. M. Buffington, and T. B. Abbe. 2003. Geomorphic Effects of Wood
in Rivers. Pages 21–47 in S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology and
Management of Wood in World Rivers. Bethesda, MD: American Fisheries Society.

Moore, M. K. 1977. Factors Contributing to Blowdown in Streamside Leave Strips on Vancouver


Island. Land Management Report No. 3. Victoria, BC: Province of British Columbia Ministry of
Forests, Information Division.

Morris, A. E. L., P. C. Goebel, and B. J. Palik. 2010. Spatial Distribution of Large Wood Jams in Streams
Related to Stream-Valley Geomorphology and Forest Age in Northern Michigan. River Research
and Applications 26:835–847.

Morrison, M. L., B. G. Marcot, and R. W. Mannon. 1998. Wildlife-Habitat Relationships. Concepts and
Applications. Madison, WI: University of Wisconsin Press.

Moscrip, A. L., and D. R. Montgomery. 1997. Urbanization, Flood Frequency, and Salmon Abundance
in Puget Lowland Streams. Journal of the American Water Resources Association 33(6):1289–
1297.

Mossop, B., and M. J. Bradford. 2004. Importance of Large Woody Debris for Juvenile Chinook
Salmon Habitat in Small Boreal Forest Streams in the Upper Yukon River Basin, Canada.
Canadian Journal of Forest Research-Revue Canadienne De Recherche Forestiere 34:1955–1966.

Moulin, B., E. R. Schenk, and C. R. Hupp. 2011. Distribution and Characterization of In-Channel Large
Wood in Relation to Geomorphic Patterns on a Low-Gradient River. Earth Surface Processes and
Landforms 36:1137–1151.

Moussalli, E., and R. Hilborn. 1986. Optimal Stock Size and Harvest Rate in Multistage Life History
Models. Canadian Journal of Fisheries and Aquatic Sciences 43(1):135–141.

MTZ Associates. 2000. River Corridor Recreation Safety Study: Sacramento River Bank Protection
Project. Prepared for the U.S. Army Corps of Engineers, Contract 42E, Sacramento, CA.

Muir, J. 1878. Forests of California, the New Sequoia. Harper's New Monthly Magazine
LVII (CCCXLII):813–827.

Mulholland, P. J., and J. R. Webster. 2010. Nutrient Dynamics in Streams and the Role of J-NABS.
Journal of the North American Benthological Society 29:100–117.

Mulholland, P. J., J. D. Newbold, J. W. Elwood, L. A. Ferren, and J. R. Webster. 1985. Phosphorus


Spiraling in a Woodland Stream - Seasonal-Variations. Ecology 66:1012–1023.

Mulholland, P. J., and 33 others. 2009. Nitrate Removal in Stream Ecosystems Measured by
15N Addition Experiments: Denitrification. Limnology and Oceanography 54:666–680.

Munn, N. L., and J. L. Meyer. 1990. Habitat-Specific Solute Retention in Two Small Streams:
An Intersite Comparison. Ecology 71:2069–2082.

Large Wood National Manual July 2015


10-42
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Murphy, D. D., and P. S. Weiland. 2010. The Route to Best Science in Implementation of the
Endangered Species Act’s Consultation Mandate: The Benefits of Structured Effects Analysis.
Environmental Management 47(2):161–172.

Murphy, M. L. 1995. Forestry Impacts on Freshwater Habitat of Anadromous Salmonids in the


Pacific Northwest and Alaska—Requirements for Protection and Restoration. U.S. Department of
Commerce Coastal Ocean Program, NOAA. Decision Analysis Series No. 7.

Murphy, M. L., and K. V. Koski. 1989. Input and Depletion of Woody Debris in Alaska Streams and
Implications for Streamside Management. North American Journal of Fisheries Management
9(4):427–436.

Murphy, M. L., J. Heifetz, S. W. Johnson, K. V. Koski, and J. F. Thendinga. 1986. Effects of Clear-Cut
Logging with and without Buffer Strips on Juvenile Salmonids in Alaskan Streams. Canadian
Journal of Fisheries and Aquatic Sciences 43:1521–1533.

Mutz, M. 2003. Hydraulic Effects of Wood in Streams. American Fisheries Society Symposium 37:93–
107.

Mutz, M., E. Kalbus, and S. Meinecke. 2007. Effect of Instream Wood on Vertical Water Flux in
Low-Energy Sand Bed Flume Experiments. Water Resources Research 43:W10424.

Nagayama, S., and F. Nakamura. 2010. Fish Habitat Rehabilitation Using Wood in the World.
Landscape and Ecological Engineering 6(2):289–305.

Nagayama, S., F. Nakamura, Y. Kawaguchi, and D. Nakano. 2012. Effects of Configuration of Instream
Wood on Autumn and Winter Habitat Use by Fish in a Large Remeandering Reach. Hydrobiologia
680:159–170.

Naiman, R. J., T. J. Beechie, L. E. Benda, P. A. Bisson, L. H. MacDonald, M. D. O’Conner, P. L. Olsen, and


E. A. Steel. 1992. Fundamental Elements of Ecologically Healthy Watersheds in the Pacific
Northwest Coastal Ecoregion. Pages 127–188 in R. J. Naiman (ed.), Watershed Management:
Balancing Sustainability and Environmental Change. Springer: New York.

Naiman, R. J., K. L. Fetherston, S. McKay, and J. Chen. 1998. Riparian Forests. Pages 289–323 in
R. J. Naiman and R. E. Bilby (eds.), River Ecology and Management: Lessons from the Pacific
Coastal Ecoregion. Springer-Verlag: New York.

Naiman, R. J., R. E. Bilby, and P. Bisson. 2000. Riparian Ecology and Management in the Pacific
Coastal Rain Forest. Bioscience 50:996–1011.

Naiman, R. J., E. V. Balian, K. K. Bartz, R. E. Bilby, and J. J. Latterell. 2002a. Dead Wood Dynamics in
Stream Ecosystems. Pages 23–48 in W. F. Laudenslayer Jr., P. J. Shea, B. E. Valentine, C. P.
Weatherspoon, and T. E. Lisle (eds.), Proceedings of the Symposium on the Ecology and
Management of Dead Wood in Western Forests. Gen. Tech. Rep. PSW-GTR-181. U.S. Forest
Service, Pacific Southwest Forest and Range Experiment Station.

Naiman R. J., S. E. Bunn, and C. Nilsson. 2002b. Legitimizing Fluvial Ecosystems as Users of Water.
Environmental Management 30:455–467.

Naiman, R. J., R. E. Bilby, D. E. Schindler, and J. M. Helfield. 2002c. Pacific Salmon, Nutrients, and the
Dynamics of Freshwater and Riparian Ecosystems. Ecosystems 5:399–417.

Large Wood National Manual July 2015


10-43
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Naiman, R. J., J. S. Becthold, T. J. Beechie, J. J. Laterell, and R. Van Pelt. 2010. A Process-Based View of
Floodplain Forest Patterns in Coastal River Valleys of the Pacific Northwest. Ecosystems 13:1–31.

Nakamura, F., and F. J. Swanson. 1993. Effects of Coarse Woody Debris on Morphology and Sediment
Storage of a Mountain Stream System in Western Oregon. Earth Surface Processes and Landforms
18:43–61.

Nanson, G. C., M. Barbetti, and G. Taylor. 1995. River Stabilisation due to Changing Climate and
Vegetation During the late Quaternary in Western Tasmania, Australia. Geomorphology
13.1(1995):145–158.

National Institute for Occupational Safety and Health. 2012. Logging Safety. Centers for Disease
Control and Prevention. Atlanta, GA. Available: http://www.cdc.gov/niosh/topics/logging/.
Accessed: July 10, 2014.

National Marine Fisheries Service. 1996. Making Endangered Species Act Determinations of Effect for
Individual or Grouped Actions at the Watershed Scale. Environmental and Technical Services
Division, Habitat Conservation Branch.

National Research Council. 1999. Committee on Health Risks of Exposure to Radon. Health Effects of
Exposure to Radon. Vol. 6. National Academies Press.

National Research Council. 2004. Adaptive Management for Water Resources Planning, The
National Academies Press. Washington, D.C.
Natural Resources Conservation Service (NRCS). 2007a. Streambank Soil Bioengineering. Technical
Supplement TS 41I in Stream Restoration Design, National Engineering Handbook Part 654.
USDA-NRCS, Washington, D.C. CD-ROM.

Natural Resources Conservation Service (NRCS). 2007b. Use and Design of Soil Anchors. Technical
Supplement TS 41E in Stream Restoration Design, National Engineering Handbook Part 654.
USDA-NRCS, Washington, D.C. CD-ROM.

National Resources Conservation Service (NRCS). 2007c. Chapter 15—Project Implementation.


In Part 654, Stream Restoration Design National Engineering Handbook. U.S. Department of
Agriculture, Washington, D.C.

National Resources Conservation Service (NRCS). 2007d. Technical Supplement 14I—Streambank


Soil Bioengineering. In Part 654, Stream Restoration Design National Engineering Handbook.
U.S. Department of Agriculture, Washington, D.C.

National Resources Conservation Service (NRCS) 2007e. Use of Large Woody Material for Habitat
and Bank Protection - Technical Supplement 14j of Part 654. The National Engineering
Handbook. 210–VI–NEH, August 2007. U.S. Department of Agriculture, Washington, D.C.

Natural Resources Conservation Service (NRCS). 2007f. Stream Hydrology. Chapter 5 in Stream
Restoration Design, National Engineering Handbook Part 654. USDA-NRCS, Washington, D.C.
CD-ROM.

Navel, S., C. Piscart, F. Meermillod-Blondin, and P. Marmonier. 2013. New Methods for the
Investigation of Leaf Litter Breakdown in River Sediments. Hydrobiologia 700:301–312.

Large Wood National Manual July 2015


10-44
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Neumann, R. M., and T. L.Wildman. 2002. Relationships Between Trout Habitat Use and Woody
Debris in Two Southern New England Streams. Ecology of Freshwater Fish 11:240–250.

New York State Department of Environmental Conservation (NYSDEC). 2014. Removal of Woody
Debris and Trash from Rivers and Streams. In Post-Flood Stream Reconstruction: Guidelines and
Best Practices. Albany, NY.

Nichols, R. A. and S. G. Sprague. 2003. The Use of Long-Line Cabled Logs for Stream Bank
Rehabilitation. Pages 422–442 in D. R. Montgomery, S. M. Bolton, D. B. Booth, and L. Wall (eds.),
Restoration of Puget Sound Rivers. University of Washington Press: Seattle.

Nickelson, T. E., and P. W. Lawson. 1998. Population Viability of Coho Salmon, Oncorhynchus kisutch,
in Oregon Coastal Basins: Application of a Habitat-Based Life Cycle Model. Canadian Journal of
Fisheries and Aquatic Sciences 55:2383–2392.

Nickelson, T. E., M. F. Solazzi, S. L. Johnson, and J. D. Rodgers. 1992. Effectiveness of Selected Stream
Improvement Techniques to Created Suitable Summer and Winter Rearing Habitat for Juvenile
Coho Salmon (Oncorhynchus kisutch) in Oregon Coastal Streams. Canadian Journal of Fisheries
and Aquatic Sciences 49:790–794.

Nickelson, T. E., M. F. Solazzi, S. L. Johnson, and J. D. Rodgers. 1993. An Approach to Determining


Stream Carrying Capacity and Limiting Habitat For Coho Salmon (Oncorhynchus kisutch). Paper
read at Proceedings of the Coho Workshop, May 1992, Nanaimo, B.C.

Niezgoda, S., and P. Johnson. 2007. Case Study in Cost-Based Risk Assessment for Selecting a Stream
Restoration Design Method for a Channel Relocation Project. Journal of Hydraulic Engineering
133(5):468–481

Nilsson, C., and K. Berggren. 2000. Alterations of Riparian Ecosystems Caused by River Regulation.
Bioscience 50:783–792.

Nooksack Tribe. 2013. ELJ Assessment.

Norris, R. H. 1995. Biological Monitoring: The Dilemma of Data Analysis." Journal of the North
American Benthological Society:440–450

North American Forest Commission. 2011. Forests of North America. Vector Digital Data. Food and
Agriculture Organization of the United Nations. Commission for Environmental Cooperation.
Montreal, Quebec, CA.

Noss, F., and R. L. Peters. 1995. Endangered Ecosystems – A Status Report on America’sVanishing
Habitat and Wildlife. 133 Pages. Defenders of Wildlife, 1101 Fourteenth Street, NW, Suite 1400,
Washington, DC 20005.

Nowinski, J. D., M. B. Cardenas, A. F. Lightbody, T. E. Swanson, and A. H. Sawyer. 2012. Hydraulic and
Thermal Response of Groundwater-Surface Water Exchange to Flooding in an Experimental
Aquifer. Journal of Hydrology 472–473:184–192.

Oakley, K. L., L. P. Thomas, and S. G. Fancy. 2003. Guidelines For Long-Term Monitoring Protocols.
Wildlife Society Bulletin 31(4):1000–1003.

Large Wood National Manual July 2015


10-45
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Occupational Safety and Health Administration (OSHA). Undated a. Logging eTool. Occupational
Safety and Health Administration. Washington, D.C. Available: https://www.osha.gov/SLTC/
etools/logging/index.html. Accessed: July 10, 2014.

Occupational Safety and Health Administration (OSHA). Undated b. Trenching and Excavation Safety.
Fact Sheet. Occupational Safety and Health Administration. Washington, D.C. Available:
https://www.osha.gov/SLTC/trenchingexcavation/construction.html. Accessed: July 10, 2014.

O'Connor, J. E., M. A. Jones, and T. L. Haluska. 2003. Flood Plain and Channel Dynamics of the
Quinault and Queets Rivers, Washington, U.S.A. Geomorphology 51:31–59.

O’Connor, M., and G. Watson, 1998. Geomorphology of Channel Migration Zones and Implications for
Riparian Forest Management. Available: http://www.oei.com/Reports/Geomorph_of_CMZ.htm.

O'Connor, R. R., and F. J. Rahel. 2009. A Patch Perspective on Summer Habitat Use by Brown Trout
Salmo trutta in a High Plains Stream in Wyoming, USA. Ecology of Freshwater Fish 18(3):473–
480.

Oglethorpe, J. 2002. Adaptive Management: From Theory to Practice. The World Conservation Union
(IUCN).

Oliver, C. D. 1980/1981. Forest Development in North America Following Major Disturbances. Forest
Ecology and Management 3:153–168.

Opperman, J. J., and A. M. Merenlender. 2004. The Effectiveness of Riparian Restoration for
Improving Instream Fish Habitat in Four Hardwood-Dominated California Streams. North
American Journal of Fisheries Management 24(3):822–834.

Oppermanh, J. J., and A. M. Merenlender. 2007. Living Trees Provide Stable Large Wood in Streams.
Earth Surface Processes & Landforms 32(8):1229–1238.

Opperman, J. J., A. M. Merenlender, and D. Lewis. 2006. Maintaining Wood in Streams: A Vital Action
for Fish Conservation. Davis, CA: University of California. Publication 8157.

Oregon Department of Fish and Wildlife (ODFW). 2010. Guide to Placement of Wood, Boulders, and
Gravel for Habitat Restoration. Oregon Departments of Forestry and Fish and Wildlife. Salem, OR.

Oregon Department of Forestry. 1995. A Guide to Placing Large Wood in Streams. Salem, OR, Forest
Practices Section.

Oregon Department of Transportation (ODOT). 2011. Hydraulics Manual, Engineering and Asset
Management. Unit Geo-Environmental Section. Salem, OR.

Oswald, E. B., and E. Wohl. 2008. Wood-Mediated Geomorphic Effects of a Jökuhlaup in the Wind
River Mountains, Wyoming. Geomorphology 100:549–562.

Pahl-Wostl, C. 1995. The Dynamic Nature of Ecosystems: Chaos and Order Entwined. Chichester:
Wiley.

Palik, B., S. W. Golladay, P. C. Goebel, and B. W. Taylor. 1998. Geomorphic Variation in Riparian Tree
Mortality and Stream Coarse Woody Debris Recruitment from Record Flooding in a Coastal Plain
Stream. Ecoscience 5:551–560.

Large Wood National Manual July 2015


10-46
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Pariset, E., R. Hausser, and A. Gagnon. 1966. Formation of Ice Covers and Ice Jams in Rivers. Journal
of the Hydraulics Division 92(6):1–24.

Parrish, R. M. and P. B. Jenkins. 2012. Natural Log Jams in the White River: Lessons for Geomimetic
Design of Engineered Log Jams. U.S. Fish and Wildlife Service, Leavenworth, WA.

Pastorok, R. A., A. MacDonald, J. R. Sampson, P. Wilber, D. J Yozzo, and J. P Titre. 1997. An Ecological
Decision Framework for Environmental Restoration Projects. Ecological Engineering 9(1–2):89–
107.

Paukert, C. P., and A. S. Makinster. 2008. Longitudinal Patterns in Flathead Catfish Relative
Abundance and Length at Age Within a Large River: effects of an urban gradient. River Research
and Applications. Available: www.interscience.wiley.com. DOI: 10.1002/rra.1089.

Pealer, S. 2012. Lessons from Irene – Building Resiliency as We Rebuild. Vermont Agency of Natural
Resources Climate Change Team, Montpelier, VT. Available: http://www.anr.state.vt.us/
anr/climatechange/Pubs/Irene_Facts.pdf.

Pearsons, T. D., and H. W. Li. 1992. Influence of Habitat Complexity on Resistance to Flooding and
Resilience of Stream Fish Assemblages. Transactions of the American Fisheries Society 121:427–
436.

Pelto, M. S. 2011. North Cascade Glacier Retreat. Nichols College, Dudley, MA. Available:
http://www.nichols.edu/departments/glacier/bill.htm.

Pess, G. R., D. R. Montgomery, E. A. Steel, R. E. Bilby, B. E. Feist, and H. M. Greenberg. 2002. Landscape
Characteristics, Land Use, and Coho Salmon (Oncorhynchus kisutch) Abundance, Snohomish
River, Washington, USA. Canadian Journal of Fisheries and Aquatic Sciences 59:613–623.

Pess, G. R., M. C. Liermann, M. L. Mchenry, R. J. Peters, and T. R. Bennett. 2012. Juvenile Salmon
Response to the Placement of Engineered Log Jams (ELJs) in the Elwha River, Washington State,
USA. River Research and Applications 28(7):872–881.

Peters, P. J., B. R. Missildine, and D. L. Low. 1998. Seasonal Fish Densities near River Banks Stabilized
with Various Stabilization Methods. First Year Report of the Flood Technical Assistance Project.
U.S. Fish and Wildlife Service, North Pacific Coast Ecoregion. Western Washington Office,
Aquatic Resources Division. Lacey, WA.

Petersen, M. S. 1986. River Engineering. Englewood Cliffs, NJ: Prentice-Hall.

Pettit, N. E., and R. J. Naiman. 2006. Flood-Deposited Wood Creates Regeneration Niches for Riparian
Vegetation on a Semi-Arid South African River. Journal of Vegetation Science 17:615–624.

Pettit, N. E., R. J. Naiman, K. H. Rogers, and J. E. Little. 2005. Post-Flooding Distribution and
Characteristics of Large Woody Debris Piles Along the Semi-Arid Sabie River, South Africa. River
Research and Applications 21:27–38.

Petts, G. E., A. L. Roux, and H. Moller (eds.). 1989. Historical Changes of Large Alluvial Rivers, Western
Europe. Chichester: John Wiley.

Phillips, E. C. 2003. Habitat Preference of Aquatic Macroinvertebrates in an East Texas Sandy


Stream. Journal of Freshwater Ecology 18(1):1–11.

Large Wood National Manual July 2015


10-47
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Phillips, J. D. 2012. Log-Jams and Avulsions in the San Antonio River Delta, Texas. Earth Surface
Processes and Landforms 37:936–950.

Phillips, J. D., and L. Park. 2009. Forest Blowdown Impacts of Hurricane Rita on Fluvial Systems.
Earth Surface Processes and Landforms 34:1069–1081.

Pickett, S. T. A., and K. H. Rogers. 1997. Patch Dynamics: The Transformation of Landscape Structure
and Function. Pages 101–127 in J. A. Bissonette (ed.), Wildlife and Landscape Ecology. New York:
Springer-Verlag.

Pianka, E. R. 1994. Evolutionary Ecology. Harper-Collins.

Pickett, S. T. A., and K. H. Rogers. 1997. Patch Dynamics: The Transformation of Landscape Structure
and Function. Pages 101–127 in J. A. Bissonette (ed.), Wildlife and Landscape Ecology. New York:
Springer-Verlag.

Piégay, H. 1993. Nature, Mass and Preferential Sites of Coarse Woody Debris Deposits in the Lower
Ain Valley (Mollon Reach), France. Regulated Rivers: Research and Management 8:359–372.

Piégay, H., and J. P. Bravard. 1997. Response of a Mediterranean Riparian Forest to a 1 in 400 Year
Flood, Ouveze River, Drome-Vaucluse, France. Earth Surface Processes and Landforms 22(1):31–
43.

Piégay, H., A. and R. A. Marston. 1998. Distribution of Coarse Woody Debris Along the Concave Bank
of a Meandering River (the Ain River, France). Physical Geography 19(4):318–340.

Piégay, H., A. Thevenet, and A. Citterio. 1999. Input, Storage and Distribution of LWD Along a
Mountain River Continuum, the Drôme River, France. Catena 35:19–39.

Pittman, S. E., and M. E. Dorcas. 2009. Movements, Habitat Use, and Thermal Ecology of an Isolated
Population of Bog Turtles (Glyptemys muhlenbergii). Copeia(4):781–790.

Plafkin, J. L., Barbour, M. T., Porter, K. D., Gross, S. K. and Hughes, R. M. 1992. Rapid Bioassessment
Protocols for Use in Streams and Rivers: Benthic Macroinvertebrates and Fish. Washington, D.C.:
EPA.

Poff, N. L., and H. K. H. Zimmerman. 2010. Ecological Responses to Altered Flow Regimes:
A Literature Review to Inform the Science and Management of Environmental Flows. Freshwater
Biology 55:194–205.

Poff, N. L., J. D. Allan, M. A. Palmer, D. D. Hart, B. D. Richter, A. H. Arthington, K. H. Rogers, J. L. Meyer,


and J. A. Stanford. 2003. River Flows and Water Wars: Emerging Science for Environmental
Decision Making. Frontiers in Ecology and the Environment 1:298–306.

Poff, N. L., B. P. Bledsoe, and C. O. Cuhaciyan. 2006. Hydrologic Variation with Land Use across the
Contiguous United States: Geomorphic and Ecological Consequences for Stream Ecosystems.
Geomorphology 79:264–285. doi:10.1016/ j.geomorph.2006.06.032.

Pohl, M. M. 2002. Bringing Down Our Dams: Trends in American Dam Removal Rationales. Journal of
the American Water Resources Association 38:1511–1519.

Pokrefke, T. J. (ed.) 2013. Inland Navigation: Channel Training Works. ASCE Manual of Practice 124.
American Society of Civil Engineers. Reston, VA.

Large Wood National Manual July 2015


10-48
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Pollen-Bankhead, N., and A. Simon. 2010. Hydrologic and Hydraulic Effects of Riparian Root
Networks on Streambank Stability: Is Mechanical Root-Reinforcement the Whole Story?
Geomorphology 116(3):353–362.

Pollock, K. H., J. D. Nichols, T. R. Simons, G. L. Farnsworth, L. L. Bailey, and J. R. Sauer. 2002. Large
Scale Wildlife Monitoring Studies: Statistical Methods for Design and Analysis. Environmetrics
13:105–119.

Pollock, M. M., and T. J. Beechie. 2014. Does Riparian Forest Restoration Thinning Enhance
Biodiversity? The Ecological Importance of Large Wood. JAWRA Journal of the American Water
Resources Association 50(3):543–559. Online publication date: June 1, 2014.

Pollock, M. M., R. J. Naiman, and T. A. Hanley. 1998. Plant Species Richness in Riparian Wetlands—
A Test of Biodiversity Theory. Ecology 79:94–105.

Pollock, M. M., T. J. Beechie., M. Liermann, and R. E. Bigley. 2009. Stream Temperature Relationships
to Forest Harvest in Western Washington. Journal of the American Water Resources Association
45(1):141–156.

Poole, G. C., J. A. Stanford, S. W. Running, and C. A. Frissell. 2006. Multiscale Geomorphic Drivers of
Groundwater Flow Paths: Subsurface Hydrologic Dynamics and Hyporheic Habitat Diversity.
Journal of the North American Benthological Society 25:288–303.

Power, M. E., and W. E. Dietrich. 2002. Food Webs in River Networks. Ecological Research 17:451–
471.

Power, M. E., A. Sun, G. Parker, W. E. Dietrich, and J. T. Wootton. 1995. Hydraulic Food Chain Models.
BioScience 45:159–167.

Prato, T. 2003. Adaptive Management of Large Rivers with Special Reference to the Missouri River.
Journal of the American Water Resources Association 39(4):935–946.

Prowse, T. D. 2001. River Ice Ecology. 1: Hydrologic, Geomorphic, and Water Quality Aspects. Journal
of Cold Regions Engineering 15(1):1–16.

Ptolemy, R. A. 1993. Maximum Salmonid Densities in Fluvial Habitats in British Columbia. Paper
read at Proceedings of the Coho Workshop, May 26–28, 1992, at Nanaimo, B.C.

Quinault Indian Nation (QIN). 2008. Salmon Habitat Restoration Plan for the Upper Quinault River.
Quinault Indian Nation Department of Fisheries. Taholah, Washington. Prepared by T. Abbe and
others.

Quinn, T. P., S. M. Carlson, S. M. Gende, and H. B. Rich. 2009. Transportation of Pacific Salmon
Carcasses from Streams to Riparian Forests by Bears. Canadian Journal of Fisheries and Aquatic
Science 87:195–203.

Raffa, K. F., B. H. Aukema, B. J. Bentz, A. L. Carroll, J. A. Hicke, M. G. Turner, and W. H. Romme. 2008.
Cross-Scale Drivers of Natural Disturbances Prone to Anthropogenic Amplification: The
Dynamics of Bark Beetle Eruptions. Bio-Science 58:501–517. doi:10.1641/b580607. Available:
http://www.jstor.org/stable/pdfplus/10.1641/B580607. pdf.

Large Wood National Manual July 2015


10-49
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Raikow, D. F., S. A. Grubbs, and K. W. Cummins. 1995. Debris Dam Dynamics and Coarse Particulate
Organic Matter Retention in an Appalachian Mountain Stream. Journal of the North American
Benthological Society 14:535–546.

Railsback, S. F., and J. Kadvany. 2008. Demonstration Flow Assessment: Judgment and Visual
Observation in Instream Flow Studies. Fisheries 33:217–227.

Railsback, S. F., H. Stauffer, and B. Harvey. 2003. What can Habitat Preference Models Tell Us? Tests
using a Virtual Trout Population. Ecological Applications 13(6):1580–1594.

Ralph, S. C., G. C. Poole, L. L. Conquest, and R. J. Naiman. 1991. Stream Channel Morphology and
Woody Debris in Logged and Unlogged Basins of Western Washington. Canadian Journal of
Fisheries and Aquatic Sciences 51:37–51.

Rapp, C., and T. Abbe. 2003. A Framework for Delineating Channel Migration Zones. Washington State
Department of Ecology Publication Number 03-06-027. Final Draft.

Raup, H. M. 1957. Vegetation Adjustment to the Instability of Sites. Proceedings and Papers of the 6th
Technical Meeting of the International Union for Conservation of Nature and Natural Resources.
Edinburgh. Pages 36–48.

Ravazzolo, D., L. Mao, L. Picco, and M. A. Lenzi 2015. Tracking Log Displacement During Floods in the
Tagliamento River Using Rfid and Gps Tracker Devices. Geomorphology 228:226-233.

Reeves, G. H., F. H. Everest, and T. E. Nickelson. 1989. Identification of Physical Habitats Limiting the
Production of Coho Salmon in Western Oregon and Washington. Portland, OR: USDA Forest
Service.

Reeves, G. H., J. D. Hall, T. D. Roelofs, T. L. Hickman, and C. O. Baker. 1991. Rehabilitating and
Modifying Stream Habitats. Pages 519–557 in Influences of Forest and Rangeland Management
on Salmonid Fishes and Their Habitats. American Fisheries Society Special Publication 19.

Reich, P. B., I. Wright, J. Cavender-Bares, J. Craine, J. Oleksyn, M. Westoby, and M. B. Walters. 2003.
The Evolution of Plant Functional Variation: Traits, Spectra, and Strategies. International Journal
of Plant Sciences 164:s143–s164.

Reid, L. M., and T. Dunne. 1996. Rapid Evaluation of Sediment Budgets. Reiskirchen, Germany: Catena
Verlag (GeoEcology paperback).

Reid, L. M., and S. Hilton. 1998. Buffering the Buffer. Pages 71–80 in R. R. Ziemer (ed.), Proceedings of
the Conference on Coastal Watersheds: The Caspar Creek Story; held May 6, 1998, in Ukiah,
California. USDA Forest Service, Pacific Southwest Research Station, General Technical Report
PSW-GTR-168.

Reisenbichler, R. R. 1989. Utility of Spawner-Recruit Relations for Evaluating the Effect of Degraded
Environment on the Abundance of Chinook Salmon, Oncorhynchus tshawytscha. Pages 21–32 in
C. D. Levings, L. B. Holtby, and M. A. Henderson (eds.), Proceedings of the National Workshop on
Effects of Habitat Alteration on Salmonid Stocks: Canadian Special Publication on Fisheries and
Aquatic Sciences 105.

Resh, V. H., R. H. Norris, and M. T. Barbour. 1995. Design and Implementation of Rapid Assessment
Approaches for Water Resource Monitoring Using Benthic Macroinvertebrates. Australian
Journal of Ecology 20(1):108–121.

Large Wood National Manual July 2015


10-50
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Reynoldson, T. B., R. C. Bailey, K. E. Day, and R. J. Norris. 1995. Biological Guidelines for Freshwater
Sediment Based on BEnthic Assessment of SedimenT (the BEAST) Using a Multivariate
Approach for Predicting Biological State. Australian Journal of Ecology 20(1):198–219.

Rheinhardt, R., M. Brinson, G. Meyer, and K. Miller. 2012. Integrating Forest Biomass and Distance
from Channel to Develop an Indicator of Riparian Condition. Ecological Indicators 23:46–55.

Richards, K. 1982. Rivers: Form and Process in Alluvial Channels. New York: Methuen.

Richmond, A. D., and K. D. Fausch. 1995. Characteristics and Function of Large Woody Debris in
Subalpine Rocky Mountain Streams in Northern Colorado. Canadian Journal of Fisheries and
Aquatic Sciences 52:1789–1802.

Rigon, E., F. Comiti, and M. A. Lenzi. 2012. Large Wood Storage in Streams of the Eastern Italian Alps
and the Relevance of Hillslope Processes. Water Resources Research 48:W01518,
doi:10.1029/2010WR009854 18 p.

Riley, A. L. 1998. Restoring Streams in Cities: A Guide for Planners, Policymakers, and Citizens.
Washington, D.C. Island Press.

Riley, S. C. and K. D. Fausch. 1995. Trout Population Response to Habitat Enhancement in Six
Northern Colorado Streams. Canadian Journal of Fisheries and Aquatic Sciences. 52:34–53.

Roadway Safety Alliance. 2005. Internal Traffic Control Plans. Laborers' Health and Safety Fund of
North America. Washington, D.C.

Roadway Safety Alliance. Undated. Internal Traffic Control Plans. Developed under a contract with
the Centers for Disease Control and Prevention contract No. 212-2003-M-02677, Laborers’
Health and Safety Fund of North America, Washington, D.C.

Robert, A. 1997. Characteristics of Velocity Profiles Along Riffle-Pool Sequences and Estimates of
Bed Shear Stresses. Geomorphology 19:89–98.

Robison, E. G. and R. L. Beschta. 1990. Identifying Trees in Riparian Areas that can Provide Coarse
Woody Debris to Streams. Forest Science 36:790–801.

Roloff, G. J., G. F. Wilhere, T. Quinn, and S. Kohlmann. 2001. An Overview of Models and Their Role in
Wildlife Management. Pages 521–536 in T. J. Johnson and T. A. O'Neil (eds.), Wildlife-Habitat
Relationships in Oregon and Washington. Corvallis, OR: Oregon State University.

Roni, P. 2002. Habitat Use by Fishes and Pacific Giant Salamanders in Small Western Oregon and
Washington Streams. Transactions of the American Fisheries Society 131(4):743–761.

Roni, P. (ed.). 2005. Monitoring Stream and Watershed Restoration. American Fisheries Society,
Bethesda, MD.

Roni, P., and T. P. Quinn. 2001. Density and Size of Juvenile Salmonids in Response to Placement of
Large Woody Debris in Western Oregon and Washington Streams. Canadian Journal of Fisheries
and Aquatic Sciences 58:282–292.

Roni, P., M. Liermann, and A. Steel. 2003. Monitoring and Evaluating Fish Response to Instream
Restoration. In D. Montgomery, S. Bolton, D. Booth, and L. Wall (eds.), Restoration of Puget Sound
Rivers. Center for Water and Watershed Studies. University of Washington Press: Seattle.

Large Wood National Manual July 2015


10-51
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Roni, P., T. Bennett, S. Morley, G. R. Pess, K. Hanson, D. Van Slyke, and P. Olmstead. 2006.
Rehabilitation of Bedrock Stream Channels: The Effects of Boulder Weir Placement on Aquatic
Habitat and Biota. River Research and Applications 22(9):967–980.

Roni, P., K. Hanson, and T. Beechie. 2008. Global Review of the Physical and Biological Effectiveness
of Stream Habitat Rehabilitation Techniques. North American Journal of Fisheries Management
28(3):856–890.

Roni, P., G. Pess, T. Beechie, and S. Morley. 2010. Estimating Changes in Coho Salmon and Steelhead
Abundance from Watershed Restoration: How Much Restoration is Needed to Measurably
Increase Smolt Production? North American Journal of Fisheries Management 30(6):1469–1484.

Roni, P., T. J. Beechie, G. R. Pess, and K. M. Hanson. 2014a. Wood Placement in River Restoration:
Fact, Fiction and Future Direction. Canadian Journal of Fisheries and Aquatic Sciences.

Roni, P., G. R. Pess, and T. J. Beechie. 2014b. Fish-Habitat Relationships and Effectiveness of Habitat
Restoration. National Marine Fisheries Service. Seattle, WA. NOAA Technical Memorandum
NMFS-NWFSC-127.

Rood, S. B., and J. M. Mahoney. 1990. Collapse of Riparian Poplar Forests Downstream from Dams in
Western Prairies: Probable Causes and Prospects for Mitigation. Environmental Management
14:451–464.

Rood, S. B., C. R. Gourley, E. M. Ammon, L. G. Heki, J. R. Klotz, M. L. Morrison, D. Mosley, G. G.


Scoppettone, S. Swanson, and P. L. Wagner. 2003. Flows for Floodplain Forests: A Successful
Riparian Restoration. Bioscience 53:647–656.

Roper, B. B., J. M. Buffington, S. Bennett, S. H. Lanigan, E. Archer, S. T. Downie, J. Faustini, T. W.


Hillman, S. Hubler, K. Jones, C. Jordan, P. K. Kauffman, G. Merritt, C. Moyer, and A. Pleus. 2010. A
Comparison of the Performance and Compatibility of Protocols Used by Seven Monitoring
Groups to Measure Stream Habitat in the Pacific Northwest. North American Journal of Fisheries
Management 30:565–587.

Rose, K. A. 2000. Why are Quantitative Relationships Between Environmental Quality and Fish
Populations so Elusive? Ecological Applications 10(2):367–385.

Rose, S. and N. E. Peters. 2001. Effects of Urbanization on Streamflow in the Atlanta Area (Georgia,
USA): A Comparative Hydrological Approach. Hydrological Processes 15:1441–1457.

Rosenberg, D. M., and V. H. Resh (eds.). 1993. Freshwater Biomonitoring and Benthic
Macroinvertebrates. Springer.

Rosenfeld, J. 2003. Assessing the Habitat Requirements of Stream Fishes: an Overview and
Evaluation of Different Approaches. Transactions of the American Fisheries Society 132:953–968.

Rosenfeld, J. S. 2014. Modelling the Effects of Habitat on Self-thinning, Energy Equivalence, and
Optimal Habitat Structure for Juvenile Trout. Canadian Journal of Fisheries and Aquatic Sciences.
71(9):1395–1406.

Rosenfeld, J. S., and S. Boss. 2001. Fitness Consequences of Habitat Use for Juvenile Cutthroat Trout:
Energetic Costs and Benefits in Pools and Riffles. Canadian Journal of Fisheries and Aquatic
Sciences 58(3):585–593.

Large Wood National Manual July 2015


10-52
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Rosenfeld, J., and T. Hatfield. 2006. Information Needs for Assessing Critical Habitat of Freshwater
Fish. Canadian Journal of Fisheries and Aquatic Sciences 63:683–698.

Rosenfeld, J. S., and L. Huato. 2003. Relationship Between Large Woody Debris Characteristics and
Pool Formation in Small Coastal British Columbia Streams. North American Journal of Fisheries
Management 23:928–938.

Rosenfeld, J. S., and R. Ptolemy. 2012. Modelling Available Habitat Versus Available Energy Flux: Do
PHABSIM Applications that Neglect Prey Abundance Underestimate Optimal Flows for Juvenile
Salmonids? Canadian Journal of Fisheries and Aquatic Sciences 69(12):1920–1934.

Rosgen, D., and H. L. Silvey 1996. Applied River Morphology. Pagosa Springs, CO: Wildland Hydrology.

Rot, B. 1993. Windthrow in Stream Buffers on Coastal Washington Streams. ITT-Rayonier Inc.

Rot, B. W., R. J. Naiman, and R. E. Bilby. 2000. Stream Channel Configuration, Landform, and Riparian
Forest Structure in the Cascade Mountains, Washington. Canadian Journal of Fisheries and
Aquatic Sciences 57:699–707.

Ruffner, E. H. 1886. The Practice of the Improvement of the Non-Tidal Rivers of the United States, with
an Examination of the Results Thereof. New York, NY: John Wiley and Sons.

Ruiz-Villanueva, V., A. Díez-Herrero, J. M. Bodoque, and E. Bladé. 2014a. Large Wood in Rivers and its
Influence on Flood Hazard. Cuadernos de Investigación Geográfica 40:229–246.

Ruiz-Villanueva, V., M. Stoffel, H. Piegay, V. Gaertner, and F. Perret. 2014b. Wood Density Assessment
to Improve Understanding of Large Wood Buoyancy in Rivers. Pages 2503–2508 in A. Schleiss,
G. De Cesare, M. Franca, and M. Pfister (eds.), River Flow. London, England: Taylor and Francis.

Rumpff, L., Duncan, D. H., P. A. Vesk, D. A. Keith, and B. A. Wintle. 2011. State-and-Transition
Modelling for Adaptive Management of Native Woodlands. Biological Conservation 144(4):1224–
1236.

Russell, I. C. 1898. Rivers of North America. New York: G.P. Putnams Sons.

Russell, I. C. 1909. Rivers of North America. New York, NY: G.P. Putnam and Sons.

Russell, K., and E. Holburn. 2009. Field Evaluation of Engineered Large Woody Debris for Structure
Performance and Habitat Value. Pages 3234–3243 in World Environmental and Water Resources
Congress 2009. American Society of Civil Engineers.

Rutherford, I., B. Anderson, and A. Ladson. 2007. Managing the Effects of Riparian Vegetation on
Flooding. In S. Lovett and P. Price (eds.), Principles for Riparian Lands Management. Land &
Water Australia, Canberra.

Ryan, S. E., E. L. Bishop, and J. M. Daniels. 2014. Influence of Large Wood on Channel Morphology and
Sediment Storage in Headwater Mountain Streams, Fraser Experimental Forest, Colorado.
Geomorphology 217:73–88.

Sabater, S., V. Acuña, A. Giorgi, E. Guerra, I. Muñoz, and A. M. Romaní, 2005. Effects of Nutrient Inputs
in a Forested Mediterranean Stream Under Moderate Light Availability. Archiv für Hydrobiologie
163:479–496.

Large Wood National Manual July 2015


10-53
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Saldi-Caromile, K., K. Bates, P. Skidmore, J. Barenti, and D. Pineo. 2004. Stream Habitat Restoration
Guidelines: Final Draft. Co-published by the Washington Departments of Fish and Wildlife and
Ecology and the U.S. Fish and Wildlife Service. Olympia, Washington.

Salmon Recovery Funding Board (SRFB). 2013. Manual 18 Salmon Recovery Grants. Washington
State Recreation and Conservation Office. Salmon Recovery Funding Board. January.

Sass, G. G., J. F. Kitchell, S. R. Carpenter, T. R. Hrabik, A. E. Marburg, and M. G. Turner. 2006. Fish
Community and Food Web Responses to a Whole-Lake Removal of Coarse Woody Habitat.
Fisheries 31:321–330.

Sauer, V. B. 1974. Flood Characteristics of Oklahoma Streams Techniques for Calculating Magnitude
and Frequency of Floods in Oklahoma, with Compilations of Flood Data Through 1971. U.S.
Geological Survey Water-Resources Investigations Report 73–52. 307 p.

Sauer, V. B., and D. P. Turnipseed. 2010. Stage Measurement at Gaging Stations: U.S. Geological Survey
Techniques and Methods Book 3, Chapter A7.

Saunders, J. W. and M. W. Smith. 1955. Physical Alteration of Stream Habitat to Improve Brook Trout
Production. Canadian Fish Culturist 16:185–188.

Sawyer, A. H., and M. B. Cardenas. 2009. Hyporheic Flow and Residence Time Distributions in
Heterogeneous Cross-Bedded Sediment. Water Resources Research 45:W08406.

Sawyer, A. H., and M. B. Cardenas. 2012. Effect of Experimental Wood Addition on Hyporheic
Exchange and Thermal Dynamics in a Losing Meadow Stream. Water Resources Research
48:W10537.

Sawyer, A. H., M. B. Cardenas, and J. Buttles. 2011. Hyporheic Exchange due to Channel-Spanning
Logs. Water Resources Research 47(8):W08502.

Sawyer, A. H., M. B. Cardenas, and J. Buttles. 2012. Hyporheic Temperature Dynamics and Heat
Exchange Near Channel-Spanning Logs. Water Resources Research 48:W01529.

Schaff, S. D., S. R. Pezeshki, and F. D. Shields Jr. 2002.The Effect of Pre-Planting Soaking on Growth
and Survival of Black Willow (Salix nigra) Cuttings. Restoration Ecology 10(2):267–274.

Scheffer, T. C. 1971. A Climate Index for Estimating Potential for Decay in Wood Structures Above
Ground. Forest Products Journal 21(10):25–31.

Scheiner S. M., and J. Gurevitch (eds.). 2001. Design and Analysis of Ecological Experiments. Oxford
University Press.

Schenk, E. R., J. W. McCargo, B. Moulin, C. R. Hupp, and J. M. Richter. 2014a. The Influence of Logjams
on Largemouth Bass (Micropterus salmoides) Concentrations on the lower Roanoke River, a
Large Sand-Bed River. River Research and Applications 2014(DOI: 10.1002/rra.2779).

Schenk, E. R., B. Moulin, C. R. Hupp, and J. M. Richter. 2014b. Large Wood Budget and Transport
Dynamics on a Large River Using Radio Telemetry. Earth Surface Processes and Landforms
39:487–498.

Scherer, R. 2004. Decomposition and Longevity of In-Stream Woody Debris: A Review of Literature
from North America. Pages 127–133 in Forest Land–Fish Conference–Ecosystem Stewardship
through Collaboration. Proceedings of Forest-Land-Fish Conference II.

Large Wood National Manual July 2015


10-54
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Scheuerell, M. D., R. Hilborn, M. H. Ruckelshaus, K. K. Bartz, K. M. Lagueux, A. D. Haas, and K. Rawson.


2006. The Shiraz Model: A Tool for Incorporating Anthropogenic Effects and Fish–Habitat
Relationships in Conservation Planning. Canadian Journal of Fisheries and Aquatic Sciences
63(7):1596–1607.

Schiff, R., J. S. Clark, G. Alexander, and M. Kline 2008a. The Vermont Agency of Natural Resources
Reach Habitat Assessment (RHA). Prepared by Milone & MacBroom, Inc. with the Vermont
Agency of Natural Resources, Departments of Environmental Conservation and Fish and
Wildlife, Waterbury, VT.

Schiff, R., J. S. Clark, and S. Jaquith 2008b. The Vermont Culvert Geomorphic Compatibility Screening
Tool. Prepared by Milone & MacBroom, Inc. with the VT DEC River Management Program,
Waterbury, VT.

Schiff, R., E. Fitzgerald, J. MacBroom, M. Kline, and S. Jaquith 2014. The Vermont Standard River
Management Principles and Practices (Vermont SRMPP): Guidance for Managing Vermont's Rivers
Based on Channel and Floodplain Function. Prepared by Milone & MacBroom and Fitzgerald
Environmental Associates for and in collaboration with the Vermont Rivers Program,
Montplelier, VT.

Schlosser, I. J. 1991. Stream Fish Ecology: A Landscape Perspective. BioScience 41:704–712.

Schlosser, I. J., and P. L. Angermeier. 1995. Spatial Variation in Demographic Processes of Lotic
Fishes: Conceptual Models, Empirical Evidence, and Implications for Conservation. Pages 392–
401 in J. L. Nielsen and D. A. Powers (eds.), Evolution and the Aquatic Ecosystem: Defining Unique
Units in Population Conservation. Bethesda, MD: American Fisheries Society.

Schmetterling, D. A., and R. W. Pierce. 1999. Success of Instream Habitat Structures After a 50-Year
Flood in Gold Creek, Montana. Restoration Ecology 7(4):369–375.

Schmidt, J. C., and P. R. Wilcock. 2008. Metrics for Assessing the Downstream Effects of Dams. Water
Resources Research 44:W04404. doi:10.1029/2006WR005092.

Scott, M. L., J. M. Friedman, G. T. Auble. 1996. Fluvial Process and the Establishment of Bottomland
Trees. Geomorphology 14:327–339.

Schuett-Hames, D., A. E. Pleus, J. Ward, M. Fox, and J. Light. 1999. TFW Monitoring Program Methods
Manual for the Large Woody Debris Survey. Prepared for the Washington State Dept. of Natural
Resources under the Timber, Fish, and Wildlife Agreement. TFW-AM9-99-004. DNR #106.
March.

Schumm, S. A. 1999. Causes and Controls of Channel Incision. Pages 19–34 in S. E. Darby and
A. Simon (eds.), Incised River Channels. Chichester, UK: Wiley.

Schumm, S. A., M. D. Harvey, and C. C. Watson. 1984. Incised Channels: Morphology, Dynamics and
Control. Water Resources Publication. Littleton, CO.

Scott, S. E., and Y. Zhang. 2012. Contrasting Effects of Sand Burial and Exposure on Invertebrate
Colonization of Leaves. American Midland Naturalist 167:68–78.

Scottish Environment Protection Agency. 2009. Engineering in Water Environment Good Practice
Guide: Temporary Construction Methods. First edition.

Large Wood National Manual July 2015


10-55
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Sear, D. A., C. E. Millington, D. R. Kitts, and R. Jeffries. 2010. Logjam Controls on Channel:Floodplain
Interactions in Wooded Catchments and Their Role in the Formation of Multi-Channel Patterns.
Geomorphology 116:305–319.

Sedell, J. R., and J. L. Frogatt. 1984. Importance of Streamside Forests to Large Rivers: The Isolation
of the Willamette River, Oregon, U.S.A., from its Floodplain by Snagging and Streamside Forest
Removal. Verhandlungen-Internationale Vereinigung für Theorelifche und Angewandte
Limnologie 22:1828–1834.

Sedell, J. R., and K. J. Luchessa. 1981. Using the Historical Record as an Aid to Salmonid Habitat
Enhancement. Symposium on Acquisition and Utilization of Aquatic Habitat Inventory
Information. October 23–28, Portland, OR.

Sedell, J. R., and K. J. Luchessa. 1982. Using the Historical Record as an Aid to Salmonid Habitat
Enhancement. Pages 222–245 in N. B. Armantrout (ed.). Acquisition and Utilization of Aquatic
Habitat Inventory Information. Proceedings of a Symposium October 28–30, 1981. Billings, MT:
The Hague Publishing.

Sedell, J. R., and F. J. Swanson. 1984. Ecological Characteristics of Streams in Old-Growth Forests of
the Pacific Northwest. Pages 9–16 in W. R. Meehan, T. R. Merrell Jr., and T. A. Hanley (eds.), Fish
and Wildlife Relationships in Old-Growth Forests. Juneau, AK: American Institute of Fisheries
Research Biologists.

Sedell, J. R., F. H. Everest, and F. J. Swanson. 1982. Fish Habitat and Streamside Management: Past
and Present. Pages 244–255 in Proceedings of the 1981 Convention of the Society of American
Foresters, September 27–30, 1981. Society of American Foresters, Publication 82–01, Bethesda,
Maryland.

Sedell, J. R., F. J. Swanson, and S. V. Gregory. 1984. Evaluating Fish Response to Woody Debris. Pages
191–221 in T. J. Hassler (ed.). Proceedings of the Pacific Northwest Streams Habitat Management
Workshop. American Fisheries Society. Humboldt State University. Arcata, CA.

Sedell, J. R., J. E. Richey, and F. J. Swanson. 1989. The River Continuum Concept: A Basis for the
Expected Ecosystem Behavior of Very Large Rivers? Canadian Special Publication of Fisheries
and Aquatic Sciences 106:49–55.

Seehorn, M. E. 1985. Fish Habitat Improvement Handbook: Atlanta, Georgia. U.S. Forest Service,
Southern Region. Technical Publication R8-TP-16. 30 pp.

Seitzinger, S., and seven others. 2006. Denitrification across Landscapes and Waterscapes:
A Synthesis. Ecological Applications 16:2064–2090.

Senter, A. E., and G. B. Pasternack. 2010. Large Wood Aids Spawning Chinook Salmon (Oncorhynchus
Tshawytscha) in Marginal Habitat on a Regulated River in California. River Research and
Applications 27:550–565.

Sharma, R., A. B. Cooper, and R. Hilborn. 2005. A Quantitative Framework for the Analysis of Habitat
and Hatchery Practices on Pacific Salmon. Ecological Modeling 18:231–250.

Shields, F. D., Jr. 2007. Scour Calculations. Technical Supplement 14B in Stream Restoration Design.
National Engineering Handbook Part 654. USDA-NRCS. Washington, D.C. CD-ROM.

Large Wood National Manual July 2015


10-56
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Shields, F. D., Jr., and C. V. Alonso. 2012. Assessment of Flow Forces on Large Wood in Rivers. Water
Resources Research 48(4):W04156.

Shields, F. D., Jr., and C .J. Gippel. 1995. Prediction of Effects of Woody Debris Removal on Flow
Resistance. Journal of Hydraulic Engineering 121 (4):341–354.

Shields, F. D., and R. H. Smith. 1992. Effects of Large Woody Debris Removal on Physical
Characteristics of a Sand-Bed River. Aquatic Conservation: Marine and Freshwater Ecosystems
2:145–163.

Shields, F. D., Jr. and A. D. Wood. 2007. The Use of Large Woody Material for Habitat and Bank
Protection. Technical Supplement 14J in Stream Restoration Design, National Engineering
Handbook Part 654. USDA-NRCS Washington, D.C. CD-ROM.

Shields, F. D., Jr., A. J. Bowie, and C. M. Cooper. 1995. Control of Streambank Erosion due to Bed
Degradation with Vegetation and Structure. Water Resources Bulletin 31(3):475–489.

Shields, F. D., Jr., R. Copeland, P. Klingeman, M. Doyle, and A. Simon. 2003. Design for Stream
Restoration. Journal of Hydraulic Engineering 129(8):575–584.

Shields, F. D., Jr., N. Morin, and C. M. Cooper. 2004. Large Woody Debris Structures for Sand-Bed
Channels. Journal of Hydraulic Engineering 130(3):208–217.

Shields, F. D. Jr., S. S. Knight, and J. M. Stofleth. 2006. Large Wood Addition for Aquatic Habitat
Rehabilitation in an Incised, Sand-Bed Stream, Little Topashaw Creek, Mississippi. River
Research and Applications 22:803–817.

Shields, F. D., Jr., S. R. Pezeshki, G. V. Wilson, W. Wu, and S. M. Dabney. 2008. Rehabilitation of an
Incised Stream with Plant Materials: The Dominance of Geomorphic Processes. Ecology and
Society 13 (2):54.

Shirvell, C. S. 1990. Role of Instream Rootwads as Juvenile Coho Salmon (Oncorhynchus kisutch) and
Steelhead Trout (O. mykiss) Cover Habitat Under Varying Streamflows. Canadian Journal of
Fisheries and Aquatic Sciences 47(5):852–861.

Sidle, R. C. 1991. A Conceptual Model of Changes in Root Cohesion in Response to Vegetation


Management. Journal of Environmental Quality 20:43–52.

Sigura, C., and Booth, D. B. 2010. Effects of Geomorphic Setting and Urbanization on Wood, Pools,
Sediment Storage, and Bank Erosion in Puget Sound Streams. JAWRA Journal of the American
Water Resources Association 46(5):972–986.

Simon, A. 1989. A Model of Channel Response in Disturbed Alluvial Channels. Earth Surface
Processes and Landforms 14:11–26.

Simon, A. 1994. Gradation Processes and Channel Evolution in Modified West Tennessee Streams:
Process, Response and Form. U.S. Geological Survey Professional Paper 1470. Washington D.C.

Simon, A., and A. J. C. Collison. 2002. Quantifying the Mechanical and Hydrological Effects of Riparian
Vegetation on Stream-Bank Stability. Earth Surface Processes and Landforms 27(5):527–546.

Simon, A., and M. Rinaldi. 2006. Disturbance, Stream Incision, and Channel Evolution: The Roles of
Excess Transport Capacity and Boundary Materials in Controlling Channel Response.
Geomorphology 79:361–383.

Large Wood National Manual July 2015


10-57
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Simon, A., A. Curini, S. E. Darby, and E. J. Langendoen. 2000. Bank and Near-Bank Processes in an
Incised Channel. Geomorphology 35(3):193–217.

Simon, A., A. Brooks, and N. Bankhead. 2012. Effectiveness of Engineered Log Jams in Reducing
Streambank Erosion to the Great Barrier Reef: The O’Connell River, Queensland, Australia. Pages
2570–2577 in World Environmental and Water Resources Congress 2012: Crossing Boundaries.
Reston, VA: ASCE.

Simon, A., R. Thomas, A. Curini, and N. Bankhead. 2014. Development of the Bank-Stability and
Toe-Erosion Model (BSTEM version 5.4). Available: www.kwo.org/reports_publications/
Presentations/pp_Development_of_BSTEM_012811_sm.pdf.

Simon, T. P., and J. Lyons. 1995. Application of the Index of Biotic Integrity to Evaluate Water
Resource Integrity in Freshwater Ecosystems. Chapter 16 in W. S. Davis and T. P. Simon.
Bioassessment and Criteria: Tools for Water Resources Planning and Decision Making. CRC Press.

Simpson, W. and A. TenWolde. 1999. Physical Properties and Moisture Relations of Wood. Chapter
3 in Wood Handbook: Wood as an Engineering Material. Report FPL-GTR-113. U.S. Department of
Agriculture Forest Service. Forest Products Laboratory. Madison, WI.

Singer, S., and M. L. Swanson. 1983. The Soquel Creek Storm Damage Recovery Plan with
Recommendations for Reduction of Geologic Hazards in Soquel Village, Santa Cruz County,
California. Unpublished USDA Soil Conservation Service report to the Santa Cruz County Board
of Supervisors.

Sinsabaugh, R. L., M. P. Osgood, and S. Findlay. 1994. Enzymatic Models for Estimating
Decomposition Rates of Particulate Detritus. Journal of the North American Benthological Society.
13:160–169.

Skidmore, P. B., C. R. Thorne, B. L. Cluer, G. R. Pess, J. M. Castro, T. J. Beechie, and C. C. Shea. 2011.
Science Base and Tools for Evaluating Stream Engineering, Management, and Restoration
Proposals. U.S. Department of Commerce. NOAA Tech. Memo. NMFS-NWFSC-112.

Sklar, L. S., J. Fadde, J. G. Venditti, P. Nelson, M. A. Wydzga, Y. Cui, and W. E. Dietrich. 2009.
Translation and Dispersion of Sediment Pulses in Flume Experiments Simulating Gravel
Augmentation Below Dams. Water Resources Research 45:W08439, doi:10.1029/
2008WR007346.

Sleeter, B., T. Wilson, W. Acevedo, W. 2012. Status and Trends of Land Change in the Western United
States—1973–2000. U.S. Geological Survey Professional Paper 1794-A.

Smith, D. G. 1979. Effects of Channel Enlargement by River Ice Processes on Bankfull Discharge in
Alberta, Canada. Water Resources Research, 15(2):469–475.

Smith, D. G., and C.M. Pearce. 2000. River Ice and its Role in Limiting Woodland Development on a
Sandy Braid-Plain, Milk River, Montana. Wetlands, 20(2):232–250.

Smith, D. G., and D. M. Reynolds. 1983. Tree Scars to Determine the Frequency and Stage of High
Magnitude River Ice Drives and Jams, Red Deer, Alberta. Canadian Water Resources Journal
8(3):77–94.

Large Wood National Manual July 2015


10-58
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Smith, D. L., J. B. Allen, O. Eslinger, M. Valenciano, J. Nestler, and R. A. Goodwin. 2011. Hydraulic
Modeling of Large Roughness Elements with Computational Fluid Dynamics for Improved
Realism in Stream Restoration Planning. Geophysical Monograph Series 194:115–122.

Smith, J. D.. 2004. The Role of Riparian Shrubs in Preventing Floodplain Unraveling Along the Clark
Fork of the Columbia River in the Deer Lodge Valley, Montana. Pages 71–85 in S. J. Bennett. and
A. Simon (eds.), Riparian Vegetation and Fluvial Geomorphology, Water Science and Application 8.
American Geophysical Union, Washington, D.C.

Smith, M. P., R. Schiff, A. Olivero, and J. G. MacBroom 2008. The Active River Area: A Conservation
Framework to Protect Rivers and Streams. Boston, MA: The Nature Conservancy.

Smith, R. D., R. C. Sidle, and P. E. Porter. 1993. Effects on Bedload Transport of Experimental
Removal of Woody Debris from a Forest Gravel-Bed Stream. Earth Surface Processes and
Landforms 18:455–468.

Smock, L. A., G. M. Metzler and J. E. Gladden. 1989. Role of Debris Dams in the Structure and
Functioning of Low Gradient Headwater Streams. Ecology 70:764–775.

Smokorowski, K. E., and T. C. Pratt. 2007. Effect of a Change in Physical Structure and Cover on Fish
and Fish Habitat in Freshwater Ecosystems - A Review and Meta-Analysis. Environmental
Reviews 15:15–41.

Sobczak, W. V., and S. Findlay. 2002. Variation in Bioavailability of Dissolved Organic Carbon among
Stream Hyporheic Flowpaths. Ecology 83:3194–3209.

Sobota, D. J., S. V. Gregory, S. V., and S. L. Johnson. 2007. Influence of Wood Decomposition on
Nitrogen Dynamics in Stream Ecosystems: Interactive Effects of Substrate Quality and Nitrogen
Loading. North American Benthological Society 55th Annual Meeting. Available:
https://nabs.confex.com/nabs/2007/techprogram/P1365.HTM.

Society for Ecological Restoration. 2002. SER International Primer on Ecological Restoration. Science
& Policy Working Group, Version 2, October. Available: http://www.ser.org/resources/
resources-detail-view/ser-international-primer-on-ecological-restoration.

Society for Ecological Restoration. 2004. The SER International Primer on Ecological Restoration.
Available: http://www.ser.org.

Solazzi, M. F., T. E. Nickelson, S. L. Johnson, and J. D. Rodgers. 2000. Effects of Increasing Winter
Rearing Habitat on Abundance of Salmonids in Two Coastal Oregon Streams. Canadian Journal of
Fisheries and Aquatic Sciences 57:906–914.

Sollins, P., S. P. Cline, T. Verhoeven, D. Sachs, and G. Spycher. 1987. Patterns of Log Decay in
Old-Growth Douglas-Fir Forests, Canadian Journal of Forest Research 17:1585–1595.

Southerland, B. S., and F. Reckendorf. 2010. Performance of Engineered Log Jams in Washington
State—Post Project Appraisal. In Joint Federal Interagency Conferences 2010: Book of Abstracts.
Joint Fed. Interagency Conf., 2010, Conference on Sedimentation and Hydrologic Modeling, June
27–July 1, Las Vegas, Nev., Government Printing Office, Washington, D. C.

Southwide Safety Committee. 2010. Timber Harvesting Safety Manual. Rockville, MD. National
Timber Harvesting and Transportation Safety Foundation. Available: http://loggingsafety.com/
content/timber-harvesting-safety-manual. Accessed: July 10, 2014.

Large Wood National Manual July 2015


10-59
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Southwood, T. R. E. 1977. Habitat, the Template for Ecological Strategies? Journal of Animal Ecology
46:337–365.

Spänhoff, B., and E. Cleven. 2010. Wood in Different Stream Types: Epixylic Biofilm and
Wood-Inhabiting Invertebrates in a Lowland Versus an Upland Stream. Annales De Limnologie-
International Journal of Limnology 46(3):169–179.

Spänhoff, B., and E. I. Meyer. 2004. Breakdown Rates of Wood in Streams. Journal of the North
American Benthological Society 23(2):189–197.

Spänhoff, B., C. Alecke, and E. Irmgard Meyer. 2001. Simple Method for Rating the Decay Stages of
Submerged Woody Debris. Journal of the North American Benthological Society 20(3):385–394.

Spies, T. A., and J. F. Franklin. 1991. The Structure of Natural Young, Mature, and Old-Growth
Douglas Fir Forests in Oregon and Washington. Pages 91–109 in L. F. Ruggiero, K. B. Aubrey,
A. B. Carey, and M. H. Huff (technical coordinators), Wildlife and Vegetation of Unmanaged
Douglas Fir Forests. USDA Forest Service. General Technical Report PNW-GTR-285.

Spies, T. A., J. F. Franklin, and T. B. Thomas. 1988. Coarse Woody Debris in Douglas-Fir Forests of
Western Washington and Oregon. Ecology 69:1689–1702.

Spooner, D. E., M. A. Xenopoulos, C. Schneider, and D. A. Woolnough. 2011. Coextirpation of


Host-Affiliate Relationships in Rivers: The role of Climate Change, Water Withdrawal, and
Host-Specificity. Global Change Biology 17:1720–1732. doi:10.1111/j.1365-2486.2010.02372.x.

Stahle, D. W., M. K. Cleaveland, R. D. Griffin, M. D. Spond, F. K. Fye, R. B. Culpepper, and D. Patton.


2006. Decadal Drought Effects on Endangered Woodpecker Habitat. Eos, Transactions American
Geophysical Union 87(12):121–125.

Stanford, J. A., and G. C. Poole. 1996. A Protocol for Ecosystem Management. Ecological
Applications:741–744.

Stanford, J. A. and J. V. Ward. 1988. The Hyporheic Habitat of River Ecosystems. Nature 335:64–66.

Stanford, J. A., and J. V. Ward. 1993. An Ecosystem Perspective of Alluvial Rivers: Connectivity and
the Hyporheic Corridor. Journal of the North American Benthological Society 12(1):48–60.

Stanford, J. A., and J. V. Ward. 1995. The Serial Discontinuity Concept: Extending the Model to
Floodplain Rivers. Regulated Rivers: Research and Management:159–168.

Statzner, B., and B. Higler. 1985. Questions and Comments on the River Continuum Concept.
Canadian Journal of Fisheries and Aquatic Science 42:1038–1044.

Steinhart, G. S., G. E. Likens, and P. M. Groffman. 2000. Denitrification in Stream Sediments in Five
Northeastern (USA) Streams. Verhandlungen Internationale Vereinigung für Theorertische und
Angewandte Limnologie 27:1331–1336.

Stelzer, R. S., L. A. Bartsch, W. B. Richardson, and E. A. Strauss. 2011. The Dark Side of the Hyporheic
Zone: Depth Profiles of Nitrogen and its Processing in Stream Sediments. Freshwater Biology
56:2021–2033.

Stewart, G. B., H. R. Bayliss, D. A. Showler, W. J. Sutherland, and A. S. Pullin. 2009. Effectiveness of


Engineered In-Stream Structure Mitigation Measures to increase Salmonid Abundance:
A Systematic Review. Ecological Applications 19(4):931–941.

Large Wood National Manual July 2015


10-60
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Stewart, P. M., S. Bhattarai, M. W. Mullen, C. K. Metcalf, and E. G. Reategui-Zirena. 2012.


Characterization of Large Wood and its Relationship to Pool Formation and Macroinvertebrate
Metrics in Southeastern Coastal Plain Streams, USA. Journal of Freshwater Ecology 27(3):351–
365.

Stewart, T. L., and J. F. Martin. 2005. Energy Model to Predict Suspended Load Deposition Induced by
Woody Debris: Case Study. Journal of Hydraulic Engineering-ASCE 131:1011–1016.

Stewart-Oaten, A., J. R. Bence, and C. W. Osenberg. 1992. Assessing Effects of Unreplicated


Perturbations: No Simple Solutions. Ecology:1396–1404.

Stiehl, R. B. 1998. Habitat Evaluation Procedures Workbook. Fort Collins, CO: U.S. Geological Survey.

Stock, J. D., D. R. Montgomery, B. D. Collins, W. E. Dietrich, and L. Sklar. 2005. Field Measurements of
Incision Rates Following Bedrock Exposure: Implications for Process Controls on the Long
Profiles of Valleys Cut by Rivers and Debris Flows. Geological Society of America Bulletin
117(11/12):174–194.

Stockner, J. G. (ed.). 2003. Nutrients in Salmonid Ecosystems: Sustaining Production and Biodiversity.
Bethesda, MD: American Fisheries Society.

Stofleth, J. M., F. D. Sheilds, Jr., and G. A. Fox. 2004 Organic Carbon Concentrations in Hyporheic Zone
Sediments: A Tool for Measuring Stream Integrity. In Proceedings of the 2004 World Water and
Environmental Resources Congress, G. Shhlke, D. F. Hayes, and D. K. Stevens (eds.). Critical
Transitions in Water and Environmental Resources, American Society of Civil Engineers: CD ROM.

Stofleth, J. M., F. D. Shields, Jr., and G. A. Fox. 2007. Hyporheic and Total Transient Storage in Small,
Sand-Bed Streams. Hydrological Processes 22:1885–1894.

Subramanya, K., 2008. Engineering Hydrology. New York: McGraw-Hill.

Suding, K. N., and R. J. Hobbs. 2009. Threshold Models in Restoration and Conservation:
A Developing Framework. Trends in Ecology & Evolution 24 (5):271–279.

Sullivan, K. J., J. Tooley, K. Doughty, J. E. Caldwell, and P.A. Knudsen. 1990. Evaluation of Prediction
Models and Characterization of Stream Temperature Regimes in Washington. TFW-WQ3-90-006,
Timber Fish & Wildlife, Department of Natural Resources, Olympia, WA.

Sundbaum, K. and I. Naslund. 1998. Effects of Woody Debris on the Growth and Behavior of Brown
Trout in Experimental Stream Channels. Canadian Journal of Zoology 76:56–61.

Svoboda, C. D. and K. Russell, K. 2011. Flume Analysis of Engineered Large Wood Structures for
Scour Development and Habitat. Pages 2572–2581 in Proceedings, World Environmental and
Water Resources Congress, ASCE, Reston, VA.

Swanson, F. J., S. V. Gregory, J. R. Sedell, and A. G. Campbell. 1982. Land-Water Interactions: The
Riparian Zone. Pages 267–291 on R. L. Edmonds (ed.), Analysis of Coniferous Forest Ecosystems in
the Western United States. US/IBP Synthesis Series, Hutchinson Ross Publishing Company:
Stroudsburg, PA.

Swanson, F. J., T. K. Kranz, N. Caine, and R. G. Woodmansee. 1988. Landform Effects on Ecosystem
Patterns and Processes. BioScience 38:92–98.

Large Wood National Manual July 2015


10-61
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Sylte, T., and C. Fischenich. 2000. Rootwad Composites for Streambank Erosion Control and Fish
Habitat Enhancement. U.S. Army Corps of Engineers. Vicksburg, MS.

Tappeiner, J. C., D. Huffman, D. Marshall, T. A. Spies, and J. D. Bailey. 1997. Density, Ages, and Growth
Rates in Old-Growth and Young-Growth Forests in Coastal Oregon. Paper 3166 of the Forest
Research Laboratory, Oregon State University, Corvallis.

Tarzwell, C. M. 1934. The Purpose and Value of Stream Improvement Method. Stream Improvement.
Bulletin R-4. Presented at the Annual Meeting of the American Fisheries Society. Ogden, UT.

Tarzwell, C. M. 1936. Experimental Evidence of the Value of Trout Stream Improvements.


Transactions of the American Fisheries Society 66:177–187.

Teels, B. M., and T. Danielson. 2001. Using a Regional IBI to Characterize Condition of Northern
Virginia Streams, with Emphasis on the Occoquan Watershed. USDANRCS. Technical Note
190-13-1. December 2001

Thomas, H., and T. R. Nisbet. 2006. An Assessment of the Impact of Flood Plain Woodland on Flood
Flows. Water and the Environment Journal 21(2):114–126.

Thompson, D. M. 1995. The Effects of Large Organic Debris on Sediment Processes and Stream
Morphology in Vermont. Geomorphology 11(3):235–244.

Thompson, D. M. 2002. Channel-bed Scour with High Versus Low Deflectors. Journal of Hydraulic
Engineering 128(6):640–643.

Thompson, D. M. 2002. Long-term Effect of Instream Habitat-improvement Structures on Channel


Morphology along the Blackledge and Salmon Rivers, Connecticut, USA. Environmental
Management 29(1):250–265.

Thompson, D. M. 2005. The History of the Use and Effectiveness of Instream Structures in the United
States. Geological Society of America Reviews in Engineering Geology XVI:35–50.

Thompson, D. M. 2006. Did the Pre-1980 Use of In-Stream Structures Improve Streams?
A Reanalysis of Historical Data. Ecological Applications 16(2):784–796.

Thompson, D. M., and Stull, G. N. 2002. The Development and Historic Use of Habitat Structures in
Channel Restoration in the United States: The Grand Experiment in Fisheries Management.
Géographie physique et Quaternaire 56(1):45–60.

Thorne, S. D., and D. J. Furbish. 1995. Influences of Coarse Bank Roughness on Flow Within a Sharply
Curved River Bend. Geomorphology 12(3):241–257.

Thorne, C., J. Townsend, and T. Ashley. 2014a. Geomorphic and Ecological Assessment and Evaluation
of Grade Building Structures on the SRS Sediment Plain, North Fork Toutle River Final Report.
Performed for the U.S. Army Corps of Engineers, Portland District, OR.

Thorne, C., J. Castro, B. Cluer, P. Skidmore, and C. Shea. 2014b. Project Risk Screening Matrix for
River Management and Rrestoration. River Research and Applications, April 2014, DOI:
10.1002/rra.2753.

Tillman, D. C., A. H. Moerke, C. L. Ziehl, and G. A. Lamberti. 2003. Subsurface Hydrology and Degree of
Burial Affect Mass Loss and Invertebrate Colonization of Leaves in a Woodland Stream.
Freshwater Biology 48:98–107.

Large Wood National Manual July 2015


10-62
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Tonglao, P., and D. Eckberg. 2012. FAQ’s about Wood Placements in Rivers. March Bulletin, Skellenger
Bender Attorneys, Seattle, Washington. Available: http://www.hallandcompany.com/
php_uploads/resources/library/2012%20March%20Bulletin%20-%20FAQ%27s%20About%
20Wood%20Placements%20in%20Rivers%20%283%29%202012.pdf .

Tonina, D., and J. M. Buffington. 2009. Hyporheic Exchange In Mountain Rivers I: Mechanics and
Environmental Effects. Geography Compass 3:1063–1086.

Torgersen, C. E., J. L. Ebersole, and D. M. Keenan. 2012. Primer for Identifying Cold-Water Refuges to
Protect and Restore Thermal Diversity in Riverine Landscapes. U.S. EPA 910-C-12-001.

Townsend, C. R. 1989. The Patch Dynamics Concept of Stream Community Ecology. Journal of the
North American Benthological Society 8(1):36–50.

Treadwell, S., J. Koehn, S. Bunn, and A. Brooks. 2007. Wood and Other Aquatic Habitat. Chapter 7 in
S. Lovett and P. Price (eds.), Principles for Riparian Lands Management. Land and Water
Australia, Canberra.

Trinity River Restoration Program. 2015. Main Web Page. Available: http://www.trrp.net/.
Accessed: February 28, 2015.

Triska, F. J. 1984. Role of Large Wood in Modifying Channel Morphology and Riparian Areas of
a Large Lowland River under Pristine Conditions: A Historical Case Study. Verhandlungen-
Internationale Vereinigung für Theorelifche und Angewandte Limnologie 22:1876–1892.

Triska, F. J., and K. Cromack, Jr.. 1979. The Role of Wood Debris in Forests and Streams. In
R. H. Waring, Forests: Fresh Perspectives from Ecosystem Analysis. Pages 171–190 in
Proceedings of the 40th Annual Biology Colloquium. Corvallis, OR: Oregon State University Press.
Corvallis, OR.

Trotter, E. H. 1990. Woody Debris, Forest-Stream Succession, and Catchment Geomorphology.


Journal of the North American Benthological Society 9(2):141–156.

Tsukamoto, Y. 1987. Evaluation of the Effect of Tree Roots on Slope Stability. Bulletin of the
Experimental Forests. 23:65–124.

Tufekcioglu, A., J. W. Raich, T. M. Isenhart, and R. C. Schultz. 2003. Biomass, Carbon and Nitrogen
Dynamics of Multi-Species Riparian Buffers within an Agricultural Watershed in Iowa, USA.
Agroforestry Systems 57(3):187–198.

Tullos, D., and C. Walter. 2014. Fish Use of Turbulence Around Wood in Winter: Physical
Experiments on Hydraulic Variability and Habitat Selection by Juvenile Coho Salmon,
Oncorhynchus kisutch. Environmental Biology of Fishes:1–15.

Turnipseed, D. P., and V. B. Sauer. 2010. Discharge Measurements at Gaging Stations: U.S. Geological
Survey Techniques and Methods Book 3, Chapter A8, U.S. Geological Survey.

Turowski, J. M., A. Badoux, K. Bunte, C. Rickii, N. Federspiel, and M. Jochner. 2013. The Mass
Distribution of Coarse Particulate Matter from an Alpine Headwater Stream. Earth Surface
Dynamics 1:1–14.

Tyler, R. N. 2011. River Debris: Causes, Impacts, and Mitigation Techniques. Prepared for Ocean
Renewable Power Company by the Alaska Center for Energy and Power, Fairbanks, Alaska.

Large Wood National Manual July 2015


10-63
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Umazano, A.M., R.N. Melchor, E. Bedatou, E.S. Bellosi, and J.M. Krause. 2014. Fluvial Response to
Sudden Input of Pyroclastic Sediments During the 2008–2009 Eruption of the Chaitén Volcano
(Chile): The Role of Logjams. Journal of South American Earth Sciences 54:140–157.

U.S. Army Corps of Engineers, 1981. The Streambank Erosion Control Evaluation and Demonstration
Act of 1974. Final Report to Congress, Main Report. Washington, D.C.

U. S. Army Corps of Engineers. 1992. Engineering and Design: Bearing Capacity of Soils. EM 1110-1-
1905. Department of the Army, U.S. Army Corps of Engineers. Washington, D.C.

U. S. Army Corps of Engineers. 1994. Engineering and Design: Hydraulic Design of Flood Control
Channels. EM 1110-2-1601. Department of the Army, U.S. Army Corps of Engineers. Washington,
D.C.

U. S. Army Corps of Engineers. 2005. Engineering and Design: Stability Analysis of Concrete
Structures. EM 1110-2-2100. Department of the Army, U.S. Army Corps of Engineers.
Washington, D.C.

U.S. Army Corps of Engineers. 2008. Safety and Health Requirements. Engineer Manual 385-1-1.
U.S. Army Corps of Engineers Headquarters, Washington, D.C

U.S. Army Corp of Engineers Institute for Water Resources (USACE IWR). 2010. IWR Planning Suite
MCDA Module User’s Guide. U.S. Army Corp of Engineers Institute for Water Resources.

U.S. Bureau of Reclamation. 2005. Watershed Conditions and Seasonal Variability for Select Streams
within WRIA 20, Olympic Peninsula, Washington. Available: http://www.ecy.wa.gov/
programs/eap/wrias/planning/docs/opendraft_wria20_final4.pdf.

U.S. Climate Change Science Program (CCSP). 2008a. Preliminary Review of Adaptation Options for
Climate-Sensitive Ecosystems and Resources. A Report by the U.S. Climate Change Science
Program and the Subcommittee on Global Change Research. (S. H. Julius and J.M. West [eds.], J. S.
Baron, B. Griffith, L. A. Joyce, P. Kareiva, B. D. Keller, M. A. Palmer, C. H. Peterson, and J. M. Scott
[Authors]). U.S. Environmental Protection Agency. Washington, D.C.

U.S. Climate Change Science Program (CCSP ). 2008b. The Effects of Climate Change on Agriculture,
Land Resources, Water Resources, and Biodiversity in the United States. A Report by the U.S.
Climate Change Science Program and the Subcommittee on Global Change Research
(P. Backlund, A. Janetos, D. Schimel, J. Hatfield, K. Boote, P. Fay, L. Hahn, C. Izaurralde,
B. A. Kimball, T. Mader, J. Morgan, D. Ort, W. Polley, A. Thomson, D. Wolfe, M. G. Ryan,
S. R. Archer, R. Birdsey, C. Dahm, L. Heath, J. Hicke, D. Hollinger, T. Huxman, G. Okin, R. Oren,
J. Randerson, W. Schlesinger, D. Lettenmaier, D. Major, L. Poff, S. Running, L. Hansen, D. Inouye,
B. P. Kelly, L. Meyerson, B. Peterson, and R. Shaw). U.S. Department of Agriculture. Washington,
D.C.

U.S. Climate Change Science Program (CCSP). 2008c. Impacts of Climate Change and Variability on
Transportation Systems and Infrastructure: Gulf Coast Study, Phase I. A Report by the U.S. Climate
Change Science Program and the Subcommittee on Global Change Research (M. J. Savonis,
V. R. Burkett, and J. R. Potter [eds.]). U.S. Department of Transportation. Washington, D.C.

U.S. Department of Agriculture (USDA). 1980. Ecoregions of the United States. U.S. Forest Service,
Washington, D.C. Miscellaneous Publication No. 1391

Large Wood National Manual July 2015


10-64
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

U.S. Department of Agriculture (USDA), Agricultural Research Service. 2013. Bank Stability and
Erosion Model. Available: http://www.ars.usda.gov/Research/docs.htm?docid=5044&page=1.

U.S. Department of Health and Human Services. 2002. Toxicological Profile for Wood Creosote, Coal
Tar Creosote, Coal Tar, Coal Tar Pitch, and Coal Tar Pitch Volatiles. September.

U.S. Department of the Interior – Bureau of Reclamation. 2011. West-Wide Climate Risk Assessments:
Bias-Corrected and Spatially Downscaled Surface Water Projections. Technical Memorandum
No. 86-68210–2011-01

U.S. Environmental Protection Agency (EPA). 1995. A Decision-Making Guide for Restoration in
Ecological Restoration. EPA 841-F-95-007 (November)

U.S. Environmental Protection Agency (EPA). 2000. Principles for the Ecological Restoration of
Aquatic Resources. EPA841-F-00-003. Available: http://www.epa.gov/owow/wetlands/
restore/.

U.S. Environmental Protection Agency (EPA). 2008. Handbook for Developing Watershed Plans to
Restore and Protect our Water.

U.S. Environmental Protection Agency. 2009. Valuing the Protection of Ecological Systems and
Services. May. Available: http://yosemite.epa.gov/sab%5CSABPRODUCT.NSF/
F3DB1F5C6EF90EE1852575C500589157/$File/EPA-SAB-09-012-unsigned.pdf. Accessed:
October 8, 2014

U.S. Environmental Protection Agency. 2011. Aquatic Indicators. Available: http://www.epa.gov/


nheerl/arm/indicators/indicators.htm.

U.S. Environmental Protection Agency (EPA). 2013. A Quick Guide to Developing Watershed Plans to
Restore and Protect Our Waters.

U.S. Environmental Protection Agency (EPA). 2014. Green Infrastructure. Available:


http://water.epa.gov/infrastructure/greeninfrastructure/index.cfm.

U.S. Environmental Protection Agency, Office of Water. 2013. Climate Change Adaptation
Implementation Plan. Available: http://epa.gov/climatechange/Downloads/impacts-
adaptation/office-of-water-plan.pdf.

U.S. Forest Service (USFS). 2008. Stream Simulation: An Ecological Approach to Providing Passage for
Aquatic Organisms at Road-Stream Crossings. Forest Service Stream-Simulation Working Group.
San Dimas, CA. May. Available: http://www.stream.fs.fed.us/fishxing/
publications/PDFs/AOP_PDFs/08771801.pdf. Accessed: February 27, 2015.

U.S. Fish and Wildlife Service (USFWS). 2008. SDM Fact Sheet. Available:
http://www.fws.gov/science/doc/structured_decision_making_factsheet.pdf. Accessed: May 15,
2015.

U.S. Global Change Research Program (USGCRP). 2009. Global Climate Change Impacts in the United
States. Edited by T. R. Karl, J. M. Melillo, and T. C. Peterson. Cambridge, MA: Cambridge
University Press.

University of New Hampshire (UNH). 2009. New Hampshire Stream Crossing Guidelines. University of
New Hampshire, Durham, NH.

Large Wood National Manual July 2015


10-65
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Vail, L. W., and R. L. Skaggs. 2002. Adaptive Management Platform for Natural Resources in the
Columbia River Basin. Pacific Northwest National Laboratory. Available:
http://www.pnl.gov/main/publications/external/technical_reports/PNNL-13875.pdf.

Valett, H. M., C. L. Crenshaw, and P. F. Wagner. 2002. Stream Nutrient Uptake, Forest Succession, and
Biogeochemical Theory. Ecology 83:2888–2901.

Valverde, R. S. 2013. Roughness and Geometry Effects of Engineered Log Jams on 1-D Flow
Characteristics. M. S. Thesis, Civil Engineering, Oregon State University, Corvallis.

Van Cleef, J. S. 1885. How to Restore Our Trout Streams. Transactions of the American Fisheries
Society 14:50–55.

Van Horne, B. 1983. Density as a Misleading Indicator of Habitat Quality. Journal of Wildlife
Management 47:893–901.

Vannote, R. L., G. W. Minshall, K. W. Cummins, J. R. Sedell, and C. E. Cushing. 1980. The River
Continuum Concept. Canadian Journal of Fisheries and Aquatic Science 37(1):130–137.

Vanoni, V. 1975. Sedimentation Engineering, ASCE Manuals and Reports on Engineering Practice—
No. 54. American Society of Civil Engineers, New York, NY.

Van Sickle, J., and S. V. Gregory. 1990. Modeling Inputs of Large Woody Debris to Streams from
Falling Trees. Canadian Journal of Forest Research 20(10):1593–1601.

Van Wilgen, B. W., and H. C. Biggs. 2011. A Critical Assessment of Adaptive Ecosystem Management
in a Large Savanna Protected Area in South Africa. Biological Conservation 144(4):1179–1187.

Veatch, A. C. 1906. Geology and Underground Water Resources of Northern Louisiana and Southern
Arkansas. Washington D.C. United States Geological Survey Professional Paper 46.

Veilleux, A. G., T. A. Cohn, K. M. Flynn, R. R. Mason, and P. R. Hummel. 2013. Fact Sheet 2013-3108:
Estimating Magnitude and Frequency of Floods Using the PeakFQ 7.0 Program. 2327-6932,
U.S. Geological Survey.

Vermont Agency of Natural Resources (VTANR). 2014. Vermont Stream Alteration General Permit.
Department of Environmental Conservation, Montpelier, VT.

Vermont Agency of Transportation (Vtrans). 2001. Hydraulics Manual. Montpelier, VT.

Vidon, P., and A. R. Hill. 2004. Denitrification and Patterns of Electron Donors and Acceptors in Eight
Riparian Zones with Contrasting Hydrogeology. Biogeochemistry 71:259–283.

Vidon, P., and nine others. 2010. Hot Spots and Hot Moments in Riparian Zones: Potential for
Improved Water Quality Management. Journal of the American Water Resources Association. DOI:
101111/j.1752-1688.2010.00420.x.

Viessman, W. J., and G. L. Lewis. 2003. Introduction to Hydrology. Prentice Hall.

Volkman, J. M., and W. E. McConnaha. 1993. Through a Glass, Darkly: Columbia River Salmon, the
Endangered Species Act, and Adaptive Management. Environmental Law 23:1249–1272.

Wadsworth, A. H., Jr. 1966. Historical Deltation of the Colorado River, Texas. Pages 99–105 in Deltas
in Their Geologic Framework. American Association of Petroleum Geologists.

Large Wood National Manual July 2015


10-66
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Wallace, J. B., and A .C. Benke. 1984. Quantification of Wood Habitat in Subtropical Coastal Plain
Streams. Canadian Journal of Fisheries and Aquatic Sciences 41:1643–1652.

Wallace, J. B., J. R. Webster, and J. L. Meyer. 1995a. Influence of Log Additions on Physical and Biotic
Characteristics of a Mountain Stream. Canadian Journal of Fisheries and Aquatic Sciences
52:2120–2137.

Wallace, J. B., M. R. Whiles, S. Eggert, T. F. Cuffney, G. H. Lugthart, and K. Chung. 1995b. Long-Term
Dynamics of Coarse Particulate Organic-Matter in 3 Appalachian Mountain Streams. Journal of
the North American Benthological Society 14(2):217–232.

Wallace, J. B., S. L. Eggert, J. L. Meyer, and J. R. Webster. 1997. Multiple Trophic Levels of a Forest
Stream Linked to Terrestrial Litter Inputs. Science 277:102–104.

Wallerstein, N. P., and C. R. Thorne. 2004. Influence of Large Woody Debris on Morphological
Evolution of Incised, Sand-Bed Channels. Geomorphology 57:53–73.

Wallerstein, N., C. R. Thorne, and M. W. Doyle. 1997. Spatial Distribution and Impact of Large Woody
Debris in Northern Mississippi. Pages 145–150 in C. C. Wang, E. J. Langendoen, and F. D. Shields
(eds.), Proceedings of the Conference on Management of Landscapes Disturbed by Channel
Incision. University of Mississippi. Oxford, MI.

Wallerstein, N. P., C. V. Alonso, S. J. Bennett, and C. R. Thorne. 2001. Distorted Froude-Scaled Flume
Analysis of Large Woody Debris. Earth Surface Processes and Landforms 26:1265–1283.

Walsh, C., and A. Roy. 2005. The Urban Stream Syndrome: Current Knowledge and the Search for a
Cure. Journal of the North American Benthological Society 24:706–723.

Walter, R. C., and D. J. Merritts. 2008. Natural Streams and the Legacy of Water-Powered Mills.
Science 319(5861):299–304.

Walters, C. 2002. Adaptive Management of Renewable Resources. The Blackburn Press.

Ward, J. V., K. Tockner, and F. Schiemer. 1999. Biodiversity of Floodplain River Ecosystems: Ecotones
and Connectivity. Regulated Rivers Research and Management 15:125–139.

Ward, J. V., K. Tockner, D. B. Arscott, and C. Claret. 2002. Riverine Landscape Diversity. Freshwater
Biology 47:517–539.

Warner, M. D., C. F. Mass, E. P. Salathé Jr. 2012. Wintertime Extreme Precipitation Events along the
Pacific Northwest Coast: Climatology and Synoptic Evolution. Monthly Weather Review,
140(7):2021–2043.

Warren, D. R., and C. E. Kraft. 2003. Brook Trout (Salvelinus fontinalis) Response to Wood Removal
from High-Gradient Streams of the Adirondack Mountains (NY, USA). Canadian Journal of
Fisheries and Aquatic Sciences 60(4):379-389.

Warren, D. R., and C. E. Kraft. 2008. Dynamics of Large Wood in an Eastern US Mountain Stream.
Forest Ecology and Management 256(4):808–814.

Warren, D. R., E. S. Bernhardt, R. O. Hall Jr., and G. E. Likens. 2007. Forest Age, Wood and Nutrient
Dynamics in Headwater Streams of the Hubbard Brook Experimental Forest. N.H. Earth Surface
Processes & Landforms 32(8):1154–1163.

Large Wood National Manual July 2015


10-67
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Warren, D. R., C. E. Kraft, W. S. Keeton, J. S. Nunery, and G. E. Likens. 2009. Dynamics of Wood
Recruitment in Streams of the Northeastern US. Forest Ecology and Management 258:804–813.

Warren, D. R., J. D. Dunham, and D. Hockman-Wert. 2014. Geographic Variability in Elevation and
Topographic Constraints on the Distribution of Native and Nonnative Trout in the Great Basin.
Transactions of the American Fisheries Society 143:205–218.

Washington Department of Fish and Wildlife (WDFW). 2012. Stream Habitat Restoration Guidelines.
Washington Department of Fish and Wildlife, Olympia, Washington, 2012.

Washington Department of Transportation. 2012. WSDOT Fish Exclusion Protocols and Standards.
Available: http://www.wsdot.wa.gov/Environment/Biology/BA/BAtemplates.htm. Washington
DOT, Olympia.

Washington State Legislature. Undated. Safety Standards—Logging Operations. Chapter 296-54


WAC. Available: http://apps.leg.wa.gov/WAC/default.aspx?cite=296-54. Accessed: July 10,
2014. Washington State Legislature. Olympia, Washington.

Watts, R. J., B. D. Richter, J. J. Opperman, and K. H. Bowmer. 2011. Dam Reoperation in an Era of
Climate Change. Marine and Freshwater Research 62:321–327.

Webb, A. A., and W. D. Erskine. 2003. Distribution, Recruitment, and Geomorphic Significance of
Large Woody Debris in an Alluvial Forest Stream: Tonghi Creek, Southeastern Australia.
Geomorphology 51:109–126.

Webster, J. R., and E. F. Benfield. 1986. Vascular Plant Breakdown in Freshwater Ecosystems. Annual
Review of Ecology and Systematics 17(1):567–594.

Webster, J. R., J. L. Tank, J. B. Wallace, J. L. Meyer, S. L. Eggert, T. P. Ehrman, B. R. Ward, B. L. Bennet,


P. F. Wagner, and M. E. McTammy. 2000. Effects of Litter Exclusion and Wood Removal on
Phosphorus and Nitrogen Retention in A Forest Stream. Verhandlungen der Internationale
Vereinigung für Limnologie 27:1337–1340.

Webster, J. R., J. A. Stanford, J. L. Chaffin, and Field Ecology Class. 2002. Large Wood Jam in a Fourth
Order Rocky Mountain Stream. Verhandlungen der Internationale Vereinigung für Limnologie
28:1–4.

Welber, M., W. Bertoldi, and M. Tubino. 2013. Wood Dispersal in Braided Streams: Results from
Physical Modeling. Water Resources Research 49:7388–7400.

Wellnitz, T., S. Y. Kim, and E. Merten. 2014. Do Installed Stream Logjams Change Benthic Community
Structure? Limnologica 49:68–72.

Welty, J. J., T. Beechie, K. Sullivan, D. M. Hyink, R. E. Bilby, C. Andrus, and G. Pess. 2002. Riparian
Aquatic Interaction Simulator (RAIS): A Model of Riparian Forest Dynamics for the Generation of
Large Woody Debris and Shade. Forest Ecology and Management 162:299–318.

Wemple, B. C., and J. A. Jones 2003. Runoff Production on Forest Roads in a Steep, Mountain
Catchment. Water Resources Research 39(8). doi:10.1029/2002WR001744

Wenger, S. J., D. J. Isaak, C. H. Luce, H. M. Neville, K. D. Fausch, J. B. Dunham, D. C. Dauwalter, M. K.


Young, M. M. Elsner, B. E. Rieman, A. F. Hamlet, and J. E. Williams. 2011. Flow Regime,
Temperature, and Biotic Interactions Drive Differential Declines of Trout Species under Climate

Large Wood National Manual July 2015


10-68
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Change. Proceedings of the National Academy of Sciences 108:14175–14180.


doi:10.1073/pnas.1103097108. Available: http://www.pnas.org/content/108/34/14175.full.
pdf+html.

Westerling, A. L., A. Gershunov, T. J. Brown, D. R. Cayan, and M. D. Dettinger. 2003. Climate and
Wildfire in the Western United States. Bulletin of the American Meteorological Society 84:595–
604. doi:10.1175/BAMS-84-5-595. Available: http://journals.ametsoc.org/doi/pdf/10.1175/
BAMS-84-5-595.

Westerling, A. L., H. G. Hidalgo, D. R. Cayan, and T. W. Swetnam. 2006. Warming and Earlier Spring
Increase Western U.S. Forest Wildfire Activity. Science 313: 940–943. doi:10.1126/science.
1128834.

Western Wood Products Association (WWPA). 1995. Ponderosa Pine Species Facts. Available:
www.wwpa.org/ppine.htm.

White, R. J., and O. M. Brynildson. 1967. Guidelines for Management of Trout Stream Habitat in
Wisconsin. Wisconsin Department of Natural Resources Technical Bulletin 39. Madison, WI.

White, P. S., and S. T. A. Pickett. 1985. Natural Disturbance and Patch Dynamics: An Introduction.
Pages 3–9 in S. T. A. Pickett and P. S. White (eds.), The Ecology of Natural Disturbance and Patch
Dynamics. San Diego, CA: Academic Press.

White, S. L., C. Gowan, K. D. Fausch, J. G. Harris, and W. C. Saunders. 2011. Response of Trout
Populations in Five Colorado Streams Two Decades After Habitat Manipulation. Canadian
Journal of Fisheries and Aquatic Sciences 68(12):2057–2063.

Whiteway, S. L., P. M. Biron, A. Zimmermann, O. Venter, and J. W. A. Grant. 2010. Do In-Stream


Restoration Structures Enhance Salmonid Abundance? A Meta-Analysis. Canadian Journal of
Fisheries and Aquatic Sciences 67:831–841.

Whiting, P. J. 2002. Streamflow Necessary for Environmental Maintenance. Annual Review of Earth
and Planetary Sciences. 30:181–206.

Whitney, G. G. 1996. From Coastal Wilderness to Fruited Plain: A History of Environmental Change in
Temperate North America from 1500 to the Present. Cambridge University Press: Cambridge, UK.

Whittaker, R. H., S. A. Levin, and R. B. Root. 1973. Niche, Habitat and Ecotope. American Naturalist
107(955):321–338.

Wiegner, T. N., L. A. Kaplan, J. D. Newbold, and P. H. Ostrom. 2005. Contribution of Dissolved Organic
C to Stream Metabolism: A Mesocosm Study Using C-13-Enriched Tree-Tissue Leachate. Journal
of the North American Benthological Society 24:48–67.

Wilcock, P. R., A. F. Barta, C. C. Shea, G. M. Kondolf, W. V. Graham Matthew, and J. Pitlick. 1996.
Observations of Flow and Sediment Entrainment on a Large Gravel-Bed River. Water Resources
Research 32:2897–2909.

Wilford, D., D. Maloney, J. Schwab, and M. Geertsema. 1998. Tributary Alluvial Fans. B.C. Ministry of
Forests Extension Note 30.

Williams, G. P. 1986. River Meanders and Channel Size. Journal of Hydrology 88(1-2):147–164.

Large Wood National Manual July 2015


10-69
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Williams, G. P., and M. G. Wolman. 1984. Downstream Effects of Dams on Alluvial Rivers. USGS
Professional Paper 1286.

Williams, B. K., R. C. Szaro, and C. D. Shapiro. 2009. Adaptive Management: The U.S. Department of the
Interior Technical Guide. Adaptive Management Working Group, U.S. Department of the Interior,
Washington, D.C.

Williams, K. L., S. W. Griffiths, K. H. Nislow, S. McKelvey, and J. D. Armstrong. 2009. Response of


Juvenile Atlantic Salmon, Salmo salar, to the Introduction of Salmon Carcasses in Upland
Streams. Fisheries Management and Ecology 16(4):290–297.

Williams, R. N. (ed.). 2006. Return to the River: Restoring Salmon Back to the Columbia River. New
York: Elsevier.

Wiltshire, P. E. J., and P. D. Moore. 1983, Paleovegetation and Paleohydrology in Upland Britain.
Pages 433–451 in K. J. Gregory (ed.), Background to Paleohydrology. Chichester, UK: John Wiley.

Winemiller, K. O., A. S. Flecker, and D. J. Hoeinghaus. 2010. Patch Dynamics and Environmental
Heterogeneity in Lotic Ecosystems. Journal of the North American Benthological Society 29:84–
99.

Wing, M. G., and A. Skaugset. 2002. Relationships of Channel Characteristics, Land Ownership, and
Land Use Patterns to Large Woody Debris in Oregon Streams. Canadian Journal of Fisheries and
Aquatic Sciences 59:796–807.

Wipf, T. J., B. M. Phares, and J. Dahlberg 2012. Debris Mitigation Methods for Bridge Piers. Iowa State
University, Ames, IA.

Wipfli, M. S., and C. V. Baxter. 2010. Linking Ecosystems, Food Webs, and Fish Production: Subsidies
in Salmonid Watersheds. Fisheries 35(8):373–387.

Wipfli, M. S., J. Hudson, and J. P. Caouette. 1998. Influence of Salmon Carcasses on Stream
Productivity: Response of Biofilm and Benthic Macroinvertebrates in Southeastern Alaska, U.S.A.
Canadian Journal of Fisheries and Aquatic Science 55:1503–1511.

Wipfli, M. S., J. Hudson, and J. P. Caouette. 2003. Marine Subsidies in Freshwater Ecosystems: Salmon
Carcasses Increase the Growth Rates of Stream-Resident Salmonids. Transactions of the
American Fisheries Society 132:371–381.

Wissmar, R. C., and R. L. Beschta. 1998. Restoration and Management of Riparian Ecosystems:
A Catchment Perspective. Freshwater Biology 40(3):571–585.

Wohl, E. E. 2001. Virtual Rivers: Lessons from the Mountain Rivers of the Colorado Front Range. New
Haven, CT: Yale University Press.

Wohl, E. 2011a. Seeing the Forest and the Trees: Wood in Stream Restoration in the Colorado Front
Range, United States. Pages 399–418 in A. Simon, S. J. Bennett, and J. Castro (eds.), Stream
Restoration in Dynamic Fluvial Systems: Scientific Approaches, Analyses, and Tools. Washington,
D.C.: American Geophysical Union Press.

Wohl, E. 2011b. What Should these Rivers Look Like? Historical Range of Variability and Human
Impacts in the Colorado Front Range, USA. Earth Surface Processes and Landforms 36:1378–
1390.

Large Wood National Manual July 2015


10-70
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Wohl, E. 2011c. Threshold-Induced Complex Behavior of Wood in Mountain Streams. Geology


39:587–590.

Wohl, E. 2013a. Redistribution of Forest Carbon Caused by Patch Blowdowns in Subalpine Forests of
the Southern Rocky Mountains, USA. Global Biogeochemical Cycles 27:1205-1213.

Wohl, E. 2013b. Floodplains and Wood. Earth-Science Reviews 123:194–212.

Wohl, E. 2014. A Legacy of Absence: Wood Removal in US Rivers. Progress in Physical Geography
38:637–663.

Wohl, E., and N. Beckman. 2014a. Controls on the Longitudinal Distribution of Channel-Spanning
Logjams in the Colorado Front Range, USA. River Research and Applications 30:112–131.

Wohl, E., and N. D. Beckman. 2014b. Leaky rivers: Implications for the loss of longitudinal fluvial
disconnectivity in headwater streams. Geomorphology 205:27–35.

Wohl, E., and D. Cadol. 2011. Neighborhood Matters: Patterns and Controls on Wood Distribution in
Old-Growth Forest Streams of the Colorado Front Range, USA. Geomorphology 125:132–146.

Wohl, E., and J. R. Goode. 2008. Wood Dynamics in Headwater Streams of the Colorado Rocky
Mountains. Water Resources Research 44:W09429.

Wohl, E., and K. Jaeger. 2009. A Conceptual Model for the Longitudinal Distribution of Wood in
Mountain Streams. Earth Surface Processes and Landforms 34:329–344.

Wohl, E., and D. J. Merritt. 2007. What is a Natural River? Geography Compass 1(4):871–900.

Wohl, E. and D. M. Merritt. 2008. Reach-Scale Channel Geometry of Mountain Streams.


Geomorphology 93(3-4):168–185.

Wohl, E., and F. L. Ogden. 2013. Organic Carbon Export in the Form of Wood During an Extreme
Tropical Storm, Upper Rio Chagres, Panama. Earth Surface Processes and Landforms 38:1407–
1416.

Wohl, E., F. L. Ogden, and J. Goode. 2009. Episodic Wood Loading in a Mountainous Neotropical
Watershed. Geomorphology 111:149–159.

Wohl E., D. A. Cenderelli, K. A. Dwire, S. E. Ryan-Burkett, M. K. Young, and K. D. Fausch. 2010. Large
in-Stream Wood Studies: A Call for Common Metrics. Earth Surface Processes and Landforms
35:618–625.

Wohl, E., L. E. Polvi, and D. Cadol. 2011. Wood Distribution Along Streams Draining Old-Growth
Forests in Congaree National Park, South Carolina, USA. Geomorphology 126:108–120.

Wohl, E., S. Bolton, D. Cado, F. Comiti, J. R. Goode, and L. Mao. 2012. A Two End-Member Model of
Wood Dynamics in Headwater Neotropical Rivers. Journal of Hydrology 462-463:67–76.

Wojan, C., A. Devoe, E. Merten, and T. Wellnitz. 2014. Web-building Spider Response to a Logjam in a
Northern Minnesota Stream. American Midland Naturalist 172(1):185–190.

Wolff, H. H. 1916. The Design of a Drift Barrier Across the White River, near Auburn, Washington.
Transactions of the American Society of Civil Engineers 16:2061–2085.

Large Wood National Manual July 2015


10-71
Bureau of Reclamation and
U.S. Army Corps of Engineers Chapter 10. Large Wood Bibliography

Wondzell, S. M. 2011. The Role of the Hyporheic Zone across Stream Networks. Hydrological
Processes 25:3525–3532.

Wondzell, S. M., and P. A. Bisson. 2003. Influence of Wood on Aquatic Biodiversity. Pages 249–263 in
S. V. Gregory, K. L. Boyer, and A. M. Gurnell (eds.), The Ecology and Management of Wood in
World Rivers. American Fisheries Society Symposium 37. Bethesda, MD: American Fisheries
Society.

Wondzell, S. M., J. LaNier, R. Haggerty, R. D. Woodsmith, and R. T. Edwards. 2009. Changes in


Hyporheic Flow Following Experimental Removal of a Small, Low-Gradient Stream. Water
Resources Research 45:W05406.

Wood, A. D., and A. R. Jarrett. 2004. Design Tool for Rootwads in Streambank Restoration. Paper
042047, Annual International Meeting, Ottawa. American Society of Agricultural Engineers. St.
Joseph, MI.

Wright, L. D. (ed.). 1995. Morphodynamics of Inner Continental Shelves. CRC Press.

Wuehlisch, G. Von. 2011. Evidence for nitrogen-Fixation in the Salicaceae Family. Tree Planters’
Notes 54(2):38–41.

Wyant, J. G., R. A. Meganck, and S. H. Ham. 1995. A Planning and Decision-Making Framework for
Ecological Restoration. Environmental Management 19(6):789–796.

Wyżga, B., and J. Zawiejska. 2005. Wood Storage in a Wide Mountain River: Case Study of the Czarny
Dunajec, Polish Carpathians. Earth Surface Processes and Landforms 30:1475–1494.

Yoccoz, N. G., J. D. Nichols, and T. Boulinier. 2001. Monitoring of Biological Diversity in Space and
Time. Trends in Ecology & Evolution 16(8):446–453.

Young, M. K., E. A. Mace, E. T. Ziegler, and E. K. Sutherland. 2006. Characterizing and Contrasting
Instream and Riparian Coarse Wood in Western Montana Basins. Forest Ecology and
Management 226:26–40.

Zarnetske, J. P., R. Haggerty, S. M. Wondzell, and M. A. Baker. 2011a. Labile Dissolved Organic Carbon
Supply Limits Hyporheic Denitrification. Journal of Geophysical Research 116:G04036.

Zarnetske, J. P., R. Haggerty, S. M. Wondzell, and M. A. Baker. 2011b. Dynamics of Nitrate Production
and Removal as a Function of Residence Time in the Hyporheic Zone. Journal of Geophysical
Research 116:G04025.

Zeng, H., J. Q. Chambers, R. I. Negron-Juarez, G. C. Hurtt, D. B. Baker, and M. D Powell. 2009. Impacts
of Tropical Cyclones on U.S. Forest Tree Mortality and Carbon Flux from 1851 to 2000.
Proceedings of the National Academy of Sciences 106(19), 7888–7892.

Zimmerman, R. C., J. C. Goodlett, and G. H. Comer. 1967. The Influence of Vegetation on Channel Form
of Small Streams, Symposium on River Morphology. International Association of Science
Hydrology Publication, Gentbrugge, Belgium 75:255–275.

Zobel, D. B., A. McKee, G. M. Hawk, and C. T. Dyrness. 1976. Relationships of Environment to


Composition, Structure, and Diversity of Forest Communities of the Central Western Cascades of
Oregon. Ecological Monographs 46:135–156.

Large Wood National Manual July 2015


10-72
Appendix A
Sample Implementation Contracts

A-1: Types of Federal Contracts Useful for Large Wood Projects ...................................................... A-1
A-2: Sample Documents for Hybrid Contracts ................................................................................. A-3
A-3: Sample Contract Language for Separate Harvest and Hauling Contract ..................................... A-9
A-4: Example—Safety and Health Provisions for Large Wood Placement Contracts ........................ A-10
This page intentionally left blank.
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

A-1: Types of Federal Contracts Useful for Large Wood


Projects
Described below are types of contracts that may be used by federal entities for large wood placement
projects. Federal procurement is government by Federal Acquisition Regulation (FAR) protocol and
procedures (http://www.acquisition.gov/far/). Often the contracting phase is a challenging step in the
implementation process, but project success may be enhanced using knowledge of the attributes of
available contracting arrangements.

Applicable contract types may be classified as follows:


 Firm fixed price
 Time and materials or labor-hour
 Hybrid (fixed price with time and materials tasks)
 Design-build

Fixed Price Contracts


Large wood (LW) restoration projects are most commonly implemented with fixed-price contracts.
Fixed price contracts allow for the contractor to determine the best method to use in order to meet the
requirements identified in the scope of work and specifications. The government can only accept or
reject work and at no time can the government direct the contractor’s labor force or equipment
operators. Fixed-price contracts place the maximum risk and full responsibility on the contractor for all
costs and resulting profit or loss associated with the work. This type of contract provides the maximum
incentive for the contractor to control costs and perform effectively and imposes a minimum
administrative burden on project sponsors.

A fixed-price contract requires the contractor to understand, in detail, what is to be constructed before
bidding to do the work. This requires a design that includes detailed drawings, specifications, and a bid
schedule containing items for each major project component. The designer must provide a cost estimate
by bid item so that the contracting officer can assess the reasonableness of the bids. Most fixed-price
contracts are awarded after contractors have submitted a sealed bid in response to an Invitation for Bids
(IFB). The IFB includes the drawings and specifications for the work and specific contract requirements.
The design effort and level of detail may be the same for simplified fixed-price contracts as it is for
formal fixed-price contracts.

Time and Materials Contracts


Time-and-materials contracts (we include labor-hour contracts within this category) are used to
procure supplies or services on the basis of direct labor and materials costs. Time-and-materials
contracts may be used only when it is not possible for the contracting agency to accurately estimate the
extent or duration of the work or to anticipate costs with any reasonable degree of confidence (FAR
16.601(c)). These types of contracts provide no positive profit incentive to the contractor for cost
control or labor efficiency; therefore, appropriate government surveillance of contractor performance is
required to give reasonable assurance that efficient methods and effective controls are being used.

Large Wood National Manual July 2015


A-1
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

These contracts include a ceiling price that the contractor exceeds at his own risk. Such contracts may be
the best choice for stream restoration projects, allowing for the most flexibility to direct the contractor’s
work. Field implementation decisions may be made as long as the scope of the contract is not modified.
It is essential to have experienced on-site construction support personnel and a field
inspection/surveillance team to support implementation of a time-and-materials contract.

Hybrid Contracts
The recommended approach for many stream restoration projects is a hybrid contract that balances the
flexibility of a time-and-materials contract with the reduced risk of a fixed-price contract. When the
project requires an activity for which it is difficult to write detailed specifications, a time-and-materials
task built in to a fixed-price framework is often the best option. One of the most important aspects of
stream restoration projects is flexibility during the construction phase. This flexibility is important for
implementation of tasks that contain a large degree of variability or intricacies that are often difficult to
define in LW projects.

Typically the scope of work details the number and type of equipment and the number and type of
personnel required present for each hourly unit of the specified task. The bid schedule will provide the
number of hours within the task.

For example,
“Typical equipment and crew composition utilized on past projects has included a Class 300
excavator with operator, a front end loader with operator, and an off-road dump truck with
operator/laborer. The above typical crew/equipment is the basis for estimation in determining
hourly units for this task, and each hour includes three pieces of equipment operating for each
individual hourly unit. Contractor must have available all applicable support equipment available
during the implementation of Task I. Examples of support equipment and hand tools are: chainsaw,
choker cable, chaps, gas/oil, etc.”

Examples of wording for specification packages, bid schedules, and scopes of work for hybrid contracts
are provided in Section A-2 below.

Design-Build Contracts
Design-build two-phase contracts are described in FAR Part 36.3
(http://www.acquisition.gov/FAR/97/pdf/36.pdf). Design-build contracts accomplish design and
construction implementation through one contract mechanism. This type of contract reduces the overall
duration of the project development by eliminating a second procurement process for construction.
Furthermore, integrating the design and construction activities can reduce the potential for design
errors and discontinuities between the design and construction efforts. Design-build contracts may yield
cost efficiencies by enabling the design-builder to propose alternate approaches to realize the
performance objectives of the project, including innovative technologies and methodologies that
leverage available government resources. By greater use of performance-based specifications that
promote creativity, design-build contracts may open opportunities to use value engineering more
frequently than in traditional design-bid-build projects. Significantly lower cost and claim frequency for
design-build projects reflect a fundamental shift in the adversarial nature of construction contracting
and bodes well for the future implementation of this procurement method, particularly for high visibility
projects where cooperation between contracting agencies and their design and construction contractors
is essential to project success.

Large Wood National Manual July 2015


A-2
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

A-2: Sample Documents for Hybrid Contracts


Specification Package
Below is example wording for a specification package for a fixed-price contract containing a time-and-
materials task.
1. Measurement and payment for miscellaneous minor changes will be made on a time-and-materials
basis in accordance with FAR clause 52.232-7, “Payments under Time-and-Materials and Labor-
Hour Contracts.” Contract Line Item Number (CLIN) X is for pricing minor changes to the original
contract work. The “minor changes” provision anticipates minor within-scope changes to work and
creates a method within the contract to more efficiently administer such minor changes that arise
during performance. It is primarily for those instances where the specifications or drawings were
incomplete, inadequate, or incorrect in the number or degree of items of work to be accomplished,
or something was clearly left off that needed to be included to accomplish the intended results(s).
a. Measurement: Measurement will be made of the actual hours worked and actual cost of
materials used in performing miscellaneous minor changes.
b. Payment: Payment for miscellaneous minor changes will be made of the actual hours worked
and actual cost of materials in accordance with the hourly rate and the material handling fee
offered in the schedule. NOTE: The estimated hours shown in the schedule will be used for
evaluation purposes only.
c. The Contractor shall invoice the Government for actual hours worked after the work is
complete. All costs of labor wages, equipment, indirect costs, general and administrative
expense, and profit shall be included in the hourly rate. Only the Contracting Officer (CO) can
approve overtime for work performed under the pay item for miscellaneous minor changes.
2. The Contractor will be reimbursed for the cost of materials and subcontracts in accordance with
FAR clause 52.232-7, “Payments under Time-and-Materials and Labor-Hour Contracts.” A material
handling fee may be included to the extent that it is clearly excluded from the hourly rate. The hours
and materials costs invoiced shall be only those required to perform miscellaneous minor changes
as directed by the Contracting Officer’s Technical Representative (COR).
3. The method that shall be used by the Contractor when a work item arises that may be addressed by
CLIN X is as follows:
a. Contractor shall notify the COR of the work he/she believes could be addressed under this CLIN
and asks for approval to perform the work under that CLIN (generally via phone or email).
b. The COR will contact the CO and request approval to allow the Contractor to perform this work
under the time and material CLIN. The CO will evaluate the information provided, and, if the
work is of minor consequence and performing the work under this CLIN would allow work to
continue forward in a timely and efficient manner, authorizes the use of the rates negotiated
under this CLIN to perform this work. If the CO does not believe the work is minor in nature or
determines another contractual method should be used, the CO will initiate the necessary action.
c. The COR will notify the Contractor and the Government Inspector that the element of work will
be performed under CLIN 12 using the contractually established rates.
d. The Contractor shall perform the work and include the necessary documentation to support the
billed hours and materials, if any, with his/her invoice. The COR will confirm with the
Government Inspector the accuracy of the proposed hours and materials, and notify the CO the
bill is correct and valid for payment.

Large Wood National Manual July 2015


A-3
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

Bid Schedule
Table A-1 below contains an example bid schedule for a hybrid contract.

Table A-1. Sample Bid Schedule for Fixed Price Contract with Time-and-Materials Items1

Item Est.
No. Supplies/Services Qty Unit Unit Price Amount
CLIN Task A1 – Reporting, Signage & Mobilization & Lump
1
001 Demobilization Sum $__________ $_________
CLIN Lump
Tasks A2 through A4 1
002 Sum $__________ $_________
CLIN Lump
Task B – Project Layout & Site Surveys 1
003 Sum $__________ $_________
CLIN Lump
Task C – Site Preparation 1
004 Sump $__________ $_________
Task D – In-Channel (IC) Features
CLIN Excavation Cut estimate: 12,200 cubic yards (cy) Lump
1
005 Boulder Estimate: 180 cy; Clean Gravel and Cobble Sum
estimate: 1,550 cy; Pit Run estimate: 3,030 cy $__________ $_________
Task E – Riverine (R) Features
CLIN Lump
Excavation Cut estimate: 40,765 cy 1
006 Sum
Infiltration Gravel Fill: 900 cy $__________ $_________
CLIN Task F – Upland (U) Features Lump
1
007 Fill & Spoil placement estimate: 47,300 cy Sum $__________ $_________
CLIN Lump
Task G – Final Site Preparation 1
008 Sum $__________ $________
Task H – Rock Material Supply
CLIN Pit Run estimate: 3,030 cy; Clean Gravel and Cobble Lump
1
009 estimate: 1,550 cy; Infiltration Rock estimate: 900 cy; Sum
Boulder estimate: 180 cy $__________ $_________
CLIN Task I – Stockpiled Material Installation
300 Hours
010 Hours assume crew $__________ $_________
CLIN
Task J – Contour Grading 60 Hours
011 $__________ $_________
CLIN Lump
Task K – Haul Large Wood 1
012 Sum $__________ $_________
CLIN Lump
Task L – Turbidity Control 1
013 Sum $__________ $_________
CLIN Lump
Task M – Plant Materials Supply 1
014 Sum $__________ $_________
CLIN Lump
Task N – Riparian Planting 1
015 Sum $__________ $_________
CLIN Lump
Task O (Optional) – Additional Rehabilitation Services 1
015 Sum $__________ $_________
1 Shaded rows are for time-and-materials items. Specific contract language for Task I is provided below.

Large Wood National Manual July 2015


A-4
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

Scope of Work
Below is an excerpt from the scope of work for an actual LW placement contract for the Trinity River
Restoration Program (TRRP). Content below corresponds to items I, J and K of the example bid schedule
depicted in Table A-1 above. Note that Tasks I and J are time-and-materials types, while K is a fixed-price
task.

Stockpiled Materials Installation – Task I


Contractor shall notify the Onsite Government Representative (OGR) at least 2 weeks before installing
stockpiled materials. Stockpiled materials including Large Wood (LW), slash, boulders, willow clumps,
topsoil, and other materials shall be placed by Contractor at locations marked by the OGR with
assistance from TRRP staff, or at locations otherwise indicated in the Performance Work Statement
(PWS). An accounting of stockpiled material installation per feature is included in Table A-2, Table A-3,
and Table A-4. Locations can be referenced per Technical Exhibit X (plan view). Measurement and
payment will be based on percentage completion as determined by measurement of hours.

Stockpiled materials shall be utilized to create habitat areas for the fishery, geomorphic, or riparian
revegetation purposes. Typical equipment and crew composition utilized on past projects has included a
Class 300 excavator with operator, a front end loader with operator, and an off-road dump truck with
operator/laborer. The above typical crew/equipment is the basis for estimation in determining hourly
units for this task, and each hour includes three pieces of equipment operating for each individual
hourly unit. Contractor must have available all applicable support equipment available during the
implementation of Task I. Examples of support equipment and hand tools are: chainsaw, choker cable,
chaps, gas/oil, etc.

Table A-2. Wood Material Accounting per Feature


12”–24” 12”–24” Tree Tops with Wood Slash Estimated
dbh tree diameter Limbs (12” (stems, Installation
stems w/ tree stem diameter and branches, brush time (3-
rootwad (log only) smaller) < 4” diameter) piece crew)
Location (each) (each) (each) (cubic yards) (hours)
IC-2 25 25 25 125 40
IC-4 @ head 20 20 20 100 40
IC-3 upper 4 6 6 25 8
IC-3 middle 5 8 8 32 8
IC-3 middle 4 6 6 25 8
IC-3 end 8 10 10 45 8
R-1 entrance 8 12 12 50 8
R-1 upper 8 10 8 45 8
R-1 mid 5 6 6 28 8
R-1 outlet 8 12 12 50 8
R-2 (multiple locations loose 40 60 60 250 40
placements)
W-1 Pond (multiple loose placements) 20 20 20 100 40

Large Wood National Manual July 2015


A-5
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

Table A-3. Boulder Accounting per Feature

Estimated Installation Time


Boulder Quantity (3 piece crew)
Location (cubic yards) (hours)
IC-1 30 10
IC-2 60 10
IC-4 60 10
IC-7 30 10

Table A-4. Salvaged Willow Clump Accounting per Feature

Willow Clump Estimated Installation Time


Quantity (3 piece crew)
Location (each) (hours)
IC-3 6 4
IC-4 4 4
R-1 10 6
R-4 6 6
R-2 16 8
R-5 8 4
W-1 6 4

Salvaged and Supplied Large Wood Debris (LWD) Installation


Contractor must provide an excavator operator with a minimum of 3 years’ experience in placing and
building large wood structures for river restoration habitat purposes, capable of working independently
with minimal direction and oversight. All LW must be anchored below grade to withstand, at a
minimum, base flow velocity conditions. Contractor must schedule wood installation into the overall
work schedule.

Salvaged Slash Material Installation


Slash is defined as woody material less than 6 inches in diameter that is stockpiled from site preparation
activities of clearing and grubbing. Slash will be used primarily in the construction of LW structures or
constructed wood jams to fill voids and create habitat. All slash not utilized for this purpose will
broadcast as slash mulch under Final Site Preparation – Task G. Slash mulch materials may be placed on
upland terraces or floodplain surfaces under Stockpiled Materials Installation – Task I as directed by
OGR.

Salvaged Boulder Installation


Contractor will place all stockpiled boulders at specified locations. Boulders may be placed in LW
structures, constructed wood jams, side channels, forced meanders, alcoves, or in the mainstem Trinity
River. Boulders are placed for both geomorphic and habitat purposes. Placement locations are indicated
in Table A-2.

Large Wood National Manual July 2015


A-6
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

Salvaged Willow Clump Installation


Salvaged clump plantings shall be replanted as quickly as feasible after salvage or removal from the
designated storage site. Clump placement locations shall comply with Technical Exhibit X.

Excavate a hole approximately the size of the rootwad along the low flow channel slope or surface. Any
competing vegetation within a 2-foot radius of the planting hole shall be removed. When digging the
hole for planting, leave a berm between the excavation and channel so not to affect turbidity. The side of
the planting hole shall be vertically lightly scarified, and the bottom shall be loosened to a minimum
depth of 6 inches. Each planting hole may be inspected by an OGR prior to planting. Planting holes shall
be filled with water at least 1 hour but not more than 2 hours before planting transplant.

Place one clump planting in the excavated hole, burying ½ to ⅔ of willow clump with ¼ to ½ of the root
mass into the groundwater. The planting hole shall be backfilled ⅔ full with the soil excavated from the
planting hole. The planting hole shall be filled with water to eliminate air pockets around roots. After the
hole has drained, add more soil and water until saturated backfill material covers the top of the root
crown to a minimum depth of 2 inches. After water has drained, Contractor shall backfill the hole with
the remaining soil to finish grade. After planting, remove ⅓ to ½ of the remaining willow stems. The
stems or trunks shall be lopped off after planting to make sure enough branches are sticking out of the
ground after setting the roots deep enough to reach the water table. Stems shall be lopped square across
the stem using sharp, clean lopping tools. Cut stem length shall be ⅓ to ½ of original stem length. After
planting, salvaged willow clumps shall be thoroughly watered.

Salvaged Topsoil Material Installation


If topsoil is encountered and stockpiled per specification, stockpiled topsoil shall be reapplied to
constructed riverine features above the waterline and included in a future executed modification.
Topsoil replacement minimizes the need for soil amendments associated with plantings, and grass seeds
shall be placed in an optimum medium for germination and establishment. Stockpiled organic
chipped/macerated material shall be spread as evenly as possible over the previously spread topsoil
before constructed surfaces are ripped. Organic material shall be applied to the spread topsoil no more
than 4 inches thick.

Contour Grading – Task J


Up to 6 acres of contour grading shall be utilized within riverine and upland areas to create topographic
complexity and provide positive drainage to the Trinity River or other hydrologic features within the
work area. The contour grading for topographic complexity and blending shall occur within disturbed
areas indicated in Table A-5.

Table A-5. Contour Grading Accounting per Feature

Time Estimate
Location Acres (hours)
C-7 2 20
C-4 2 20
C-5 1 10
W-1 1 10

Large Wood National Manual July 2015


A-7
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

Supply Large Wood – Task K


Contractor shall harvest 200 trees from within U-3 project footprint. Trees for harvest will be clearly
marked by OGR. Trees taken from outside the U-3 feature footprint shall not be measured toward
completion of this task. Wood materials generated from construction of other project features outside
the U-3 project footprint, and wood materials generated from clearing, are incidental to those tasks and
will not count to completion of task 5.11.

All trees must be removed with rootwad intact. The ground surface at the tree wells created from
removing rootwads must be smoothed and graded into adjacent ground to provide downhill drainage of
surface water. After removal, trees may be cut to a length between 30–40 feet without prior approval by
OGR, excluding length of attached rootwad. Limbs and branches will be left intact to the greatest extent
practicable. All slash generated from wood material shall be retained for use. It is anticipated that each
tree will generate three distinct products: one rootwad with attached stem 30–40’ in length, one stem
30–40’ in length, and one tree-top of varying length with intact branches.

LW materials will be stockpiled for placement as described under Stockpiled Materials Installation –
Task I. At least one stockpile will be created on each bank of the Trinity River and within a Contractor
use area, at a location mutually agreeable to the OGR and the Contractor. On the left bank of the Trinity
River, the product (rootwads with stem, stems, and tree-tops) of 90 trees will be stockpiled. On the right
bank of the Trinity River, the product of 105 trees will be placed in the stockpile. Materials taken across
the Trinity River must be backhauled to the greatest practical extent to reduce river crossing. Slash
materials generated during performance of this task will also be placed in the stockpiles. All materials
must be stored in piles or decks of similar product (i.e., one log deck of stems with rootwads, one log
deck of stems, one log deck of tree-tops, and one pile of slash). Stockpiles shall be organized to allow
direct access to load and transport each distinct material with a minimum of sorting and handling.

Contractor must take special care in handling wood materials so as not to damage during loading or
transport. No root balls will be removed to create more efficient hauling. Measurement and payment will
be based on percentage completion as determined by count of actual number of trees present in
stockpile.

Large Wood National Manual July 2015


A-8
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

A-3: Sample Contract Language for Separate Harvest


and Hauling Contract
Task 12: Large Wood Supply
12.1. Purpose: Under this task, the contractor will supply the TRRP with large wood logs stockpiled at a
secure location along the Trinity River for later use during channel rehabilitation activities. Location is
anticipated to be Douglas City or Lower Junction City Project Sites.

12.2. Statement of Work: The Contractor shall locate timber sources, secure permits, and harvest, haul,
and deck large wood within a 20-mile radius of the harvest area. The following are the types of materials
that will be paid for under this task:
 32 feet long x 12”–20” diameter at breast height (DBH) with root wad
 32 feet long x 12”–20” DBH without root wad
 32 feet long x 20” or greater DBH with root wad
 32 feet long x 20” or greater DBH without root wad
 Semi-end dump load of brush/limbs (slash)
 Optional – Additional haul distance greater than 20 miles from harvest site
 Optional – Secure stockpile decking location in Weaverville, CA, or Junction City, CA

Locate trees, obtain appropriate permits (federal, state, or local), fall trees, limb/stockpile slash,
load/haul trees and slash, and deck/store trees and slash at a designated location for later use by the
TRRP for channel rehabilitation activities. Below are the assumptions related to this task:
 Scope does not include reloading trees at stockpile area and hauling to U.S. Bureau of Reclamation
(USBR)-directed sites
 Scope assumes up to a 20-mile one-way haul distance between loading site and stockpile site
 Stockpile site is at agreed-upon location that is a secured gated area.
 If USBR identifies, permits, and pays royalties for a site, then the deduction for USBR source logs is
applicable. (Site must be comparable for access as other sites.)
 Basic fire equipment is included. If trees are required on short notice during high-fire season, there
may be added costs for fire watch labor.
 The harvesting area and haul road will be maintained and left in good condition.

12.3. Safety: The Contractor will contact Trinity County and any other applicable local, state, or federal
agency regarding constraints, weight limits, and other restrictions for roads, bridges, and other
requirements to implement the job. Roads subjected to interference by the work shall be kept open. The
contractor shall provide, erect, and maintain all necessary barricades, suitable and sufficient flasher
lights, flagmen, danger signals, and signs, and shall take all necessary precautions for the protection of
the work and the safety of the public within the roadway and when crossing the Trinity River. Specific
signs, barricades, and flagmen requirements are detailed the American National Standards Institute’s
“Manual on Uniform Traffic Control Devices for Streets and Highways” (ANSI 06.1). The Contractor shall
fully comply with Reclamation Safety and Health Standards (RSHS).

Large Wood National Manual July 2015


A-9
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

A-4: Example—Safety and Health Provisions for Large


Wood Placement Contracts
1.4.1 A Site-Specific Safety and Health Plan shall be submitted as part of the Work Plan required in
Section 5.1.2, Task A2. The general requirements stated below shall be addressed in the Site-
Specific Safety and Health Plan.
1.4.2 No one employed in management or performance of the contract (including subcontracts) shall
be required to work under conditions that are unsanitary, hazardous, or dangerous to the
employee’s health or safety. The Contractor shall fully participate in a Contractor Safety
Program Review meeting, according to Reclamation Safety and Health Standards (RSHS) Section
3.4.1, prior to mobilization. The Contractor shall comply with the clause titled “Accident
Prevention” in the U.S. Bureau of Reclamation (USBR) RSHS. The minimum work crew at any
time on the implementation site shall be no less than two people.
1.4.3 One copy of the USBR RSHS will be provided to the Contractor, at no charge, for use in
accordance with the notice titled “Notice of Safety and Health Requirements and of Safety
Handbook Availability – Reclamation.” Additional copies may be obtained from Superintendent
of Documents, item stock No. 024-003-00190-2, phone number 202-512-1800 or online at:
http://www.usbr.gov/ssle/safety/RSHS/rshs.html. Implementation Safety and Health
Standards promulgated by the Secretary of Labor may be obtained from any regional or area
office of the Occupational Safety and Health Administration of the U.S. Department of Labor.
1.4.4 The Contractor shall be cognizant of and ensure compliance with the requirements set forth in
the paragraphs above. The Contractor’s responsibility applies to all operations, including those
of the Contractor’s Subcontractors. When violations of safety and health requirements contained
in these specifications or referenced standards are called to the Contractor’s attention by the
Contracting Officer’s Technical Representative (COR), the Contractor shall immediately correct
the condition to which attention has been directed. Either oral or written notice will be deemed
sufficient. When the Contractor fails or refuses to promptly correct a compliance directive, the
COR may issue an order to stop all or any part of the work. When satisfactory corrective action is
taken, an order to resume work will be issued. The Contractor shall not be entitled to extension
of time, or to claims for damage, or to additional compensation by reason of either the directive
or the stop order. Failure of the Contracting Officer (CO) or the COR to order discontinuance of
any or the entire Contractor’s operations will not relieve the Contractor of the responsibility for
the safety of personnel and property.
1.4.5 The Contractor shall maintain an accurate record of, and report to the COR in the manner
prescribed by the CO, all cases of death, occupational diseases, or traumatic injury to employees
or the public involved, and property damage in excess of $2,500 occurring during the
performance of work under this contract. The rights and remedies of the Government provided
in this section are in addition to any other rights and remedies provided by law or under this
contract. In the event there is a conflict between requirements contained in USBR RSHS, this
Performance Work Statement (PWS), Contractor’s reviewed Safety Program, referenced safety
and health codes and standards, or the U.S. Department of Labor Implementation Safety and
Health Standards, promulgated under Section 107 of the Contract Work Hours and Safety
Standards Act (40 U.S.C. 327 et seq.), as amended, the more stringent requirement will prevail.
1.4.6 The Contractor shall comply with the noise levels in Table A-6 below:

Large Wood National Manual July 2015


A-10
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

Table A-6. Exterior Noise Level Standards at Rehabilitation Site Boundary

Measurement 7:00 AM to 7:00 PM 7:00 PM to 7:00 AM


Hourly equivalent sound level (Leq dB) 55 45
Maximum level dB 75 45

1.4.7 The work areas described in this PWS are popular recreational destinations for rafting,
kayaking, inner-tubing, and fishing. The public has, in the past, accessed the work areas by foot,
horseback, mountain bike, motorized vehicles, rafts, kayaks, and drift-boats. The Contractor
shall keep the public out of areas actively being worked, or in various degrees of completion, via
signs or other effective means as reviewed by the COR and in accordance with requirements
contained in USBR RSHS. Access through the work areas by watercraft on the Trinity River shall
be available to the public continuously for the performance period of the contract. Non-
motorized access to the sites shall be maintained outside of normal working hours and on
weekends and holidays, when work is not being performed. The Contractor shall provide, erect,
and maintain any and all necessary barricades and warning signs and take all necessary
precautions to protect the work and the safety of the public. A boater warning sign placed
upstream will be mandatory during in-channel implementation activities from July 15–
September 15 as required in contractors Work Plan – Task A2, section 5.1.2.
1.4.8 The Contractor shall develop Job Hazard Analyses (JHA) for each distinct phase of work as
directed by the Onsite Government Representative (OGR). Each JHA shall be given to the OGR for
review and acceptance. Work will not begin on the phase of work until the JHA is acceptable to
the OGR.
5.1.2 Work Plan – Task A2
5.1.2.1 The Contractor shall prepare and submit a Work Plan that will be used by the Contractor and
the Government to plan and manage the work to be performed. The Work Plan shall include an
overall description and schedule of all required activities including project tasks, milestones,
and management strategies. The Work Plan shall clearly describe the overall approach for
implementing and reporting on all required work. The responsibility and authority of all
organizations and key personnel involved in conducting each task will be outlined. The Work
Plan shall be submitted complete, and no partial submittal of Work Plan sections will be
allowed. Elements of the Work Plan shall include, but not be limited to, the following:
 Description of all tasks and subtasks and overall implementation approach to complete
these tasks;
 Calculations showing quantity of earthwork to be moved to meet the design digital terrain
modeling (DTM) lines and grades;
 Project site drawings with representative plan views, cross sections, and profiles for each
feature, and maps that will be used for implementation;
 Description of how the Contractor intends to comply with the requirements in the Water
Quality Certification 401 Permit;
 Project management strategy for achieving timely completion of all required work;

Large Wood National Manual July 2015


A-11
Bureau of Reclamation and
U.S. Army Corps of Engineers Appendix A. Sample Implementation Contracts

 Proposed detailed Critical Path schedule, including a bar chart timeline for completion of all
required tasks showing predecessor and successor relationships and critical milestone
dates. Schedule shall be prepared in Microsoft Project;
 Proposed composition and individual qualifications of a technical team or teams of
personnel;
 Proposed Contractor key personnel, work crew size, equipment, and supplies needed to
implement the contract;
 List of sub-contractors and responsibilities;
 Site-Specific Safety and Health Plan – See Section 5.1.2.2;
 Quality Control Plan – See Section 5.1.2.4;
 River Crossing Plan – See Section 5.1.2.5; and
 Traffic Control Plan – See Section 5.1.2.6.
5.1.2.2 Site-Specific Safety and Health Plan: As part of the Work Plan, develop a Site-Specific Safety
and Health Plan according to Section 3 of the USBR Reclamation Safety and Health Standards
(RSHS) manual. Cover all aspects of onsite and applicable offsite operations and activities
associated with this contract. Follow the outline in Appendix B of RSHS. The Plan will not be
accepted for review unless it addresses, in order, lettered and numbered per Appendix B, a
narrative for each item in the outline. Mark any item included in the outline that is not
applicable to this project as N/A after the item listing. The Plan shall also provide a list of Job
Hazard Analyses (JHA) anticipated throughout the project and a statement that additional JHA
shall be provided as required as the project progresses. The Safety and Health Plan shall include
a noise monitoring plan. Develop JHA for each distinct phase of work under the contract and as
directed by OGR. Activities involving hazardous materials shall have the appropriate Material
Safety Data Sheet(s) attached to the JHA. A generic Company Safety Plan is not acceptable. The
Safety Program shall be sit- specific for the requirements in this PWS.

Large Wood National Manual July 2015


A-12

You might also like