0% found this document useful (0 votes)
648 views628 pages

SHRP2 - S2-R19A-RW-2 Design Guide For Bridges For Service Life

book

Uploaded by

KY Peng
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)
648 views628 pages

SHRP2 - S2-R19A-RW-2 Design Guide For Bridges For Service Life

book

Uploaded by

KY Peng
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/ 628

Design Guide

for Bridges for


Service Life

S2-R19A-RW-2
TRANSPORTATION RESEARCH BOARD 2014 EXECUTIVE COMMITTEE*

OFFICERS
Chair: Kirk T. Steudle, Director, Michigan Department of Transportation, Lansing
Vice Chair: Daniel Sperling, Professor of Civil Engineering and Environmental Science and Policy; Director, Institute of Transportation
Studies, University of California, Davis
Executive Director: Robert E. Skinner, Jr., Transportation Research Board

MEMBERS
Victoria A. Arroyo, Executive Director, Georgetown Climate Center, and Visiting Professor, Georgetown University Law Center,
Washington, D.C.
Scott E. Bennett, Director, Arkansas State Highway and Transportation Department, Little Rock
Deborah H. Butler, Executive Vice President, Planning, and CIO, Norfolk Southern Corporation, Norfolk, Virginia (Past Chair, 2013)
James M. Crites, Executive Vice President of Operations, Dallas–Fort Worth International Airport, Texas
Malcolm Dougherty, Director, California Department of Transportation, Sacramento
A. Stewart Fotheringham, Professor and Director, Centre for Geoinformatics, School of Geography and Geosciences, University of
St. Andrews, Fife, United Kingdom
John S. Halikowski, Director, Arizona Department of Transportation, Phoenix
Michael W. Hancock, Secretary, Kentucky Transportation Cabinet, Frankfort
Susan Hanson, Distinguished University Professor Emerita, School of Geography, Clark University, Worcester, Massachusetts
Steve Heminger, Executive Director, Metropolitan Transportation Commission, Oakland, California
Chris T. Hendrickson, Duquesne Light Professor of Engineering, Carnegie Mellon University, Pittsburgh, Pennsylvania
Jeffrey D. Holt, Managing Director, Bank of Montreal Capital Markets, and Chairman, Utah Transportation Commission, Huntsville, Utah
Gary P. LaGrange, President and CEO, Port of New Orleans, Louisiana
Michael P. Lewis, Director, Rhode Island Department of Transportation, Providence
Joan McDonald, Commissioner, New York State Department of Transportation, Albany
Abbas Mohaddes, President and CEO, Iteris, Inc., Santa Ana, California
Donald A. Osterberg, Senior Vice President, Safety and Security, Schneider National, Inc., Green Bay, Wisconsin
Steven W. Palmer, Vice President of Transportation, Lowe’s Companies, Inc., Mooresville, North Carolina
Sandra Rosenbloom, Professor, University of Texas, Austin (Past Chair, 2012)
Henry G. (Gerry) Schwartz, Jr., Chairman (retired), Jacobs/Sverdrup Civil, Inc., St. Louis, Missouri
Kumares C. Sinha, Olson Distinguished Professor of Civil Engineering, Purdue University, West Lafayette, Indiana
Gary C. Thomas, President and Executive Director, Dallas Area Rapid Transit, Dallas, Texas
Paul Trombino III, Director, Iowa Department of Transportation, Ames
Phillip A. Washington, General Manager, Regional Transportation District, Denver, Colorado

EX OFFICIO MEMBERS
Thomas P. Bostick, (Lt. General, U.S. Army), Chief of Engineers and Commanding General, U.S. Army Corps of Engineers,
Washington, D.C.
Timothy P. Butters, Acting Administrator, Pipeline and Hazardous Materials Safety Administration, U.S. Department of Transportation
Alison J. Conway, Assistant Professor, Department of Civil Engineering, City College of New York, New York, and Chair, TRB Young
Members Council
T. F. Scott Darling III, Acting Administrator and Chief Counsel, Federal Motor Carrier Safety Administration, U.S. Department of Transportation
David J. Friedman, Acting Administrator, National Highway Traffic Safety Administration, U.S. Department of Transportation
LeRoy Gishi, Chief, Division of Transportation, Bureau of Indian Affairs, U.S. Department of the Interior, Washington, D.C.
John T. Gray II, Senior Vice President, Policy and Economics, Association of American Railroads, Washington, D.C.
Michael P. Huerta, Administrator, Federal Aviation Administration, U.S. Department of Transportation
Paul N. Jaenichen, Sr., Acting Administrator, Maritime Administration, U.S. Department of Transportation
Therese W. McMillan, Acting Administrator, Federal Transit Administration
Michael P. Melaniphy, President and CEO, American Public Transportation Association, Washington, D.C.
Gregory Nadeau, Acting Administrator, Federal Highway Administration, U.S. Department of Transportation
Peter M. Rogoff, Under Secretary for Policy, U.S. Department of Transportation
Craig A. Rutland, U.S. Air Force Pavement Engineer, Air Force Civil Engineer Center, Tyndall Air Force Base, Florida
Joseph C. Szabo, Administrator, Federal Railroad Administration, U.S. Department of Transportation
Barry R. Wallerstein, Executive Officer, South Coast Air Quality Management District, Diamond Bar, California
Gregory D. Winfree, Assistant Secretary for Research and Technology, Office of the Secretary, U.S. Department of Transportation
Frederick G. (Bud) Wright, Executive Director, American Association of State Highway and Transportation Officials, Washington, D.C.
Paul F. Zukunft, Adm., U.S. Coast Guard, Commandant, U.S. Coast Guard, U.S. Department of Homeland Security.

* Membership as of October 2014.


The Second Strategic Highway Research Program

Design Guide
for Bridges for
Service Life

SHRP 2 Report S2-R19A-RW-2

Atorod Azizinamini
Florida International University

Edward H. Power
HDR Engineering, Inc.

Glenn F. Myers
Atkins North America Inc.

H. Celik Ozyildirim
Virginia Center for Transportation Innovation and Research

Eric Kline
KTA-Tator, Inc.

David W. Whitmore
Vector Corrosion Technologies Ltd.

Dennis R. Mertz
University of Delaware

Transportation Research Board

Washington, D.C.
2014
www.TRB.org
Subject Areas
Bridges and Other Structures
Highways
Maintenance and Preservation
Materials
The Second Strategic Highway SHRP 2 Report S2-R19A-RW-2
ISBN: 978-0-309-27326-8
Research Program
© 2014 National Academy of Sciences. All rights reserved.
America’s highway system is critical to meeting the mobility
and economic needs of local communities, regions, and the Copyright Information
nation. Developments in research and technology—such as Authors herein are responsible for the authenticity of their
advanced materials, communications technology, new data materials and for obtaining written permissions from pub-
collection technologies, and human factors science—offer lishers or persons who own the copyright to any previously
a new opportunity to improve the safety and reliability of published or copyrighted material used herein.
this important national resource. Breakthrough resolution
of significant transportation problems, however, requires The second Strategic Highway Research Program grants
concentrated resources over a short time frame. Reflecting permission to reproduce material in this publication for
this need, the second Strategic Highway Research Program classroom and not-for-profit purposes. Permission is given
(SHRP 2) has an intense, large-scale focus, integrates mul- with the understanding that none of the material will be
tiple fields of research and technology, and is fundamentally used to imply TRB, AASHTO, or FHWA endorsement of a
different from the broad, mission-oriented, discipline-based particular product, method, or practice. It is expected that
research programs that have been the mainstay of the high- those reproducing material in this document for educa-
way research industry for half a century. tional and not-for-profit purposes will give appropriate ac-
The need for SHRP 2 was identified in TRB Special knowledgment of the source of any reprinted or reproduced
Report 260: Strategic Highway Research: Saving Lives, material. For other uses of the material, request permission
Reducing Congestion, Improving Quality of Life, pub- from SHRP 2.
lished in 2001 and based on a study sponsored by C ­ ongress Note: SHRP 2 report numbers convey the program, focus
through the Transportation Equity Act for the 21st Cen- area, project number, and publication format. Report num-
tury (TEA-21). SHRP 2, modeled after the first Strategic bers ending in “w” are published as web documents only.
Highway Research Program, is a focused, time-constrained,
management-driven program designed to complement NOTICE
existing highway research programs. SHRP 2 focuses
on applied research in four areas: Safety, to prevent or The project that is the subject of this report was a part of the
reduce the severity of highway crashes by understanding second Strategic Highway Research Program, conducted by
driver behavior; Renewal, to address the aging infrastruc- the Transportation Research Board with the approval of
ture through rapid design and construction methods that the Governing Board of the National Research Council.
cause minimal disruptions and produce lasting facilities; The members of the technical committee selected to moni-
­Reliability, to reduce congestion through incident reduc- tor this project and to review this report were chosen for
tion, management, response, and mitigation; and Capacity, their special competencies and with regard for appropriate
to integrate mobility, economic, environmental, and com- balance. The report was reviewed by the technical commit-
munity needs in the planning and designing of new trans- tee and accepted for publication according to procedures
portation capacity. established and overseen by the Transportation R ­ esearch
SHRP 2 was authorized in August 2005 as part of Board and approved by the Governing Board of the
the Safe, Accountable, Flexible, Efficient Transportation ­National Research Council.
Equity Act: A Legacy for Users (SAFETEA-LU). The pro-
The opinions and conclusions expressed or implied in this
gram is managed by the Transportation Research Board
report are those of the researchers who performed the re-
(TRB) on behalf of the National Research Council (NRC).
search and are not necessarily those of the Transportation
SHRP 2 is conducted under a memorandum of understand-
Research Board, the National Research Council, or the pro-
ing among the American Association of State Highway and
gram sponsors.
Transportation Officials (AASHTO), the Federal Highway
Administration (FHWA), and the National Academy of Sci- The Transportation Research Board of the National Acad-
ences, parent organization of TRB and NRC. The program emies, the National Research Council, and the sponsors of
provides for competitive, merit-based selection of research the second Strategic Highway Research Program do not en-
contractors; independent research project oversight; and dorse products or manufacturers. Trade or manufac­turers’
dissemination of research results. names appear herein solely because they are considered
essen­tial to the object of the report.

SHRP 2 Reports
Available by subscription and through the TRB online
bookstore: www.mytrb.org/store
Contact the TRB Business Office: 202.334.3213
More information about SHRP 2: www.TRB.org/SHRP2
The National Academy of Sciences is a private, nonprofit, self-perpetuating society of distinguished schol-
ars engaged in scientific and engineering research, dedicated to the furtherance of science and technology
and to their use for the general welfare. On the authority of the charter granted to it by the Congress in
1863, the Academy has a mandate that requires it to advise the federal government on scientific and techni-
cal matters. Dr. Ralph J. Cicerone is president of the National Academy of Sciences.

The National Academy of Engineering was established in 1964, under the charter of the National Academy
of Sciences, as a parallel organization of outstanding engineers. It is autonomous in its administration and
in the selection of its members, sharing with the National Academy of Sciences the responsibility for advis-
ing the federal government. The National Academy of Engineering also sponsors engineering programs
aimed at meeting national needs, encourages education and research, and recognizes the superior achieve-
ments of engineers. Dr. C. D. Mote, Jr., is president of the National Academy of Engineering.

The Institute of Medicine was established in 1970 by the National Academy of Sciences to secure the ser-
vices of eminent members of appropriate professions in the examination of policy matters pertaining to
the health of the public. The Institute acts under the responsibility given to the National Academy of Sci-
ences by its congressional charter to be an adviser to the federal government and, upon its own initiative,
to identify issues of medical care, research, and education. Dr. Victor J. Dzau is president of the Institute
of Medicine.

The National Research Council was organized by the National Academy of Sciences in 1916 to associate
the broad community of science and technology with the Academy’s purposes of furthering knowledge
and advising the federal government. Functioning in accordance with general policies determined by the
Academy, the Council has become the principal operating agency of both the National Academy of Sci-
ences and the National Academy of Engineering in providing services to the government, the public, and
the scientific and engineering communities. The Council is administered jointly by both Academies and the
Institute of Medicine. Dr. Ralph J. Cicerone and Dr. C. D. Mote, Jr., are chair and vice chair, respectively,
of the National Research Council.

The Transportation Research Board is one of six major divisions of the National Research Council. The
mission of the Transportation Research Board is to provide leadership in transportation innovation and
progress through research and information exchange, conducted within a setting that is objective, inter­
disciplinary, and multimodal. The Board’s varied activities annually engage about 7,000 engineers, sci-
entists, and other transportation researchers and practitioners from the public and private sectors and
academia, all of whom contribute their expertise in the public interest. The program is supported by
state transportation departments, federal agencies, including the component administrations of the U.S.
Department of Transportation, and other organizations and individuals interested in the development of
transportation. www.TRB.org

www.national-academies.org
SHRP 2 STAFF
Ann M. Brach, Director
Stephen J. Andrle, Deputy Director
Neil J. Pedersen, Deputy Director, Implementation and Communications
Cynthia Allen, Editor
Kenneth Campbell, Chief Program Officer, Safety
JoAnn Coleman, Senior Program Assistant, Capacity and Reliability
Eduardo Cusicanqui, Financial Officer
Richard Deering, Special Consultant, Safety Data Phase 1 Planning
Shantia Douglas, Senior Financial Assistant
Charles Fay, Senior Program Officer, Safety
Carol Ford, Senior Program Assistant, Renewal and Safety
James Hedlund, Special Consultant, Safety Coordination
Alyssa Hernandez, Reports Coordinator
Ralph Hessian, Special Consultant, Capacity and Reliability
Andy Horosko, Special Consultant, Safety Field Data Collection
William Hyman, Senior Program Officer, Reliability
Linda Mason, Communications Officer
Matthew Miller, Program Officer, Capacity and Reliability
David Plazak, Senior Program Officer, Capacity and Reliability
Rachel Taylor, Senior Editorial Assistant
Dean Trackman, Managing Editor
Connie Woldu, Administrative Coordinator
ACKNOWLEDGMENTS

This work was sponsored by the Transportation Research Board of the National Acad-
emy of Sciences and was conducted as part of the Second Strategic Highway Research
Program (SHRP 2). The project (R19A) was managed by program officers Monica
Starnes (December 2007 through January 2011), Mark Bush (January 2011 through
November 2011), and Jerry DiMaggio (December 2011 onward), and the authors
greatly appreciate their input and guidance throughout the project. The project princi-
pal investigator was Atorod Azizinamini, Chair­person of the Civil and Environmental
Engineering Department at Florida International University.
Don White of the Georgia Institute of Technology is greatly acknowledged for
being a member of the research team and for his contributions to the development of
the Guide.
Several consultants and industry representatives provided input into the research
and development of various parts of the Guide. In particular, the authors thank ­Martin
Burke, private consultant; Reid W. Castrodale, Carolina Stalite Company; David
­Darwin, University of Kansas; Simon Greensted, Sterling Lloyd; Mark ­Kaczinski,
D.S. Brown Company; Ralph Oesterle, CTL Group; Duncan Paterson, HDR Engi-
neering, Inc.; Charles Roeder, University of Washington; and Ronald J. Watson,
R.J. Watson, Inc.
Throughout the research studies, various bridge committees within the AASHTO
Subcommittee on Bridges and Structures (SCOBS) provided valuable input and review
comments. The authors especially thank AASHTO Technical Committee T-9 and
the leadership and help provided by Mr. Bruce Johnson, Chair of the A ­ ASHTO T-9
Committee.
The authors also thank the members of the Project R19A TETG for their valuable
technical comments and guidance throughout the project.
Several research associates and graduate students assisted in conducting the
project. Nima Ala, Saeed Doust, Marcelo Da Silva, and Ardalan Sherafati obtained
their Ph.D. degrees by carrying out various research tasks within Project R19A, and
Luke Glaser and Kyle Burner obtained their M.S. degrees. The authors thank Kromel
Hanna, Research Associate at the University of Nebraska–Lincoln (Omaha campus),
for his contributions.
The authors especially thank Aaron Yakel, Research Associate at Florida Inter-
national University, for his assistance with various research topics and with putting
the Guide in its final format. Special thanks are also due to Christine Boyer of Boyer
Associates for her professional editing of the Guide and for her coordination with the
editorial staff at SHRP 2, headed by Dean Trackman.
The authors acknowledge the contribution of the University of Nebraska–Lincoln
for providing the testing facility to carry out the experimental work while the P.I. was
a member of the faculty.
FOREWORD

Jerry A. DiMaggio, D.GE, PE


SHRP 2 Senior Program Officer, Renewal

This report, Design Guide for Bridges for Service Life, is the main product of SHRP 2
Project R19A, Bridges for Service Life Beyond 100 Years: Innovative Systems, Sub-
systems, and Components. Compared with traditional approaches, which are based
solely on strength considerations, this guide advances an emerging approach to bridge
design, rehabilitation, and preservation that is based on service life considerations.
Both the Guide and its companion report, Bridges for Service Life Beyond 100 Years:
Innovative Systems, Subsystems, and Components, are available at the Transportation
Research Board website (http://www.trb.org/Design/Blurbs/168760.aspx).
The main objective of the Guide is to provide information and guidance and to
define procedures to systematically approach service life and durability for both new
and existing bridges. The Guide equips users with knowledge to develop specific solu-
tions for a bridge under consideration in a systematic manner by using a standard
framework. In some respects, the R19A Guide may be considered a foundational refer-
ence that will be built on, expanded, modified, and progressively embraced at different
project and program levels by the bridge and structures community. The future path of
development and mainstream acceptance and implementation may be similar to that of
load resistance factor design (LRFD) specifications.
Providing safety for the public by having adequate strength is the cornerstone
of the framework used by engineers for bridge design. This approach has not been
restricted to bridges; for example, it has also been the framework used in various
building codes. Significant changes to our contemporary bridge design practice have
also been mainly related to strength issues. The transition to LRFD is a well-known
recent example. A review of bridges that have lasted more than 100 years provides
valuable lessons related to achieving long service lives. These bridges—which have
proved to be maintainable and well maintained over their lives and adaptable to func-
tional changes—were all originally overdesigned.
As limited resources demand enhancing the service life of existing and new
bridges, the design for service life is gaining more importance. The cost of addressing
service life issues at the design stage is significantly lower than taking maintenance
and preservation actions while the bridge is in service. In general, design for service
life is approached by using individual strategies, each capable of enhancing the ser-
vice life of a particular bridge element that historically has experienced unsatisfactory
performance.
The R19A Guide addresses service life in a systematic manner by using a frame-
work that is general and applicable for all bridges, while having specifics that differ
from one bridge to another. These differences reflect that design for service life is a
context-sensitive problem-solving method that necessarily considers local experiences,
practice, and owner preferences. The Guide includes both well-proven and new con-
cepts and approaches capable of enhancing the service life of bridges. The Guide’s
objective is achieved through 11 chapters, each devoted to certain parts of a bridge or
aspects of the service life design process.
Contents

1 Chapter 1 D
 esign for Service Life:
General Framework
1 1.1 Background
5 1.2 Objectives of the Guide
6 1.3 Bridge Service Life Terminology and Relationships
7 1.4 Guide Approach to Design for Service Life
13 1.5 Organization of the Guide
14 1.6 Categories of Information Provided in Guide Chapters
18 1.7 Quantifying Service Life of Bridge Element, Component,
Subsystem, and System
25 1.8 Owner’s Manual
26 1.9 Independent Review of Design for Service Life Process
26 1.10 Summary of Steps for Design for Service Life for Specific Bridge
Element, Component, and Subsystem
27 1.11 Approaches to Using the Guide
47 1.12 Future Development of the Guide

49 Chapter 2 B
 ridge System Selection
49 2.1 Introduction
50 2.2 Bridge System Description
72 2.3 Factors Affecting Service Life
103 2.4 Options for Enhancing Service Life
111 2.5 Strategy Selection
121 2.6 Bridge Management
123 Chapter 3 Materials
123 3.1 Introduction
124 3.2 Description of Material Types
154 3.3 Concrete and Steel Distresses and Solutions
165 3.4 Fault Tree Analysis of Factors Influencing Service Life
179 3.5 Individual Strategies to Mitigate Factors Affecting Service Life
183 3.6 Overall Strategies for Enhanced Material Service Life

186 Chapter 4 Bridge Decks


186 4.1 Introduction
186 4.2 Description of Bridge-Deck Types
194 4.3 Factors Influencing Bridge-Deck Service Life
212 4.4 Individual Strategies to Mitigate Factors Affecting Service Life
225 4.5 Overall Strategies for Enhanced Bridge-Deck Service Life

235 Chapter 5 Corrosion of Steel in


Reinforced Concrete Bridges
235 5.1 Introduction
236 5.2 Description of Corrosion
240 5.3 Factors Influencing Corrosion
242 5.4 Strategies for Addressing Corrosion
250 5.5 Case Studies Addressing Corrosion in Existing Structures

258 Chapter 6 Corrosion Prevention of


Steel Bridges
258 6.1 Introduction
259 6.2 Description of Methods for Corrosion Prevention
284 6.3 Factors Adversely Affecting Service Life
295 6.4 Options for Enhancing Service Life: Corrosion Performance of
Steel
306 6.5 Strategy Selection Process

316 Chapter 7 Fatigue and Fracture of


Steel Structures
316 7.1 Introduction
316 7.2 Background
321 7.3 Crack Detection Techniques
321 7.4 Repair and Retrofit Methods
325 7.5 Fatigue Caused by Secondary Stresses
335 7.6 Retrofit Validation of Secondary Stress Fatigue
336 7.7 Load-Controlled Fatigue Crack Repair
343 Chapter 8 Jointless Bridges
343 8.1 Introduction
343 8.2 History of Jointless Bridges
344 8.3 Types of Jointless Bridges
349 8.4 Factors Affecting Performance of Jointless Bridges
354 8.5 Strategy Selection Process
358 8.6 Design Provisions for Jointless Bridges
397 8.7 Details
406 8.8 Construction
408 8.9 Maintenance and Repair
411 8.10 Retrofits

415 Chapter 9 Expansion Devices


415 9.1 Introduction
416 9.2 Descriptions of Various Expansion Joint Devices
428 9.3 Factors and Considerations Influencing Expansion Joint Service
Life
435 9.4 Overall Strategies for Enhanced Bridge Expansion Joints
Service Life

445 Chapter 10 Bridge Bearings


445 10.1 Introduction
445 10.2 Bearing Types
454 10.3 Factors Influencing Service Life of Bearings
471 10.4 Options for Enhancing Service Life of Bearings
488 10.5 Strategies for Bearing Selection and Design
498 10.6 Bridge Management Related to Bearings

502 Chapter 11 Life-Cycle Cost Analysis


502 11.1 Introduction
503 11.2 LCCA Defined
505 11.3 Elements of LCCA

521 References
543 Appendix A Design Provisions for
Self-Stressing System for
Bridge Application with
Emphasis on Precast Panel
Deck System
544 A.1 Construction Procedure Overview
545 A.2 Design Considerations
546 A.3 Design Procedure and Implementation Details
556 A.4 Design Flowchart
557 A.5 Design Aids for Two-Span Bridges

559 Appendix B Displacement of Skewed


Bridge
559 B.1 Background
563 B.2 Analyses for Transverse Response to Thermal Expansion
566 B.3 Expected Transverse Movement with Typical Integral
Abutment

570 Appendix C Design of Piles for Fatigue


and Stability
570 C.1 Estimation of Maximum Allowable Strain
573 C.2 Pushover Analysis Example

576 Appendix D Restraint Moments


576 D.1 Background
581 D.2 Design Recommendations

587 Appendix E Design Steps for Seamless


Bridge System Developed by
SHRP 2 Project R19A
590 E.1 Structural Analysis
590 E.2 Design of System Components
593 E.3 Cracked Section Analysis

596 Appendix F Curved Girder Bridges


596 F.1 Background
597 F.2 Calculating Magnitude and Direction of End Displacement
600 F.3 Optimum Pile Orientation
602 Appendix G Design Provision for Sliding
Surfaces Used in Bearing
Devices for Service Life
602 G.1 Introduction
603 G.2 Elements of Design Provisions
609 G.3 Design Process for Sliding Surfaces
1
Design for Service Life:
General Framework

The design for service life is gaining more importance as limited resources demand
enhancing the service life of existing and new bridges. As part of SHRP 2 Project
R19A, Bridges for Service Life Beyond 100 Years: Innovative Systems, Subsystems,
and Components, a systematic and general approach to design for service life has been
developed. The major product of this project is this document, titled Design Guide for
Bridges for Service Life, and referred to here as the Guide. This chapter provides the
general framework used in the Guide, which primarily pertains to bridges with spans
of less than about 300 ft. However, the framework is general and can be adapted and
customized for major and complex bridges, including those with much longer spans.
It can also be adapted to suit bridges located in any region in the United States, rec-
ognizing, however, that although the framework remains the same for all bridges, the
resulting details for service life could be significantly different.
1.1
Background
Providing safety for the public by having adequate strength for constructed facilities
has been the cornerstone of the framework used by engineers for bridge design. This
design for strength approach has not been restricted to bridges: it has also been the
framework one could find in various building codes. In the case of buildings, however,
most structural elements are protected from environmental-type loads, and as a result
the strength framework has served this sector of the industry very well. In the case of
bridges or pavements, which are constructed facilities exposed to environmental loads,
the story is different.
Significant changes to contemporary bridge design specifications have also been
mainly related to strength issues. The transitions from allowable stress design to load
factor design, and more recently to load and resistance factor design (LRFD), reflect

1
this line of thinking. It is important to note that in the early 1970s, bridge engineers
developed criteria for steel bridge details to protect against fatigue and fracture failure.
These were indeed service life design provisions.
The strength framework did not prevent visionary engineers such as John R ­ oebling
from thinking in terms of service life. A review of bridges that have lasted more than
100 years provides valuable lessons. These bridges are not so much innovative in sys-
tem or material, but they have proved to be

• Maintainable and well maintained over their 100-year lives as a result of their
extreme importance or high capital replacement cost;
• Adaptable to changes in functional use, as well as service limit state demands; and/
or
• Originally overdesigned.

Examples of bridges with long service lives are New York City’s oldest East River
bridges, the Brooklyn Bridge (the longest bridge in the world when opened to traffic
in 1883), the Williamsburg Bridge (the longest bridge in the world when opened in
1903), and Saint Louis’s Eads Bridge (the first steel bridge, opened in 1874).
The Brooklyn Bridge has been well maintained and rehabilitated in a timely manner
throughout its lifetime. Initial coatings to protect the bridge’s steel from corrosion did
not provide a 100-year life, but cleaning and repainting the bridge did. The metal deck
of the Brooklyn Bridge has not survived its 100-plus-year service life, but replacement
of the replaceable metal decking has. Figure 1.1 shows the Brooklyn Bridge circa 1890.
The Williamsburg Bridge was not as well maintained, as evidenced by its emer-
gency closing in 1988. In April of that year, after a thorough inspection revealed cor-
rosion of the cables, beams, and steel supports, the Williamsburg Bridge was closed
to all vehicular and train traffic for nearly 2 months. After engineers performed emer-
gency construction on the bridge and reopened it to traffic, a panel of design experts
convened to determine if the Williamsburg Bridge should be replaced or if it should be
rehabilitated. In November 1988, after evaluating several alternatives, the New York
City Department of Transportation (DOT) determined that the Williamsburg Bridge

Figure 1.1.  The


­Brooklyn Bridge from the
South Street Seaport, circa
1890.
Source: Photo courtesy
Library of Congress, Prints
and Photo­graphs Division,
Detroit Publishing ­Company
Collection, LC-D4-90107.

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


should be repaired while kept open to traffic. This option was deemed to have the least
detrimental impact on motorists and nearby communities. In 1991, the New York
City DOT began a major rehabilitation of the Williamsburg Bridge. The program was
designed to undo the effects of age, weather, increased traffic volumes, and deferred
maintenance. Figure 1.2 shows the Williamsburg Bridge circa 1904.
The decision to rehabilitate the Williamsburg Bridge instead of undertaking a costly
in-place replacement in downtown Manhattan was made possible by the original con-
servative design of the bridge cables. The need to rehabilitate the cables was necessi-
tated by a poor corrosion-protection choice. The 1988 inspection of the Williams­burg
Bridge cables revealed significant corrosion, proving the choice of linseed oil by ­Leffert
L. Buck, the designer of the bridge, to be a relatively poor one. For the Brooklyn
Bridge, John Augustus Roebling chose a coating of graphite to protect the individual
wires of the bridge cable from corrosion, a choice that provided over 100 years of
corrosion protection. Fortunately, the cable design for the Williamsburg Bridge used a
factor of safety of resistance divided by load of about five. After the significant loss of
section as a result of corrosion was observed in 1988, the factor of safety was deemed
adequate, and a cable rehabilitation program to arrest the corrosion was initiated
instead of a cable replacement. Thus, the original overdesign allowed the bridge and
its cables to continue in service.
The Eads Bridge, completed in 1874 and named for its designer and builder, James
Buchanan Eads, has proved long lived by being well maintained and readily adaptable.
Figure 1.3 shows the Eads Bridge circa 1983. The scale of the bridge was unprece-
dented: the more than 500-ft span of the center arch exceeded by some 200 ft any arch
built previously. The arch ribs were made of steel, its first extensive use in a bridge. An
additional innovation was the cantilever erection of the arches without falsework, the
first example of this type of construction for a major bridge.
An interesting feature of the history of the Eads Bridge has been its adaptability to
varying use. (The Brooklyn and Williamsburg bridges have also seen varied use.) The
bridge was originally a railway bridge carrying pedestrians on an upper deck with two
rail lines below. It eventually carried vehicular and rail traffic; the last train crossed

Figure 1.2. The W­illiamsburg Bridge, circa 1904.


Source: Photo courtesy Library of Congress, Prints and
Photographs Division, Detroit Publishing Company
Collection, LC-D4-17414.

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Figure 1.3.  The Eads Bridge looking toward Saint
Louis and the Gateway Arch, circa 1983.
Source: Photo courtesy Library of Congress, Prints and
Photographs Division, Detroit ­Publishing Company
Collection.

the bridge in 1974. By the early 1990s, traffic on the bridge had dwindled to about
4,000 cars a day, and in 1991, the Eads Bridge was closed. It was briefly unused until
1993, when MetroLink, the region’s new light rail system, began to use the lower deck,
which originally served passenger and freight train traffic. In 2003, the upper deck
reopened to buses and automobiles. Today, a new lane for pedestrians and bicyclists
on the south side of the bridge provides a great place to look at the river and the Saint
Louis skyline.
The examples of these three 100-plus-year-old bridges illustrates that for bridges
to serve a long life, they must be

• Resistant to environmental and human-caused hazards;


• Maintainable (and subsequently maintained) or relatively maintenance free; and
• Adaptable to changes in traveled-way cross section and usage.

Traditional approaches for enhancing the service life of bridges used in various
codes and specifications, such as American Association of State Highway and Trans-
portation Officials (AASHTO) specifications, Eurocodes, or British Standards, are
mainly in an indirect form, specifying the use of certain details or properties such as
cover thickness, maximum crack width, and concrete compressive strength.
Recognizing the importance of design for service life has motivated different agen-
cies to undertake new initiatives for developing more formal design approaches for
service life, similar to those used for design for strength. However, to date the majority
of these efforts have concentrated on addressing concrete durability and service life,
and significant advances have been achieved in this field. Designing bridges for service
life, however, is more than just addressing service life and durability of concrete.
The design for service life for bridges needs to be approached in a systematic,
all-inclusive manner rather than as a series of isolated tasks, each independently
addressing the service life of a particular portion of a bridge. The interaction among
4

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


strategies for enhancing the service life of different bridge elements, components, and
subsystems must be given critical consideration. In addition, a maintenance program,
retrofit or replacement options, and management plan should all be part of this sys-
tematic service life design approach. In summary, at the design stage the design for
service life should be approached as a comprehensive plan capable of providing the
owner with a complete picture of what will be necessary for the bridge to achieve its
specified service life.
The most notable efforts to develop a scientific approach for service life and dura-
bility of concrete elements (covering buildings, bridges, and tunnels) were a series of
studies carried out between 1996 and 1999 in Europe for the Fédération International
du Béton (International Federation for Structural Concrete). One of the products of
these efforts was the publication of fib Bulletin 34, Model Code for Service Life Design
(FIB 2006). Bulletin 34, however, focused only on concrete service life and durability.
Further, caution must be exercised when applying the recommendations of this publi-
cation to concrete placed in a horizontal configuration, such as a bridge deck. Although
Bulletin 34 has many useful recommendations for designing concrete elements for
service life and durability, the application of these recommendations to bridge compo-
nents such as bridge decks remains a point of debate (in particular, the use of various
solutions to Fick’s second law to predict the rate of chloride ingress through deck con-
crete). Bulletin 34 recommendations are believed to be most applicable for concrete
in vertical configuration and under compression, such as in sub­structure columns or
sides of concrete box girders. This same debate can also be extended to the use of some
of the available commercial and noncommercial programs that use the fundamental
concepts stated in Bulletin 34.
Efforts to address service life of bridges are not limited to Europe. A significant
number of research studies have been carried out and continue to be performed to
develop solutions for various service life issues related to different bridge types.
One of the missing elements for designing bridges for service life has been a frame-
work that would approach the problem in a systematic manner and provide a com-
plete solution in a format that could ensure long-lasting bridges. Individual solutions
to problems that historically have reduced service life or to issues involving main-
tenance plans, retrofit or replacement plans, bridge management, and life-cycle cost
analysis (LCCA) are only components of this systematic framework: they are not the
framework itself. The steps within this framework should start at the design stage and
should provide the owner with complete information for ensuring the serviceability of
the bridge for a specified target service life. It is important for the plan to be transpar-
ent and identify the challenges for the period of specified service life at the design stage
so that the owner will encounter no surprises.
1.2
Objectives of the Guide
The main objective of the Guide is to provide information about and define procedures
for systematically designing for service life and durability for both new and existing
bridges. The cost of addressing service life issues at the design stage is significantly
lower than taking maintenance and preservation actions while the bridge is in service.
5

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


1.3
Bridge Service Life Terminology and Relationships
The following sections define service life–related terms and describe the relationships
used in the Guide.
1.3.1
Service Life and Design Life
Service life. The time duration during which the bridge element, component, subsys-
tem, or system provides the desired level of performance or functionality, with any
required level of repair or maintenance.
Target design service life. The time duration during which the bridge element,
component, subsystem, and system is expected to provide the desired function with a
specified level of maintenance established at the design or retrofit stage.
Design life. The period of time on which the statistical derivation of transient loads
is based: 75 years for the current version of AASHTO LRFD Bridge Design Specifica-
tions (2012), referred to throughout the Guide as LRFD specifications.
1.3.2
Bridge Element, Component, Subsystem, and System
The term bridge subsystem is introduced by the Guide. The terms bridge element,
component, and system are the same as those defined by the Federal Highway Admin-
istration (FHWA) National Bridge Inventory.
Bridge element. Individual bridge members such as a girder, floor beam, stringer,
cap, bearing, expansion joint, railing, and so forth. Combined, these elements form
subsystems and components, which then constitute a bridge system.
Bridge component. A combination of bridge elements forming one of the three
major portions of a bridge that makes up the entire structure. The three major compo-
nents of a bridge system are substructure, superstructure, and deck.
Bridge subsystem. A combination of two or more bridge elements acting together
to serve a common structural purpose, such as a composite girder, which could consist
of girder, reinforcement, and concrete.
Bridge system. The three major components of the bridge (deck, substructure, and
superstructure) combined to form a complete bridge.
1.3.3
Service and Design Life: Basic Relationships
Several basic relationships exist between the service lives of bridge components, ele­
ments, subsystems and systems, and bridge design life. These relationships are d
­ escribed
as follows:

• Predicting the service life of bridge systems is accomplished by predicting the ser-
vice life of its elements, components, or subsystems.
• The design life of a bridge system is a target life in years that is set at the initial
design stage and specified by the bridge owner.
• The service life of a given bridge element, component, subsystem, or system could
be more than the target design service life of the bridge system.

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


• The end of service life for a bridge element, component, or subsystem does not
necessarily signify the end of bridge system service life as long as the bridge ele-
ment, component, or subsystem could be replaced or resume its function with a
retrofit.
• A given bridge element, component, or subsystem could be replaced or retrofitted,
allowing the bridge as a system to continue providing the desired function.
• The service life of a bridge element, component, or subsystem ends when it is no
longer economical or feasible to repair or retrofit it, and replacement is the only
remaining option.
• The service life of a bridge system ends when it is not possible to replace or retrofit
one or more of its components, elements, or subsystems economically or because
of other considerations.
• The service life of a bridge system is governed by the service life of its critical ele-
ments, components, and subsystems. The critical bridge elements, components, or
subsystems are defined as those needed for the bridge as a system to provide its
intended function.

In general, the service life (ts) of the bridge elements, components, and subsystems
should be equal to or greater than the design life (tD) of the bridge system, as defined
by Equation 1.1:

(ts)C, E, SS ≥ (tD)BS (1.1)

where
(ts)C, E, SS = service life of bridge component (C), bridge element (E), or bridge
subsystem (SS); and
(tD)BS = design life of bridge system (BS).

The service life of the bridge system is less than or equal to the service life of its
governing elements, components, or subsystems, as described by Equation 1.2:

(ts)BS ≤ [(ts)C, E, BS]critical (1.2)

The service life of the bridge system must exceed or be equal to the target design
life of the bridge system, as described by Equation 1.3:

(ts)BS ≥ (tD)BS (1.3)

1.4
Guide Approach to Design for Service Life
The Guide approach to design for service life is to provide a body of knowledge re-
lating to bridge durability under different exposure conditions and constraints and
to establish an array of options capable of enhancing service life. A solution for a

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


particular service life issue is highly dependent on many factors that vary from location
to location and state to state. A solution also depends on local practices and prefer-
ences. Consequently, the Guide is not intended to dictate a unique solution for any
specific service life problem or identify the best and only solution. Rather, it equips the
reader with a body of knowledge for developing specific solutions best suited to stated
conditions and constraints.
In applying the Guide framework to a particular bridge, including long-span
bridges, an array of solutions can be identified for enhancing the service life of a bridge
element, component, or subsystem, and an optimum solution can be identified through
the LCCA. The solutions can be based on data collected by local DOTs or agencies
responsible for maintaining the bridge and, in order to be complete, the LCCA should
include maintenance, retrofit, replacement, and user costs. It is important that the list
of assumptions and feasible solutions considered for a particular bridge element, com-
ponent, and subsystem be communicated and shared with the owner, especially with
respect to the LCCA, so that the entire process is fully transparent.
The Guide recognizes that not all bridges can or need to have 100 years of service
life. Therefore maintenance, rehabilitation, and replacement are part of the service life
design process. The Guide provides the general framework to achieve this objective in
a systematic manner that considers the entire bridge system and all project demands.
Enhancing the service life of existing and new bridges can be achieved in differ-
ent ways. Two examples include using improved, more durable materials and systems
during original construction that will require minimal maintenance and improving
techniques and optimizing the timing of interventions such as preventive maintenance
actions. Interventions can be planned and carried out based on the assessment of indi-
vidual bridge conditions and needs, or they can be based on a program of preventive
maintenance actions planned for similar elements on a group of bridges. A simple
example of a preventive, planned maintenance program might include the following
activities:

• Washing deicing salts off bridge decks in the spring;


• Cleaning debris from bridge-deck expansion joints;
• Cleaning debris from bearings and truss joints;
• Cleaning drainage outlets;
• Spot painting steel structures;
• Sealing decks or superstructures in marine environments; and
• Sealing substructures on overpasses where deicing salts are used on the roadways
below.

By acknowledging that service life can be extended by either using more durable,
deterioration-resistant materials or by planned intervention, a cost comparison can be
made to determine the most cost-effective approach for various environmental expo-
sure levels and various levels of available maintenance and preservation actions.

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


The following sections provide an overview of the general approach used in the
Guide. The discussion is customized for new bridges; however, the Guide approach
can be used for existing bridges by eliminating some of the steps used for new bridges.
Because the discussions are general and use very simple examples to demonstrate the
point of discussion, many intermediate steps are eliminated for the sake of clarity.
More detailed procedures and examples are provided in subsequent chapters of the
Guide.
Three related flowcharts (Figures 1.4, 1.5, and 1.6) are used to demonstrate the
general approach used in the Guide. Blocks within each flowchart are numbered and
described under the corresponding step number. For example, Step 1 corresponds to
Block Number 1 in Figure 1.4. After each flowchart, a brief discussion explains the
intent of each block within that flowchart.
Customization of the framework introduced in the Guide for a particular bridge
could be achieved by developing similar flowcharts, making each step of the process
transparent to the owner. For major and complex bridges, various elements of the
flowcharts need to reflect specific project requirements.

Step 1. The design for service life starts by considering all project demands set by
the owner, including the service life requirements, as stated in Figure 1.4. Chapter 2
provides examples of local operational and site requirements, as well as service life
considerations needing attention.
Step 2. All feasible and preliminary bridge alternatives that satisfy project demands
should be developed. For example, one might want to consider steel, concrete, and seg-
mental bridge alternatives for a particular bridge. The development of the potential
bridge system is carried out in a conventional manner, meeting all the provisions of
the LRFD specifications. It is good practice to consider potential service life problems,
even at this stage of the design process. It is also feasible to use bridge technologies that
do not have a specific design guideline within the LRFD specifications. In such cases,
the best available design approach could be used, subject to owner approval.
Steps 3 and 4. The next steps in the process consist of evaluating each bridge sys-
tem alternative one at a time and considering service life issues related to each element,
component, and subsystem of that bridge system. For each bridge element, compo-
nent, and subsystem, the Guide provides a framework for incorporating the changes
and modifications needed to meet service life requirements.
For example, assume that one of the alternative bridge systems to be considered
for a particular project is a steel bridge. The designer will first develop the prelimi-
nary bridge configurations by using the conventional approaches that meet all LRFD
specifications. Using procedures depicted in Blocks 4a, 4b, and 4c, each element, com-
ponent, or subsystem of the steel alternative will then be checked against the service
life requirements by using the fault tree approach described later in the Guide. These
evaluation requirements may lead to changes in the details of the element, component,
or subsystem under consideration. For example, the preliminary deck configuration
may indicate that 8-in.-thick concrete is sufficient from a strength standpoint. Going
through the fault tree corresponding to bridge deck and described in Chapter 4 of the

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


General Steps in
Design for Service Life

1. Identify the job and service life requirements.

2. Identify feasible bridge system alternatives that


satisfy design provisions of LRFD specifications.

3. Evaluate all components, elements, and


subsystems of the selected bridge system
C
alternatives against service life requirements in the
Guide stated in various Guide chapters.

Yes 4b. Identify mitigation


4a. Does procedure and incorporate
specific service changes to bridge
life apply? configurations.

No

No 4c. Are all


service life
requirements
considered?

Yes

Go to A

Figure 1.4.  General flowchart demonstrating the Guide’s approach for service life design.

10

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Guide, the designer may change the deck thickness to 9 in. to address potential over-
loads, or may specify sealing the bottom of the deck to protect it from salt spray, if the
bridge is located along the coastline. For major and complex bridges, most of these
fault trees must be customized to meet specific needs and preferred practices. Examples
of fault trees and how they work are provided in later sections of this chapter.
Steps 5 through 8. At the end of Step 4 and after going through appropriate fault
trees for various bridge elements, components, and subsystems, the designer will have
developed a bridge system that meets both strength and service life requirements, as
illustrated by Step 5 in Figure 1.5. To some extent, changes to configurations of various

5. Develop a modified bridge system that meets


both LRFD specifications and service life
requirements.

6. Determine if the draft configuration of the


selected bridge alternative satisfies requirements
and that the incorporated changes are compatible.

7b. Make appropriate modifications in


7a. Is No components, subsystems, or elements
configuration for compatibility.
okay?

Yes
8. Develop final configuration
of the selected bridge
alternative.

Go to B

Figure 1.5.  General flowchart demonstrating the Guide’s approach for service life design,
starting with A from Figure 1.4.

11

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


bridge elements, components, and subsystems are carried out separately. Thus, there
is a need to make sure that these changes are compatible and are not contradictory
or overly conservative. Steps 6 and 7 in Figure 1.5 depict this process. For example,
in the steel bridge example discussed previously, service life requirements may dictate
the use of a jointless, integral abutment system and require metalizing the end of the
girder. The designer may then want to consider not metalizing the end of the girder,
because leaking joints would be eliminated. Finally, for the selected bridge system
alternative under consideration, a final configuration is developed (Step 8) that meets
both strength and service life requirements.

9b. Is service
9a. Predict service life of No
life of the parts 9c. Identify
various components,
greater than rehabilitation or
subsystems, and
service life of replacement plan.
elements.
the system?

Yes
10. Compile all
requirements for bridge 9d. Identify
system alternative and maintenance plan.
compute life-cycle cost.

11b. Consider the No 11a. Are all


next bridge system alternatives
alternative. considered?

Go to C Yes

12. Compare advantages and


disadvantages of all final
alternatives and select final
bridge system.

Figure 1.6. General


Figure flowchart
General flowchart demonstrating
demonstrating the Guide’s
the Guide’s approach
approach for service
for service lifestarting
life design. design.with B from
Figure 1.5.

12

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Steps 9 through 12. The next step in the process is to evaluate the service lives of
the various bridge elements, components, and subsystems of the bridge alternative
under consideration and compare its overall service life with the owner-specified target
service design life of the bridge system. For example, the owner may require that the
bridge provide 100 years of service life, but the life of a particular bridge element, such
as the sliding surface for a bearing, may be limited to 20 years. This situation would
require a plan for timely replacement of the sliding surfaces. Regardless, there will be
a need for a systematic maintenance plan that could require the designer to identify
“hot areas” requiring more detailed inspection and maintenance. Blocks 9a through
9d depict the development of a maintenance plan and/or rehabilitation and replace-
ment plan for the bridge system alternative under consideration. The result of this
process, as illustrated in Step 10, is a bridge system alternative that meets both strength
and service life requirements with an associated maintenance and/or rehabilitation or
replacement plan for the bridge. Step 10 also includes an LCCA to consider the final
configuration of the select bridge alternative and maintenance plan. The same steps are
repeated for all bridge alternative systems, as shown by Step 11. After comparing all
alternatives, the designer can recommend which alternative should be used, allowing
the owner to make the final selection.
As described, the selection of the final bridge system within the framework pro-
moted in the Guide is mainly based on service life requirements. Some of the details
included in the steps presented, such as fault tree analysis, will be described in later
sections of this chapter.
A summary of steps for design for service life is provided in Section 1.10 of this
chapter.
1.5
Organization of the Guide
Included in the Guide are 11 chapters, each devoted to particular bridge elements,
components, subsystems, or systems. The following is a brief description of informa-
tion included in each chapter.
Chapter 1. Design for Service Life: General Framework. This chapter provides an
overview of the approach used in the Guide for design for service life and describes ter-
minology used throughout the Guide and various relationships that exist between the
service life of bridge element, component, subsystem, and system and bridge design life
as used in AASHTO specifications. The chapter introduces the different philosophies
used to predict service life.
Chapter 2. Bridge System Selection. This chapter describes various bridge systems
and factors that affect their service life. Included is a description of a general strategy
and rational procedure for selecting the optimum bridge system, subsystems, compo-
nents, and elements that consider specific project limitations and requirements, such as
climate, traffic, usage, and importance.
Chapter 3. Materials. This chapter provides general properties and durability
characteristics of the two most commonly used materials in bridge systems, steel and
concrete. For each material, a general description of variables affecting the service life
is provided, followed by strategies used to mitigate them. Chapter 3 comprises the
13

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


basic information for materials used in bridge subsystems and elements specifically
addressed in other chapters of the Guide.
Chapter 4. Bridge Decks. This chapter describes various bridge-deck types and
essential information related to their service life, such as modes of deterioration and
strategies to mitigate them. The chapter concentrates on cast-in-place and precast con-
crete bridge decks.
Chapter 5. Corrosion of Steel in Reinforced Concrete Bridges. This chapter looks
at basic mechanisms causing corrosion of reinforcement embedded in concrete and
provides strategies for preventing corrosion of reinforcement in concrete bridges.
Chapter 6. Corrosion Prevention of Steel Bridges. This chapter describes various
coating systems using paint, galvanizing and metalizing, and corrosion-resistant steel,
along with factors affecting service life. Various options for preventing corrosion of
steel bridges and general approaches that could lead to bridge coatings with enhanced
service life are presented.
Chapter 7. Fatigue and Fracture of Steel Structures. This chapter provides the
basics of fatigue and fracture and the factors that cause fatigue and fracture in steel
bridges. Various available options for repairing observed cracking in steel bridges are
also presented.
Chapter 8. Jointless Bridges. This chapter describes various jointless bridge sys-
tems, considers their advantages and disadvantages, and provides complete steps for
the design of jointless integral abutment bridges. Design procedures to extend the
application of jointless integral bridges to curved girder bridges are provided. Also
introduced are new details and integral abutment systems, in which expansion joints
are completely eliminated, even at the end of approach slabs.
Chapter 9. Expansion Devices. The Guide encourages eliminating the use of
expansion joints. However, expansion joints may be needed when the total bridge
length exceeds the practical limits of jointless bridges. This chapter describes various
expansion joints used in practice, observed modes of failure for each, and potential
strategies to mitigate those failures.
Chapter 10. Bridge Bearings. This chapter describes various bearing types, lists
the factors that affect the service life of the various bearings, and offers strategies to
mitigate such factors. New high-performing sliding surfaces capable of providing long
service life are introduced, as well as deterioration models for sliding surfaces. The
Guide emphasizes use of elastomeric bearing pads.
Chapter 11. Life-Cycle Cost Analysis. This chapter provides essential information
for incorporating LCCA into bridge system, subsystem, component, and element selec-
tion. It concentrates on general features and elements of incorporating LCCA into the
design process, emphasizing consideration of project costs throughout the service life.
1.6
Categories of Information Provided in Guide Chapters
Typically, each chapter consists of the five major categories of information described
in this section. Closer examination of the type of data included in each chapter could
also assist in developing customized information for addressing design for service life
for major and complex bridges.
14

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


1.6.1
Description of Bridge Elements, Components, and Systems
These sections of each chapter provide brief descriptions of, and essential information
related to, both commonly used and more recently developed types of bridge compo-
nents, elements, subsystems, and systems.
1.6.2
Factors That Affect Service Life
The factors affecting service life are identified using a fault tree approach, which pro-
vides a systematic method of identifying factors that can affect the service life of a
particular bridge element, component, or subsystem in various categories and suc-
cessive subcategories. Most chapters have fault trees applicable for the types of ele-
ments, components, or subsystems covered within that chapter. In the case of major
and complex bridges, designers should develop customized fault trees that reflect the
specifics associated with location and traffic conditions. A customized fault tree can be
developed using data and experiences available from local agencies.
The fault tree starts with the identification of major factors that can reduce the
service life of a particular bridge element, component, or subsystem. Each major factor
can then be broken down into more detailed subcomponents, each capable of reducing
the service life. The fault tree continues branching until each branch ends with factors
at the lowest or base levels of influence. The factors with subcomponents are placed
inside rectangles, and the identified lowest or base factors are placed inside circles.
Figure 1.7 shows a portion of the fault tree used in Chapter 4 for a bridge deck.
In Figure 1.7, either of two main factors (obsolescence or deficiency) is shown
to be capable of contributing to reduced service life of a bridge deck. The elliptical
symbol just above these two factors is referred to as an “or gate,” which signifies that
either one of the factors below it could result in reduced service life. The fault tree
shown in Figure 1.7 continues to list the major categories of factors (i.e., those related
to induced loads, natural or man-made hazards, and production or operation defects)
that could result in reduced service life of a bridge deck.
Figure 1.8 shows the continuation of the fault tree for a breakdown of factors
related to load-induced factors for a bridge deck.
In Figure 1.8, load-induced factors are subcategorized into traffic-induced loads or
loads induced by system-dependent load factors, such as restraints provided by shear
studs. Two factors are further divided into subcomponents, each capable of reducing
the service life of a bridge deck. The factors inside the circles are the basic factors
without any further subcomponent. They represent the end of that branch of the fault
tree and require the development of individual strategies to mitigate them. This aspect
of the process is described later.
In the Guide, each element of the fault tree is described immediately after intro-
ducing each branch of the fault tree. It is advisable to do the same when develop-
ing customized fault trees for major and complex bridges. Documentation of factors
affecting the service life of bridge elements, components, or subsystems in the form of
a fault tree should be part of the overall plan for design for service life and provided to
the owner for future reference.

15

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Reduced Service Life of
Cast-in-Place Bridge Deck

Caused by
Caused by Deficiency
Obsolescence

Natural or Man-Made Production/


Load-Induced
Hazards Operation Defects

Figure 1.7.  Starting point for fault tree for a bridge deck.

1.6.3
Mitigation Strategies
When possible, each chapter provides provable solutions for major factors affecting
the service life of a particular bridge element, component, subsystem, or system. Some
chapters also include technology tables that summarize major characteristics associ-
ated with each solution and provide the potential solutions to factors affecting service
life in a form that is easier to comprehend. For example, Figure 1.9 shows a technol-
ogy table that summarizes solutions for enhancing the service life of bridge decks; it
is related to traffic-induced loads as shown in Figure 1.8. In Figure 1.8, as part of the
fault tree for a bridge deck, traffic-induced loads are identified as one factor capable
of reducing service life. In Figure 1.8, below traffic-induced loads, three basic factors
capable of reducing bridge-deck service life are identified: fatigue, overload, and wear
and abrasion. Each basic factor needs to be mitigated using a select strategy, and in
almost all cases, there is more than one strategy to mitigate these basic factors. It is
good practice to collect these strategies in table form and select the optimal strat-
egy, considering its interaction with other parts of the bridge. The technology tables
16

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Load-Induced

Traffic-Induced System-
Loads Dependent Loads

System-
Differential
Fatigue Overload Wear Thermal Framing
Shrinkage
Restraint

Figure 1.8. Continuation


Figure 1.8 showsofthe
the continuation
fault tree for a of
breakdown
the faultoftree
factors
for arelated to load-induced
breakdown of factors related to load-induced facto
factors for a bridge deck.

Service Life
Issue Mitigating Strategy Advantage Disadvantage
Fatigue Design per LRFD specifications Minimizes the possibility of May increase the area of steel
reinforcement failure
Overload Increase deck thickness Minimizes cracking Adds weight to bridge
structure, increases cost
Wear and Implement concrete mix See Chapter 3, Materials See Chapter 3, Materials
abrasion design strategies
Implement membranes and Protects surface from direct Requires rehabilitation every
overlays contact with tires 10 to 20 years

Figure 1.9.  Technology table for mitigating factors related to traffic-induced loads
affecting the service life of bridge decks.

provided in various chapters of the Guide summarize strategies that can be used to
mitigate various basic factors capable of reducing service life. For major and complex
bridges, the list of strategies could be different and based on local preferences and
experi­ences. Most agencies have access to field data collected over the years that could
be used to construct customized strategy tables for the purpose of mitigating basic fac-
tors c­ apable of reducing service life of bridge elements, components, and subsystems.
17

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


The sample table in Figure 1.9 lists the advantages and disadvantages for each
possible solution capable of mitigating the adverse service life consequences of traffic-
induced loads. In some cases, more information than just advantages and disadvan-
tages is provided, such as qualitative assessment of maintenance cost. For major and
complex bridges additional considerations may be included in technology tables.
1.6.4
Optimum Selection Strategies
Overall strategies are provided for achieving enhanced service life. The overall strategy
approach provided depends on the particular bridge component, element, subsystem,
or system. Figure 1.10 shows an example taken from Chapter 10 of the Guide of
the overall strategy for selecting a bearing that meets both strength and service life
requirements.

Chapter 10, Bridge Bearings, identifies factors affecting the service life of bearings
and provides potential solutions for each. This information, combined with the steps
outlined in the flowchart, can be used as a rational approach for selecting an appropri-
ate bearing that meets project requirements with emphasis on service life.
1.6.5
Examples and Tools
Most chapters include examples demonstrating the application of strategies in that
chapter.
1.7
Quantifying Service Life of Bridge Element, Component,
Subsystem, and System
One of the important steps in developing a systematic, comprehensive service life d­esign
plan for bridges is the capability of predicting the expected service life of various bridge
elements, components, and subsystems, which in turn will dictate the service life of the
bridge system. This process is Step 9a in Figure 1.6. Service life prediction capability is
important for developing maintenance, retrofit, and replacement plans, which are an
integral part of the service life design process. The objective of this section is to provide
an overview of the methodology used in the Guide for predicting service life.
Bridge elements, components, subsystems, and systems are subject to the effects of
traffic and the environment. These external sources of deterioration act through vari-
ous mechanisms to cause actual deterioration and eventually failure of bridge elements.
The mechanisms of deterioration are the physical laws that govern such deterioration.
Deterioration rates can be described using mathematical expressions or empirical and
semiempirical models, which are developed using data collected by field monitoring
of bridges, laboratory-generated data, expert opinions, or combinations of available
data. Service life is also affected by risk to damage from either traffic or extreme envi-
ronmental occurrences. The acceptability of this damage is evaluated based on risk.
Service life can be extended by minimizing risk or designing for appropriate levels of
extreme occurrences.

18

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Identify DEMAND requirements.

Compare with SUPPLY parameters for


bearing alternatives.

Identify potential bearing alternative(s)


that accommodate loads and
movements.
No

Perform preliminary design to confirm Is bearing


suitability of potential bearing type. suitable?

Yes
Identify factors affecting service life
and mitigation requirements.

Evaluate potential service life of


Is bearing No
alternatives and determine if bearing
has potential for achieving desired life. suitable?

Yes

Compare life-cycle costs considering Determine expected


initial, maintenance, and replacement frequency of bearing
costs. replacement.

Perform final bearing selection.

Figure 1.10.  Overall strategy for bearing design considering service life.

Enhanced service life for bridge elements, components, subsystems, and systems
can be achieved through

• Use of durable materials;


• Use of passive or active protection systems;

19

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


• Optimum selection of details;
• Optimum maintenance and repair;
• Reduced service level;
• Increased factor of safety or reduction in stress levels; and
• Isolation from risk damage.

To estimate the service life of bridge elements, components, or subsystems quanti-


tatively, the following information is needed:

• Source of deterioration;
• Deterioration mechanism;
• Deterioration models; and
• Failure modes.

The following sections provide information on each of these items of information.


1.7.1
Source of Deterioration
Traffic-related or environmental effects are the basic external causes of deterioration.
For example, deicing compounds, an external source of deterioration, can result in
corrosion of reinforcement in bridge elements.
1.7.2
Deterioration Mechanism
Deterioration is governed by a process called the deterioration mechanism. For ex-
ample, sliding surfaces in bearings experience deterioration through horizontal move-
ment and friction between sliding materials created by truck passages or temperature
fluctuations. The horizontal movement and friction in this instance is the deterioration
mechanism. In the case of concrete elements, ingress of chloride through concrete
causes initiation of corrosion in unprotected steel reinforcement. In this instance, the
ingress of chloride is the deterioration mechanism.
1.7.3
Deterioration Models
Deterioration models are used to describe the rate of deterioration. They describe the
relationship between the condition of the bridge (or its element) and its time of use
and show how the bridge deteriorates over time. A deterioration model assumes that
no replacements or major repairs are made, but it usually implies that scheduled main-
tenance actions are performed as planned. The basic model applies either to a bridge
system as a whole or to any of its subsystems, components, or elements.
An example of a deterioration curve is presented in Figure 1.11. For a bridge
placed in service at period T0, the deterioration curve represents the gradually declin-
ing condition of the bridge over time. Initially the condition is good, but after a period
of wear and aging, it eventually (at time Tf) reaches an unacceptably low condition
(Cf). The time period between T0 and Tf is called the service life of the bridge.

20

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Bridge Deterioration

Condition

Service Life
Cf

T0 Time of Use Tf

Figure 1.11.  An example of a bridge deterioration curve.

In practice, the development of realistic behavioral deterioration models is a data-


intensive process complicated by lack of knowledge of the underlying physical and
chemical processes fostering deterioration, as well as by data availability. Currently
available deterioration models, which are based on long-term data collection, are
very limited. To complicate the issue, the quality of bridge design and construction
improves over time. As a result, application of data collected from existing bridges to
predict the performance of future bridges should be practiced with caution.
Deterioration models capable of quantitatively predicting the service life of bridge
elements, components, subsystems, or systems are very limited or nonexistent. The
most acceptable deterioration model is in the form of the solution to Fick’s second law,
which is used to predict the rate of chloride ingress through concrete cover. This model,
including its limitations, is described in Chapter 5 of the Guide. It is expected that with
time, more deterioration models will become available and will greatly enhance quan-
tification of the service life of bridge elements, components, subsystems, or systems.
As shown on Figure 1.12, if left alone a bridge will deteriorate over the period of
its service life. However, in most cases a bridge is not left to follow the basic deterio-
ration path and reach an unacceptable condition without interruption. The agency
responsible for the bridge will from time to time undertake repairs, rehabilitations,
and renewals that return conditions to higher levels and extend its service life. During
these interventions, the condition of the bridge improves, as depicted in Figure 1.12.
Deterioration models can also be based on some level of understanding of the
mechanism governing the deterioration and the capability of expressing the process
through a mathematical expression. An example is deterioration of concrete elements
caused by chloride-induced corrosion of reinforcement. The assumption is that ingress
of chloride through the concrete element is governed by Fick’s second law, which
assumes a homogeneous material.

21

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Bridge Condition

Service Life

Condition

Cf

T0 Time of use Tf

Figure 1.12.  Bridge condition life cycle.

In the case of chloride- and carbonation-induced corrosion, there is some level of


agreement within the scientific community as to the existence of deterioration models.
However, for other deterioration modes, such as sulfate attack, alkali-silica reactiv-
ity (ASR), and freeze–thaw or wear and abrasion, there is a lack of adequate models.
Further, as described previously, the use of deterioration models to predict the time to
initiate corrosion of reinforcement embedded within certain distances of the concrete
surface because of chloride ingress should be approached with caution. The following
paragraphs briefly describe a deterioration model for chloride-induced corrosion.
There are different approaches to solving Fick’s second law. A finite-difference
approach, or the use of error functions, is reported in the published literature. Equa-
tion 1.4 is an error function solution of Fick’s second law that is capable of predicting
the chloride concentration level at various depths within the concrete element.
 a − ∆x 
( )
Ccrit = C ( x, t ) = C0 + CS, ∆x − C0 . 1 − erf

 (1.4)
2. Dapp,C t 

where
Ccrit = critical chloride content (wt.-%/c),
C(x,t) = content of chlorides (wt.-%/c) in the concrete at depth x
(structure surface: x = 0 m) and time t,
C0 = initial chloride content (wt.-%/c) of the concrete,
CS,Dx = chloride content (wt.-%/c) at depth Δx and certain point of time t,
x = depth (mm) with a corresponding content of chlorides [C(x,t)],
a = concrete cover (mm),
Δx = depth of convection zone (concrete layer, up to which the process of
chloride penetration differs from Fick’s second law of diffusion) (mm),
Dapp,C = apparent coefficient of chloride diffusion through concrete (mm2/years),
t = time (years), and
erf = error function.
22

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Equation 1.4 should be used in conjunction with probabilistic approaches to
account for the variability of several parameters, such as apparent coefficient of dif-
fusion, chloride concentration, and critical chloride level to start corrosion. Further-
more, the diffusivity of concrete through different layers of the concrete element is not
uniform. Equation 1.4 predicts the chloride content in the structure at a given depth x
and time t. This number is given by the left-hand side of the equation, C(x,t).
The C(x,t) obtained from Equation 1.4 is then compared with the critical chloride
content (Ccrit), which is the value determined to be the point at which corrosion starts.
When the chloride level at a given depth (x) of the structure is reached (i.e., the critical
value), the corrosion is assumed to initiate. The service life of the concrete element can
then be assumed to consist of the time period to initiate corrosion plus the time period
for propagation of the corrosion to the point that will limit the functionality of the
concrete element. This process is depicted in Figure 1.13.
1.7.4
Failure Modes
Sources of deterioration (such as deicing compounds) acting through deterioration
mechanisms (such as ingress of chloride through concrete cover) and described by
deterioration models (such as a solution to Fick’s second law) result in failure modes
(corrosion of reinforcement, causing corrosion-induced cracking and loss of strength).
The final failure could consist of several stages, such as start and propagation phases.
1.7.5
Service Life Estimation
In the Guide, two general philosophies are presented to estimate the service life of
bridge elements, components, and subsystems. Quantifying the service life of bridge ele-
ments, components, or subsystems establishes ts, the design service life of bridge ele-
ments, components, or subsystems, which can then be compared with the specified
service life of the bridge system to determine whether retrofit or replacement strategies
are needed.

Figure 1.13.  Relationship between


­damage and service life.
Source: Edvardsen 2008.

23

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Two general design approaches for service life are the finite service life approach
and the target service life approach.
When the design service life (ts) of the bridge element, component, or subsystem
established through one of the two design approaches for service life philosophies is
less than the specified design life of the bridge system (tD), the bridge element, com-
ponent, or subsystem under consideration could be replaced to achieve the specified
design life of the bridge system.
The major difference between the two approaches for service life design is the need
for well-accepted deterioration models in the finite service life approach.
1.7.5.1
Finite Service Life Approach
Bridge elements, components, and subsystems designed using the finite service life ap-
proach should have a service life that is greater than or equal to the specified bridge
system service life. Otherwise, the bridge element, component, or subsystem under
consideration must be retrofitted or replaced to allow the bridge to continue provid-
ing its intended function until reaching the specified bridge system service life. In the
finite service life approach, the service life of the bridge components, elements, or
sub­systems is estimated using well-accepted deterioration models. The existence of
deterioration models is therefore essential for using the finite service life approach.
The deterioration models are generally developed using one of the following
approaches:

• Mathematical models that describe the deterioration rate. These models could be
approximate or based on the laws of physics;
• Empirical or semiempirical models developed using data collected from laboratory
or field performance of bridges. Fatigue models used in the LRFD specifications
are examples of empirical deterioration models; or
• Empirical models based on expert opinions or experiences. Examples include vari-
ous models used in Pontis.

When deterioration models exist, the service life design could be in the form of a
full probabilistic approach or a semiprobabilistic or partial load factor approach. The
full probabilistic approach requires having probability distribution functions for all
variables used in the deterioration model. The semiprobabilistic or partial load factor
approach is developed using the full probabilistic approach. It is equivalent to using
load and resistance factors in the LRFD specifications versus using the full p
­ robabilistic
approach (such as using Monte Carlo simulation) to design or rate bridges.
1.7.5.2
Target Service Life Approach
In many instances deterioration models are not available, or their applicability is ques-
tionable. In these situations, available alternatives are (1) the use of high-­performing
material that does not deteriorate, such as stainless steel (this approach is generally
referred to as the avoidance of deterioration method within European practice) or
(2) the use of material that, based on experience or expert opinion, could provide

24

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


a specified or target service life. If the estimated service life of the bridge element,
component, or subsystem is less than the specified service life of the bridge, retrofit
or replacement strategies must be specified, allowing the bridge system to continue
providing its intended function.
The major difference between the finite service life and target service life design
approaches is that in the finite service life design approach, the condition of the bridge
element, component, and subsystem can be traced over time by using deterioration
models; in contrast, in the target service life design approach, only the total expected
service life is estimated. The specified target service life of the bridge element, compo-
nent, or subsystem is mainly established on the basis of experience or expert opinion
and could vary significantly from assumed values. Nevertheless, specifying a target ser-
vice life for a given bridge element, component, or subsystem allows the bridge owner
to plan and anticipate necessary maintenance actions and places demands on the
designer to incorporate necessary design features as needed. For example, in response
to an assumed service life of about 10 years (i.e., target service life of 10 years) for
polytetrafluorethylene sliding surfaces in bearing devices, the designer must incorpo-
rate necessary mechanisms to lift the bridge and replace the sliding surfaces, prefer-
ably while maintaining traffic, and the bridge owner must plan for and anticipate the
replacement of sliding surfaces every 10 years.
1.8
Owner’s Manual
When specified by the owner and for major and complex bridges, the final step in the
design for service life process is the development of a bridge Owner’s Manual, which
summarizes the processes used for the design for service life and provides complete
descriptions of outcomes and recommendations. The Owner’s Manual is intended to
equip the owner with the necessary knowledge to keep the bridge operational for the
specified service life period. It should be provided to the owner at the time of opening
the bridge to traffic, after an independent review process as described in Section 1.9.
The entire process used for design for service life should be well documented and
include assumptions, limitations, and any other information about which the owner
should be aware, including complete information with respect to “hot spots” within
various bridge elements, components, or subsystems that will require closer inspection,
maintenance, retrofit, or replacement. The Owner’s Manual should include a complete
management plan with respect to service life, including information on timely mainte-
nance actions, and identify replacement items and methodologies for replacement with
information on the required level of traffic interruption, if any. In the case of major
and complex bridges it is suggested that a bridge instrumentation and monitoring plan
be developed and tied to the bridge service life management plan. Additional informa-
tion to be incorporated into the Owner’s Manual after construction should include the
actual material properties of critical bridge elements versus the assumed values used in
the design process. Such information is important for future bridge rating.
For major and complex bridges, the designer should use sound engineering judg-
ment for determining the level and extent of information to be included in the Owner’s
Manual. The bridge Owner’s Manual is analogous to the design calculations that are
25

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


customarily provided to the bridge owner, except that the Owner’s Manual contains
much more detailed information.
1.9
Independent Review of Design for Service Life Process
The design for service life processes, results, and recommendations as summarized in
the bridge Owner’s Manual should be checked by an independent and knowledgeable
third party. This independent check is analogous to an independent design check typi-
cally conducted for bridge design.
1.10
Summary of Steps for Design for Service Life for Specific
Bridge Element, Component, and Subsystem
This section summarizes the steps in the design for service life. Detailed descriptions of
individual steps are provided in Section 1.4.
Bridge elements, components, and subsystems can deteriorate at different rates
and have different service lives. This variable deterioration governs the service life of a
bridge system, which is reached when the service life of critical bridge elements, com-
ponents, or subsystems is exhausted beyond being repaired or replaced economically
or because of other considerations.
The general steps in design for service life for a particular bridge element, compo-
nent, or subsystem can be summarized as follows:

Step 1. Identify the project requirements, particularly those that will influence the
service life.
Step 2. Identify feasible bridge systems capable of meeting the project demand.
Step 3. Select each feasible bridge system one at a time and complete Steps 4
through 10.
Step 4. Identify the factors that influence the service life of bridge elements, com-
ponents, and subsystems, such as traffic and environmental factors.
Step 5. Identify modes of failure and consequences (e.g., the corrosion of rein-
forcement that causes corrosion-induced cracking and loss of strength).
Step 6. Identify suitable approaches for mitigating the failure modes or assessing
risk of damage through LCCA (e.g., use of better-performing materials for sliding sur-
faces in bearings or use of material prone to deterioration at lower initial cost).
Step 7. Modify the bridge element, component, or subsystem under consideration
by using the selected strategy and ensure compatibility of different strategies used for
various bridge elements, components, or subsystems. This step may involve the need
to develop several alternatives.
Step 8. For each modified alternative, estimate the service life of the bridge ele-
ment, component, or subsystem using the finite or target service life design approaches.
Step 9. For each modified alternative, compare the service life of the bridge ele-
ment, component, or subsystem with the service life of the bridge system and develop
appropriate maintenance, retrofit, and/or replacement plans.

26

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Step 10. For each modified alternative, develop design, fabrication, construction,
operation, maintenance, replacement, and management plans for achieving the speci-
fied design life for the bridge system.
Step 11. For each modified alternative, conduct LCCA for each feasible bridge
system meeting strength and service life requirements and select the optimum bridge
system.
Step 12. When specified by the owner or in cases of major and complex bridges,
document the entire design for service life processes in an Owner’s Manual. Arrange
for an independent review of the document and provide it to the bridge owner at the
time of opening the bridge to traffic.
1.11
Approaches to Using the Guide
This section provides a limited example demonstrating the use of the Guide and how
to implement systematic approaches for design for service life. The example is not
inclusive and considers an isolated component of the bridge without considering the
remaining bridge elements, components, or subsystems. Further, the example, for the
sake of demonstration, uses Life-365, which has limitations when applied to horizon-
tal surfaces, such as bridge decks. Life-365 uses the solution to Fick’s second law to
predict deterioration of concrete elements subjected to chloride ingress. Although this
approach has merits for vertical surfaces, such as columns under compression, its ap-
plicability to horizontal surfaces, such as bridge decks, is not warranted. In the case of
bridge decks, the existence of cracks violates the assumption of a homogeneous mate-
rial in Fick’s second law. The use of Life-365 for the bridge-deck example here is for
demonstration purposes, as it includes LCCA in addition to predicting time to initiate
corrosion and propagation.
The following sections provide an overall description of the bridge used for the
example and illustrate the steps taken in the design for service life for a single isolated
component of the bridge.
1.11.1
Example Bridge Description
The example bridge is a 1,400-ft-long structure carrying four lanes of high-volume
traffic with pedestrian sidewalks and bicycle lanes. The bridge crosses over low-­volume
urban local roads, a railroad, and a navigable waterway. Refer to Figures 1.14 and
1.15 for a rendering of the project concept. Figure 1.16 shows the bridge-deck system
that will be used for this example.
The following characteristics of the bridge setting influence the service life:

• Located in a cold environment where deicing salts are used and multiple freeze–
thaw cycles are anticipated;
• Located in an area where studded tires are used in the winter;
• Subjected to potential overloads with 20-kip tire loads in an HL 93 truck
configuration;

27

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Figure 1.14.  Aerial conception of bridge project.

Figure 1.15.  Typical superstructure and substructure configuration.

28

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 1.16.  Superstructure section of bridge deck with cast-in-place option designed
according to LRFD specifications.

• Spans over a navigable waterway with primarily brackish conditions, and located
adjacent to a park with water access for jet skis;
• Located near the coastline with possible salt water storm surge and potential hur-
ricane force winds with gusts up to 150 mph; and
• Located in an area where local aggregates are subject to ASR.

1.11.2
Steps in Addressing Service Life Design
The first step in the process for addressing the service life design issue for a bridge
deck is to follow the flowchart shown in Figure 1.17; this flowchart also appears as
Figure 4.18 in Chapter 4 of the Guide.
Table 1.1 is used to extract the information needed to address the requirements of
Steps 1a and 1b shown in Figure 1.17.
The information in Table 1.1 was developed from project requirements and exem-
plifies the type of information necessary for layout and service life evaluation of the
entire bridge system. It is well beyond that information needed for simply evaluating a
bridge deck, but it is provided here for completeness. The issues in Table 1.1 pertinent
to the bridge deck are indicated by an arrow at the right side of the table.
The next step is to identify the possible bridge-deck alternatives (Step 2 in Fig-
ure 1.17). Information provided in Section 2.3.2 of Chapter 2 and Section 4.2.1 of
Chapter 4 of the Guide can be used to obtain advantages and disadvantages of vari-
ous deck systems. In the case of major and complex bridges the designer may consider
feasible bridge-deck systems on the basis of local preferences and experiences. In the
29

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Bridge-Deck System Component
Selection Process

1a. Identify local operational and 1b. Identify local factors affecting
site requirements. service life.

2. Identify feasible deck alternatives satisfying design


provisions of LRFD specifications, operational, site, and
bridge system requirements.

3. For each alternative, identify factors affecting service life


by following fault tree.

4.
Go to
A

B
8a. Go to the
next alternative.

5. Is deck
service life No 5a. Identify rehabilitation
greater than or
or replacement
equal to the
requirements.
system
TDSL?
Yes

6. Identify maintenance requirements.

7. Develop life-cycle costs.

Yes
8. Additional deck
alternative?

No

9. Compare alternatives and select deck system.

Figure 1.17.  Bridge-deck system component selection process. For steps from A to B,
see Figure 1.18. (TDSL means target design service life.)

30

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 1.1.  Operational and Local Factors to Be Considered: Steps 1a and 1b in Figure 1.17
Operational Category Operational Criteria to Be Specified
Traffic capacity requirements Urban arterial, four lanes, 40 mph
Traffic volumes and required capacity 24,000 average daily traffic northbound and southbound á
Truck volumes 10% á
Special vehicle uses Overload possible á
Local environment or man-made hazard Maintain two existing lanes á
category
Mixed-use requirements Traffic, pedestrians, bicycle lane
Vehicle loads and special vehicle load HL 93 with typical legal and permit loads á
requirements No special construction loads
Overload with 20-kip tire loads (HL 93 truck configuration)
Studded tires used in winter
   
Service Life Category Service Life Criteria to Be Specified
Identify bridge importance Critical
Identify target design service life 100 years á
Potential for future bridge widening Not applicable
Potential for future widening of crossed Not applicable
roadways
Vertical clearance requirements related Not applicable
to future bridge widening or widening of
crossed roadways
   
Local Site Category Local Site Criteria to Be Specified
Geometry See plan. Overall bridge length: 1,400 ft
Curvature: Curve 1 R = 1,150 ft; Curve 2 R = 1,300 ft, reversing
Cross slope: superelevation transition from +2% to –2%
Skew: 35°
Features crossed Road A: two lane urban, 25 mph
Road B: one lane urban, 25 mph (two lanes temporary)
Railroad: one existing track, four additional future tracks,
potential commuter rail corridor with mixed freight usage
Navigable waterway: U.S. Coast Guard jurisdiction, small vessels
and jet skis
Horizontal clearances Standard clear zones with no barrier protection
No railroad crash walls allowed
No piers in waterway
(continued)

31

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Table 1.1.  Operational and Local Factors to Be Considered: Steps 1a and 1b in Figure 1.17
(­continued)
Local Site Category (continued) Local Site Criteria to Be Specified (continued)
Vertical clearances Road A: 16.5 ft plus 2 ft for raising of tracks (for railroad bridge
replacement)
Road B: 16.5 ft
Railroad: 23.5 ft plus 2 ft for raising of tracks (for railroad bridge
replacement)
Navigable waterway: 15 ft minimum (bank to bank)
Hydraulic or waterway requirements Greatest flood (500-year) elevation 7.50 (storm surge)
Natural bend in channel
Historic erosion of south bank
Navigation requirements No piers in waterway (bank to bank)
Utility issues, carried or crossed Bridge lighting
Relocate underground fiber optic lines
Other physical boundary conditions Maintain access to adjacent park
Geotechnical considerations All foundation types acceptable
Environmental considerations Water in waterway tests with low chlorides, but adjacent
mangroves indicate that water may be brackish at times
Subject to salt water intrusion from storm surge
Access for construction Limited to existing right-of-way and railroad agreement
Aesthetics and sustainability Closed box system required by city (I-girders unacceptable)
   
Local Environmental or Man-Made Local Environmental or Man-Made Hazard Criteria to Be
Hazard Category Specified
Thermal climate Cold climate, solar radiation, Zone 3
Deicing salts used, multiple freeze–thaw cycles, ice flow
Coastal climate Brackish conditions
Chemical climate ASR susceptible
Hydraulic action hazard, flood, or scour 500-year water velocity/discharge: 4.8 fps/11,200 cfs
Wind action hazard Hurricane zone (150 mph), Exposure C
Drainage requirements 50-year storm
Vehicle and vessel collision susceptibility Vehicle collision to be addressed
Vessel collision potential negligible (mostly pleasure craft)
Fire or blast susceptibility Minimal combustible materials on route
Seismic susceptibility Seismic Zone 1
   
Construction Constraints Category Construction Constraints Criteria to Be Specified
Constructability requirements Phasing not required
Construction schedule requirements Accommodate short schedule demands
Special local construction preferences Do not use cast-in-place concrete boxes

Note: Rows identified by á point out items used in developing the design example listed in this section.

32

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


current example, it is assumed that only the cast-in-place option is selected, as indi-
cated in the typical girder cross section shown in Figure 1.16.
Figure 1.18 shows the next steps in identifying factors that affect service life; the
same flowchart appears as Figure 4.19 in Chapter 4 of the Guide.
Figure 1.18 aids in identifying factors that affect the service life of the bridge
deck and in selecting possible strategies capable of mitigating the adverse effects of
these factors. Identifying the factors that affect bridge-deck service life can be accom-
plished using the fault trees provided in Chapter 4, Section 4.2 of the Guide. Navigat-
ing through the fault tree can be simplified by using software. Figure 1.19 shows an
example of what an Excel-based solution could look like. Using the software shown in
Figure 1.19, the user selects applicable factors from the first fault tree layer and then
continues through successive layers until reaching the last levels, which are depicted as
circles. The items listed in each circle are the factors that will have to be addressed in

4.
A

1A. Identify individual factors affecting service life


considering each branch of fault tree.

Yes 2A.a. Identify consequences and


3A.a. Go to the
2A. Does factor apply? determine appropriate strategies
next factor.
for avoidance or mitigation.

No

No
3A. Are all factors 2A.b. Modify bridge deck
considered? configuration.

Yes

4A. Modified bridge deck configuration for deck alternative under


consideration.

Go To
B

Figure 1.18. Flowchart
Figure 1.18. Flowchartto
toidentify
identify factors
factors affecting
affectingservice
servicelife.
life (Figure 4.19).
33

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Figure 1.19.  Screenshot of navigation through fault tree using an Excel worksheet.

the design for service life process. Each factor has the capability of reducing the service
life of the bridge deck. Chapter 4 identifies one or more strategies capable of mitigating
the effect of each particular factor listed in the circles.
Figure 1.19 shows branches of the fault tree that are applicable to the example
under consideration. In the first layer, based on the project requirements, a decision is
made that the service life of a bridge deck can only be reduced due to deficiencies. The
second layer states that either loads or natural or man-made hazards can cause bridge-
deck deficiencies. Both are judged to be applicable and are therefore selected. Figure
1.19 then shows the progression through the fault tree on the branch related to the fac-
tor “Due to Loads.” The fourth layer in Figure 1.19 states that either traffic-induced
34

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


loads or system-dependent loads can reduce bridge-deck service life. Again, both fac-
tors are judged to be applicable, and the boxes are checked. Further, the reduction
in service life of the bridge deck as a result of traffic-induced loads can be caused by
fatigue, overload, or wear, as depicted by the circles. All circles are applicable to the
example. Figure 1.19 also identifies each factor that is capable of reducing the service
life of the bridge deck as a result of system-dependent loads, as shown in circles and
identified as differential shrinkage, thermal, and system-framing restraint. Based on
project requirements, these factors are also judged applicable for consideration when
addressing service life design.
The circles in the fault tree signify issues capable of reducing service life and issues
for which the designer needs to develop mitigating strategies.
The strategies to address the individual items listed in each circle are provided in
Chapter 4 on bridge decks. For most factors listed in the circles, the Guide identifies
more than one possible strategy.
To complete the process of navigating through the fault tree, all branches appli-
cable to the problem need to be considered, and applicable circles checked. Figure 1.20
shows the remainder of the fault tree with the “Due to Loads” branch collapsed for
clarity. The decision for selecting the applicable circles is based on specific project con-
ditions and requirements.

Figure 1.20.  Segment of the fault tree with applicable circles checked for the bridge-deck example.
Figure 1.20. Segment of the fault tree with applicable circles checked for the bridge-deck example.
35

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


1.11.3
Developing Strategies and Alternative Solutions
Once the fault tree is completed and all applicable factors are identified, the individual
strategies capable of mitigating the factors can be collected. If software is used to work
through the fault tree, then this step can be automated based on the selections made.
Table 1.2 summarizes the list of individual strategies capable of mitigating the factors
developed for the example bridge-deck problem that are identified as checked circles
in Figures 1.19 and 1.20.
At this point, the designer has developed a complete list of strategies that can be
used for mitigating various factors capable of affecting the service life of the bridge
deck that are based on project requirements. Table 1.2 provides sections of the Guide
to which the user can refer for more information, as well as listing the advantages and
disadvantages of the defined strategies. Some of the strategies may contradict each
other, and others may result in similar results. Intentionally, the Guide does not pro-
vide a single strategy or attempt to identify the best strategy. In many cases strategies
to mitigate the individual factors capable of reducing service life are context sensitive,
meaning that the best strategy is very much dependent on such factors as local prac-
tice, environment, or owner preferences.
As mentioned, some of the strategies may contradict each other and some may
be more preferable because of local practices or owner preferences. Consequently, the
next step for the designer is to select strategies that are desirable for each factor affect-
ing the service life. Table 1.3 shows a narrower list of strategies extracted from the
complete list given in Table 1.2. The first row in Table 1.3 shows the applicable factors
that can reduce the service life of the bridge deck under consideration. These factors
were obtained by navigating through branches of the fault tree.
In developing the information shown in Table 1.3, the designer may consider many
factors and ensure that there are no contradicting strategies. For instance, appropriate
concrete mix is specified as one strategy to address wear, differential shrinkage, freeze–
thaw cycles, humidity, ASR, and alkali-carbonate reactivity (ACR). The designer must
ensure that a concrete mix capable of addressing all of these issues can be developed.
Otherwise, for a particular factor the designer may be forced to use another strategy.
For differential shrinkage, for example, a low modulus of elasticity concrete is needed,
but for wear, a high modulus of elasticity is needed. Consequently, to address wear and
differential shrinkage, the same concrete mix cannot be used to mitigate both factors,
and within a given deck alternative, one of these factors should be mitigated using a
different strategy.
The next step in the process is to develop possible deck alternatives that meet both
LRFD specifications and Guide requirements.
Using the information provided in Table 1.3 and ensuring there is no contradic-
tion among strategies to mitigate various factors, Table 1.4 shows four possible deck
alternatives capable of mitigating all factors affecting the service life of the bridge for
the example under consideration. The four alternatives shown in Table 1.4 are project-
specific solutions. It is also possible to automate this step by first identifying all possible
deck alternatives based on all possible combinations of strategies listed in Table 1.3, and
eliminating those judged not feasible because of contradiction among strategies.
36

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 1.2.  List of Individual Strategies Capable of Mitigating Factors Affecting Bridge Deck Service Life for Example Problem
Service Life Corresponding
Issue Project Requirements Section Mitigating Strategy Advantage Disadvantage
Overload HL 93 with 20-kip 4.3.1.1 Increase deck thickness. Minimizes cracking. Adds weight to bridge
wheel load, applied structure, increases cost.
once a month Minimize bar spacing for given Improves crack control. More labor to install and higher
amount of steel. cost.
Fatigue 24,000 average daily 4.3.1.1 Design per LRFD specifications. Minimizes possibility of May increase area of steel.
traffic northbound and reinforcement failure.
southbound
and 10% truck volume
Wear and Studded tires on high 4.3.1.1 Implement concrete mix design Identified in Chapter 3. Identified in Chapter 3.
abrasion level of service bridge strategies.
Implement membranes and Protects surface from direct Requires periodic rehabilitation
overlays. contact with tires. every 10 to 20 years.
Differential No requirements for 4.3.1.1 Use low-modulus concrete mix Allows additional strain to be Typically lower in strength and
shrinkage example problem design for composite decks. accommodated up to cracking may be subject to wear and
stress. abrasion.
Use high-creep concrete mix Reduces locked-in stresses. Uncommon mix design.
designed for composite decks. Difficult to assess stress relief.
Develop composite action after Allows slippage between deck Little experience with
concrete has hardened. and supporting members, experimental systems.
minimizing locked-in stresses. Friction reduction difficult to
assess. Introduces numerous
construction joints. Grout
integrity issues in closed void
systems.
Use precast deck panels. Allows slippage between deck Introduces numerous
and supporting members, construction joints.
minimizing locked-in stresses.

(continued)

37

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


38
Table 1.2. List of Individual Strategies Capable of Mitigating Factors Affecting Bridge Deck Service Life
for Example Problem (­continued)
Service Life Corresponding
Issue Project Requirements Section Mitigating Strategy Advantage Disadvantage
System- Deck shrinkage restraint 4.3.1.1 Develop accurate system model. Identifies design criteria for Restraining force may cause
framing from shear studs establishing stresses. cracking in deck. Refer to
restraint Chapter 8.

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Reactive Local aggregates are 4.3.1.2 Use materials and mix designs Refer to Chapter 3. Refer to Chapter 3.
ingredients: reactive that are not sensitive to
ASR/ACR aggregate.
Coastal Relative humidity 4.3.1.2 Use materials that are not Refer to Chapter 3. Refer to Chapter 3.
climate: average 70% sensitive to moisture content.
Humidity
Thermal Multiple freeze–thaw 4.3.1.2 Refer to Chapter 3 for strategies Refer to Chapter 3 for strategies Refer to Chapter 3 for strategies
climate: cycles expected relating to freeze–thaw. relating to freeze–thaw. relating to freeze–thaw.
Freeze–thaw
Thermal Potential for high 4.3.1.2 Use impermeable concrete. Increases passivity around High initial shrinkage, which
climate: chloride concentrations reinforcement. Refer to can result in cracking.
Deicing Chapter 5.
salts Use corrosion-resistant Eliminates deck spalls, High cost. Limited availability.
reinforcement. delaminations, and cracking Some performance issues as
from reinforcement. noted in Chapter 3.
Use waterproof membranes or Minimizes intrusion of dissolved Requires periodic rehabilitation
overlays. chlorides into deck. Easily every 10 to 20 years.
rehabilitated.
Use external protection Reduces corrosion. Refer to High cost. Requires extensive
methods, such as cathodic Chapter 5. maintenance and anode/battery.
protection.
Use effective drainage to keep Minimizes intrusion of dissolved Requires maintenance of
surface dry, minimize ponding. chlorides into deck. drainage.
Use periodic pressure washing Minimizes intrusion of dissolved Requires dedicated maintenance
to remove contaminants. chlorides into deck. Low cost. staff and appropriate budget.
Use nonchloride-based deicing Eliminates corrosion from High cost.
solution. chlorides.

(continued)
Table 1.2. List of Individual Strategies Capable of Mitigating Factors Affecting Bridge Deck Service Life
for Example Problem (­continued)
Service Life Corresponding
Issue Project Requirements Section Mitigating Strategy Advantage Disadvantage
Coastal Splash potential from 4.3.1.2 Use impermeable concrete. Increases passivity around Typically lower in strength and
climate: jet skis (roostertails) reinforcement. Refer to may be subject to wear and
Salt spray Chapter 5. abrasion.
Use corrosion-resistant Eliminates deck spalls, High cost. Limited availability.
reinforcement. delaminations, and cracking Some performance issues as
from reinforcement corrosion. noted in Chapter 3.
Refer to Chapter 3.
Use waterproof membranes/ Minimizes intrusion of dissolved Requires periodic rehabilitation
overlays on travel services of chlorides into deck. every 10 to 20 years.
bridge deck.
Use external protection Reduces corrosion. Refer to High cost. Requires extensive
methods, such as cathodic Chapter 5. maintenance and anode/
protection. battery.
Use sealers on nontravel surfaces Minimizes intrusion of dissolved Requires periodic rehabilitation
of bridge deck. chlorides into deck. every 5 to 10 years.
Use corrosion-resistant stay- Minimizes intrusion of dissolved Difficult to inspect.
in-place forms on bottom of chlorides into deck.
bridge deck.
Use effective drainage to keep Minimizes intrusion of dissolved Requires maintenance of
surface dry. chlorides into deck. drainage and periodic cleaning.
Use periodic pressure washing Minimizes intrusion of dissolved Requires dedicated
to remove contaminants. chlorides into deck. maintenance staff and
appropriate budget.

Note: ACR = alkali-carbonate reactivity; ASR = alkali-silica reactivity.

39

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Table 1.3.  List of Strategies Specific for Developing Deck Alternatives
Issue Strategy
Overload Increase deck thinness
Fatigue Design per AASHTO
Wear Concrete mix
Membrane and overlay
Increase thickness
System restraint Accurate modeling during structural analysis of bridge system
Differential shrinkage Concrete mix: Use mix with low modulus
Deicing Impermeable concrete
Stainless steel
Specify nonchloride-based deicing
Membrane and overlay
Freeze–thaw Concrete mix: Air content
Salt spray Stainless steel
Stay-in-place metal deck to protect bottom
Deck bottom sealer and top membrane
Humidity Use aggregate that is not sensitive to humidity
ASR/ACR Concrete mix nonreactive aggregate

For each alternative shown in Table 1.4, Rows 2 through 11 show the service life
design factors identified in Table 1.3 and corresponding strategies selected for each
alternative. Incorporating all of the select strategies listed in Rows 2 through 11 for
each of the four alternatives results in the modified deck configurations shown in the
bottom row of Table 1.4. Development of the four deck alternatives signifies the com-
pletion of Step 4a from Figure 1.18. Although the strategies listed in Table 1.3 could
lead to the development of additional deck alternatives, for the sake of simplicity only
four alternatives are shown in Table 1.4.
Alternative 1 in Table 1.4 represents a design that meets the strength requirements
stated in LRFD specifications. The total deck thickness is 8 in., with no consideration
for any of the factors capable of reducing service life. The main feature of Alternative
1 is having impermeable concrete, with 5% silica fume and 10% fly ash. The addition
of fly ash is assumed to affect the rate of reduction in the diffusivity of concrete, a
parameter used in estimating the time to initiate corrosion.
Alternative 2 in Table 1.4 relies mainly on the use of stainless steel reinforcement,
in this case Grade 316 stainless steel, to prevent corrosion. Alternative 3 in Table 1.4
uses regular concrete with increased cover to delay the time to initiate the corrosion,
and Alternative 4 uses a membrane and overlay to address corrosion.
All alternatives use increased thickness to address overload, increasing the deck
thickness by 0.5 in. Table 1.4 also presents additional strategies to address factors
capable of reducing service life.

40

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 1.4. Developing Deck Alternatives Meeting Both AASHTO and Guide Requirements
Configuration per
AASHTO Design
Issue Requirements Alternative 1 Alternative 2 Alternative 3 Alternative 4
Overload Base design Increase thickness by Increase thickness by Increase thickness by Increase thickness by
0.5 in. 0.5 in. 0.5 in. 0.5 in.
Fatigue Base design Design per AASHTO. Design per AASHTO. Design per AASHTO. Design per AASHTO.
Wear Base design Increase thickness by Increase thickness by Increase thickness by Membrane and overlay.
0.5 in. 0.5 in. 0.5 in.
System restraint Base design Accurate modeling Accurate modeling Accurate modeling Accurate modeling
during analysis of the during analysis of the during analysis of the during analysis of the
system. system. system. system.
Differential Base design Concrete mix: Use mix Concrete mix: Use mix Concrete mix: Use mix Concrete mix: Use mix
shrinkage with low modulus. with low modulus. with low modulus. with low modulus.
Deicing Base design Impermeable concrete: Stainless steel Increase cover by 1 in. Membrane and overlay.
5% silica fume, 10% fly reinforcement.
ash.
Freeze–thaw Base design Concrete mix: Air Concrete mix: Air Concrete mix: Air Concrete mix: Air content
content. content. content. and membrane and
overlay.
Salt spray Base design Impermeable concrete: Stainless steel Seal the bottom using Seal the bottom using
5% silica fume, 10% fly reinforcement. stay-in-place metal deck; stay-in-place metal deck;
ash. top is protected by top is protected by
increasing cover. membrane and overlay.
Coastal: Base design Use aggregates that are Use aggregate that are Use aggregate that are Use aggregate that are
Humidity not sensitive to humidity. not sensitive to humidity. not sensitive to humidity. not sensitive to humidity.
ASR and ACR Base design Concrete mix: Concrete mix: Concrete mix: Concrete mix:
Nonreactive aggregate. Nonreactive aggregate. Nonreactive aggregate. Nonreactive aggregate.
Strategy As designed As designed with As designed with As designed with As designed with black
(LRFD strength) thickened deck and thickened deck thickened deck, black steel reinforcement, deck
impermeable concrete. and stainless steel steel reinforcement, deck bottom sealed by metal
reinforcement. bottom sealed with stay- deck and top of deck
in-place form. membrane.
Configuration

41

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


1.11.4
Evaluating Alternatives
The next step is predicting the service life of each alternative (Step 5 in Figure 1.17)
and comparing it with the design service life of the bridge system as specified by the
owner and project requirements. Based on the outcome, the development of rehabili-
tation or replacement requirements (Step 5a in Figure 1.17) and a maintenance plan
(Step 6 in Figure 1.17) may be necessary. The last step for the bridge-deck alternative
under consideration (Step 7 in Figure 1.17) is to perform LCCA for comparison.
As described in Chapter 2 on system selection, the designer must also consider
the interactions that might exist among various parts of the bridge. This step is not
covered for this example.
The potential service life of each deck alternative can be calculated based on the
assumption that the main mode of deterioration is ingress of chloride into concrete,
which can result in corrosion of reinforcement. One approach is to use the solution to
Fick’s second law as shown in the following equation.
In a one-dimensional case, Fick’s law can be expressed and illustrated as shown
by Equation 1.5:
 x 
C( x,t ) = C0  1 − erf  (1.5)
 2 Dc t 

where
C(x,t) = chloride concentration at depth x and time t,
C0 = surface chloride concentration (kg/m3 or lb/yd3),
Dc = chloride diffusion constant (cm2/year or in.2/year), and
erf = error function (from standard mathematical tables).

The use of Fick’s law to determine the time of corrosion initiation is described
in Chapter 5, Section 2 of the Guide. Equation 1.5 can be used to assess ingress of
chloride through the concrete cover. As an example, Figure 1.21 indicates the type of
information that can be developed, which shows chloride concentration through the
deck thickness for three time periods after a deck is cast.
The information shown in Figure 1.21 can be used to predict the time when corro-
sion will be initiated, which in turn can be used to estimate the service life of the bridge
deck if corrosion of reinforcement is the main mode of deterioration. The Fick’s law in
this case is the deterioration model.
To complete the example, Life-365, a free program developed by the concrete
industry, is used to conduct the LCCA and assist in selecting an optimum solution.
Life-365 uses a finite-difference approach to solve Fick’s second law and to estimate
the time to initiation of corrosion. Other approaches, such as an error function solu-
tion to Fick’s second law, Equation 1.5, could also be used.
The solution to Fick’s second law estimates the time to initiation of corrosion (ti).
For the example, it is assumed (assumption within Life-365) that once the corrosion
is initiated, the time to propagate the corrosion to the point at which repair is needed
(tr) is a constant 6 years, regardless of concrete mix used. After the time period ti + tr,

42

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 1.21.  Chloride concentration within
­concrete over time.

Life-365 assumes that repair action is at set time intervals, say, every 10 or 20 years,
and set cost per unit area in square feet. Further, within each repair cycle, it is assumed
that only a portion of the deck area will be repaired. For instance, within each repair
cycle, only 10% or 20% of the deck will need repair.
The time to initiate corrosion depends on concrete mix and preventive measures,
such as use of stainless steel, concrete cover, or membranes. Life-365 follows the guid-
ance and terminology in ASTM E-917, Standard Practice for Measuring the Life-Cycle
Costs of Buildings and Building Systems. The final number that can be used to select
the optimal deck alternative can be the life-cycle cost, which is the initial cost plus the
present value of all future rehabilitation costs over the desired service life, in this case
100 years.
Table 1.5 shows the input parameters used within Life-365 to conduct an LCCA
for each of the four alternatives shown in Table 1.4.
It is assumed that the bridge is located in Boston, Massachusetts; has a required
service life of the deck of 100 years; and, for the sake of comparison, a total surface
area of the deck of 10,000 square ft. Table 1.5 also shows the yearly temperature pro-
file used. The diffusion coefficient and ingress of chloride are influenced by tempera-
ture fluctuation. The input parameters shown in Table 1.5 are applicable to all four
alternatives shown in Table 1.4. The specific input and end results for each alternative
are shown in Table 1.6.
As indicated in Table 1.6, Alternatives 1, 2, 3, and 4 use the same concrete mix,
referred to here as regular mix, which was used for the base option designed in accor-
dance with LRFD specifications (this option is designated as AASHTO Design in
Table 1.4). Alternative 1 uses 5% silica fume to make the concrete impermeable. Alter-
native 2 uses stainless steel. Alternative 3 uses increased concrete cover to delay the
adverse effects of corrosion, and Alternative 4 uses a membrane and overlay. A brief
description of the results for each alternative follows.
Deck Design Based on LRFD Specifications. As shown in Table 1.6, in the case
of the AASHTO base design, corrosion starts after 6.6 years (shown as Initiation).
Thereafter, the propagation of corrosion to the point of needed repair is 6 years. At the
end of 12.6 years, repair and maintenance actions are assumed to start and continue

43

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Table 1.5. Input Parameters in Life-365 for Four Alternatives AND YEARLY
TEMPERATURE PROFILE
Parameter Value
Base units U.S. units
Concentration units Weight concentration (%)
Type of structure Slabs and walls (1-D)
Third dimension (ft )2
10,000
Base year 2011
Study period (years) 100
Inflation rate (%) 1.6
Discount rate (%) 2
Location Massachusetts
Sublocation Boston
Exposure type Urban highway bridges
Maximum surface concentration (% weight of concrete) 0.68
Time to buildup (years) 7.1

Month Temperature Profile (°F)


January –1.9
February –0.9
March 3.7
April 8.9
May 14.6
June 19.8
July 23.1
August 22.2
September 18.2
October 12.7
November 7.4
December 0.9

every 10 years, during which 20% of the surface area is repaired. These assumptions
are used for the sake of demonstration and will vary based on various DOT prefer-
ences and practices. Figure 1.22 shows the total life-cycle cost based on present value.
The total life-cycle cost for the AASHTO base design is $774,676, which is shown
in the bottom row of Table 1.6. An inflation rate of 1.6% and a discount rate of 2%
were used in developing the total life-cycle cost. As indicated in Table 1.6, the initial
cost of the AASHTO base design is the lowest of the alternatives ($37, 215). However,
the total life-cycle cost is the highest ($774,676). It should be mentioned that these
LCCAs ignore the user costs, or any other cost to society during repair and closure to
traffic, and these costs can be significant.
44

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 1.6  Parameters Specific to Each Alternative from Table 1.4
AASHTO
Analysis Parameter Design Alternative 1 Alternative 2 Alternative 3 Alternative 4
Concrete mix type Regular Silica fume Regular Regular Regular
Water-cement ratio 0.42 0.35 0.42 0.42 0.42
Slag (%) 0 0 0 0 0
Fly ash (%) 0 10 0 0 0
Silica fume (%) 0 5 0 0 0
Steel type Black steel Black steel 316 Stainless Black steel Black steel
Steel (%) 1.2 1.2 1.2 1.2 1.2
Propagation period (years) 6 6 6 6 6
Inhibitor none none none none none
Barrier none none none none Membrane
D28 (in. × in./s) 1.38E-08 4.09E-09 1.38E-08 1.38E-08 1.38E-08
m (diffusion decay coefficient) 0.2 0.28 0.2 0.2 0.2
Initiation (years) 6.6 34.8 100 15.2 22.4
Propagation (years) 6 6 6 6 6
Service life (years) 12.6 40.8 106 21.2 28.4
Use user mix cost? (true or false) False False False False False
User mix cost ($/yd ) 3
0 0 0 0 0
Depth (in.) 8 9 9 10 8.5
Depth to reinforcement (in.) 2 2.5 2.5 3.5 2
Unit Cost
Area to repair (%) 20 10 10 20 5
Repair cost ($/ft ) 2
50 50 50 50 20
Repair interval (years) 10 10 10 10 10
Base mix cost ($/yd3) 80 90 80 80 80
Black steel cost ($/lb) 0.45 0.45 0.45 0.45 0.45
Epoxy steel cost ($/lb) 0.6 0.6 0.6 0.6 0.6
Stainless steel cost ($/lb) 2.99 2.99 2.99 2.99 2.99
Inhibitor cost ($/lb) 5.68 5.68 5.68 5.68 5.68
Membrane cost ($/ft2) 7 7 7 7 7
Sealant cost ($/ft2) 0.65 0.65 0.65 0.65 0.65
Result
Repair interval (years) 10 10 none 10 10
Base cost ($) 37,215 44,645 152,753 46,519 39,541
Barrier cost ($) 0 0 0 0 70,000
Repair cost ($) 737,461 232,905 0 644,595 62,711
Life-cycle cost ($) 774,676 277,550 152,753 691,114 172,252

45

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Cumulative Present Value
90

Constant Dollars ($ per sq. ft)


80
70
60
50
40
30
20
10
0
2005 2015 2025 2035 2045 2055 2065 2075 2085 2095 2105 2115
Year

Base Case

Figure 1.22.  Total life-cycle cost for LRFD specifications–based design.

Alternative 1. This alternative uses impermeable concrete by incorporating 5%


silica fume into the mix. As a result, the initial unit cost of concrete is assumed to
increase by $10/yd3. The initiation of corrosion starts at 34.8 years after casting, and
the propagation phase lasts 6 years. Therefore, after 40.8 years the repair procedure
is assumed to start. It is assumed that the repair procedure is to be conducted every
10 years and further assumed that during each repair cycle 10% of the surface area
is to need repair. Other repair alternatives can consist of complete replacement of the
deck every 40 years. The total life-cycle cost of Alternative 1 for the assumptions made
is $277,550 (Table 1.6). The initial cost of Alternative 1 ($44,645) is slightly higher
than the initial cost for the AASHTO base design ($37,215).
Alternative 2. This alternative uses Grade 316 stainless steel to address the issue of
reinforcement corrosion. The time to initiation of corrosion for this alternative is more
than 100 years, and consequently no repair action is needed. The initial cost of using
stainless steel is the highest among all alternatives ($152,753). However, the total life-
cycle cost associated with Alternative 2 ($152,753 as shown in Table 1.6) is the lowest
among all the alternatives.
Alternative 3. This alternative uses increased cover to delay the initiation of cor-
rosion. Using increased cover delays the corrosion initiation from 6.6 years for the
AASHTO base design to 15.2 years (Table 1.6). The total cost of Alternative 3 is rela-
tively high, $691,114. There does not seem to be much benefit in using this alternative,
especially considering that increasing the concrete cover will subject the substructure
and foundations to higher dead loads.

46

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Alternative 4. Alternative 4 uses a membrane and overlay to prevent corrosion of
reinforcement. The total deck thickness is only 8.5 in. compared with 9 in. and 10 in.
for Alternatives 1, 2, and 3. The repair cost per square foot for this alternative is
assumed to be lower at $20/ft2 as compared with $50/ft2 for others when a membrane
is used. It is assumed that a high-quality membrane at $7/ft2 is used and that it will
last 75 years. Consequently, the repair will involve replacing damaged overlay areas,
which are assumed to be 5% of the total surface area during each repair cycle—every
10 years starting 28.4 years after the initial installation (Table 1.6). At $172,252, the
total life-cycle cost of this alternative, using membrane, is very low, but the initial cost
($39,541 + $70,000) is more than twice the AASHTO base design cost of $37,215.
The use of a membrane could be much more economical than that indicated by this
example. For instance, the calculation leads to the conclusion that corrosion will start
after 28.4 years, which is not realistic. The concrete deck below the membrane could
last a long time without any need for repair, and any needed repair action would only
be for replacing the thin overlay, which could be achieved quickly with minimal inter-
ruption to traffic. These factors were not considered in conducting the LCCA for this
alternative.
1.11.5
Summary and Conclusion
Table 1.7 summarizes the results for all alternatives. Using this information, it is fea-
sible to conclude that the use of stainless steel or membrane plus overlay can provide
the best economy.
1.12
Future Development of the Guide
The Guide provides a general, comprehensive framework for designing new bridges
and rehabilitating existing bridges for service life. The approach presented by the
Guide is flexible and can be adapted as new information becomes available. The Guide
also provides a platform for developing customized manuals by state DOTs or for
developing a customized and systematic approach for the service life design of major
and complex bridges.

Table 1.7.  Alternative Summary


Alternative Main Feature to Address Corrosion Initial Cost Life-Cycle Cost
AASHTO base design Not applicable $37,215 $774,676
1 Impermeable concrete using silica fume $44,645 $277,550
2 Use of 316-stainless steel $152,753 $152,753
3 Increasing concrete cover $46,519 $691,114
4 Using membrane and overlay $109,541 $172,252

47

Chapter 1. DESIGN FOR SERVICE LIFE: GENERAL FRAMEWORK


Design for service life is a context-sensitive problem: local agency practices and
preferences are important. Customizing the Guide can be achieved by using the gen-
eral framework outlined in these pages and incorporating strategies and solutions pre-
ferred by each DOT for factors affecting the service life of its bridges.
One of the challenges in developing true service life design is the lack of reliable,
available deterioration models that are based on either field data or laws of physics
governing the deterioration. Several studies are under way to develop deterioration
models for various bridge elements, components, and subsystems. Such information
can be incorporated into the Guide as it becomes available. These models are needed
to further develop reliable LCCAs.
There is a need to develop specific LCCA tools dedicated to bridges, with the
ability to incorporate user costs when applicable. These tools must be flexible enough
to allow incorporating new information and deterioration models as they become
available.
There is a further need to develop more comprehensive examples that take into
account the interaction between solutions that may seem appropriate for an individual
bridge element, component, or subsystem when viewed in isolation, and yet are less
than optimum when considering service life solutions for the combined bridge system.
Finally, the significant amount of information provided in the Guide is time
consuming to comprehend in its entirety. There is a need to automate the use of the
Guide by developing tools that would facilitate navigating through all the included
information.

48

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


2
Bridge System Selection

2.1
Introduction
Selecting the proper bridge system and incorporating service life design principles into
the planning and design process are critical steps in achieving long-term bridge service
life. As it is more cost-effective to address service life at the design stage, the design for
service life must be approached in a systematic, coherent manner.
This chapter provides essential information, steps, and guidelines for selecting and
designing optimum bridge systems for both existing and new bridges. More specific
details for certain bridge elements, components, subsystems, and materials are pro-
vided in other chapters to which the reader is directed.
Commonly used bridge systems are examined along with their associated chal-
lenges and solutions, with a focus on durability and service life. The discussion covers
conventional bridge systems and newer, innovative systems involving accelerated and
modular construction. Steel and concrete bridge superstructure types are discussed,
but they are not directly compared. Instead, the discussion addresses various service
life issues within both steel and concrete superstructures.
Section 2.2 provides general information and the advantages and disadvantages of
various bridge elements, components, subsystems, and systems currently in use.
Section 2.3 summarizes the factors affecting the service life of bridge elements,
components, subsystems, and systems using a fault tree analysis approach. (A detailed
description of the fault tree is provided in Chapter 1.)
Section 2.4 provides strategies that can be used to avoid or mitigate most of the
factors affecting service life, along with options for enhancing the service life of those
factors.

49
Section 2.5 describes a systematic approach for selecting the most appropriate
bridge systems that will accommodate operational requirements and site conditions
while achieving the desired target design service life. In addition to primary system selec-
tion factors relating to function and initial cost, the necessity of considering service life
factors (e.g., importance, potential for obsolescence, element and material durability,
element maintenance and possible replacement, and life-cycle cost) is also stressed.
Above all, the strategies emphasize that durability and service life of all bridge
elements, components, subsystems, and systems must be addressed during the system
selection process for new bridges as part of a comprehensive approach to service life
design. A similar approach should be implemented for existing bridges as part of a
comprehensive plan for extending service life.
2.2
Bridge System Description
2.2.1
Bridge System Terminology
Guide definitions for various bridge terms—bridge element, component, subsystem,
and system—are presented in Chapter 1. The bridge system is the total of all elements,
components, and subsystems that make up an entire bridge.
As shown in Figure 2.1, a bridge system is initially subdivided into three main
components: deck, superstructure, and substructure. These are the primary categories
or groupings of subsystems and elements within a bridge that define specific purpose
and function.
The deck component supports and receives live load and must provide a safe,
smooth riding surface for traffic. It transfers live load and deck dead load to other
components, which in most cases is to the superstructure. The superstructure compo-
nent supports the deck and transmits loads across the span(s) to the bridge supports.
The substructure component includes all elements that support the superstructure.
It transfers vertical and horizontal loads from the superstructure to the foundation
material, such as soil or rock. At abutments, additional vertical and horizontal loads
applied from the roadway embankment are also resisted.
Often bridge systems are categorized or named by the superstructure type and
material. Superstructure components are discussed further in Section 2.2.3.

Bridge System

Superstructure Substructure
Deck Component
Component Component

Figure 2.1.  Bridge system composition.

50

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


2.2.2
Deck Component
2.2.2.1
Deck Elements
Figure 2.2 shows the various elements that make up the deck component, which in-
cludes the deck–slab element itself along with other related elements including over-
lays and wearing surfaces, expansion joints, drainage elements, railings, and curbs and
sidewalks. There are various types of deck–slab elements, including concrete decks
[either cast-in-place (CIP) or precast], steel orthotropic decks (including open or con-
crete-filled steel grids), and other types including timber and fiber-reinforced polymer.
A detailed discussion of bridge decks and related service life issues is included in
Chapter 4. Most decks are composite CIP concrete types, but other types composed
of precast concrete panels (both partial depth and full depth) and posttensioning have
been used, particularly with accelerated construction techniques. Steel deck types,
including steel orthotropic decks, are also discussed in Chapter 4. A thorough look at
materials used in bridge decks is provided in Chapter 3, and Chapter 9 examines deck
expansion devices and joints.
2.2.2.2
Bridge-Deck Drainage
The deck drainage subsystem includes inlets or scuppers, pipes and downspouts, and
outlets. The main requirement of this subsystem is to remove rainfall-generated runoff
from the bridge deck before it collects and spreads excessively in the gutter to encroach
on the traveled roadway. The deck drainage subsystem must be designed to deter flow
and accompanying corrosive deicing chemicals from contacting vulnerable structural

Deck Component

Deck/Slab Overlays

Concrete Deck Expansion Joints


Open/Sealed

Steel/Orthotropic
Drainage Subsystems
Deck
Deck/Open Exp Joint

Other: Timber, FRP


Railings

Curbs and Sidewalks Figure 2.2.  Deck component.

51

Chapter 2. BRIDGE SYSTEM SELECTION


members. Proper maintenance of deck drainage elements is essential to avoid clog-
ging and malfunction, and such maintenance requirements must be considered in the
design.
Open expansion joint drainage includes collection troughs, pipes, and attachments
below open expansion joints, such as tooth or sliding plate dams, which collect drain-
age, debris, and deicing chemicals that flow through the openings and protect adjacent
structural elements. Again, proper maintenance is essential and must be factored into
the design.
2.2.2.3
Bridge Railings
Materials used in bridge railing designs include various combinations of metal and
reinforced concrete. Crash testing requirements have been established by FHWA and
AASHTO to provide adequate strength depending on vehicle size and speed. Three
general categories of bridge railings are typically considered: traffic railings, pedestrian
or bicycle railings, and combination railings.

• Traffic railings are designed to contain and safely redirect vehicles.


• Pedestrian or bicycle railings are generally located on the outside edge of a bridge
sidewalk and are designed to safely contain pedestrians or bicyclists. AASHTO
specifications require certain heights and limit the opening sizes between members.
• Combination railings are dual-purpose railings designed to contain both vehicles
and pedestrians or bicyclists; these railings are generally located at the outside edge
of a bridge sidewalk. With this type of railing, there is usually no other barrier
between the roadway and sidewalk.

Bridge railings are often located in high-splash zones and often subject to harsh
environments that effect steel element corrosion, concrete deterioration, and reinforc-
ing bar corrosion. Special protection is necessary to ensure long-term service life of
these elements.
Bridge rails are usually cast, following the deck casting. In these instances, special
attention should be paid to the cold joint that will be created between the deck and
CIP rail as it provides a natural path for ingress of moisture and causes reinforcement
corrosion.
2.2.2.4
Curbs and Sidewalks
Curbs and sidewalks are affected similarly to deck slabs. Information relative to these
elements is provided in Chapter 4 on bridge decks and Chapter 3 on materials.
2.2.3
Superstructure Component
The superstructure component includes the structural subsystem and bearings. A de-
tailed discussion of bearing elements is given in Chapter 10. Figure 2.3 shows the vari-
ous subsystems and elements that make up the superstructure component.

52

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 2.3. Super­
Superstructure
Component structure component.

Structural Subsystem
Bearings
Steel/Concrete

Girder Arch

Truss Cable Supported

Superstructures are often categorized by

• Material type. Steel or concrete is most commonly used.


• Structure subsystem type. Girder subsystems are most often used for common
span lengths within the 300-ft limit. Longer spans typically use girders, trusses,
arches, or cable-supported types, depending on span length.
• Superstructure continuity. Many older bridges were simple spans, but more mod-
ern bridges are fully continuous or continuous for live load. Continuous spans
provide structural continuity that helps distribute traffic loads in case of excessive
deterioration of some of the bridge elements. Structural continuity is especially
important in instances of bridges with fracture-critical elements.
• Jointless systems. Integral abutment construction is gaining popularity in many
states. Integral pier construction is used only occasionally. Additional information
on integral construction is provided in Chapter 8 on jointless bridges.
• Modular construction. Modular systems using prefabricated superstructure ele-
ments such as “topped girders” or preconstructed spans are becoming more popu-
lar in situations requiring accelerated construction. Durability of connection de-
tails is a concern for the long-term service life of these systems. These types of
connections are addressed in Chapter 8 on jointless bridges.

These categories are often combined in an overall classification of the superstructure,


which is frequently used to define the entire bridge system.
The most common steel and concrete superstructure types are briefly discussed
below.

53

Chapter 2. BRIDGE SYSTEM SELECTION


2.2.3.1
Steel Superstructures
2.2.3.1.1
Steel Girder Superstructures
The most common steel bridge superstructures today are composite multigirder sub-
systems that use either rolled beams, plate girders, or tub girders. These systems can be
single span or multispan and can be either straight or curved. Either of these can also
be skewed. Rolled-beam superstructures using W-shapes are used in shorter spans up
to about 100 ft for simple spans and up to about 120 ft for continuous spans. Recently,
deeper rolled shapes (44 in.) for bridge applications have become available. When
combined with the simple for dead load and continuous for live load (SDCL) concept,
these W-shapes can be used for longer spans. Welded plate girders are usually used for
spans over 120 ft (NSBA 2008). Figure 2.4 shows typical steel I-girder and tub girder
systems. Recently, folded-plate beam sections have been developed for short-span ap-
plications. See Section 2.2.3.1.4 for steel modular systems.
Until the 1970s, many bridges were designed with systems using two main deck
girders combined with transverse floor beams and longitudinal stringers. The perceived
notion that two-girder systems are not redundant led to a significant decrease in their
use within the United States. However, two-girder systems are very popular in Europe.
Multigirder bridges with inherent redundancy are currently preferred by many bridge
owners (NSBA 2008). Use of high-performance steels with greater fracture toughness,
however, has led to a reevaluation of two-girder systems. Further, a FHWA memo
dated June 20, 2012, has paved the way to more use of two-girder systems (FHWA
2012).

Deck I-girder system. Tub girder system.

Sources: Courtesy (left) HDR Engineering, Inc. and (right) Palmer Engineering.
Figure 2.4.  Typical steel girder superstructures.
Figure 2.4. Typical steel girder superstructures.
Sources: Courtesy (left) HDR Engineering, Inc., and (right) Palmer Engineering.
54

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


A variation to the typical multigirder system is the girder–substringer system,
which has been used as an economical concept for longer spans beyond approximately
275 ft. This system uses several heavy girders with wide girder spacing and rolled-beam
stringers supported midway between the main girders by truss K-type cross frames.
2.2.3.1.2
Continuity in Steel Systems
For many years, bridges were designed as a series of simple spans with expansion joints
at each pier because they were easy to design and construct. Leaking joints, however, be-
came a leading cause of structural deterioration, and the desire to eliminate joints became
prevalent. Multispan steel girder systems were also shown to be much more efficient
when designed as continuous systems, making continuous design more commonplace.
Multispan systems have typically been fully continuous for both dead load and
live load, but new systems, typically with spans up to about 150 ft, have been intro-
duced with the SDCL concept. These systems combine the advantage of simple-span
construction with the efficiency of live load continuity and the durability of not having
joints that can ultimately leak.
Recently, extensive research studies have developed practical details for SDCL steel
bridge systems (Azizinamini et al. 2003, 2005a, 2005b; Azizinamini 2014; Lampe et
al. 2014; Farimani et al. 2014; Yakel and Azizinamini 2014; Javidi et al. 2014). These
studies demonstrate that for the SDCL steel bridge system, continuity for live load
can be provided by using steel reinforcement placed over the pier, before casting the
deck; however, in order to provide continuity, various girder connection details have
been used in practice. Figure 2.5 shows two details in use. Splice plates are sometimes
used for top flange connections, which are in tension. The SDCL research studies,

(a) (b)

Figure 2.5. 
Source: Steel bridge
Courtesy systems
(left) HDR using SDCL concept:
Engineering, continuity
Inc., and (a) with top flange splice and (b) without top
(right) UNL.
flange splice.
Figure
Sources:2.5. Steel(left)
Courtesy bridge
HDRsystems using
Engineering, Inc.,SDCL concept:
and (right) UNL. continuity (a) with and (b) without top flange splice.

55

Chapter 2. BRIDGE SYSTEM SELECTION


however, do not recommend using such detail. Bottom flanges in compression are typi-
cally butted with plates and wedges.
The disadvantage of the continuity detail with the top flange splice is that the bolts
for connecting the top plate have to be tightened after casting the deck. This require-
ment creates additional construction sequencing with a separate closure pour over the
pier.
2.2.3.1.3
Long-Span Superstructures
Figure 2.6 shows examples of long-span girder, truss, arch, and cable-stayed bridge
systems. Steel-plate girder systems have been used for spans up to approximately
500 ft. Spans up to 400 ft have been designed economically with parallel flanges.
Variable-depth haunched girders have been used in the 350- to 500-ft range. Use of

Long-span plate girder. Continuous truss.

Tied arch. Cable supported.

Figure
Source: 2.6.  Long-span
All photos steel HDR
courtesy bridgeEngineering,
superstructures.
Inc.; cable-supported bridge photo by Vince Streano.
Source:
Figure All
2.6.photos courtesy
Long-span HDRbridge
steel Engineering, Inc.; cable-supported bridge photo by Vince Streano.
superstructures.

56

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


high-performance steel (HPS 70W) has shown economy for plate girder and tub girder
systems in most span ranges over 150 ft, particularly in hybrid combinations. Studies
have shown that hybrid configurations using conventional-grade 50W steel in webs
and HPS 70W steel in top and bottom flanges in negative moment regions and bottom
flanges in positive moment regions are typically the most economical (Horton et al.
2002). Top flanges in positive moment regions are affected by composite action with
the deck and cannot realize enough benefit from the use of higher-strength steel to
be economical, except for longer spans. Use of HPS 70W steel in long-span negative
moment ranges can also permit economical parallel flange design without expensive
haunches. Trusses, arches, and cable-stayed and suspension systems have also been
used for longer-span applications, typically over 500 ft. For spans up to 300 ft, deck
girder systems are the most applicable.
Long-span structures can have special needs for addressing long-term service life
relating to unique details, inspection, and maintenance. Access for inspection and
maintenance can require elaborate systems of inspection walkways and access ladders,
particularly for access to fracture-critical members. Older trusses typically require
intensive maintenance because of large surface area–to–weight ratios and riveted,
built-up members with lacing bars subject to pack-out and other surface corrosion.
Truss joint details typically have moisture and debris traps that initiate corrosion, but
newer trusses have cleaner surface details that are more easily painted and maintained.
Through structures are subject to splash-zone wetting environments for all struc-
tural elements near roadway edges. This wetting needs to be considered in a corrosion-
protection and maintenance plan. Long-span bridges have large thermal movement
requirements that result in large expansion joints. This situation requires additional
attention to joint maintenance to prevent deck drainage from spilling through. Heavy
loads and large thermal movements also require special bearing designs. Navigation
channel crossings are subject to vessel collision and need to be protected.
2.2.3.1.4
Steel Modular Systems
New steel systems that provide for accelerated construction include modular construc-
tion with a pretopped deck. Modular orthotropic deck systems are also a consideration.
These modular systems require special attention to both transverse and longitudinal
connection details for achieving long-term durability.
Pretopped modular bridge systems are best suited for accelerated bridge construc-
tion applications. In these systems, several units consisting of pretopped steel or con-
crete girders are placed side by side and joined together using longitudinal closure
pours. The service life of these pretopped modular systems is significantly influenced
by the service life of the longitudinal closure pour. Pretopped steel modular systems
present two major advantages: (1) the use of steel girders significantly reduces the
creep and shrinkage deflections, and (2) pretopped steel modular units weigh less.
The folded-plate bridge system is a new modular system that offers an economical
solution for many short-span bridge applications. The system consists of a series of
standard shapes that are built by bending flat plates into inverted tub sections by using
a press break, as shown in Figure 2.7.

57

Chapter 2. BRIDGE SYSTEM SELECTION


Figure 2.7.  Making of folded-plate
girder using break press.
Source: Courtesy University of Nebraska–
Lincoln (UNL).

The maximum span length for this system is currently about 60 ft, and it is limited
by the press break length capacity available in the industry.
The process of bending the plate to form a girder can take less than an hour. Geo-
metrical variations are obtained simply by changing the bend locations. Span-length
requirements are accommodated by varying the depth of the web and the width of the
girder top and bottom flanges, thereby providing additional capacity.
The ability to provide rapid delivery is one of the major advantages of this system,
which uses only 0.375- or 0.5-in.-thick plates. Minimizing plate thicknesses allows for
stocking standard plate sizes, which means the girders can be produced and delivered
quickly (Azizinamini 2009).
Another advantage is that this system can be adapted for accelerated bridge con-
struction techniques as well as conventional construction methods. In the case of
conventional construction procedures, the deck can be easily formed and constructed
using readily available construction equipment, as shown in Figure 2.8 (Azizinamini
2009).
In the case of accelerated construction, this system easily allows for rapid con-
struction of short-span bridges using prefabricated, pretopped-girder elements. This
capability supports the recent trend within the bridge construction industry toward
minimizing the interruption of traffic by reducing the amount of construction activity
at the bridge site.
In the pretopped-girder concept, the tributary width of the concrete deck for each
folded-plate girder is cast on the girder before being shipped to the site. In this case,
each prefabricated girder unit is a folded-plate girder with a precast deck, as shown
in Figure 2.9. The steel girder can be supported at the ends or continuously supported
along the length during casting, in which case all dead loads are carried by the com-
posite section, thereby reducing deflections (Azizinamini 2009).

58

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure
Source:2.8.  Deck UNL.
Courtesy forming using conventional approach.
Source:
Figure Courtesy UNL.
2.8. Deck forming using conventional approach.

Figure
Source:2.9.  Precast
Courtesy folded-plate girder
Massachusetts unit. of Transportation (MassDOT).
Department
Figure Courtesy
Source: 2.9. Precast folded-plate
Massachusetts DOT.girder unit.

A typical two-lane rural-road bridge would require only three or four prefabri-
cated folded-plate girder units placed side by side and connected longitudinally at the
deck, as shown in Figure 2.10. The units can be connected by a variety of methods.
Further, construction can use relatively lightweight cranes as a 40-ft-long folded-plate
girder with a precast deck segment weighs only about 24,000 lb (Azizinamini 2009).

59

Chapter 2. BRIDGE SYSTEM SELECTION


Figure
Source:2.10.  Folded-plate
Courtesy MassDOT. girder system.
Figure Courtesy
Source: 2.10. Folded-plate girder
Massachusetts DOT.system.

2.2.3.2
Concrete Bridge Superstructures
Several reinforced concrete bridge systems are commonly used in the United States.
The type of system implemented at a particular site is generally dictated by economics
and the system’s ability to accommodate the required span or geometric requirements
such as curvature.
The most commonly used concrete bridge superstructures are

• CIP concrete slabs;


• Precast concrete box beams, including both spread and adjacent box beams;
• Precast concrete I-girders, including standard I-girders, bulb-tee girders, and
U-beams;
• Precast concrete spliced girders, including spliced I-girders, U-beams, and box
girders;
• CIP posttensioned box girders;
• Segmental posttensioned concrete box girders, including both precast and CIP;
• Concrete arches; and
• Modular pretopped concrete girder units, which are typically used for accelerated
bridge construction.

2.2.3.2.1
Cast-in-Place Concrete Slabs
Full-depth, CIP concrete slab superstructures consist of a concrete slab that spans
substructure units without the aid of supporting stringers, as shown in Figure 2.11.
Concrete slab bridges commonly span less than 50 ft and are typically used over minor

60

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


CIP concrete flat slab bridge. Transversely posttensioned prestressed slabs.

Source:2.11. 
Figure Courtesy Atkins North
Short-span America,
concrete Inc.
bridge applications.
Figure Courtesy
Source: 2.11. Short-span concrete
Atkins North bridge
America, Inc. applications.

water crossings. This bridge system was traditionally constructed as a series of simple
spans, but in recent years, the use of continuous spans has gained favor, eliminating
the joints over the substructure units. This system is commonly reinforced convention-
ally, but it can also be posttensioned to increase the span-length range. The haunched
posttensioned concrete slab system used can span up to about 100 ft. Many states,
especially in the Midwest, own many older concrete slab bridges, mainly constructed
in the 1930s, which have a very good performance history. When rated, these older
concrete slab bridges usually demand posting. However, research results indicate that
older concrete slab bridges possess reserve capacity significantly more than that in-
dicated by routine rating calculations (Azizinamini et al. 1995a, 1995b). The main
reason for the high capacity of older concrete slab bridges is the higher yield strength
of the reinforcement used versus the assumed value in rating calculations. This higher
capacity of existing older concrete slab bridges coupled with their good performance
record can be advantageous when developing maintenance plans.
2.2.3.2.2
Precast Concrete Box Beams
This type of superstructure consists of adjacent precast concrete box beams with a
noncomposite deck, adjacent box beams with a composite CIP concrete deck, and
spread box beams with a composite CIP concrete deck. Shallower precast solid and
voided rectangular slabs also fall into this category. Precast concrete box beams are
typically plant-manufactured standard AASHTO-PCI (Precast/Prestressed Concrete
Institute) sections that range in depth from 27 to 42 in. and are available in 36- or
48-in. widths. These precast girders are plant produced, which generally results in
high-quality products.

61

Chapter 2. BRIDGE SYSTEM SELECTION


NCHRP 2009 provides a synthesis of current practice relating to precast adjacent
box beam bridges. This superstructure type is the most prevalent box girder system for
short- and medium-span bridges, typically 20 to 127 ft, especially on secondary road-
ways. These bridges consist of multiple precast concrete box beams that are butted
against each other to form the bridge deck and superstructure. Their advantage is that
they eliminate the need for forming when using a composite CIP deck, or they can
be used directly with a bituminous overlay in the noncomposite state. Adjacent box
beams are generally connected using partial or full-depth grouted shear keys along
the sides of each box. Transverse ties are usually used in addition to the grouted shear
keys and may vary from a limited number of threaded rods to several posttensioned
tendons. In some cases, no topping is applied to the structure, but in other cases a
noncomposite topping or a composite structural slab is added. Problems have been
encountered with adjacent noncomposite box beam superstructures; these problems
are discussed in Section 2.3.3.2.2 (Hanna et al. 2009; NCHRP 2009).
2.2.3.2.3
Precast Concrete Girders
Precast concrete I-girders with a composite CIP deck are commonly used concrete
bridge superstructures in the 50- to 150-ft-plus span range. These girders are made of
high-performance, plant-produced materials and are generally very durable and result
in high-quality products. In a bridge system consisting of I-girders with composite CIP
slabs, commonly referred to as beam-slab bridges (see Figure 2.12), the longitudinal
stringers are often prestressed concrete I-girders using one of six standard AASHTO-
PCI sections, Types I through VI, which vary in depth from 28 to 54 in. In addition,

Figure 2.12. Prestressed
concrete I-girder bridge with
composite concrete deck.
Source: Courtesy Atkins North
America, Inc.

62

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


newer standard AASHTO-PCI bulb-tee shapes are used in 54-, 63-, and 72-in. depths.
These standard I- and bulb-tee shapes accommodate various span requirements up to
about 170 ft.
Bulb-tee shapes were developed to provide increased efficiency over the original
I-shapes. They have wide top flanges, similar to Type V and VI girders, which increase
stability for handling and shipping and reduce deck forming. However, bulb-tees also
have thinner top flanges, webs, and bottom flanges that reduce weight, and have other
flange geometric and proportioning modifications that optimize the sections. A num-
ber of states, including Washington, Colorado, Florida, and Nebraska, have developed
special bulb-tee shapes by modifying the standard AASHTO-PCI bulb-tee shapes to
accommodate local needs and practice.
A noteworthy I-girder advancement is the NU I-girder, which was developed by
the University of Nebraska–Lincoln (UNL) in cooperation with the Nebraska Depart-
ment of Roads and has a series of eight standard shapes with depths ranging from 29.5
to 94.5 in. (Geren and Tadros 1994; Hanna et al. 2010). The NU girders have several
section efficiency enhancements, such as wide and thick bottom flanges that enable
increased strand capacity for simple spans and provide increased negative moment
capacity for continuous spans. The wide bottom flanges also provide increased stabil-
ity in shipping and handling. Curved fillets in the top and bottom flanges reduce stress
concentration and aid the flow of concrete during fabrication. With the increased
section efficiencies, these girders have been used for spans greater than 200 ft (see
Figure 2.13).

Figure 2.13.  A 9-ft- × 3-in.-deep, 213-ft-long, 130-ton NU I-girder being shipped.


Source: Courtesy Con-Force Structures, a division of Armtec Limited Partnership, Calgary,
Alberta, Canada.

63

Chapter 2. BRIDGE SYSTEM SELECTION


In many instances, precast concrete I-girders are erected as simple spans and then
connected over the piers to form continuous for live load systems that eliminate deck
joints.
A newer alternative to concrete I-girders is the U-beam, or concrete tub girder, first
developed in Texas and now used in other states (including Florida and Washington
State), which provides economic and aesthetic spread beam systems. The Texas U54
beam is 54 in. deep, similar to an AASHTO Type IV beam, and can span up to about
140 ft (Ralls et al. 1993). The Florida U-beams have four depths ranging from 48 to
72 in. and can be used in spans ranging from about 100 to 160 ft (Florida DOT 2012).
Washington State U-beams are similar and have four depths varying from 54 in. to 72
in. and bottom flanges that are either 4 or 5 ft wide. Similar to the Florida U-beams,
these beams can accommodate span lengths up to 160 ft. With a composite concrete
deck, U-beams form a trapezoidal box shape, similar to steel tub girders. These beams
are typically designed to act as simple spans under both dead load and live load, even
when the deck is placed continuous across intermediate supports. As with I-girders,
these beams are plant produced and result in high-quality products.
2.2.3.2.4
Precast Concrete Spliced Girders
Spliced girders are precast concrete girders fabricated in several segments that are then
assembled longitudinally, typically using posttensioning, into a single simple-span or
continuous girder for the final bridge structure. They have been used to extend the
span lengths of regular short- to medium-span precast concrete girders and are de-
signed to utilize the economy and high quality of plant-produced precast girders for
longer-span applications. The length and weight of typical precast girders prevents
them from being effectively used on spans greater than about 150 ft because of the
limitations of transportation equipment and available cranes. However, with spliced
girders, precast girder segments with manageable weights and lengths are transported
to the construction site and then joined together. This procedure can either be done by
splicing girder segments on the ground and erecting them into their final position or
by placing girder segments on temporary supports and then splicing them in their final
position. Spliced girders have been used for simple spans up to about 220 ft and for
continuous spans up to about 320 ft; they have been found to provide an economical
concrete superstructure type in span ranges between that of conventional precast gird-
ers and segmental box girders. Figure 2.14 shows a typical spliced girder span under
construction using temporary supports.
Spliced girders are typically used on relatively straight alignments; however, in
recent years they have also been used for curved alignments in Nebraska and Colo-
rado. Figure 2.15 shows a spliced box girder bridge recently built in Denver, Colorado.
Precast spliced girders have some similarity with segmental box girders in that
both structure types consist of smaller girder segments that are assembled and con-
nected by posttensioning to form a final, longer girder, and both types are erected
by staged construction. However, spliced girders and segmental box girders are quite
different in the length of segments, types of splices, types of sections, tendon loca-
tions, and construction methods. A composite concrete deck is typically cast on spliced

64

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 2.14.  Spliced I-girder construction.
Source: Courtesy HDR Engineering, Inc.

Figure 2.15.  Spliced concrete curved boxes.


Source: Courtesy Summit Engineering Group.

65

Chapter 2. BRIDGE SYSTEM SELECTION


girders, but the deck slab is typically cast as an integral part of a segmental box girder.
Spliced girders use bulb-tee, I-beam, U-beam, or box shapes; segmental box girders are
typically box shapes.
2.2.3.2.5
Cast-in-Place Posttensioned Concrete Box Girders on Falsework
Posttensioned concrete box girders cast continuously on falsework have become very
popular in several states, particularly California, Arizona, and Nevada, and have been
used in spans up to about 350 ft. This type of construction lends itself to local con-
struction industry practices in which contractors can economically provide the re-
quired falsework. Similar to segmental construction, CIP on falsework offers the ad-
vantage of longer spans than conventional girders and can easily accommodate curved
alignments. CIP construction also allows clean lines and architectural finishes that
improve the aesthetics. The use of posttensioning further enhances concrete durabil-
ity by providing a superstructure that will remain essentially crack free under service
loads. Designing the structures as a frame and using monolithic connections between
the superstructure and piers also eliminates bearings, which further eliminates associ-
ated future maintenance. A potential disadvantage of this type of construction is diffi-
culty in replacing the deck or widening the bridge. Figure 2.16 shows a CIP box girder
bridge under construction.
2.2.3.2.6
Segmental Posttensioned Concrete Box Girders (CIP and Precast)
Segmental concrete box girder systems have been used when span requirements are
greater than what can be achieved with conventional stringer-type girders or spliced
girders, in instances of sharp horizontal curvature, or when special aesthetics are re-
quired. They have been economical in span ranges from about 250 to 500 ft. This
system is further divided into cantilever construction and span-by-span construction
and can be either precast or CIP. They can be cast to match the shape of any alignment,

Figure 2.16.  Cast-in-place box girder bridge on falsework.


Source: Courtesy Atkins North America, Inc.

66

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


making them particularly suited to curvature. Figure 2.17 shows a CIP segmental
bridge recently built in Florida by using balanced cantilever construction.
Segmental box girder bridges have been observed to improve deck performance as
a result of precompression of the deck. Refer to Chapter 3 on materials for additional
information on these bridge-deck systems and durability issues relating to details cur-
rently in use.
2.2.3.2.7
Concrete Arches
Concrete arches have been used for bridges with short spans of about 100 ft to long
spans of over 1,000 ft. They are typically considered today only in certain long-span
applications because of the relative economy of I-girder and segmental box girders in
shorter spans, or when special aesthetics are required. True arches are efficient struc-
tural systems because vertical dead load produces axial member compressive forces
that are resisted by a thrust at the arch abutments. Concrete has been useful for arches
because of its inherent efficiency in compression.
Concrete arches have typically been used in deck-type systems in which the arch
ribs are below the deck, but they have also been used in some through-type applica-
tions in which the arch ribs extend above the deck. Deck arch systems are either closed
spandrel types or open spandrel types. Closed spandrel types typically use ­barrel arches
with longitudinal walls along the outside edges of the arches that are either filled or
unfilled. Open spandrel types have a series of spandrel columns that transmit deck
loads to the arches.
Concrete arches in the United States have typically been constructed using either
CIP on falsework methods or cable-stayed segmental methods. Figure 2.18 shows the
cable-stayed, CIP segmental construction method used for the Hoover Dam bypass
concrete arch bridge.

Figure
Source:2.17.  CIP segmental
Courtesy concrete box
HDR Engineering, system.
Inc.; photo on right by John Rupe.
Source:
Figure Courtesy HDR
2.17. CIP Engineering,
segmental Inc.; photo
concrete box on right by John Rupe.
system.
67

Chapter 2. BRIDGE SYSTEM SELECTION


Figure 2.18.  Hoover Dam
bypass concrete arch bridge
constructed using cable-stayed
segmental methods.
Source: Courtesy HDR
Engineering, Inc.; photo by
Keith Philpott.

2.2.3.2.8
Modular Pretopped Concrete Girders
These types of systems use precast beam elements that are fabricated with a portion
of the deck in place as a unit and are erected side by side and connected with a CIP
closure joint, posttensioning, composite concrete topping, or a combination of these
methodologies. The precast elements commonly consist of conventionally reinforced or
prestressed sections that include T-beams, double Ts, and deck bulb-tees. This system
is expected to gain popularity with accelerated bridge construction as pressure mounts
to expedite construction and to minimize field forming and placing of concrete. Refer
to Chapter 4, Bridge Decks, for information concerning CIP closure connections.
A recent concept is the NEXT (Northeast EXtreme Tee) beam, which was devel-
oped by PCI Northeast along with the departments of transportation for New York,
Connecticut, Massachusetts, Vermont, Maine, New Hampshire, and Rhode Island
(Culmo and Seraderian 2010). The NEXT beam is a precast, prestressed double-tee
section with 8- or 12-ft deck widths and beam depths from 24 to 36 in. that is appli-
cable for approximately 40- to 90-ft spans. It is available with a thick top flange that
comprises the deck, or with a top flange that creates a form for a composite CIP deck.
The NEXT beam is considered as an alternative to traditional adjacent concrete box
beams, providing improved durability, lower cost, easier inspection, and accelerated
bridge construction.
2.2.4
Substructure Component
The substructure component includes all structural elements required to support
the superstructure and is typically defined from the underside of the bridge bearings
down through the foundation. The function of these elements is to transfer all verti-
cal loads from the superstructure to the foundation-supporting strata and to resist
horizontal forces acting on the bridge. The transfer of load to the supporting ground
can be through spread foundations, piles, or drilled shafts, depending on the strength

68

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


and stability of the subsurface geotechnical conditions. This component typically in-
cludes pier and abutment subsystems, each including several elements, as shown in
Figure 2.19.
2.2.4.1
Piers and Bents
Piers are intermediate supports for multispan bridges. They can have multiple con-
figurations, but typically fall into two major categories, piers and bents, as illustrated
in Figure 2.20. A pier subsystem consists of several elements, including a cap beam
supporting the main load-carrying elements of the superstructure, which in turn is

Substructure
Component

Pier Subsystem Abutment Subsystem

Cap, Column(s), Backwall, Cap Beam,


Footing(s)/Pile Cap(s), Stem/Breast Wall,
Piles/Drilled Shafts Wingwalls,
Footing/Pile Cap,
Piles/Drilled Shafts,
Reinforced Soil

Figure 2.19.  Substructure component.

(a) (b)
Figure 2.20.  Pier types: (a) multicolumn supported pier and (b) pile bent.
Source:
Source: Courtesy
Courtesy Atkins
Atkins North America,
North America, Inc.
Inc.
Figure 2.20. Pier types: (a) multicolumn supported pier and (b) pile bent.
69

Chapter 2. BRIDGE SYSTEM SELECTION


supported on one or more columns. The columns are supported by foundations that
are typically located at or below the finish grade of the adjacent ground. The founda-
tion can be a footing bearing directly on rock or soil, or a deep foundation using piles
or drilled shafts.
T-piers are examples of single-column piers with a cap element. Solid- or wall-type
piers are also single-column piers, but are wide enough to support the superstructure
without having a separate cap element.
A bent consists of a cap beam supporting the main load-carrying elements of the
superstructure, which in turn is supported directly on deep foundation elements such
as piles or drilled shafts that extend up from the finish grade. In some cases, the term
“bent” is also used to describe a multicolumn pier.
Common practice is to construct piers with reinforced concrete, although some
steel piers with pier caps have been used. When deep foundations are required to sup-
port the bent caps, they normally consist of timber; prestressed concrete square, solid
round, or hollow cylinder piles; CIP concrete drilled shafts; or high-performance steel
or steel pipe sections.
Modular, precast concrete pier elements have also been used for accelerated
construction.
Integral pier cap construction was developed as a way to avoid sharp skews or
associated longer spans in interchange ramp bridges and to lower overpass profiles.
Integral pier caps also have the advantage of eliminating bearings, which can minimize
future maintenance requirements. Figure 2.21 shows a ramp bridge with conventional
stacked T-pier construction and a similar ramp bridge with integral pier construction.
An integral pier system is advantageous in seismic areas because it integrates the super-
structure and substructure and creates frame action.

(a) (b)

Figure
Source: 2.21.  Ramp
Courtesy bridge
HDR with (a) conventional
Engineering, Inc. T-piers and (b) integral piers.
Figure Courtesy
Source: 2.21. Ramp
HDRbridge with (a)
Engineering, Inc.conventional T-piers and (b) integral piers.

70

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


2.2.4.2
Abutments
Abutments are provided in multiple configurations, but they can be defined in two
major categories, as illustrated in Figure 2.22: stub or spill-through abutments and full
abutments.
Stub abutments are characterized by sloped embankments under the end span of
the bridge. They provide support to the superstructure through a shallow bent cap
resting on a pile foundation.
Traditionally, full abutments have been characterized by a vertical wall that retains
the embankment fill and also transfers the bridge load to the supporting foundation
at the base of the wall. Full abutments can also be in the form of a mechanically
stabilized earth (MSE) system, which employs a fascia wall connected to a system of
reinforcing elements in multiple layers that work with the backfill material to create a
composite soil mass. This composite soil mass can then support vertical load and/or
act as an earth retention system. There are two types of MSE abutments: true or mixed
(Anderson and Brabant 2010). In a true MSE abutment, the bridge superstructure is
supported on spread footings bearing directly on the top of the reinforced soil mass.
In a mixed MSE abutment, a shallow bent cap with a row of piles is used to support
the superstructure behind the MSE fascia wall, and the reinforced soil mass is used to
retain the fill behind and adjacent to the end of the bridge. An MSE full abutment is
pictured in Figure 2.22.
Another recent form of abutment system is the geosynthetic-reinforced soil inte-
grated bridge system, which is described in a recent FHWA report (Adams et al. 2011).
This is a relatively new abutment system that has been used for accelerated bridge con-
struction, typically for short spans up to about 140 ft. The abutment uses alternating

(a) (b)

Figure
Source:2.22.  (a) Stub
Courtesy or spill-through
Atkins North America,abutment
Inc. and (b) MSE full abutment.
Figure Courtesy
Source: 2.22. (a)Atkins
Stub North
or spill-through abutment
America, Inc. and (b) MSE full abutment.

71

Chapter 2. BRIDGE SYSTEM SELECTION


thin layers of compacted fill and geosynthetic reinforcement sheets that combine to
form a reinforced soil mass foundation that directly supports the bridge superstructure
without the need for piles. The geosynthetic reinforcement is connected into layers of
precast facing blocks that are placed with the reinforcement and soil backfill. Once
completed, the reinforced soil mass is ready to support the bridge.
Traditional abutments are typically concrete construction. When deep foundations
are required to support the bent caps, they normally consist of timber; prestressed
concrete square, solid round, or hollow cylinder piles; CIP concrete drilled shafts; or
high-performance steel or pipe pile sections.
The types of abutments used also characterize the way the entire bridge system
responds to thermally induced longitudinal movements. There are three distinct abut-
ment types:

1. Integral abutment;
2. Semi-integral abutment; and
3. Abutment using expansion devices.

In integral abutment systems, piles are attached directly to the abutment, and ther-
mally induced longitudinal movements are accommodated by the flexibility of the
piles. The piles are subject to both axial and flexural moments. In semi-integral abut-
ment systems, girders and piles are not directly connected, and the bearings used to
support the girders and piles are mainly subject to axial loads. Integral and semi-
integral abutment systems are part of different continuous bridge systems. Chapter 8
provides a more in-depth discussion, as well as detailed design provisions, for integral
and semi-integral abutment systems.
2.3
Factors Affecting Service Life
All the elements, components, and subsystems that make up the overall bridge system
are adversely affected to various degrees by external and internal factors that contrib-
ute to reduced service life. External factors typically refer to loads or hazards, which
can be both natural and human caused. Internal factors can pertain to such items as
structure type (e.g., fracture critical), materials, and design and details.
A discussion using a fault tree analysis of critical factors that affect bridge ser-
vice life is presented in this section. Figure 2.23 shows the initial fault tree diagram,
which identifies factors affecting the service life for a complete bridge system, and
Section 2.3.1 discusses these factors. Sections 2.3.2, 2.3.3, and 2.3.4 address specific
factors affecting the deck, superstructure, and substructure components, respectively.
Section 2.4 addresses options to avoid or mitigate these various factors.
2.3.1
Bridge System Fault Tree Analysis
The fault tree diagram in Figure 2.23 identifies and categorizes causes of reduced ser-
vice life or factors affecting service life. (A detailed discussion of fault tree analysis is
included in Chapter 1.) In the following sections, these categories are successively sub-
categorized at descending levels to identify multiple contributing factors. The fault tree

72

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Reduced Service Life of
Bridge System

Caused by
Caused by Deficiency
Obsolescence

Due to Capacity/
Safety Issues Of Deck Of Superstructure Of Substructure
Component Component Component

Due to Span Layout/


Clearance Issues

Due to Increase in Due to Due to Natural or Due to Production/


Design Live Load Loads Man-made
- Hazards Operation Defects

Figure 2.23.  Factors affecting service life.


Figure 2.23. Factors affe cting service life.

analysis is continued until the basic events or lowest levels of resolution are reached
and discussed. The lowest level or basic events require strategies for mitigation.
2.3.1.1
Obsolescence or Deficiency
At the highest fault level, reduced service life of a bridge system can be attributed to
either obsolescence or deficiency.
Obsolescence refers to reduced service life of a bridge as a result of issues related
to how it functions, which can be further categorized as

• Operational issues related to reduced traffic capacity and safety;


• Physical issues related to span layout and clearances; or
• Loading issues related to increases in design live load.

Many bridges are replaced because of functional issues well before their full poten-
tial service life is achieved. Significant increases in corridor traffic demand, caused by
such factors as urban planning, land use, and development, can ultimately result in
the functional inability of a bridge to provide a required level of service, necessitat-
ing bridge widening or replacement. Vertical clearance limitations sometimes prevent

73

Chapter 2. BRIDGE SYSTEM SELECTION


existing bridges from being widened. Increased corridor traffic can also require
replacement of overpass bridges to accommodate widened roadways and increased
span requirements below. Major interchanges are sometimes reconstructed because of
the need for increased traffic capacity.
Often, safety issues relating to inadequate lane and shoulder widths, sharp curves,
and inadequate sight distances have a significant effect on service life. Changes in
design live load over the life of a bridge can affect service life as it relates to the struc-
ture’s ability to safely accommodate increased load.
Service life design considerations should evaluate the potential of future opera-
tional needs and how those needs might affect the service life of the planned facility.
Risks of future obsolescence should be considered, and appropriate choices should be
made concerning mitigation or acceptance. Those choices should be incorporated into
the design as appropriate considering life-cycle cost analysis.
Deficiency refers to reduced service life of a bridge as a result of damage or dete-
rioration that can be caused by a variety of primary factors and subfactors, each of
which can lead to reduced service life if unmitigated. Deficiency can occur in all three
bridge components: deck, superstructure, and substructure. The fault tree in Fig-
ure 2.23 continues below the superstructure component, but it applies equally to all
three components.
Within a bridge system, the interaction between components, deficiencies, or fail-
ures within a particular component can have a significant effect on other components.
A primary example is deterioration of superstructure and substructure below leaking
joints in the deck component (see Section 2.3.1.3.1). Another example is damage to
substructure and other superstructure elements caused by failed bearings in the super-
structure component (see Section 2.3.4.4).
Deficiency can be further attributed to any of three major causes:

1. Load-induced deficiencies;
2. Natural or man-made hazards; or
3. Defects in production and operation.

2.3.1.2
Reduced Service Life Caused by Loads
Load-induced deficiencies can be further categorized as those caused by traffic-­induced
loads or system-dependent loads (see Figure 2.24). The fault tree is continued for each
of these load types to identify the basic or lowest levels causing damage or deterioration.
2.3.1.2.1
Traffic-Induced Loads
Traffic-induced loads include the effects of truck and other vehicular traffic that are
applied to the bridge deck and transmitted throughout the bridge system. Traffic load
can ultimately cause damage to bridge elements through fatigue, overload, or wear.
Fatigue is structural damage to an element resulting from cyclic loading that results
in the initiation and propagation of cracks; it can occur at stress levels considerably
below the yield stress. Although fatigue can occur in reinforced concrete and structural

74

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Load-Induced

Traffic-Induced System-Dependent
Loads Loads

Time-
System-
Dependent
Fatigue Overload Wear Thermal Framing
Material
Restraint
Properties

Figure
Figure 2.24.  Load-induced deficiencies.
2.24. Load-induced deficiencies.

steel elements, it is more predominant in steel elements. Chapter 3 on materials dis-


cusses fatigue deterioration in reinforced concrete.
Early-welded steel structures have a history of cracking at certain types of weld
details as a result of load-induced and distortion-induced fatigue. Newer design provi-
sions and recommended details have been developed that provide solutions for both
load-induced and distortion-induced fatigue that will achieve desired service life. Sec-
tion 2.3.3.1.1 provides additional information of fatigue in steel structures, and Chap-
ter 7 provides a comprehensive discussion of fatigue and fracture in steel bridges.
Overload refers to element overstress or damage resulting from vehicles that
exceed maximum gross vehicle weight restrictions or individual axle or tire restric-
tions. Overload often results from illegal, nonpermitted vehicles, and it is the third
leading cause of bridge failure in the United States behind hydraulic and impact causes
(Wardhana and Hadipriono 2003). Overload produces higher stress in members than
what was considered in design; it can significantly reduce safety factors against failure,
and it can cause cracking in concrete elements. Multiple applications can also affect
fatigue behavior and also result in excessive deflection that can affect certain elements,
particularly in cases of differential deflection.

75

Chapter 2. BRIDGE SYSTEM SELECTION


Because overload occurs on many bridges, the risk of overload should be consid-
ered on certain vehicular routes when planning new bridges. It may also be necessary
to consider special owner-specified loads to avoid or mitigate this risk.
Wear refers to element damage and gradual loss of material caused by friction or
rubbing. Decks are susceptible to wear from vehicle tires, especially with the use of
studs or chains. Deck wear and abrasion is further discussed in Chapter 4. Wear has
been a factor in steel structures, particularly on pins and pin plates in connections, and
in bearings, with surface wear in sliding bearings, brass sealing ring wear in pot bear-
ings, and pin wear in steel bearings.
2.3.1.2.2
System-Induced Loads
System-induced loads include the effects of the bridge system configuration on the
behavior of the structure. These effects are accentuated by restraints provided through
bridge boundary conditions and can result in significant locked-in stresses. The sys-
tem-induced loads can be the result of movements due to time-dependent material
properties, thermal movements, or system-framing restraint.
Time-dependent material properties refer mainly to shrinkage and creep-related
deformations in restrained concrete elements that can result in concrete cracking. This
phenomenon is discussed further in Chapter 4 for bridge decks and in Chapter 3 for
concrete materials in general.
Thermal conditions refer to effects caused by temperature change, which can
result in significant stresses in restrained structural members, and in some cases can be
as damaging as live load stresses. The effects can be the result of uniform stress across
a bridge member or the result of a temperature gradient throughout the depth of a
member.
System-framing restraint refers to effects caused by boundary condition restraints
that prevent normal or intended structural behavior. Improper function or seizing of
bearings can result in unintended movement restraint, which can further cause pier
cracking and distress. Another example occurs at the ends of skewed integral abut-
ments, where lateral movement resulting from the skew can cause cracking and dis-
tress in corner details if adequate clearance is not provided to allow for the movement.
2.3.1.3
Reduced Service Life As a Result of Natural or Man-Made Hazards
Environmental hazards from both natural and man-made sources can have a signifi-
cant influence on bridge service life. These hazards also include effects from areas
with adverse thermal climate, coastal climates, and chemical climates, as well as
from chemical properties of the materials themselves. Other hazards such as hy-
draulic action, collisions, fire or blast, or seismic events can also have considerable
effect. These natural and man-made hazards are introduced in the fault tree shown
in Figure 2.25.

76

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Natural or
Man-made
Hazards

Thermal Coastal Chemical Reactive Hydraulic Extreme


Climate Climate Climate Materials Action Events

Salt Water/ Corrosion- Flood/ Vehicle/


Deicing Salts Vessel Seismic
Spray Inducing ASR Storm
Corrosion Collision
Corrosion Chemicals Surge

Freeze– Sulfate
Humidity ACR Scour Fire/Blast
Thaw Attack

Figure2.25.
Figure 2.25.  Natural
Natural oror man-madehazards.
man-made hazards.

2.3.1.3.1
Thermal Climate
Deicing salts corrosion. Bridge service life is typically severely affected in cold, wet
climates because of the heavy use of roadway deicing salt. Salt-contaminated mois-
ture penetration directly affects bridge-deck service life by initiating and propagating
corrosion in unprotected reinforcing steel and by accelerating concrete deterioration
caused by cracking and freeze–thaw damage. All unprotected bridge elements below
open or leaking deck joints are subject to salt-contaminated roadway drainage, which
causes unprotected structural steel corrosion, concrete reinforcing bar corrosion, and
associated concrete cracking and spalling.
On overpass bridges, salt spray rising from roadways below affects super­structures
and the undersides of decks, causing concrete penetration and corrosion of unpro-
tected reinforcing and corrosion of unprotected structural steel. The salt spray from
vehicles passing underneath the bridge can also affect the service life of weathering
steel by keeping the steel continuously wet.
Decks, barriers, and deck joints are also susceptible to damage from snow plows
used to clear roadways for traffic.
Freeze–thaw. Water absorbed into concrete surfaces and contained in cracks can
freeze in cold weather conditions. The frozen water tends to expand, causing stresses
within the concrete. Cyclic freezing and thawing of the water absorbed in the deck
surface fatigues the concrete and results in cracking, scaling, and spalling. Refer to
Chapter 3 on materials for additional information on freeze–thaw in concrete.
77

Chapter 2. BRIDGE SYSTEM SELECTION


2.3.1.3.2
Coastal Climate
Salt water and spray corrosion. Coastal saltwater environments also have severe ­effects
on bridge service life as a result of wetting and chloride penetration causing corrosion
of unprotected reinforcing and unprotected structural steel.
Structures in these areas are subjected to a chloride-laden saltwater environment
and a combination of wind and wave action that causes these chlorides to become
airborne as salt spray. The susceptibility of various bridge components to these envi-
ronmental influences depends on their degree of direct contact or their height above
the water. Pier columns with direct contact in areas of continual wetting are most
susceptible to damage.
Wave action hitting substructure units and seawalls or abutments under the bridge
tends to cause the salt spray to explode upward, wetting the bottoms of lower-level
superstructures and decks. On windy days the salt spray can also land directly on the
bridge-deck surface. The salt spray wets the surfaces, leaving a chloride residue that
can be absorbed into the concrete, resulting in reinforcement corrosion that in turn
can cause cracking, spalling, or delamination. The salt spray can also strike the sides of
structural steel members, affecting the service life of coatings and thus directly causing
and accelerating steel corrosion.
Humidity. High humidity in coastal regions also results in cyclic wetting and dry-
ing of bridge surfaces. Concrete materials sensitive to repeated wetting, such as those
in which reactive aggregates are used, can have an adverse effect on concrete elements.
Continuous wetting and drying also affects coatings on structural steel members and
causes steel corrosion.
2.3.1.3.3
Chemical Climate
Corrosion-inducing chemicals. Chemical climate influences on bridge service life per-
formance can be attributed primarily to airborne corrosion-inducing chemicals from
nearby industrial facilities such as chemical plants or oil- and coal-burning f­acilities.
Chapter 3 provides further discussion on chemical influences on concrete, and
Chapter 6 discusses the influence of corrosion-inducing chemicals with respect to steel
coatings and steel corrosion.
Sulfate attack. Exposure to sulfates can cause expansion of concrete material that can
cause spalling and cracking and the loss of bond strength between the cement paste and
aggregate. Refer to Chapter 3 for additional information on sulfate attack in concrete.
2.3.1.3.4
Reactive Materials
Reactive ingredients with the concrete mix can affect concrete service life performance
by altering the volumetric stability of the concrete mix design. These influences pri-
marily occur naturally.
Alkali-silica reactivity (ASR) results in swelling of aggregate particles within con-
crete that can lead to spalling, cracking, and general concrete deterioration. Chapter 3
provides additional information on ASR in concrete.

78

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Alkali-carbonate reactivity (ACR) results in aggregate expansion within concrete
that can lead to spalling, cracking, and general concrete deterioration. Refer to Chap-
ter 3 for additional information on ACR in concrete.
2.3.1.3.5
Hydraulic Action
Hydraulic action is the leading cause of bridge failure in the United States (Wardhana
and Hadipriono 2003). The principal elements of hydraulic action are flood and storm
surge and scour.
Flood and storm surge. Floods and storm surges can significantly affect bridge
service life, including dislodging spans from their bearings and washing them away.
Storm surges during major hurricanes are most often the cause of bridge damage. In
2005, for example, Hurricane Katrina devastated the Gulf coastline from ­Louisiana to
the Florida Panhandle and damaged nearly 45 bridges (Padgett et al. 2008). Most of
the damaged bridges were adjacent to water, and damage resulted from storm surge–
induced loading. Much of the damage was to superstructures, on which typical dam-
age included unseating or shifting of decks and failure of bridge parapets. Several
bridges suffered damage caused by impacts from loose barges and debris. The most
common severe failure was unseating, which often occurred in low-elevation spans.
Deck displacements were attributed primarily to a combination of buoyant forces
and pounding waves. Superstructure damage largely depended on the connection type
between the decks and bents, and the bearings often provided no apparent positive
connection between the superstructure and substructure.
Figure 2.26 shows the I-10 bridges across Escambia Bay in Florida, which were
dislodged during a storm in 2006. Damage caused by superstructure unseating was
similar to that experienced by bridges during Katrina.
Bridges with low vertical clearance over a waterway can also be vulnerable to
damage resulting from debris flow in a flood.
Scour. Scour is defined as the erosion or removal of streambed or bank material
from bridge foundations as a result of flowing water. Although scour can occur at any
time, bridge scour most often occurs during floods when swiftly flowing water has

Figure 2.26.  Florida I-10 Escambia Bay bridges


washed out during a storm in 2006.
Source: Courtesy Florida DOT.

79

Chapter 2. BRIDGE SYSTEM SELECTION


more energy than calm water to lift and carry sediment downriver. A hole is created
adjacent to the pier or abutment when material is washed away from the river bottom,
exposing or undermining footings, a situation which can compromise the integrity of
the structure and lead to failure. Figure 2.27 shows an example of abutment scour. See
Section 2.3.4 for factors affecting service life of bridge substructures.
2.3.1.3.6
Extreme Events
The following sections describe extreme events that can seriously and abruptly reduce
the service life of the bridge system: vehicle or vessel collision, fire, blast, and seismic
activity.
2.3.1.3.6a
Vehicle or Vessel Collision
Vehicle or vessel collision is second to hydraulic effects as the leading cause of bridge
failures (Wardhana and Hadipriono 2003). Bridges crossing other roadways with min-
imum or low clearance are subject to various types of vehicle collision, particularly
involving overheight vehicles. Figure 2.28 shows the effects of a collision in which a
truck transporting a hydraulic crane with the boom inadvertently raised struck a con-
crete bridge and cut halfway through the entire width of the superstructure. Piers with
minimum offset from the edge of the roadway or shoulder are also subject to vehicle
collision if not adequately protected by barriers.
The risk of vehicle impact should be considered in the design of new bridges,
particularly bridges crossing heavy truck routes where a greater probability exists for
overheight vehicles, or where there has been a history of impacts from overheight
vehicles. Possible mitigation strategies include

• Using higher clearances;


• Using sacrificial beams to protect load-carrying members; and
• Using laser detection systems that set off warning signals if an overheight vehicle
is detected.

Figure 2.27.  Abutment scour.


Source: Courtesy U.S. Geological Survey; photo by
Bill Colson.

80

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 2.28. 
Source: Bridge
Courtesy damaged
Kansas by truck transporting
Department hydraulic crane.
of Transportation.
Source: Courtesy
Figure Kansasdamaged
2.28. Bridge DOT. by truck transporting hydraulic crane.

Bridges crossing water bodies or waterways are subject to ships colliding with
either piers or superstructure. These are rarely occurring extreme events, but they have
potentially high consequences. Figure 2.29 shows the aftermath of a ship collision with
the original Sunshine Skyway Bridge in Florida. The ship collided with one of the end
piers in the main channel three-span unit and took out the pier and subsequently the
superstructure unit.
Considerations for new bridges should evaluate the span openings required for
safe navigation, including horizontal and vertical clearances, and consider appropriate
mitigation measures to reduce the risk of collision. Adequate fender systems or other
pier protection devices also need to be considered when there is a risk of ship collision.
The current LRFD Bridge Design Specifications (LRFD specifications) (AASHTO
2012) provides requirements for new bridge design for both vehicle and ship impacts.
2.3.1.3.6b
Fire
Fires, as extreme-event hazards for bridges, have a low probability of occurrence, but
they can cause significant damage to affected bridge components, including the deck,
superstructure, and substructure, and can cause collapse of entire spans. Although
fires are considered a low-risk hazard, a recent study by the New York Department of
Transportation (DOT) in 2008 showed that nearly three times more bridges have col-
lapsed because of fire than earthquakes (Kodur et al. 2010).
Fires affecting bridges most typically occur as a result of vehicle accidents either on
a bridge or on a roadway or railway crossing below a bridge, but they can also result
from fires in adjacent buildings or facilities. Fires can vary in intensity; the most intense

81

Chapter 2. BRIDGE SYSTEM SELECTION


Figure 2.29.  Sunshine Skyway Bridge collapse from ship collision.
Source: Courtesy Tampa Bay Times.

are caused by accidents with tanker trucks or railroad tanker cars carrying large quan-
tities of highly flammable fuels or chemicals. The temperature of a recent fire below a
bridge that was caused by a railroad tanker car collision loaded with 30,000 gallons of
methyl alcohol was estimated to be approximately 3,000°F (Stoddard 2002). Recent
bridge fires involving tanker trucks carrying diesel fuel and gasoline were reported to
have reached temperatures over 2,000°F (Kodur et al. 2010).
The extreme high temperatures generated in these types of bridge fires for prolonged
periods of time can significantly affect both steel and concrete structures. Figure 2.30
shows an example of a dramatic bridge fire caused by a gasoline tanker truck accident.
Steel bridge elements are especially vulnerable to high temperatures because of
steel’s high thermal conductivity: the temperature of unprotected steelwork will vary
little from that of the fire. These cases can result in loss of strength, significant sagging,
and possible collapse. Steel starts to lose strength at about 600°F, and its strength is
reduced to about half its yield strength at about 1,100°F (Brandt et al. 2011). At about
1,700°F, the yield strength is only about 10% or less. When fires at steel bridge ele-
ments reach these extreme temperatures, significant deformation and sagging usually
occurs (if not total collapse), and the affected bridge elements will typically have to be
replaced. Figure 2.31 shows extreme sagging in a steel bridge span and heavy concrete
pier deterioration after a gasoline tanker truck fire.
82

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 2.30.  Intense bridge fire resulting from a
tanker truck accident.
Source: Courtesy U.S. Fish and Wildlife Service.

Figure 2.31.  Steel bridge heavily damaged by fire after gasoline tanker truck collision.
Source: Courtesy Alabama DOT.

In cases in which damage to steel bridges sustained during a fire is not obvi-
ous (i.e., no clear signs of distress, such as sagging or buckling) the question is often
raised as to whether permanent material property effects in heat-affected areas have
occurred. It has been reported that steel will begin to encounter phase changes at
temperatures greater than 1,300°F, whereby undesirable material properties such as
reduced ductility and toughness can result during uncontrolled cooling. The Penn­
sylvania Department of Transportation (PennDOT) sponsored a study to examine the
effects of fire damage on the structural properties of steel bridge elements (Brandt et al.
2011). The study performed fire tests on painted steel-plate specimens at various tem-
peratures up to 1,200°F and exposure times to evaluate changes in surface conditions

83

Chapter 2. BRIDGE SYSTEM SELECTION


and discoloration and then tested for changes in material properties after cooling.
The results showed that up to steel surface temperatures of 1,200°F, the fire-exposed
material after cooling still satisfied AASHTO material specifications. The researchers
concluded that if excessive distortions or deformations occurred, the steel would likely
have been subjected to steel temperatures well in excess of 1,200°F, and the corre-
sponding sections would require replacement.
Concrete bridge elements are typically able to withstand high temperatures with
less distress than unprotected steel elements. Concrete has inherent fire-resistant prop-
erties because of its relatively low thermal conductivity, which insulates interior por-
tions of the member, including reinforcement and prestressed steel, from high surface
temperatures. However, concrete does experience a reduction in strength and modulus
of elasticity with high temperature. Strength reduction is largely a function of type
and size of aggregate. Concrete with siliceous aggregate (materials consisting of silica
and including granite and sandstone) begins to lose strength at about 800°F, and it is
reduced to about 55% at 1,200°F. Concrete containing lightweight aggregates (manu-
factured by heating shale, slate, or clay) and carbonate aggregates (including limestone
and dolomite) retains most of its compressive strength up to about 1,200°F (Bilow
and Kamara 2008). The following list compares concrete temperature with typically
encountered signs of distress and concrete color (Shutt 2006):

• Up to 200°F—Little or no concrete damage;


• 500°F—Surface crazing, localized cracks, iron-bearing aggregates begin to acquire
pink or red color;
• 700°F—Cracks appear around aggregate, numerous microcracks present in
­cement paste;
• 900°F—Purple or gray color appears if enough iron and lime are present;
• 1,000°F—Serious cracking of paste and aggregates occurs because of expansion.
Purple or gray color may become more pronounced;
• 1,500°F—Cement paste is completely dehydrated with severe shrinkage cracking
and honeycombing. Concrete may begin to become friable and porous;
• 2,200°F—Some components of concrete begin to fail; and
• 2,500°F—Concrete fails completely.

In all cases, fire-damaged structures should be evaluated as quickly as possible


once the fire is extinguished to determine the extent and severity of damage. The limits
of concrete damage can often be tested with an impact–rebound hammer. Concrete
core samples can be taken for petrographic examination, which will determine the
extent of damage within the overall concrete matrix. Steel coupons can be taken to
evaluate changes in material properties.

84

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


2.3.1.3.6c
Blast
The possibility of terrorism against our nation’s bridges is an ever-increasing threat.
The risk of blast attack is typically considered very low for most bridges, but major
bridges or bridges along major corridors that have high economic or sociopolitical
impact can have greater risk. By their nature, bridges are very accessible to vehicles
carrying explosive devices traveling either on the bridge or below on crossed road-
ways. They are also susceptible to ships or boats carrying explosive devices below.
Because bridges vary in type and size, the assessment of blast vulnerability can be very
complicated. Until recently, there has been little work done, and methods of analysis
and information available concerning the effects of blasts on highway bridges have
been scarce.
Extensive research is being undertaken by FHWA (Duwadi and Munley 2011) to
further understand the behavior and effects of blast loadings on bridge elements. Part
of the work of these studies is to develop methods for evaluating risk and risk-mitiga-
tion strategies. The most significant research in the area of blast-resistant design guide-
lines for highway bridges is being conducted under National Cooperative Highway
Research Program (NCHRP) Project 12-72, which has been recently documented in
NCHRP Report 645 (Williamson et al. 2010). The response evaluation of reinforced
concrete bridge columns was a key part of this investigation. Other recent research by
Agrawal and Yi (2008) dealing with blast-load effects on highway bridges developed
computer models and showed through simulation analysis that seismic capacities and
blast-load effects are strongly correlated. Kiger et al. (2010) focused on the response
of posttensioned box girder bridges under blast loads in their report on bridge vulner-
ability assessment and mitigation against explosions.
Blast loads are considered one of the extreme hazards affecting bridges; even a
small amount of explosive can produce severe localized damage to a bridge element.
In some cases, this localized damage can potentially progress to global collapse of the
structure (Kiger et al. 2010). Various factors affect the potential damage to a bridge
from a blast, including

• Size and type of explosive charge. Small explosive devices can have varied e­ ffects
depending on their placement and the size of bridge element, but large truck
bombs can be disastrous. The Oklahoma City Federal Building bombing in 1995
is an example of the devastating effect of a large truck bomb.
• Proximity to blast (standoff distance). The distance from the blast to a bridge ele-
ment is a critical parameter in determining the blast effect. For a given size blast,
the effect will reduce significantly with relatively small increases in distance from
the blast.
• Depending on the size and standoff distance, three blast categories exist: contact,
close in, and plane wave (Williamson et al. 2010). Contact blasts are very close
and create high-intensity, nonuniform loads in which breaching, or complete loss
of material at a section in a bridge element, can occur. In this case, there can be
extensive local destruction. A close-in blast is farther away but still results in a

85

Chapter 2. BRIDGE SYSTEM SELECTION


localized spherical shock wave striking the structure to produce a nonuniform
load and impulse-dominated response. A plane wave blast is far enough away to
produce essentially planar shock waves and a uniform load on the structure. In
this case, the structure will be loaded in a manner that leads to global deformation
and will be resisted by the entire structure or a number of combined elements.
• Location of blast. Blasts can occur above or below the deck. Above-deck blasts
can affect the deck itself and any structural elements above the deck, such as in a
through truss, arch, or cable-supported bridge. Blasts below the deck would typi-
cally have more effect on pier columns, but a sufficiently large blast can affect the
superstructure. Below-deck blasts can also have greater intensity because of the
enclosed effect created by the overhead structure; above-deck blasts have more
freedom to dissipate without shock wave reflection.
• Type and size of bridge element. Members with greater mass, hardness, and flex-
ibility have greater blast resistance.
• Structural redundancy. Having multiple load paths is a key factor in resisting over-
all structural collapse with any type of individual member failure. Multicolumn
piers or multigirder superstructures are typically able to redistribute internal forces
and provide greater resistance to overall structure collapse.

A risk-management approach can be taken for bridges with greater potential of


fire or blast damage. These bridges can be identified by reviewing major corridors that
would experience the greatest economical and sociopolitical impact if damaged by
these extreme-event hazards. Potential mitigation including local protective measures,
alternative routes, or accelerated reconstruction strategies can be evaluated for these
higher-risk bridges.
2.3.1.3.6d
Seismic Events
Earthquakes, including those of moderate intensity, are extreme-hazard events that
can cause significant damage to bridges, particularly to existing bridges that were de-
signed under older codes and have not been retrofitted.
The 1971 San Fernando earthquake in California, which resulted in numerous
bridge collapses, has often been cited as a watershed event in bridge engineering
because it demonstrated the inadequacy of the seismic bridge design practices of the
time (Buckle et al. 2006). FHWA became a major sponsor of bridge seismic research
shortly afterward, including research on retrofitting existing bridges.
Later major California seismic events such as the 1989 Loma Prieta and 1994
Northridge earthquakes, and the 1995 Kobe, Japan, earthquake caused significant
bridge damage and collapse, which also led to further research and understanding of
bridge seismic behavior (Azizinamini and Ghosh 1997).
The observed damage and knowledge gained from these previous events, along
with extensive research undertaken since 1971, have led to significantly improved seis-
mic bridge design specifications, including the Guide Specifications for LRFD Seismic
Bridge Design (AASHTO 2011) and LRFD Bridge Design Specifications (AASHTO

86

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


2012); advanced concepts for seismic retrofit, such as Seismic Retrofitting Manual for
Highway Structures: Part 1—Bridges (Buckle et al. 2006); and guidance for seismic
design of foundations, such as LRFD Seismic Analysis and Design of Transportation
Geotechnical Features and Structural Foundations (Kavazanjian et al. 2011).
The current approach adopted in the LRFD specifications (AASHTO 2012) is
to design new conventional (or ordinary) bridges for a design earthquake, or level of
ground motion, that represents the largest motion that can be reasonably expected
during the life of the bridge. It implies that ground motions larger than the design
earthquake could occur during the life of the bridge, but their likelihood of happening
is small. This likelihood is usually expressed as the probability of exceedance, and it
may also be described by a return period in years. The specifications call for a design
earthquake that has a 7% probability of exceedance in 75 years (a return period of
approximately 1,000 years). Bridges designed and detailed under these provisions may
suffer damage, but they should have a low probability of collapse. Key principles used
for the development of these specifications are that small to moderate earthquakes
should be resisted within the elastic range of the structural components without sig-
nificant damage, and large earthquakes should not cause collapse of all or part of the
bridge, although they may cause significant damage requiring replacement.
One of the key considerations in seismic design is repairability of the damage to
bridges during moderate seismic events. Oftentimes the so-called minor damage may
require complete replacement of the bridge. During the 1995 Hyogoken-Nanbu earth-
quake in Kobe, Japan (Bruneau et al. 1996; Chung 1996; Shinozuka et al. 1995;
­Azizinamini and Ghosh 1997), steel bridges suffered damage to superstructure ele-
ments, including inadequate cross-frame detailing leading to lateral bending of the
girder webs near the girder ends. The damage resulted in major retrofit activities and
the closing of major highways, such as the Hanshine Expressway, which was closed for
more than a year. The Kobe experience demonstrated that even minor damage to steel
bridges during seismic events can result in damage that could be very difficult to repair.
Among the lessons learned is that critical elements of the bridge that are difficult to
inspect and repair must be protected from any level of damage and remain elastic dur-
ing the entire seismic excitation.
Service life design philosophy needs to be considered when following seismic
design principles by examining the effects of repair on traffic interruption after small
to moderate earthquakes. In particular, the areas with potential to form plastic hinges,
as described in this section, must be detailed so that the repair can proceed with little
or no interruption to traffic. The major areas of concern are substructure elements, in
which most plastic hinges are anticipated. The superstructure elements of the bridge
are mainly kept elastic during the entire seismic event.
Seismic load behavior is largely unknown; consequently, the design philosophy for
buildings and bridges is to work on the behavior of the structure under known condi-
tions. Specifically, the design objective is to predefine the damage locations and design
them accordingly by providing adequate levels of ductility. In the case of bridges,
the preferred damage locations are at the ends of pier columns (formation of plastic
hinges). In the direction of traffic, it is preferred to put columns in double curvature,

87

Chapter 2. BRIDGE SYSTEM SELECTION


as shown in Figure 2.32. This arrangement allows larger portions of the pier column
(two plastic hinges versus one for single curvature) to participate in energy dissipation.
In the transverse direction, pier columns are usually designed to act in single cur-
vature, as shown in Figure 2.33.
Under longitudinal excitation, plastic hinges are located near the top and bot-
tom of the columns; under transverse excitation, the plastic hinge is located near the
­bottom of the pier column.
The main design feature in the seismic design of bridges is to keep the superstruc-
ture elements completely elastic during an entire seismic event. Repairing any elements
of the superstructure, even “minor” damage, could be very time consuming and result
in a major interruption to traffic. In a capacity design approach, which is routinely
used for designing bridges in seismic regions, the elements that should remain elastic
are referred to as protected elements. Inelasticity is then forced to take place at pre-
defined locations within the substructure. The predefined damage locations are the
weak links, or fuses, that control the level of forces to be transmitted to the super­
structure elements.

Figure 2.32.
Figure 2.32. Deflected
Deflected shape
shapeofofa athree-span bridge
three-span under
bridge a longitudinal
under (along (along
a longitudinal traffic) direction.
traffic) direction.

(a) (b)
Figure 2.33.  Deflected shape of pier column in (a) longitudinal and
(a) (b)
(b) transverse directions.
88

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


In the capacity design approach, protected elements are designed for the largest
possible force effects they might experience; the design considers the overstrength that
may exist because of higher actual material strength than that specified in design.

Areas with Seismic Risk


Although earthquakes are sometimes considered primarily a California or West Coast
problem in the continental United States, data produced by the U.S. Geological Survey
(USGS) National Seismic Hazard Mapping indicates that at least 40% of the United
States is subject to damaging, ground-shaking levels of seismic risk (Kavazanjian et al.
2011). Since 1996, USGS has developed and updated maps that have depicted areas
in the United States with various levels of seismic risk. These maps display earthquake
ground motions for various risk levels, including a 2%, 5%, and 10% probability of
being exceeded in 50 years. Figure 2.34 shows the USGS seismic hazard map depicting
peak ground acceleration levels with a 2% probability of being exceeded in 50 years, or
a return period of approximately 2,500 years. Along with areas along the West Coast,
these maps show areas of high seismic risk in the central and eastern United States.

Performance During Earthquakes


Moehle and Eberhard (2000) discuss various causes and types of damage that bridges
experience during earthquakes. Key factors that affect the type and severity of bridge
damage include the following.

Figure 2.34.  Seismic hazard map of peak ground acceleration levels with a 2% probability
of being exceeded in 50 years.
Source:
Source: Courtesy
Courtesy U.S. Geological
U.S. Geological Survey.
Survey.
Figure 2.34. Seismic hazard map of peak ground acceleration levels with a 2% probability of being
exceeded in 50 years. 89

Chapter 2. BRIDGE SYSTEM SELECTION


• Close proximity to fault rupture. Such proximity results in ground motions having
high horizontal and vertical ground accelerations and large-velocity pulses.
• Soil conditions. Soft soil sites can significantly amplify ground motion. Soil lique-
faction and lateral spreading results in permanent substructure deformation and
loss of superstructure support.
• Structural configuration. Bridges are most vulnerable that have excessive defor-
mation demands in rigid, nonductile elements; complex or nonuniform structural
configuration; curved or skewed configuration; or nonredundancy.

Major types of damage include

• Unseating at joints. Superstructure expansion joints introduce a structural irregu-


larity that can have catastrophic consequences. These joints can occur within a
span or at substructure supports. Irregular ground shaking can induce superstruc-
ture movements that can cause a span to unseat. Unrestrained superstructures can
be toppled from their supporting substructures as a result of shaking or differential
support movement associated with ground motion. Bridges with short seats are
especially vulnerable. Use of restrainers has been effective in minimizing this risk.
Figure 2.35a shows a span unseating failure on the Oakland Bay Bridge in San
Francisco during the Loma Prieta earthquake in 1989.
• Superstructure damage. Superstructures typically have sufficient strength to r­ emain
elastic during earthquakes and are unlikely to be the primary cause of collapse
of a span. However, certain types of superstructure damage have been observed,
including bearing damage, pounding of adjacent units at expansion joints, and
transverse bracing or diaphragm damage.
• Substructure damage. Substructures typically sustain the most damage, which can
be categorized by column failure and abutment damage:
–– Column failure. The lateral load capacity of a pier is limited by the shear or
flexural strength of its columns. For nonductile reinforced concrete columns,
shear failure is often the primary mode of failure when the column is subject
to large inelastic demands during strong earthquakes. Column failure is often
the primary cause of bridge collapse during earthquakes (Moehle and Eberhard
2000).
Most damage to columns can be attributed to inadequate detailing, which
limits the ability of the column to deform inelastically. In concrete columns, de-
tailing inadequacies can produce flexural, shear, splice, or anchorage failures. In
steel columns, local buckling has been observed to lead progressively to collapse
(Moehle and Eberhard 2000).
Figure 2.35b shows a nonductile column shear failure on the Cyprus Street
Viaduct in San Francisco that occurred during the Loma Prieta earthquake in
1989.
–– Abutment damage. Damage to shear keys and wingwalls is often prevalent.

90

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


(a) (b)
Figure 2.35.  1989 Loma Prieta earthquake damage near San Francisco, California: (a) Oakland Bay Bridge
upper roadway span unseating and collapse and (b) Cyprus Street viaduct support column collapse.
Source: Courtesy U.S. Geological Survey.

2.3.1.4
Reduced Service Life Resulting from Production or Operation Defects
Decisions made for the production of a bridge and activities during its operation can
have a significant influence on overall bridge service life. These production and opera-
tion influences (shown in the fault tree in Figure 2.36) include decisions made dur-
ing the design and detailing of the bridge, quality of fabrication and manufacturing,
quality of construction, the level of inspection performed during operations, and the
level and quality of maintenance. Each of these categories can be further developed to
identify the lowest or basic levels causing deficiency, but these lower levels can vary
significantly for each bridge system, component, or element type. The discussion in
this section addresses general issues that are common to all.

91

Chapter 2. BRIDGE SYSTEM SELECTION


Production/Operation
Defects

Design/ Fabrication/
Construction Inspection Maintenance
Detailing Manufacturing

Figure2.36.
Figure 2.36.  Production
Production or operation
or operation defects.
defects.

2.3.1.4.1
Design and Detailing
Decisions made during the system selection, design, and detailing phase of a bridge
project can significantly affect the service life of the bridge. It is incumbent on design-
ers to understand the implications of these decisions in order to help in making ratio-
nal choices that will improve service life.
Examples of design and detail issues causing reduced service life include

• Using bridge systems with deck joints that can ultimately leak and cause service
life issues below. See Chapter 8 on jointless bridges;
• Providing inadequate drainage systems that allow moisture to remain on bridge
decks, leading to deterioration; or improper layout, capacity, or slopes on drainage
elements that ultimately lead to clogging and malfunction;
• Dealing with moisture trap details on steel bridges that hold water and debris,
resulting in coating damage and steel corrosion. See Chapter 6 on corrosion pro-
tection of steel bridges;
• Using fatigue-prone details; and
• Not considering design errors. Effective quality assurance and quality control in
design process is necessary to avoid errors.

2.3.1.4.2
Fabrication and Manufacturing
Defects in fabrication or manufacturing can lead to reduced service life in steel or
concrete bridge elements. Undetected fabrication defects can lead to fatigue damage in
steel structures.

92

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


2.3.1.4.3
Construction
Defects or damage in construction can reduce service life in steel or concrete bridge
elements. Poor concrete placement and curing practices can have significant effects.
See Chapter 3 on materials and Chapter 4 on decks for further discussion of concrete
placement.
The transportation and erection of both steel and concrete girders can become an
issue if not handled properly. As high-performance materials (high-performance steel
and high-performance concrete) are increasingly used for bridge construction, girders
tend to become longer and their webs slimmer. Transportation of these g­ irders can
cause higher stresses or out-of-plane bending, which can result in cracking. Girder
stability during erection, particularly in curved steel girders, needs to be carefully
addressed.
2.3.1.4.4
Inspection
Proper inspection during bridge operation is essential to identify defects and issues
early, before more serious conditions develop.
2.3.1.4.5
Maintenance
Lack of maintenance or inadequate maintenance can allow deterioration to initiate
throughout the bridge system and develop into serious conditions for which the only
alternative is costly component or element replacement. Applying the appropriate
bridge preservation treatments and activities at the appropriate time can extend bridge
service life at a lower lifetime cost (FHWA 2011b).
The Bridge Preservation Guide (FHWA 2011b) provides general guidance on the
importance and benefits of preventive maintenance as part of an overall bridge preser-
vation program. Examples of various cost-effective preventive maintenance activities
that can be applied to decks, superstructure, and substructure components are further
discussed in Section 2.4.1.3.
2.3.2
Factors Affecting Service Life of Deck Component
The deck component includes several elements as described in Section 2.2.2.
Bridge-deck service life and the factors affecting service life are described in detail
in Chapter 4. Concrete bridge decks are particularly affected by thermal and coastal
environments in which chloride penetration can cause reinforcing steel corrosion
­leading to concrete cracking and spalling. This cracking can cause the concrete sur-
rounding the steel reinforcement to reach the corrosion threshold limit in environ-
ments in which the top of the bridge deck is exposed to chlorides, such as deicing salts.
Other concrete deck issues include wear and freeze–thaw damage.
Bridge expansion devices, commonly referred to as bridge joints, are discussed in
Chapter 9.
Drainage systems are most affected by lack of maintenance that causes clogging
and malfunction.

93

Chapter 2. BRIDGE SYSTEM SELECTION


Bridge railings are affected by wet chloride environments that cause corrosion of
reinforcing steel and concrete cracking and spalling in concrete railings. This condi-
tion is exacerbated at cold joints between the concrete barrier and the top of the slab,
where salt moisture can easily penetrate and cause reinforcing corrosion. The same
environments cause corrosion of steel railings.
2.3.3
Factors Affecting Service Life of Superstructure Component
2.3.3.1
Steel Superstructures
The principal causes of steel element deterioration in steel bridge systems are fatigue
and fracture (addressed in additional detail in Chapter 7) and corrosion (discussed in
Chapter 6).
2.3.3.1.1
Load-Induced Deficiency: Fatigue and Fracture
Early-welded steel structures have a history of cracking at certain types of weld details
as a result of load-induced and distortion-induced fatigue. Cracking at I-beam cover
plate terminations or at other longitudinal weld terminations in tension zones has been
particularly evident. Cracking in girder webs as a result of out-of-plane bending within
stiffener web gap regions next to cross-frame attachments also became a common
problem. Subsequently, extensive research and laboratory testing have provided an
understanding of fatigue behavior, and different weld detail types were found to have
varying levels of fatigue susceptibility. Newer design provisions and recommended
details were developed that provide solutions for both load-induced and distortion-
induced fatigue that will achieve desired service life.
Steel bridges can fail by fracture, which is the rapid, unstable propagation of a
larger flaw, most likely the result of fatigue. Fatigue crack initiation is independent of
steel type and strength, but possible brittle fracture is influenced by steel toughness,
among other variables. Early steels were more susceptible to brittle fracture, but in
recent years, new high-performance steels—HPS 50W, 70W, and 100W—have been
developed with very high toughness characteristics. Although somewhat more costly
than conventional-grade steels, high-performance steels are now encouraged when
applicable, particularly in nonredundant or fracture-critical applications. Use of high-
performance steel allows time for any fatigue cracks that may have developed to be
found during regular bridge safety inspections, before fracture can occur.
Fatigue should not be an issue in new steel bridges designed in accordance with
the latest LRFD specifications. Extensive research in recent years has identified causes
and solutions for fatigue- and fracture-related problems. When using proper details
and fabrication methods, both load-induced and distortion-induced fatigue problems
should not be an issue in achieving desired service life.

94

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


2.3.3.1.2
Deficiency Caused by Natural or Man-Made Hazards: Corrosion
Corrosion, the result of exposure to oxygen and moisture, is a fundamental limitation
of steel as a construction material. The process is greatly accelerated in the presence of
chloride ions from roadway deicing salt or salt spray in a marine environment. Deck
drainage with deicing salt leaking through open deck joints is a leading cause of steel
element corrosion in bridges.
Corrosion control should be designed into the overall steel bridge system. Use of
systems that eliminate or minimize deck joints will have a significant effect in reducing
corrosion. Details that serve to protect and keep the steel dry should be included in
the design. Among these are bridge system solutions that eliminate deck joints, thus
preventing salt-contaminated drainage from reaching steel elements below.
Salt marine environments or locations subject to deicing salt spray from below
also create harsh environmental conditions conducive to corrosion; thorough clean-
ing and zinc-rich primer coating systems can provide long-term protection. However,
requirements for related long-term coating maintenance cannot be overlooked and
must also be designed into an overall corrosion protection plan.
To achieve long-term bridge durability, a corrosion-resistance plan must be a
design requirement for every new or rehabilitated steel structure. This plan should
include the use of best painting practices and a maintenance plan that addresses paint-
ing priorities and timetables.
Best painting practices now include paint systems that contain metallic zinc as
the corrosion-resistant pigment. Zinc coatings provide galvanic protection to the steel
because zinc (the more noble metal) will oxidize (corrode) in preference to the steel.
To protect the zinc-coating layer from oxidation, additional coating layers are applied
over the zinc-rich primer.
Many studies have demonstrated the value of zinc coatings as a steel protection
system. In addition to zinc-rich paint, these zinc coatings also include galvanizing and
metalizing. Their use should be considered carefully as part of the best plan for achiev-
ing extended service life.
Weathering steel has also found widespread use in steel bridges in both unpainted and
painted applications. Weathering steel is corrosion resistant in some circumstances, but it
is adversely affected by continual drainage and roadway salts, particularly below joints.
Typically, special coatings are applied in these locations when using weathering steel.
In addition to weathering steel, a new structural stainless steel for bridges, ASTM
A1010, has been developed for use in severe corrosive environments.
2.3.3.2
Concrete Superstructures
2.3.3.2.1
General Deficiencies
There are numerous causes of deterioration in concrete superstructure elements. Chap-
ter 3 on materials discusses the various factors influencing concrete durability. Typi-
cally, the deficiencies are caused by three main factors that were described in the fault
tree analysis in Section 2.3.1:
95

Chapter 2. BRIDGE SYSTEM SELECTION


• Load-induced influences
–– Traffic causing vibration, impact, or wear; and
–– Restrained thermal movement causing internal stress and cracking.

• Natural or man-made hazards


–– Environmental influences, including effects from moisture and freezing and
thawing, and reinforcing corrosion caused by chloride exposure; and
–– Chemical influences, including exposure sulfates, carbon dioxide, alkalis, and
various acids.

• Defects in production or operation, primarily defective placement, curing, and


maintenance.
The degree and severity of concrete deterioration depends on the level of load and
environmental influences to which the bridge is subjected.
The durability of concrete exposed to these influences is highly dependent on
design practice, materials, and their proportioning and workmanship during construc-
tion. Although all of these influences are important, the principal deterioration of
concrete elements is the corrosion of steel reinforcement, which results in severe crack-
ing, spalling, and delamination of the surrounding concrete. Cracking of the concrete
often compromises the protection provided by the depassivated zone around the steel
reinforcement.
Concrete superstructures in thermal or marine environments are not exposed to
the same concentration of chlorides as the top of a bridge deck that may be directly in
contact with deicing salts. They are, however, susceptible to the same type of corro-
sion, only at a slower rate through the same mechanism of failure.
2.3.3.2.2
Adjacent Box Beam Deterioration
Adjacent box beam bridges have been used for many years and have generally per-
formed well. However, a commonly reported problem has been with cracking in the
longitudinal grouted joints between adjacent beams, which results in reflective cracks
in the top wearing surface (Hanna et al. 2009). This has been reported in both compos-
ite and noncomposite bridges (Hanna et al. 2009; NCHRP 2009; Russell 2011). The
longitudinal cracking can significantly affect the overall durability and structural be-
havior of these types of bridges. Open surface cracking allows penetration of roadway
drainage, often with de-icing chemicals, which can penetrate the sides and bottoms
of beams and cause corrosion of both nonprestressed and prestressed reinforcement
and concrete freeze–thaw drainage (see Figure 2.37). In addition, shear key cracking
can adversely affect the load distribution among the beams, resulting in loaded beams
carrying a greater proportion of load than what was assumed in the design (Hanna et
al. 2009).
In some cases, these adverse conditions have led to significant deterioration and
actual beam failure, requiring replacement of the entire bridge. In December of 2005,
a fascia beam on the Lakeview Drive Bridge over I-70 in Washington, Pennsylvania,

96

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


(a) (b)

Figure
Source:2.37.  Longitudinal
Courtesy cracking in Inc.
HDR Engineering, adjacent box beam bridge: (a) underside of beam cracking and (b) top of
roadway reflective cracking .
FigureCourtesy
Source: 2.37. Longitudinal cracking
HDR Engineering, Inc. in adjacent box beam bridge: (a) underside of beam cracking and (b) top

of roadway reflective cracking .


failed near midspan and fell to the highway below, as shown in Figure 2.38. Inspection
immediately afterwards revealed heavy concrete spalling and corrosion of strands on
the failed beam bottom flange. The inspection also found strand corrosion on other
box beams, and the bridge was permanently closed and ultimately replaced (Hanna et
al. 2009; Naito et al. 2006).

Figure 2.38.  Failure of


adjacent box beam bridge
in Pennsylvania.
Source: Courtesy
Pennsylvania DOT.

97

Chapter 2. BRIDGE SYSTEM SELECTION


2.3.4
Factors Affecting Service Life of Substructure Component
The numerous causes of substructure deterioration can be categorized in three areas:

• Improper detailing and improper consideration of appropriate forces resulting


from applied mean recurrence-level event forces, such as scour, vessel collision,
and earthquake;
• Deterioration caused by corrosion and section loss, primarily from chloride intru-
sion; and
• Seized bearings and unintended movement restraint.

2.3.4.1
Mean Recurrence-Level Event Forces
As bridge service life increases, bridges are subjected to environmental conditions for
a longer period of time. Many of these conditions are accounted for in design by the
use of traditional recurrence values of extreme environmental conditions such as those
for hydraulic stages, wind loads, and seismic events. Design for vessel impact is treated
similarly. The increased service life from a 50-year (pre-LRFD) to a 75-year (LRFD)
to a 100-year service life increases the statistical probability of exceeding the design
recurrence-level event.
The importance of considering these events has been exemplified by numerous
bridge incidents, including, but not limited to

• The 1987 collapse of the I-90 Bridge over Schoharie Creek in New York State,
which emphasized the importance of providing adequate hydraulic openings and
flow characteristics under bridges. Undermining of embankments, such as shown
in Figure 2.39, and undermining of pier foundations as a result of scour remain
primary causes of bridge failures around the nation.

Source:2.39. 
Figure Courtesy Atkins
Scour North America,
undermining of bridgeInc.
abutment.
Figure 2.39. Scour undermining of bridge abutment.
Source: Courtesy Atkins North America, Inc.

98

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


• Multiple catastrophic vessel and bridge accidents around the world from the 1960s
to the mid 1980s (Knott and Larsen 1990), including the 1964 and 1974 collapses
of the Pontchartrain Bridge in Louisiana and the 1980 collapse of the Sunshine
Skyway Bridge in Florida.
• Impact forces from motor vehicles, similar to vessel collision, and subsequent fires
that have resulted in structural damage to substructure units. Susceptibility to
damage from these aberrant vehicle impacts underscores the importance of pier
protection.
• The 1971 San Fernando earthquake, and subsequent major earthquakes identified
in Section 2.3.1.3.6d, which caused catastrophic damage to numerous bridges and
stimulated research relating to bridge performance during seismic events, such as the
Standard Specifications for Seismic Design of Highway Bridges (AASHTO 1983).

2.3.4.2
Substructure Deterioration Caused by Material Deterioration, Corrosion,
and Section Loss
Corrosion deterioration in substructure elements stems from numerous causes,
including

• Chloride intrusion from leakage of expansion joints and bridge drainage where
deicing salts are used to remove snow and ice from bridge decks;
• Chloride intrusion from direct salt splash from traffic traveling on roadways below
the bridge where deicing salts are used to remove snow and ice from the pavement;
• Chloride intrusion found in marine and brackish water environments affecting
exposed elements (such as those shown in Figure 2.40); and
• Corrosion from concrete cracking induced by ASR and other concrete quality
issues.

Many of the issues that affect the durability of the substructure are similar to the
issues affecting the bridge in general. Leakage of expansion joints and bridge drainage
create a major problem for the superstructure and substructure below the leak. Strate-
gies to address the causes and possible relief of these leaks are addressed in Chapter 9
on joints. Strategies to address concrete quality issues, such as ASR and others, are
addressed in Chapter 3. Issues related to the substructure in a marine environment or
in grade separations in which deicing salts can be splashed on supporting members are
addressed in this chapter.
Degradation of concrete and steel structures in aggressive corrosive environments,
such as the splash zone in a marine environment, has historically led to a reduction in
service life of numerous structures. The areas particularly susceptible to chloride intru-
sion are the splash zone, areas of poor concrete consolidation, spalls, and pile splices.
The degradation of piles and other deep foundation elements in marine environ-
ments has spawned an enormous concrete and steel protection and repair industry that
has developed numerous products dealing with the preservation of these deteriorating
structural elements (Heffron 2007). Many of these products are relatively new, and

99

Chapter 2. BRIDGE SYSTEM SELECTION


(a) (b)
Figure 2.40. 
Source: Marine
Courtesy pileNorth
Atkins degradation:
America,(a)Inc.
jacket failure on concrete pile and (b) corroded steel pile.
Source: 2.40.
Figure Courtesy Atkinspile
Marine North America, Inc.
degradation: (a) jacket failure on concrete pile and (b) corroded steel pile.

their long-term effectiveness and expected life are not verifiable with historic data.
Some of the more promising techniques include

• Cathodic protection with embedded sacrificial anodes;


• Pile jacketing, as shown in Figure 2.41;
• Metalized coatings;
• Crystalline admixtures for crack sealing;
• Repassivation through the removal of chloride ions; and
• Various combinations of these techniques.

2.3.4.3
Potential Effect of Climate Change and Service Life of Bridges in Coastal
Areas
The greatest potential impact of climate change on the service life of bridges in coastal
areas is the potential rise in sea level and the associated increased risk of flooding and
erosion. In general, limate-related impacts are already being observed in the United
100

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 2.41.  Pile restoration with pile encapsulation and epoxy grout fill.
Source: Courtesy Atkins North America, Inc.
Source: Courtesy Atkins North America, Inc.
Figure 2.41. Pile restoration with pile encapsulation and epoxy grout fill.

States and its coastal waters, and empirical projections suggest that many of these im-
pacts will grow in severity in the future (U.S. Global Change Research Program 2009).
Evidence shows that sea level has increased along most of the U.S. coast over the past
50 years. In some areas along the Atlantic and Gulf coasts, these increases have been
greater than 8 in. (America’s Climate Choices: Panel on Adapting to the Impacts of
Climate Change 2010).
Predicting the potential future sea level rise, however, has a great deal of uncer-
tainty. Vermeer and Rahmstorf (2009) developed a model to determine the projected
sea level rise from assumed levels of greenhouse gas emissions and subsequent tempera-
ture increases, and they evaluated three scenarios of assumed greenhouse gas emissions
as compiled by the Intergovernmental Panel on Climate Change (IPCC). Figure 2.42
shows the projected sea level rise through 2100 that was developed for each of these
scenarios (marked as B1, A2, and A1F1 on the right side of the graph). The shaded
areas represent the potential variation in each projection. The outer light-gray area
represents additional uncertainty in the projections resulting from uncertainty in the fit
between temperature rise and sea level rise. From this graph, the overall range in poten-
tial sea level rise (considering uncertainties) is from about 2 ft to about 6 ft by 2100. All
these projections are considerably larger than earlier sea level rise estimates for 2100
provided in a previous IPCC report (AR4), which is shown by the vertical bars on the
right side of the graph. This earlier report, based on data from or before 2005, did not
account for potential changes in ice sheet dynamics and resulted in predictions ranging
from about 0.6 ft to 2 ft by 2100. Also the first part of the graph shows the observed
annual global sea level rise over the past half-century relative to 1990 (Vermeer and
Rahmstorf 2009; America’s Climate Choices: Panel on Advancing the Science of Cli-
mate Change 2010).

101

Chapter 2. BRIDGE SYSTEM SELECTION


Figure 2.42.  Projection of sea
level rise from 1990 to 2100.
Sources: Vermeer and Rahmstorf
2009 and America’s Climate
Choices: Panel on Advancing the
Science of Climate Change 2010.

Although there is still debate on the level of severity that will be experienced over
the next decades because of climate change, it is important to consider this factor when
planning for new bridges or retrofitting existing bridges located in areas that could be
affected.
Currently there are no organized, formal plans for considering the possible effects
of climate changes on existing and new bridges located in coastal areas. However, the
major impact could be on the substructure and splash zone of the columns located in
the water.
2.3.4.4
Seized Bearings and Unintended Movement Restraint
The structural design of substructures is in part based on a distribution of longitudinal
and transverse forces associated with the allowable movement of the superstructure.
Fixed bearings provide an anchor that is intended to restrict “walking” of the super-
structure that can result from shrinkage and cycling of expansion and contraction. The
bearings are usually located longitudinally near the point of zero movement of a sup-
ported multispan superstructure unit. Care must be exercised in locating the so-called
point of zero movement, especially in the case of curved girder bridges. The existence
of such a point could be viewed as an assumption more than a reality. Chapter 8 on
jointless bridges provides a detailed discussion of the point of zero movement for
curved girder bridges. The point of zero movement is more meaningful in instances of
straight bridges with zero skew. Existence of skew or curvature complicates the deter-
mination of the point of zero movement. These fixed bearings, used in combination
with bearings designed to allow the superstructure either to move or slide over the
top of the substructure, reduce restraining forces that would otherwise be required to
resist the movement. Improper function or seizing of the bearings results in u­ nintended
movement restraint that can raise the force resisted by the substructure well above

102

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


the intended design. This unintended restraint can cause unanticipated cracking with
greater potential for corrosion. Proper bearing performance, which is essential to sub-
structure durability, is addressed in Chapter 10.
2.4
Options for Enhancing Service Life
This section describes options for enhancing service life and addresses the factors iden-
tified in Section 2.3.
2.4.1
System-Related Options
2.4.1.1
Functional Options
Sometimes bridges are replaced because of functional issues well before their full
­potential service life is achieved. The following considerations should be incorporated
into a bridge system when there is a probability that future bridge widening or crossed-
roadway widening may be necessary:

• Use bridge system types that can be widened, particularly superstructures. Multi­
girder superstructures typically lend themselves to widening, but CIP concrete
structures provide additional challenges for this modification.
• Consider longer spans when crossing roadways that have the potential for
widening.
• Consider additional vertical clearance when setting limits for bridges that have
the potential for future widening. Bridge widening along a deck cross slope can
infringe on minimum vertical clearances if additional clearance is not provided at
the beginning.

2.4.1.2
System Configuration Options
How the bridge system is configured will significantly affect the service life. Leaking
deck joints have been identified as one of the leading causes of system deterioration.
The considerations discussed in this section should be incorporated into the system
selection to avoid or mitigate this risk (see Chapter 8 on jointless bridges).
2.4.1.2.1
Integral Abutments
Consider integral abutments to eliminate joints at abutments. Fully integral abutments
eliminate deck joints and bearings, and semi-integral abutments eliminate deck joints.
In addition to their many other advantages with respect to enhanced service life, joint-
less integral abutment bridges provide lower initial cost. This type of bridge is increas-
ing in popularity, and its use is encouraged when appropriate. Chapter 8 on jointless
bridges provides step-by-step design provisions for jointless integral abutment bridges.
Unlike the current practice, there is no need to impose arbitrary limits on maximum
length of bridges.

103

Chapter 2. BRIDGE SYSTEM SELECTION


2.4.1.2.2
Maximum Length Limits for Continuity
Using the procedure specified in Chapter 8, establish the maximum lengths for con-
tinuity to minimize the number of joints in long, multispan viaducts. Length in the
following types of structures should be considered:

• Superstructure with integral abutments and no joints. Use design provisions stated
in Chapter 8 on jointless bridges to establish the maximum length for integral
abutments.
• Long continuous superstructure with joints only at abutments. Consider the maxi-
mum length for the structure type that can accommodate joints only at abutments
without any intermediate joints.
• Multiple continuous units with interior joints (viaduct construction). Consider the
maximum length for unit layout between interior joints to minimize the number
of joints.
• Structures requiring deck joints. Consider joint systems that are more leak resis-
tant. See Chapter 9 on expansion joints.

2.4.1.2.3
Continuity over Piers
Various systems for providing continuity over piers that eliminate deck joints should
be evaluated:

• Fully continuous. These systems are continuous for dead and live loads and are
suitable for all span lengths, but they are typically more economical for longer
spans, on which the benefit of dead load continuity is better realized.
• Simple for dead load, continuous for live load. These systems are becoming very
popular in the 150-ft span range because of the ease and speed of their construc-
tion, but girders must carry all dead load in positive bending.
• Link slab. Link slabs are economical and popular when the deck is made continu-
ous over intermediate supports while the beams remain simple span without any
continuity. This concept is further discussed in Chapter 8.
• Integral pier caps. The use of integral pier caps eliminates joints and bearings while
lowering the roadway profile, which can add further economic benefit. However,
it requires special details for the integral connection, and the system interaction
between superstructure and substructure must be considered.

2.4.1.2.4
Fixed-Pier and Expansion Pier Layout
Proper layout of fixed- and expansion-pier locations can help balance loads to piers
while minimizing superstructure thermal movements. Considerations include

• Traditional layout (single fixed pier, others expansion). Providing a single fixed
pier near the center of the bridge focuses longitudinal loads to one location, which

104

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


is usually acceptable for minimum height bridges and balances thermal move-
ments at adjacent piers and abutments as much as practicable.
• Multiple pier fixity. This method offers a benefit in taller pier situations, in which
longitudinal loads can be distributed to additional piers. In addition, tall pier flex-
ibility minimizes temperature loads that develop. The relative stiffness of multiple
fixed piers must be considered in distributing longitudinal loads and in determin-
ing temperature forces.
• Integral piers. The use of integral piers creates a fixed-pier condition and has the
benefit of eliminating joints and bearings and lowering the roadway profile. De-
pending on the type of detail, it can also provide longitudinal frame action in
resisting longitudinal loads.
• Orientation of expansion bearings on curved and skewed alignments.

2.4.1.3
Maintenance Considerations
The Bridge Preservation Guide (FHWA 2011b) provides general guidance on the im-
portance and benefits of preventive maintenance as part of an overall bridge preser-
vation program. Examples of various cost-effective preventive maintenance activities
that can be applied to decks, superstructure, and substructure components are pre-
sented, including

• Sealing or replacing leaking deck joints before deterioration can begin on elements
below;
• General bridge cleaning, including decks, joints, drainage systems, bearings, tops
of piers, and all elements below-deck joints;
• Placing deck overlays on aging decks;
• Installing cathodic protection or electromechanical chloride extraction;
• Applying concrete sealants or coatings;
• Spot painting or zone painting steel elements;
• Retrofitting fatigue-prone details;
• Lubricating bearings;
• Jacketing concrete piles in marine environments;
• Installing scour countermeasures; and
• Removing large debris from stream channels.

2.4.1.4
Access Considerations
Proper accessibility to all components and elements below deck for inspection and
future maintenance is essential for achieving long-term service life. Accessibility and
maintainability considerations must be included as part of the overall bridge system
configuration.

105

Chapter 2. BRIDGE SYSTEM SELECTION


2.4.2
Deck Component Options
For options related to deck components, see Chapter 4. For options related to joint
elements, see Chapter 9 on expansion joints.
2.4.3
Superstructure Component Options
2.4.3.1
Steel Superstructures
For options related to controlling fatigue, see Chapter 7 on fatigue and fracture. For
options related to corrosion resistance, including paint systems, galvanizing, metal-
izing, and use of corrosion-resistant steels, see Chapter 6 on corrosion protection of
steel bridges.
2.4.3.2
Concrete Superstructures
2.4.3.2.1
General Strategies
Several strategies have been developed to address the durability of concrete systems,
subsystems, and components. These strategies, which are fully described in Chapter 3
on materials, include the following:

• Proportioning concrete to provide low permeability and low cracking potential;


• Using noncorrosive materials, such as stainless steel, for reinforcement, and pro-
tective coatings, such as epoxy coating;
• Prestressing or posttensioning elements to eliminate cracking;
• Applying other solutions, such as cathodic protection and electrochemical chloride
extraction; and
• Using various combinations of these strategies.

The use and application of these strategies is highly dependent on the environment
to which the concrete systems will be exposed. A single strategy that fits all conditions
within the United States does not exist. These strategies must be reviewed for applica-
bility by each governing agency.
2.4.3.2.2
Solutions for Adjacent Concrete Box Beams
Adjacent box beam bridges have experienced service life issues in recent years that have
resulted in failures, as illustrated in Section 2.3.3.2.2. Hanna et al. (2009) reported on a
survey conducted by a Precast/Prestressed Concrete Institute sub­committee that inves-
tigated current practices in the design and construction of adjacent box girder bridges
in the United States and Canada (PCI 2008). NCHRP (2009) reported on a synthesis
of highway practice of using these types of bridges and presented more results of this
survey. Russell (2011) summarized the NCHRP report in a PCI State of the Prac-
tice Report. The survey found that 29 states and three provinces are currently using
adjacent box girder bridges. Most of these transportation agencies have experienced
106

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


premature reflective cracks in the wearing surface on the bridges built in the late 1980s
and early 1990s. These agencies have emphasized the importance of eliminating these
cracks because they allow the penetration of water and deicing chemicals, which leads
to the corrosion of reinforcing steel in the sides and bottoms of concrete boxes. The
following are examples of the preventive actions that the states and provinces have
recommended on the basis of lessons learned in the last two decades, as reported by
Hanna et al. (2009) and repeated by NCHRP (2009):

• Use of CIP deck on top of the adjacent boxes to prevent water leakage and to uni-
formly distribute the loads on adjacent boxes;
• Use of nonshrink grout or appropriate sealant instead of the conventional sand–
cement mortar in the shear keys, in addition to blast cleaning of key surfaces be-
fore grouting. A few states have also recommended the use of full-depth shear keys
because of their superior performance over traditional top flange keys;
• Use of transverse posttensioning to improve load distribution and minimize dif-
ferential deflections among adjacent boxes. Adequate posttensioning force should
be applied after grouting the shear keys to minimize the tensile stresses that cause
longitudinal cracking at these joints;
• Use of end diaphragms to ensure proper seating of adjacent boxes and intermedi-
ate diaphragms to provide the necessary stiffness in the transverse direction;
• Use of wide bearing pads under the middle of the box (centered) to eliminate the
rocking of the box while grouting the shear keys. NCHRP also recommended con-
sideration of a three-point bearing system (2009). Using sloped bearing seats that
match the surface cross slope is also recommended;
• Use of adequate concrete cover and corrosion inhibitor admixtures in the concrete
mix to resist the chloride-induced corrosion of reinforcing steel; and
• Eliminating the use of welded connections between adjacent boxes and avoiding
dimensional tolerances that result in inadequate sealing of the shear keys.

In addition to the major recommendations listed above from Hanna et al. (2009),
NCHRP (2009) listed a number of other design, fabrication, and construction recom-
mendations that would provide increased overall performance based on the PCI survey
and review of other literature.
Hanna et al. (2011) reported on a study that developed a new transverse connec-
tion system for adjacent box girder bridges that would perform better than methods
currently used. The study evaluated the potential of eliminating internal box girder
diaphragms combined with the use of non-posttensioned transverse connection sys-
tems that are designed to transfer shear and moment between adjacent box girders.
The study discussed that although the use of transverse posttensioned diaphragms
to connect adjacent box girders is an effective and practical solution in many cases, it
has some disadvantages. Posttensioning of skewed bridges is difficult and may have
to be staggered and done in stages. Staged construction leads to a significant increase
in construction cost and duration because of the variation in diaphragm location, the

107

Chapter 2. BRIDGE SYSTEM SELECTION


large number of posttensioning operations, and the excessive traffic control required
for replacement projects. Moreover, posttensioned diaphragms depend on the shear
keys to achieve the desired continuity. Shear keys need to be properly cleaned, sand-
blasted, sealed, and grouted, which adds complexity to the system and makes it sus-
ceptible to cracking and leakage.
Hanna et al. developed a revised approach that eliminates diaphragms and uses
top and bottom transverse ties. Two systems were developed and referred to as the
wide-joint system and the narrow joint system. The wide-joint system uses full-depth
and full-length shear keys, 5 in. wide, between boxes that are then filled with self-­
consolidated concrete. The narrow joint system also uses full-depth and full-length
shear keys, but the keys are narrower and filled with nonshrink grout. Both sys-
tems use transverse high-strength rods through the top and bottom flanges to pro-
vide transverse connection. A more detailed description of the narrow joint system is
included below.
Figures 2.43 and 2.44 illustrate the narrow joint system concept. It uses Grade 75
threaded rods with coupling nuts for connection over the joints between boxes, and the
rods are end-anchored by washers and nuts. The rods are spaced every 8 ft and pro-
vide continuous connection that transfers shear and moment between adjacent boxes
more efficiently than do middepth transverse ties at discrete diaphragm locations. A
slight modification is made to the standard box cross section by developing full-length
horizontal and full-depth vertical shear keys, as shown in Figure 2.43. The boxes are
fabricated with a plastic duct at the top and bottom flanges to create openings for the
threaded rods, as shown in Figure 2.44. The bottom duct is inserted between the two
layers of prestressing strands, and the top plastic duct is located 3 in. from the top
surface to provide adequate concrete cover. Vertical vents are provided at one side
from each box to allow the air to escape while grouting the ducts (Hanna et al. 2011).

Figure 2.43. Narrow joint system


box dimensions.
Source: Hanna et al. 2011.

108

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


h
4″ × 4″ × ½″ Washer Coupling nut 5″ long Figure 2.44.  Narrow joint
system connection details.
Source: Hanna et al. 2011.

0.5″ Vent 0.5″ Vent

Plastic du
uct

Nut for 0.75″ diameter rod 0.75″ diameter × 3′ × 10″


Threaded Rods

Hansen et al. (2012) developed a modified version of the narrow joint system
described above, which allows for posttensioning of the transverse rods while still
eliminating diaphragms. The rods are placed in ducts located inside the box girder
voids, adjacent to the inside surfaces of the top and bottom flanges, and are not
through the center of the flanges as shown for the narrow joint system above. It was
determined that posttensioning increased the capacity and efficiency of the section
because joints are placed under compression and are less likely to experience reflective
cracking and leakage.
2.4.3.3
Bearing Options
See Chapter 10 for a complete discussion of bearing options. Use of steel-reinforced
elastomeric bearings is considered the best option for achieving long-term service life.
2.4.4
Substructure Component Options
Substructure deficiencies are primarily the result of operational and natural or man-
made hazards, including the following.

109

Chapter 2. BRIDGE SYSTEM SELECTION


• Element material deterioration due to thermal, coastal, or chemical climate and
reactive materials. Strategies for mitigating these effects are discussed in
–– Chapter 3 on materials,
–– Chapter 5 on corrosion protection of reinforced concrete, and
–– Chapter 6 on corrosion protection of steel bridges.
• Hydraulic action, which includes flood and storm surge and scour.
• Vessel collision.

Operational issues include frozen or locked expansion bearings, which are discussed in
Chapter 10 on bridge bearings.
2.4.4.1
Hydraulic Action
2.4.4.1.1
Flood and Storm Surge
Following hurricanes Ivan in 2004 and Rita in 2005, which damaged numerous bridges
along the Gulf Coast, FHWA and 10 states sponsored a study that culminated in the
Guide Specifications for Bridges Vulnerable to Coastal Storms (AASHTO 2008). This
guide specification recommends that when practical, a vertical clearance of at least
1 ft above the 100-year design wave crest elevation, which includes the design storm
water elevation, should be provided. In response to the large uncertainty in the basic
wave and surge data needed to determine the wave crest elevation, the AASHTO study
further recommends additional freeboard. If this vertical clearance is not possible, the
bridge should be analyzed and designed to resist storm wave forces, and other wave
force mitigation measures should be implemented, such as venting to reduce buoyancy
forces.
The Florida DOT issued Temporary Design Bulletin C09-08 (Florida DOT 2009),
which required the implementation of the AASHTO Guide Specifications for Bridges
Vulnerable to Coastal Storms. In the Florida DOT bulletin, the importance or critical-
ity of bridges, as described in Table 2.1, is a factor in evaluating the risk of damage
and potential consequences.
The Florida DOT further recommends that for all bridges subject to coastal
storms, simple and inexpensive measures that enhance a structure’s capacity to resist
storm forces should be implemented. For example, placing vents in all diaphragms and

Table 2.1. Bridge Importance Level


Importance Level of Design
Extremely critical Strength limit state for little or no damage
Critical Extreme-event level state for repairable damage
Noncritical Evaluation of wave forces not required

110

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


venting all bays will reduce the effects of buoyancy forces on a structure. Anchoring
the superstructure to the substructure to reduce or prevent damage from storm surges
should also be considered.
2.4.4.1.2
Scour
The LRFD specifications require that scour at bridge foundations be designed for
100-year flood events or from an overtopping flood of a lesser recurrence interval. In
addition, the bridge foundations are to be checked for stability for a 500-year flood
event or from an overtopping flood of lesser recurrence level.
FHWA’s Hydraulic Engineering Circular No. 18: Evaluating Scour at Bridges
(Arneson et al. 2012) provides guidelines for designing bridges to resist scour and
improving the estimation of scour at bridges.
Riprap remains the countermeasure most commonly used to prevent scour at
bridge abutments. A number of physical additions to the abutments of bridges can
help prevent scour, such as the installation of gabions and stone pitching upstream
from the foundation. The addition of sheet piles or interlocking prefabricated concrete
blocks can also offer protection. These countermeasures do not change the scouring
flow, and they are temporary as the components are known to move or to be washed
away in certain flood events.
FHWA recommends design criteria and countermeasures in Arneson et al. (2012)
and in Bridge Scour and Stream Instability Countermeasures: Experience, Selection,
and Design Guidance (Lagasse et al. 2009), such as avoiding unfavorable flow pat-
terns; streamlining the abutments; designing pier foundations resistant to scour with-
out depending on the use of riprap; and other countermeasures that may be available.
To reduce the potential for scour, the bottom of spread footings should be placed
below the scour design depth, and piles or drilled shafts should be designed by assum-
ing all material above the maximum scour depth is unavailable for load resistance.
Floods also place extreme lateral loads on piers and bents, which should be consid-
ered in design for bridges in locations with a high risk of flooding. In these cases, the
presence of soil and any corresponding load or resistance should only be considered
below the minimum scour elevation.
2.5
Strategy Selection
This section outlines an approach for selecting the most appropriate bridge systems to
accommodate operational requirements and site conditions, while also achieving the
desired target design service life. The process combines the requirements for selecting
bridge systems on the basis of operational needs and initial construction cost with
requirements for service life and life-cycle cost. The approach presented in this section
must be developed in conjunction with strategies presented in subsequent chapters,
which address materials and specific components and elements, such as bridge deck,
joints, or bearings, in additional detail.

111

Chapter 2. BRIDGE SYSTEM SELECTION


Providing bridge systems with enhanced service life requires a complete under-
standing of the potential deterioration mechanisms, or factors affecting service life.
These mechanisms, described in Section 2.3, are associated with load-induced con-
ditions, local environmental hazards, production-created deficiencies, and lack of
effective operational procedures. The avoidance or mitigation of these deterioration
mechanisms through the appropriate selection of enhancement techniques is described
in Section 2.4. The overall system selection process involves a detailed evaluation of
these mechanisms as they would affect each major bridge component, subsystem, and
element and identification of a group of individual strategies that together define an
optimum bridge system configuration. This integrated approach of combining opera-
tional and service life requirements will result in the optimum bridge system with the
greatest potential for enhanced service life.
2.5.1
Service Life Design Methodology
Chapter 1 provides information concerning design methodologies for service life.
2.5.2
System Selection Process Outline
A process for selecting the optimum bridge system is shown in flowcharts in Figures
2.45 through 2.47. The outlined process involves four major steps, which are de-
scribed in the various numbered flowchart blocks:

1. Block 1: Identifying demand, which includes requirements that the bridge must
satisfy;
2. Block 2: Identifying feasible bridge system alternatives that satisfy requirements;
3. Blocks 3 through 12: Evaluating alternatives for service life; and
4. Block 13: Comparing and selecting the optimum alternative.

Each of these steps and related flowchart blocks are described in Section 2.5.3.
For ease of reading, the final flowchart step in Figures 2.45 and 2.46 is repeated at the
beginning of Figures 2.46 and 2.47, respectively. Further examples corresponding to
the blocks are provided in Table 2.2.
2.5.3
System Selection Process Description
The following is a discussion of the steps in the selection process relating to the flow-
chart blocks in Figures 2.45 through 2.47.

Block 1. The first step is to identify demand parameters that the bridge must sat-
isfy, including

a. Operational requirements. These relate to issues such as the type of corridor,


traffic and truck percentages, vehicle loads, and mixed-use requirements. This
information establishes requirements for capacity, number of lanes, and other
operational issues.

112

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 2.45.  Blocks 1 to
Bridge System Selection 3b of integrated system
Process selection and design process.

1. Identify local operational and site requirements


that affect bridge layout and service life
requirements.

2. Identify feasible system alternatives that satisfy


design provisions of LRFD specifications and satisfy
operationaland site requirements.

Go to the
next 3. For system alternative being considered,
alternative. evaluate all components, subsystems, and
From elements against factors affecting service life
Block 12 by conducting fault tree analysis.

3a. Follow each branch of


fault tree.

3b. Identify individual


factors affecting service
life.

b. Site requirements. These typically relate to issues that affect bridge layout, in-
cluding features crossed, geometrics involving curvature or skew, geotechnical
data, and other constraints.
c. Service life requirements. Based on the type of corridor and traffic demand,
a judgment is made, usually by the bridge owner, as to the importance of the
bridge and the target service life for which it should be designed.
d. Future considerations. Based on the type of corridor, an evaluation is made
regarding the potential for future needs. This might include the probability of
future traffic demand that would require bridge widening or the potential of
having to widen any crossed roadways that might affect span layout.

Overall, these requirements relate to how the bridge must function and how long it
should last. They may also include limitations on how a bridge might be constructed,
how much it should cost, or how it might look in a given setting. Further examples of
these items are listed in Table 2.2.

113

Chapter 2. BRIDGE SYSTEM SELECTION


3b. Identify individual
factors affecting service life.

Does Yes 4. Identify consequence and


factor determine appropriate strategy
apply? for avoidance or mitigation.

No

Go to the No Are all


next factors
factor. considered?

Yes

5. Collect all strategies and incorporate into


bridge system being considered to develop draft
configuration.

6. Determine if the draft configuration satisfies


requirements and has compatibility between all
components, subsystems, and elements.

Is No 7. Make appropriate modifications in


configuration components, subsystems, or elements for
okay? compatibility.

Yes
8. Develop final configuration
of system alternative being
considered.

Figure 2.46.  Blocks 3b to 8 of integrated system selection and design process.

114

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


8. Develop final configuration
of system alternative being
considered

Is service life
9. Predict service life of greater than or No 10a. Identify
various components, equal to the rehabilitation or
subsystems, and elements. system service replacement.
life?

Yes
11. Incorporate
requirements for 10b. Identify
rehabilitation, replacement, maintenance.
and maintenance into final
configuration of system
being considered, and
compute life-cycle cost.

Go to the No
next 12. Are all
alternative alternatives
Block 3 considered?

Yes

13. Compare all final


alternatives. Evaluate
advantages and disadvantages
and select most cost-effective
alternative.

Figure 2.47.  Blocks 8 to 13 of integrated system selection and design process.

115

Chapter 2. BRIDGE SYSTEM SELECTION


Table 2.2.  System Selection Process for Operational and Service Life Requirements
Stage and Major Steps Process
Preliminary Planning or Type, Size, and Location Stage
1. Identify demand Operational Demand (Functionality) Requirements and Corridor-Related Items
requirements Identify corridor-, function-, and traffic-related demand requirements:
• Corridor type such as interstate, urban arterial rural, or other;
• Traffic volumes and required capacity;
• Truck volumes;
• Special vehicle uses such as oversize vehicles or tanker trucks;
• Traffic maintenance requirements;
• Mixed-use requirements; and
• Vehicle loads and special vehicle load requirements.
Service Life Requirements (Durability and Long-Term Performance) and Corridor-
Related Items
Identify service life requirements:
• Bridge importance;
• Target design service life;
• Potential future needs according to corridor type, including
–– Potential for future bridge widening,
–– Potential for future widening of crossed roadways, and
–– Vertical clearance requirements related to future bridge widening or widening of
crossed roadways.
Local Site-Related Requirements
Identify local site-related issues that the bridge must accommodate:
• Geometrics, curvature, and skew;
• Features crossed;
• Horizontal and vertical clearances;
• Hydraulic or waterway requirements;
• Navigation requirements;
• Utility issues, either carried or crossed;
• Other physical boundary conditions;
• Geotechnical issues;
• Environmental issues;
• Drainage requirements and special criteria;
• Access for construction; and
• Aesthetics and sustainability.

(continued)

116

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 2.2.  System Selection Process for Operational and Service Life Requirements (continued)
Stage and Major Steps Process
1. Identify demand Local Items
requirements Identify local relevant environmental or man-made hazards that affect service life (follow
(continued) fault tree):
• Climate type
–– Thermal,
–– Coastal, and
–– Chemical;
• Potential for hydraulic action hazard
–– Flood, and
–– Scour;
• Potential for wind hazard;
• Potential for other extreme-event hazards
–– Vehicle or vessel collision,
–– Fire or blast, and
–– Seismic event potential.
Construction Constraints
Identify construction requirements that can affect service life:
• Construction phasing requirements;
• Construction schedule requirements (such as accelerated bridge construction); and
• Special local construction preferences.
2. Identify alternative Identify feasible alternative bridge systems including span layouts, structure types, and
solutions materials that accommodate operational and site requirements. Alternative solutions
should
• Accommodate span requirements and constraints;
• Accommodate curvature, profile, and skew requirements;
• Provide optimal span balance;
• Provide optimal superstructure and substructure cost balance;
• Identify superstructure options;
• Identify alternative bridge-deck options;
• Identify superstructure–substructure connection options;
• Identify expansion joint location and type options;
• Identify bearing type options;
• Identify substructure options; and
• Identify foundation options.
3. Evaluate and compare Identify system service life improvement strategies that avoid future potential
alternatives obsolescence by considering
• Bridge types that can be widened;
• Span lengths that accommodate widening of crossed roadways; and
• Additional vertical clearance for bridges that may need to be widened.
Evaluate factors affecting service life and identify strategies to avoid or mitigate all
hazards.
For each alternative bridge system, estimate potential service life of components,
subsystems, and elements and compare with target design service life of bridge system.
Determine rehabilitation, replacement, and maintenance requirements over target
design service life.
Determine estimated life-cycle costs over target design service life.
(continued)

117

Chapter 2. BRIDGE SYSTEM SELECTION


Table 2.2.  System Selection Process for Operational and Service Life Requirements (continued)
Stage and Major Steps Process
4. Select optimal alternative Compare operational advantages and disadvantages of bridge system alternatives:
• Identify local preferences for structure types and construction;
• Compare estimated life-cycle costs; and
• Select optimal cost-effective system considering operational and service life
requirements and cost–benefit analysis.
Final Design Stage
1. Design Design in accordance with strength and serviceability provisions of LRFD specifications.
Include specific design and details to address service life issues identified in preliminary
stage:
• Design provision and details that allow for potential future deck replacement;
• Plan and access for inspection and future maintenance and rehabilitation;
• Drainage plan;
• Bridge-deck protection plan;
• Reinforcing bar or prestressing steel corrosion protection plan for all reinforced or
prestressed concrete elements;
• Concrete element protection plan;
• Fatigue- and fracture-resistance plan for steel superstructure;
• Corrosion protection plan for steel superstructure;
• Design provisions and details for potential bearing replacement; and
• Mitigation plan for applicable extreme-event hazards.
2. Maintenance issues Identify recommended future maintenance requirements for achieving design service
life related to deck, superstructure, and substructure components:
• Deck maintenance plan;
• Drainage system maintenance plan;
• Steel superstructure coating maintenance plan;
• Expansion joint maintenance plan;
• Bearing maintenance plan; and
• Substructure maintenance plan.

Block 2. The next step is to identify feasible system alternatives that satisfy opera-
tional and site requirements, while also satisfying various provisions of the LRFD
specifications. This step would typically include

a. Preparing alternative bridge span arrangements that accommodate features


crossed, horizontal and vertical clearances, geometric requirements, and other
construction constraints. Geotechnical requirements affect foundation costs;
when possible, span layouts need to consider the relative costs between super-
structure and substructure in achieving optimum span lengths. Optimum span
lengths will also vary for different superstructure types, and the interaction
between optimal span and superstructure type needs to be considered in set-
ting span layouts.
b. Identifying feasible deck alternatives, such as CIP concrete, precast concrete,
or other. These choices are further discussed in Chapter 4.

118

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


c. Identifying feasible superstructure alternatives that accommodate geometric
and span-length requirements. Various bridge systems are described in Section
2.2. Feasible superstructure alternatives might include steel or concrete girders
or concrete segmental box girders, among others.
d. Identifying feasible substructure and foundation types that are compatible
with superstructure and geotechnical requirements.
e. Verifying preliminary member or element sizes at this stage through prelimi-
nary design.

Block 3. After feasible alternative bridge systems are identified, the next step is to
evaluate each alternative against factors that affect service life by using the fault tree
analysis described in Section 2.3. Within each system, each component, sub­system, and
element should be considered. Procedures in other chapters also need to be followed
in evaluating certain specific components or elements such as bridge deck, joints, and
bearings or in evaluating effects on materials such as ASR or freeze–thaw. A system-
level evaluation of these components and elements is necessary first to assure compat-
ibility among all components within the system.
Block 3a. In evaluating the various components, subsystems, and elements within
a system alternative, each branch of the fault tree illustrated in Figure 2.23 should be
followed. The initial categories to be considered include obsolescence, mostly pertain-
ing to inadequate capacity, or deficiency, pertaining to damage or deterioration. Defi-
ciency is further subdivided into deficiency caused by loads, natural or human-caused
hazards, or production or operation defects, which are shown in Figures 2.24, 2.25,
and 2.36, respectively. The fault tree branches end with basic events or the lowest
­levels of resolution, which are the individual factors to be considered. The individual
factors within fault trees for which the specific mitigation strategies need to be devel-
oped are placed in circle symbols.
Block 3b. Each factor is systematically examined and evaluated in regard to its
application to the various bridge system components, and a decision is made as to
whether that factor applies. For example, in the case of natural or human-caused haz-
ards, is the bridge in a thermal climate, which is a cold, wet climate with heavy use of
roadway deicing salt? If the bridge is not in this type of climate, this factor would not
apply, and the evaluation continues to the next factor. If the bridge is in this type of
climate, the evaluation proceeds to the next step.
Block 4. If the individual factor applies, the potential consequence of that fac-
tor should be evaluated and an appropriate strategy or strategies should be identified
to either avoid the factor or mitigate its influence. For example, if the bridge is in a
thermal climate, strategies to avoid or mitigate the potential of corrosion caused by
deicing salts would need to be identified. Special strategies such as stainless steel deck
reinforcement or metalizing the ends of steel girders under the deck expansion joints
need to be included. Strategies to mitigate various factors affecting service life are
given in Section 2.4. After determining the appropriate mitigation strategy, the evalu-
ation proceeds to consider the next factor.

119

Chapter 2. BRIDGE SYSTEM SELECTION


Block 5. After all factors are considered, the identified strategies are summarized
and integrated into the bridge system to develop a draft final configuration.
Block 6. This draft configuration then needs to be checked to be certain all require-
ments are satisfied and whether all identified strategies to improve service life are con-
sistent with one another and are compatible between various components and elements.
Block 7. If inconsistencies or incompatibilities are found, they need to be resolved
by making appropriate modifications to the strategies or to the affected components
or elements. For example, in evaluating a steel superstructure in a thermal climate, a
strategy might be to metalize the girder ends below the deck expansion joints at abut-
ments to mitigate the possibility of corrosion. A substructure evaluation, however,
might identify a strategy to eliminate expansion joints and use integral abutments.
These two strategies are inconsistent in that if the integral abutment strategy is imple-
mented, the strategy for metalizing steel girder ends could be unnecessary. A resolution
is made as to which strategy to implement.
Block 8. After all inconsistencies are identified and resolved and the final con-
figuration satisfies all requirements, the final configuration of the system alternative is
developed. It is feasible to develop more than one configuration capable of meeting the
service life requirements. See Chapter 1 for an example.
Block 9. The next step in the process is to predict the service life of the various
bridge components, subsystems, and elements within the final bridge system alterna-
tive under consideration. Deterioration models to quantitatively predict service life
are limited or nonexistent, so often the prediction is made on the basis of experience
or expert opinion. The predicted service lives of the various components, subsystems,
and elements are then checked to see if they will be equal to or greater than the owner-
specified service life of the bridge system.
Block 10a. If the predicted service lives of the various components, subsystems,
and elements are not equal to or greater than the owner-specified service life of the
bridge system, future rehabilitation and/or replacement of these components, sub­
systems, or elements will have to be anticipated at certain intervals during the system
service life. For example, if the bridge deck is predicted to have a service life less than
the system service life, the rehabilitation plan might include milling and overlaying at
an anticipated interval. A replacement plan would include the anticipated extent of
replacement and interval.
Block 10b. Future maintenance requirements will have to be identified whether
or not the component, subsystem, or element service life meets the system service life.
In the case of bridge decks, maintenance might include washing the deck surface to
remove salt or cleaning gutters and drains to permit proper drainage.
Block 11. After all requirements for replacement and/or rehabilitation and main-
tenance are determined, these requirements are incorporated into the final alternative
system configuration, which now includes the system layout with a compatible and
consistent set of service life strategies and a replacement, rehabilitation, and main-
tenance plan. The corresponding initial construction cost and life-cycle cost for this
complete alternative system configuration is then computed.
Block 12. When the evaluation is completed for this alternative, the engineer
should return to Block 3 and repeat the evaluation steps for the next identified bridge
120

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


system alternative. When all identified system alternatives have been evaluated, the
designer should proceed to the last step.
Block 13. The last step is to compare the final alternative bridge system configura-
tions and select the optimum alternative. Often this is done in a matrix-type evaluation
in which various key performance categories, determined specifically for the bridge,
are weighted and evaluated for each alternative. Example performance categories
might include service life, traffic impact, construction duration, construction complex-
ity, site suitability, local preference, or aesthetics. In this process, the advantages and
disadvantages of the various alternatives are compared. Initial construction costs and
life-cycle costs are also compared. With this type of evaluation, the optimum bridge
system can be identified as part of a complete cost–benefit selection process.
2.5.4
System Process Tables
The system selection process is further expanded in Table 2.2, which illustrates the
process design phases of the preliminary planning stage (also called the type, size, and
location stage) and the final design stage for a new bridge. Table 2.2 supplements the
information and examples provided in Sections 2.5.2 and 2.5.3.
2.5.5
Existing Bridges
Many of the service life considerations for new design are also applicable to existing
bridges. An inherent difference for existing bridges, however, is that the bridge system
has already been selected, designed, and constructed, and may have been in service for
a considerable time. An existing bridge has been subjected to factors affecting service
life and may have already experienced some level of deterioration. The level of deterio-
ration will dictate the type of rehabilitation required for restoration. The process for
restoring and extending the service life of an existing bridge involves a detailed system
evaluation rather than a system selection. The process for existing bridges will follow a
similar flowchart to that shown in Section 2.5.2, except that it will start at Block 3 and
will entail an evaluation of the existing system components, subsystems, and elements
against the various factors in the fault tree branches. Strategies for mitigating the factors
are more limited and are more focused on protecting the existing elements and materials.
However, they could also consider various forms of retrofit, such as converting simple
spans with deck joints to continuous spans without deck joints. The final evaluation is to
determine the optimal protection strategy that will extend the existing bridge service life.
2.6
Bridge Management
A major component of the Guide and this chapter addresses the type of data that should
be maintained for a bridge from design through fabrication, construction, and opera-
tion. The framework for this documentation is the introduction of the Owner’s Manual
described in Chapter 1. A bridge Owner’s Manual should be provided for unique bridges
or when requested by a bridge owner. The information to be included in the Owner’s
Manual is essential for proper future inspection and maintenance of the bridge in order
to achieve the bridge’s target design service life. The Owner’s Manual should be provided

121

Chapter 2. BRIDGE SYSTEM SELECTION


to the bridge owner just before opening the bridge to traffic. The bridge Owner’s ­Manual
is similar to the design calculation document that is usually provided to the bridge owner.
It is recommended that the Owner’s Manual be reviewed by an independent engineer.
Chapter 1 provides a more detailed description of the bridge Owner’s Manual.
Engineering judgment must be exercised in identifying the type of information to
be included in the bridge Owner’s Manual. A partial list of information for the bridge
Owner’s Manual includes the following:

• Target design service life of the bridge as determined by the owner;


• Overall process and philosophies used to address the service life design;
• List of all assumptions and special data used in the service life design process;
• All factors affecting service life that were identified in the initial and final service
life design, with adequate justification;
• All strategies that were designed into the bridge to avoid or mitigate factors affect-
ing service life;
• Procedure used to estimate the service lives of the bridge elements, components,
and subsystems;
• Description of any special steps or requirements that must be followed during
construction;
• Specific maintenance needs for various bridge elements, components, and sub­
systems in order to achieve their expected service lives;
• All considerations that were incorporated into the overall bridge system design to
accommodate rehabilitation or replacement of those items, including the expected
schedule;
• Identification of all “hot spot” areas of the bridge that would require special in-
spection or data to be collected during inspection that could be coordinated with
the FHWA-sponsored Long-Term Bridge Performance Program; and
• Health monitoring of unique bridges to develop a comprehensive bridge manage-
ment system that might be needed and should be described in detail.

The bridge Owner’s Manual should also describe how the bridge was designed,
constructed, and intended to function from an operational perspective, including

• Design loads, particularly any special vehicle types;


• Expected superstructure deflections;
• Expected movements at expansion joints and bearings;
• Relevant as-built data; and
• Construction methods and procedures.

In summary, the bridge Owner’s Manual should provide a clear picture of the pro-
cedure used to address service life design and what will be needed to keep the bridge
operational for the intended service life.

122

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


3
Materials

3.1
Introduction
This chapter provides essential information on materials used for constructing du-
rable bridge structures. It focuses on durability and service life issues related mainly to
concrete and steel, materials widely used in bridge construction, and provides limited
information on other material types used in bridge construction.
Information for enhancing the service life of materials used in bridge systems, sub-
systems, and components is summarized. Figure 3.1 identifies the materials-enhance-
ment selection process developed in this chapter, which begins with developing an
understanding of the types of materials. These viable materials are further evaluated
for the factors that adversely affect their service life, and individual strategies are devel-
oped to mitigate these adverse effects. The overall strategy selection is then developed,
blending these individual strategies that are sometimes in conflict with one another.
The components of an overall strategy should

• Identify appropriate design methodologies;


• Select durable material types, considering life-cycle costs;
• Consider additional protective measures, such as cathodic protection and electro-
chemical chloride extraction;
• Specify best practices for construction; and
• Develop an effective maintenance plan.

Sections 3.2 and 3.3 provide a general description of material types, primarily
concrete, reinforcement, and structural steel, used for bridge systems, subsystems, and
components. Section 3.4 addresses factors known to adversely affect the service life
of these materials and presents a fault tree approach for considering these factors.
123
Strategy Selection
(Section 3.6)
Material Factors Strategies Design
Types Affecting for methodology,
Evaluate Mitigate Develop
(Sections Service Life Enhanced material selection,
3.2 and 3.3) (Section 3.4) Service Life protective measures,
(Section 3.5) best practices for
construction,
maintenance plan.

Figure 3.1.  Materials enhancement selection process.


Figure 3.1. Materials enhancement selection process.

Section 3.5 provides available solutions, methods, technologies, and other informa-
tion helpful in developing individual strategies to mitigate factors adversely affecting
service life. Section 3.6 identifies a process to select the overall material selection and
protection strategy for providing materials with enhanced service life; however, much
of this selection process is highly dependent on the application and is addressed in
subsequent chapters.
3.2
Description of Material Types
This section describes concrete, reinforcement, and structural steel, the materials pri-
marily used for bridge systems, subsystems, and components.
Different concretes for bridge elements require different degrees of durability,
depending on the exposure environment and the properties desired. There are many
examples of longevity of concrete from ancient times; Rome’s Pantheon, for example,
was built around 126 CE and remains intact. The Confederation Bridge, which joins
the eastern Canadian provinces of Prince Edward Island and New Brunswick, was
constructed in 1997 and is a good example of a modern concrete bridge designed to
resist a harsh marine environment for at least 100 years (Figure 3.2).
Concrete has high compressive strength but low tensile strength, which makes
it prone to cracking. Early concrete structures were designed to be subjected only to
compression in order to avoid tensile failures. Arch shapes were used to span distances.

Figure 3.2.  Confederation Bridge.

124

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


In modern structures, reinforcement is common for providing tensile capacity, crack
control, and ductility. The service life of reinforced concrete structures depends on the
durability of the concrete and the durability of the reinforcement.
The major distress in reinforced concrete is due to the corrosion of the reinforce-
ment. Reinforcement must be protected from aggressive environments, either through
measures such as the use of the low-permeability concrete, adequate cover, and corro-
sion inhibitors, or the use of noncorrosive reinforcement. Preventive measures could
also be employed to prolong the service life of concrete structures, such as incorporat-
ing cathodic protection systems.
Structural steel is another common material used for bridge systems, subsystems,
and components. It provides high compressive and tensile strengths with consider-
able ductility, which makes it particularly suitable for long-span bridges. The most
common steel bridge systems used today are composite multigirder deck systems that
use either rolled beams, plate girders, or tub girders. These systems can be single or
multispan, and either straight or curved. Simple-span systems were often used in the
past, but most multispan systems today are continuous. Rolled beam bridges using
W-shapes are used in shorter spans up to about 100 ft for simple spans and up to about
120 ft for continuous spans. Welded deck plate girders are most often used for spans
over 120 ft.
3.2.1
Concrete
Concrete consists of cementitious material, aggregate, water, and admixtures. Con-
crete may also include fibers. The following sections provide general descriptions of
each concrete ingredient.
3.2.1.1
Cementitious Material
Cementitious materials include portland cements, blended cements, other hydraulic
cements, specialty cements for repairs, and supplementary cementitious materials
(SCMs). These cementitious materials have different chemical and physical properties
that affect the durability of concrete.
3.2.1.1.1
Portland Cement
Portland cement is produced from a combination of calcium, silica, aluminum, and
iron (Kosmatka and Wilson 2011). During processing, the raw materials reach tem-
peratures of 2,600ºF to 3,000ºF, forming clinker. Clinker and gypsum are ground to
a fine powder such that nearly all the material passes a No. 200 mesh (75-μm sieve).
Cement has four main compounds: tricalcium silicate (3CaO·SiO2 = C3S), dicalcium
silicate (2CaO·SiO2 = C2S), tricalcium aluminate (3CaO·Al2O3 = C3A), and tetracal-
cium aluminaferrite (4CaO·Al2O3Fe2O3 = C4AF). These compounds form the different
types of portland cements conforming to the requirements of ASTM C150. Ten types
of portland cement are summarized in Table 3.1.
Some cements are designated with a combined classification, such as Type I/II,
indicating that the cement meets all the requirements of the specified types.

125

Chapter 3. MATERIALS
Table 3.1. Cement Types and Uses
Type of Cement Use
Type I General purpose
Type IA Type I + air entraining
Type II General use, moderate sulfate resistance
Type IIA Type II + air entraining
Type II(MH) General use, moderate heat of hydration, moderate sulfate resistance
Type II(MH)A Type II(MH) + air entraining
Type III High early strength
Type IIIA Type III + air entraining
Type IV Low heat of hydration
Type V High sulfate resistance

3.2.1.1.2
Blended Cement
Blended cements are produced by intergrinding or blending two or more types of fine
material such as portland cement, ground granulated blast furnace slag, fly ash, silica
fume, calcined clay, other pozzolans, hydrated lime, and preblended combinations of
these materials (Kosmatka and Wilson 2011). ASTM C595 includes three classes of
blended cement:

1. Type IS (X)—portland blast furnace slag cement;


2. Type IP (X)—portland–pozzolan cement; and
3. Type IT(AX)(BY)—ternary blended cement.

The letters X and Y stand for the percentage of SCM included in the blended
cement, and A and B are the types of SCMs (S for slag and P for pozzolan). Type IS(X)
can include up to 95% slag cement. Type IP(X) can include up to 40% pozzolans. Spe-
cial properties are added after the percentages and are designated in the following list:

1. A—air entrainment;
2. MS—moderate sulfate resistance;
3. HS—high sulfate resistance;
4. MH—moderate heat of hydration; and
5. LH—low heat of hydration.

For example, IP(25)(HS) indicates 25% pozzolans and high sulfate resistance. As an
example for ternary cements, Type IT(S25)(P15) contains 25% slag and 15% p
­ ozzolans.
Type IT can meet MS, HS, and LH options.

126

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


3.2.1.1.3
Other Hydraulic Cement
All portland and blended cements are hydraulic cements. ASTM C1157 is a perfor-
mance specification that includes portland cement, modified portland cement (­specialty
cement providing characteristics of more than one type of portland cement), and
blended cements. ASTM C1157 recognizes six types of hydraulic cements:

1. Type GU—general use;


2. Type HE—high early strength;
3. Type MS—moderate sulfate resistance;
4. Type HS—high sulfate resistance;
5. Type MH—moderate heat of hydration; and
6. Type LH—low heat of hydration.

3.2.1.1.4
Expansive Cements
Expansive cements are modified hydraulic cements that expand slightly during the
early hardening period after setting. They are used to compensate for volume decreases
attributable to drying shrinkage, to induce tensile stress in reinforcement, and to sta-
bilize the long-term dimensions of posttensioned concrete. ASTM C845 designates
E-1 with three varieties (K, M, and S), and only K is available in the United States.
Type E-1 (K) contains portland cement, anhydrous tetracalcium trialuminosulfate, cal-
cium sulfate, and uncombined calcium oxide (lime).
It has been demonstrated that using lightweight aggregate (LWA) in concrete
with expansive cements is very beneficial for achieving the expansion from expansive
cements (Russell 1978). This effect is attributed to the continued presence of internal
moisture that has been absorbed by the LWA before batching. This internal moisture
provides a prolonged reaction time for the expansive cement, allowing it to achieve
greater expansion. It is suggested that LWA be used in conjunction with expansive
cements.
3.2.1.1.5
Specialty Cements for Repairs
Specialty cements for repairs are needed to achieve high early strengths, and cements
other than portland cements may be used. They may be rapid-setting hydraulic c­ ement,
gypsum-based cement, magnesium phosphate cement, or high-alumina cement for use
in partial depth concrete repairs (Caltrans 2008; ACPA 1998).
Gypsum-based cement mixtures contain calcium sulfates and gain strength r­ apidly.
They are placed easily, can be used at cold temperatures above freezing, and are toler-
ant to high ambient temperatures (Good-Mojab et al. 1993; NCHRP 1977). However,
they exhibit poor performance when placed in rainy and freezing weather conditions
(NCHRP 1977), and the presence of free sulfates has been found to promote corrosion
(Smith et al. 1991).

127

Chapter 3. MATERIALS
Magnesium phosphate cement mixtures have rapid setting time, high early strength,
low permeability, and good bonding to clean, dry surfaces. They are extremely sensi-
tive to water content, and extra water reduces their strength (NCHRP 1977). Also,
they are not used to repair concrete with limestone aggregates because the carbon
dioxide that forms weakens the bond (ACI 546, 2004; Smith et al. 1991).
High-alumina cement mixtures provide rapid strength gain, good bonding to
dry surfaces, and very low shrinkage (Smith et al. 1991). However, they experience
strength loss due to chemical conversions of some of their calcium aluminate hydrate
components (ACPA 1998).
Other rapid-setting materials are available but should be used with caution to
avoid adverse chemical reactions. For example, some rapid-hardening repair materials
contain high-alkaline–bearing materials, which may react with certain siliceous aggre-
gates to cause alkali-silica reactivity (ASR) (ACPA 1998).
3.2.1.1.6
Supplementary Cementitious Material
SCMs generally consist of by-products from other processes or natural materials. They
exhibit hydraulic or pozzolanic activity and contribute to the properties of concrete.
Pozzolanic materials do not possess cementitious properties, but when used with port-
land cement they form cementitious compounds. SCMs modify the microstructure of
concrete and reduce its permeability. They can reduce internal expansion caused by
chemical reactions. They can also reduce the heat of hydration, which can cause ther-
mal cracking. Typical examples are natural pozzolans, fly ash, slag cement, and silica
fume. They are used individually with portland or blended cements or in different
combinations. Fly ash is a finely divided residue that results from the combustion of
ground or powdered coal and is transported by flue gases. It is a by-product of power-
generating stations. Natural pozzolans found in nature may be calcined to induce sat-
isfactory properties. Fly ash and natural pozzolans are covered by ASTM C618. Slag
cement is a by-product of iron blast furnaces and consists essentially of silicates and
alumina silicates of calcium and other bases. Slag cement conforms to ASTM C989. It
is classified by performance in the slag activity test into three grades: Grade 80, Grade
100, and Grade 120. Silica fume is a finely divided residue resulting from the produc-
tion of elemental silicon or ferro-silicon alloys that is carried from the furnace by the
exhaust gases. It conforms to ASTM C1240. ASTM C1697 covers blended SCMs that
result from the blending or intergrinding of two or three ASTM-compliant SCMs.
SCMs are summarized in Table 3.2. SCMs improve durability by reducing permeabil-
ity and improving chemical resistance; however, the level of durability improvement
depends on the type, chemical and physical characteristics, and amount used. For ex-
ample, concrete with Class F fly ash is expected to have better sulfate resistance than
concrete with Class C fly ash.
In some areas, a superfine (ultrafine) grade of fly ash is available that exhibits char-
acteristics midway between normal fly ash and silica fume in cost, effectiveness, and
desirable dose rate. Unlike normal fly ash, it does not require large-volume batching
facilities, and it is not as difficult a material to handle and disperse effectively as silica
fume (Day 2006).

128

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 3.2.  SCMs: ASTM Standard Specifications and Use
SCM ASTM Standard Use
Fly ash, Class F C618 Low permeability, chemical resistance
Fly ash, Class C C618 Low permeability, chemical resistance
Fly ash, Class N C618 Low permeability, chemical resistance
Slag cement C989 Low permeability, chemical resistance
Silica fume C1240 Low permeability, chemical resistance
Blended SCMs C1697 Low permeability, chemical resistance

Superfine fly ash is generally processed from a Class F fly ash by passing the parent
ash through a classifier in which the coarse and fine particles are separated. The aver-
age particle size of raw or unprocessed fly ash is around 20 to 30 μm, and the largest
size is about 100 μm; however, an ultrafine fly ash has a maximum size less than 10 μm
and an average particle size of 2 to 4 μm. The finer size provides additional reactive
surface area, which contributes to the high-early-strength and low-permeability con-
crete. Early strengths and durability measures similar to those of silica fume concrete
were observed when a slightly higher dosage of ultrafine Class F fly ash was used with
a small reduction (10%) in water content (Obla et al. 2003).
3.2.1.2
Aggregates
Aggregates are granular material, such as sand, gravel, crushed stone, or iron blast
furnace slag, used in a cementing medium to form hydraulic-cement concrete. Ag-
gregates are divided into two categories: coarse and fine. Coarse aggregates are those
predominantly retained on a No. 4 sieve. Fine aggregates pass the No. 4 sieve and are
predominantly retained on the No. 200 sieve. However, the combination of coarse and
fine aggregates is important in concrete. The combined grading indicates the amount
of paste needed to affect the water demand and cement content. In comparison to
aggregates, paste is more porous, enabling easier transport of water and solutions,
which can be detrimental to durability. Normal-density fine and coarse aggregates
meet the requirements of ASTM C33. Structural low-density aggregates conform to
the requirements of ASTM C330. The aggregate characteristics that affect the proper-
ties of concrete are the grading, durability, particle shape and surface texture, abrasion
and skid resistance, density, absorption, and surface moisture. Angular, elongated, and
rough-textured aggregates have a high water demand that can lead to a high water-­
cementitious materials ratio (w/cm). The absorption and porosity of the aggregate may
affect the resistance to freezing and thawing. Aggregates should be free of potentially
deleterious materials such as clay lumps, shales, or other friable particles that can
affect the water demand and bond. The chemical composition of the aggregates is
important because of the possibility of expansive reactions.

129

Chapter 3. MATERIALS
3.2.1.2.1
Normal-Weight Aggregates
Most commonly used normal-weight aggregates (NWAs)—sand, gravel, and crushed
stone—produce concrete with a density of 140 to 150 lb/ft3. Aggregates must be clean,
hard, strong, durable particles, free of absorbed chemicals, coatings, and other fine
material that can adversely affect hydration and bond of the cement paste (Kosmatka
and Wilson 2011). The grading, shape, and texture of aggregates affect the water de-
mand. Concretes with angular and poorly graded aggregates are also difficult to pump.
Aggregates may also be reactive, causing ASR or alkali-carbonate reactivity (ACR). If
reactive aggregates are used, pozzolans and lithium-based admixtures can be added
to minimize the expansion and provide resistance to ASR. The space in the aggregate
also provides a place for the reaction products, if any expansion occurs as a result of
the other aggregates in the concrete. In ACR, diluting the aggregates or changing the
source, or selective quarrying, could minimize the expansion. In selective quarrying,
the layers that contain reactive aggregate are avoided (Ozol 2006). Reactivity is mea-
sured using ASTM C1105.
3.2.1.2.2
Lightweight Aggregates
In the United States, LWA is typically manufactured by expanding shale, clay, or slate
by firing the materials at high temperatures in a rotary kiln. At high temperatures,
gases are evolved within the pyroplastic mass, forming bubbles that remain after cool-
ing (ACI 213R, 2003b). The cellular structure of LWA results in a density that is lower
than that of NWA; when used in concrete, LWA reduces the density of concrete.
The porous structure of LWA absorbs more water than NWA. Pores close to
the surface are readily permeable, but interior pores are hard to fill. Many of the
interior pores are not connected to the surface at all. Water absorbed in the surface
pores by prewetting before batching is released into the paste during hydration of the
cement, which provides internal curing (Holm and Ries 2006), which in turn results in
improved properties and reduced cracking within the concrete.
LWA used in structural lightweight concrete (LWC) must be capable of producing
concrete with a minimum 28-day compressive strength of 2,500 psi with an equilib-
rium density between 70 and 120 lb/ft3 (ACI 213R, 2003b). The strength of LWA
varies with type and source and will affect the strength of LWC that can be achieved.
The density of LWA varies with particle size, increasing in density for the smaller
particles. Because of this density variation, the grading requirements for LWA (ASTM
C330) deviate from those of NWA (ASTM C33) by requiring a larger mass of the
LWAs to pass through the finer sieve sizes. This modification yields the same volumet-
ric distribution of aggregates retained on a series of sieves for both LWA and NWA.
3.2.1.3
Water
ASTM C1602 covers mixing water used in the production of hydraulic-cement con-
crete. Mixing water consists of batch water, ice, water added by truck operator, free
moisture on the aggregates, and water introduced in the form of admixtures. Potable
and nonpotable water are permitted to be used as mixing water in concrete. Nonpotable
130

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


water, including treated wash water and slurry water, is not used in concrete unless it
produces 28-day concrete strengths equal to at least 90% of a control mixture using
100% potable water or distilled water and time of set meets the limits in C1602. Some
excessive impurities may cause durability problems; therefore, optional chemical lim-
its for combined mixing water are given for chloride, sulfate, alkalis, and total solids.
Small solid particles increase the water demand as a result of large surface area.
3.2.1.4
Admixtures
Chemical admixtures are the ingredients in concrete other than portland cement, wa-
ter, and aggregate that are added to the mix immediately before or during mixing. Ad-
mixtures are primarily used to achieve the following objectives (Kosmatka and ­Wilson
2011):

• To reduce cost of concrete construction;


• To achieve the properties of hardened concrete effectively;
• To maintain the quality of concrete during mixing, transporting, placing, and cur-
ing in adverse weather conditions; and
• To overcome certain emergencies during concrete operations.

Chemical admixtures must conform to the requirements of ASTM C494 or, when
flowing concrete is applicable, to C1017. ASTM 494 covers the materials and the
test methods for use in chemical admixtures to be added to hydraulic-cement con-
crete mixtures in the field. The admixtures given in ASTM C494 and C1017 are
summarized in Table 3.3. The terms superplasticizer and high-range water-reducing
admixture are used interchangeably. Other admixtures, such as shrinkage-reducing or
viscosity-­modifying admixtures, are covered by Type S, specific-performance admix-
ture, in ASTM C494. Corrosion-inhibiting admixtures are covered by ASTM C1582,
and the cold-weather admixtures by ASTM C1622.

Table 3.3.  Admixtures: ASTM Standard Specifications and Use


Type ASTM Use
A C494 Water reducing
B C494 Retarding
C C494 Accelerating
D C494 Water reducing and retarding
E C494 Water reducing and accelerating
F C494 High-range water reducing
G C494 High-range water reducing and retarding
S C494 Specific-performance admixtures
I C1017 Plasticizing
II C1017 Plasticizing and retarding

131

Chapter 3. MATERIALS
Admixtures help in achieving workable, low w/cm, and low-permeability con-
cretes. Entrained air voids provide resistance to freezing and thawing, improved work-
ability, and reduced bleeding and segregation. Shrinkage-reducing admixtures reduce
shrinkage and cracking potential. Viscosity-modifying admixtures improve stability of
the mixture by minimizing segregation. Corrosion-inhibiting admixtures increase the
resistance to corrosion. They can form a protective layer at the steel surface as a result
of chemical reaction with ferrous ions (as with inhibitors containing calcium nitrite)
or they can provide a protective layer and reduce chloride ion ingress (as with inhibi-
tors containing amine/ester). ASTM specifications are available to provide guidance in
using admixtures in concrete, as summarized in Table 3.3.

3.2.1.5
Fibers
Fibers are added to concrete to control cracking. Steel fibers conform to ASTM A820,
and synthetic fibers meet the requirements of ASTM C1116, 4.1.3, Type III. Fibers, gen-
erally synthetic, are added to concrete at a low-volume dosage, about 0.1%, to reduce
plastic shrinkage cracking. At high-volume percentages, up to 2%, they can increase
resistance to cracking in hardened concrete and decrease crack width (­Kosmatka and
Wilson 2011). Plastic shrinkage cracks are addressed in bridge structures by proper
curing. Crack control in hardened concrete has been attempted in decks and overlays
(Ozyildirim 2005; Sprinkel and Ozyildirim 2000; Baun 1993). The cost and handling
of fibers have limited their use in bridge structures.
Synthetic fibers offer one other advantage: they reduce spalling during fires, which
has been a concern, especially in tunnels (Parsons et al. 2006). In a fire, spalling of
concrete can occur as a result of high vapor pressure. Spalling becomes more severe
with an increase in strength. Synthetic fibers such as polypropylene fibers reduce the
risk of spalling. At high temperatures these fibers melt, leaving pores or channels in
the concrete for the vapor to escape.
3.2.1.6
Concrete Types
Types of concrete used in bridge structures include normal-weight concrete (NWC);
high-performance concrete, including self-consolidating concrete (SCC); ultrahigh-
performance concrete (UHPC); fiber-reinforced concrete (FRC); and LWC. In repairs,
overlay concretes (latex-modified concrete, low-slump concrete, silica fume, polymer
concrete, and very-early-strength concretes), shotcrete, and grouts have been used.
The following sections describe different types of concretes and provide information
on their characteristics that can be considered in selecting the proper type of concrete
for a given application.
3.2.1.6.1
Normal-Weight Concrete
NWCs have a wide range of ingredients and performance characteristics including air
content, slump, and temperature. The density (unit weight) of NWC is approximately
145 lb/ft3. The density of concrete varies, depending on the amount and density of the

132

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


aggregate and the air, water, and cement contents. Generally, strength is the selected
parameter.
Concrete ingredients, proportions, handling, placing, curing practices, and the ser-
vice environment determine the ultimate durability and life of concrete (Ozyildirim
2007).
The durability of concrete depends largely on its ability to resist the infiltration of
water and aggressive solutions. Concretes that are protected from the environment can
provide a long service life. For longevity, concretes with low permeability are needed.
Concrete may deteriorate when exposed to cycles of freezing and thawing, especially
in the presence of deicing chemicals, and become critically saturated. The addition of
an air-entraining admixture to critically saturated concrete can improve the resistance
to freezing and thawing.
Concrete can resist most natural environments and many chemicals. However,
certain chemicals can attack concrete and cause deterioration. For example, sulfate
attack, ASR, ACR, acid attack, corrosion, and wear can damage concrete and reduce
its service life. However, in such environments proper material selection, proportion-
ing, and construction practices can protect concrete from unwanted attack and distress.
For example, if reactive aggregates are used, pozzolans and lithium-based admixtures
can be added to minimize the expansion caused by ASR.
3.2.1.6.2
High-Performance Concrete
High-performance concretes exhibit high workability, strength, and/or durability.
High-performance concrete has been used in decks, superstructures, and substruc-
tures to extend the service life. FHWA proposed defining high-performance concrete
by using long-term performance criteria (Goodspeed et al. 1996). The goal was to
stimulate the use of higher-quality concrete in highway structures. The definition of
high-­performance concrete developed for FHWA (Goodspeed et al. 1996) had four
performance parameters related to durability (see Table 3.4):

• Resistance to freezing and thawing;


• Resistance to scaling;
• Resistance to abrasion; and
• Resistance to chloride ion penetration.

The four structural design characteristics were compressive strength, modulus of


elasticity, shrinkage, and creep. The tensile strength, which is related to compressive
strength, modulus of elasticity, shrinkage, and creep, is an important factor affecting
the cracking potential. For each characteristic, standard laboratory tests, specimen
preparation procedures, and grades of performance were presented. Later additions
to this definition and changes to the grade limits were recommended (Russell and
Ozyildirim 2006) and are summarized in Table 3.5. These additions included resis-
tance to ASR and sulfate. Workability was also added as a characteristic; workability
affects durability because concrete should be well consolidated (either through vibra-
tion or self consolidation) in order to achieve the desired hardened concrete properties.

133

Chapter 3. MATERIALS
Table 3.4. Grades of Performance Characteristics for High-Performance Structural Concrete
Performance Standard Test FHWA High-Performance Concrete Performance Grade
Characteristic Method 1 2 3 4
Freeze–thaw durability AASHTO T 161 60% ≤ x ≤ 80% 80% ≤ x    
(x = relative dynamic
modulus of elasticity after ASTM C666
300 cycles) Procedure A
Scaling resistance (x = visual ASTM C672 x = 4, 5 x = 2, 3 x = 0, 1  
rating of the surface after
50 cycles)
Abrasion resistance (x = ASTM C944 2.0 > x ≥ 1.0 1.0 > x ≥ 0.5 0.5 > x  
average depth of wear in
mm)
Chloride permeability AASHTO T 277 3,000 ≥ x > 2,000 ≥ x > 800 800 ≥ x  
(x = Coulombs) 2,000
ASTM C1202
Strength (x = compressive AASHTO T 22 41 ≤ x < 55 55 ≤ x < 69 69 ≤ x < 97 x ≥ 97 MPa (x ≥
strength) MPa (6 ≤ x < MPa (8 ≤ x < MPa (10 ≤ x < 14 ksi)
ASTM C39 8 ksi) 10 ksi) 14 ksi)
Elasticity ASTM C469 24 ≤ x < 40 GPa 40 ≤ x 50 GPa x ≥ 50 GPa (x ≥  
(x = modulus of elasticity) (4 ≤ x < 6 × 106 (6 ≤ x < 7.5 × 7.5 × 106 psi)
psi) 106 psi)
Shrinkage ASTM C157 800 > x ≥ 600 600 > x ≥ 400 400 > x  
(x = microstrain)
Creep (x = microstrain/ ASTM C512 0.52 ≥ x > 0.41 0.41 ≥ x > 0.31 0.31 ≥ x > 0.21 0.21 ≥ x
pressure unit)

Source: Goodspeed et al. 1996 (available at http://www.fhwa.dot.gov/bridge/HPCdef.htm).

Table 3.5.  Additional Grades of Performance Characteristics


FHWA High-Performance Concrete
Standard Test Performance Grade
Performance Characteristic Method 1 2 3
Alkali-silica reactivity (ASR = expansion at ASTM C441 0.20 ≥ ASR > 0.15 ≥ ASR > 0.10 ≥ ASR
56 days, %) 0.15 0.10
Sulfate resistance (SR = expansion, %) ASTM C1012 SR ≤ 0.10 at 6 SR ≤ 0.10 at 12 SR ≤ 0.10 at 18
months months months
Workability (SL = slump, SF = slump flow) AASHTO T 119 SL > 7½ in. 20 ≤ SF ≤ 24 in. 24 in. < SF
ASTM C143 SF < 20 in.
ASTM C1611

Source: Russell and Ozyildirim 2006.

134

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Another characteristic to be considered is density. Concretes with densities varying
from light weight to normal weight can be prepared to address the dead load in span-
ning long distances.
3.2.1.6.3
Self-Consolidating Concrete
SCC is a highly flowable, nonsegregating concrete that can spread into place, fill the
formwork, and encapsulate the reinforcement without any mechanical consolidation
(ACI 237R, 2007b). SCC contains a large amount of fine material to obtain a stable
mixture; sometimes a viscosity-modifying admixture is used to provide the stability
instead of, or in combination with, the fine material. The high-range water-reducing
admixtures used are generally based on polycarboxylate ethers; they provide large
water reductions and slow slump loss. Eliminating the consolidation problem would
enhance the strength and reduce the permeability of concretes, essential characteristics
for longevity. SCC has been used in Japan since the late 1980s (ACI 237R, 2007b) and
is used widely in the precast industry in North America, but its use in the ready-mixed
concrete industry has been slower. Some of the benefits of SCC are decreased l­abor
requirements, increased construction speed, improved mechanical properties and
­durability characteristics, the ability to be used in heavily reinforced and congested
areas, consolidation without vibration, and a reduced noise level at manufacturing
plants and construction sites (Okamura and Ouchi 1999). The flow characteristic of
SCC is measured using the slump flow test (ASTM C1611). The stability of the mix-
ture can be qualitatively assessed in accordance with ASTM C1611 by using the visual
stability index. A visual stability index value of 0 indicates a highly stable mix with no
evidence of segregation or bleeding; 1 is a stable mix with no evidence of segregation
but with slight bleeding; and values of 2 and 3 indicate unstable mixtures. To deter-
mine the ability of SCC to pass through the reinforcement, a J-ring test is used (ASTM
C1621). ASTM C1610 addresses the determination of static segregation of SCC by
measuring the coarse aggregate content in the top and bottom portions of a column.
SCC is placed in a column separated into three sections, and the coarse aggregate in
the top and bottom sections is removed by washing. Another ASTM test method,
C1712, covers the rapid assessment of static segregation resistance of normal-weight
SCC. This test does not measure static segregation resistance directly, but rather pro-
vides an assessment of whether static segregation is likely to occur.
There are some concerns about the use of SCC, among them segregation due to a
high flow rate; a poor air void system and form pressure and tightness due to the high
fluidity; and high amounts of high-range water-reducing admixture resulting in coarse
air bubbles, shrinkage due to smaller maximum size aggregate, and lower amount
of coarse aggregate (ACI 237R, 2007b). SCC has been used successfully in beams
(Ozyildirim 2008), drilled shafts (Schindler and Brown 2006), substructure repairs,
and tunnel sections (ACI 237R, 2007b).

135

Chapter 3. MATERIALS
3.2.1.6.4
Ultrahigh-Performance Concrete
UHPC has high strength and high ductility. A proprietary UHPC that is commonly
available is discussed in this section; this UHPC is formulated by combining portland
cement, silica fume, quartz flour, fine silica sand, high-range water reducer, water, and
steel or organic fibers. Its superior durability and negligible permeability is expected
to reduce maintenance and extend the service life. Current applications (as of 2012),
including beams and connections, are explained in this section.
UHPC is expected to achieve compressive strengths greater than 21.7 ksi
(150 MPa), and it contains fiber reinforcement for improved ductile behavior (AFGC
2002). Small brass-coated steel fibers with a diameter of 0.007 in. (0.185 mm) and a
length of 0.55 in. (14 mm) are commonly used as fiber reinforcement in UHPC. A syn-
thetic fiber (polyvinyl alcohol) has also been used (Parsekian et al. 2008). To achieve
very high strengths, exceeding 30 ksi, UHPC is steam cured. UHPC exhibits strain
hardening that results in numerous tight cracks rather than one large crack prior to
failure. It has a very low w/cm (about 0.2) and a very dense matrix, leading to negli-
gible permeability. The high amount of binder and very low w/cm in UHPC make it
very cohesive; however, it flows within the forms without the need for vibration. The
high strength and low permeability of UHPC are attributed to a very dense packing of
fine material and a low w/cm (Graybeal 2006). UHPC contains no coarse aggregate.
Fine sand, typically between 0.006 and 0.024 in. (150 and 600 μm) is the largest
particle, followed by crushed quartz, cement, and silica fume. The resulting gradation
allows for tight packing of these particles. The unit weight of UHPC is approximately
155 lb/ft3. The coefficient of thermal expansion is about 50% higher than that of con-
ventional concrete (Graybeal 2006).
Whether steam cured or not, commonly available UHPC exhibits enhanced dura-
bility compared with NWC (Graybeal 2006; Graybeal and Tanesi 2007). Graybeal
and Tanesi (2007) studied UHPC characteristics under four curing regimes: steaming
at 60°C (194°F) and 95% relative humidity for 48 h (recommended by the manufac-
turer), untreated (no steam curing, specimens kept in the standard laboratory environ-
ment from demolding until testing), tempered steam treated (temperature in steam
chamber limited to 60°C [140°F]), and delayed steam treated (steam treatment initiated
after 15 days of casting). Regardless of the curing treatment applied, UHPC exhibited
enhanced durability properties over normal concretes. Thus, field casting and curing
can provide the desired properties without the need for steam curing. Steam curing can
improve UHPC properties even more, but are relevant to precast operations.
UHPC also exhibits strong freeze–thaw resistance. The relative dynamic modulus
of UHPC was at least 96% after being subjected to 690 cycles of freezing and thaw-
ing (more than two times the normal number of 300 cycles indicated in ASTM C666).
The concrete was innocuous to ASTM C1260 ASR deterioration, to ASTM C672
scaling deterioration, and to AASHTO T259 chloride penetration. The ASTM C1202
­Coulomb test result was negligible, less than 40°C, if any steam-based curing treat-
ment was applied, and ranged from very low (averaging 360°C) at 28 days to negligi-
ble (averaging 76°C) at 56 days in the absence of any steam curing. The ASTM C1202

136

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


test is based on electrical conductance of concrete, and the presence of steel fibers
affects the charge passed; however, the very dense matrix of UHPC isolates the steel
fibers and provides very high resistance to current flow. Ponding test results (­AASHTO
T259) have shown that the volume of chlorides that penetrate is extremely low. No
scaling was observed when tested in accordance with ASTM C672. The abrasion resis-
tance (ASTM C944) of steam-treated UHPC was very low; it abraded 0.1 to 0.3 g.
The untreated UHPC lost 10 times more (1 to 3 g) in the abrasion test. Although
air-entraining admixture is not added during casting, UHPC exhibits satisfactory resis-
tance to cycles of freezing and thawing.
The first bridge with UHPC beams was constructed in Wapello County, Iowa
(Moore and Bierwagen 2006). The three 110-ft beams were modified 45-in. Iowa
bulb-tee beams. To save material in the beam section, the web width was reduced by
2 in., the top flange by 1 in., and the bottom flange by 2 in. The Virginia Department
of Transportation (DOT) used five 45-in.-tall bulb-tee beams with UHPC in the bridge
on Route 624 over Cat Point Creek (Ozyildirim 2011). The bridge has ten 81.5-ft
spans, one of which contained UHPC beams. The steel fibers provided adequate shear
resistance, so the UHPC beams did not contain the conventional stirrups normally
used as shear reinforcement; however, confinement steel was included at the beam ends
(Ozyildirim 2011). Other applications in bridge structures have been accomplished or
recommended. Garcia (2007) detailed UHPC flexural behavior and offered a design
methodology for two-way ribbed, precast bridge decks. Under the FHWA Highways
for LIFE Program, a two-lane bridge on a secondary road in Wapello County, Iowa,
was constructed using prestressed concrete girders and 14 UHPC waffle deck bridge
panels (Heimann and Schuler 2010). UHPC has very high bond strength. At the
­Virginia project, the extra UHPC flowing to the top of the steel form had bonded
strongly to the form after hardening, making the removal of the form very difficult
(Ozyildirim 2011). The high bond strength, low permeability, and crack resistance
make UHPC highly desirable in joints and connections, and such applications have
been successfully completed (Perry and Royce 2010).
3.2.1.6.5
Fiber-Reinforced Concrete
FRC is expected to improve tensile strength; provide crack control; and increase dura-
bility, fatigue life, resistance to impact and abrasion, shrinkage, and fire resistance (ACI
544.1R, 1996). Crack control is critical for longevity in reinforced concrete structures,
and one effective solution is to use fiber-reinforced concrete. FRCs containing steel
and synthetic fibers are common. Two or more fiber types can also be used to produce
hybrid FRC to obtain improved properties or cost reduction. For example, steel and/
or macrosynthetic fibers that enhance toughness and postcrack load-carrying capacity
can be combined with microfibers that help control plastic shrinkage cracking.
The aspect ratio of fibers, which is the ratio of length to diameter, affects the work-
ability and the hardened concrete properties. Typical aspect ratios range from about
20 to 100 (ACI 544.1R, 1996). Fibers with high aspect ratios can improve the ductile
behavior of concrete and also increase strength and stiffness. However, fibers with high
ratios tend to interlock to form a mat, or ball, which is very difficult to separate by

137

Chapter 3. MATERIALS
vibration alone. In contrast, short fibers with ratios less than 50 are not able to inter-
lock and can easily be dispersed by vibration.
FRC can reduce the amount of cracks in concrete; however, wide cracks (>0.004 in.
[0.1 mm]) still exist. Cracks less than 0.004 in. (0.1 mm) wide inhibit the intrusion
of corrosive chemicals and are expected to hinder the intrusion of harmful solutions
(Wang et al. 1997; Lawler et al. 2002). To obtain tight cracks, high-performance FRC
is needed.
High-performance FRC undergoes large deflection and exhibits deflection or
strain hardening, causing multiple microcracks instead of one large localized crack.
As deflection occurs, strain hardening is exhibited after the first crack is initiated, and
an increase in stress occurs with further deformation. One such concrete is SIFCON
(slurry infiltrated fiber concrete) (Naaman and Homrich 1989), which is produced by
filling an empty mold with loose steel fibers (about 10% by volume) and filling the
voids with high-strength cement-based slurry. The resulting composite exhibits high
strength and ductility. UHPC with steel or polyvinyl alcohol fibers and engineered
cementitious composite with polyvinyl alcohol fibers also exhibit tight cracks (Li 2002,
2003). Both the UHPC and engineered cementitious composite are m ­ ortar mixtures
without the coarse aggregate. High-performance FRC with hybrid fibers (steel and
polyvinyl alcohol) containing #8 coarse aggregate with a nominal maximum size of
3
/8 in. was developed and exhibits deflection hardening needed for tight crack forma-
tion (Blunt and Ostertag 2009).
FRC can be used in bridge decks, deck repairs, and overlays. For example, the
Ohio DOT has used steel fibers in bridge-deck overlays containing silica fume or dense
concrete (Baun 1993). High-performance FRC can be used in joints, connections, and
link slabs.
3.2.1.6.6
Lightweight Concrete
LWC is typically used to reduce the dead load of a structure in order to improve
structural efficiency, thus allowing reduced element sizes, less reinforcement, increased
span lengths, fewer piers, or reduced foundation elements (ACI 213R, 2003b). As
prefabrication becomes more widely used for bridge elements, LWC can save money
by reducing handling, shipping, and erection costs. It has also been shown that LWC
provides enhanced durability by reducing the permeability and cracking tendency of
concrete and has a higher fire resistance than conventional concrete.
LWC consists of LWA or a blend of lightweight and normal-density aggregates.
Standard procedures and admixtures are used to proportion LWC mixtures. Standard
batching and transporting equipment are also used for LWC. LWC can have an equi-
librium density between 90 and 125 lb/ft3, but values from 110 to 120 lb/ft3 are most
common. Equilibrium density is defined in ASTM C567 as the density reached after
exposure to relative humidity of 50% + 5% and a temperature of 73.5°F + 3.5°F for
a period of time sufficient to reach constant mass. The fresh density is used for quality
control in the field and should also be used to compute precast element weights for
handling and shipping.

138

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


LWA is sometimes used in combination with NWA to create structural concretes
with densities between 120 and 145 lb/ft3. The properties such as strength and modu-
lus of elasticity would vary and should be addressed by performance requirements.
Although design compressive strengths of 3,000 to 5,000 psi are common for
LWC, design strengths up to 10,000 psi have been used in bridge beams (Liles 2010).
The maximum strength that can be achieved using an LWA source may be increased by
reducing the maximum aggregate size. As is typical with NWC, the use of a high-range
water-reducing admixture enables w/cm reduction, and with the addition of SCMs,
LWC with high workability, strength, and durability can be achieved.
LWA costs more than NWA because of the thermal processing used to manufac-
ture it. The impact of the increased cost of LWA on the difference in cost between LWC
and NWC depends on the cost of the LWA and the cost of the NWA that it is replacing.
As sources of good NWA dwindle, the use of LWA will become more cost-effective
because NWAs will also be shipped greater distances. In spite of the increased cost of
LWC, the benefits that can be achieved using LWC can make it a cost-effective solution
for concrete structures.
LWC can be placed and finished using conventional equipment. It exhibits a lower
slump than NWC with the same workability as a result of the reduced aggregate
density. In high-workability mixtures, lower-density LWA particles may rise to the
surface, unlike NWA, in which aggregates segregate by settling to the bottom. There-
fore, LWC mixtures should be designed to be cohesive, and excess vibration should
be avoided. Slump, air content, and temperature requirements for LWC are similar to
those for NWC. LWA should be prewetted before use in concrete that will be delivered
by pumping. Without adequate moisture, the aggregate may absorb mixing water and
cause slump loss during pumping. With proper preparation, LWC has been success-
fully pumped for long horizontal and vertical distances (Valum and Nilsskog 1999).
LWA suppliers can provide additional guidance in preparing for pumping LWC.
Experience shows that LWC with a proper air void system can be durable and
exhibit the satisfactory performance expected of NWCs (Holm and Ries 2006). Resis-
tance to freezing and thawing of LWC in the presence of deicing salts is similar to
that of NWC (ACI 213R, 2003b). Because LWC contains more absorbed water than
NWC, LWC should be allowed to dry before it is subjected to freezing and thaw-
ing. ASTM C330 requires 14 days of drying for the LWC tested in accordance with
ASTM C666.
Because of the porous structure of LWA particles, the resistance to wearing forces
may be less than that of a solid particle. However, in many cases, LWC bridge decks
have exhibited wearing performance similar to that of NWC (ACI 213R, 2003b)
because the aggregate is a vitrified material with hardness comparable to quartz. LWA
is nonpolishing, so it has excellent skid resistance.
The hardened properties of LWC are equal to or somewhat lower than NWC in
many cases. The modulus of elasticity of LWC is reduced from NWC. The splitting
tensile strength and the modulus of rupture values are lower for LWC, about 60% to
85% of NWC (ACI 213R, 2003b). Poisson’s ratio may be assumed as 0.20, which is
similar to NWC. The creep and shrinkage of LWC are similar to or a little higher than

139

Chapter 3. MATERIALS
that of NWC (Davis 2008). In general, the creep and shrinkage are higher at low-
strength LWC (i.e., for a bridge deck). However, it has been found that conventional
methods can be used to estimate prestress losses for LWC bridge girders (Kahn and
Lopez 2005). Although values for some properties may be lower for LWC than for
NWC, designs can usually be adjusted to account for the differences. In some cases,
differences may be offset by the reduced dead load in the structure.
The increased modulus of elasticity of LWC results in a high ultimate strain capac-
ity, which is beneficial in reducing the cracking tendency of concrete. The low level of
microcracking observed in LWC provides high resistance to weathering and corrosion.
ASR is not expected in concretes with LWAs, as the surface of the aggregate acts as
a source of silica. Silica reacts with the alkalis at an early stage to help counteract any
potential long-term disruptive expansion. The space in the aggregate also provides a
place for the reaction products if any expansion occurs as a result of other aggregates
in the concrete.
LWC is more fire resistant than NWC because of its lower thermal conductivity,
lower coefficient of thermal expansion, and the fire stability of the aggregate, which
is already exposed to high temperatures (over 2,000ºF) during processing (ACI 213R,
2003b).
3.2.1.6.7
Overlay Concrete
The use of overlays is described in additional detail in the bridge-deck section. Five
types of concretes that have been used in overlays are briefly described here.
Latex-modified concrete consists of a conventional concrete supplemented by a
polymeric latex emulsion (styrene–butadine–latex) (Russell 2004). The water in the
emulsion contributes to hydrating the cement. It has low permeability and, conse-
quently, good durability, and also has good bonding characteristics. It outperforms
conventional and low-slump dense concrete overlays and can be expected to last up to
25 years. Latex-modified concrete requires special mobile mixers, and proper curing is
needed. It is typically applied in thicknesses of 1.5 to 2 in.
Low-slump dense concrete has a moderate to high cement content and low w/cm
ratio (ACI 546R, 2004; Russell 2004). It has increased resistance to chloride ion pen-
etration. The main problems are that it is difficult to place, expensive, and prone to
surface cracking. It requires special equipment for proper consolidation, and proper
curing is critical. The standard has been to apply a 2-in. overlay.
Silica fume concrete is widely used to produce concrete with greater resistance to
chloride penetration (ACI 234R, 2006). It can be used effectively in thin overlays, in
similar thickness as the latex-modified concrete, and like latex-modified concrete, it
provides resistance to the penetration of chlorides. It is mixed in stationary mixers at
the plant or in ready-mix concrete trucks. Proper curing is necessary.
Polymer concrete is a composite material in which the aggregate is bound together
in a dense matrix with a polymer binder (ACI 546R, 2004). It provides low permeabil-
ity to water and aggressive solutions. Performance is highly dependent on the strength
of the bond between the overlay and the concrete underneath, which is also dependent
on surface preparation, cleanliness, and field application techniques. Most failures are

140

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


attributed to workmanship or improper handling of materials. Epoxy polymers have
a minimum thickness of 0.25 in. and are expected to last 10 to 15 years. Polymer con-
crete has been used in applications up to 1.5 in. Certain types have an expected life of
20 years, depending on the thickness of treatment.
Very-high-early-strength concrete is achieved using special blended cement with
high fineness and high aluminum oxide (Al2O3) and sulfur trioxide (SO3) (Sprinkel
1998). Very-early latex-modified concrete provides a reliable driving surface within a
few hours and reduces traffic interruption.
3.2.1.6.8
Shotcrete
Shotcrete is mortar or concrete pneumatically projected at high velocity onto a surface
(ACI 506R, 2005b). The high velocity of the material striking the surface provides
the compactive effort necessary to consolidate the material and develop a bond to the
existing surface. It contains coarse and fine aggregates, water, admixtures, and fibers.
The use of an air-entraining admixture improves the resistance of shotcrete to freezing
and thawing. The use of fibers improves toughness and gives load-bearing capacity
after cracking. Fibers also help in minimizing plastic shrinkage cracking. F ­ ibers used
in shotcrete are generally divided into two groups by their diameter (ACI 506.1R,
2008b). Fibers with equivalent diameters greater than 0.012 in. (0.3 mm) are known
as macrofibers; they are either steel or synthetic fibers. Macrofibers reduce crack prop-
agation, increase flexural toughness, and improve ductility and impact resistance. They
can provide resistance to drying shrinkage cracking and control crack widths at dos-
ages as low as 0.25% by volume. Fibers with diameters less than 0.012 in. (0.3 mm)
are known as microfibers. Microfibers used in shotcrete are mainly polypropylene or
nylon. They can provide resistance to plastic shrinkage cracking due to excessive mois-
ture loss at early ages at volume percentages as low as 0.1%.
Shotcrete is capable of being placed in vertical and overhead applications without
the use of forms (ACI 546R, 2004). There are two basic shotcrete processes: wet mix
and dry mix. In wet-mix shotcrete, ingredients are mixed and pumped through a hose
to a nozzle where air is added to project the material onto the surface. In dry-mix
shotcrete, cementitious material and aggregate are premixed and pumped and water is
added at the nozzle and projected onto a surface.
Shotcrete is frequently used for repairing deteriorated concrete bridge substructures.
It is also used for reinforcing structures by encasing additional reinforcing steel added to
beams, placing bonded structural linings on walls, and placing additional concrete cover
on existing concrete structures (ACI 546R, 2004). The success of shotcrete depends on
the material used and the skill of the nozzle operator. In repairs with irregular shapes,
shotcrete may be preferred because formwork is not needed. Shotcrete failures occur
mainly due to inadequate preparation of the existing surface, poor workmanship, and
not accounting for the relatively impermeable nature of shotcrete. The material may trap
moisture and contribute to critical saturation, which is harmful during cycles of freez-
ing and thawing. Wet-mix shotcrete is generally used where high production rates are
needed. Concrete trucks usually supply concrete for wet-mix shotcreting. In substructure
repairs, with small quantities of material, dry mix is commonly used.

141

Chapter 3. MATERIALS
3.2.1.6.9
Grout
Grout is a mixture of cementitious material and water, with or without aggregate, pro-
portioned to produce a pourable consistency without segregation of the constituents.
Grout is a common material used in repairs to fill cracks, honeycombed areas, and
interior voids and as a bonding agent. In new construction, it is used in open joints
and to fill tendon ducts (ACI 546R, 2004). It can be a hydraulic-cement grout or other
chemical grout such as a polymer–cement slurry, epoxy, urethane, or high-molecular-
weight methacrylate. Grouting can strengthen a structure, arrest water movement, or
both. Grout can be injected into an opening from the surface of a structure or through
holes drilled to intersect the opening in the interior. When injected from the surface,
short entry holes (ports), a minimum of 1 in. in diameter and a minimum of 2 in.
deep, are drilled into the opening. The surface of the opening is sealed between ports,
and grout is injected under pressure. Grouting is usually started with a relatively thin
grout, but is thickened when possible. Narrow cracks would be filled by injection
under pressure; however, wider cracks can be filled by gravity. Even though grouts
are used as a bonding agent, research has shown that concrete bonds well to existing
concrete, provided that the surface has been properly prepared.
Selection of type of grout depends on the magnitude of stress at the location, the
movement of the crack, the presence of solutions, crack width, required internal grout
pressure, setting characteristic, heat liberation (high for epoxy types) cost, compatibil-
ity with the existing concrete, penetrability, and bonding in the presence of moisture
(ACI 546R, 2004). Chemical grouts have different mechanical properties and are more
expensive than cement grout, and a high degree of skill is needed for satisfactory use
of chemical grouts. Some epoxy systems do not bond in the presence of moisture.
Chemical grouts can fill cracks as narrow as 0.02 in. (0.5 mm); for cement grouts the
minimum crack width is 3 mm. ASTM C1107 covers three grades of packaged, dry,
hydraulic-cement (nonshrinkable) grouts intended for use under applied load, such as
to support a structure or a machine, where change in thickness below initial placement
thickness is to be minimized. Rigid chemical grouts, such as epoxies, exhibit excellent
bond to clean, dry substrates, and some bond to wet concrete. These grouts can restore
the full strength of a cracked concrete member. ASTM C881 covers two-component,
epoxy–resin bonding systems for application to concrete that are able to cure under
humid conditions and bond to damp surfaces.
Grouts in tendon ducts hinder the penetration of chlorides to reach the steel and
bond the internal strands to the structure (Corven and Moreton 2004). Complete fill-
ing of the duct is essential for proper protection. The primary constituent of grout is
ordinary portland cement (Type I or II). SCMs are added to lower the permeability.
Prepackaged materials are preferred because a more uniform product can be obtained.
Total chlorides in grouts should be less than the specified limit of 0.08% by weight
of cementitious material as specified in Table 10.9.3-2 of Construction Specifications
in the LRFD Bridge Construction Specifications (AASHTO 2010a). Chlorides are
limited to ensure that the protective layer over the strand is not compromised. The
bleeding and resulting voids in grouts have been a concern. Usually the voids are

142

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


interconnected and facilitate the intrusion of harmful solutions. Similarly, shrinkage
should be controlled by using nonshrink grouts, so that cracks that can facilitate the
intrusion of chlorides are eliminated or minimized. Grout properties are given in Con-
struction Specifications, Table 10.9.3-2, along with the maximum total chloride ions
(AASHTO 2010a).
Grouts can be placed by pumping and vacuum injection. Vacuum injection is gen-
erally used after initial grouting and requires that the duct system be sealed. To ensure
that the duct is filled completely during construction, thixotropic grouts are used. These
grouts have very low viscosity after agitation, making them easy to pump. They stay
fluid when mechanically agitated or moved during pumping, but stiffen when at rest.
3.2.2
Reinforcement Material
Carbon steel (black steel) bars are commonly used as reinforcement in concrete. In the
high-alkaline environment of concrete (pH ≈ 13.0 to 13.8), an oxide layer forms on the
steel that protects the reinforcement from corrosion (Hartt et al. 2009). Chlorides pen-
etrating the concrete break down this protective layer and initiate corrosion. Carbon-
ation can lower the pH, increasing the corrosion rate. The iron corrosion products that
form on steel have much greater volume than the metal that is consumed in the corro-
sion reaction. This increase in volume causes tensile stresses in the concrete. When the
stresses exceed the strength of the concrete, cracks, spalls, and delaminations occur.
Alternatives to black steel are introduced to minimize the corrosion distress. These
corrosion-resistant reinforcements are expected to extend the service life of structures.
They include epoxy-coated steel, galvanized steel, low-carbon chromium steel, stain-
less steel (solid or clad), nickel-clad reinforcement and copper-clad reinforcement, tita-
nium, and fiber-reinforced polymers (FRPs).
3.2.2.1
Carbon Steel
Reinforcing steel is available in different grades and specifications. The grades vary in
yield strength, ultimate tensile strength, chemical composition, and percentage elonga-
tion. The grade designation is equal to the minimum yield strength of the bar in kips
per square inch (1 ksi = 1,000 psi). For example, Grade 60 rebar has a minimum yield
strength of 60 ksi.
Carbon steels are covered by ASTM specifications as follows: ASTM A615/A615M,
deformed and plain carbon steel bars for concrete reinforcement (covers Grades 40 and
60); ASTM A996 (replaces ASTM A616), rail steel (covers Grades 50 and 60) and ASTM
A617, axle steel (covers Grades 40 and 60); and ASTM A706 (Grade 60) for enhanced
weldability. A706 includes Grade 80; however, its use may be outside of consideration of
consensus design codes and specifications as noted in the ASTM specification.
3.2.2.2
Epoxy-Coated Reinforcement
Organic coatings were first investigated as a possibility for rebar protection in the
1970s, when FHWA began searching for an organic coating that would be best suited
for that purpose (Ramniceanu et al. 2008). After looking at 47 coatings, researchers

143

Chapter 3. MATERIALS
determined that the best candidates were four fusion-bonded epoxy powders (Kepler
et al. 2000).
3.2.2.3
Galvanized Reinforcement
Hot-dipped zinc-coated, or galvanized, steel reinforcement may provide superior per-
formance to that of uncoated steel (Kepler et al. 2000; Xi et al. 2004; Virmani and
Clemeña 1998).
3.2.2.4
Low-Carbon Chromium Steel
Typical carbon steels form a matrix of chemically dissimilar materials: carbide at the
grain boundaries and ferrite. In a moist environment, a microgalvanic cell forms that
initiates corrosion. Low-carbon chromium steel matrix is almost carbide free, mini-
mizing the galvanic action. Low-carbon chromium steel has low carbon content, less
than 0.15%, and contains from 8% to 10.9% chrome (ASTM A1035). It exhibits
strength and toughness (not brittle), and it is significantly more corrosion resistant
than conventional steel. The chloride corrosion threshold for initiation of corrosion
was found to be four times that of black bar (Hartt et al. 2009).
3.2.2.5
Stainless Steel
Stainless steels are chromium-containing steel alloys with a minimum chromium con-
tent of 10.5% (Markeset et al. 2006). Corrosion-resistant stainless steel contains a
minimum of 12% chromium (Scully and Hurley 2007). An iron-chromium-rich oxide
layer protects the stainless steel. Additional alloying elements such as nickel, molybde-
num, and titanium are added for improved corrosion resistance.
ASTM Specification A955 covers three grades of stainless steel bars with minimum
yield strengths of 40, 60, and 75 ksi. Another type of stainless steel reinforcement,
stainless steel–clad bars, is available. The cladding is a barrier coating used to prevent
corrosive agents from coming in contact with the core of the bar. The improved cor-
rosion resistance of stainless steel is due to a thin chromium oxide film that is formed
on the steel surface. Oxygen is required for the film formation. Other typical alloying
elements are molybdenum, nickel, and nitrogen. Nickel is mostly alloyed to improve
the formability and ductility of stainless steel. The four major types of stainless steel
(Markeset et al. 2006) are as follows:

• Martensitic (not for reinforcement);


• Ferritic. This has properties similar to mild steel but with better corrosion resis-
tance, even though in the lower range of corrosion resistance for reinforcement;
• Austenitic. This is the most widely used type of stainless steel and is rated in the
higher range of corrosion resistance for reinforcement; and
• Austenitic-ferritic (duplex). The duplex structure delivers both strength and ductil-
ity, and is rated in the very high range of corrosion resistance.

144

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


It is apparent that stainless steel has varying properties and corrosion-resisting
potential, and studies are continuing to identify them for use as reinforcement (Hartt
et al. 2009).
3.2.2.6
Nickel-Clad Reinforcing Bars
Nickel coatings at least 0.025 mm (0.001 in.) thick provide good corrosion resistance
(Virmani and Clemeña 1998). Even if the coating is damaged, the underlying diffusion
zone of alloyed nickel and iron provides additional corrosion protection. A concern is
the high cost of nickel.
3.2.2.7
Stainless Steel–Clad Reinforcing Bars
Stainless steel–clad bars are an attractive alternative to solid stainless steel from the
standpoint of both corrosion mitigation and cost. One would ideally gain the resis-
tance to corrosion of solid stainless steel at a fraction of the cost of solid single-phase
stainless steel (Scully and Hurley 2007). Concerns about using stainless steel–clad rein-
forcement include (1) the difficulty encountered bonding the cladding to the bars and
(2) the possibility that if areas of carbon steel are exposed because of mechanical dam-
age (e.g., construction site handling or unsealed cut ends), those localized areas will
corrode (Darwin et al. 1999; Scully and Hurley 2007). The corrosion resistance of clad
bars is dependent on any defects that expose the carbon steel core. The performance is
similar to solid stainless steel when intact, and when defective, it is similar to that of
carbon steel rebar (Scully and Hurley 2007).
3.2.2.8
Copper-Clad Reinforcement
Copper coatings of about 0.5 mm (0.02 in.) provide good corrosion resistance ­(Virmani
and Clemeña 1998). However, copper can retard the hydration of cement. This retar-
dation can lead to unhydrated cement particles around the bars, even though the paste
has hardened. Copper-clad bars are expected to be cost-effective.
3.2.2.9
Titanium
Titanium is highly resistant to corrosion; however, the cost can be five times that of
stainless steel (MacDonald 1995).
3.2.2.10
Fiber-Reinforced Polymer
FRP composite bars and strands are commercially available and have been used in
structures; their use in highway infrastructure is discussed in NCHRP Report 503
(Mertz et al. 2003).
FRP composite materials are light in weight and easy to construct; provide excellent
strength-to-weight characteristics; and can be fabricated for made-to-order strength,
stiffness, geometry, and other properties (ACI E2, 2000; Mertz et al. 2003). In addi-
tion, FRP is nonmagnetic. FRP composite materials may be the most cost-effective

145

Chapter 3. MATERIALS
solution for repair, rehabilitation, and construction of portions of the highway infra-
structure (Mertz et al. 2003).
FRP is composed of a polymer matrix, either thermoset or thermoplastic, rein-
forced with fiber or other reinforcing material (ACI 440R, 2007a). An anisotropic
material, its most favorable properties are in the direction of the fibers. FRP perfor-
mance depends on the fiber, the matrix, and the interaction of the two. The resin or the
polymer holds the fibers in place, the fibers provide the mechanical strength, and the
fillers and additives aid in processing and performance. Principal types of fibers used
in structural applications are glass, carbon, and aramid.
FRP bars are generally made of glass fibers embedded in a matrix (thermoset or
thermoplastic resins). The properties of the FRP bars, such as high-temperature perfor-
mance, corrosion resistance, dielectric properties, flammability, and thermal conduc-
tivity, are mainly derived from the properties of the matrix. Depending on the matrix
used, the mechanical properties of FRP bars (e.g., tensile strength, ultimate strain, and
Young’s modulus of elasticity) might degrade under specific environmental conditions
such as alkaline environment and moisture (water and salt added to water) as follows:

• Effects of the alkaline environment. The alkaline water contained in concrete pores
may cause degradation, which might cause damage to the polymeric matrix; and
• Effects of moisture (water and salt added to water). The main effects of moisture
absorption by the matrix may result in strength reduction and stiffness reduction
(less pronounced) in the FRP. The absorption of moisture depends on the type of
polymeric matrix, the matrix interface, and the quality of the bars.

Glass fibers are available as E-glass (electrical grade), high-strength (S-2) glass, and
improved acid resistance (epoxy-coated reinforcement (ECR)) and alkali-resistant (AR)
glass (ACI 440R, 2007a). These glass fibers generally range in diameter from 9 to 23 μm.
Fibers are drawn in at high speed through small holes in electrically heated bushings to
form the individual filaments. The filaments are coated with a chemical binder or sizing
for protection and to enhance the composite properties at the fiber–matrix interface.
The filaments are gathered into groups or bundles called strands or tows.
E-glass fibers are the most susceptible to degradation caused by alkalinity and
moisture and must be protected by the appropriate resin. Carbon fibers are inert to the
environment. Different fiber systems and resin systems provide different levels of resis-
tance to environmental conditions such as moisture, alkaline solution, UV radiation,
or extreme temperatures. Selection of the proper reinforcement and the proper resin is
needed for longevity. The coefficient of thermal expansion of the FRP composites with
glass fibers is higher than that of the concrete. The difference should be considered
when FRP is in direct contact with concrete. At low temperatures and exposure to
cycles of freezing and thawing fibers are not affected, but the resin and the fiber–resin
interface are affected. Cycles of freezing and thawing and the presence of road salts
may result in microcracks and gradual degradation.
Polymer resins exhibit high creep and relaxation behavior; addition of fibers
increases the creep resistance. Thermosetting resins are more resistant to creep than the
thermoplastic resins. FRP generally has good fatigue performance because the fibers

146

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


have minimal defects and the matrix resists the propagation of cracks. Several draw-
backs of FRP are that it does not exhibit ductile failure mode as steel, it is the lower
elastic modulus, and the strands are difficult to grab, making anchorage details critical.

3.2.3
Structural Steel
This section describes the various steel types, including currently used types and older
types found in older bridges, and describes various characteristics that are important
for durability and long-term service life.

3.2.3.1
Current Steel Grades
Currently, there are seven structural steel grades for bridges covered by ASTM A709
(AASHTO M-270) specifications with four yield strength levels (ASTM A709-11).
Table 3.6 identifies these current types. Included are ASTM reference standard grades,
along with the steel descriptions and yield and tensile strengths. The letter W is at-
tached to the grade number to designate steel that has weathering capability. The letter
S is attached to designate special steel grade for structural shapes.
Grades 36, 50, 50W, HPS 50W, and HPS 70W are available with the properties
shown in plate thicknesses up to 4 in. Properties for HPS 100W will vary depending
on plate thickness, as shown in Table 3.6.
Grade 36 steel, which has been a highly used grade over the years, includes basic
carbon steel shapes, plates, and bars for use in riveted, bolted, or welded construction.
Grade 50 and 50S steel includes high-strength, low-alloy columbium-vanadium
structural steel for shapes, plates, bars, and sheet piling and is intended for riveted,
bolted, or welded construction.

Table 3.6. Current ASTM A709 Steel Grades (ASTM A709-11)


Additional
Reference Yield Tensile
Grade Standard Name or Designation Strength (ksi) Strength (ksi)
36 A36 Structural carbon steel 36 (min) 58 to 80
50 A572 High-strength low-alloy 50 (min) 65 ksi (min)
columbium-vanadium steel
50S A992 Structural steel shapes 50 to 65 65 (min)
50W A588 High-strength low-alloy 50 (min) 70 (min)
weathering steel
HPS 50W High-performance steel 50 (min) 70
HPS 70W High-performance steel 70 (min) 85 to 110
HPS 100W (to 2.5 in. thick) High-performance steel 100 (min) 110 to 130
HPS 100W (over 2.5 in. to 4 in.) High-performance steel 90 (min) 100 to 130

Note: min = minimum.

147

Chapter 3. MATERIALS
Grade 50W includes high-strength, low-alloy weathering steel shapes, plates, and
bars for welded, riveted, or bolted construction, but it is intended primarily for use in
welded bridges in which savings in weight and added durability are important.
New high-performance steels have been added with three grades as shown in
Table 3.6. These new steels are further described in Section 3.2.3.6.
Table 3.7 identifies older steel grades that are still available, but AASHTO and
A709 have replaced them with HPS 70W and HPS 100W grades because of improved
properties.
3.2.3.2
Steel Properties
Steel’s mechanical properties are those that characterize its elastic and inelastic behav-
ior under stress and strain. Such properties include parameters that are related to the
material’s strength, ductility, and toughness. Other parameters, such as weldability,
machinability, and weathering, are also important in regard to the ease with which
the material can be fabricated and achieve long-term durability (Barson 1994). The
following discussion of various steel properties is adapted from Barson (1994) and
Errera (1964).
Strength is represented by the material’s yield strength and ultimate tensile strength.
Ductility is the ability of a material to undergo large plastic deformations without
fracture. It is represented by the amount of elongation that the material can experience
after yielding and before reaching its ultimate strength. Ductility is an important mate-
rial property because it allows redistribution of high local stresses that occur in welded
connections and at regions of stress concentration such as holes or geometric changes.
Toughness is the capacity of a material to absorb large amounts of energy before
fracture. It is related to the area under the stress–strain curve and is dependent on
strength and ductility. The larger the area under the stress–strain curve, the tougher
the material.
Weldability is the material’s capability to withstand welding without seriously
impairing its mechanical properties. Weldability varies considerably for different types
of steels and different welding processes.
Machinability is the ease with which a material can be sawed, drilled, or otherwise
shaped without seriously impairing its mechanical properties.
Weathering is the material’s capability to resist corrosion in a given environment.

Table 3.7. Older Structural Steel Grades Still Available


Additional Yield Tensile
Reference Strength Strength
Grade Standard Name or Designation (ksi) (ksi)
70W A852 High-strength low-alloy quenched and tempered steel 70 (min) 90 (min)
100/100W A514 High-strength quenched and tempered alloy steel 100 (min) 110 to 130

Note: min = minimum.

148

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


The steel’s mechanical properties are affected by the following three factors:

1. Chemical composition,
2. Processes used to transform the base metal into the final plate or structural shape
product, and
3. Heat treatment.

The following sections describe these three factors.


3.2.3.3
Chemical Composition
Chemical composition is a primary factor in determining the properties of a steel type.
Structural steels are made up of iron and carbon with varying amounts of other ele-
ments, primarily manganese, phosphorus, sulfur, and silicon. Specific properties are
achieved by the addition or presence of these and other elements in various combina-
tions and proportions (Barson 1994)
In carbon steels, the elements carbon and manganese have a controlling influ-
ence on strength, ductility, and weldability. Grade 36 carbon steel is more than 98%
iron, with about 0.25% carbon and about 1% manganese by weight. These percent-
ages change dramatically among the various grades. Carbon increases the hardness
or tensile strength, but it has adverse effects on ductility and weldability. To counter-
act these adverse effects, small amounts of various alloying elements are sometimes
used to increase the ability of a steel to achieve higher strength while maintaining a
lower percentage of carbon content. Phosphorus and sulfur typically have a harmful
effect, especially on toughness, and the fractional percentages of these elements must
be kept low. Small amounts of copper increase corrosion resistance and help develop
the weathering steel grades; fractional percentages of silicon are used mainly to elimi-
nate unwanted gases from the molten metal; and nickel and vanadium have a generally
beneficial effect on steel behavior (Errera 1964).
Table 3.8 lists advantages and disadvantages of commonly used alloy elements and
includes approximate chemical composition used in plate steel grades for comparison.
It is interesting to note how each element percentage changes in progressing from
Grade 36 to 50 to 50W to HPS 50W and finally to HPS 100W. The progression in steel
grades increases in strength, weathering characteristics, and ultimately in toughness
and weldability, and the various chemical percentages illustrate how these properties
are partially achieved. The chemical composition of Grade HPS 70W is same as HPS
50W. Grade 50S is not shown.
3.2.3.4
Rolling and Shaping Processes
The process used to transform the base metal into the final plate or structural shape
product is another factor affecting steel’s mechanical properties. Hot rolling is typi-
cally the process used to shape steel. It consists of passing the material between two
rolls revolving in opposite directions, with the distance between the rolls less than
the thickness of the original entering material. The rolls grip the piece and reduce its
cross-sectional area, increasing its length and to a lesser extent, its width. Because of

149

Chapter 3. MATERIALS
Table 3.8. Effects and Composition of Alloying Elements
Alloy Typical
Element Advantage Disadvantage Steel Grade Composition
Carbon Principal hardening element in Decreases ductility, toughness, 36 0.25%
steel. Increases strength and and weldability. 50 0.23%
hardness. 50W 0.19%
HPS 50W 0.11%
HPS 100W 0.08%
Manganese Increases hardness and High amounts can cause 36 0.8% to 1.2%
strength. Controls harmful embrittlement and reduce 50 1.35%
effects of sulfur. weldability. 50W 0.8% to 1.25%
HPS 50W 1.1% to 1.35%
HPS 100W 0.95% to 1.5%
Phosphorus Increases strength and Generally considered an 36 0.04%
hardness. Improves corrosion impurity. Decreases ductility 50 0.04%
resistance. and toughness. 50W 0.04%
HPS 50W 0.02%
HPS 100W 0.015%
Sulfur Improves machinability. Generally undesirable. 36 0.05%
Decreases ductility, toughness, 50 0.05%
and weldability. Adversely 50W 0.05%
affects surface quality. HPS 50W 0.006%
HPS 100W 0.006%
Silicon Used to deoxidize molten Decreases weldability. 36 0.4%
steel. Increases strength and 50 0.4%
hardness. 50W 0.3% to 0.65%
HPS 50W 0.3% to 0.5%
HPS 100W 0.15% to 0.35%
Aluminum Used to deoxidize molten None. HPS 50W 0.01% to 0.04%
steel. Refines grain size, thus HPS 100W 0.02% to 0.05%
increasing strength and
toughness.
Vanadium Used to refine grain size. Small High amounts reduce hardness. 50 0.01% to 0.15%
additions increase strength and At high finishing temperature 50W 0.02% to 0.1%
toughness. may be detrimental to HPS 50W 0.04% to 0.08%
finishing.  HPS 100W 0.04% to 0.08%
Columbium Small additions produce finer None. 50 0.005% to 0.05%
grain size, which increases HPS 100W 0.01% to 0.03%
strength and toughness.
Nickel Increases strength and Cost. 50W 0.4%
toughness and improves HPS 50W 0.25% to 0.4%
corrosion resistance. HPS 100W 0.65% to 0.9%
Chromium Increases strength and Cost. 50W 0.4% to 0.65%
corrosion resistance. Used in HPS 50W 0.45% to 0.7%
weathering steel. HPS 100W 0.4% to 0.65%

(continued)

150

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 3.8. Effects and Composition of Alloying Elements (continued)
Alloy Typical
Element Advantage Disadvantage Steel Grade Composition
Molybdenum Increases strength, weldability, Cost. HPS 50W 0.02% to 0.08%
toughness, and corrosion Delays rather than eliminates HPS 100W 0.4% to 0.65%
resistance. temper embrittlement.
Copper Increases corrosion resistance. May adversely affect notch 50W 0.25% to 0.4%
Used in weathering steel. toughness.  HPS 50W 0.25% to 0.4%
HPS 100W 0.90% to 1.2%
Nitrogen Increases strength and Decreases ductility and HPS 50W 0.015%
hardness. toughness. HPS 100W 0.015%

Note: Advantages and disadvantages are from the Highway Structures Design Handbook (Vol. 1, Chapter 3, Properties of
Bridge Steels, Barson 1994) and the Metals Handbook (ASM 1998); percent composition is from ASTM A709-11.

the rolling, the steel fracture toughness is greater in the direction of rolling than in the
direction perpendicular to it. Further, a greater reduction of cross section during the
normal hot-rolling process will produce a greater yield and tensile strength (Barson
1994).
Some steels are purposely cold rolled to obtain higher strength levels. The cold
working strain hardens the material.
3.2.3.5
Heat Treatment
A variety of heat treatments used in the steel-making process develop certain desirable
characteristics in steel. These heat treatments can be divided into two groups: slow
cooling and rapid cooling. Slow-cooling treatments include annealing, normalizing, and
stress relieving, which decrease hardness, promote uniformity of microstructure, and
improve ductility and fracture toughness. They also improve machinability or facili-
tate cold forming, and they relieve internal stresses. Rapid-cooling treatments, such as
quenching and tempering, increase strength, hardness, and toughness (Barson 1994).
Quenching and tempering consist of heating steel to a high temperature (about
1,650°F) long enough to produce a desired microstructural change and then quenching
by immersion in water. After quenching, the steel is tempered by reheating to a specific
temperature (usually between 800°F and 1,250°F), holding for a specified time, and then
cooling under suitable conditions to obtain the desired properties. The rapid cooling by
quenching increases strength but reduces ductility. Tempering restores part of the duc-
tility, but gives up some of the strength gained by the quenching. This process permits
attainment of higher strengths while retaining relatively good ductility (Barson 1994).
3.2.3.6
High-Performance Steel
A cooperative research program between FHWA, the American Iron and Steel In-
stitute, and the U.S. Navy was launched in 1994 to develop new high-performance
steels for bridges. The driving force was the need to develop improved higher-strength,

151

Chapter 3. MATERIALS
higher-toughness steels with improved weldability. Improvements in these character-
istics would improve the overall quality and ease of fabrication of bridge steels used
in the United States. Further, the steel would have weathering characteristics, with the
added designation “W.” Three strength grades were developed and are now available
for general use: HPS 50W, HPS 70W, and HPS 100W. Grades HPS 70W and HPS
100W have now replaced the older high-strength, low-alloy quenched and tempered
steel (AASHTO M270 Grade 70W) and the high-strength quenched and tempered
alloy steels (Grades 100 and 100W). It is the intent that the newer high-performance
steels should be used at these higher strength levels because of their enhanced proper-
ties. The older steels are still available, but their use is discouraged (FHWA 2002).
HPS 70W is the most widely used grade in the HPS group because its reduced
weight can make it more economical than conventional Grade 50W steel. HPS 70W is
produced by quenching and tempering or thermal-mechanical controlled processing.
Because quenching and tempering processing limits plate lengths to 50 ft in the United
States, thermal-mechanical controlled processing practices have been developed to
produce HPS 70W up to 2 in. thick and up to 125 ft long, depending on the weight.
HPS 50W steel, which has the same chemistry as HPS 70W, is produced using con-
ventional hot rolling or controlled rolling up to 4 in. thick in lengths similar to Grade
50W steel (FHWA 2002).
A major advantage of high-performance steel is its increased fracture toughness,
which is much higher than that of conventional bridge steels. This is evident from
Figure 3.3, which shows the Charpy V-notch transition curves for HPS 70W and con-
ventional Grade 50W steel. The brittle-ductile transition of HPS 70W occurs at a
much lower temperature than conventional Grade 50W steel. This means that HPS
70W remains fully ductile at lower temperatures, while conventional Grade 50W steel
begins to show brittle behavior (FHWA 2002).
The current AASHTO Charpy V-notch toughness requirements are specified to
avoid brittle failure in steel bridges above the lowest anticipated service tempera-
ture. The HPS 70W steels tested so far show ductile behavior at the extreme service

Figure 3.3.  Charpy V-notch transition curves


for 50W and HPS 70W steels.
Source: FHWA 2002.

152

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


temperature of –60°F, which corresponds to Zone 3, with a minimum service tempera-
ture of below –30°F to –60°F (FHWA 2002).
With higher fracture toughness, high-performance steels have much higher crack
tolerance than conventional-grade steels. Full-scale laboratory fatigue and fracture
tests of precracked I-girders made of HPS 70W steel showed that the girders were able
to resist the full design overload without fracturing even when the initial crack was
large enough to cause a 50% loss in tension flange net section. This is a major feature
that can increase service life of bridges using HPS steels because a large crack tolerance
increases the time for detecting and repairing fatigue cracks before the bridge becomes
unsafe (FHWA 2002).
3.2.3.7
Weathering Steels
Uncoated weathering-grade steels have been used for bridge applications since the mid-
1960s. When used in the right environment, these steels result in both initial and life-
cycle cost savings as they eliminate the need for shop and field painting and can provide
over 100 years of service life with minimal maintenance. Weathering-grade steels are
currently supplied under AASHTO Specification M270, ASTM A709 Grade 50W, and
in high-performance steel Grades HPS 50W, 70W, and 100W.
These steels have a special alloy composition with small amounts of copper, phos-
phorus, chromium, nickel, and silicon that makes them able to resist corrosion with-
out any applied coating. During initial exposure to the elements, these steels form a
dense, tightly adhering patina that is essentially a natural oxide coating, about the
same thickness as a heavy coat of paint. It is this patina that protects the steel and
enables it to resist further atmospheric corrosion (McEleney 2005).
Weathering steel bridges initially look orange-brown in color; the color darkens as
the patina forms. In 2 to 5 years, depending on the climate, the steel will attain a dark,
rich, purple-brown color that many think is attractive (McEleney 2005).
Guidelines for proper application of weathering-grade steels in highway structures
and recommendations for maintenance to ensure continued successful performance
were issued by FHWA in Technical Advisory T-5140.22 (FHWA 1989).
Despite its many advantages, there are certain conditions where the use of uncoated
weathering-grade steel should be considered with caution. Proper weathering behavior
of these steels depends on alternating cycles of wet and dry conditions that allow a
proper patina or protective layer to form (McEleney 2005). In environments of contin-
ual wetting, such as areas with frequent high rainfall, high humidity, or persistent fog
or with low-level water crossings, the patina development can be adversely affected,
resulting in continuing corrosion (FHWA 1989).
Exposure to salt is also detrimental to proper weathering. This can occur in marine
coastal areas where the steel is subject to salt-laden air or spray. It can also occur in
certain grade separation conditions where a tunnel effect is created by the combina-
tion of a bridge with minimum clearance over a depressed roadway section between
vertical approach retaining walls and deep abutments that are directly adjacent to
shoulders. In these conditions, roadway spray cannot be dissipated by air currents, and
where heavy deicing salt is used, it can result in excessive salt deposits on the overhead

153

Chapter 3. MATERIALS
bridge steel (FHWA 1989). These conditions are often found in urban grade separation
settings.
Exposure to roadway drainage, especially when combined with roadway deicing
salts, is also detrimental. This often occurs below open or leaking expansion joints,
and FHWA recommends that the ends of weathering steel girders below roadway
joints be painted within a length equal to 1.5 times the depth of the girder for protec-
tion (FHWA 1989). This girder end painting also helps prevent the staining of concrete
piers and/or abutments below (McEleney 2005).
Heavy industrial environments where steel may be subject to concentrated chemical
fumes (especially from sulfur dioxide) that may drift directly onto the structure can also
be detrimental. These environments should be evaluated before making decisions about
using uncoated weathering steel in them. Moderate industrial environments are not det-
rimental and can actually speed up the weathering process, however (McEleney 2005).
3.2.3.8
Structural Stainless Steel
A new steel was developed in recent years, ASTM A1010, which is a 12% chromium
structural steel with corrosion resistance that is superior to conventional weathering
steels’. Currently, A1010 is widely used in aggressive structural applications such as
coal rail cars and coal processing equipment, but because of A1010’s superior corrosion
resistance, it is also being considered for bridge applications in corrosive environments.
A1010 steel is available in typical plate sizes in 50- and 70-ksi strength levels, and data
on thicknesses up to 4 in. are now available (Fletcher et al. 2003; Wilson 2005).
Laboratory corrosion testing to the SAE 52334 standard has shown that A1010
performs in a superior fashion in a wet–dry saltwater environment when compared
with weathering or galvanized specimens. Figure 3.4 summarizes test results of A1010
compared with carbon steel, A588 weathering steel, and galvanized carbon steel in
a salt, wet–dry 8-week test. In addition, long-term exposure in seaside locations has
shown that A1010 performs significantly better than a variety of weathering steels.
Although this material is more expensive than conventional steels, its ability to per-
form in highly corrosive environments may give it an advantage when considering life-
cycle cost over a long-term service life (Fletcher et. al. 2003; Wilson 2005).
Precautions must be taken, however, when using stainless steel to avoid galvanic
corrosion when in contact with carbon steel, zinc, or aluminum.
3.3
Concrete and Steel Distresses and Solutions
The common distress factors that affect durability and the relevant solutions are given
in Section 3.3.1, and methods for protecting reinforcement against corrosion are
given in Section 3.3.2.
3.3.1
Common Concrete Distress Factors and Solutions
Concrete is subject to deterioration due to physical factors (moisture, temperature,
freeze and thaw); chemical factors (ASR, ACR, carbonation, chlorides, sulfates, and
acids) and functional factors (vibrations, impact, concrete consolidation, concrete

154

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 3.4.  Corrosion resistance of A1010 steel.
Source: Wilson 1999.

Source: Wilson (1999).


curing,
Figure 3.4.and concreteresistance
Corrosion placement).
ofStresses
A1010 induced
steel. in concrete caused by volumetric
changes resulting from these chemical, physical, and/or functional factors can exceed
[Composition:
the strength Please delete “(1)
of the material, salt, wet/dry,
leading to cracks,8-week test” at the
delaminations, andlower leftProper
spalling. of the figure, if
possible.]
steps should be taken to protect concrete from cracking.
3.3.1.1
Factors
The most common outcome of a distress is the formation of cracks in concrete. Volu-
metric changes that occur in concrete; the restraint in the bridge system; and the elastic
modulus, creep, and tensile strength characteristics lead to cracking (TRB 2006). The
lower the modulus of elasticity, the lower will be the amount of the induced elastic
tensile stress for a given magnitude of volumetric change, thus reducing the possibility
of cracking. In contrast, high elastic modulus is associated with brittle concretes that
are prone to cracking and crack propagation. In concretes with low elastic modulus,
such as lightweight concrete (LWC), stresses are less for a given strain compared with
conventional normal-weight concrete (NWC) with higher elastic modulus. Reduced
stresses minimize the occurrence and severity of cracks (number and width). Cracks
facilitate the intrusion of aggressive solutions that adversely affect the durability of
concrete. In columns and prestressed beams reduced modulus of elasticity would in-
dicate higher deformation and prestress losses. Concrete can be engineered to have
tight cracks (Sahmaran and Li 2001). In such concrete cracks formed are numerous
and very tight, less than 0.004 in. (0.1 mm), in width. It is difficult for harmful solu-
tions to penetrate such tight cracks. Research has shown that concrete with crack
widths less than 0.004 in. (0.1 mm) performs like sound concrete and does not allow
water to penetrate the cracks easily (Wang et al. 1997; Lawler et al. 2002). ACI 224R

155

Chapter 3. MATERIALS
(ACI 2001a) provides guidance on tolerable crack widths for reinforced concrete ex-
posed to different exposure conditions. For example, the reasonable crack width for
reinforced concrete exposed to seawater and seawater spray wetting and drying is
0.006 in. (0.15 mm). A note to the table in ACI 224R (ACI 2001a) indicates that a
portion of the cracks in such a structure are expected to exceed these values; with time,
a significant portion can exceed these values. The provisions contained in the LRFD
Bridge Design Specifications (AASHTO 2012) are based on a maximum crack width
of 0.004 in. (0.1 mm) for reinforced concrete exposed to a marine environment.
Volumetric changes can occur because of moisture loss, as in plastic shrinkage
or drying shrinkage, or in response to high temperatures and high temperature dif-
ferentials. As cement and water react, heat is generated; this phenomenon is known
as heat of hydration (Kosmatka and Wilson 2011). The rate of heat generation is
greatest at early concrete ages. For thin concrete elements heat generation is gener-
ally not a concern because the heat is dissipated quickly. However, in mass concrete
(Figure 3.5), heat is not easily dissipated, and a significant rise in temperature occurs.
As a general rule, any placement of structural concrete with a minimum dimension
equal to or greater than 3 ft should be considered mass concrete (Gajda 2007). Simi-
lar considerations should be given to other concrete placements that do not meet this
minimum dimension, but contain ASTM C150 Type III or HE (high-early-strength)
cement, accelerating admixtures, or cementitious materials in excess of 600 lb/yd3 of
concrete (Gajda 2007). Mass concrete specifications, generally, limit the temperature
difference between the center and surface of the concrete to 35°F. This is a conserva-
tive limit that has been effective (Gajda 2007). A performance-based temperature dif-
ference limit tailored to the particular job can indicate temperature differences higher
than 35°F without adverse effects. Finite element modeling and detailed calculations
are often used to develop temperature difference limits.
High temperatures exceeding 160°F during the first few hours following place-
ment can lead to delayed ettringite formation at later ages. Ettringite forms when gyp-
sum and other sulfate compounds react with calcium aluminate in cement in the first

Figure 3.5.  Mass concrete: footings and


columns.
Source: Courtesy Virginia DOT.

156

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


few hours after mixing with water (Kosmatka and Wilson 2011). However, at high
temperatures the normal formation of ettringite during the first few hours is impeded,
and delayed ettringite formation occurs in hardened concrete, which can crack the
concrete. Heat-related cracking can also occur in bridge decks, which are typically
not considered to be mass concrete (Figure 3.6). In this case the bridge beams restrain
temperature-related movement of the deck. The heat of hydration causes the deck to
expand and then contract during cooling, and the beams restrain the movement. To
prevent the strains and resulting tensile stresses, a maximum temperature differential
of 22°F between the beams and the deck is recommended for at least 24 hours follow-
ing concrete placement (Babaei and Fouladgar 1997). As with mass concrete, higher
temperatures may be justified, but they require detailed analysis to justify. To minimize
the potential for temperature-related cracking, the amount of portland cement should
be minimized, concrete delivery temperature reduced, and pozzolans and slag included
in the mixture. With proper temperature management thermal cracks can be mini-
mized and controlled.
Volumetric changes also occur as a result of freezing and thawing cycles and chem-
ical (corrosion, ASR, and ACR) reactions.
3.3.1.2
Prevention of Distress
3.3.1.2.1
Surface Treatments
Surface treatments, membranes, sealers, and overlays are widely used to reduce so-
lution infiltration. They protect the deck concrete from deterioration induced by
freeze–thaw cycles and protect the reinforcement from corrosion (Kepler et al. 2000).
Currently, there are three types of waterproofing membranes used in North America:
preformed sheets, liquid membranes, and built-up systems (Russell 2012). Preformed
systems are labor intensive and require good workmanship. Liquid systems are easier
to place. Built-up systems are generally labor intensive and expensive (Manning 1995).

Figure 3.6.  Thermal cracks in bridge deck.


Source: Courtesy Virginia DOT.

157

Chapter 3. MATERIALS
Sealers can be a pore blocker, forming a microscopically thin (up to 2 mm) protec-
tive layer on the concrete surface, or a penetrating liquid that acts as a hydrophobic
agent (Zemajtis and Weyers 1996). When used as a pore blocker on bridge decks, the
effect on skid resistance should be considered (Sherman et al. 1993). Sealers should be
breathable, enabling escape of moisture to prevent high vapor pressures that can cause
blistering and peeling (Sherman et al. 1993).
3.3.1.2.2
Control of Volumetric Changes Caused by Moisture and Temperature
To minimize volumetric changes caused by shrinkage and temperature, low cementi-
tious material, low water content, and low paste content are desirable. This can be
achieved by using well-graded aggregates, reducing mix temperature, and selecting the
appropriate admixtures.
3.3.1.2.3
Control of Wear and Abrasion
Traffic on bridge decks causes abrasion. Moving water also carries objects that can
cause abrasion. Studded tires are very damaging to the surface of the concrete and are
not permitted in some states, and many other states have seasonal restrictions. Chains
also have a damaging effect, although not as much as studs. Chains are encouraged in
winter in many areas, and some states require chains for commercial trucks. Abrasion
resistance of concrete is a function of the water-cementitious materials ratio (w/cm)
at the surface and the aggregate quantity and quality. The Los Angeles abrasion test
indicates the abrasion resistance of aggregates; in general, siliceous aggregates provide
satisfactory abrasion resistance. Proper finishing and curing are also factors that influ-
ence abrasion resistance.
3.3.1.2.4
Control of Freeze–Thaw Damage
Concrete that can become critically saturated and exposed to cycles of freezing and
thawing must be properly air entrained, have sound aggregates, and have the maturity
to develop a compressive strength of about 4,000 psi to avoid cracking and scaling
(Mather 1990). Air voids must be small in size, closely spaced, and uniformly distrib-
uted to ensure adequate resistance to freezing and thawing and satisfactory strength.
A spacing factor less than 0.008 in. is needed for adequate protection during freezing
and thawing. However, with the use of high-range water-reducing admixtures and
low-permeability concretes, higher values may be acceptable and should be verified by
relevant tests, such as ASTM C666 Procedure A.
3.3.1.2.5
Control of Chemical Reactions
Disruptive chemical reactions can occur in concrete that adversely affect durability.
The chemical composition of the cementitious material affects the rate of hydration
and pozzolanic reaction and heat generation and contributes to the formation of dis-
ruptive products. The fineness of the cementitious material affects the rate of reac-
tions and water demand. At high water w/cm, permeability is increased and durability
is reduced. The presence of certain chemicals in sufficient amounts in the cements

158

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


contributes to the expansion. Chemical reactions involving alkalis with aggregates, sil-
icates, and carbonates; carbonation; chlorides; sulfates; acids; and salts are explained
in the following subsections.
3.3.1.2.5a
Alkali-Aggregate Reaction
Alkali-aggregate reactions are the reactions between the hydroxide ions in the pore
fluid of concrete, usually associated with alkalis from the cement or from outside
sources such as deicing salts, and the reactive constituents of the aggregates (ACI
221.1R, 1998). This reaction results in expansion and cracking.
3.3.1.2.5b
Alkali-Silica Reaction
A chemical reaction between aggregates containing reactive silica and the alkalis in
concrete can produce an alkali-silica gel that swells when water is absorbed. The high
pressures generated within the concrete lead to cracking (Figure 3.7). To prevent ASR,
nonreactive aggregates, or pozzolanic materials or slag, lithium nitrate, low-permea-
bility concrete, and cements with low alkali contents are used.
3.3.1.2.5c
Alkali-Carbonate Reaction
Some argillaceous, dolomitic aggregates can react with alkalis, causing the aggregates
to expand. Figure 3.8 shows a joint closing due to ACR expansion. A common preven-
tion for ACR is to avoid using reactive aggregate or to dilute the aggregate by blending
with nonreactive aggregate.
3.3.1.2.5d
Carbonation
Carbon dioxide (CO2) produced by plants penetrates concrete and reacts with the
hydroxides, such as calcium hydroxide (lime), to form carbonates. In this carbonation
process the pH is reduced to less than nine and influences the protective layer over the
steel (Neville 1995).

Figure 3.7.  Alkali-silica reaction.


Source: Courtesy Virginia DOT.

159

Chapter 3. MATERIALS
3.3.1.2.5e
Chlorides
Chlorides that penetrate the concrete and reach the steel surface destroy the protec-
tive oxide layer, making reinforcement prone to corrosion. Without the protective
layer, steel will corrode rapidly in the presence of water and oxygen. The corrosion
of steel is accompanied by expansive pressures, which lead to cracking (Figure 3.9).

Figure 3.8.  Closing of joints as a result of ACR.


Source: Courtesy Virginia DOT.

Figure 3.9. Corrosion.
Source: Courtesy Virginia
DOT.

160

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Corrosion-­inhibiting admixtures can be used to stabilize the passive oxide layer on
the reinforcement, or viscosity-modifying admixtures can be added to improve the
stability of the mix.
3.3.1.2.5f
Sulfates
Sulfates react with the calcium hydroxide [Ca(OH)2] and the calcium aluminate hy-
drates, causing expansive reactions that can affect the cement paste (Neville 1995).
Figure 3.10 shows the loss of material after sulfate attack. To increase the resistance of
concrete to sulfate attack, a low tricalcium aluminate (C3A) content and low quanti-
ties of Ca(OH)2 in the cement paste and low-permeability concrete are needed. These
results can be obtained by using sulfate-resistant cements and pozzolans.
3.3.1.2.5g
Acids
Pollutants cause acid rain that can cause deterioration (Neville 1995). In damp condi-
tions, SO2, CO2, SO3, and other acid forms that are present in the atmosphere may
attack concrete by dissolving in water and removing parts of the cement paste. These
acids will leave a soft and mushy mass (Eglinton 1975). To minimize acid attack, low-
permeability concrete and barrier coatings may be used. A barrier material separates
the concrete surface from the environment.
3.3.1.2.5h
Salts
Chloride-bearing deicing chemicals initiate and accelerate corrosion. Magnesium salts
are very damaging to concrete, causing crumbling (Lee et al. 2000). Calcium mag-
nesium acetate has been suggested to prevent corrosion, but it was found to be very
damaging to concrete. Deicing chemicals can also aggravate freeze–thaw deteriora-
tion. Osmotic pressure occurs because moisture tends to move toward zones with
higher salt concentrations. In addition, the salts increase the rate of cooling, causing an
increase in the potential for freeze–thaw deterioration at the concrete surface. Proper
air entrainment and maturity are needed to provide the necessary protection.

Figure 3.10.  Sulfate attack.


Source: Courtesy Virginia DOT.

161

Chapter 3. MATERIALS
3.3.1.2.6
Functional Considerations
Functional considerations include vibrations, impact, concrete consolidation, concrete
curing, and concrete placement.
3.3.1.2.6a
Vibrations
Vibration of fresh concrete using vibrators may cause loss of air in mixtures with
a high sand content and could result in freeze–thaw damage. In these mixtures, the
frequency and duration of vibration should be reduced to prevent segregation (loss of
stability) and loss of air. A recommended practice is to prepare mixtures with a large
amount of coarse aggregate content. However, in some regions where D-cracking (Fig-
ure 3.11) is an issue, the size and amount of coarse aggregate is reduced, which leads
to mixtures with an excessive sand content. In D-cracking, the aggregate has a pore
structure that hinders the expulsion of water from the aggregate pores during freezing,
which results in the cracking of the aggregate and concrete.
Concrete can exhibit fatigue behavior when subjected to cyclic loading of a given
level (below its short-term static strength) and will eventually fail.
3.3.1.2.6b
Impact
Concrete can be subjected to extreme loads from the impact of falling rocks, snow
avalanches, landslides, or vehicle crashes. The impact resistance is related to the com-
pressive strength and aggregate type (Kosmatka and Wilson 2011). Fiber-reinforced
concretes are used to minimize the effect of impact in certain applications.
3.3.1.2.6c
Concrete Consolidation
Proper consolidation ensures that undesirable entrapped air voids are eliminated. Voids
in concrete reduce strength, and when interconnected or in large amounts, they can
increase permeability. Currently, workable concretes are made using water-­reducing
admixtures. It is also possible to make stable concretes that have high flow rates and
are self-consolidating.

Figure 3.11. D-cracking.
Source: Courtesy Portland Cement Association,
Skokie, Illinois.

162

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


3.3.1.2.6d
Concrete Curing
Curing ensures that reactions occur and volumetric changes are minimized. Curing
should continue until a certain level of the desired properties is achieved. Concrete
should stay wet during the curing period, and the temperature should be managed to
eliminate large differentials. ACI 308R provides recommendations for adequate cur-
ing. ACI 305R and ACI 306R provide information on curing of concrete in hot and
cold weather, respectively.
3.3.1.2.6e
Concrete Placement
Concrete should be placed without causing segregation and loss of moisture. Forms
should be adequately set, clean, tight, and adequately braced. Forms should be oiled or
treated with a form-release agent. Reinforcing steel should be clean and free of rust or
mill scale. In cold weather, concrete should not be placed in contact with metal forms
and embedments, such as steel structural members or reinforcement, which can freeze
the concrete (ACI 306R, 2010). If the frozen concrete does not thaw before the bulk
of the concrete sets, bond may be significantly reduced, and concrete quality adjacent
to cold metal would be poor. Ideally, the adjacent metal should be heated to the tem-
perature of the concrete immediately before concrete placement (ACI 306R, 2010).
Concrete should be deposited continuously and as near as possible to its final posi-
tion without objectionable segregation (Kosmatka and Wilson 2011). Concrete should
be deposited in areas free of standing water. However, in some applications, such as
drilled shafts, standing water may be present. In such applications, concrete should be
placed in a manner that it displaces the water ahead of the concrete without mixing
the water with the concrete. Pumps and tremies with ends buried in the fresh concrete
can be used to maintain a seal below the rising surface. Pumps are widely used to place
concrete in bridge structures. Care should be exercised in delivering with the pump
because free fall within the pump hose can result in loss of slump and air.
3.3.2
Protection of Reinforcing Steel Against Corrosion
Methods for protecting reinforcing steel elements from corrosion include the use of
corrosion-resistant reinforcing steel and the use of nonintrinsic protection of the re-
inforcement (such as admixtures, cathodic protection systems, and electrochemical
chloride extraction [ECE] techniques) as described in the following section. Chapter 5
covers the cathodic protection system and ECE in additional detail.
3.3.2.1
Corrosion-Resistant Reinforcement
Corrosion-resistant reinforcement allows chlorides to penetrate around the reinforce-
ment without causing significant damage to the reinforcing steel. These systems in-
clude epoxy-coated reinforcement (ECR), galvanized reinforcement, titanium rein-
forcement, stainless steel reinforcement, stainless steel–clad reinforcement, nickel-clad
reinforcement, and copper-clad reinforcement and are described in Section 3.2.2.

163

Chapter 3. MATERIALS
3.3.2.2
Nonintrinsic Corrosion Protection of Reinforcement
3.3.2.2.1
Admixtures for Corrosion Protection
Chemical admixtures that are added to concrete during batching to protect against
corrosion of embedded steel reinforcement due to chlorides are available. There are
two main types: corrosion inhibitors and physical barrier admixtures. Some corrosion
inhibitors also act as physical barrier admixtures (ACI 222.3R, 2003a).
Corrosion inhibitors, although known for many years, are only now beginning
to be actively marketed as a preventative treatment in a concrete repair program
(­Macdonald 2003). Calcium nitrite admixtures are the most researched inorganic
inhibitor and the most widely used (Berke and Rosenberg 1989). Inhibitors do not
create a physical barrier to chloride ion ingress. Rather, they modify the steel surface,
either electrochemically (anodic, cathodic, or mixed inhibitor) or chemically (chemical
barrier) to inhibit chloride-induced corrosion above the accepted chloride corrosion
threshold level. They are added to the concrete at the time of batching (ACI 201.2R,
2008a). Calcium nitrite is also an accelerating admixture. If the accelerating effect is
undesirable, a retarding admixture can be added.
Admixtures with organic compounds protect steel from chloride-induced corro-
sion. They include alkanolamines and an aqueous mixture of amines and fatty acid
esters (Nmai et al. 1992; Nmai and Krauss 1994). They are claimed to both reduce
ingress of chlorides (physical barrier) and enhance the passivating layer on the steel
surface (corrosion inhibitor). Other similar amine products are claimed to migrate
through concrete in the vapor phase to provide protection to embedded steel. Cor-
rosion inhibitors are attractive from a conservation point of view as they are almost
invisible on application, although their long-term visual effect is unknown (­Macdonald
2003).
Physical barrier admixtures reduce the rate of ingress of corrosive agents (chlo-
rides, oxygen, and water) into the concrete. These admixtures belong to one of two
groups. One group comprises waterproofing and damp-proofing compounds. The
second group consists of agents that create an organic film around the reinforcing
steel, supplementing the passivating layer. They typically contain bitumen, silicates,
and water-based organic admixtures consisting of fatty acids, such as oleic acid; ­stearic
acid; salts of calcium oleate; and esters, such as butyloleate (ACI 222R, 2001b). A
liquid admixture containing a silicate copolymer in the form of a complex, inorganic,
alkaline earth may also be effective in reducing the permeability of concrete and pro-
viding protection against corrosion of reinforcing steel (Miller 1995).
3.3.2.2.2
Cathodic Protection
Cathodic protection can effectively stop corrosion in contaminated reinforced con-
crete structures and can reduce the concentration of chloride ions at the steel surface
of protected reinforcement (Kepler et al. 2000). Cathodic protection is applicable to
areas that are contaminated chlorides but are still sound (Polder 1994).

164

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


The wide use of cathodic protection has been hampered by the high cost and the
maintenance of the power source or the protective material. Impressed current can be
used on bridge decks and substructures; galvanic anodes are limited to substructures.
3.3.2.2.3
Electrochemical Chloride Extraction
ECE is applied to concrete structures containing reinforcement in order to extract chlo-
rides from the concrete (Clemeña and Jackson 2000) and can be used for decks and sub-
structures. The steel acts as a cathode and is connected to the negative pole of the power
source. The anode, which is either steel or a titanium mesh, is temporarily placed on the
concrete cover and is connected to the positive pole of the power source. The electrolyte
is placed on the concrete cover and allows the current flow. Due to the electric field, the
negative ions, like chlorides, migrate from the rebar to the concrete cover.
During ECE, hydroxyl ions are generated at the reinforcement, increasing the alka-
linity and making the concrete susceptible to ASR. ASR can be mitigated by lithium
salts (Velivasakis et al. 1997).
3.3.2.2.4
Other Protective Methods
Other protective methods include the use of drainage design and stay-in-place metal
forms. Posttensioning of members would also eliminate the formation of cracks.
3.4
Fault Tree Analysis of Factors Influencing Service Life
3.4.1
Service Life of Concrete
Concrete deterioration is caused by deficiencies in one or more factors, listed as ma-
terials, design, and workmanship; the effect of external factors; and the occurrence of
cracks. These factors are shown in the main fault tree (Figure 3.12) and are explained
in additional detail in the fault trees for each factor shown in Figures 3.13 to 3.18.

 
Figure3.12.
Figure 3.12.  Mainfault
Main faulttree
treefor
for reduced
reduced service
servicelifelife
of of
concrete.
concrete.
165

Chapter 3. MATERIALS
3.4.1.1
Materials
Material-related factors are shown in Figure 3.13. They include the ingredients of the
concrete: the cementitious material, aggregates, water, admixtures, and fibers. The
w/cm is also included in this module. Information on materials is given in Section 3.2
­under description of material types. The cementitious materials have different chemi-
cal and physical properties that affect the hydration reaction, harmful chemical re-
actions, and volumetric changes that relate to the durability of concrete. The type,
quality, grading, and texture of the aggregates affect the water content and durability.
Aggregates may cause D-cracking, ASR, and ACR; precautions are needed to inhibit
such occurrences. Water, whether it is potable or recycled, may have some excessive
impurities that may cause durability problems as a result of expansive chemical re-
actions and increased w/cm. Small solid particles in water have large surface area;
they increase the water demand and the w/cm if the cement content is kept the same.
Admixtures help in achieving workable, low w/cm, and low-permeability concretes,
leading to improved durability. Fibers are used to control cracking.

Figure 3.13.  Materials fault tree.


Figure 3.13. Materials fault tree.

166

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


3.4.1.2
Design
The design of the structure through the selection of the geometry, detailing, and flex-
ibility affects the performance of the concrete. The design fault tree is shown in Fig-
ure 3.14. In the geometry, the cover depth over the reinforcement has a large influence
on salt intrusion to the level of the steel. Thicker decks provide more rigidity and less
cracking potential. Long-span length causes more flexibility, increasing the possibility
of cracking.
Design details that minimize saturation with water would lead to improved dura-
bility. Eliminating joints is desirable because joints cause leakage problems.
3.4.1.3
Workmanship
Workmanship is important in achieving the desired durability; the materials have to
be fabricated properly under adequate inspection to achieve the desired performance.
Frequent maintenance to correct deficiencies is also needed to eliminate premature
failure of the elements. Fabrication of components includes mixing, consolidation,
finishing, and curing, as shown in the workmanship fault tree in Figure 3.15.

!
Figure 3.14.  Design fault tree.
Figure 3.14. Design fault tree.

167

Chapter 3. MATERIALS
 

Figure 3.15.
Figure 3.15. Workmanship fault
Workmanship fault tree.
tree.

The concrete mixer should be in good working order, the capacity given by the
manufacturer should not be exceeded, and enough mixing time at the specified mixing
rate should be maintained. Following the proper mixing guidelines will ensure a uni-
form, consistent concrete mixture. Consolidation, which is achieved by proper internal
or external vibration, eliminates the large air voids that adversely affect the strength
and durability of concrete.
Bridge decks require minimal finishing operations. The vibratory screed with
proper speed, vibration, and setup of the auger and the roller can provide adequate
vibration if the workability of the concrete is adequate. The sides and ends that the
168

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


vibratory screed cannot reach are finished by hand. Extra hand finishing is detrimental
and may delay the curing operation and cause loss of entrained air voids near the top
surface. Curing is essential for the continuation of the hydration reactions and the
control of cracking due to volumetric changes. The best curing process is a water cure
that enables moisture retention and temperature management. Curing compounds are
also used to maintain satisfactory moisture and temperature. The Virginia DOT uses
curing compounds after a 7-day wet curing of decks.
3.4.1.4
External Factors
External factors that affect the service life of concrete are loads and the environment.
Loads are illustrated in Figure 3.16, and environment factors are shown in Figure 3.17.
3.4.1.4.1
Loads
Loads can be traffic induced, age dependent, or system dependent. Traffic-induced
loads are a result of vehicle loads, the frequency of traffic imparting fatigue stresses,
and overloads due to unexpected high loads. The adverse effects of loads are wide
and frequent cracks that facilitate the intrusion of harmful solutions. The duration of
loading has to be considered, as loads can be instantaneous or time dependent. Time-
dependent loads cause additional distress and result in additional cracks. System-­
dependent loads are a result of moisture and temperature variation and the available
restraint. When restrained, deformations lead to stresses that can exceed the strength
of the material, leading to cracking. When cement reacts with water, an exothermic
reaction takes place: temperature rises, heat is given off, and concrete expands. As
the reaction slows, cooling takes place and thermal contraction occurs, subjecting the
restrained concrete to tensile stresses. When the stresses exceed the strength of the ma-
terial, cracks occur. This thermal effect is more pronounced in mass concrete because
the dissipation of heat is difficult in a large mass. High heat can also result in delayed
ettringite reaction, which can cause cracks in hardened concrete. In hot weather, ther-
mal effects can be detrimental; however, in a cold environment, the heat of hydration
can provide the favorable temperature needed for the hydration reactions.
3.4.1.4.2
Environment
The environment can provoke damage in structures caused by physical and/or chemi-
cal factors. The physical factors are the result of freezing and thawing of critically satu-
rated concrete, scaling of surfaces due to salt concentrations or excessive wear, seismic
activity, settlement, volumetric changes due to moisture and temperature variation, or
stresses due to wind velocity. Chemical factors involve corrosion; carbonation, which
reduces pH and makes steel vulnerable to corrosion; sulfate attack; or expansion due
to alkali-aggregate reactions or ASR.

169

Chapter 3. MATERIALS
Loads

Traffic-Induced Age System-Dependent

Figure 3.16.  Load fault tree.

3.4.1.5
Cracking
Cracking of concrete can cause serious and costly damage to concrete structures. The
cracking fault tree is shown in Figure 3.18. The factors affecting cracking can be due
to design, materials, and/or construction. The design factors of interest include the
restraint, span type, deck thickness, girder type, and the steel alignment and location.
When concrete is restrained, deformations result in stresses. Flexible structures and
low rigidity lead to additional cracks. For example, decks on steel beams exhibit addi-
tional cracking compared with decks on rigid concrete beams. High modulus and low
creep, which can be beneficial in reducing prestress losses and deflections in beams,
can lead to additional cracking in decks. High modulus leads to brittle structures, and
low creep does not enable relaxation that reduces stresses. It can be expected that low
w/cm, high paste contents, and high heat of hydration can cause additional cracking.
Concretes with low w/cm are brittle and more sensitive to curing. Concretes with high
170

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Environment

Physical Chemical

Figure 3.17.  Environment fault tree.

water, cement, and paste content exhibit additional shrinkage and additional tempera-
ture rise. During construction and afterward, the weather conditions, curing, time of
setting, consolidation, and curing sequence and length affect the cracking pattern and
severity.
3.4.2
Service Life of Reinforcement
Reduced service life of reinforcement can be attributed to three causes: load-induced,
man-made, or natural hazards; causes resulting from production defects in construc-
tion processes and/or design details; or operational procedures. These deficiencies are
illustrated in the main fault tree shown in Figure 3.19.

171

Chapter 3. MATERIALS
 
Figure 3.18.
Figure 3.18.  Crackingfault
Cracking fault tree.
tree.

Reduced Service Life


of Reinforcement

Natural or Production/
Load-Induced
Man-Made Hazards Operation Defects

Figure 3.19.  Main fault tree for reduced service life of reinforcement.

172

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


3.4.2.1
Load-Induced Factors
Load-induced bridge-deck deterioration can be attributed to fatigue, strength and brit-
tleness, or thermal incompatibility. These load-induced factors are listed in the fault
tree in Figure 3.20.
3.4.2.1.1
Fatigue
Fatigue is caused by the repetition of applied loads that result in a degradation of
the strength resistance of the reinforcement. Information on reinforcement material is
summarized in Section 3.2.2. Corrosion-resistant reinforcement has fatigue properties
similar to that of carbon steel reinforcement when tested in the atmosphere. How-
ever, in a corrosive environment, corrosion-resistant reinforcement performs better
than the carbon steel because carbon steel is expected to corrode, lose material area,
and develop corrosion pits. The fatigue limit is related to the tensile strength of the
steel; hence corrosion-resistant reinforcement with increased strength has an increased
­fatigue limit.
3.4.2.1.2
Strength and Brittleness
Strength and brittleness affect cracking and failure. In general, corrosion-resistant rein-
forcement has higher tensile strength and ductility than the carbon steels. Cold-formed
austenitic reinforcement has a combination of high strength and good ductility; a yield
stress level of 70 ksi or higher and elongation at maximum force higher than 15% is
achieved (Bourgin et al. 2006). Duplex grades exhibit strengths exceeding 70 ksi in
hot-rolled and 90 ksi in cold-rolled bars. At the higher strengths, ductility is at least as
good as the carbon steels.

Load-Induced
Load-Induced

Strength and Thermal


Fatigue
brittleness incompatibility

Figure 3.20.  Load-induced deficiency fault tree.

173

Chapter 3. MATERIALS
3.4.2.1.3
Thermal Compatibility
Thermal compatibility affects cracking potential. Temperature changes in a material
result in deformations that can cause significant stress when restrained by the sur-
rounding material. However, in reinforced concrete the carbon steel and concrete have
similar coefficients of thermal expansion; this similarity means that stresses caused by
temperature changes in the structure are negligible. Carbon steels have a coefficient
of thermal expansion of about 5.5 × 106/ºF; corrosion-resistant reinforcements with
austenitic steels have a coefficient of thermal expansion of approximately 8.9 × 106/ºF;
and austenitic-ferritic duplex steels have a coefficient of thermal expansion of about
7.2 × 106/ºF (Markeset et al. 2006). No problems stemming from these differences
have been reported.
3.4.2.2
Natural or Man-Made Hazards
Natural or man-made hazards include effects from areas with adverse thermal cli-
mates, coastal climates, and chemical climates, and fire. These natural and man-made
hazards are listed in the fault tree provided in Figure 3.21.
3.4.2.2.1
Thermal Climate
The thermal climate affects corrosion activity. In cold climates, chloride-bearing
­deicing salts are commonly used to prevent ice buildup on roads. Chlorides destroy
the protective iron oxide layer over the carbon steels, exposing the reinforcement to
corrosion. Corrosion of prestressing steel is generally a greater concern than corrosion
of nonprestressed reinforcement because of the possibility that corrosion may cause a
local reduction in cross section and failure of the prestressing steel (ACI 222R, 2001b).
The typical higher stresses in the prestressing steel also render it more vulnerable to
stress-corrosion cracking and to corrosion fatigue. Because of the potentially greater
vulnerability and the consequences of corrosion of prestressing steel, chloride limits for
prestressed concrete are lower than those for reinforced concrete (ACI 222R, 2001b).
Corrosion rate is dependent on the temperature, humidity, and chloride content.
An increase in temperature in dry environments results in reduced corrosion activity
and an increase in expected life (Lopez et al. 1993). However, in environments with
high humidity, increasing temperatures result in increased corrosion activity, leading
to reduced expected life.
Low-temperature, cryogenic applications can cause brittle failure. Carbon steel
reinforcement exhibits brittle behavior below 0°F when exposed to sudden loading
and seismic actions. Austenitic stainless steels do not present such a transition and can
be used in cryogenic applications; their toughness remains very high at temperatures
as low as –320°F. Duplex stainless steel may not be used below –60°F (Markeset et al.
2006).

174

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 3.21.  Natural or man-made hazard fault tree.
Figure 3.21. Natural or man-made hazard fault tree.

3.4.2.2.2
Coastal Climate
Coastal climates introduce salt spray and high humidity, and salt and moisture accel-
erate the corrosion rate. Salt spray in coastal climates provides high chloride buildup
that can destroy the protective iron oxide coating over the steel reinforcement.
Humidity affects corrosion activity. In the absence of chloride ions, little corrosion
activity takes place when the relative humidity is under 60% (Jung et al. 2003). The
corrosion activity increases as the relative humidity is increased up to a fully saturated
state (>95% relative humidity) and then begins to decrease again. When concrete is
fully saturated, the corrosion rate is reduced because the oxygen level in the concrete
pores is too low (Qian et al. 2002). When chlorides are present at relative humidity
below 60%, corrosion activity may still develop.

175

Chapter 3. MATERIALS
3.4.2.2.3
Chemical Climate
Chemical climates influence the performance of reinforcement. The main effect can be
attributed to corrosion-inducing chemicals that occur naturally and can be man-made.
Chlorides are the main chemical substance that adversely affects the corrosion process.
3.4.2.2.4
Fire
Fire generates heat that affects mechanical properties. Cold-worked steel subjected to
temperatures of less than 850°F typically recovers all of its yield strength after cooling
(Suprenant 1996). Hot-rolled steel can be exposed to temperatures as high as 1,100°F
and recover its yield strength. Higher temperatures may cause rapid strength loss in
reinforcing steel and lead to excessive deflections in reinforced members. The effect
of fire is more critical on prestressing steel; at temperatures of 750°F, the strength of
prestressing steel can be reduced by more than 50% (Suprenant 1996).
Austenitic stainless steels maintain their strengths at considerably higher tempera-
tures than carbon steel and thus are more resistant and robust under fire loading than
carbon steel (Markeset et al. 2006).
Heating also adversely affects the bond between concrete and reinforcement
(Suprenant 1996). At 570°F the bond strength is no greater than 85% and at 930°F no
greater than 50% of the strength at ambient temperatures.
3.4.2.3
Production and Operational Defects
Production and operational defects are shown in the fault tree in Figure 3.22 They
include design and detailing, construction, inspection, and maintenance issues.
3.4.2.3.1
Design and Detailing
Design and detailing factors are listed in the fault tree shown in Figure 3.23. They
include the design philosophy, mix design, and drainage.
The design philosophy of providing proper concrete cover, eliminating joints, min-
imizing cracking through geometry (e.g., large skews exhibit additional cracking), and
the selection of corrosion-resistant reinforcement affect the corrosion potential. The
redundancy and ductility design aspects in structures should be improved to confine
the damage to a small area in the event a major supporting element is damaged or an
abnormal loading event has occurred (ACI 318-11). The following considerations and
relationships should be adopted in the concrete mixture design:

• Low-permeability concretes hinder the penetration of aggressive solutions to the


level of reinforcement;
• High alkalinity of the concrete passivates the steel through its protective cover;
• Cracking resistance of concretes can be improved by optimizing the strength (high-
strength concrete exhibits brittle behavior), minimizing paste, and adding latex
modifiers;

176

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Production/
Operation Defects

Design/Detailing Construction Inspection

Figure 3.22.  Production and operation defects fault tree.

• The use of corrosion-inhibiting admixtures increases the passivation state of the


reinforcement, extends the time to corrosion, and reduces the corrosion rate of
embedded metal;
• Increased creep and reduced elastic modulus and reduced shrinkage are helpful in
reducing the cracking; and
• Cracks facilitate the intrusion of aggressive solutions into concrete; chlorides initi-
ate and accelerate the corrosion process.

3.4.2.3.2
Construction
Construction-related parameters affect the performance of structures. It is critical that
the correct amount of reinforcement is placed in the right location within the specified
tolerances. Reinforcement during concrete placement should be free from mud, oil,
177

Chapter 3. MATERIALS
Design/Detailing

Design Philosophy Mix Design Drainage

Figure 3.23.  Design and detailing defects fault tree.

or other nonmetallic coatings that decrease bond (ACI 318-11). A normal amount of
rust or mill scale that is not loose on the reinforcement is not detrimental to the bond
between the concrete and the bars.
In the field, bending to proper bend diameters is needed to ensure that there is no
breakage and no crushing of the concrete inside the bend (ACI 318-11).
To minimize the corrosion potential, dissimilar metals should not be in contact.
Further, reinforcement should be protected from the weather to minimize contamina-
tion and corrosion. If there is a coating over the reinforcement, special care is needed
to avoid damage to the coating during handling and placement.

178

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


3.4.2.3.3
Inspection
Visual inspection can indicate the condition of the reinforcement and determine if
there are any gross mistakes in the reinforcement selection and placement. The avail-
ability of a large number of reinforcement types makes it difficult to identify the rein-
forcement visually; nondestructive evaluation is beneficial in this respect.
Rebar corrosion in existing structures can be assessed by different methods, such
as open circuit potential and surface potential, concrete resistivity, linear polarization
resistance, Tafel extrapolation, galvanostatic pulse transient method, electrochemical
impedance spectroscopy, harmonic analysis, noise analysis, embeddable corrosion-
monitoring sensor, cover thickness, ultrasonic pulse velocity technique, X-ray, gamma
radiography, infrared thermograph, electrochemical method, and visual inspection
(Song and Saraswathy 2007).
3.4.3
Service Life of Structural Steel
The primary factors affecting service life of structural steel are fatigue and fracture
and corrosion. These factors are covered in Chapter 7, Fatigue and Fracture of Steel
Structures; and Chapter 6, Corrosion Prevention of Steel Bridges.
3.5
Individual Strategies to Mitigate Factors Affecting Service Life
3.5.1
Concrete
The durability of concrete depends largely on its permeability. For longevity, concrete
must be designed, proportioned, and constructed properly. In addition to the concrete
material issues, the structural design must also be performed properly to avoid high
stresses and load-related cracking. Section 3.3.1 provides additional information on
distresses due to physical (volumetric changes, freezing and thawing), chemical (­alkali–
aggregate reaction, ASR, ACR, carbonation, chlorides, sulfates, acids, and salts), and
functional (impact, concrete consolidation, curing, and placement) factors.
Table 3.9 is a technology table that informs the designer of the most common
types of service life issues and distresses related to concrete materials. For each service
life issue, solutions are identified, along with their advantages and disadvantages. The
technology table for concrete durability also includes concrete-related issues involved
in resisting the corrosion of reinforcement, such as low permeability, w/cm, aggregates,
chemical admixtures, shrinkage, modulus of elasticity, cover, overlays, and corrosion
inhibitors.
3.5.2
Steel Reinforcement
Corrosion of reinforcement is a major problem requiring costly repairs. Methods for
protecting reinforcing steel elements from corrosion include as the use of corrosion-­
resistant reinforcing steel, admixtures, cathodic protection systems, and ECE
techniques.

179

Chapter 3. MATERIALS
Table 3.9. Technology Table for Concrete Durability
Service Life
Issue Solution Advantage Disadvantage
Freeze and Good air void system High resistance to freezing and Reduction in strength due to
thaw thawing extra air
Sound aggregates Durable aggregates Availability
Strength of 4,000 psi and up Used to overcome stresses Increased strength makes
concrete more brittle
Drainage design Minimizes saturation Ingress of water
Low w/cm Reduces infiltration of water Can produce high-strength
concrete that is brittle
Abrasion and Hard aggregates Attain high concrete strengths Hard to obtain in some areas
wear and increased resistance to
abrasion and wear
High-strength concrete Reduces wearing Concrete more brittle
Add cover Provides new surface Extra weight
Chemical Nonreactive siliceous Reduces ASR Hard to obtain in many areas
reactions aggregates
(ASR) SCMs Reduces permeability, reduces Quality fly ash or slag missing in
ASR, limits alkalis from outside many areas
Low w/cm Reduces infiltration of solutions, Can produce high-strength
limits alkalis from outside concrete that is brittle
Chemical admixtures Improved properties Cost, incompatibility, side effects
Lithium-based admixtures Inhibits ASR Cost
Limestone sweetening Limits expansion Reduced skid resistance
(blending with limestone)
Chemical Nonreactive carbonate Reduces ACR Hard to obtain in some areas
reactions aggregates
(ACR) Reduce infiltration of Can produce high-strength Cracking
solutions, limit alkalis from concrete that is brittle
outside
Blend aggregate Limits expansion Hard to obtain in some areas
Limit aggregate size to Limits expansion Rich mixes with high paste
smallest practical content
Sulfate attack Low C3A contents Reduces sulfate attack na
SCMs Reduces permeability, reduces Quality fly ash or slag missing in
sulfate attack, limits sulfates from many areas
outside
Low w/cm Reduces infiltration of solutions, Can produce high-strength
limits sulfates from outside concrete that is brittle

(continued)

180

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 3.9. Technology Table for Concrete Durability (continued)
Service Life
Issue Solution Advantage Disadvantage
Corrosion of Low permeability Reduces infiltration of aggressive Can produce high-strength
reinforcement solutions concrete that is brittle
Membranes and coatings Reduces infiltration of aggressive Difficult to apply in the field,
solutions wear of traffic
Sealers for pore lining and Reduces infiltration of aggressive Difficult to apply in the field,
blocking solutions concrete may be difficult to
penetrate
Low w/cm High strength, low permeability Excessive cracking, shrinkage
Low shrinkage Minimizes cracking Low water content may
adversely affect workability
Low modulus of elasticity High deformation, minimizes Reduces stiffness
deck cracking
SCMs Reduces permeability Quality fly ash or slag missing in
many areas
Large maximum aggregate Less surface area, less water, Less bond
size cement, and paste
Well-graded aggregates Less paste Problem when good shape is
missing
Chemical admixtures Reduced permeability Cost, incompatibility, side effects
Cover More resistance to penetration of Wider cracks, extra weight and
solutions cost
Overlays Creates a low-permeability Difficult to place, expensive, and
protective layer over the prone to cracking; proper curing
conventional concrete is critical.
Corrosion inhibitors Stable protective layer on the Cost
steel

Note: SCM = supplementary cementitious material; na = not applicable.

181

Chapter 3. MATERIALS
Table 3.10 is a technology table that summarizes the solutions to reinforcement
corrosion. This table includes other protective methods mentioned previously in this
chapter, such as epoxy-coated, Z-bar, low-carbon chromium steel, and stainless steel.

Table 3.10. Technology Table for Corrosion of Reinforcement


Service Life
Issue Solution Advantage Disadvantage
Corrosion of Electrochemical chloride Extracts chlorides from the Extraction depends on the
reinforcement extraction concrete, or used in new depth and location, risk of
structures to increase corrosion embrittlement (prestressed).
threshold Difficult to predict service life.
Cathodic protection Prevents corrosion from High cost involved in maintaining
initiating, advantage as a repair the power source and sacrificial
method mesh anode. Embrittlement of
strand and softening of concrete
(prestressed structures).
Sealers Prevents solutions from Difficult to ensure adequate
penetrating the concrete, easy coverage. Varying performance
to apply either during or after and cost. Short service life.
construction Abrasion, sunlight, and
environment affect the sealer’s
efficiency.
Membrane Prevents moisture infiltration Varying performance. Difficult to
install on curved or rough decks
and to maintain quality and
thickness during field installation.
Stay-in-place metal form for Prevents infiltration of aggressive Cost.
marines structures solutions
Stainless steel High resistance to corrosion Initial cost.
Fiber-reinforced polymer High resistance to corrosion Fiber-reinforced polymer
prone to degradation from
environmental factors.
Z-bars (galvanizing over High resistance to corrosion Cost.
epoxy coating)
Epoxy-coated steel Creates protective layer over the Epoxy coating can be damaged
steel and increases the electrical during handling, shipping;
resistance and storage, and corrosion can
initiate under the coating.
Low-carbon chromium steel High resistance to corrosion High strength, no yield point.
Drainage design Minimize saturation Continuous maintenance.
Posttension Puts the concrete in compression, Posttensioning ducts and
minimizing cracks that facilitate grout are concerns in resisting
the penetration of chlorides corrosion.

182

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


3.5.3
Structural Steel
Individual strategies to mitigate factors affecting service life of structural steel are dis-
cussed in Chapter 6.
3.6
Overall Strategies for Enhanced Material Service Life
The introduction to this chapter describes a process for developing a strategy selection
to enhance material service life. This process is summarized in Figure 3.1.
Providing materials with enhanced service life requires a complete understanding
of the potential deterioration mechanisms. These mechanisms, described in Section
3.3, are associated with load-induced conditions, local environmental hazards, pro-
duction-created deficiencies, and lack of effective operational procedures. Mitigation
of these deterioration mechanisms through the selection of enhancement techniques,
described in Section 3.5, requires a thought process that combines the individual strat-
egies to define a single family of symbiotic strategies. This process will produce the best
approach to providing materials with enhanced service life.
This chapter provides guidelines for selecting the most appropriate individual
strategy to achieve the desired service life. Although the individual strategies provide
solutions to many of the material durability issues, the majority of the strategies must
be developed in conjunction with the material’s application, such as bridge decks. Sub-
sequent chapters will refer to this chapter as needed.
3.6.1
Design Methodology
With limited funds available for bridge construction, cost is often an overriding factor
in critical material selection decisions. However, to take advantage of the long-term
advantages of durable materials, service life enhancement strategies must be applied to
a cascading series of economic, design, construction, and maintenance measures. Suc-
cess of the strategy selection process is dependent on the ability to predict service life
and the incorporation of best practices to enhance service life.
3.6.2
Material Selection and Protection Strategies
The selection of the type of concrete for a particular application depends on many
factors, including the design of the structure (span length and slenderness of columns),
availability of the type of concrete, subsurface conditions, and the environmental con-
ditions (temperature and chemical exposures). Examples of factors to consider in the
selection of concrete include the following:

• If poor soil conditions exist and longer spans are planned, or if the substructure
is to be kept but additional or wider lanes and shoulders are planned, lightweight
concrete would be the material of choice.
• If there is severe exposure to salts or marine spray, high-performance concretes
with low permeability would be appropriate.

183

Chapter 3. MATERIALS
• In areas with congested reinforcement or intricate formwork, high-performance con-
crete with high workability, such as self-consolidating concrete, would be preferred.
• In bridge decks, self-consolidating concrete can lead to difficulty in maintaining
the grade or the cross slope due to high flow rates; normal-weight concrete may be
preferable unless durability or weight is of concern.
• Ultrahigh-performance concrete can be used when very small cross sections or
height restrictions exist or if high bond strengths and low permeabilities are
needed, as in connections.

Care should be exercised to select or specify only the necessary criteria for the
subject application. Additional criteria can cause undesirable distresses that adversely
affect performance and also increase the cost of construction. For example, for a
bridge deck, if a low w/cm (less than 0.40) is specified to achieve lower permeability,
high strengths will be obtained that would make the bridge-deck concrete more prone
to cracking. High strengths are accompanied by high stiffness (elastic modulus) and
low creep, which are instrumental in increased cracking potential. Cracks will facili-
tate the intrusion of chlorides, negating the benefits obtained by low w/cm. A better
approach would be to use moderate w/cm (0.40 to 0.45) with pozzolanic material to
reduce permeability. In addition, to achieve a low w/cm, high cement factors are used
that would increase the cementitious material and paste contents, thus making the
concrete more vulnerable to shrinkage and thermal problems.
Table 3.11 summarizes durability strategies for concrete materials. The strategies
for each potential deterioration mode must be compared for conflicts in order to estab-
lish the overall strategy to be deployed. For example, a designer faced with a bridge
deck having the potential for deterioration from wear and abrasion and differential
shrinkage should not specify concrete with both high strength and a low modulus. In
this case, using an overlay or membrane would be more appropriate.
Because the selection of appropriate material and protection strategies is highly
dependent on the application of the materials, material selection and protection strate-
gies considering the overall structure (not only materials) are provided in subsequent
chapters.
3.6.3
Construction Practice Specifications
Once the materials are selected, a proper set of specifications must be developed to en-
sure that the highest standard of care is used during construction. These specifications
and procedures are fairly well established and documented by FHWA and various state
agencies.
3.6.4
Maintenance Plan
An effective maintenance plan should be developed to ensure the maintenance assump-
tions regarding upkeep made in the material selection process are properly identified
for staff and budget requirements. If the bridge owner cannot commit to such a pro-
gram, then strategies for low-maintenance life-cycle costs should be recommended.

184

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 3.11. Concrete Durability Strategies
Potential Life-Cycle Costs
Deterioration Material Selection and Protective
Mode Measures Selection Maintenance Mode Initial Long Term
Freeze and Minimum 6% air entrainment None Low Low
thaw Sound aggregates
Strength >3.5 ksi
Proper drainage and cover None Low Low
Membrane or overlay Continual overlay Medium Medium
Replacement every 20 years
ASR Nonreactive aggregates None Medium Low
Low-alkali portland cement None Medium Low
Blended aggregates None Medium Low
Low-alkali portland cement
SCMs (e.g., fly ash, slag)
Blended aggregates None Medium Low
Low-alkali portland cement
Lithium nitrate
Proper drainage None Low Low
Membrane or overlay Continual overlay Medium Medium
Replacement every 20 years
ACR Nonreactive aggregates None Medium Low
Blended aggregates None Medium Low
Proper drainage None Low Low
Sulfate attack Cement with low C3A content, early None Low Low
curing temperature <160°F
Pozzolans, low w/cm, proper drainage None Low Low
Delayed Cement with low C3A content, early None Low Low
ettringite curing temperature <160°F
formation Pozzolans, low w/cm, proper drainage None Low Low

185

Chapter 3. MATERIALS
4
BRIDGE DECKS

4.1
Introduction
This chapter provides essential information and steps to be considered in developing a
bridge-deck system for a particular project in order to meet both strength and service
life requirements. Section 4.2 describes various deck systems and their known advan-
tages and disadvantages, as summarized in Table 4.1.
Section 4.3 summarizes factors that affect the service life of bridge decks by using
the fault tree format. Refer to Chapter 1 for a description of fault trees and how they
are constructed.
Section 4.4 provides strategies that can be used to mitigate most of the factors
affecting the service life of bridge decks, as described in Section 4.3.
Section 4.5 provides a framework for systematically addressing the service life
design of bridge decks designed for strength, based on design provisions stated in the
LRFD Bridge Design Specifications (LRFD specifications) (AASHTO 2012).
4.2
Description of Bridge-Deck Types
The primary function of a bridge deck is to provide a safe riding surface for traffic,
ensuring direct structural support of wheel loads. Two principal superstructure types
are considered as bridge decks in this section: (1) bridge decks cast on top of beams or
stringers, acting either compositely or noncompositely with superstructure supporting
elements; and (2) superstructure systems in which the top of the superstructure ele-
ment forms the top of the riding surface.
By definition, numerous types of systems qualify as bridge decks, including con-
crete deck systems, metal deck systems, timber deck systems, and fiber-reinforced poly-
mer (FRP) deck systems. The factors affecting service life of concrete deck systems,

186
the main system used in the United States, are further described in this chapter. Metal,
timber, and FRP bridge-deck systems are not addressed. Major bridge-deck systems
are summarized in Table 4.1 and are described in this section.
4.2.1
Concrete Bridge-Deck Systems
Concrete bridge-deck systems can consist of cast-in-place (CIP) systems and precast
systems.
The predominant bridge-deck system in the United States consists of CIP rein-
forced concrete. CIP concrete systems are defined as concrete bridge decks that are cast
in their final position. Typical CIP systems include

• Bridge decks on beams or stringers;


• Full-depth concrete slab superstructure;
• Multicell box girders; and
• CIP segmental construction.

Precast concrete systems are defined as concrete bridge decks that are cast remotely
and then brought to the bridge site for assembly into the final structure. Typical precast
systems include

• Adjacent member;
• Deck panel over beams or stringers; and
• Precast segmental construction.

Table 4.1. Bridge-Deck Systems


Type Advantage Disadvantage
Cast-in-place concrete deck Readily available material. Susceptible to cracking and corrosion.
systems Accommodates tolerances.
Low cost.
Precast concrete deck Readily available material. Requires construction joints between
systems Typically prestressed, reducing cracking. components.
Higher initial cost.
Metal deck systems Lightweight system. Requires protective coatings.
Prefabricated system. Difficult tolerance adjustments.
High cost.
Timber deck systems Lightweight system. Limited span range.
Constructible with unskilled labor. Susceptible to wear without overlays.
Low cost. Susceptible to moisture degradation.
FRP deck systems Lightweight system. High cost.
Noncorrosive system. Limited history.
Requires overlay for traction.

187

Chapter 4. BRIDGE DECKS


4.2.1.1
Cast-in-Place Concrete Systems on Beams or Stringers
As shown in Figure 4.1, CIP bridge decks on beams or stringers are typically reinforced
with mild steel reinforcement. They are generally 7.5 to 9 in. in thickness and are cast
on forms that span between beams or stringers. These forms can either be removable
or stay-in-place.
4.2.1.2
Full-Depth Cast-in-Place Concrete Systems
Full-depth CIP concrete deck slab superstructures are a classification of bridge decks
that span between pier supports without the aid of supporting beams and stringers.
These deck slabs can be solid, can contain circular voids (as shown in Figure 4.2),
or can contain more trapezoidal-shaped voids such as those used in CIP multicell
box structures and CIP segmental structures. Voids are introduced to reduce the dead

Figure 4.1.  Cast-in-place concrete deck over


longitudinal beams and stringers.
Source: Courtesy Atkins North America, Inc.

Figure 4.2.  Full-depth cast-in-place concrete slab posttensioned with voids.


Source: Courtesy Atkins North America, Inc.

188 Source: Courtesy Atkins North America, Inc.

Figure 4.2. Full-depth cast-in-place concrete slab posttensioned with voids.


DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE
weight of the bridge. These bridge systems are usually conventionally reinforced with
mild steel reinforcing, but they can be posttensioned longitudinally and transversely to
achieve longer span lengths.
4.2.1.3
Precast Adjacent-Member Concrete Systems
One of the most commonly used superstructure systems is the adjacent-member su-
perstructure system that consists of prefabricated beam elements placed side-by-side in
close proximity. This system has been used in various forms to expedite construction
and minimize field forming and placing of concrete. These members are predominantly
prestressed concrete beam elements in the form of prestressed solid and hollow-cored
slab units (as shown in Figure 4.3), deck bulb-tees, double Ts, channels, and adjacent
box beams. These deck systems are typically built as simple spans, but they can be
constructed as continuous members. Typically, the members are tied together with a
continuous longitudinal grout- or concrete-filled shear key that allows for the trans-
verse distribution of applied vertical forces across the joint and prevents differential
movement between adjacent members. The members may also be either transversely
connected with conventional reinforcement or posttensioned together to develop the
moments across the joint.
4.2.1.4
Precast Concrete Deck Panels
As shown in Figure 4.4, precast concrete deck panel systems employ a series of precast
concrete panels that are usually full-depth in thickness and have a length and width
determined by specific bridge geometry. The length of the panel along the roadway is
approximately 8 to 12 ft and is typically dictated by transportation limitations and
crane capacity. Panels span the supporting girders and are designed with conventional
reinforcement or as prestressed concrete. The general preference of precasters and
contractors is to use prestressed concrete to eliminate possible cracking from handling
and shipping.

Figure 4.3.  Example of adjacent-member slab


unit superstructure system.
Source: Courtesy Atkins North America, Inc.

189

Chapter 4. BRIDGE DECKS


Figure 4.4.  Full-depth precast panel system.
Source: Courtesy University of Nebraska, Omaha.

Precast concrete deck panel systems include both transverse and longitudinal slots
for connections. The transverse slots are typically grout-filled keyways connected in
a manner similar to the adjacent-member bridge systems. The longitudinal slots may
consist of grouted (or concreted) pockets or block-outs to accommodate the shear
connections to the girder. The system may also require temporary support and forms
along the girder to retain the grout and some type of overlay to improve pavement
ride quality. Longitudinal posttensioning is typically included in the system to tie the
panels together; however, systems without posttensioning have been used. At the time
of writing this guide (2012), new nonposttensioned connections were being developed.
4.2.1.5
Precast Segmental Concrete Superstructure Systems
This structural system consists of numerous precast bridge elements that are post-
tensioned together to form either simple-span units or, more commonly, continuous
spans. Segmental construction has gained favor in locations where access is challeng-
ing, such as in deep valleys, environmentally sensitive areas, across existing roadways,
and where accelerated construction is warranted. The basic cross section of a segmen-
tal bridge is usually a box shape with a top slab serving as the bridge-deck riding sur-
face, as shown in Figure 4.5. The primary longitudinal reinforcement consists of either
posttensioning tendons or bars that can be installed either internal to the web or ex-
ternally inside the box section. The bridge deck is typically posttensioned transversely.
4.2.2
Metal Deck Systems
Metal deck systems are bridge-deck systems that rely on a metal such as steel or alumi-
num to provide the structural resistance to vehicle wheel loads. Metal deck systems can
consist of metal grid decks, orthotropic steel decks, or orthotropic aluminum decks.

190

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 4.5.  Segmental superstructure.
Source: Courtesy Atkins North America, Inc.

4.2.2.1
Metal Grid Decks
A metal grid deck system is a prefabricated module system consisting of main I- or
T-shaped sections and secondary crossbars combined to form a rectangular or diago-
nal pattern. These members can be either steel or aluminum, and the main elements
span between beams, stringers, or other crossbeams. This system is typically used for
movable bridges and for long-span structures in which a reduced bridge-deck weight
is demonstrated to have an economic advantage. It has also been used in deck replace-
ment projects. The system consists of open grid deck or can be combined with con-
crete to form a partially or fully filled grid deck. The partially or fully filled concrete is
typically cast flush with the grid service, or it can be cast above the unfilled deck. This
system is known as the Exodermic™ bridge-deck system. The addition of concrete in
these systems reduces noise, improves fatigue performance, and improves the ability to
channelize and collect storm water.
4.2.2.2
Steel Orthotropic Decks
Bridge structures can utilize the orthotropic steel plate as one of the key structural sys-
tems in the distribution of deck traffic loads and for stiffening the supporting slender
plate elements in compression. Generally, the orthotropic system consists of a flat, thin
steel plate stiffened by a series of closely spaced longitudinal ribs at right angles or or-
thogonal to intermediate floor beams (see Figure 4.6). The orthotropic deck is typically
made integral with the supporting bridge superstructure as a common top flange to the
floor beams and girders. This arrangement results in cost savings in the design of these
other components. The defining characteristic of the orthotropic steel bridge is that it
results in a nearly all-steel superstructure.
The orthotropic system has been used for many bridges worldwide, especially in
Europe, Asia, the Far East, and South America. The United States has not yet fully
embraced this technology and currently has fewer than 100 such bridges in inventory.

191

Chapter 4. BRIDGE DECKS


Source:
Figure Wolchuk
4.6.  (1963).
Orthotropic steel deck bridge.
Source: Wolchuk 1963.
Figure 4.6. Orthotropic steel deck bridge.

The orthotropic deck has been most commonly used in the United States for long-
span bridges in which the minimization of dead load is paramount and for redecking
bridges on urban arterials. Orthotropic construction has tremendous potential for use
in short- to medium-span girder bridges. The system has not been used more exten-
sively for economic reasons; however, its light weight makes it beneficial for increasing
a bridge’s load rating during a deck replacement when replacement of the bridge may
have been the only other alternative.
4.2.2.3
Aluminum Orthotropic Decks
The aluminum orthotropic deck system configuration is similar to the steel orthotropic
deck described in Section 4.2.2.2. The use of aluminum provides a corrosion resistance
advantage that can result in lower maintenance costs, as it does not need periodic
painting. Although aluminum is lighter than steel, its additional cost has often deterred
its use in the United States. Other factors to carefully consider that make aluminum
orthotropic decks different from the steel orthotropic deck system include differences
in thermal expansion coefficients, reactions with dissimilar materials, lower modulus
of elasticity, and lower fatigue strength of the material, particularly at weld locations.

192

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


4.2.3
Timber Bridge Decks
Timber bridge decks have been used for hundreds of years, but increases in vehicle
loads have typically restricted their use to low-volume roadways. The materials used
for these bridge decks can be rough sawn timbers or glue-laminated panels. Their per-
formance can be enhanced through the use of protective coatings that can minimize
water absorption, which can be detrimental to the service life of the timber.
The timbers can be posttensioned together to form stress-laminated decks or
nailed together to form spike-laminated decks. Overlays are typically provided on
these bridges to improve skid resistance; however, the overlay requires extensive main-
tenance due to the flexibility of the timbers and the numerous connections between
members.
4.2.4
Fiber-Reinforced Polymer Bridge Decks
FRP bridge decks and superstructure systems are an emerging technology. FRP decks
have been used for short-span bridges and for deck replacement on bridges. The prin-
cipal advantages of FRP as a material are that it is lightweight and does not corrode
under the same conditions as steel materials. It has shown promise for use in projects
for which deck replacement is needed (as shown in Figure 4.7), particularly if total
load capacity is relatively low.
FRP bridge decks and superstructures have been constructed in many states. Com-
parisons with traditional CIP concrete bridge-deck systems have shown that they
exhibit lower dead loads, higher live-load fatigue ranges, and lower dynamic allow-
ance (impact) (Albers et al. 2007).
Because the surface of the FRP material has low skid resistance and the material
itself is soft, overlay systems are required to provide a safe riding surface that has ade-
quate surface friction and can withstand daily traffic wheel-load abrasion. Failure of

Figure 4.7.  FRP bridge-deck and superstructure applications.


Source: Aboutaha 2001.

193

Chapter 4. BRIDGE DECKS


overlay adherence to the FRP material was noted in early applications of this technol-
ogy. Connections for crashworthy barriers for FRP decks present additional challenges.
Further research is recommended to study the long-term behavior of this new
material to demonstrate its ability to provide a sufficiently long service life.
4.3
Factors Influencing Bridge-Deck Service Life
Bridge decks are one of the most costly maintenance items within a typical bridge
system. The reduced service life of bridge decks can be attributed to two causes:
(1) obsolescence, which is a functional planning issue and not a factor relating to
durability issues; and (2) material service life performance deficiencies, which may be
load induced; caused by human activity or natural hazards; or result from produc-
tion defects in construction processes, design details, or operational procedures. These
­deficiencies are illustrated in the fault tree shown in Figure 4.8.

Reduced Service Life of


Cast-in-Place Bridge Deck

Caused by
Caused by Deficiency
Obsolescence

Natural or Man-Made Production/


Load-Induced
Hazards Operation Defects

Figure 4.8.  Bridge-deck reduced service life fault tree.

194

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Many interrelated factors during the design, construction, and management
phases of a bridge deck’s service life must be considered in developing long-lasting,
cost-effective bridge decks. These factors vary depending on the bridge-deck system
used, which can be arranged in four broad categories:

• Concrete bridge-deck systems, including


–– CIP concrete bridge-deck systems, and
–– Precast concrete bridge-deck systems;
• Metal deck systems;
• Timber deck systems; and
• FRP bridge-deck systems.

The factors affecting the service life of CIP concrete and precast concrete bridge-
deck systems are further described in this chapter. Metal, timber, and FRP bridge-deck
systems are not addressed in the Guide.
4.3.1
Cast-in-Place Concrete Bridge Decks
CIP bridge-deck systems are systems in which the concrete for the bridge deck is cast
in the field as an integral part of the final superstructure. This bridge-deck system is
one of the most common systems used in the United States today. These decks provide
a major constructability advantage in that the casting process easily molds the bridge
deck to meet geometric requirements (such as skews, lane tapering, and super­elevation
transitions) and to match existing locations of supporting elements that are not pre-
cisely located in accordance with the plans. The main disadvantages of these decks
include the quality of concrete produced as a result of workmanship and the curing
processes.
Inspections of bridge decks have revealed numerous performance issues with CIP
concrete, including cracking, corrosion of reinforcement, spalling, delamination, and
concrete deterioration evidenced by scaling, wear, and abrasion. Although concrete in
compression is considered a very durable construction material, tension introduced
through various loading and bridge restraint conditions can result in significant ten-
sion that can exceed the material’s tension strength limits, resulting in cracking. Crack-
ing of bridge-deck concrete reduces the integrity of the passivated concrete layer that
surrounds the reinforcing steel, significantly reducing the encased reinforcement’s
resistance to corrosion.
The following subsections discuss factors affecting the service life of CIP bridge
decks.
4.3.1.1
Load-Induced Bridge Deck Considerations
Load-induced bridge-deck deterioration can be attributed either to loads induced by
the traffic or by characteristics dependent on the overall bridge system. These load-
induced factors are shown in the fault tree in Figure 4.9.

195

Chapter 4. BRIDGE DECKS


Load-Induced

Traffic-Induced System-Dependent
Loads Loads

Wear and System-


Differential
Fatigue Overload Abrasion Thermal Framing
Shrinkage
Restraint

Figure 4.9.  Cast-in-place bridge-deck load-induced deficiency fault tree.


Figure 4.9. Cast-in-place bridge-deck load-induced deficiency fault tree.

4.3.1.1.1
Traffic-Induced Loads
Traffic-induced loads include the effects of truck and other vehicular traffic on the rid-
ing surface of the bridge. Bridge-deck loading has a degree of uncertainty that must
be addressed during the design of the bridge, especially when achieving long service
life is an objective. Typically the service life of bridge decks will be affected by fatigue,
overload, and wear and abrasion.
Fatigue. CIP concrete deck consists of two materials, steel and concrete, both of
which can fail by fatigue. Design provisions for fatigue are addressed in the LRFD
specifications.
Overload. Despite weight limit regulations in most states that define load limits
for permit and legal truck configurations, overloads exceeding these limits do occur.
Overload is one of the main reasons for reduced service life of bridges.
Overloads result in additional flexural stresses in bridge decks that can cause
excessive cracking not accommodated by the original design. Heavier tire loads may
also affect the wear and abrasion on the structure, and multiple applications of these
loads can affect the fatigue behavior of the deck.

196

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Wear and abrasion. Wear and abrasion is typically affected by high traffic volume,
high tire loads, and the types of tires used on the facility. Tires in cold climates may
have features to aid in traction, such as deep grooves, studs, and chains. These added
tire features, while aiding traction, can abrade the surface of the bridge deck.
Wear and abrasion can result in reduced thickness of the bridge deck, which in turn
reduces the concrete cover protecting the reinforcement from corrosion; can reduce the
load-resisting section, resulting in higher stresses and cracking; and can change the deck
stiffness assumed for distribution of loads between superstructure elements.
4.3.1.1.2
System-Dependent Loads
System-induced loads include the effects of the bridge system configuration on the be-
havior of the bridge deck, such as restraint of integral abutment systems.
Differential shrinkage. Differential shrinkage occurs when bridge-deck concrete
is cast over previously cured concrete or over steel girders. The shrinkage of fresh
concrete is restrained by the cured concrete or steel stringers, which results in a set of
equal and opposite forces causing tension in the deck and compression in the girders.
Heat of hydration also contributes to development of tensile forces in the freshly cast
concrete deck.
Thermal. Temperature changes can result in the development of axial forces in the
bridge deck. These thermal forces are due to uniform internal temperature changes
and temperature gradient. The level of these thermally induced axial forces is a func-
tion of bridge system boundary conditions.
System-framing restraint. Bridge decks can be subject to additional axial forces cre-
ated by bridge boundary conditions. Bridge system boundary conditions are set at the
design stage, during bridge system selection. For instance, in integral abutment joint-
less bridge systems, the elimination of expansion devices and reliance on flexibility of
piles to resist the bridge expansion and contraction add axial forces to the deck. Other
examples of boundary conditions capable of creating axial forces in the bridge deck
include choices for bearings and connections between superstructure and substructure
made during design. These axial forces range from compression during system expan-
sion to tensile forces during system contraction. Resistance to system contraction can
create tensile forces in the bridge deck and cause cracking. Calculation of these addi-
tional axial forces is important and can be achieved through conducting proper analysis
methods that correctly model the bridge boundary conditions.
Improper function, or seizing of the bearings, results in unintended movement
restraint that can raise the force resisted by the substructure well above the intended
design. This unintended restraint can cause unanticipated cracking with greater poten-
tial for corrosion. Proper bearing function, which is addressed in Chapter 10, is essen-
tial to substructure durability. Lack of maintenance may also result in bearings losing
the movement capability intended by their design.

197

Chapter 4. BRIDGE DECKS


4.3.1.2
Natural or Man-Made Hazard Bridge-Deck Considerations
The environment to which the bridge deck is subjected can have a significant influence
on its service life. Environmental influences comprise hazards from both natural and
sources and include effects from areas with adverse thermal climate, climates, and
chemical climates, as well as from chemical properties of the materials and outside
agents, such as fire. These natural and man-made hazards are listed in the fault tree in
Figure 4.10.
4.3.1.2.1
Thermal Climate
Thermal climate influences on bridge-deck service life performance are primarily due
to cold weather. These influences are both man-made, from the application of deicing
salts, and natural, in the case of freeze–thaw.
Agencies in cold weather climates that deal with ice and snow on roadways and
bridges have traditionally applied deicing salts to melt the ice and snow to facilitate
tire traction. The application of these deicing salts is viewed as a safety enhancement
for the traveling public; however, these chloride-laden compounds tend to ingress into
the concrete deck either through porosity in the concrete or through open deck cracks.
The chloride ingress into the bridge deck continues to reduce the effectiveness of the
passivating layer around the reinforcing steel, eventually initiating reinforcement cor-
rosion. The reinforcement corrosion process causes the bar to expand, resulting in
deck cracking, spalling, and delamination. The cross slope built into bridge decks for
drainage purposes causes the salt to wash down toward the bridge gutter adjacent

Natural or Man-
Made Hazards

Thermal Coastal Chemical Reactive Hydraulic Extreme


Climate Climate Climate Materials Action Events

Salt
Deicing Corrosion- Flood/ Vehicle/
Water/ Seismic
Salts Inducing ASR Storm Vessel
Spray
Corrosion Chemicals Surge Collision
Corrosion

Freeze– Sulfate
Humidity Attack ACR Scour Fire/Blast
Thaw

Figure 4.10.  Cast-in-place bridge deck natural or man-made hazards fault tree.
[LANDSCAPE IN FINAL]

198 Figure 4.10. Cast-in-place bridge deck natural or man-made hazards fault tree.

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


to the traffic railing barriers bounding the bridge. Removal equipment that scrapes
snow from the bridge deck also deposits residual snow laden with deicing salts at this
location, resulting in a very high concentration of chlorides. Construction joints at
this location are particularly susceptible to chloride intrusion, and in many cases this
susceptibility has led to corrosion of the barrier reinforcement.
Water absorbed into the concrete deck surface and contained in cracks can freeze
in cold weather conditions. The frozen water tends to expand, causing stresses within
the concrete. Cyclic freezing and thawing of the water absorbed in the deck surface can
result in bridge-deck deterioration in the form of cracking, scaling, and spalling. Refer
to Chapter 3 for additional information on freeze–thaw in concrete.
4.3.1.2.2
Coastal Climate
Coastal climate influences on bridge-deck service life performance are primarily due to
the introduction of chlorides through salt spray and from the effects of high humidity.
Both of these influences occur naturally.
Coastal regions are subjected to a chloride-laden saltwater environment and a
combination of wind and wave action that causes these chlorides to become airborne
as salt spray. The susceptibility of the bridge deck to these environmental influences
depends on the height of the bridge deck above the water elevation and the distance
to coastal areas. The action of waves hitting substructure units and seawalls or abut-
ments under the bridge tends to cause the salt spray to explode upwards, wetting
the bottoms of lower-level bridge decks. The salt spray can also be deposited on the
bridge-deck surface, particularly on windy days. When the salt spray wets the surfaces,
it leaves a chloride residual that can absorb into the concrete, resulting in reinforce-
ment corrosion.
High humidity in coastal regions also results in cyclical wetting and drying of con-
crete surfaces. Concrete materials sensitive to repeated wetting, such as those where
reactive aggregates are used, can have an adverse effect on the bridge-deck service life.
4.3.1.2.3
Chemical Climate
Chemical climate influences on bridge-deck service life performance can be attributed
to corrosion-inducing chemicals and sulfate attack. These influences can occur natu-
rally or can be man-made.
Corrosion-inducing chemicals can be introduced to the bridge deck from adjacent
industries, where residuals from pollution can contribute to reduction in bridge-deck
service life. For example, oil- and coal-burning facilities release sulfur dioxide and
nitrogen oxide into the air, which causes acid rain consisting of sulfuric and nitric
acids. These acids can dissolve cement compounds in the cement paste and calcareous
aggregates and can leave crystallized salts on concrete surfaces that can lead to spalling
and the corrosion of reinforcing bars.
Exposure to sulfates can cause expansion of the concrete material and conse-
quently result in spalling and cracking of the bridge deck. Refer to Chapter 3 for addi-
tional information on sulfate attack in concrete.

199

Chapter 4. BRIDGE DECKS


4.3.1.2.4
Reactive Ingredients
Reactive ingredients within the mix used for bridge decks can affect service life perfor-
mance as the reactive ingredients alter the volumetric stability of the concrete. These
influences primarily occur naturally.
Alkali-silica reactivity (ASR) results in swelling within concrete that can lead to
spalling, cracking, and general concrete deterioration. Refer to Chapter 3 for addi-
tional information on ASR in concrete.
Alkali-carbonate reactivity (ACR) results in aggregate expansion within concrete
that can lead to spalling, cracking, and general concrete deterioration. Refer to Chap-
ter 3 for additional information on ACR in concrete.
4.3.1.2.5
Fire
A key factor in the amount of damage caused to concrete by fire is the duration of the
fire and the heat levels generated. Because of the low thermal conductivity of concrete,
it takes considerable time for the interior of concrete to reach damaging temperatures.
When concrete is exposed to the extreme heat of a fire, the chemical bonds between
the water molecules in the concrete break, resulting in dehydration and the destruc-
tion of the cement binder. The concrete loses its mechanical properties, exhibiting
cracking and spalling, and exposes steel, leaving it unprotected (ACI 216, 1989). Once
the reinforce­ment has become exposed, it conducts heat and accelerates this action.
Reinforcing steel in bridge decks subjected to temperatures above 550°C (1,022°F)
exhibits a rapid reduction of strength, which can lead to collapse. In addition, spalling
can result from the rapid quenching of hot fires by fire hoses.
4.3.1.3
Design, Construction (Production), and Operation Bridge-Deck Considerations
Decisions made for the design and construction of bridge decks and the activities that
will occur during their operation can significantly influence service life. These influ-
ences, which are listed in the fault tree in Figure 4.11, include decisions made during
the design and detailing of the bridge deck, the quality of construction, the level of
inspection, and the testing performed during operations and maintenance.
4.3.1.3.1
Design and Detailing Bridge-Deck Considerations
Decisions made during the design and detailing phase of a bridge project can signifi-
cantly affect the service life of the bridge. It is incumbent on designers to understand
the implications of these decisions in order to make rational choices that will improve
the service life of bridge decks. These decisions are listed in the fault tree in Figure 4.12
and include choices in design philosophy, expansion joints, construction joints, con-
crete mix design, and bridge-deck drainage.
4.3.1.3.1a
Design Philosophy
There are two principal methods presented in the LRFD specifications for the design
of bridge decks: the traditional design method and the empirical design method.

200

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Production/
Operation Defects

Design/Detailing Construction Inspection Maintenance

Field Visual
Placement
Bending

Protection
Dissimilar Nondestructive
and
Metals Testing
Repair

Figure 4.11.  Cast-in-place bridge-deck design, construction (production) and operation defects fault tree.
Figure 4.11. Cast-in-place bri dge-deck design, construction (production), and operation defects fault tree.

The traditional design method assumes flexural action to describe the behavior of
bridge-deck spanning between supporting girders and ignores the axial forces created
in the bridge deck as a result of arching action. Under this assumption, the amount of
reinforcement needed in the bridge deck will generally far exceed the demand, usually
by more than a factor of two. Providing additional reinforcement in the bridge deck
creates additional means for corrosion and deterioration of the bridge deck.
The empirical design method provides better estimation of bridge-deck resistance
to applied traffic loads than the traditional design method. Test results (Fang 1985;
Holowka et al. 1980) show that the principal mechanism for resisting the applied
traffic loads in the bridge deck is the creation of axial compressive loads, commonly
referred to as arching action. These axial compressive loads are resisted by supporting
longitudinal beams. Consequently, the use of the empirical method is not applicable
to cantilever portions of the deck. The axial compressive loads in the bridge deck

201

Chapter 4. BRIDGE DECKS


Design/
Detailing

Design Expansion Construction


Mix Design Drainage
Philosophy Joints Joints

Target
Design Empirical Permeability Passivity
Life Modular
Construction

Composite Cracking
Traditional Workability
Action Resistance

Phasing

Other Creep and


Methods Shrinkage

Figure 4.12.  Cast-in-place bridge-deck design and detailing deficiency fault tree.
Figure 4.12. Cast-in-place bridge-deck design and detailing deficiency fault tree.

significantly reduce the need for reinforcement in the bridge deck, and reduction of
reinforcement in the bridge deck significantly reduces the sources of corrosion.
Research in Canada in the past 20 years (Newhook and Mufti 1996) has focused
on eliminating bridge-deck reinforcement corrosion through the development of
“steel-free” bridge decks. Similar to the methodology employed by the empirical
design method, this concept provides the tension tie required to resist the compressive
forces created in the bridge deck by arching action. Under this concept, the tension ties
are attached to top flanges of the supporting beams or stringers. These bridges have
experienced some temperature and shrinkage cracking. To control this cracking to
acceptable levels, recent recommendations suggest supplementing the steel tension ties
below the deck with a mat of FRP reinforcing bars (Memon and Mufti 2004).

202

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


4.3.1.3.1b
Expansion Joints
Expansion joints are provided to relieve system-framing restraints that can cause a
buildup of tension stresses in the superstructure and the bridge deck. Refer to Section
4.3.1.1.2 for additional information on system-framing restraints and to Chapter 9 for
additional information on expansion joints.
Dirt and other deleterious material, such as deicing salts, can collect on expan-
sion devices within the bridge deck and produce an adverse effect on the service life
of bridge decks. Impact from vehicles and from snow removal equipment can cause
spalling, which reduces the protective concrete cover over reinforcement or exposes
the reinforcement, which leads to corrosion.
4.3.1.3.1c
Construction Joints
Construction joints are surface discontinuities at which successive concrete placement
regions meet; they are generally specified by the designers or construction contrac-
tors. Typical construction methods for which construction joints are required include
modular construction and phasing of construction.
The modular construction used to accelerate bridge construction requires con-
struction joints within adjacent-member superstructure systems. Typically these joints
are designed to transfer shear and moment across the interface, and if not properly
designed and detailed they may lead to cracking along the adjacent-member interface.
Phasing of bridge construction is often a result of public pressure and demand for
uninterrupted traffic flow during bridge construction. A phased approach is used both
for widening of existing bridges and the construction of new bridges. In the case of
new construction, a portion of the new bridge is constructed (Phase 1) while the exist-
ing bridge carries the traffic. Traffic is then transferred to the new bridge (Phase 1), the
existing bridge is demolished, and the new bridge is completed (Phase 2). Finally, the
two phases are typically joined using a closure pour. The phases will experience differ-
ent deflections at the time of placing the closure pour; this differential deflection can
result in major construction problems.
One of the characteristics of phased-constructed bridges is that transverse and
longitudinal cracks form near the points at which Phases 1 and 2 are joined. Forma-
tion of these cracks is a well-known feature of these bridge types. Further, closure-pour
regions need to be water proofed to prevent deterioration of the deck in these regions
as a result of chloride ingress and initiation of reinforcement corrosion. Placing addi-
tional reinforcement in the deck will not prevent formation of these cracks; it will only
make the crack width smaller.
A well-constructed phase bridge can perform very satisfactorily (Azizinamini et al.
2003b). Nevertheless, several factors (described in this subsection) need to be taken
into consideration.
Composite sections experience long-term displacement because of creep and
shrinkage. Figure 4.13 shows an exaggerated difference between elevations of Phase 1
and Phase 2 for newly constructed bridges. The same phenomenon also exists for
widening projects.

203

Chapter 4. BRIDGE DECKS


Figure 4.13.  Exaggerated differential displacement between phases.
Source: Azizinamini et al. 2003b.

The reason for the observed differential displacement shown in Figure 4.13 is
illustrated in Figure 4.14, which shows the displacement of the Phase 1 girders due to
creep and shrinkage. At about 90 days after completion of Phase 1, the girders experi-
ence maximum creep and shrinkage displacement. As shown along the horizontal axis,
construction of Phase 2 starts after Phase 1 is completed. Phase 2 also experiences the
creep and shrinkage displacements and, depending on the time of casting the closure-
pour region, differential displacement will exist between Phase 1 and Phase 2 girders.
In the case of steel bridges, this differential deflection between the two phases has
resulted in major fit-up problems for the cross frame in the bay containing the closure
pour. Once the closure-pour region is cast, the two systems are locked in, but Phase 2
continues to experience additional displacements that subject the deck in the closure-
pour region to additional stresses.

Figure 4.14.  Displacement of


Deflection

Phase 1 and Phase 2 portions of


the bridge.
Source: Azizinamini et al. 2003b.

Time

204

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


When the deck elevations of the two phases do not match, the contractor may
attempt to force the two separate superstructure portions together. This practice sub-
jects the deck in the closure-pour region to additional stresses, which are difficult to
estimate and may also jeopardize the service life of the deck concrete.
There is debate among bridge engineers on conditions under which the closure-
pour region should be cast. Some contractors prefer to close traffic completely until
the concrete in the closure-pour region is set, but others believe that vibration caused
by traffic is helpful for better consolidation of the concrete in the closure-pour region.
Generally, traffic closure is not an option because of public demand that traffic inter-
ruptions be minimized.
Another major condition that could facilitate the construction of phased bridges
is the elimination of cross frames in the bay containing the closure pour, if possible.
However, the pros and cons of such action need to be investigated.
4.3.1.3.1d
Mix Design
Chapter 3 provides detailed information concerning factors that affect the service life
of concrete. The following paragraphs briefly describe mix design factors that affect
the service life of bridge decks: permeability, passivity around reinforcement, crack
resistance, workability, and creep and shrinkage.
Permeability. The durability of concrete largely depends on its ability to resist the
infiltration of water and aggressive solutions. Concretes with high permeability pro-
vide less resistance to aggressive solutions or water penetrating the concrete and pos-
sibly causing expansive forces due to physical (freeze–thaw) or chemical (corrosion,
ASR, sulfate attack) factors.
Passivity around reinforcement. The loss of passivity of the outer layer of the
reinforcing steel initiates a corrosion process that deteriorates the steel. This corrosion
process begins by the diffusion of chloride ions to the depth of the reinforcing steel
and/or carbonation, which reduces the pH of the concrete to the passivating layer sur-
rounding the concrete.
Crack resistance. Mix design can affect the extent of cracking for all CIP concrete
bridge decks and slab superstructures. Mixtures with high water and paste content are
prone to shrinkage cracks that occur over time. The use of large aggregate sizes and
well-graded aggregates reduces the water and paste content and minimizes shrinkage.
In fresh concrete, when the rate of evaporation exceeds the rate of bleeding, plastic
shrinkage occurs. Concrete with low-bleed water, stiff consistency, and a low water-
cement ratio is prone to plastic shrinkage cracking. Prevention of plastic shrinkage
cracking depends on prompt, effective curing.
Workability. Concrete mix designs must include good-quality aggregates and
appropriate admixtures to facilitate construction. Mix designs with poor workability
can cause uncontrolled field adjustments to the mix through the addition of water
in the field, resulting in higher water-cement ratios and overvibration that can cause
aggregate segregation. Proper workability must be ensured for the integrity of the mix
design to provide concrete with the intended properties.

205

Chapter 4. BRIDGE DECKS


Creep and shrinkage. The creep and shrinkage properties of concrete mixes can
affect the service life performance of bridge decks. The adverse effects of concrete’s
shrinkage characteristics are discussed in conjunction with system-dependent loads
in Section 4.3.1.1.2. In contrast to the effects of shrinkage, high levels of creep can
be either beneficial or detrimental depending on the application. High creep levels
are beneficial when differential shrinkage occurs, particularly within bridge decks. In
instances of higher creep, the restraining force between the bridge deck and the sup-
porting superstructure will reduce significantly, consequently reducing the potential
for cracking. However, high creep levels are detrimental when the creep results in
unrestrained volumetric changes that cause significant, unintended movements of the
structure, such as in posttensioned structures.
4.3.1.3.1e
Drainage
In design, poor drainage details can result in ponding and prolonged exposure of
bridge components to moisture and aggressive solutions, causing corrosion and other
environmental distress. When concrete gains moisture, it expands slightly or swells.
When concrete loses moisture, the concrete contracts or shrinks. As drying occurs, the
portion of concrete near the surface dries and shrinks faster than the inner portion of
the concrete. This process results in a differential moisture condition in which tensile
stresses that can cause cracks may occur on the surface.
The degree of frost damage to concrete is also highly dependent on the degree of
saturation. Ponding of water on bridge decks can cause critical saturation that results
in bridge damage. High moisture is also detrimental to concrete susceptible to ASR
expansion, which can cause spalling and cracking.
4.3.1.3.2
Construction
Attention to good practices during construction is crucial to the long-term durability
of reinforced concrete. A well-qualified and well-trained work force and work that
is well executed increase productivity, reduce material waste, and provide expected
service life. The proper use of appropriate equipment provides better workability by
increasing efficiency, and well-planned construction schedules reduce overall costs by
providing set times for equipment rental and reducing downtime. The correct imple-
mentation of test methods ensures quality concrete.
4.3.1.3.2a
Placement and Curing of Concrete
Good construction practices, which ensure the proper location of reinforcing steel for
proper cover depth, consolidation, and curing, are essential for longevity. Proper con-
solidation minimizes entrapped air voids, which can reduce strength and durability.
Proper curing is necessary for formation of the binder and control of volumetric
changes and includes both moisture and temperature control. In bridge structures, the
deck surfaces require special attention because of their large surface areas, where loss
of moisture is a concern.

206

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Handling of concrete affects the final product. Delay in placement, particularly
on hot days, should be avoided as it can lead to stiffening of the concrete that can
cause tearing of the deck surface during finishing, resulting in a poor surface finish and
reduced durability.
4.3.1.3.2b
Formwork
The type of concrete formwork can affect the surface finish of the concrete. Imperme-
able forms can allow surface voids to occur, resulting in increased surface permeability,
reduced strength, and an overall decrease in durability.
CIP bridge decks on beams or stringers are cast on forms that span the beams or
stringers. These forms can be either removable or stay-in-place. Stay-in-place forms
are typically made of galvanized steel (as shown in Figure 4.1), precast concrete panels
(either conventionally reinforced or prestressed), and FRP. Many owners do not allow
the use of stay-in-place forms, citing the inability to inspect the bottom surface of
the concrete deck and the potential for the collection of water intruding through the
cracks.
The use of precast concrete stay-in-place panels designed to be composite or non-
composite with the deck pour above has resulted in reflective cracking over the panel
joints in past applications and has raised questions about their long-term durability.
Research is needed to develop acceptable details for control or elimination of reflective
cracks.
4.3.1.3.2c
System Vibration During Construction
Excessive traffic-induced vibrations during construction can occur during the widening
of a new bridge deck that is being cast against an existing bridge deck subjected to ac-
tive traffic. The effect of these vibrations on the quality of the finished deck, especially
in the closure-pour regions, is not well understood and requires further investigation.
4.3.1.3.2d
Casting Schedule
The casting schedule for bridge decks should carefully consider weather conditions,
particularly hot days. The rise of heat from hydration in concrete can be exacerbated
by a concurrent rise in the ambient temperature, resulting in a greater cooling differ-
ential that can cause bridge-deck cracks. It takes approximately 18 hours for heat of
hydration to reach its peak value. After casting, the concrete at the surface is always at
ambient temperature, while within the deck, the temperature varies due to the develop-
ment of heat of hydration. The center of the deck cools last. As a result, the maximum
temperature differential takes place between the center of the deck and the deck sur-
face. When the maximum temperature differential exceeds a certain limit, which most
departments of transportation limit to about 30°F, the deck can crack. Therefore, this
maximum temperature differential needs to be controlled. When casting is performed
at night, the peak ambient temperature generally occurs 12 to 18 hours later, when the
center of the deck is at its highest temperature. Timing the casting in this way results
in a minimum temperature differential between the center and outside surfaces of deck

207

Chapter 4. BRIDGE DECKS


concrete. In contrast, when the deck is cast in the morning, the temperature differential
between the center of the deck and the surface of deck is much higher. In conclusion,
although casting a deck at night is ideal, the common practice is to cast the concrete
deck in the morning, the most undesirable time.
4.3.1.3.2e
Casting Sequence
Construction joints within bridge decks control the sequence of casting bridge decks,
either as a result of concrete volume placement constraints, or, in the case of continu-
ous steel girders, in order to minimize bridge-deck tension in the negative moment
area over substructure elements. These locations form a discontinuity that can open as
cracks within the bridge deck and allow ingress of moisture.
4.3.1.3.3
Visual Inspection
Although inspections are valuable tools for identifying deficiencies in bridge decks,
they are typically visual, making them subject to the ability, training, and disposition
of the individual inspector. Often a deficiency is not easily detectable and may show
only subtle signs that can easily be missed by cursory inspections or by inexperienced
inspectors. Deficiencies can also be located below the undamaged surface or in inac-
cessible areas. The inability to see the deficiency leads to inadequate identification of
repair methods, scope, and material selection and could cause failure of the structure
without visible signs, because surficial repairs may cover the damaged area.
4.3.1.3.4
Maintenance
Lack of preventive maintenance reduces the service life of bridge decks. Sometimes
simple maintenance tasks are delayed until a problem becomes a safety issue, at which
time the required repairs may be either significantly more extensive or ultimately
irreparable.
4.3.2
Precast Concrete Bridge-Deck Systems
Precast concrete bridge-deck systems are systems in which the concrete components
for the bridge deck are produced in a controlled environment, minimizing v­ ariability
in concrete uniformity of both material behavior and construction personnel per-
formance. Use of these systems can minimize traffic disruption caused by prolonged
­concrete casting operations over active roadway facilities; such systems are a key con-
sideration for accelerated bridge construction.
Inspections of precast concrete bridge decks have revealed many of the same per-
formance issues experienced with CIP concrete as described in Section 4.3.1, including
cracking, corrosion of reinforcement, spalling, delamination, and concrete deterio-
ration evidenced by scaling, wear, and abrasion. Precast concrete bridge-deck com-
ponents introduce numerous joints in the superstructure that are usually the source
of many bridge-deck service life issues, particularly when the material provided to
seal these construction joints breaks down, causing cracking, leakage, and eventually
reinforce­ment corrosion.

208

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Rather than repeat the discussion of the many performance-related service life
issues inherent with concrete systems described in Section 4.3.1, this section describes
the factors affecting service life specific only to precast concrete bridge-deck systems.
The user of this Guide should also become familiar with the other factors described in
Section 4.3.1 when considering precast concrete bridge-deck systems.
Load-induced (Section 4.3.1.1) and natural or man-made hazard (Section 4.3.1.2)
bridge-deck considerations for precast bridge-deck systems are the same as those for
CIP bridge-deck systems. The following subsections discuss the factors affecting pre-
cast concrete bridge-deck service life in production and operation.
4.3.2.1
Design, Construction (Production), and Operation Bridge-Deck Considerations
Decisions regarding the production of a bridge deck and the activities that will occur
during its operation can have a significant influence on service life. These produc-
tion and operation influences are listed in the fault tree in Figure 4.15. They include

Production/
Operation
Defects

Fabrication/ Visual
Design/Detailing Construction Maintenance
Manufacturing Inspection

Vibration
Placement During
Construction
Tolerances/
Field Fit-up

Casting
Curing
Schedule
Lifting
Embedments

Connection
Formwork
Integrity

Figure 4.15.  Precast concrete bridge-deck design, construction (production) and operation defects fault tree.
Figure 4.15. Precast concrete bridge-deck design, construction (production),and operation defects fault tree.
209

Chapter 4. BRIDGE DECKS


decisions made during the design and detailing of the bridge deck, fabrication and
manufacturing requirement, quality of construction, and decisions concerning the
level of inspection and testing performed during future operation and maintenance.
4.3.2.1.1
Design and Detailing Bridge-Deck Considerations
Decisions made during the design and detailing phase of a bridge project can signifi-
cantly influence the service life of the precast bridge deck. It is incumbent on design-
ers to understand the implications of these decisions in order to help make rational
choices that will improve service life. These influences, which are listed in the fault
tree in Figure 4.16, include choices regarding design philosophy, expansion joints,
construction joints, concrete mix design, and bridge-deck drainage. Again, many of
these influences are similar to those for a CIP concrete deck, except for the addition of
composite action considerations for precast decks.

Design/Detailing

Design Expansion Construction


Mix Design Drainage
Philosophy Joints Joints

Target
Design Empirical Permeability Passivity
Life
Modular
Construction

Composite Cracking
Traditional Workability
Action Resistance

Phasing
Creep
Other
and
Methods
Shrinkage

Figure 4.16.  Precast concrete bridge-deck design and detailing deficiency fault tree.

210

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


4.3.2.1.2
Composite Action
For precast systems such as full-depth deck panels, the design philosophy addresses
either a composite or a noncomposite connection of the deck panels to the supporting
superstructure element.
Composite bridge-deck systems add to the stiffness of the overall bridge system,
reducing deflection and vibration and improving bridge-deck performance. The con-
nection requirements for composite systems are developed through field casting of
concrete or grout around shear connectors or studs accessed through continuous
full-depth open pockets (as shown in Figure 4.4); through localized, full-depth open
­pockets; or through continuous or localized embedded channels in the panels under
the deck surface. Open-pocket systems introduce construction joints, forming a dis-
continuity that can open as cracks. Embedded channel systems require pressure grout-
ing to fill the void. Improper grout installation can lead to entrapped air voids that can
fill with bleed water and water intruding through deck cracks, leading to freeze–thaw
issues and increased potential for reinforcement corrosion.
In noncomposite systems, excessive flexibility and inconsistent friction between
the deck panels and the superstructure along the length of the supporting stringer
could result in localized stress that can cause cracking and delamination in the con-
crete. Excessive vibration can also lead to fatigue issues.
4.3.2.1.3
Modular Joint Construction
In precast systems consisting of adjacent members or segmental construction, the de-
sign and detailing of the joint is essential to its proper performance. These joints can
open as cracks within the bridge deck if not properly considered, leading to leakage,
spalling, and reinforcement corrosion.
4.3.2.1.4
Precast Component Fabrication and Manufacturing Considerations
Decisions relating to the fabrication and manufacturing of precast components for
bridge decks can significantly affect the service life of the bridge. It is incumbent on
the fabricators to understand the implications of these decisions in order to help make
rational choices for improving the service life of bridge decks. These decisions are
listed in the fault tree in Figure 4.15; they include choices in field fit-up and casting
tolerances, as well as methods for lifting the precast elements in the precast yard and
for erection in the field.
4.3.2.1.4a
Tolerances and Field Fit-Up
Careful planning is needed to incorporate precast components as bridge decks on
bridges. These components must be aligned in the field fairly accurately to ensure
a smooth, safe ride for the traveling public. Often the construction sequence and
schedule must be assessed to establish casting dimensions for the precast compo-
nents. ­Liberal tolerances and insufficient control in the precasting facility can result

211

Chapter 4. BRIDGE DECKS


in ill-fitting pieces that require unintended adjustments in the field, leading to spalling
and other structural failures caused by localized, nonuniform contact surfaces that
were not anticipated during design.
4.3.2.1.4b
Transportation and Lifting Methods
Precast components must be moved from the casting facility to their final position in
the bridge. This transport may consist of multiple lifts, depending on storage require-
ments and transportation schedules. The components must also be transported be-
tween the casting bed, storage facilities, and the project site. Improper consideration of
the stresses imposed on the precast components during these critical events can result
in cracking, spalling, and sometimes failure.
Lifting precast components also requires devices for attaching slings or cables (or
both), usually consisting of steel embedded in the finished surface of the concrete.
Improper removal and treatment of these embedments once the precast component is
set in the field can result in localized spalls due to steel corrosion.
4.3.2.1.5
Construction
Attention to good practices during construction is crucial to the long-term service
life of reinforced concrete. Well-qualified and well-trained workers and well-executed
workmanship increase productivity, reduce material waste, and provide expected ser-
vice life. Proper use of adequate equipment provides better workability by increasing
efficiency, and a well-planned construction schedule reduces the overall cost of the
project by providing set times for equipment rental and reducing downtime. Proper
use of test methods is needed to ensure that quality concrete is achieved. These prac-
tices, which are listed in the fault tree in Figure 4.15, include workmanship related to
the connectivity of the precast components.
Connection of precast components is performed in several ways. Match-cast com-
ponents are typically joined with epoxy and prestressed across the joint. Improper
epoxy material for the temperature of application, inadequate epoxy set time, and
inconsistent, nonuniform application can cause spalling of the joint as the prestressing
compresses the joint across a nonuniform contact surface.
Components that are not match cast are typically detailed with open joints to
be filled with concrete or grout. Improper surface preparation and incomplete filling
of these joints can result in early breakdown of the joint filler material, resulting in
cracks, leakage, reinforcement corrosion, and a reduction in load distribution charac-
teristics that can be detrimental to the carrying capacity of the bridge.
4.4
Individual Strategies to Mitigate Factors Affecting Service Life
Section 4.3 defines numerous factors affecting the service life of bridge decks. This sec-
tion provides individual strategies to mitigate those factors. Table 4.2 summarizes the
areas in which strategies are provided.

212

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 4.2. Mitigation Categories
Section Mitigation Category
4.4.1 Strategies to mitigate load-induced effects
4.4.2 Strategies to mitigate system-dependent loads
4.4.3 Strategies to mitigate natural or man-made environment deterioration
4.4.4 Strategies to improve production and operations

4.4.1
Strategies to Mitigate Load-Induced Effects
This section addresses concrete bridge decks. Load-induced effects are created from
the traffic using the bridge and from system-dependent framing restraints. Strategies
for mitigating deterioration from these effects are provided in this section.
4.4.1.1
Strategies to Mitigate Traffic-Induced Loads
A complete understanding of the characteristics of the traffic on the structure is re-
quired to define the strategies required for enhancing the service life of a bridge deck.
These characteristics include vehicle configuration, such as axle and wheel spacing
and individual wheel weights; type of wheel or tire; potential for overloads; type of
suspension system; traffic volumes and frequency of truck and overload application;
and vehicle location on the deck.
In order to establish criteria to adequately address fatigue response, overload,
wear, and abrasion, these characteristics must be understood. Table 4.3 identifies the
strategies for these service life issues.
Bridge-deck systems can be adequately designed for fatigue by considering indi-
vidual wheel loads, dynamic impact effects, and the frequency of load application
developed from the volume of truck traffic to which the bridge deck will be subjected.
The fatigue design of reinforcing steel within the concrete deck is adequately addressed
by the threshold design methods provided in the LRFD specifications.

Table 4.3. Mitigating Strategies for Traffic-Induced Loads


Service Life Issue Mitigating Strategy Advantage Disadvantage
Fatigue Design per LRFD specifications Minimizes the possibility of May increase the area of
reinforcement failure steel
Overload Increase deck thickness Minimizes cracking Adds weight to bridge
structure, increases cost
Wear and abrasion Implement concrete mix design See Chapter 3, Materials See Chapter 3, Materials
strategies
Implement membranes and Protects surface from direct Requires rehabilitation
overlays contact with tires every 10 to 20 years

213

Chapter 4. BRIDGE DECKS


Table 4.4. Mitigating Strategies for System-Dependent Loads
Service Life Issue Mitigating Strategy Advantage Disadvantage
Differential Use low-modulus concrete mix Allows additional strain to Typically lower in strength
shrinkage designed for composite decks be accommodated up to and may be subject to wear
cracking stress and abrasion
Use high-creep concrete mix Reduces locked-in stresses Uncommon mix design.
designed for composite decks Difficult to assess stress
relief.
Develop composite action after Allows slippage between Very limited knowledge on
concrete has hardened deck and supporting available systems capable
members, minimizing of developing composite
locked-in stresses action after hardening of
bridge deck
Use precast deck panels Allows slippage between Introduces numerous
deck and supporting construction joints
members, minimizing
locked-in stresses
Thermal restraint Develop an accurate system Identifies design criteria for Analysis is time consuming
model for analysis purposes establishing stresses
System-framing Develop an accurate system Identifies design criteria for Analysis is time consuming
restraint model for analysis purposes establishing stresses

Bridge decks can also be designed for overload conditions with adequate determi-
nation of the potential for overload and the frequency of its application. Additional
strength requirements for overloads can be addressed by increasing the thickness of
the deck.
Additional information on wear and abrasion can be found in Chapter 3. Mem-
branes and overlays can be used to separate the wheel contact surface from the deck
surface.
4.4.2
Strategies to Mitigate System-Dependent Loads
Bridge-deck performance can be enhanced by the proper selection of a system to ac-
commodate bridge movements, whether they are caused by differential shrinkage or
from system-framing restraints of movements from thermal expansion and contrac-
tion or creep and shrinkage. Table 4.4 identifies the strategies for these service life
issues; these strategies are expanded below.
4.4.2.1
Differential Shrinkage
Several enhancements are viable for addressing the restraint forces at the interface of
the bridge deck and supporting beam or stringer superstructure elements. Potential
enhancements include the following:

214

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


• Using low-modulus concrete mix design to allow the deck to accommodate the
shrinkage strain with less tension force, which can reduce cracking. Refer to Chap-
ter 3 for additional discussion of low-modulus concrete mix designs.
• Using a high-creep concrete mix in the supporting superstructure element, which
continues to reduce the locked-in tension force in the bridge. Refer to Chapter 3
for additional discussion of high-creep concrete mix designs.
• Using delayed composite action systems in which the interface of the bridge deck
and the supporting beam or stringer superstructure elements is not made compos-
ite until a majority of the deck shrinkage has occurred.

4.4.2.2
System-Framing Restraint
Superstructure and substructure systems must be designed to provide either movement
or restraint of the structure, with proper consideration of internally induced forces.
For additional information on the proper system selection, see Chapter 2.
4.4.2.2.1
Fully Integral Deck Systems
Eliminating expansion joints at abutments and over piers can enhance bridge-deck
performance. This bridge system, commonly referred to as a jointless bridge, is ad-
dressed in Chapter 8.
4.4.2.2.2
Semi-Integral Deck Systems
A significant number of states use a semi-integral approach to bridge-deck systems.
This system provides expansion joints at the beginning and end bridge abutments and
no joints (or limited joints) in the remainder of the bridge. Bridge-deck performance
is improved by eliminating joints. Separating the bridge deck from the substructure at
the abutment locations reduces the tensile forces that could otherwise be generated in
the bridge deck during deck contraction as a result of traffic and thermal loads.
4.4.3
Strategies to Mitigate Natural or Man-Made Environment
Deterioration
Proper studies for identifying environmental exposures detrimental to bridge-deck per-
formance should be performed. Understanding the causes of deterioration leads to
proper consideration during design.
Tables 4.5, 4.6, and 4.7 describe the strategies developed for natural and man-
made environment deterioration.
Tables 4.4 through 4.7 present strategies for addressing the various factors affect-
ing service life presented in Section 4.3. Proper incorporation of design features and
materials is important for enhancing the service life of CIP and precast bridge decks.
Likewise, there are numerous protection strategies for enhancing the service life of
concrete bridge decks. These include providing adequate concrete cover, proper con-
crete mix design, proper reinforcement selection and protection, proper drainage,

215

Chapter 4. BRIDGE DECKS


Table 4.5. Mitigating Strategies for Thermal Climate Environment Deterioration
Service Life Issue Mitigating Strategy Advantage Disadvantage
Thermal deicing Use impermeable concrete. Increases passivity around High initial shrinkage,
salts reinforcement. which can result in
Refer to Chapter 6, cracking.
Corrosion Prevention of
Steel Bridges.
Use corrosion-resistant Eliminates deck spalls, High cost.
reinforcement. delaminations, and cracking Limited availability.
from reinforcement Some performance issues
corrosion. as noted in Chapter 3,
Materials.
Use waterproof membranes or Minimizes intrusion of Requires rehabilitation to
overlays. dissolved chlorides into replace riding surface every
deck. 5 to 20 years.
Easily rehabilitated.
Use external protection methods, Reduces corrosion. High cost.
such as cathodic protection. Refer to Chapter 5, Requires extensive
Corrosion of Steel in maintenance and anode
Reinforced Concrete and battery replacement.
Bridges. Could have limited
effectiveness.
Use effective drainage to keep Minimizes intrusion of Requires maintenance of
surfaces dry and minimize dissolved chlorides into drainage.
ponding. deck.
Use periodic pressure washing to Minimizes intrusion of Requires dedicated
remove contaminants. dissolved chlorides into the maintenance staff and
deck. appropriate budget.
Low cost.
Use nonchloride-based deicing Eliminates corrosion from High cost.
solution. chlorides.
Freeze–thaw Refer to Chapter 3, Materials, for strategies relating to freeze–thaw deterioration.

application of prestressing, and the use of external protection systems. These enhance-
ment strategies are described in the following sections.
4.4.3.1
Concrete Cover
Concrete cover is recognized as an effective method for protecting steel from corro-
sion. A minimum top concrete cover of 2 in. is required by the LRFD specifications.
Generally, 2.5 in. of cover is used to allow for 0.5 in. of wear over the life of the deck.
Cracking of the deck, however, can allow chlorides to quickly penetrate to the level of
the reinforcement, initiating the corrosion process.

216

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 4.6. Mitigating Strategies for Coastal Climate Environment Deterioration:
Natural or Human-Caused
Service Life Issue Mitigating Strategy Advantage Disadvantage
Salt spray Use impermeable concrete. Increases passivity around Higher cost.
reinforcement. Not effective at transverse
Refer to Chapter 3, cracking locations.
Materials.
Use corrosion-resistant Eliminates deck spalls, High cost.
reinforcement. delaminations, and cracking Limited availability.
from reinforcement Some performance issues
corrosion. as noted in Chapter 3,
Refer to Chapter 3, Materials.
Materials.
Use waterproof membranes or Minimizes intrusion of Requires rehabilitation
overlays on travel surfaces of dissolved chlorides into the every 5 to 20 years.
bridge deck. deck.
Use external protection methods, Reduces corrosion. High cost.
such as cathodic protection. Refer to Chapter 5, Requires extensive
Corrosion of Steel in maintenance and anode
Reinforced Concrete and battery replacement.
Bridges.
Use sealers on nontravel surfaces Minimizes intrusion of Requires rehabilitation
of bridge deck. dissolved chlorides into every 5 to 10 years.
deck.
Use corrosion-resistant stay-in- Minimizes intrusion of Difficult to inspect.
place forms on bottom of bridge dissolved chlorides into
deck. deck.
Use effective drainage to keep Minimizes intrusion of Requires maintenance
surface dry. dissolved chlorides into of drainage and periodic
deck. cleaning.
Use periodic pressure washing to Minimizes intrusion of Requires dedicated
remove contaminants. dissolved chlorides into maintenance staff and
deck. appropriate budget.
Humidity Use materials that are not Refer to Chapter 3, Materials.
sensitive to moisture content.

4.4.3.2
Concrete Mix Design
The impermeability of concrete enhances the protection of bridge-deck reinforcement.
Concrete mix design is addressed earlier in this section and is further discussed in
Chapter 3.
Enhanced service life of bridge decks can be achieved by implementing a mix
design to obtain desirable properties for mitigating the potential for deficiencies.
The desirable properties for enhanced performance include crack resistance through
improved tension capacity, low permeability to delay chloride intrusion, low modulus

217

Chapter 4. BRIDGE DECKS


Table 4.7. Mitigating Strategies for ENVIRONMENTAL DETERIORATION BY Chemical Climate,
Reactive Ingredient, and Fire: Natural or Human-Caused
Service Life Issue Mitigating Strategy Advantage Disadvantage
Corrosion-inducing Use materials and mix designs Refer to Chapter 3, Materials.
chemicals and that are not sensitive to chemical
sulfate attack.
Reactive Use materials and mix designs Refer to Chapter 3, Materials.
Ingredients: ASR that are not sensitive to
and ACR aggregate reactivity.
Extreme events: Incorporate fire rating. The height of the concrete Can result in high cost.
fire structure and concrete Increases weight with
cover can protect increased fire rating.
reinforcement from
softening significantly,
lessening collapse risk.
Provide fire-protective coatings. Increases fire rating, Aesthetically unappealing.
reduces collapse risk. Subject to deterioration in
an exposed environment.

of elasticity to allow deck strain with lower tension force, and high creep to allow
reduced locked-in stresses over time.
Bridge-deck concrete can also be enhanced by incorporating proper materials and
admixtures:

• Proper cement selection. In areas where sulfate attack may be a concern, Type II or
Type V cements may be used to provide added resistance to its detrimental effects.
Heat of hydration, which adds to the differential shrinkage strain, may be reduced
by using a Type IV cement.
• Proper aggregate selection. Some readily available aggregates may be reactive to
the internal concrete chemistry and be more susceptible to ASR and ACR. Ad-
mixtures and/or proper blending with nonreactive aggregate can minimize these
effects. Using high-quality aggregates also enhances abrasion resistance.
• Proper air entrainment agent. Air entrainment increases workability in the field
and also enhances concrete performance when concrete is subjected to freeze–
thaw cycles.
• Proper admixture selection. Numerous admixtures are available to enhance the
properties of concrete and improve concrete durability substantially. In particular,
admixtures for concrete decks can inhibit corrosion, improve workability (pro-
viding a proper durable concrete finish), delay initial set to provide time for con-
crete placement and finishing, and reduce water requirements to improve concrete
strength and density.

Some of these enhancements may conflict, and therefore a mix design must incor-
porate desired features with the understanding that all enhancement strategies cannot
be achieved. For more information on mix designs, refer to Chapter 3.
218

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


4.4.3.3
Reinforcement Selection
The selection of bridge-deck reinforcement can enhance the service life of the bridge
deck and increase resistance from corrosion and section loss, particularly in marine en-
vironments or in areas where deicing salts are used. Enhanced reinforcing steel includes

• Corrosion-resistant reinforcing, such as FRP, stainless steel, and titanium bars;


• Reinforcement protection systems, such as epoxy coating and galvanizing; and
• Multiple posttensioning protection strategies, such as those defined by FHWA
in the Post-Tensioning Tendon Installation and Grouting Manual (Corven and
Moreton 2004).

Refer to Chapter 3 for more information on reinforcing materials.


4.4.3.4
Bridge-Deck Drainage
Eliminating prolonged exposure to moisture and allowing the bridge deck to be main-
tained in a dry condition can enhance the performance of bridge decks. Proper deck
slopes, both transversely and longitudinally, should be provided to channel ­water to
appropriate collection points. Construction joints at these collection points, as shown
on the left in Figure 4.17, should be eliminated or moved away from the collection
point, as shown on the right in Figure 4.17, to minimize contaminant intrusion, which
can lead to the deterioration of reinforcement. Bridge drains, drain grates, and piping
should be sized appropriately to be self-flushing, minimizing maintenance requirements.
4.4.3.5
Application of Compression to Relieve Tension
The elimination of tension in a CIP bridge deck enhances the performance of the
bridge deck by eliminating or significantly reducing deck cracking. Compression is
typically introduced by posttensioning the concrete; however, the durability of the

Figure 4.17.  Construction joint at barrier–bridge deck interface.

219

Chapter 4. BRIDGE DECKS


posttensioning is contingent on the proper incorporation of durability enhancements,
as well. Refer to Chapter 3 for additional information on the durability concerns of
posttensioned systems.
Compression can also be introduced into the bridge deck of continuous girder
bridge systems through the use of a self-stressing bridge-deck system developed by
SHRP 2 Project R19A (da Silva 2011). The compression in this system, typically appli-
cable to a two-span steel girder bridge unit, is introduced by casting the bridge deck
with the intermediate support higher than required. After the deck has cured, the inter-
mediate support is lowered to its final position, thereby introducing compression into
the area of the deck usually subjected to tension forces from negative moments. This
system requires additional research to establish a history of satisfactory performance.
Refer to Appendix A for more information on this system.
4.4.3.6
Membranes
Membranes are placed on top of the concrete and are protected by an asphalt layer
that also functions as a riding surface. Effective waterproofing enhances the service life
of the membrane system, and in turn, the bridge deck.
4.4.3.7
Overlays
The purpose of concrete overlays is to create a low-permeability protective layer over
the conventional concrete on bridge decks. An overlay serves as a barrier to chloride
ions and thus increases the time required for the concentration of the ions at the level of
the reinforcement to reach the threshold for corrosion. Low-permeability overlays also
decrease water penetration into a structure, allowing it to dry out, which reduces chlo-
ride ion mobility. Overlays can be applied to new decks or as a rehabilitation method
to existing decks. However, overlays are not as effective when applied to existing decks
because if chloride ions are already present in the deck when the overlay is placed, then
the only protection that the overlay can offer is a decrease in moisture infiltration.
The most common type of overlay has been a low-slump dense concrete overlay,
which has been effective in extending the service life of damaged bridge decks in some
states. Special equipment is required to handle the very stiff concrete; special atten-
tion to placement and consolidation are needed; and the overlays are prone to rapid
loss of moisture, necessitating extra care in curing. Recently, silica–fume concrete and
latex-modified concrete overlays have been successfully used in extending the service
life of contaminated structures; these concretes have improved workability compared
with low-slump concrete overlays. Polymer concrete overlays are also available, gen-
erally as a temporary repair method on damaged bridge decks. Refer to Chapter 3,
Materials, for more information on overlays.
­

4.4.3.8
Sealers
Sealers are expected to minimize the intrusion of aggressive solutions into concrete. The
primary purpose of sealers is to prevent water and chloride ions from penetrating the con-
crete and thereby reduce the corrosion of reinforcement or the deterioration of concrete.

220

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


An important property of a sealer is its vapor transmission characteristics. Mois-
ture within the concrete needs to pass through the sealer and escape in order to prevent
high vapor pressures from building up in the concrete during drying periods, which
could cause the sealer to blister and peel.
Sealers can be either pore blockers or water repellents. Pore-blocker sealers work
by forming a microscopically thin (up to 2 mm) impermeable layer on the concrete
surface. Most pore-blocker sealers are not appropriate for use on bridge decks because
they do not offer good skid resistance and do not hold up under traffic wear. Water-
repellent sealers, in contrast, work by penetrating slightly into the concrete and act-
ing as hydrophobic agents. Hydrophobic sealers for bridge decks include silanes and
siloxanes.
Sealers can protect all of the exposed concrete surfaces of the structure, including
bridge decks, superstructure members, substructure members, and deck undersides.
Proper surface preparation and consideration of application rates are key factors to
be considered during installation of the sealer. Abrasion, sunlight, and the environ-
ment can affect the effective life of sealers, and resealing of the bridge deck could be
expected every 2 to 5 years.
4.4.3.9
External Protection Systems
Several external protection systems are available to enhance the service life of CIP
bridge decks, including electrochemical chloride extraction and cathodic protection
systems.

• Electrochemical chloride extraction involves the application of a direct current to


an existing bridge deck for a 4- to 8-week period. Electrochemical chloride extrac-
tion extracts chlorides from concrete and enhances the passivated zone around the
reinforcement. An anode is provided by a titanium mesh or steel anode, which is
temporarily placed on the concrete cover. This removes an average of 40% to 90%
of the initial free chlorides. The extraction depends on the depth and location of
the reinforcement. The pH of the concrete is increased, and the remaining chloride
contents are typically below threshold levels near the reinforcement and increase
with distance from the rebar. Prior application of this technology has resulted in
more than 20 years of a passive noncorroding condition. Typically applied to ex-
isting bridge decks, this system also has the potential to pretreat a new concrete
bridge deck, enhancing the passivated zone around the reinforcement.
• Cathodic protection is used to prevent corrosion from initiating, thereby reducing
the concentration of chloride ions. This method is mainly used to prevent further
corrosion after repair of damaged structures, and it has recently been used to
prevent corrosion from initiating in new structures. The most common impressed-
current anode uses a titanium mesh anode in conjunction with a concrete overlay
or titanium ribbon. The current must be uniformly distributed, and the system
must be regularly monitored and inspected to ensure that polarization is in the
desired range.

221

Chapter 4. BRIDGE DECKS


4.4.4
Strategies to Improve Production and Operations
Improving the performance of bridge decks relies on following proper methods and
procedures during construction. The strategies to improve production and operations
are included in Table 4.8.
The methods and procedures used to mitigate production and operation defects
include proper construction joint selection, proper choice of formwork and bar
supports, specification of enhanced placement procedures, and other maintenance
considerations.
4.4.4.1
Construction Joint Selection
4.4.4.1.1
Sequence of Deck Casting
Bridge-deck performance can be enhanced through proper selection of the deck-­casting
sequence. Concrete mix designs that delay initial set until the weight and vibration of
casting and finishing machinery have passed reduce concrete cracking potential from
movement of the supporting structure below.

Table 4.8.  Strategies for Mitigating Production and Operation Defects


Service Life Issue Mitigating Strategy Advantage Disadvantage
Design philosophy Use empirical design. Uses less reinforcement. Limited application for
future bridge widening.
Expansion joints Eliminate expansion joints. See Chapter 8, Jointless Bridges.
Construction joints Minimize construction joints. Reduces corrosion Requires larger casting
potential. volumes.
Minimal cost.
Locate joint away from areas of Minimizes saturation at Could require additional
ponding. joint location. Reduces formwork.
corrosion potential.
Minimal cost.
Phased Use minimum-width closure pour Minimizes cracking. High cost.
construction with UHPC. Difficult to finish.
Tight tolerances.
Use wider closure pour, Minimizes water intrusion Slower construction, more
conventional concrete with and closure pour cracks. cracking from differential
waterproofing, and overlay. Shortens construction time. shrinkage and restraint.
Accommodates residual
differential deflection.
Allow sufficient time for creep Minimizes differential More construction time and
deformation to stabilize before elevation between adjacent higher cost.
casting closure pour. bridge sections.

Note: UHPC = ultrahigh-performance concrete.

222

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


If the bridge deck cannot be cast in one operation, the appropriate location of con-
struction joints and the proper sequencing of the deck pour can improve performance.
Performance is enhanced by delaying the casting of those sections of the bridge suscep-
tible to tension from adjacent casting operations. This is typical of casting sequences
for continuous steel girders in which the positive moment areas are cast first, f­ ollowed
by the negative moment areas. The casting sequence should be specified by the designer
and noted on the design plans.
4.4.4.1.2
Adjacent Members
Proper sealing (making the construction joint waterproof and preventing water intru-
sion between the bridge-deck elements) enhances construction joints between adja-
cent CIP members, such as transverse construction joints in CIP segmental structures.
Water­proofing strategies are addressed in the FHWA Post-Tensioning Tendon Installa-
tion and Grouting Manual (Corven and Moreton 2004) and in Chapter 3, Materials.
Numerous details have been used by many states for these types of structures.
Many details transfer only the vertical shear across the construction joint, causing
the adjacent members to act together vertically, but allowing the bridge system to flex
at these joints. This design at times has resulted in a breakdown of the material in
the field construction joint. In general, details in which the connection between these
adjacent members is designed to transmit the applied vertical shears and transverse
moments have performed significantly better over time.
Enhanced service life of filled construction joints between adjacent CIP mem-
bers can be achieved through the use of ultrahigh-performance concrete. UHPC,
which is described in Chapter 3, provides high strength and stiffness with negligible
­permeability and improved durability. It is expected to reduce maintenance require-
ments and extend service life. When UHPC is used in bridge-deck construction joints,
consideration should be given to two issues: grinding the surface due to the higher
strength of UHPC and dissimilarities in color.
4.4.4.1.3
Staged or Phased Construction
Construction joints at the interface between adjacent phases of construction, such as in
bridge widening, can be enhanced by a combination of the following:

• Properly locating the construction joint away from areas in which water and
water­borne contaminants can collect, such as at the construction joint between
traffic railing barriers and the bridge deck, as shown in Figure 4.17;
• Ensuring proper reinforcement through the construction joint to control cracking;
• Applying epoxy to bond the surfaces together to prevent water intrusion, such as
at the construction joint between traffic railing barriers and the bridge deck, as
shown in Figure 4.17;
• Limiting live-load influence near the joint to prevent vibration and joint flexing
until concrete has attained the appropriate resistance to tension;

223

Chapter 4. BRIDGE DECKS


• Using admixtures in the concrete design to increase the time to initial set until
all construction activities affecting the deflection of adjacent supporting members
have been completed;
• Delaying the casting of the deck between adjacent phases of construction by add-
ing a closure pour to be completed after the casting of the deck on the supporting
members for both phases; and
• Addressing differential shrinkage between the phase-constructed closure pours and
the adjacent completed bridge phases using procedures identified in Section 4.4.2.1.

4.4.4.2
Formwork
Formwork for bridge decks can be either removable or stay-in-place. Removable forms
can be made of various materials, but usually consist of some type of plywood. Stay-
in-place forms can consist of steel panels supporting a full-depth CIP deck or precast
panels that can be either composite or noncomposite with the CIP deck above.
The use of improved formwork technologies can improve the quality of the con-
crete surface, increasing its impermeability. Controlled permeability formwork is a
special class of lined formwork that increases the strength and durability of the con-
crete surface (Malone 1999). The formwork liner allows trapped air and excess water
to pass through during concrete placement and consolidation. The result is a surface
free of voids (bug holes), which increases the strength and durability of the surface.
CIP concrete with metal stay-in-place forms has gained popularity nationwide.
However, several states are reluctant to adopt it because the underside cannot be easily
inspected. The steel forms are susceptible to corrosion from salt spray and should be
limited to areas where this type of corrosion is not an issue.
4.4.4.3
Bar Supports
Bar supports typically rest on concrete surfaces that are exposed to natural and man-
made environmental hazards. Enhanced service life can be achieved through the use of
noncorroding materials or noncorroding coatings on chair legs.
4.4.4.4
Bridge-Deck Construction Procedures
Concrete placement procedures are fairly well established. Improvements in concrete
mix design, placement, and curing specifications appear to have adequately addressed
many of these service life issues, except for cracking and corrosion of reinforcement.
Service life is enhanced through proper planning of the bridge-deck–casting pro-
cess. Concrete placement is enhanced by ensuring that sufficient vibration equipment
is used; that vibration is effective at areas of congestion, such as at expansion joints;
and that concrete is not dropped from excessive heights.
Curing is among the most important factors in developing durable deck concrete
(Darwin et al. 2010) and is essential for the continuation of hydration reactions and
the control of cracking due to volumetric changes. Curing of concrete is enhanced
by ensuring that moisture is not lost, which can be accomplished by using curing

224

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


compounds and maintaining a wet curing environment, such as under a moist burlap
covering, for 7 to 10 days. Performance enhancements are maximized with longer wet
cure periods. The wet curing process also helps maintain thermal control of the bridge
deck in its critical early stages of hydration. Refer to Chapter 3 on materials for more
information on concrete curing.
4.4.4.5
Maintenance Considerations for Existing Bridge Decks
To extend concrete bridge service life in existing structures, preventive maintenance
should be emphasized in a maintenance plan, and proper repairs should be performed
before extensive damage occurs and costly rehabilitation is required. The scope of re-
pair or rehabilitation work can vary significantly, from sealing cracks to applying over-
lays to replacing large components such as bridge decks. Preventive maintenance may
also include tasks as simple as washing the structure to eliminate chloride buildup.
4.5
Overall Strategies for Enhanced Bridge-Deck Service Life
This section provides tools for selecting the most appropriate individual strategy to
achieve the desired bridge-deck service life. It provides a template to the designer for
selecting the optimum solution available and quantitatively predicting the service life,
when applicable.
The flowchart in Figure 4.18 shows the bridge-deck system component selection
process. The flowchart in Figure 4.19 shows the service life factor mitigation process.
An identifying number in each step designates each activity within the flowchart.
These identifying numbers are used in the following discussion of the various elements
of the flowcharts.
Steps 1a and 1b. Development of Design Criteria
The activities in Steps 1a and 1b of Figure 4.18 identify the project’s local operational
site requirements and the local factors affecting service life. They are crucial to the
development of design criteria used to identify and evaluate bridge-deck alternatives.
Examples of the information to be gathered during these activities are provided in
Table 4.9.
Step 2. Identification of Feasible Bridge-Deck Systems
The next step in the process is the selection of various bridge-deck–type alternatives
that can meet the project requirements and the design provisions stated in the LRFD
specifications. For example, CIP and precast deck systems could be identified as poten-
tial alternatives. Table 4.1 in Section 4.2 provides the advantages and disadvantages
of these various decks. The selection of potential deck alternatives should consider the
requirements of other bridge subsystems, components, and elements and their interac-
tion. Refer to Chapter 2 for more information on overall bridge system requirements.
The most prominently used bridge-deck system is the CIP or precast concrete
bridge deck. This bridge deck can either be self-supporting as part of the overall super-
structure system, (e.g., voided slabs and segmental structures) or supported on beam
or stringer superstructure elements. All these concrete deck systems are subject to
cracking, which typically results in reduced life from corrosion.

225

Chapter 4. BRIDGE DECKS


Bridge-Deck System Component
Selection Process

1a. Identify local operational and 1b. Identify local factors affecting
site requirements. service life.

2. Identify feasible deck alternatives satisfying design


provisions of LRFD specifications, operational, site, and
bridge system requirements.

3. For each alternative, identify factors affecting service life


by following fault tree.

4.
Go to
A

B
8a. Go to the
next alternative.

5. Is deck
service life No 5a. Identify rehabilitation
greater than or
or replacement
equal to the
requirements.
system
TDSL?
Yes

6. Identify maintenance requirements.

7. Develop life-cycle costs.

Yes
8. Additional deck
alternative?

No

9. Compare alternatives and select deck system.

Figure 4.18.  Bridge-deck system component selection process. For steps from A to B,
see Figure 4.19.(TDSL means target design service life.)
226

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


4.
A

1A. Identify individual factors affecting service life


considering each branch of fault tree.

Yes 2A.a. Identify consequences and


3A.a. Go to the
2A. Does factor apply? determine appropriate strategies
next factor.
for avoidance or mitigation.

No

No
3A. Are all factors 2A.b. Modify bridge deck
considered? configuration.

Yes

4A. Modified bridge deck configuration for deck alternative under


consideration.

Go To
B

Figure 4.19.  Mitigation of factors affecting service life process.


Figure 1.18. Flowchart to identify factors affecting service life (Figure 4.19).

Selection of the overall concrete bridge-deck system may be affected by the follow-
ing factors:

• Need for accelerated construction to shorten overall user impacts;


• Maintenance of traffic requirements that may dictate construction staging;
• Commitments made during the NEPA process, such as acceptable noise levels,
access limitations, or environmental and biological limitations that may dictate a
precast system;
• Availability of special mix designs to provide a more durable concrete;
• Availability and construction expertise to incorporate prestressing to compress the
concrete, minimizing or eliminating tension in the concrete;

227

Chapter 4. BRIDGE DECKS


Table 4.9. Identify Bridge-Deck Demands: Local Operational, Site, and Service Life Requirements
Demand Examples
Identify owner requirements Legal and permit loads
Commitments made during the NEPA process
Noise
Access limitations
Environmental and biological
Other design directives
Acceptable risk
Bridge-deck target design service life
Contingency planning for future expansion
Identify traffic load demands Potential for overloads
Construction loads
Impact considerations from suspension systems
Frequency of load application
Tire type for wear and abrasion
Identify system-dependent demands Coordinate with bridge system selection
Differential shrinkage effects
Boundary conditions for system-framing restraint
Thermal exposure
Identify natural and human-made Deicing requirements
hazard demands Freeze–thaw potential
Local aggregate reactivity
Susceptibility to fire
Susceptibility to collision
Chloride concentrations (natural)
Chloride concentrations (applied)
Sulfate concentrations
Humidity levels
Identify other general demands Traffic maintenance requirements
Construction phasing requirements
Need for accelerated construction
Drainage and storm water requirements
Identify local construction practice expertise

Note: NEPA = National Environmental Policy Act.

• Availability of and the construction expertise to incorporate internal and/or ancil-


lary protective systems for the bridge deck; and
• Ability to provide an alignment and/or a bridge drainage system to prevent pond-
ing of water and soluble pollutants on the bridge deck.
Step 3. Identification of Factors Affecting Service Life
After feasible bridge-deck alternatives are selected and designed based on applicable
provisions in the LRFD specifications (Step 2 in Figure 4.18), the next step is to use
fault tree analysis to identify the factors affecting service life for each feasible deck
alternative (Step 3 in Figure 4.18). The fault tree is described in Section 4.3.

228

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


After each feasible bridge-deck alternative has undergone fault tree analysis to
identify all possible factors that may affect service life, the procedure continues with
Process A, which is developed in Figure 4.19 and described next.
Process A. Refinement of Alternatives: Mitigation of Factors
Affecting Service Life
Process A (Figure 4.19) develops mitigating design features for each factor affecting
service life identified under Step 3. An alternative selection process would be to evalu-
ate these strategies by examining material selection and protection strategies, construc-
tion practice specification requirements, and maintenance requirements. A summary
example of this alternative process is provided in Table 4.10. Further explanation of
some for these elements is provided

Table 4.10.  Alternative Bridge-Deck System Development Process


Demand Examples
Determine function Accommodate span requirements and constraints
Strength design Accommodate curvature and skew requirements
Accommodate proposed deck joint layout and its effect on system restraint
Identify deck deterioration modes Concrete cracking
Reinforcement corrosion
Wear and abrasion
Aggregate reactivity
Freeze–thaw
Assess risk and deterioration Loss of required strength
consequences Safety concerns from pot holing (large concrete spalls)
Loss of skid resistance
Develop mitigation strategies (Section Traffic-induced loads
4.4) System-dependent loads
Natural, man-made, or environmental deterioration
Develop design requirements Concrete cover
Concrete mix design
Reinforcement selection
Bridge-deck drainage
Introduction of compression to relieve tension
Application of membrane, overlay, and sealers
External protection systems
Identify strategies to improve Sequence of deck casting
construction Construction joint location and detailing
Construction phasing and staging details
Formwork selection
Bar support selection
Specification of construction procedures
Quantification of maintenance requirements

229

Chapter 4. BRIDGE DECKS


Process A includes the following steps:
Step 1A
Identify the individual factors affecting service life by considering each branch of the
fault tree defined in Section 4.3.
Step 2A
For each identified factor, use the design criteria to evaluate whether the factor has an
effect on the service life of the bridge deck. If the factor identified in Step 2A has an
effect on the service life of the bridge deck, proceed to Steps 2A.a and 2A.b.
Step 2A.a
Identify the consequences of the factor and determine appropriate strategies to miti-
gate or avoid the effects of deterioration. Refer to Section 4.4 for mitigation and avoid-
ance strategies. There may be more than one strategy alternative to consider for each
factor. For example, bridge decks subjected to wear and abrasion can be mitigated
through concrete mix design alternatives and/or the use of waterproofing membranes
and overlays. Note that the applicable factors identified in the circle symbols are basic
factors that demand development of strategies to mitigate them.
Step 2A.b
Modify the bridge-deck alternative under consideration as needed to address the incor-
poration of the chosen strategy or strategies.
Steps 3A and 3A.a
Continue with the evaluation process until every factor has been considered that may
affect the service life for each feasible bridge alternative. This iterative process contin-
ues (Step 3A.a) to ensure that all factors are considered.
Step 4A
Step 4A, the last step in Process A, is to finalize the modifications to the bridge-deck
configuration for the bridge-deck alternative under consideration. Verify that strength
and service performance have not been affected.
A sample series of strategies for a bridge with a CIP concrete deck supported
on stringers is shown in the next three tables. Table 4.11 provides strategies for CIP
bridge-deck systems, and Table 4.12 provides strategies for precast bridge-deck sys-
tems. The identified systems can achieve long service life with the proper inspection and
maintenance as indicated. Table 4.13 provides strategies for rehabilitation of existing
bridge decks, which may be used in the case of a bridge widening. Note that deteriora-
tion modes addressed through proper design, such as fatigue and overload potential,
are not addressed with a strategy in these tables. Other material deterioration modes,
such as freeze–thaw and sulfate attack, are addressed in Chapter 3.
The selection of appropriate material and protection strategies is fairly consistent
for all of the concrete bridge-deck systems included in this chapter, with only minor
differences in the durability performance selection process between CIP and precast
deck systems. The appropriate strategies for each potential deterioration mode must
be compared for conflicts in order to establish the overall strategy to be deployed for
a specific project. For example, a bridge deck with the potential for deterioration from

230

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 4.11. Bridge-Deck Selection Strategies: CIP Systems
Potential Material Selection and Life-Cycle Costs
Deterioration Protective Measures
Mode Selection Maintenance Mode Initial Long-Term
Differential shrinkage Low-modulus concrete None Low Low
and thermal restraint Proper bearing design
Wear and abrasion Overlay and/or membrane Continual overlay replacement Medium Medium
5 to 20 years
Wear and abrasion High concrete strength None Medium Low
Hard aggregates
Wear and abrasion Sacrificial thickness and overlay Continual overlay replacement Low Medium
5 to 20 years
Reinforcement Overlay and/or membrane Continual overlay replacement Medium Medium
corrosion 5 to 20 years
Reinforcement Corrosion-resistant rebar None High Low
corrosion
Reinforcement External protection systems Continual inspection and High High
corrosion system maintenance
ASR Refer to Chapter 3, Materials na na na
ASR Blended aggregates, proper None Medium Low
drainage
ASR Blended aggregates, Continual overlay replacement Medium Medium
waterproof membrane, proper 5 to 20 years
drainage
ACR Refer to Chapter 3, Materials na na na
Note: na = not applicable.

wear and abrasion and differential shrinkage cannot easily use concrete with both
high strength and a low modulus. In this case, the wear and abrasion strategy using an
overlay or membrane is more appropriate.

Process B. Identification of Rehabilitation, Maintenance, and Life-


Cycle Costs for System Alternatives and Final System Selection
Step 5. Check Service Life
At this step in the process, the service life for the bridge-deck alternative is determined
either through the use of deterioration models or through empirical evidence based on
past performance. Refer to Chapter 1 for methods available for predicting service life.
If the bridge-deck service life does not equal or exceed the bridge system design service
life, then rehabilitation and/or replacement requirements (Step 5a) should be added to
increase the longevity of the bridge.
Step 6. Identify Maintenance Requirements
The next step is to identify maintenance requirements for the proposed alternative and
their associated costs. Table 4.14 identifies example maintenance issues. The cost for
developing a maintenance plan should also be included in the bridge Owner’s Manual.
Refer to Chapter 1 for a detailed description of the bridge Owner’s Manual.
231

Chapter 4. BRIDGE DECKS


Table 4.12. Bridge-Deck Selection Strategies: Precast Systems
Potential Material Selection and Life-Cycle Costs
Deterioration Protective Measures
Mode Selection Maintenance Mode Initial Long-Term
Differential shrinkage Low-modulus concrete, proper None Low Low
and thermal restraint bearing design
Wear and abrasion Overlay and/or membrane Continual overlay replacement Medium Medium
5 to 20 years
Wear and abrasion High concrete strength, hard None Low Low
aggregates
Wear and abrasion Sacrificial thickness and overlay Continual overlay replacement Low Medium
5 to 20 years
Reinforcement Overlay and/or membrane Continual overlay replacement Medium High
corrosion 5 to 20 years
Reinforcement Corrosion-resistant rebar None High Low
corrosion
Reinforcement External protection systems Continual inspection and High High
corrosion system maintenance
Reinforcement Application of compression by None Medium Low
corrosion design to eliminate tension
Reinforcement Application of compression Continual inspection, Medium Medium
corrosion through posttensioning supplemental posttensioning
ASR Refer to Chapter 3 for material na na na
component–based solutions
ASR Blended aggregates, proper None Medium Low
drainage
ASR Blended aggregates, Continual overlay replacement Medium Medium
waterproof membrane, proper 5 to 20 years
drainage
ACR Refer to Chapter 3 for material na na na
component–based solutions
Note: na = not applicable.

Table 4.13. Bridge-Deck Selection Strategies: Existing Bridge Decks


Potential Material Selection and Life-Cycle Costs
Deterioration Protective Measures
Mode Selection Maintenance Mode Rehabilitation Long-Term
Wear and Overlay and/or membrane Continual overlay replacement Medium Medium
abrasion 5 to 20 years
Reinforcement Overlay and/or membrane Continual overlay replacement Medium High
corrosion 5 to 20 years
Reinforcement External protection systems Continual inspection and High High
corrosion system maintenance
Reinforcement Epoxy injection of deck cracks Continual inspection and Medium Medium
corrosion system maintenance
ASR Waterproof membrane, Continual overlay replacement Medium Medium
proper drainage 5 to 20 years

232

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 4.14. Maintenance Requirements
Demand Examples
Maintenance issues Inspection requirements and intervals
Drainage system maintenance
Membrane, overlay, and sealer maintenance
Expansion joint maintenance
Health monitoring
Scheduled maintenance
Examples of items to be included in Describe how bridge was designed, constructed, and intended to function
bridge Owner’s Manual from an operational perspective. Include the following:
• Design loads
• Expected movements at expansion joints
• Relevant as-built data, including, but not limited to, the following:
–– Concrete mix design
–– Slump test results
–– Chemical content of materials
–– Curing methods used
–– Compression cylinder test results
–– Reinforcing-steel material certifications
–– Coating tests on reinforcement
–– Formwork materials
–– Actual construction procedures
–– Temperature of concrete
–– Ambient temperature and time of casting
–– Timing of casting sequence and concrete delivery
–– Concrete cover measurements

For each component, describe what is needed to achieve design service life
for specific elements. Include the following:
• Required maintenance
• Expected rehabilitation and/or replacement of bridge elements with
service life less than overall bridge system design service life
• Areas for inspection and types of adverse behavior to watch for

Step 7. Life-Cycle Cost Analysis


In Step 7 the bridge-deck alternative is evaluated to establish its life-cycle cost. Life-
cycle cost analysis assesses the overall long-term cost of a strategy throughout the
target service life of the structure. See Chapter 11, Life-Cycle Cost Analysis.
Step 8. Consideration for Additional Bridge-Deck Alternatives
Once the life-cycle cost is established, the bridge deck alternative under investigation
is complete and the process resumes at Step 3 until all feasible bridge-deck alternatives
are addressed.

233

Chapter 4. BRIDGE DECKS


Step 9. Alternatives Comparison and Deck System Selection
The final step in the process is the comparison of bridge-deck system alternatives.
The life-cycle cost of these alternatives is combined with the analysis performed for
each bridge subsystem, component, and element for a specific bridge system. Refer to
Chapter 2 for the overall bridge system selection process. In general, the most cost-
effective bridge-deck system should be chosen; however, the cost-effectiveness has to
be weighed against the requirements of the overall bridge system. The selection should
also be presented to the bridge owner for acceptance.

234

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


5
Corrosion of Steel in
Reinforced Concrete
Bridges

5.1
Introduction
This chapter of the Guide provides essential information for addressing corrosion of
reinforcing steel in conventionally reinforced concrete structures. The focus is on con-
trolling and mitigating corrosion for extended durability and service life. Corrosion in
prestressed or posttensioned concrete structures is not discussed.
The description of corrosion in Section 5.2 covers the diffusion process that enables
the penetration of chlorides through concrete and the creation of corrosion cells once
the chlorides infiltrate. It also addresses the patch-accelerated corrosion commonly
referred to as ring anode corrosion in repairs. Section 5.3 describes factors influencing
corrosion, including chloride contamination and carbonation. Section 5.4 summarizes
strategies for addressing corrosion in new and existing structures; different levels of
corrosion protection are considered, such as corrosion prevention, corrosion control,
corrosion passivation, and electrochemical treatments. Section 5.5 summarizes case
studies that address corrosion in existing structures.
Faced with rising maintenance costs, many engineers and owners recognize the
need to protect existing structures from future corrosion damage. As a result, accord-
ing to Ball and Whitmore (2005), the use of corrosion mitigation systems to delay the
need for future concrete rehabilitation is increasing. Selecting the appropriate corro-
sion mitigation approach is based on many factors, including the amount and depth
of contamination (chloride ingress or carbonation), amount of concrete cracking and
concrete damage, severity and location of corrosion activity (localized or widespread),
expected environmental exposure, use and service life of the structure, and the cost
and design life of the corrosion-protection system.

235
Deicing salts applied during winter months generally contain chlorides. Chloride
solutions penetrate existing cracks and diffuse through the concrete cover to the rein-
forcing steel, initiating corrosion. Corrosion products exert stresses that can crack the
concrete and cause delaminations and spalling.
One approach to mitigating the problem is to prevent or minimize chloride pen-
etration of chlorides by minimizing cracking using low-permeability concretes and
providing adequate concrete cover over the steel, membranes, sealers, or overlays.
Another approach is to prevent the steel from corroding or to minimize the rate of cor-
rosion by using corrosion-resistant reinforcement or cathodic protection. Depending
on the specifics of a project, one or a combination of both of these approaches may
be desirable.
5.2
Description of Corrosion
5.2.1
Corrosion Process
The source for this section is Ball and Whitmore (2005). The corrosion of steel in
concrete is an electrochemical reaction similar to that in a battery. The corrosion rate
is influenced by various factors including chloride ion content, pH level, concrete per-
meability, and availability of moisture to conduct ions within the concrete. For cor-
rosion to initiate in reinforced concrete, four elements are required to complete the
corrosion cell: an anode, a cathode, ionic continuity between the anode and cathode
through an electrolyte, and a metallic (electrical) connection between the anode and
cathode.
The anodic site becomes the site of visible oxidation (corrosion); the cathode is
the location of the reduction reaction, which is driven by the activity at the anode. In
reinforced concrete, the metallic path can be provided by the mild steel reinforcing or
embedded prestressing strands. The ionic path is provided by the concrete matrix with
sufficient moisture due to the permeability of concrete.
At the anode, iron is oxidized to ferrous ions:

Fe → Fe2+ + 2e–

At the cathode, a reduction reaction takes place. In an alkaline environment, the


reduction reaction is typically

2H2O + O2 + 4e– → 4OH–

For the corrosion process to be initiated, the passive oxide film on the reinforc-
ing steel must be broken. In most cases breaking the oxide film occurs as a result
of the presence of sufficient quantities of chloride ions in the concrete matrix at the
level of the steel. Chloride-induced corrosion is most commonly observed in structures
exposed to roadway deicing salts or in marine environments with direct exposure to
salt water or wind-borne sea spray. Chlorides can also be introduced into the con-
crete during the original construction by the use of contaminated aggregates, water, or
chloride-containing admixtures.

236

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Over time, the corroding area (anode) will become more acidic as hydroxyl ions
(OH–) are consumed from the concrete in contact with the corroding area, and the
cathode will become more alkaline by the generation of hydroxyl ions (OH–).
5.2.2
Diffusion Process
The main source for this section and the next (5.2.3) is Whitmore (2002). To the ­casual
observer, uncracked concrete is a solid and impenetrable material. Viewed under a
micro­scope, however, concrete is a labyrinth of fine capillaries, pores, and voids be-
tween the individual cement and aggregate particles. The degree of porosity depends
on the quality and density of the concrete mix.
As a result of this porosity, liquids can soak into the exposed surfaces of concrete
and carry contaminants such as chloride ions with them. Over time, the concentration
of chloride ions within the concrete will tend to equalize as governed by Fick’s law.
In a one-dimensional case, Fick’s law can be expressed as shown in Equation 5.1:
 x 
C( x,t ) = C0  1 − erf  (5.1)
 2 Dc t 

where
C(x,t) = chloride concentration at depth x and time t,
C0 = surface chloride concentration (kg/m3 or lb/yd3),
Dc = chloride diffusion constant (cm2/year or in.2/year), and
erf = error function (from standard mathematical tables).

This expression indicates that over time the chloride concentration within the con-
crete will tend to equalize with the chloride concentration exposed to the surface. As
expected, the chloride concentration within the concrete is greater near the exposed
surface and increases with time at any point within the concrete. Concrete with a
lower chloride diffusion constant (Dc) will provide longer-term protection to reinforc-
ing steel located at depth x from the surface of the concrete.
The diffusion constant for a particular point in a concrete structure may be deter-
mined if chloride data are available for one location at two points in time or if a com-
plete chloride profile is available some time after the structure has been constructed.
With current chloride data and an estimate of the diffusion coefficient, future chloride
profiles can be predicted using the formula in Equation 5.1. Figure 5.1 displays the
chloride concentration within concrete over time.
Figure 5.2 shows the chloride contents with depth. Based on the chloride profile,
the calculated (best-fit) diffusion coefficient Dc is 4.38 × 10–13 cm2/s.
If the concrete element is cracked, chloride penetration at crack locations may
greatly exceed chloride levels in the surrounding concrete. This level of chloride pen-
etration can lead to corrosion initiation at crack locations long before general corro-
sion may otherwise occur.

237

Chapter 5. CORROSION OF STEEL IN REINFORCED CONCRETE BRIDGES


Figure 5.1.  Chloride concentration
within ­concrete over time.

Test Values Fit Using Equation 5.1; C0 fixed


Figure 5.2.  Chloride contents with depth.

5.2.3
Corrosion Cells
Once the chloride concentration at the depth of the reinforcing steel exceeds threshold
levels, the passive oxide film will begin to degrade, and corrosion may be initiated.
Chlorides act similarly to a catalyst in the corrosion process: the chlorides are involved
in the corrosion reaction, but they are generally not consumed by the corrosion reac-
tion itself, such that a single chloride ion can be responsible for the corrosion of many
atoms of iron.

238

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


On a localized basis, corrosion cells can be formed as a result of differences in
chloride concentration at various locations along a single bar. These variations can
result in localized pitting-type corrosion. Similarly, if entire sections of a reinforced
concrete structure become contaminated relative to other adjacent areas, an overall
corrosion cell or “macrocell” can be created, as illustrated in Figure 5.3. Macrocell
corrosion can be very aggressive and is responsible for much of the severe structural
damage seen on bridges and other structures. Both pitting-type corrosion and general
corrosion result from corrosion cells (Whitmore 2002).
The corrosion products (rust) occurring as a result of macrocell corrosion occupy
a large volume and cause cracking, concrete delamination, and spalls.
5.2.4
Patch-Accelerated Corrosion
Commonly referred to as ring anode corrosion or halo effect, patch-accelerated cor-
rosion is a phenomenon specific to concrete restoration projects (Figure 5.4). When
repairs are completed on corrosion-damaged structures, abrupt changes in the con-
crete surrounding the reinforcing steel are created. Typical concrete repair procedures
call for the removal of the concrete around the full circumference of the reinforcing
steel within the repair area, cleaning corrosion byproducts from the steel, and refill-
ing the cavity with new chloride-free, high-pH concrete. These procedures leave the

Figure 5.3.  Corrosion macrocell in a concrete


Fe Fe 2+ + 2e - Chloride-Contami
C inated
Fe 2+ + 2Cl - FeCl 2 Concrete deck.

FeCl 2 + 2OH - Fe(O


OH) 2 + 2Cl -
2Fe(O
OH) 2 + 1/2O 2 Fe 2O 3 + 2H 2O

2O
OH -
2e -

1/
2O 2 + H 2O + 2e 2O
OH -
-

Figure 5.4.  Patch-accelerated corrosion.

239

Chapter 5. CORROSION OF STEEL IN REINFORCED CONCRETE BRIDGES


reinforcing steel embedded in adjacent environments with abruptly different corrosion
potentials. This difference in corrosion potential (voltage) is the driving force behind
new corrosion sites forming in the surrounding contaminated concrete. The evidence
of this activity is the presence of new concrete spalling adjacent to previously com-
pleted patch repairs.
5.3
Factors Influencing Corrosion
One of the leading causes of concrete rehabilitation is corrosion-induced concrete
damage and spalling in reinforced concrete structures (Figure 5.5). In steel-reinforced
concrete, the concrete matrix must be sufficiently strong to resist applied forces from
a structural standpoint and to serve as a corrosion-protection mechanism for the em-
bedded reinforcing steel. The ability of concrete structures to resist corrosion attack is
not related to the mechanical strength of the concrete alone. Instead, two important
factors limit the ability of concrete structures to resist corrosion: the presence of cracks
and the porosity of the concrete.
The presence of cracks and the ability of chlorides to permeate the concrete allow
chlorides to get to the reinforcing steel, thus compromising the corrosion resistance
provided by concrete’s naturally high alkalinity.
As discussed in Ball and Whitmore (2005), numerous factors may influence the
durability of concrete, including the water-cement ratio, permeability, curing, shrink-
age and cracking, ingredients including admixtures, and the severity of environmental
exposure. Due to the high alkalinity of the concrete pore water solution, a thin passive
oxide layer is formed and maintained on the surface of the embedded steel that pro-
tects it from corrosion activity. Until this passive film is destroyed by the intrusion of
aggressive elements or a reduction in the alkalinity of the concrete, the reinforcement
will remain in a passive, noncorroding state.
The causes of corrosion are summarized in Figure 5.6. Chloride contamination
and carbonation are explained in the sections that follow.

Figure 5.5.  Corrosion-induced delamination on


a bridge pier.

240

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Causes of Corrosion
of Steel in Concrete

Chlorides Carbonation Dissimilar Metals

Surface-Applied
Cast-In Chlorides
Chlorides

Figure 5.6.  Causes of corrosion of steel in concrete.

5.3.1
Chloride Contamination
Destruction of the protective oxide film on reinforcing steel is most often caused by the
presence of elevated levels of chloride ions. The chloride threshold that initiates corro-
sion is generally considered to be around 1.0 to 1.4 lb of water-soluble Cl– per cubic
yard of concrete (at the level of the steel). This chloride threshold varies depending on
the pH of the concrete. For example, concrete that has experienced a loss of alkalinity
requires less chloride to initiate corrosion. Chloride-induced corrosion, illustrated in
Figure 5.7, is common in structures exposed to deicing salts, marine environments, or
certain industrial processes. In some cases, sufficient amounts of chlorides capable of
causing corrosion have been introduced during construction by the use of chloride-
containing admixtures or contaminated aggregates.

Figure 5.7.  Chloride-induced corrosion of


­reinforcing steel.

241

Chapter 5. CORROSION OF STEEL IN REINFORCED CONCRETE BRIDGES


Non-chloride-bearing salts, including calcium magnesium acetate, magnesium
acetate, and calcium acetate, lower the freezing temperature of water and can be used
for ice control. However, magnesium-bearing solutions cause severe paste deteriora-
tion by forming brucite and noncementitious magnesium silicate hydrate (Lee et al.
2000). The detrimental effects caused by calcium acetate are much less severe than
those caused by magnesium, but the use of these non-chloride-bearing salts has not
gained wide acceptance due to cost and the distress they cause.
5.3.2
Carbonation
The main source for this section is Ball and Whitmore (2005). The passive condition
of the reinforcing can also be disrupted by the loss of alkalinity in the concrete matrix
surrounding the reinforcing steel. It is generally accepted that a pH greater than 10 is
sufficient to provide corrosion protection in chloride-free concrete. The reduction in
alkalinity is generally caused by carbonation, a reaction between atmospheric carbon
dioxide and calcium hydroxide in the cement paste in the presence of water. The result
is a reversion of the calcium hydroxide to calcium carbonate (approximate pH 8.5),
which has insufficient alkalinity to maintain the passive oxide layer.
The zone of carbonation begins on the surface of atmospherically exposed con-
crete. The amount of time for the carbonation zone to reach the level of the reinforcing
is a function of the thickness of concrete cover, presence and extent of cracks, concrete
porosity, humidity levels, and the level of exposure to carbon dioxide. Carbonation-
induced corrosion is more likely to be observed in structures situated in industrial envi-
ronments, where airborne pollutants are commonplace; in old or historic structures
with a high degree of concrete porosity; or in structures with low concrete cover over
the reinforcement.
In bridge structures, there is generally good-quality concrete cover of sufficient
thickness (about 2.5 in. in decks) to resist carbonation for up to 100 years.
5.4
Strategies for Addressing Corrosion
Because of the magnitude of the corrosion problem, both the public and private sectors
have ongoing activities aimed at reducing or eliminating corrosion damage to concrete
structures. Although many technologies and materials have been developed for the
prevention and repair of corrosion-induced damage, the challenge is to select durable,
cost-effective technologies and materials from the numerous choices available.
Durability and desired service life must be considered during design. In the case
of new structures, it is desirable to avoid, prevent, or delay the initiation of corrosion
through the use of low-permeability concretes, proper precautions against cracking
(see Chapter 3), and other corrosion-prevention techniques. For existing structures,
the condition of the structure should be evaluated to determine whether it is corroding.
If the structure is corroding and the chloride content is high in the concrete, the con-
taminated concrete is removed; if rust is forming on the surface of the reinforcement,
it is cleaned or removed and overlaid with a low-permeability concrete overlay. If the

242

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


existing structure is not corroding, some of the techniques used on new structures,
such as applying sealers and membranes, may also be used on the existing structure.
An overview of the available options is provided in Figures 5.8 through 5.11.
Many strategies have been used successfully to improve the corrosion resistance
and durability of new structures. These strategies include

• The use of low-permeability concrete;


• The use of increased concrete cover;
• The use of improved construction methods such as curing to minimize cracking;
• The use of corrosion-resistant reinforcement;
• The use of corrosion inhibitors to increase the corrosion initiation threshold;
• The use of membranes, coatings, and sealers; and
• The use of improved design details to keep elements dry and to prevent exposure
to chlorides.

In the United States, epoxy-coated reinforcement has been widely used as a


c­ orrosion-protection system for concrete bridges. However, because recent work and
observations in the field have shown that the longevity desired (75 years and beyond)
may not be achievable, other corrosion-resistant reinforcements are being considered
(see Chapter 3 on materials).
Each of these methods can be effective if it significantly extends the time for cor-
rosion to initiate. In many cases it is preferable to employ more than one technique,
which will generally reduce the overall risk of corrosion.
Additional information on these topics is provided in Chapter 3 on materials.

Reduced Service Life


of Reinforced
Concrete Structures

New Structures
Existing Structures
(Not Corroding)

Avoid Prevent/Delay Electro chemical Currently Not Currently


Corrosion Corrosion Initiation Passivation Corroding Corroding
(Immunity)

Figure 5.8.  Reduced service life of reinforced concrete.


Figure 5.8. Reduced service life of reinforced concrete.

243

Chapter 5. CORROSION OF STEEL IN REINFORCED CONCRETE BRIDGES


Avoid Corrosion

Use Corrosion-
Eliminate Cause of
resistant Reinforcing
Corrosion
Materials

Fiber-reinforced Internal—
Corrosion- External Exposure
Stainless Steel Polymer (FRP) No Contamination.
resistant Steel (CI– or Carbonation)
Reinforcing No Cast-in CI–.

Figure 5.9.  Options for avoiding corrosion.


Figure 5.9. Options for avoiding corrosion.

Prevent/Delay
Corrosion Initiation

External— Internal—
Delay Corrosion
Eliminate Exposure No Contamination.
Initiation
No Cast-in Cl–

Use Low- Reduce/Eliminate Increase Concrete Increase Corrosion


Use Protective
Permeability Cracks Cover Thickness Threshold
Barrier
Concrete (i.e., Inhibitor)

Figure 5.10.  Options for preventing or delaying corrosion initiation.

Figure 5.10. Options for preventing or delaying corrosion initiation.


Electrochemical
Passivation (Immunity)

Impressed
Galvanic Treatment
Current

Figure 5.11.  Electrochemical passivation.

244

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


5.4.1
Existing Structures
Options for protecting structures from corrosion and extending their service life are
much more limited when dealing with existing structures, as many of the physical
parameters are already defined and cannot be changed or easily altered. In particular,
the concrete and reinforcing steel of existing structures are already in place, and the
characteristics of these materials, including type, quality, cover thickness, permeability,
resistance to corrosion initiation, and presence of cracks, are already fixed. Because of
these limitations, many of the options that are viable for new (noncorroding) construc-
tion, as shown in Figure 5.12, are not possible or may not be economically practical
for use with existing (corroding) structures. Figure 5.13 shows the options for corrod-
ing structures.
Despite the reduced number of options available for existing structures, it is still
beneficial to be as proactive as possible. Preventing or delaying corrosion is generally
preferable to managing corrosion after it has initiated. Additional options (and often
more economical options) exist to prevent or delay corrosion activity if the structure is
not chloride contaminated or has not already started to corrode.

Structures not
Currently
Corroding

Cl– Contaminated
Carbonated to Not Cl–
Steel Contaminated

Treatment to
Corrosion Treatment to Slow
Delay Corrosion Remove Source Do Nothing
Prevention Deterioration
Initiation

Galvanic

Impressed Current

Electrochemical
Treatment

Figure 5.12.  Noncorroding structures.


Figure 5.12. Noncorroding structures.
245

Chapter 5. CORROSION OF STEEL IN REINFORCED CONCRETE BRIDGES


Options for
Currently
Corroding Structures

Treatment
to Reduce Rate Partial Removal and Cathodic Protection Electrochemical
Do Nothing
of Deterioration Replacement Treatment

Impressed Current Galvanic

Figure
Figure5.13. 
5.13.Options
Optionsfor
forcorroding structures.
corroding structures.

5.4.2
Levels of Corrosion Protection for Existing Structures
5.4.2.1
Selecting an Active Corrosion-Protection Strategy for Reinforced Concrete
Structures
The selection of the appropriate level of corrosion protection is based on many fac-
tors, such as the level of chloride contamination and carbonation, amount of concrete
damage, location of corrosion activity (localized or widespread), the cost and design
life of the corrosion-protection system, and the expected service life of the structure
(Ball and Whitmore 2005).
The levels of corrosion protection described in this section are summarized in
Table 5.1.

Table 5.1.  Summary of Levels of Corrosion Protection for Electrochemical


Corrosion Mitigation Systems
Level of Protection Description
Cathodic protection Stops active corrosion by applying ongoing electrical current
Corrosion prevention Prevents new corrosion activity from initiating
Corrosion control Significantly reduces active corrosion
Corrosion passivation– Stops active corrosion by changing the chemistry of the
electrochemical treatment concrete around the steel

246

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


5.4.2.2
Cathodic Protection
Cathodic protection provides proven corrosion protection and is intended to effec-
tively stop ongoing corrosion activity. It should be selected when the highest level of
protection is necessary and the cost is economically justified. Cathodic protection sys-
tems are grouped into two general categories: impressed current and galvanic.
Impressed current systems may use titanium- or zinc-based anodes and an outside
power source. For long-term performance, these systems should be monitored and
maintained. Discrete anodes are ideal to protect heavily reinforced concrete, thick
structural sections such as columns or beams, or steel-framed masonry buildings; tita-
nium ribbon or mesh anodes are placed in slots cut into the concrete surface or cast
into a concrete overlay.
Galvanic systems may be designed to provide corrosion control or cathodic protec-
tion. These systems are self-powered and typically require less monitoring and main-
tenance than impressed current systems. Galvanic jackets are used to protect marine
pilings and other structures. Galvanic anodes may be arc-sprayed zinc or otherwise
applied to the concrete surface, or they may be cast into a concrete overlay, jacket, or
encasement to provide galvanic cathodic protection over a desired area. If galvanic
anode systems are cast into concrete or are not directly exposed to a marine envi-
ronment, they should be activated to ensure that sufficient current is supplied to the
­reinforcing steel to provide long-lasting corrosion protection.
Cathodic protection systems are generally designed to meet National Associa-
tion of Corrosion Engineers (NACE) cathodic protection standards and typically use
100-mV depolarization as the acceptance criteria (NACE 2000). The current density
required to achieve cathodic protection is higher than the current required for corro-
sion prevention or corrosion control applications. Typical cathodic protection systems
operate in the range of 2 to 20 mA/m2 of steel surface area. At these current densities
and polarization levels, cathodic protection has demonstrated a very high level of cor-
rosion protection (Ball and Whitmore 2005).
Galvanic cathodic protection systems were evaluated in the research phase of
SHRP 2 Project R19A, the report for which is forthcoming. In this study, four gal-
vanic anodes were evaluated for the purpose of minimizing corrosion in reinforced
concrete members: a commonly used (ordinary) anode, an anode with a zinc surface
area four times larger than ordinary (OA4) anodes, and two high-voltage anodes with
different degrees of output voltage (a higher-level high-voltage anode and a lower-level
high-voltage anode). Concrete test slabs were cast in two layers; the concrete in the
upper layer was contaminated with salt to accelerate the corrosion activity, but the
lower layer was not modified. (Salt was not added to the concrete in the lower layer.)
The test results indicated that there was no corrosion in any of the specimens in the
given time period. The testing further indicated that specimens with the higher-level
high-voltage anode provided increased corrosion protection by having higher current
and generating more negative potential values than specimens with the lower-level
high-voltage anode. OA4 anodes provided additional current and corrosion protec-
tion compared with the ordinary anodes. Lower-level high-voltage anodes exhibited

247

Chapter 5. CORROSION OF STEEL IN REINFORCED CONCRETE BRIDGES


current and potential values similar to the OA4 anodes. The two high-voltage anodes
and the OA4 anodes provided higher current and generated more negative potential
values, indicating better corrosion protection, than ordinary anodes. Because of time
constraints, the tests were terminated without observing corrosion in the specimens.
Further research with an extended time frame is recommended.
5.4.2.3
Corrosion Prevention
Corrosion prevention is used to keep corrosion activity from initiating in contaminated
concrete. In concrete repair projects, the removal and replacement of damaged con-
crete, if completed in accordance with industry guidelines (ICRI 2008), will address
the areas with the highest levels of corrosion. However, new corrosion sites are likely
to form in the surrounding contaminated concrete that was passive before the repairs.
To mitigate new corrosion activity from occurring around concrete repairs or at other
interfaces between new and old concrete, such as bridge widening, joint repairs, and
slab replacements, a localized corrosion prevention strategy may be employed using
embedded galvanic anodes to extend the life of the concrete repairs. Size and spacing
of embedded galvanic anodes should be adjusted to suit site conditions, including
quantity and existing condition of reinforcing steel, level of chloride contamination,
and environmental conditions.
Ball and Whitmore (2005) point out that there has been a significant amount of
research in corrosion prevention, some of which has indicated that applied current
densities as low as 0.5 to 2.0 mA/m2 of steel surface area are effective at preventing the
initiation of corrosion for concrete with chloride concentrations up to at least 10 times
the chloride threshold (Pedeferri 1996). Other research has shown beneficial effects of
applied currents between 0.25 and 1.0 mA/m2 (Bertolini et al. 1996).
5.4.2.4
Corrosion Control
Corrosion control systems are used when corrosion has initiated but has not yet pro-
gressed to the point of causing concrete damage. The use of corrosion control systems
will either stop ongoing corrosion activity or provide a significant reduction in the
corrosion rate and thus increased service life for the rehabilitated structure. In many
cases, this level of protection can be provided with low incremental cost, as the protec-
tion can be targeted at specific areas of contamination or corrosion activity. Galvanic
anodes embedded in drilled holes or installed on a grid pattern in a concrete repair or
overlay can be used to provide targeted galvanic corrosion control to columns, beams,
decks, posttensioned anchorages, and other areas where ongoing corrosion activity
threatens the service life or serviceability of the structure.
The applied current necessary to control active corrosion is significantly higher
than the current required for corrosion prevention (Ball and Whitmore 2005). Davison
et al. (2003) achieved corrosion control of specimens with very high initial corrosion
rates. Their research indicated that the typical current density to control active corro-
sion is in the range of 1 to 7 mA/m2 (Davison et al. 2003). Some polarization of the
reinforcing steel will typically occur at these current densities, although the level of

248

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


polarization may be significantly less than the NACE 100-mV depolarization criteria
for cathodic protection.
In many situations corrosion activity is moderate or localized so that corrosion
control is an appropriate approach because large-scale cathodic protection may not
be economically justifiable (Ball and Whitmore 2005). Examples of such situations
include localized areas beneath leaking expansion joints or decks with isolated areas of
high corrosion potentials. In these cases, targeting the protection to address the specific
contaminated zone, or hot spot, rather than the entire structure may make sense from
a cost–benefit point of view.
5.4.2.5
Corrosion Passivation by Electrochemical Treatment
Corrosion passivation is provided by electrochemical treatments that are aimed at
directly addressing the cause of the corrosion activity. Electrochemical chloride extrac-
tion (ECE) is used to address corrosion caused by chlorides in chloride-contaminated
structures. Electrochemical realkalization is used to address corrosion resulting from
carbonation of the concrete. These systems are installed on the structure, operated for
a short duration, and then dismantled and removed, leaving the structure in a pas-
sive condition. Electrochemical treatments provide many of the long-term corrosion
mitigation benefits of cathodic protection systems, but without the need for long-term
system maintenance and monitoring. Additional information about the implementa-
tion and evaluation of the two systems follows.
ECE is an electrochemical treatment in which an electric field is applied between
the reinforcement in the concrete and an externally mounted mesh (Figure 5.14). The
mesh is embedded in a conductive media, generally a sprayed-on mixture of lime,
water, and cellulose fiber.
During treatment, as pointed out by Ball and Whitmore (2005), the concrete is
kept saturated, which allows chlorides to go into solution within the pores of the
concrete. The negatively charged chloride ions (Cl–) are repelled from the negatively
charged rebar and attracted toward the positively charged external electrode mesh
as a result of the applied electric field. This process lowers the amount of chloride in
the concrete, particularly adjacent to the steel. An ECE treatment generally takes 4 to
8 weeks to complete.
In instances of structures with carbonation-induced corrosion, a different electro-
chemical treatment process, realkalization, can be used to increase the pH of the con-
crete cover. The installation for realkalization is essentially the same as for ECE except
the conductive media is saturated with an alkaline potassium carbonate solution. The
potassium (K+) ions in the alkaline solution are transported into the concrete by the
application of the electric field. A realkalization treatment generally takes 4 to 7 days
to complete and will not recarbonate (Ball and Whitmore 2005).
Electrochemical treatment was evaluated in the research phase of SHRP 2 Project
R19A; detailed results will be presented in the forthcoming final report. In the SHRP 2
study, corrosion-resistant reinforcement and electrochemically treated black bars were
evaluated to determine if the chloride threshold levels increased to a level to resist
corrosion. Both low and high levels of electrochemical treatment were used. The test

249

Chapter 5. CORROSION OF STEEL IN REINFORCED CONCRETE BRIDGES


Figure 5.14.  Schematic diagram showing ECE treatment process.
Figure 5.14. Schematic diagram showing ECE treatment process.

matrix included black bars, electrochemically treated black bars, stainless steel bars,
and titanium bars, each subjected to salt solution. Because of time constraints, testing
was terminated after 26 cycles consisting of 4 days of wet cycle and 10 days of dry
cycle for a total duration of 1 year. At termination, only one specimen from a set of
three with black bars and electrochemically treated black bars showed an increase in
current or potential values indicative of uncertain corrosion activity. The remaining
specimens indicated no corrosion activity. Thus, initial observations indicate that elec-
trochemically treated black bars may not provide the protection expected of stainless
steel or titanium; however, whether they provide benefits over the black bars without
treatment could not be concluded from this study due to time constraints. Further
research with an extended time frame is recommended.
5.5
Case Studies Addressing Corrosion in Existing Structures
5.5.1
Project Overview: Impressed Current Cathodic Protection,
­Corrosion Control, and Electrochemical Chloride Extraction
During summer 1989 and continuing until 1994, the Ontario Ministry of Transporta-
tion completed various restoration and protection projects on the reinforced concrete
piers of the Burlington Skyway, a major viaduct located between Toronto and Niagara
250

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Falls, Ontario. This work was monitored by the Ministry’s Materials and Research
Branch. Some of this research was conducted under Project SHRP-C-620 and is docu-
mented in Evaluation of NORCURE Process for Electrochemical Chloride Removal
from Steel-Reinforced Concrete Bridge Components (Bennet and Schue 1993).
An impressed current cathodic protection system was installed on more than
200,000 ft2 of reinforced concrete substructure. The system was designed to operate
as an impressed current system with an applied current density of 10 mA/m2 (1 mA/
ft2). After an initial period of operation at the design current density, and because
of operational issues, the system was run at an average current density of 1 mA/m2
(0.1 mA/ft2). Thus, instead of being operated at a cathodic protection current density
(2 to 20 mA/m2), the system was effectively operated at a corrosion control current
density (1 to 7 mA/m2). The low applied current density meant that much of the area
did not meet the 100-mV NACE criteria for cathodic protection. Despite operating
at approximately 10% of the intended cathodic protection current density, the sys-
tem operation fell within typical corrosion control current densities and experienced
a significant reduction in concrete deterioration and damage. Over the study period,
concrete delamination within the protected area was reduced by 96% compared with
the rate of deterioration of unprotected concrete piers.
In 1989 the Ontario Ministry of Transportation also completed an ECE trial on
a section of the same structure. The treated portion comprised rectangular piers and
bents. The piers were contaminated with chlorides from long-term leakage of the deck
joints above.
ECE was used to reduce the chloride content of the concrete to below the thresh-
old level of corrosion at the rebar. This process eliminated high corrosion-potential
readings, and corrosion potentials in the treated area were shifted into the passive
range, as shown in Table 5.2; the ECE-treated section exhibited a high percentage of
area with a low risk of corrosion. Corrosion current as measured by linear polariza-
tion was greatly reduced, as shown in Table 5.3; the ECE-treated sections exhibited
a high percentage of passive area. Thus, corrosion potentials and corrosion currents
were reduced to the passive, noncorroding range as a result of the ECE treatment for a
20-year duration (Tables 5.2 and 5.3). These readings have remained stable and show
no significant changes over the period.
Additional piers were treated during the fall of 1997 as part of the next phase of
work on this project. The long-term results from these more recent tests are expected
to be similar to the ECE trial completed in 1989.
5.5.2
Project Overview: Galvanic Cathodic Protection Using Galvanic
Encasement of Severely Corroded Elements
The source for this section is Sergi (2009). Galvanic anodes have also been developed
for more global corrosion control. One such configuration is the system installed to
repair and protect severely damaged and corroding bridge abutments in the Midwest
(Figure 5.15). The abutments had been contaminated with chlorides causing corrosion
of the reinforcing steel and significant concrete damage.

251

Chapter 5. CORROSION OF STEEL IN REINFORCED CONCRETE BRIDGES


Table 5.2. Corrosion Potential Measurements
Untreated (Control) ECE Treated
Area Area
Area (%) (%) with Area (%) Area (%) (%) with Area (%)
with Low Moderate with High with Low Moderate with High
Risk of Risk of Risk of Risk of Risk of Risk of
Time Since Corrosion Corrosion Corrosion Corrosion Corrosion Corrosion
Treatment <200 mV 200–350 mV >350 mV <200 mV 200–350 mV >350 mV
Pretreatment 0 85 15 0 96 4
1 Year after 41 59 0 98 2 0
2 Years after 41 59 0 100 0 0
3 Years after 26 74 0 96 4 0
4 Years after 26 70 4 98 2 0
6 Years after 26 59 15 96 4 0
8 Years after 11 78 11 96 4 0
10 Years after 15 78 7 96 4 0
15 Years after 20 70 10 98 2 0
20 Years after 15 70 15 96 4 0

Note: Corrosion potentials measured in –mV versus Cu-CuSO4. Values represent percentage of readings within range.

Table 5.3. Corrosion Current Measurements


Untreated (Control) ECE Treated
Area (%) Area (%) Area (%) Area (%)
Area (%) with Low with High Area (%) with Low with High
with Passive Corrosion Corrosion with Passive Corrosion Corrosion
Time Since 0 <0.22 0.22–1.08 >1.08 <0.22 0.22–1.08 >1.08
Treatment mA/cm2 mA/cm2 mA/cm2 mA/cm2 mA/cm2 mA/cm2
Pretreatment 0 46 54 0 87 13
1 Year after 0 52 48 87 13 0
2 Years after 0 84 16 78 20 2
3 Years after 4 88 8 89 11 0
4 Years after 8 86 6 90 10 0
6 Years after 0 40 60 65 35 0
8 Years after 0 29 71 63 37 0
15 Years after 0 62 38 65 34 1
20 Years after 15 70 15 96 4 0

Note: Values represent percentage of readings within range.

252

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 5.15.  A distributed anode
­system used for corrosion control
bridge abutment in the Midwest.

Distrributed
anoddes

As part of the rehabilitation, which also included enlargement and strengthening


of the abutment, the cracked and spalled concrete was removed. Elongated anodes
were connected to the existing reinforcing steel and encased in a new layer of concrete
to reface the abutment wall. The purpose of the anodes was to protect the existing
steel from chloride-induced corrosion. This allowed uncracked chloride-contaminated
concrete to remain in place and thereby reduce concrete breakout and the need for
structural shoring. The cross-sectional configuration of the repaired abutment wall
and adjoining structural elements is shown in Figure 5.16.

Figure 5.16. Cross-sectional
detail of the abutment
rehabilitation system.

Distrib
buted anode system

253

Chapter 5. CORROSION OF STEEL IN REINFORCED CONCRETE BRIDGES


The current output, shown in Figure 5.17, appears to be strongly related to tem-
perature. Its magnitude varied considerably with temperature on an annual basis, with
the mean current density gradually reducing year by year. After supplying an initial
current of over 35 mA/m2 of steel area in the first few days, the output averaged over 8
mA/m2 during the first year, decreasing gradually to around 5 mA/m2 in the fourth year.
These levels of current density are within the design limits of 2 to 20 mA/m2 of steel
area for cathodic protection as specified in European Standard BS EN 12696:2012
(BSI 2012). Current densities in impressed current cathodic protection systems are
also normally reduced with age as the steel becomes easier to polarize. Depolarization
levels were measured to be well in excess of 100 mV, as specified in the same standard,
suggesting that the galvanic system was able to satisfy the criteria for cathodic protec-
tion of steel reinforcement. Depolarization and corrosion-protection status are given
in Table 5.4.
5.5.3
Project Overview: Corrosion Prevention Using Galvanic Anodes
The oldest site trial of discrete galvanic anodes is over 10 years old (Figure 5.18). To
verify the performance of the anodes, 12 anodes were installed in an otherwise con-
ventional patch repair on the soffit of a bridge beam (Figure 5.19). The performance
of these anodes was monitored over time.

60 120
Galvanic Current
55
Manual Current
50 Temperature 100

45
Galvanic Current, mA

Temperature, °F
40 80

35

30 60

25

20 40

15

10 20

0 0
May-05 Nov-05 May-06 Nov-06 May-07 Nov-07 May-08 Oct-08 May-09 Oct-09 May-10 Oct-10 May-11 Oct-11 May-12

Date
Figure 5.17.  Current output of anode system and its relationship to temperature.

Figure 5.17. Current output of anode system and its relationship to temperature.
254

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 5.4. Depolarization and Corrosion-Protection Status
Date Temperature (°F) Current Density (mA/ft2) Depolarization (mV) Status
May 6, 2005 88 >3.5 na Corrosion protected
Aug. 16, 2005 87 1.23 344 Corrosion protected
Oct. 26, 2005 54 0.52 368 Corrosion protected
Dec. 7, 2005 51 0.28 310 Corrosion protected
May 1, 2006 57 0.70 313 Corrosion protected
Dec. 20, 2006 40 0.36 459 Corrosion protected
May 30, 2007 79 0.72 449 Corrosion protected
Sept. 20, 2007 75 0.88 482 Corrosion protected
Dec. 19, 2008 40 0.34 450 Corrosion protected
July 9, 2009 74 0.27 471 Corrosion protected
May 11, 2010 54 0.33 485 Corrosion protected
Oct. 16, 2011 72 0.57 488 Corrosion protected

Note: na = not applicable.

Figure 5.18.  North side of bridge with discrete


galvanic anodes.
Source: Sergi 2009.

Figure 5.19.  Installation of anodes within the


repaired area of a beam; control box and wiring
are also shown.
Source: Sergi 2009.

255

Chapter 5. CORROSION OF STEEL IN REINFORCED CONCRETE BRIDGES


The anodes were inserted around the perimeter of the repair area between 600-
and 700-mm centers (Figure 5.19). The anodes were installed to enable monitoring
by connecting a single wire from each anode to a control box such that the anodes
could be monitored via the box. Monitoring of the anodes consisted of measuring the
current output for each installed anode and, on occasion, performing a depolarization
test over a 4- or 24-hour period after disconnection of the anodes. Monitoring started
in April 1999.
The main source for the rest of this chapter is Sergi (2009). Ten-year results of the
current output of each anode are presented in Figure 5.20. They indicate a variable
current depending on the ambient temperature and moisture content in the concrete.
For example, the same anode could generate up to 400 to 600 μA of current during
hot periods and less than 100 μA during cold spells. Corrosion of the steel is expected
to have similarly varying corrosion rates so that the current output of the anodes is
thought to be self-regulating, producing higher levels when the steel is corroding most.
Systems such as this are designed to prevent the onset of corrosion of the reinforce-
ment. The current density required for corrosion prevention (referred to as cathodic
prevention in Europe) is 0.2 to 2 mA/m2, as reported by Bertolini et al. (1993) and
Pedeferri (1996), and adapted in the European Standard BS EN 12696:2012 (BSI
2012). Based on the steel surface area within the repaired and affected adjacent areas,
the mean current density ranged between 0.6 and 3.0 mA/m2 through the duration of

Figure 5.20.  Protective current from galvanic anodes over a 10-year period.

256
Figure 5.20. Protective current from galvanic anodes over a 10-year period.
DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE
this trial, with a mean current density of around 1.4 mA/m2 over the 10-year period.
This current density is within the suggested range of 0.2 to 2.0 mA/m2 for corrosion
prevention (cathodic prevention).
Monitoring the depolarized potential of the steel in the vicinity of the repair with
time may be another way of determining the effectiveness of the system. Figure 5.21,
which shows the mean depolarized potential with time both within and outside the
repaired area, indicates that the mean potential is moving to a more noble level with
time. This change indicates increasing passivation of the steel over time.

300 mm Outside Repair 50 mm Outside Repair Within Repair

100

50

0
Steel Potential (mV vs. Cu-CuSO4)

-50

-100

-150

-200

-250

-300

-350

-400
10 100 1000 10000
Time (Days)
Figure 5.21.  Mean depolarized steel potentials with time (4 or 24 hours after disconnection of the anodes).
Source:
Source: SergiSergi
2009.(2009).

Figure 5.21. Mean depolarized steel potentials with time (4 or 24 hours after disconnection of the

anodes).

257

Chapter 5. CORROSION OF STEEL IN REINFORCED CONCRETE BRIDGES


6
Corrosion Prevention
of Steel Bridges

6.1
Introduction
This chapter is a best-practices guide and discussion for preventing corrosion of ex-
posed structural steel for bridges and includes factors to be considered for design
through installation, inspection, and maintenance. Various types of coatings, includ-
ing painting, galvanizing, and metalizing, are discussed along with other methods of
corrosion prevention that include the use of steels with higher resistance to corrosion,
such as weathering steel.
Figure 6.1 shows the structural steel elements susceptible to corrosion. The focus
of this chapter is on superstructure elements; however, much of the discussion is also
applicable to deck and substructure elements.
Within the superstructure component, structural steel subsystem elements include
all configurations of steel shapes and plates that alone or in combination comprise
members used as supporting steel on various types of structures, including trusses,
beams, haunch parallel-flange welded plate girders, multiple web and single-bottom-
flange tub girders, and square or rectangular cross-section box girders. Also included
are all angles, channels, fasteners, sole plates, diaphragms, shims, and bearings.
There are three primary methods for preventing corrosion of structural steel:

1. Use of coating systems,


2. Use of corrosion-resistant steel (weathering steel) or noncorrosive steel, and
3. Avoidance of corrosive environments or corrosion-prone details.

Each of these methods is discussed in this chapter. The use of partially painted
weathering steel is also addressed.

258
Structural Steel Elements
Subject to Corrosion

Deck Superstructure Substructure

Steel/Orthotropic Deck Structural Steel Steel Pier Caps or


Subsystem Elements Columns

Expansion Joints
Open/Sealed Bearings Steel Piles

Drainage Elements

Railings

Figure 6.1.  Structural steel elements subject to corrosion.


Figure 6.1. Structural steel elements subject to corrosion.

6.2
Description of Methods for Corrosion Prevention
This section discusses the corrosion process and describes the three main methods used
to prevent corrosion of steel bridges.
6.2.1
Corrosion of Steel: General Discussion
In its simplest form, the corrosion of steel results from exposure to oxygen and mois-
ture. Corrosion is accelerated in the presence of salt from roadway deicing, salt water,
or perhaps salt deposited from other sources. The fact that steel corrodes is one of the
few fundamental limitations of steel as a material of construction.
Although steel corrodes readily in the presence of oxygen and moisture, the rate
of corrosion is accelerated in the presence of chloride ions or other corrosive chemi-
cals. Chloride ions result mainly from the use of deicing agents composed of materials
with readily soluble chloride ions. These ions create an atmosphere in which unpro-
tected steel corrodes very quickly. In order to ameliorate corrosion issues, engineers
have used protective coatings as one means of protecting steel from the impact of the
environment.

259

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


6.2.2
Description of Coating: Painting
In a cost-effective, multicoat paint system, the primary purpose of the coating layer
closest to the steel surface is to provide corrosion protection for the steel surface. Any
special aesthetic considerations are accommodated in the subsequent coating layers,
principally the topcoat. Aesthetics, while important for some applications, is not the
focus of this chapter.
The performance of protective coatings is dependent on the environment in which
they are exposed. In some dry climate areas of the country where corrosion is not an
issue, aesthetic considerations can play a more compelling role. The survey of state
departments of transportation (DOTs) conducted as part of SHRP 2 Project R19A
(final report available at http://www.trb.org/Design/Blurbs/168760.aspx) confirms
that a bridge system that could be expected to provide 50+ years of service life in
relatively dry states such as Arizona would last only 20 to 30 years in states with a
moister climate.
It follows, therefore, that a sure way to protect steel from corrosion is to keep
it from getting wet, and a key means of accomplishing this is by coating the steel to
provide a barrier to the elements, protect the steel from moisture, and keep it dry. The
ability of the topcoat, or outmost layer, to shed water is the key to using coating as
corrosion protection.
When water penetrates the outer coating layer(s) and comes in contact with the
steel substrate, the primer acts to inhibit corrosion as the steel surface is subjected to
repeated wet–dry cycles.
From the earliest years in the steel bridge era in the United States, beginning
around 1874 with the Eads Bridge in Saint Louis, Missouri (see Figure 6.2), lead and
chromium rust-inhibitive pigments were added to paint to supplement the ­barrier
protection offered by a coating film. For almost 100 years, the use of lead- and
­chromium-pigmented, multilayer coatings was the norm during new bridge construc-
tion, maintenance overcoating, and maintenance repainting.
After about 1965, bridge coating engineers began to turn away from coatings con-
taining these toxic heavy-metal pigments and instead started using coatings containing
metallic zinc as the corrosion-inhibitive pigment. The DOT survey conducted as part

Figure
Source: 6.2.  Eads KTA-Tator,
Courtesy Bridge. Inc.
Source: Courtesy KTA-Tator, Inc.
Figure 6.2. Eads Bridge.
260

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


of SHRP 2 Project R19A indicated that all states responding use a system consisting of
a zinc-rich primer. As long as the zinc pigment in the coating is in close metal-to-metal
contact with the steel substrate, the coating provides galvanic protection to the steel.
Galvanic protection is provided when zinc and steel (iron) are connected (i.e., have a
conductive pathway between them) in the presence of air (oxygen) and moisture. In
this coupling of materials, zinc (the less noble metal) will oxidize (corrode) in prefer-
ence to the iron (steel). The preferential oxidation of zinc provides protection for the
steel as long as there is nearby zinc left to be consumed in the chemical reaction, which
takes place at the anode. When the zinc is consumed, the steel beneath will be subject
to corrosion (oxidation). The method used to resist corrosion attack since about the
mid 1960s has been to use a multicoat, “belt and suspenders” approach. In a multi­
coat system, the outer layer or layers resist the effects of weather and protect the
zinc from being consumed in the atmosphere, while the zinc-rich primer inhibits cor-
rosion from occurring at the steel beneath in locations where the coating is breached.
Even in instances in which steel is painted with a coating system using a zinc-
rich primer, when the protected steel surface is bathed in salt water and is subjected
to many wet–dry cycles, discontinuities in the coating inevitably provide a pathway
through the coating for moisture to reach the zinc-coated steel surface beneath. As
a result, the zinc begins to react to protect the steel from corroding. Eventually, the
metallic zinc in the zinc-rich primer is consumed, and corrosion in the form of red rust
(iron oxide) results. The corrosion protection offered by the zinc may last many years
before there is evidence of corrosion of the steel; the rate of corrosion is dictated by
the local factors surrounding the steel (e.g., wet–dry cycles, chloride contamination,
humidity).
The following general discussion of coatings introduces the use of protective coat-
ings in the prevention of corrosion of structural steel. Figure 6.3 identifies the various
items discussed.
6.2.2.1
Paint Coating Composition
Figure 6.4 illustrates the basic ingredients of an industrial protective coating. The chart
divides a coating into two major components: pigmentation and vehicle. The pigmen-
tation typically consists of corrosion inhibitors, colorants, and extenders, although
other raw materials may also be included. The vehicle typically consists of the resin
or binder, solvents, and any additives that may be included in the formulation. The
vehicle may also contain other raw materials to provide additional or different perfor-
mance characteristics. The vehicle carries the pigmentation to the surface and binds it
into the coating film.
The ingredients can also be categorized as nonvolatile components and volatile
components, indicated in Figure 6.4 by (NV) and (V), respectively. Nonvolatile com-
ponents remain in the coating and on the surface once applied. Conversely, the volatile
components evaporate from the coating into the air once the coating is applied to the
surface. The nonvolatile components typically include the resin or binder, the pigmen-
tation, and any additives that may be incorporated into the formulation. The volatile

261

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


Bridge Coatings
General Considerations

Coating Components Curing Application


Coating Systems
and Ingredients Mechanisms Methods

Solvent Surface Conventional


Pigmentation
Evaporation Preparation Air Spray

Vehicle Coalescence Primer Airless Spray

Plural Component
Nonvolatile
Oxidation Intermediate Coat Spray

Volatile Electrostatic Spray


Polymerization Top Coat

Brushes, Daubers,
and Rollers
Moisture Cure

Figure 6.3.  Paint system general considerations.


Figure 6.3. Paint system general considerations.

component is the solvent system used in the formulation that is a component of the
wet film, but not the dry film, of the coating.
6.2.2.1.1
Vehicle Resin
The vehicle resin (or binder) portion of the coating vehicle comprises both volatile and
nonvolatile components. That is, it is both part of the wet film and the dry film. Often
a coating is identified generically by the type of resin used in the formulation. For ex-
ample, a two-coat epoxy is a commonly specified coating system. In this case, “epoxy”
is used to describe both the coating type and the raw material resin system used to
formulate the coating. The resin system is the film-forming component of a coating. It
cohesively bonds the pigmentation together and adhesively bonds the coating to the
underlying substrate or coating layer. It is essentially the glue of the coating. In many
cases, the resin system dictates the performance properties of a coating.
6.2.2.1.1a
Pigmentation
The pigment is also a nonvolatile component of the coating formulation and is es-
sentially an insoluble raw material. It suspends in the resin and solvent rather than
dissolving. Although some believe that the pigment merely gives the coating its color,
that is only one of several potential functions.

262

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Industrial Coatings

Pigment Vehicle

Inhibitors Colorants Extenders Resin Solvent Additives


(NV) (NV) (NV) (NV) (V) (V)

Zinc Red Clay Acrylic Thickeners

Primary Secondary
UV
Chromate White Silica Alkyd
Absorbers

Phosphate Blue Mica Oil Plasticizers

Borate Vinyl Catalysts

Wetting
Epoxy
Agents

Urethane

Siloxirane

Figure 6.4.  Industrial coating elements.


Source: Courtesy KTA-Tator, Inc.

263

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


The pigment in a coating may also provide corrosion protection. If used for this
purpose, the pigmentation must be formulated into the primer layer (the layer adjacent
to the steel substrate). Inhibitors like barium, phosphorous, and others formulated
into a primer inhibit the corrosion process. Zinc powder added to a primer in suffi-
cient quantities galvanically protects the underlying steel. Because of the shape of the
pigment particles, certain pigments even provide barrier protection; in other words,
their inherent shape and the way in which they orient themselves in the dry film create
a barrier to moisture penetration through the coating. Examples include micaceous
iron oxide and leafing aluminum pigments. These raw materials are lamellar, mean-
ing they are plate-like, tend to lie flat in the coating film, and cause any moisture that
penetrates the coating film to take a considerably longer pathway to the substrate.
Extenders such as silica, mica, and clay may be incorporated into the formulation to
improve film build, increase the solids content of the coating, and/or provide added
barrier protection.
6.2.2.1.1b
Additives
Additives formulated into the coating also become part of the dry film. Various quan-
tities of additives are used by the formulator to adjust the consistency, flow-out, sur-
face wetting, color, ultraviolet light (sunlight) resistance, and flexibility or to prevent
settling in the can (suspending agents). For example, an alkyd coating that typically
chalks and fades on exposure to sunlight can be formulated with silicone (minimum
30%) to provide better color and gloss retention characteristics. In this case, the sili-
cone is an additive. Polyurethane coatings are formulated with hindered amine light
stabilizers to help preserve gloss and color on exposure to sunlight, and plasticizers
formulated into a coating provide film flexibility.
6.2.2.1.1c
Solvents
The solvent system in a coating is the volatile component. Although the solvent sys-
tem is part of the wet film during application, it is not intended to be part of the dry
film once the coating dries or cures. This component is referred to as a solvent system
because it is uncommon for a coating to be formulated with just one type of solvent.
Typically, a blend of solvents is used, and each type of solvent in the blend may per-
form a different function. As a general rule, primary solvents are formulated into the
coating to reduce the viscosity of the resin, pigment, and additives so that the coating
can be properly atomized through a spray gun or applied by brush and roller. Second-
ary solvents typically stay in the wet coating film a little longer than the primary sol-
vents, as they are more slowly evaporating solvents that help the coating to flow out to
form a uniform, continuous film.
6.2.2.1.1d
Volatile Organic Compounds
Solvents have been used in coatings for many decades because they have been useful
and affordable. Many solvent systems in a coating (and thinners added to a coat-
ing by the applicator) are categorized as volatile organic compounds (VOCs) by the

264

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Environmental Protection Agency (EPA). The type and amount of solvent(s) used in
an industrial coating may be regulated by the EPA because as they evaporate from the
coating film into the air, they can photochemically react with sunlight and become a
precursor to ozone (a component of smog). As part of the Clean Air Act, federal and
state environmental agencies have developed regulations to control ozone-producing
operations. The amount of VOCs that can be legally emitted into the atmosphere
­varies considerably from location to location. For example, densely populated areas
such as Southern California and Houston, Texas, have very strict VOC regulations,
but less populated areas typically comply with the federal limit, which represents a
considerably higher threshold. California has led the nation in the march toward coat-
ings with ever-lower amounts of VOCs. Coatings suppliers have been reformulating
and retesting their coatings as VOC regulations have tightened, and eventually, the use
of such materials in coatings may diminish to the point that they are not a significant
part of coatings used on bridges.
The VOC content of a coating is expressed in pounds per gallon (or grams per
liter) and is reported on the manufacturer’s product data sheet. Many manufacturers
also recalculate the VOC content of a coating after the addition of thinner, and this
information is also commonly provided on the product data sheet.
When painting a structure in the field, the VOC limit is typically dictated by the
specification or the local air pollution agency for the project. Conversely, fixed facili-
ties such as painting shops are sometimes required to log the number of gallons of
paint used over a specific period (say 90 days) and the VOC content of each type.
6.2.2.2
Curing Mechanisms
The method in which a coating converts from a liquid to a solid state is known as the
curing mechanism. Many liquid-applied coatings dry by solvent evaporation, but cure
by employing a separate reaction.
6.2.2.2.1
Solvent Evaporation
Coatings that cure by solvent evaporation actually only dry. That is, the resin, pig-
ment, and additives are suspended in a solvent system. When the solvent evaporates
from the applied film into the air, the resin, pigment, and additives remain on the sur-
face. Because there is no subsequent curing reaction, the resin can be redissolved by the
solvent system that evaporated from the coating film. This lack of subsequent curing is
why a coating that cures by solvent evaporation should not be overcoated with a coat-
ing containing strong solvents, as such solvents may redissolve the underlying coating
film. A vinyl coating is an example of a coating that cures by solvent evaporation.
6.2.2.2.2
Coalescence
Waterborne acrylic coatings cure by solvent evaporation and form a coating film by
a process known as coalescence. Water, the primary solvent in these coatings, first
evaporates from the acrylic-containing emulsion coating film. As the water evaporates,
a special coalescing solvent (e.g., propylene glycol) aids in fusing the acrylic emulsion

265

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


particles to form a solid film. The coalescing solvent then evaporates from the coating
film. Without this coalescing solvent, the acrylic particles will not impinge and fuse
together, which can result in a poorly performing coating film. The coalescing process
typically requires a minimum 50°F air temperature. Should the air temperature fall be-
low 50°F before the coalescing process is complete, curing may stop and may not start
again once the temperature recovers. This major concern with industrial waterborne
acrylic coatings should be carefully considered by the specifier.
6.2.2.2.3
Oxidation
Coatings that cure by oxidation react with oxygen (air) to form a film. This oxidation
process never stops as long as the coating is exposed to oxygen. For example, long-
used alkyd coatings, which typically contain unsaturated oils, pigments, and driers,
cure by oxidation. Many aged alkyd systems, even those that have been in service for
decades, become very brittle, as the resin continues to oxidize long after the coating is
fully cured.
6.2.2.2.4
Polymerization
“Poly” means “many.” Many monomers or “mers” are used to create a polymer. These
monomers are formulated into components, and the components are packaged sepa-
rately by the coating manufacturer. It is only when these components are blended in
the correct proportions that a chemical reaction known as polymerization occurs, gen-
erating a very resilient coating layer. Coatings that cure by polymerization are multi­
component, typically packaged in two or three containers. Prior to application, these
components are blended in the correct ratio. Generally, only complete, pre­packaged
kits are blended. Once blended, the chemical reaction begins. Coatings that cure by
polymerization have a limited pot life. That is, the blended components must be ap-
plied before that pot life expires. The pot life will vary from a few minutes to several
hours, depending on the formulation and temperature of the coating. Many coatings
cure by polymerization; epoxy coatings and aliphatic acrylic or polyester polyurethane
coatings are a few of the more common types used on bridges.
6.2.2.2.5
Moisture Cure
Hydrolysis is the reaction of a coating with moisture in order to cure. Only a few
industrial coatings hydrolyze in the curing process. These include moisture-cure ure-
thanes and ethyl silicate–type inorganic zinc-rich primers, which require a minimum
amount of moisture to cure. In this process, moisture-cure urethanes release carbon
dioxide, and inorganic zinc-rich primers release ethyl alcohol. The moisture cure pro-
cess results in a very resilient coating layer, similar to that achieved by polymerization.
(Zinc coatings are further discussed in a subsequent section.)

266

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


6.2.2.3
Coating Systems Defined
In many cases, coatings can be combined to create a coating system, which includes
both the surface preparation and the application of one or more coating layers. If mul-
tiple coats of the same product are specified, contrasting colors are sometimes used to
help ensure the coverage of the applicator.
Although coating layers usually consist of a primer and topcoat, in many instances
an intermediate coat may also be specified. When multiple coatings are used to create
a system, they must be compatible. In addition, each coating layer has a function that
is performed at a given thickness. Accordingly, adding extra thickness of an epoxy
intermediate coat cannot make up for an inadequate zinc-rich primer thickness. Each
layer should be applied at the optimum thickness (i.e., neither too thick nor too thin)
and verified for proper thickness before the application of subsequent layers.
6.2.2.3.1
Surface Preparation
The Society for Protective Coatings (SSPC) and the National Association of Corrosion
Engineers International (NACE) have developed standard requirements that include
the end condition of the surface and materials and procedures necessary to achieve and
verify the end condition.
The level of surface preparation to be performed is an integral component to the
coating system. For example, there would be little point in applying a zinc-rich primer
to a marginally prepared surface, because the zinc must maintain intimate contact
with clean steel to provide galvanic protection. If zinc were to be applied over an old
coating, the desired galvanic protection would not develop. Conversely, applying a
surface-tolerant coating to a surface prepared to SSPC-SP 5/NACE No. 1, White Metal
Blast Cleaning, would be excessive, as equivalent performance could be achieved over
a much lesser degree of cleaning, usually at a much lower cost.
6.2.2.3.2
Primer Function
The function of the primer is to bond the coating system to the substrate. It also pro-
vides corrosion protection for the steel substrate by using one or more methods of
barrier, inhibitive, or galvanic protection. The primer must also be tolerant of the level
of surface preparation performed and must be compatible with the next layer applied.
If the primer is the only layer, as with a single-coat system, it must be resistant to the
service environment and provide corrosion protection to the steel beneath.
6.2.2.3.3
Intermediate Coat Function
An intermediate coat is typically incorporated into a coating system for the purpose of
adding barrier protection. It must be compatible with both the primer and the topcoat.

267

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


6.2.2.3.4
Topcoat Function
The topcoat, or finish coat, is the first line of defense in a corrosion-protection system. It
must also be aesthetically compatible with the specifier’s priorities and is often required
to maintain color and gloss levels for long periods of time. Naturally, the topcoat must
be resistant to the service environment and must be compatible with the underlying
layer (i.e., primer or intermediate coat, as appropriate). In addition, the topcoat must
be able to accept a maintenance overcoat.
6.2.2.4
Coating Application Methods
6.2.2.4.1
Conventional (Air) Spray
Conventional or air-atomized spray uses compressed air to transport the paint from a
pressurized pot to the spray gun in order to atomize the coating into a fine spray and
then propel the atomized coating to the surface. As the ratio of air to paint is quite high
(~600:1), compressed air cleanliness is critical, and transfer efficiency is relatively low.
The primary reason conventional (air) spray is used to apply industrial coatings is the
ability to precisely control the amount of paint that exits the spray gun and to control
the shape of the spray pattern. The apparatus used for the application of metalizing
spray is a special variation of a conventional spray gun.
Conventional air spray equipment consists of a pressure pot equipped with two
regulators. The first regulator is used to control the amount of pressure in the pot itself
(pot pressure), and the second is used to control the volume of atomization air that
is used to break up the stream of paint exiting the spray tip. Coating manufacturers
provide recommended pot and atomization pressures, which often have to be adjusted
slightly based on project conditions (e.g., amount of thinner addition, temperature,
and so forth). Two hoses connect the spray gun to the pot; one contains the paint and
the other contains the atomization air. The spray gun has two controls. The lower
control regulates the amount of paint that comes out of the spray tip, and essentially
adjusts how far the operator can pull back the spray gun trigger, which regulates the
amount of paint that exits the spray tip. The upper control regulates the shape of the
fan pattern, from a small circle for striping of corners and other small areas to a larger
oval for spraying flat surfaces. A conventional spray gun can be half-triggered—that
is, the trigger can be pulled back part way—so that the atomization air, without paint,
exits the spray nozzle. This compressed air can be used to perform a final blow-down
of the surface immediately before the coating application.
The conventional spray gun is held 6 to 10 in. from the surface, with variations in
distance dependent on the type of coating and prevailing spraying conditions, such as
air temperature and wind speed.

268

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


6.2.2.4.2
Airless Spray
Airless spray has long been the most common method used for applying bridge coat-
ings. If the equipment is operating properly and the applicator employs good spray
technique, the finish quality can approach that which is created by conventional
spray, but at much higher production levels. Because airless spray does not use com-
pressed air to atomize the coating, the transfer efficiency is relatively higher than
conventional spray, reducing airborne emissions. Further, because airless spray does
not incorporate compressed air into the paint stream, the cleanliness of the com-
pressed air is not critical.
Airless spray equipment consists of a paint pump that is operated using an air
compressor equipped with a regulator. Coating manufacturers provide recommended
airless spray pressures, which frequently have to be adjusted slightly based on project
conditions (e.g., amount of thinner addition and temperature). A single hose contain-
ing the paint connects the spray gun to the pump.
The airless spray gun is held 18 to 24 in. from the surface, with variations in dis-
tance dependent on the type of coating and prevailing spraying conditions.
6.2.2.4.3
Plural Component Spray
Plural component spray is used for coating materials with a relatively short pot life
or coating materials that do not contain viscosity-reducing solvents (e.g., 100% solids
materials). Plural component spray does not require premixing the coating compo-
nents. Rather, the individual components are pumped to a mixer or manifold at the
correct ratio. They are then mixed and delivered to the spray gun using a short mate-
rial hose that can be flushed with solvent in a solvent purge system. This is known as
an internal mix process. There are also external mix plural component systems that
send each component to the spray gun in separate material hoses. The components
blend as they exit the spray tip. It is common for the material hoses to be heated for
plural component spray in order to reduce the viscosity of the components and allow
for easier transport and improved atomization.
Plural component spray is available in two basic designs: fixed ratio and variable
ratio proportioning pumps. Fixed ratio pumps can only proportion the components in
a set ratio (e.g., 1:1), but a variable ratio pump can proportion the materials according
to the required ratio (e.g., 2:1, 4:1, 8:1, 16:1). Plural component spray equipment can
be complex and typically requires a technician to set up the equipment and monitor
the mix ratio so that the coating materials are not applied off-ratio. This equipment is
set up to spray similar to airless spray equipment.
6.2.2.4.4
Electrostatic Spray
Electrostatic spray is sometimes used to apply bridge coatings. Because of its poten-
tially high transfer efficiency rate and the resulting reduction in material usage, it can
be an attractive application method. In principle, the paint particles are energized (+)
as they exit the spray gun. The electrical charge is imparted by a small wire protrud-
ing from the spray nozzle. The surface to be coated is grounded (–). The particles are
269

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


“attracted” to the component or part having the opposing electrical charge, signifi-
cantly reducing overspray and material usage. The coating, however, must be able to
accept an electrical charge, and the addition of a polar solvent is sometimes required
to inhibit the coating’s natural resistivity. Electrostatic spray can also result in a wrap-
ping of the coating on the opposite side of the direction of spray, making it an attrac-
tive alternative for coating inaccessible areas. Because of the difficulty in achieving a
uniform ground on large structures, electrostatic spray has not been widely used to
coat bridges. Electrostatic spray equipment is typically set up to spray as an airless
operation, and electrostatic spray can be used to coat smaller members and hard-to-
reach areas.
6.2.2.4.5
Brushes, Rollers, and Daubers
The use of brushes for bridge projects is typically limited to striping, the application
of a layer of coating to surfaces where it is difficult to achieve a normal film build. For
the same reason, brushes are also used on pitted or rough surfaces around rivet heads,
welds, and bolt-and-nut assemblies and to cut-in inside and outside corners. Daubers
are often used to coat surfaces within crevices like back-to-back angles. Rollers have
high coating-transfer efficiency and can be used to coat large flat surfaces with limita-
tions: roller nap can become embedded in the dry coating film and act like a wick to
pull moisture into the coatings, and film thickness is hard to control with a roller.
Choosing the method of coating application depends on many factors, including
the size and configuration of the surfaces to be coated, the proximity to other struc-
tures, environmental regulations, and the specification and coating manufacturer’s rec-
ommendations. Although airless spray is no doubt the most frequently used method,
there are other issues for the contractor or applicator to consider when choosing an
application method, including speed and control.
6.2.3
Description of Coating: Hot-Dip Galvanizing
The material in this section is from the American Galvanizers Association (www.­
galvanizeit.org) and adapted from AGA (2006). Hot-dip galvanizing (HDG) is a pro-
cess in which fabricated steel, structural steel, castings, or small parts, including fas-
teners, are immersed in a kettle or vat of molten zinc, resulting in a metallurgically
bonded alloy coating that protects the steel from corrosion. HDG is often referred to
as simply “galvanizing,” a term that is often used incorrectly to describe steel coated
with zinc by other methods such as paint or plating. These other methods of applying
zinc to steel for corrosion protection are very different from HDG.
The AGA HDG Specifier’s Guide (2006) states, “Galvanizing forms a metallurgical
bond between the zinc and the underlying steel or iron, creating a barrier that is part of
the metal itself. During galvanizing, the molten zinc reacts with the surface of the steel or
iron article to form a series of zinc–iron alloy layers. [Figure 6.5] is a photomicrograph
of a galvanized steel coating’s cross section and shows a typical coating microstructure.”
The HDG coating consists of four layers. The first three layers above base steel
have a mixture of iron and zinc, and the external top layer is typically composed of
100% zinc.
270

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 6.5.  Magnified cross section of galvanized coating.
Source: Photo courtesy of the American ­Galvanizers Association.

The Specifier’s Guide continues:

Below the name of each layer in [Figure 6.5] its respective hardness, expressed
by a diamond pyramid number (DPN), appears. The DPN is a progressive mea-
sure of hardness (i.e., the higher the number, the greater the hardness). Typi-
cally, the Gamma, Delta, and Zeta layers are harder than the underlying steel.
The hardness of these inner layers provides exceptional protection against coat-
ing damage through abrasion. The Eta layer of the galvanized coating is quite
ductile, providing some resistance to impact. The galvanized coating is adher-
ent to the underlying steel on the order of several thousand pounds per square
inch. . . . Hardness, ductility, and adherence combine to provide the galvanized
coating with unmatched protection against damage caused by rough handling
during transportation to and/or at the project site, as well as in service. The
toughness of the galvanized coating is extremely important since barrier pro-
tection is dependent upon the integrity of the coating.

6.2.3.1
Hot-Dip Galvanizing Process
Though the process may vary slightly from plant to plant, the fundamental steps in the
galvanizing process are surface preparation, galvanizing, and finishing.
6.2.3.1.1
Surface Preparation
Degreasing/Caustic Cleaning. A hot alkaline solution removes dirt, oil, grease, shop
oil, and soluble markings.

271

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


Pickling. Dilute solutions of either hydrochloric or sulfuric acid remove surface
rust and mill scale to provide a chemically clean metallic surface.
Fluxing. Steel is immersed in liquid flux, usually a zinc ammonium chloride solu-
tion, to remove oxides and prevent oxidation before the item is dipped into the bath
of molten zinc. In the dry galvanizing process, the item is separately dipped in a liq-
uid flux bath, removed, allowed to dry, and then galvanized. In the wet galvanizing
process, the flux floats on top of the molten zinc and the item passes through the flux
immediately before galvanizing.
6.2.3.1.2
Galvanizing
The article is immersed in a bath of molten zinc between 815°F to 850°F (435°C to
455°C). During galvanizing, the zinc metallurgically bonds to the steel, creating a se-
ries of highly abrasion-resistant zinc–iron alloy layers, commonly topped by a layer of
impact-resistant pure zinc.
6.2.3.1.3
Finishing
After the steel is withdrawn from the galvanizing bath, excess zinc is removed by
draining, vibrating, or, in the case of small items, centrifuging. The galvanized item is
then air cooled or quenched in liquid.
6.2.3.2
Coating Uniformity
The AGA HDG Specifier’s Guide (2006) states

The galvanizing process naturally produces coatings that are at least as thick
at the corners and edges as the coating on the rest of the substrate. As coating
damage is most likely to occur at the edges, this is where added protection is
needed most. [Figure 6.6] is a photomicrograph showing a cross section of an
edge of a galvanized piece of steel.

Because the galvanizing process involves total immersion of the material, . . .


all surfaces are coated. . . . Galvanizing provides protection on both exterior
and interior surfaces of hollow structures. . . . Hollow structures must be
detailed in a way that allows zinc to drain from the interior when the item is
removed from the kettle.

The inspection process for galvanized items is simple, fast, and requires mini-
mal labor. . . . Galvanizing continues at a factory under any weather or hu-
midity conditions. . . . The galvanizer’s ability to work in any type of weather
allows a higher degree of assurance of on-time delivery. . . . A turnaround time
of two or three days is common for galvanizing.

ASTM, ISO, CSA (the Canadian Specification Association), and AASHTO


specifications establish minimum standards for thickness of galvanized coat-
ings on various categories of items. These minimum standards are routinely
exceeded by galvanizers due to the nature of the galvanizing process. Factors

272

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 6.6.  Cross section of corner
of galvanized steel section.
Source: Courtesy American G
­ alvanizers
Association.

influencing the thickness and appearance of the galvanized coating include


chemical composition of the steel, steel surface condition, cold working of
steel before galvanizing, bath temperature, bath immersion time, bath with-
drawal rate, and steel cooling rate.

6.2.3.3
Effect of Amount of Silicon in Steel on Galvanized Coating
The Specifier’s Guide (AGA 2006) explains this effect:

The chemical composition of the steel being galvanized is perhaps the most
important, [influencing factor]. The amount of silicon and phosphorus in the
steel strongly influences the thickness and appearance of the galvanized coat-
ing. Silicon, phosphorus, or combinations of the two elements can cause thick,
brittle galvanized coatings. The coating thickness curve shown in [Figure 6.7]
relates the effect of silicon in the base steel to the thickness of the zinc coating.
The carbon, sulfur, and manganese content of the steel also may have a minor
effect on the galvanized coating thickness.

The combination of elements mentioned above, known as “reactive steel”


in the galvanizing industry, tends to accelerate the growth of zinc–iron alloy
layers. This may result in a finished galvanized coating consisting entirely of
zinc–iron alloy. Instead of a shiny appearance, the galvanized coating will have
a dark gray, matte finish. This dark gray, matte coating will provide as much
corrosion protection as a galvanized coating having a bright appearance.

It is difficult to provide precise guidance in the area of steel selection [for gal-
vanizing] without qualifying all of the grades of steel commercially available.
[However,] the guidelines discussed below usually result in the selection of
273

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


Figure 6.7.  Galvanized coating thickness curve.
Source: Courtesy American Galvanizers Association.

steels that provide good galvanized coatings: Levels of carbon less than 0.25%,
phosphorus less than 0.04%, or manganese less than 1.35% are beneficial;
Silicon levels less than 0.04% or between 0.15% and 0.22% are desirable.

Silicon may be present in many steels commonly galvanized even though it


is not part of the controlled composition of the steel. This occurs primarily
because silicon is used in the deoxidization process in for the steel [and is
found in continuously cast steel]. The phosphorus content should never be
greater than 0.04% in steel [that is] intended for galvanizing. Phosphorus acts
as a catalyst during galvanizing, resulting in rapid growth of the zinc–iron
alloy layers. [This growth is virtually uncontrollable during the galvanizing
process.]

Because the galvanizing reaction is a diffusion process, higher zinc bath


temperatures and longer immersion times will generally produce somewhat
heavier alloy layers. Like all diffusion processes, the reaction proceeds rapidly
at first and then slows as layers grow and become thicker. [However,] contin-
ued immersion beyond a certain time will have little effect on further coating
growth. When galvanizing reactive steels, the diffusion process significantly
changes [and proceeds at a faster rate, producing thicker coatings].

The thickness of the outer pure zinc layer is largely dependent upon the rate of
withdrawal from the zinc bath. A rapid rate of withdrawal causes an article to
carry out more zinc and generally results in a thicker coating.

274

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


ASTM, CSA, and AASHTO specifications and inspection standards for galva-
nizing recognize that variations occur in both coating thickness and composi-
tions. Thickness specifications are stated in average terms. Further, coating
thickness measures must be taken at several points on each inspected article to
comply with ASTM A123/A123M for structural steel and A153/A153M for
hardware.

Figure 6.8 shows a thickness measurement being taken.


Fortunately, many grades of steel commonly used in steel bridges meet the chemi-
cal requirements and are readily galvanized. When in doubt the owners or engineers
should be queried, or the galvanizer’s advice should be sought.
6.2.3.4
Inspection
Inspections for coating thickness and surface condition complete the process. Inspec-
tion of structural steel will normally fall under ASTM A123/A123M–12, Standard
Specification for Zinc (Hot-Dip Galvanized) Coatings on Iron and Steel Products, or
ASTM A153/A153M–09, Standard Specification for Zinc Coating (Hot-Dip) on Iron
and Steel Hardware.
6.2.4
Description of Coating:
Thermally Sprayed Metal Coating or Metalizing
The material in this section was adapted from Ellor et al. (2004). When sizes and
shapes of steel members will not fit in a galvanizing kettle or when the schedule simply
does not allow time to transport items to a galvanizer’s plant, there is the option to
metalize the item.

Figure 6.8.  Coating thickness


measuring.
Source: Courtesy American
­Galvanizers Association.

275

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


Thermally sprayed metal coating (TSMC), referred to as metalizing, is the process
of applying metallic zinc in wire form to clean steel by feeding it into a heated gun,
where it is heated, melted, and spray applied by using combustion gases or auxiliary
compressed air to provide ample velocity. Metalizing may be used on any size steel
object, which eliminates limitations due to vat size and awkward shapes. Applying a
consistent coating in recesses, hollows, and cavities adds a measure of complexity. Pure
zinc can be used, but often zinc is alloyed with 15% aluminum to provide a smoother
abrasion-resistant film.
6.2.4.1
Thermally Sprayed Metal Coating Processes
Two similar processes are used to apply the metallic zinc to the steel surface. These
processes, which are differentiated by the manner in which the zinc metal is melted, are
described in the following sections, adapted from Ellor et al. (2004).
6.2.4.1.1
Flame Spray Process

The flame spray process can be used to apply a wide variety of feedstock
materials, including metal wires, ceramic rods and metallic and nonmetallic
powders. In flame spraying, the feedstock material is fed continuously into the
tip of the spray gun or torch, where it is then heated and melted in a fuel gas/
oxygen flame and accelerated toward the substrate being coated in a stream
of atomizing gas. Common fuel gases used include acetylene, propane, and
methyl acetylene–propadiene (MAPP). Oxyacetylene flames are used exten-
sively for wire-flame spraying because of the degree of control and the higher
temperatures attainable with these gases. The lower-temperature oxygen–pro-
pane flame can be used for melting metals such as aluminum and zinc, as well
as polymer feedstock. The basic components of a flame spray system include
the flame spray gun or torch, the feedstock material and a feeding mechanism,
oxygen and fuel gases with flow meters and pressure regulators, and an air
compressor and regulator.

With wire-flame spraying, the wire-flame spray gun or torch [shown in Figure
6.9] consists of a drive unit with motor and drive rollers for feeding the wire
and a gas head with valves, gas nozzle, and an air cap that controls the flame
and atomization air. Compared with wire-arc spraying, wire-flame spraying
is generally slower and more costly because of the relatively high cost of the
oxygen-fuel gas mixture compared with the cost of electricity. However, flame
spraying systems are generally simpler and less expensive than wire-arc spray-
ing systems. Both flame spraying and wire-arc spraying systems are field porta-
ble and may be used to apply quality metal coatings for corrosion protection.

276

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 6.9.  Schematic of a typical
flame-wire spray gun.
Source: Ellor et al. 2004.

6.2.4.1.2
Wire-Arc Process

Due to its high deposition rates, excellent adhesion, and cost-effectiveness,


wire-arc spray is the preferred process for applying TSMCs to steel [bridges].
In the wire-arc spray process, two consumable wire electrodes of the metal be-
ing sprayed are fed into a gun such that they meet at a point located within an
atomizing air, or other gas, stream. An applied direct current (DC) potential
difference between the wires establishes an electric arc between the wires that
melts their tips. The atomizing air flow subsequently shears and atomizes the
molten droplets to generate a spray pattern of molten metal directed toward
the substrate being coated. Wire-arc spray is the only thermal spray process
that directly heats the material being sprayed, a factor that contributes to its
high energy efficiency.

The wire-arc spray system consists of a wire-arc spray gun or torch [shown in
Figure 6.10], atomizing gas, flow meter or pressure gauge, a compressed air
supply, DC power supply, wire guides/hoses, and a wire feed control unit. Op-
eration of this equipment must be in strict compliance with the manufacturer’s
instructions and guidelines.

6.2.4.2
Thermally Sprayed Metal Coating Guidelines
Table 6.1 provides a TSMC selection guide for 20- to 40-year life and shows the
TSMC thickness typically applied under various environmental conditions.

277

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


Figure 6.10.  Schematic of a typical wire-arc spray gun.
Source: Ellor et al. 2004.

Table 6.1. TSMC Coating Guide


Thickness
Environment Coating mil (µm) Sealer
Atmospheric
Rural Zinc or zinc–aluminum 6–8 (150–200) No
Industrial Zinc or zinc–aluminum 12–15 (305–308) Yes
Marine Aluminum or zinc–aluminum 12–15 (305–308) No
Immersion
Fresh water Zinc–aluminum 12–15 (305–308) Yes
Brackish water Aluminum 12–15 (305–308) No
Seawater Aluminum 12–15 (305–308) No
Alternate Wet–Dry
Fresh water Zinc–aluminum 10–12 (250–305) Yes
Seawater Aluminum 12–15 (305–308) Yes
Abrasion Zinc–aluminum 14–16 (355–405) Yes
Condensation Zinc or zinc–aluminum 10–12 (250–305) Yes

Source: Ellor et al. 2004.

Again quoting from Ellor et al.,

TSMCs should always be applied to “white” metal (SSPC-SP 5/NACE # 1,


[White Metal Blast Cleaning]). It is common practice in fieldwork to apply
the TSMC during the same work shift in which the final blast cleaning is
performed. The logical endpoint of the holding period is when the surface

278

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


cleanliness degrades or a change on performance (as per bend or tensile test)
occurs. If the holding period is exceeded, the surface must be re-blasted to
establish the correct surface cleanliness and profile.

Thermal spraying should be started as soon as possible after the final anchor
tooth or brush blasting and completed within 6 hours for steel substrates sub-
ject to the temperature to dew point and holding period variations. In high-
humidity and damp environments, shorter holding periods should be used.

In low humidity environments, or in controlled environments with enclosed


structures using industrial dehumidification equipment, it may be possible
to retard the oxidation of the steel and hold the near-white-metal finish for
more than 6 hours. With the concurrence of the purchaser, a holding ­period
of greater than 6 hours can be validated by determining the acceptable
­temperature–­humidity envelope for the work enclosure by spraying and ana-
lyzing bend test coupons, or tensile adhesion coupons, or both. Should the
sample fail the bend test, the work must be re-blasted and re-tested.

When specified, . . . a flash coat of TSMC equal to or greater than 1 mil


(25 μm) may be applied within 6 hours of completing the surface preparation
in order to extend the holding period for up to 4 hours beyond the applica-
tion of the flash coat. The final TSMC thickness, however, should be sprayed
within 4 hours of the application of the flash coat. This procedure should be
validated using a tensile adhesion test, or bend test, or both, by spraying a
flash coat and waiting through the delay period before applying the final coat-
ing thickness.

For small and movable parts, if more than 15 minutes is expected to lapse
between surface preparation and the start of thermal spraying, or if the part
is moved to another location, the prepared surface should be protected from
moisture, contamination, and finger/hand marks. Wrapping the part with
clean, print-free paper is normally adequate.

If rust bloom, blistering, or a degraded coating appears at any time during the
application of the TSMC, the following procedure should be performed:

1. Stop spraying.
2. Mark off the satisfactorily sprayed area.
3. Repair the unsatisfactory coating (i.e., remove the degraded coating and reestab-
lish the minimum “white metal” finish and anchor-tooth profile depth as per the
maintenance and repair procedure).
4. Record the actions taken to resume the project in the project documentation.
5. Contact the coating inspector to observe and report the remedial action to the
purchaser.

279

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


The materials and application costs of a TSMC system are higher than the cost of
conventional liquid-applied coatings; however, the dominant factor may not be either
material or application labor cost. Instead, the cost of taking the facility out of service,
contractor mobilization, environmental constraints, and monitoring costs are often
the major considerations of the total cost. In some complex coating work, the actual
cost of materials and application is less than 20% of the total process. Therefore, if
service life is increased by the use of TSMC, the process can pay for itself. Neverthe-
less, TSMC is used most effectively on broad flat surfaces, and complexities occur
when the gun or hoses are maneuvered around elements that are mounted at an angle
to the main flat surface. For example, TSMC is most efficient when used on a girder
web or flange, but progress may be slower when connection plates or stiffeners are
encountered. The current premium for TSMC can be as little as 40% compared with
other protective treatments. In many cases, this is a small price to pay for a material
that has proven performance and an estimated service life measured in decades.
Conditions that can damage a TSMC system should also be considered as part of
the overall coating selection process. According to Ellor et al. (2004),

Impact and abrasion are significant environmental stresses for any coating sys-
tem. Abrasion is primarily a wear-induced failure caused by contact of a solid
material with the coating. Examples include foot and vehicular traffic on floor
coatings, ropes attached to mooring bitts, sand particles suspended in water,
and floating ice. When objects of significant mass and velocity move in a direc-
tion normal to the surface as opposed to parallel, as in the case of abrasion,
the stress is considered to be an impact. Abrasion damage occurs over a period
of time, whereas impact damage is typically immediate and discrete. Many
coating properties are important to the resistance of impact and abrasion,
including adhesion to the substrate, cohesion within the coating layers, tough-
ness, ductility, and hardness. Thermally sprayed coatings of zinc, aluminum,
and their alloys are very impact resistant. Zinc metalizing has only fair abra-
sion resistance in immersion applications because the coating forms a weakly
adherent layer of zinc oxide. This layer is readily abraded, which exposes
more zinc, which in turn oxidizes and is abraded; 85:15 wt% zinc/aluminum
is more impact/abrasion resistant than pure zinc or pure aluminum.

6.2.4.3
Concerns Related to Performance of Thermally Sprayed Metal Coating
Coating selection may be limited by the degree or type of surface prepara-
tion that can be achieved on a particular structure or structural component.
Because of physical configuration or proximity to other sensitive equipment
or machinery, it may not always be possible to abrasive blast a steel substrate.
In such cases, other types of surface preparation, such as hand tool or power
tool cleaning, may be necessary, which, in turn, may place limits on the type
of coatings that may be used. In some cases, it may be necessary to remove the
old coating by means other than abrasive blasting, such as using power tools,

280

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


high-pressure water jetting, or chemical strippers. These surface preparation
methods do not impart the surface profile that is needed for some types of
coatings to perform well. In the case of thermally sprayed coatings, a high
degree of surface preparation is essential. This kind of preparation can only be
achieved by abrasive blast cleaning using a good-quality, properly sized angu-
lar blast media. Thermal spraying should never be selected for applications in
which it is not possible to provide the highest-quality surface preparation. . . .

Angular blast media must always be used. Rounded media such as steel
shot, or mixtures of round and angular media, will not produce the appropri-
ate ­degree of angularity and roughness in the blast profile. The adhesion of
TSMCs can vary by an order of magnitude as a function of surface rough-
ness profile shape and depth. TSMCs adhere poorly to substrates prepared
with rounded media and may fail in service by spontaneous delamination.
Hard, dense, angular blast media such as aluminum oxide, silicon carbide,
iron ­oxide, and angular steel grit are needed to achieve the depth and shape of
blast profile necessary for good TSMC adhesion. Steel grit should be manu-
factured from crushed steel shot conforming to SAEJ827. Steel grit media
composed of irregularly shaped particles or mixtures of irregular and angular
particles should never be used. Steel grit having a classification of very angu-
lar, angular, or subangular . . . by the American Geological Institute should
be used [Hansink 1994].

For additional information, see Joint Standard SSPC-CS 23.00/AWS C2.23M/


NACE No. 12, which is discussed in Specification for the Application of Thermal
Spray Coatings (Metallizing) of Aluminum, Zinc, and Their Alloys and Composites
for the Corrosion Protection of Steel (SSPC 2003).
6.2.5
Description of Corrosion-Resistant Steels: ASTM A1010
ASTM A1010 steel is a 10.5% to 12.5% chromium structural steel with superior
corrosion resistance when compared with traditional weathering or galvanized steels
(Fletcher et al. 2005). A1010 is widely used in thicknesses from 1/8 to ½ in. for struc-
tures subjected to aggressive service conditions, such as coal railcars and coal-process-
ing equipment. Because of its superior corrosion resistance, A1010 is also a candidate
for challenging bridge applications. The steel can meet the strength and impact proper-
ties of AASHTO M270 Grades 50W and HPS 50W up to a thickness of 4 in., making
it an attractive steel for traditional plate girder bridges.
6.2.6
Description of Weathering Steels
Uncoated weathering-grade steels contain small amounts of copper, phosphorus, chro-
mium, nickel, and silicon to attain their weathering or corrosion-resistant properties.
Weathering-grade steels are currently supplied under AASHTO Specification
M270, ASTM A709 Grade 50W, and in high-performance steel grades HPS 50W, 70W,
and 100W. When used in the right environment, these steels are very cost-effective in
both the short and long term as they eliminate the need for shop and field painting.
281

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


Weathering steels have been successfully used on coal hopper cars, buildings, and
electric transmission towers and began appearing in bridges on a large scale in the mid
1960s (McEleney 2005). Currently, thousands of weathering steel bridges are provid-
ing trouble-free service across the United States.
According to Bill McEleney in A Primer on Weathering Steel (2005),

unpainted weather steel, properly designed and detailed, can realize bridge life
cycles up to 120 years with minimal maintenance. This high-strength, low-
alloy steel forms a tightly adhering “patina” during its initial exposure to
the elements. The patina is essentially an oxide film of corrosion by-products
about the same thickness as a heavy coat of paint.

The initial corrosion of weather steel depends on the presence of mois-


ture and oxygen. As corrosion continues, a protective barrier layer forms that
greatly reduces further access to oxygen, moisture, and contaminants. This
stable barrier layer greatly resists further corrosion, reducing it to a low value.
Under appropriate conditions, weather steel will generally corrode at a rate
of less than 0.3 mils per year. Corrosion of conventional steels, on the other
hand, forms rust layers that eventually dis­engage from the surface, exposing
“fresh” metal below, thereby continuing the corrosion cycle. . . .

Bridges constructed of weathering steel in suitable environments, and with


proper detailing, have all the qualities of conventional steel, plus they offer the
following benefits:

• Initial cost savings compared to conventional painted alternatives;


• Low maintenance consisting of periodic inspection and cleaning, which
reduces direct operating costs;
• Minimal indirect costs from traffic delays for major maintenance
operations;
• Faster construction resulting from elimination of shop and field painting;
• Good aesthetics, since weathering steel bridges eventually achieve an
­attractive dark brown color that blends well with the environment and
improves with age;
• Low impact on the environment, compared to painted alternatives that
emit undesirable volatile organic compounds (VOCs);
• Minimal health and safety issues relating to initial and future painting;
• A good track record for long-term performance based on various state
and federal studies.

Certain environments with high moisture, salt, or pollution levels can have unde-
sirable effects on the performance of weathering steel and can inhibit the proper devel-
opment of the protective patina. These environments include

282

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


• Locations of continual or persistent rainfall (or wetting), fog, or high humidity,
or where heavy, close vegetation can hold moisture against the steel and prevent
drying.
• Locations that have contact with salt-laden roadway drainage such as below leak-
ing deck joints.
• Locations subject to airborne salt water spray such as in marine coastal areas.
• Locations that create a tunnel effect such as with recessed, minimum clearance
under­passes with close, high abutments and where deicing salts are used fre-
quently. In these situations, rising salt water spray from the lower roadway can be
deposited on the girders above.
• Locations that contain high concentrations of pollution and industrial fumes,
­especially those containing sulfur dioxide.

Poor detailing can also have detrimental effects on weathering steel performance.
Steel detailing should permit all parts of the steelwork to dry, avoid moisture and debris
retention, and promote adequate ventilation (McEleney 2005). The FHWA Technical
Advisory T-5140.22, Uncoated Weathering Steel in Structures (FHWA 1989) provides
the following recommendations for detailing bridges that contain weathering steel:

• Eliminate bridge joints where possible through use of continuous girders


and integral abutments.
• Control water on the deck near the expansion joints deck. Consider the
use of a trough under the deck joint to divert water away from vulnerable
elements.
• Paint all superstructure steel within a distance of 1½ times the depth of
girder from bridge joints.
• Locate welded drip bars in areas of low stress.
• Minimize the number of bridge-deck scuppers (holes cut near the edge of a
deck to drain water below). Fewer scuppers result in a higher amount of
flow through each, minimizing the chance for blockage.
• Eliminate geometries that serve as water and debris “traps.”
• “Hermetically seal” box members when possible, or provide weep holes
to allow proper drainage and circulation of air.
• Cover or screen all openings in boxes that are not sealed.
• Consider protecting pier caps and abutment walls with drip pans and
plates to minimize staining.

Proper inspection and maintenance are also necessary for achieving desired weath-
ering steel performance. “Inspectors should specifically look for leaking expansion
joints, blocked drains, buildups of debris and other moisture traps, sealant failure, and
bulging joints and overlaps,” according to McEleney (2005). Any of these conditions,
if found, should be addressed with appropriate maintenance.

283

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


McEleney further summarizes research performed by the TxDOT that recom-
mends the following periodic measures:

• Flush debris, dirt, and bird and bat droppings from the bridge structure.
• Clear vegetation from pier and abutment areas to enhance air circulation.
• Reseal deteriorating joints.
• Unblock drains and troughs.

6.3
Factors Adversely Affecting Service Life
6.3.1
General Discussion
One of the most important tasks for developing corrosion-prevention systems is prop-
erly identifying the prevailing service environment, for existing structures, or the pro-
jected service environment, for new structures. To what will the system and the bridge
be subjected? Whether the structure already exists or is being planned, answering this
question can be a challenge. Service environments can be both predictable (e.g., deic-
ing salt exposure on a bridge in the winter) and unpredictable (e.g., hurricanes and
other like storms may bring unexpected conditions). Figure 6.11 shows some of the
factors that can influence the service life of steel bridge elements related to corrosion;
it is followed by a brief discussion of these factors.

Environmental Factors

Detail Exposure Moisture


Temperature
Exposure Type Type

Coastal Particle Nearby Outside


Regions Impact Industries Products

Figure 6.11.  Factors that can influence service life of steel bridge elements related to
corrosion.

284

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Detail exposure. Effects will vary for details having interior versus exterior expo-
sure. In cases of interior exposure, entire structures or parts of structures are sheltered
and exposed to a less aggressive environment. Coatings on the interior beams or the
interior of box members are examples.
Exposure type. Effects will vary for atmospheric versus immersion-like exposure.
In cases of immersion-like exposure, it must be determined whether the exposure is
constant or intermittent (i.e., splash) or from condensation.
Moisture type. For immersion-like exposure, the medium must be considered (i.e.,
fresh water, salt water).
Temperature. Normal operating and extreme conditions must be considered.
Coastal regions. Prevailing environment (i.e., coastal airborne sea saltwater mist)
must be considered.
Particle impact. The likelihood and type of physical damage (i.e., impact damage
from traffic or traffic-propelled debris) must be considered.
Nearby industries. Surrounding operations (e.g., an adjacent chemical plant) must
be considered.
Outside products. The type and concentration of product that will be stored or
transported in tank cars or vessels over or beneath the structure must be considered.
The specifier should consider these and other likely potential environments before
selecting a coating system. Any coating manufacturer will almost certainly request this
type of information before recommending a coating system. Note also that there may
be multiple service environments for a given structure, and interviewing nearby facility
owners and plant maintenance personnel may provide added insight into the actual
service environment that may be less than obvious.
6.3.2
Factors Affecting Service Life of Steel Bridge Elements Specific to
Paint Coating
Figure 6.12 shows factors that affect the service life of steel bridge elements specific to
paint coating. The factors are then described further.
6.3.2.1
Moisture and Debris Traps
The creation of moisture and debris traps in new structures is an area of obvious con-
cern, as the presence of such areas will certainly shorten the service life of any organic
coating system. The design of new structures should focus on eliminating the creation
of these corrosion-prone conditions. If such areas are absolutely essential to the design,
corrosion-mitigation strategies must be developed.
During maintenance recoating or overcoating projects on existing structures, the
consideration of debris and moisture traps is even more critical. If residual contaminants
remain on the surface and are overcoated or recoated in any area, the service life of virtu-
ally any system will be shortened. This effect is especially exacerbated by the presence of
chloride-laden residue from seawater or snow and ice removal activities. The service life
of any coating will be extended by extra cleaning efforts in these moisture trap locations.
In addition, owners should consider the use of zinc spray metalizing in these areas, mak-
ing use of the best protection in the most aggressive corrosion locations.
285

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


Factors Affecting Paint
Coatings

Moisture
Splash-
and Roadway Deck
Zone
Debris Joints Cracks
Exposure
Traps

Wind-
Exposed Back-to-
Blown
Steel Back Chlorides
Rain or
Bearings Angles
Salt

Figure 6.12.  Factors affecting service life of steel bridge elements related to paint
coatings.

As a result of this type of exposure, a surface that is normally expected to be dry


is in effect an area of severe exposure. These areas are categorized by the SSPC as
SSPC Category 2A (frequently wet by fresh water), 2B (frequently wet by salt water),
2C (fresh water immersion), or 2D (saltwater immersion). Coatings for these areas
should be chosen with care. SSPC currently recommends that zinc-rich, primer-based
­materials be used.
6.3.2.2
Roadway Joints
The design of roadway joints is discussed in Chapter 9. In the past, leaky expansion
joint seals were one of the principle reasons that steel below bridge decks became wet
and corroded. These leaks allow water from the bridge deck to cascade from the deck
and pour onto the steel members beneath the deck. These leaks change the exposure
conditions in such areas from an exposure zone that is designed to be dry (exposure
Zone 1B exterior, normally dry) to one of the following: Zone 2A (frequently wet by
fresh water), 2B (frequently wet by salt water), 2C (fresh water immersion), or 2D
(saltwater immersion). The type of corrosion protection used in these areas is described
in Section 6.2. See also the discussion about composite protection in Section 6.5.1.6.
6.3.2.3
Deck Cracks
Cracks in the deck and/or in the wearing surface allow water, especially salt water, to
penetrate through the deck and pour salt water onto steel surfaces below. When the
steel becomes wet, corrosion almost always follows.

286

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


6.3.2.4
Splash-Zone Exposure
When traffic travels beneath a steel overpass or through a steel truss or similar struc-
ture, an area of the steel above and often beside the traffic is bathed in water from the
roadway. This water is deposited on the steel surfaces. Figure 6.13 shows the effect
of roadway splash on coated railings. Splash from automobiles and trucks can travel
vertically as high as about 20 ft and horizontally 10 ft or more. Painted steel surfaces
within that envelope will be in an environmental Zone 2A, 2B, 2C, or 2D (as previ-
ously described) and perhaps in a zone requiring special treatment with zinc spray
metalizing when possible, galvanizing any steel that is replaced and any steel that can
be removed, galvanized, and returned to service.
6.3.2.5
Exposed Steel Bearings
Steel bearings are often sheltered and isolated from water or salt water by the steel and
roadway deck directly above, as shown in Figure 6.14. On some structures, steel bear-
ings can be the target of corrosion. At times bridges must be closed to traffic and
entire bearings must be replaced, which can require extensive shoring, as shown in
Figure 6.15. Steel bearings can be exposed to an immersion-like environment as a re-
sult of leaks from joint areas above.
6.3.2.6
Back-to-Back Angles
The use of back-to-back angles should be discontinued when new or replacement steel
configurations are encountered.
In rehabilitation maintenance, the use of back-to-back angles presents a configura-
tion that is very difficult to clean or coat effectively; great care in cleaning and coating
such areas is recommended. Special tools, often developed by the contractor for these
special situations, are needed for effective protection, and even with their use effective
protection is usually unattainable.

Figure 6.13.  Effect of roadway


splash on coated railing.

287

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


Figure 6.14.  Bearings below roadway joints are difficult to protect.
Source: Courtesy KTA-Tator, Inc.

Figure 6.15.  Shoring to support bridge during bearing replacement operation. Note heavy corrosion on
­bearings in right photo.
Sources: Courtesy (left) KTA-Tator, Inc., and (right) District 11-0, Pennsylvania DOT.
Source: Courtesy (left) KTA-Tator, Inc., and (right) District 11-0, Pennsylvania Department of
Transportation.
6.3.2.7
Figure 6.15. Shoring to support bridge during bearing replacement operation. Note heavy corrosion on
Wind-Blown
bearings in right-hand photo. Rain or Salt Spray
During rain, water can be blown onto steel surfaces even when there is a bridge side-
walk above. The ability of a coating to withstand this occasional exposure to fresh,
nonbrackish water should not present a problem for zinc-rich-based coating systems.
In fact, rain water can provide a benefit because it can cleanse exposed surfaces.
288

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


During storms, salt water from nearby brackish water or from ice- or snow-melt
can be blown onto steel surfaces. Repeated exposure to salt water is very corrosive,
and the duration of wetness and the number and length of wet–dry cycles affect the
degree and extent of corrosion. In areas with long periods of wetness, such as those
caused by frequent wet–dry cycles, hot-dip galvanizing (HDG) or zinc spray metalizing
should be considered for primary corrosion protection.
6.3.2.8
Chlorides
In a landmark literature survey compiled by Alblas and von Londen published in the
Journal of Protective Coatings and Linings in February 1997, the researchers stated
that “It has been clearly established that soluble salts on the surface of steel will
­increase the rate of corrosion and paint breakdown for many [coating] systems now in
use.” This conclusion is as true in 2012 as it was in 1997. The authors offered several
conclusions:

• From available data, it is not possible to establish a definitive allowable level of


chloride contaminants.
• In relation to the durability of the paint system, a maximum chloride level of 10 to
50 µg/cm2 is thought to be permissible, depending on the use and exposure condi-
tions. This is only a rough guideline.
• Under specific conditions, higher maximum levels of chloride (up to hundreds of
micrograms per square centimeter) are allowed for special, durable paint systems
(e.g., zinc silicate).
• Exposure to marine conditions and/or industrial environments considerably in-
creases the chloride contamination on steel.
• Abrasive blast cleaning does not remove all the chloride.
• Results of detection methods for soluble chlorides are affected by temperature,
mechanical forces, and the chemicals and type of analytical method used.
• The effect on steel of the hydrochloric acid generated as a consequence of the cor-
rosion reaction is notable. Therefore, the removal of as much chloride as possible
during blast cleaning and other surface preparation efforts is crucial. Although
there is still not complete agreement as to the precise level of chloride residue
that is acceptable, the Surface Preparation Commentary for Steel and Concrete
Substrates (Subsection 4.3.6, Soluble Salts) (SSPC 2004a) identifies three levels of
chloride removal:
–– 0 µg/cm2;
–– Less than 7 µg/cm2; and
–– Less than 50 µg/cm2.
• A level of chloride removal commonly specified for conventional mild steel is
7 µg/cm2.

289

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


An FHWA-sponsored study (Appleman et al. 1995) concluded that the maximum
safe level of chloride to remain when cleaning weathering steel is 50 µg/cm2.
In accordance with SSPC-SP 10, abrasive blast cleaning on noncorroded areas will
reduce chloride levels to an acceptable <7 µg/cm². It will not always do so on heavily
rusted, rust scale–covered, or pitted or pack rust–affected areas.
In some areas, it is believed that complete removal of all chlorides via only dry
abrasive blast cleaning is at best unlikely, and at worst, provides a false sense of secu-
rity, even if white metal blast cleaning (SSPC-SP 5, White Metal Blast Cleaning) is
specified. Cleaning efforts beyond abrasive blast cleaning are usually needed. The use
of high-pressure water cleaning (5,000 to 10,000 psi) has been found to significantly
reduce residual chlorides to a very low level. High-pressure waterjetting (10,000 to
30,000 psi) has also been used to reduce residual chloride levels. As noted, the effect
of chlorides on the corrosion rate of steel has been studied and is well documented.
6.3.3
Factors Affecting Service Life of Galvanized or Painted Steel
Bridge Elements Specific to Galvanizing Coating
The corrosion protection of unpainted galvanizing comes from the formation of a
thin, invisible layer of insoluble corrosion products. Zinc, an active metal, reacts with
oxygen in the air; zinc oxide starts forming within 24 to 48 hours after galvanizing
and takes about a year to cover the entire galvanized surface. The zinc oxide converts
to zinc hydroxide on exposure to moisture in the form of rain, dew, or high humidity.
The final step is the reaction of zinc oxide and zinc hydroxide with carbon dioxide in
the air to form zinc carbonate. This reaction requires free-flowing air. Zinc carbonate
is the dense insoluble material that forms the protective layer, sometimes called the
­patina. Zinc oxide and zinc hydroxide are water soluble and not very dense. They
adhere loosely to the surface, so painting over zinc oxide or zinc hydroxide does not
provide good adhesion of the coating to the surface. The practical problem is that zinc
oxide, zinc hydroxide, and zinc carbonate are all white and cause the galvanized sur-
face to appear a dull, matte gray, which does not allow a visual determination of what
form of zinc compound is present. Knowing the compound is important, because some
forms are not corrosion resistant and are unsuitable for painting over.
6.3.3.1
Reactivity of Zinc
The reactivity of zinc is well known to galvanizers. For instance, they know that if
pieces are closely stacked together for shipment, there will be no access to carbon
dioxide in free-flowing air to form the zinc carbonate. In such a case, only loose zinc
oxide and zinc hydroxide will form, causing rapid consumption of the zinc. For this
reason, closely spaced galvanized pieces should be unpacked, after which the loose
white deposit on the surface should be noted. If this reaction process is allowed to
continue, it can consume all of the zinc by reaction with the moisture caught between
the pieces. Although rare, rusting of the unprotected steel may then occur, resulting in
the presence of rust beneath the deposit.

290

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


6.3.3.2
Prevention or Passivation of White Storage Stain
The white deposit described in Section 6.3.3.1 is called wet storage stain. Galvanizers
apply a light coating of oil to prevent the stain. The oil forms a barrier to keep mois-
ture from reaching the zinc, thus preventing the zinc from being converted to oxide
and hydroxide forms. However, as paints do not stick to oil, painting the surface with-
out first removing the oil is unacceptable. This is true no matter what type of coating is
applied. Another process used to prevent wet storage stain is quenching or passivating
with chromates or phosphates. Quenching (i.e., cooling and water bath) is not harmful
in itself. However, the quenching bath may contain small amounts of oil and grease
on the surface of the water that are picked up when pieces are removed. Coatings also
do not stick to chromate-quenched galvanizing, but the phosphate improves adhesion.
Although wet storage stain can damage galvanizing, the methods used to prevent it can
affect painting results. It is always recommended to consult the galvanizers concerning
the process employed, especially if the galvanized items are to be painted.
6.3.3.3
Repair of Defects in the Galvanized Surface
The next step in surface preparation is to repair any defects or handling damage. Gal-
vanizing can leave high spots and zinc droplets, which occur when a galvanized piece
is withdrawn from the bath and excess zinc runs down the edges onto a protrusion
or irregular edge. Droplets form at edges where zinc drains from the piece and can be
removed with hand tools. High spots are usually ground down with power tools. Care
is required to avoid removing so much zinc that the remaining thickness is below the
specified minimum. SSPC-Guide 14, Guide for Repair of Imperfections in Galvanized
or Inorganic Zinc-Rich Coated Steel Using Organic Zinc-Rich Coating (SSPC 2004b)
should be consulted.
Unstable zinc oxide or zinc hydroxide may not have been entirely removed during
the initial cleaning process. There is no simple method for identifying the presence of
either, so the surface must be further treated.
Galvanizing can be eroded if exposed to very strong acids or alkali, which may
cause the zinc to dissolve as metallic zinc is soluble in very strong acid or alkali envi-
ronments. In these unusual circumstances, if regalvanizing is not possible, repairs can
be made with coatings in accordance with SSPC-Guide 14 described above. After
restoring the zinc protection, a decision can be made as to whether painting is desired.
6.3.3.4
Preparing Weathered Galvanizing for Painting
Fully weathered galvanizing (i.e., galvanizing that has been outdoors for at least 1 year
and preferably about 2) should have a fully formed layer of protective zinc carbonate.
Nothing is required to prepare the surface for normal atmospheric exposure, and its
service life will not be limited in the normal course of exposure events. Bare galvanized
surfaces will be subjected to the vicissitudes of the local weather environment, and
this unknown, complex exposure may mean that it makes sense to combine galvaniz-
ing and painting. A so-called duplex system (i.e., galvanizing and paint) should be

291

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


considered. Such systems are said to provide 1.5 to 2.5 times the service life of the
sum of both galvanizing and paint if each is considered separately, and an aesthetically
pleasing palette of colors is available.
If a surface of weathered galvanizing is to be painted, the surface must normally
be power washed with clean water at about 1,450 psi. Spot repairs of any damage in
accordance with SSPC-Guide 14 are all that is necessary.
6.3.3.5
Surface Preparation of New Galvanizing for Painting
The often-made claim that galvanized surfaces cannot be painted is incorrect. In fact,
galvanized surfaces are routinely painted successfully. Several steps are described in
this section in which errors of omission or commission are routinely made, with rem-
edies for each. There is no reason why so-called duplex systems cannot perform for
decades.
New galvanizing means galvanized steel that is between 1 or 2 days new and up
to about 2 years old. Wet storage stain, if present, must be removed before surface
preparation. Removal can be done by brushing the stain with a 1% to 2% ammonia
solution such as diluted household ammonia. After treatment, ammonia should be
removed by rinsing with warm water.
The first step in the surface preparation is to wash off oil, grease, and dirt. This
cleaning is performed in accordance with SSPC-SP 1, Solvent Cleaning. Water-based
emulsifiers or alkaline cleaners work best. A mildly alkaline cleaner should be used.
The cleaning solution should be applied by dipping, spraying, or brushing with soft-
bristle brushes. A temperature range of 140°F to 185°F works well. Afterwards, the
surface should be thoroughly rinsed with hot water and allowed to dry. One helpful
tip for determining if oil was applied to the galvanized surface to prevent wet storage
stain is to contact the galvanizer; another way is to perform a water bead test in which
a drop of water is placed on the surface. If it beads, oil will probably be present. The
best advice is that when in doubt, the entire surface should be washed as described.
After the surface is washed, it should be examined for zinc ash, a residue that consists
of particles of oxidized zinc that float on the surface of the galvanizing bath. The ash
can be removed by washing the surface with a 1% to 2% ammonia solution.
Common methods for treating the surface in the field before painting are phos-
phating by the use of wash primers or sweep-blast cleaning.
6.3.3.5.1
Phosphating Preparatory to Painting
Phosphating is often accomplished by using a wash primer, a coating that neutralizes
the surface oxides or hydroxides and etches the galvanized surface. The most common
wash primer is polyvinyl butyral (e.g., SSPC-Paint 27). These materials are very thinly
applied (0.3 to 0.5 mil) by brush or spray. The galvanized surface should shadow
through the coating at this thickness. If the galvanized surface is completely hidden,
the wash primer is too thick. Wash primers have poor cohesive strength and will split
apart if they are too thick, resulting in paint disbondment.

292

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Phosphating is not recommended if a zinc-rich primer is going to be applied. Zinc-
rich primers require intimate contact between the zinc particles in the paint and the
zinc metal on the galvanized surface. The zinc phosphate acts as an insulator in the
same way that iron oxide (i.e., rust) acts as an insulator on steel surfaces.
6.3.3.5.2
Sweep-Blast Cleaning Preparatory to Painting
Sweep blasting is a method of light blast cleaning that can remove zinc oxides on
the surface and roughen the surface without significantly removing the galvanizing.
Sweep-blast cleaning should be performed with abrasives that are softer than the gal-
vanized surface. The use of materials with a Mohs scale hardness of five or less is
suitable. Sweep-blast cleaning should be performed in accordance with SSPC-SP 7,
Brush-Off Blast Cleaning.
6.3.4
Factors Affecting Service Life of Steel Bridge Elements
Specific to Metalizing Coating
Metalized coatings provide corrosion protection to steel by both sacrificial and barrier
protection. The coating itself provides a barrier between the environment and the steel
surface, especially when applied in combination with conventional sealer coatings (e.g.,
epoxies, polyurethanes, and acrylics) as topcoats. Due to the electro­chemical reaction
between steel and zinc or aluminum in an aqueous or salt-contaminated envi­ronment,
these coatings sacrifice themselves to protect the steel at the site of any damage, or
holes, in the coating. This sacrificial protection is akin to the protection p
­ rovided by
zinc-rich primers or galvanizing.
6.3.4.1
Metalized Coatings
Metalized coatings can be applied in the shop or in the field by using a variety of
techniques and equipment. The metal or metal alloy is applied in wire form and is fed
through a source and liquefied. The source may be either flame (i.e., oxygen–acetylene)
or electric arc. The liquefied metal is immediately propelled onto the prepared steel
surface by using air spray in a manner similar to that used in painting. Once on the
surface, the liquid metal cools and dries very quickly to form a continuous protective
coating over the steel surface.
6.3.4.2
Cost of Metalizing
Cost estimates made in 2012 place metalizing as two to three times per square foot
the cost of conventional painting. A recent project at a large fabricator queried by the
authors indicates that a significant differential currently exists. On a best-case basis, it
is estimated that metalizing costs at least 40% to 50% more than painting.
6.3.4.3
Salt-Contaminated Areas
Metalized coatings consist of spray droplets that have solidified and overlapped, pro-
viding a somewhat coarse matrix. This matrix is a barrier coating, as well as a chemi-
cally active one, as a result of the anode–cathode relationship between zinc and iron.
293

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


The primary benefit of metalizing over other coating technologies is its durability and
corrosion resistance, especially in salt-rich environments. For this reason, metalizing
should be considered as an option for bridge structures in salt-rich environments or
for areas or components of bridge structures that receive considerable exposure to salt
and moisture from drainage and runoff. Although there are cost differences between
metalizing and painting, in many cases metalizing should be specified. Based on the
performance of metalizing over a long period of time, repairs and renovations on steel
bridges would benefit by its use.
Metalized coatings have been shown to perform very well in studies when applied
over steel that has been blast-cleaned in accordance with SSPC-SP 5, White Metal Blast
Cleaning, or SSPC-SP 10, Near-White Blast Cleaning. These coatings have a dull gray
appearance with a rough texture as applied, but may be sealed and topcoated with
most conventional paints. Sealing is recommended by many existing guidelines as it
tends to increase coating life, reduce the deleterious effects of metalized coating poros-
ity, and improve aesthetics.
Metalized coatings provide the benefit of defect tolerance. The sacrificial nature
of these coatings provides corrosion protection to the underlying steel at the site of
breaches in the coating film. Metalized coatings, particularly aluminum and aluminum
alloys, also tend to be quite abrasion resistant.
6.3.4.4
Bond Strength
The bond between the metalized coating and the steel surface is mechanical in nature.
As such, the bond is sensitive to surface contaminants and to the shape of the surface
profile. Surface preparation should be specified as above (SSPC-SP 10 or SSPC-SP 5)
with an angular 2- to 4-mil anchor-tooth profile.
Because blast cleaning with rounded steel shot has produced deficient adhesion
results, steel shot abrasives should not be used on surfaces that will or may be metalized.
As a solventless coating application method, metalizing is less forgiving than con-
ventional paint application. Applicators must be properly trained and experienced
with the specific equipment and metals or alloys to be used.
Because metalized coatings are inherently porous, achieving an adequate coat-
ing thickness (6- to 8-mil minimum) in an overlapping spray pattern is critical to
­coating life.
6.3.4.5
Field Versus Shop Application
Metalizing technology may also be applicable to field maintenance coating operations
when a long-term, durable corrosion-protection coating system is required. Applica-
tions of metalized coatings in the shop, and particularly in the field, require technically
sound specifications and practices.

294

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


6.3.5
Factors Affecting Service Life of Steel Bridge Elements
Specific to Weathering and Noncorrosive Steels
6.3.5.1
Corrosion-Resistant Weathering Steel
With the introduction of steel as a material of construction for bridges in the late
1800s, the industry has sought to find a form of steel that can overcome its most basic
limitation: corrosion. It was believed that an answer had been found in the 1970s with
the introduction of weathering steel. Since the 1970s, the search for the ideal material
has evolved through improved higher-strength weathering steel, or high-performance
steel. Each step in this evolution has produced incremental improvements in the per-
formance of weathering steel in normal weathering environments. Unfortunately, there
has been even wider use of weathering steel in bridges at locations that are not recom-
mended for the best use of weathering steel. These areas are often in heavily salted
areas or areas where the steel is sheltered or exposed to other conditions so that the
corrosion-resistant patina simply does not form. These areas are discussed elsewhere
in this chapter.
6.3.5.2
A1010 Structural Stainless Steel
Although initial corrosion studies performed on A1010 steel have been favorable, the
use of A1010 steel is currently inhibited somewhat by its premium cost. It is believed
that with sufficient production, volume costs will come down. Although testing has
produced promising results, its performance in aggressive, salt-laden areas is not com-
pletely known. Even with these unknowns, however, it is hopeful that a solution to
the 125+ year-old problem of dealing with the corrosion characteristics of steel will
be determined. As additional bridges are constructed using A1010 steel, time will tell.
As of 2012, one bridge has been completed in California and two others are under
construction in Oregon.
6.4
Options for Enhancing Service Life:
Corrosion Performance of Steel
6.4.1
General Categories of Solutions for Preventing Corrosion of Steel
Bridge Elements
In general, there are three options for developing corrosion-prevention systems for
steel bridges:

1. The use of a coating system, which can consist of paint, galvanizing, or metalizing
systems;
2. The use of corrosion-resistant steel (weathering steel) or noncorrosive steel; and
3. Avoidance of corrosive environments or corrosion-prone details.

295

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


6.4.2
General Strategies for Producing an Effective Corrosion-Protection
System
Regardless of the option selected from the list provided in Section 6.4.1, there are
five major strategies that can result in an effective corrosion-protection system. These
strategies are described in the following sections and listed here:

1. Review design to assure that the best protection is designed into the structure,
2. Use composite protection,
3. Use corrosion-proof materials,
4. Employ superdurable coatings, and
5. Use ongoing engineered maintenance painting.

6.4.2.1
Designing Corrosion Protection into the Structure
Through the first 125 years of the steel bridge era, steel bridges have benefitted from
the corrosion control foresight of their designers. The elimination or minimization of
corrosion on such structures has resulted in a knowledge base that, if systematically
applied to every structure, can benefit each one. As these lessons learned are applied,
the corrosion-resistance features, principles, experiences, and insights should be de-
signed into every new and rehabilitated steel bridge. Actions taken during the earliest
project design stages can cause a dramatic lengthening of the coatings part of the main-
tenance–repair–replace cycle by eliminating areas likely to corrode early in the ser-
vice life of a structure. If corrosion resistance is designed into a structure by carefully
managing the configuration and details of bridge design and detailing while using the
current coatings systems, bridge corrosion resistance should improve dramatically. As
seen in Figure 6.16, lack of attention to the relationship between design and potential
corrosion can lead to unwanted exposure of the metal to corrosion. The design review
should be considered a design hold point. In this instance, hold point means that fur-
ther progress on the design would depend on having a corrosion review performed and
a corrosion-resistance control plan initiated.
This design phase review is a major means for creating a 100-year life for a new
steel bridge. In order to attain 100 years of service life, it will be necessary to develop
and use preferred details, which will serve as a way to lengthen the time before any
maintenance painting is needed during the structure’s expected 100-year service life.
6.4.2.2
Composite Protection
Currently the use of zinc to protect steel from corrosion is the gold standard of care.
The use of a composite protection strategy is based on the premise that there is an
order of efficacy in terms of corrosion protection provided by zinc as delivered in its
various forms.
HDG is considered the most efficacious protection because of the iron–zinc alloy
that is formed on the steel surface closest to the outside of the HDG part. Even if the
HDG surface is later nicked, the alloy layer will afford substantial protection from
296

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


(a) (b)

Figure 6.16.  Details difficult to paint. (a) Lack of stiffener clearance to back wall results in an almost unpaint-
able detail. (b) Coating failure at splice, especially around fasteners.
Source: Courtesy KTA-Tator, Inc.

corrosion. Many smaller bridge elements, steel bearings, cross frames, bolts, and
expansion devices can be protected with HDG.
Metalizing, as noted previously, has been tested for decades and also found to be
an excellent means of protecting steel from corrosion. The spray-applied zinc does not
form an alloy layer like HDG, but it does provide zinc in intimate contact with steel
(iron) in order to provide effective galvanic protection. Metalizing has been tested
repeatedly in both the laboratory and field and found to provide a very high level of
corrosive protection.
Zinc-rich, primer-based coatings systems have been the workhorse of the steel
bridge industry for over 40 years, and coating systems based on zinc-rich coatings have
a successful track record on countless bridges.
Uncoated weathering steel also has a 40+ year history of providing successful cor-
rosion protection in certain exposure areas.
Structures and parts of structures can be protected using combinations of protec-
tive steps. For example, steel bearings or cross frames can be hot-dip galvanized or
metalized and then painted. Some fasteners (ASTM A-325 bolts) are available with
either an HDG or mechanically galvanized coating and either can be coated or not as
required.
297

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


A composite approach is often adopted on weathering steel bridges when girder
ends are cleaned and coated. Figure 6.17 shows several structural details that are suit-
able for composite protection.
In the future, ASTM A1010 or a similar material may conceivably be routinely
employed in coastal areas or where salt usage is a certainty. It may prove feasible to use
A1010 in combination with weathering steel. In such a case, perhaps girder ends could
be made of A1010 steel, while the remainder of the girder is composed of weathering
steel or painted regular mild steel.
Table 6.2 summarizes the comparative functionality of galvanizing, metalizing,
and zinc-rich paint in terms of cost, protection, and durability. “Duplexable” in
Table 6.2 refers to the particular coating’s ability to be combined with other types to
provide a composite coating system. In the Duplexable column, G represents galvanize,

(a) (b)

(c)

Figure 6.17.  Details suitable for composite protection. (a) Welded cross frame could not be galvanized.
(b) Mill to bear stiffener leaves crack in the coating design. (c) Unique opportunity for composite protection
presented by the configuration of a large trunnion girder for a lift bridge.
Source: All photos courtesy KTA-Tator, Inc.
298

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 6.2. Three Ways to Apply Zinc to Steel
Method Efficacy Relative Cost Durability Duplexable?
Galvanize Best Best ($1.76/ft2) Best Yes, G/P
Metalize Better Better ($4.10/ft ) 2
Better Yes, M/P
Zinc-rich paint system Good Good ($2.27/ft )2
Good Yes, P/P/P

Source: KTA-Tator, Inc.


Note: The best (strongest), second-best, and third-best performers are identified as Best,
Better, and Good, respectively. Galvanizing costs are from the American Galvanizers
Association. Other costs are from experience and research among fabricators.

M represents metalize, and P represents a paint layer. For example, galvanizing can be
combined with a top paint coat, and zinc-rich paint as a primer can be combined with
multiple additional paint layers.
6.4.2.3
Use of Corrosion-Resistant or Corrosion-Proof Materials
6.4.2.3.1
Corrosion-Resistant Steel
Weathering steel (coated or uncoated) has been the subject of much research and dis-
cussion since its initial use on bridges in about 1970. Weathering steel’s roots lie in the
improvement in the corrosion resistance of steel when small amounts of copper, chro-
mium, nickel, phosphorous, silicon, manganese, or combinations of these elements are
added to carbon steel. When weathering steel is properly exposed, a rusty red-orange
to brown or purple-tinted patina forms. When the patina is formed, the corrosion rate
of the steel stabilizes within about 3 to 5 years. The formation of the protective patina
requires a series of wet and dry periods. In certain situations, the protective patina does
not form completely or not at all. For example, if the steel is sheltered from the rain,
the dark patina cannot form. In areas with high concentrations of corrosive industrial
or chemical fumes, weathering steel may exhibit a much higher corrosion rate. In
a saltwater marine environment or in areas heavily exposed to chloride-containing
­deicing materials, the protective patina does not form. The use of uncoated weathering
steel in such locations is not recommended.
When weathering steel is used in locations where regular wet–dry cycles occur,
the steel is corrosion resistant to the point that no coating is necessary. In some loca-
tions, weathering steel enjoys a vastly enhanced corrosion resistance that can render it
relatively impervious to corrosion. The exact degree of corrosion resistance afforded
is dependent on a number of variables, including climatic conditions, pollution levels,
and the degree of sheltering from the atmosphere, as well as the composition of the
steel itself. These variables influence the areas in which the use of weathering steel is
appropriate.
In a survey conducted as a part of the SHRP 2 R19A Project (final report available
at http://www.trb.org/Design/Blurbs/168760.aspx), about one-third of the 16 DOTs
responding reported the use of weathering steel on over 50% of their steel bridges.
299

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


Many states use Technical Advisory FHWA T5140.22, Uncoated Weathering Steel
in Structures (FHWA 1989), as a guidance document. This technical advisory for weath-
ering steel use is in the process (as of 2012) of being revised and updated by FHWA.
In order to shield the weathering steel in areas likely to experience chloride-laden
water exposure, many states paint the end of the weathering steel members for a dis-
tance of 1.5 to 2 times the depth of the web. The same zinc-rich, primer-based coating
systems used by the various DOTs for nonweathering steel are employed. In other loca-
tions, weathering steel is used for its other characteristics and is coated in its entirety
for protective and/or aesthetic reasons.
6.4.2.3.2
Noncorrosive Steel
The perfect weathering steel would be a material that is corrosion proof in all environ-
mental conditions. Such a material would present a desirable solution to addressing
the main limitation of steel as a construction material, that is to say, steel rusts. And
steel rusts even more in the presence of chlorides. Therefore, the “perfect” material
would be a true unpainted, corrosion-resistant solution for any environment.
A steel product believed by some to meet these stringent requirements has now
been developed. The corrosion resistance has been accomplished by chemically aug-
menting the steel’s metallurgical composition. This new grade of weathering steel,
ASTM A1010, contains 10.5% to 12.5% chromium and is said by its vendors to be
immune to corrosion based on testing performed in Kure Beach, North Carolina, in a
25-m test site (see Figure 6.18).

Kure Beach 25m


120
A36
Corrosion Loss, microns

100 50W

HPS-70W

80 100W

A1010
60

40

A1010
20

0
0123456
Years
Figure 6.18.  Corrosion resistance of A36, 50W, HPS-70W, 100W, and A1010 grades
of steel exposed at Kure Beach, North Carolina [25.4 μm = 1 mil (0.001 in.)].
Source: Fletcher et al. 2003.
300

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


This “corrosion-proof” steel is just beginning to be used on bridges. One small
structure was constructed in 2004 in Colusa County, California. This project was the
subject of a report presented at the 2004 Prefabricated Bridge Elements and Systems
Conference. The bridge design was a prefabricated lightweight section referred to as a
multicell box girder. Less than 23 tons of A1010 were needed to form the structure of
this short-span bridge with overall dimensions of 72 ft long by 32 ft wide.
Construction of two bridges using A1010 steel is under way in Oregon, and the
possibility of constructing other bridges has been mentioned in other states.
Although A1010 steel remains a somewhat costly alternate material choice, it does
demonstrate that it may be possible to anticipate the use of such a material in the future.
6.4.2.4
Use of Superdurable Coatings
The search for the Superman of coatings continues:

• FHWA’s Research, Development, and Technology Program on coatings has re-


cently focused research on two-coat systems and their ability to perform as well as
the traditional three-coat system that has been in use since around 1965. Testing
of one-coat system candidate materials for steel bridges has also been under way.
FHWA released its final report on that testing, Performance Evaluation of One-
Coat Systems on New Steel Bridges, as Report No. HRT-11-046 (FHWA 2011c).
Although there were some promising prospects among the materials tested, none of
them approached the performance of the three-coat control system in the testing.
• Ongoing efforts to identify new resins and pigments for improved coatings are
under way in the private sector.
• Developing new superdurable coating materials via both basic and applied re-
search efforts, including industry-to-industry technology transfer, is under way.
The use of nano particles in coatings has begun but has not yet spread to the bridge
industry. The potential use of nano-sized (a billionth of an inch) pigment particles
that can dramatically alter the performance of a coating is much anticipated.
• In new construction, the coating systems and procedures are basically unchanged
since the late 1960s. For new bridges the steel is coated using a system consisting
of a zinc-rich primer, usually an epoxy midcoat and usually a urethane topcoat,
applied over steel cleaned in accordance with SSPC-SP, 10 Near-White Metal Blast
Cleaning. Initial cleaning and priming is normally done in the fabrication shop.

6.4.2.5
Maintenance Painting
As in many other areas of the construction industry, quality in bridge painting must
be built in and cannot simply be added after the fact. Properly selected and applied
coatings can often last for many decades with periodic planned maintenance painting.
A comprehensive approach to maintenance painting requires considerations of surface
preparations, inspection, and proper planning.

301

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


6.4.2.5.1
Surface Preparation
The initial condition of the surface to be cleaned will determine the amount of work,
time, and money required to achieve any particular degree of surface cleanliness. It is
more difficult to remove contaminants from rusty steel and to remove mill scale from
new steel than it is to wash surface film off steel in good condition. Therefore, it is
necessary to consider the surface condition before selecting the method of cleaning.
The method of cleaning is an integral part of how the coating system may be expected
to perform in any given environment. The initial condition of the steel may determine
the choice of abrasives. Steel shot is an economical and effective choice for removing
intact mill scale. Although their use in the field is not unknown, steel abrasives are
usually recycled and therefore find their most common use in the shop. However, if the
steel is rusted or pitted, an angular abrasive, such a steel grit or a nonmetallic mineral
abrasive, will more effectively scour out the rust.
The initial conditions encountered can be broadly divided into three categories as
follows:

1. New construction—steel not previously coated;


2. Maintenance—repainting of previously coated, painted, metalized, or galvanized
steel; and
3. Contaminated surfaces—common to both new construction and maintenance.

Typical contaminants that should be removed during surface preparation are rust,
corrosion products, mill scale, grease, oil, dirt, dust, moisture, soluble salts (e.g., chlo-
rides and sulfates), paint chalk, and loose, cracked, or peeling paint. Each of these
contaminants is discussed in the following paragraphs.
Rust contaminants include rust, rust scale, and pack rust. Rust consists primar-
ily of iron oxides, the corrosion products of steel. Whether loose or relatively tightly
adherent, rust must be removed for satisfactory coating performance. Rust resulting
from the corrosion of steel is not a good base for applying coatings because it expands
and becomes porous.
Ideally, rust and rust scale should be removed, even when using the lowest degrees
of hand and power tool cleaning (SSPC-SP 2, Hand Tool Cleaning, and SSPC-SP 3,
Power Tool Cleaning). Judgment should be used on an individual project basis whether
the cost and effort required to remove the stratified rust, rust scale, and to a greater
or lesser extent, pack rust, can be justified by the expected increase in the life of the
coating system. To effectively repair pack-rusted joints, it may be necessary to remove
rivets, separate the plies of steel, clean, paint, and refasten with bolts.
On riveted and bolted connections, bridge management practices are required that
cause surfaces to be repaired long before such inefficient, costly repairs are necessary.
It is obvious that many square feet of steel can be cleaned and recoated before the cost
of disassembly and reassembly of bridge connections is equaled.

302

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


The existence and treatment of rust and particularly pack rust can make bridge
repair so expensive that bridge demolition may appear to be a feasible option. When
such matters are expected to be at issue, agreement about the extent of removal of
these materials should be reached before work begins.
There is a trade-off between repair cost and extended service life. For maintenance
repainting, the degree of surface preparation required depends on the new coating
system and on the extent of degradation of the surface to be painted. The amount of
rusting on the surface is based on the numerical scale of zero to 10 given in SSPC-VIS
2, Standard Method of Evaluating Degree of Rusting on Painted Steel Surfaces (SSPC
2000), in which a reading of 10 indicates no rust and a rating of zero indicates more
than 50% rusting. SSPC-PA Guide 4, Guide to Maintenance Repainting with Oil Base
or Alkyd Painting Systems (SSPC 2004c), suggests the minimum surface preparation
needed for each degree of rusting. This guide includes a description of accepted prac-
tices for recleaning old, sound paint, removing rust, and feathering the edges of sound
coating around the area and recoating. Additional information on the subject may be
found in SSPC-SP COM, Surface Preparation Commentary for Steel and Concrete
Substrates (SSPC 2004a).
Mill scale is a bluish, normally slightly shiny outside residue that forms on steel
surfaces during hot rolling at the steel mill. Although initially tightly adherent, it even-
tually cracks, pops, and disbonds. As a general rule, unless it is completely removed
before painting, at some point it will most likely crack and cause the coatings to crack,
exposing the underlying steel surface. In addition, steel is anodic to mill scale, and as
the anode in the resultant dissimilar metals corrosion cell (with oxygen from the air
and moisture) will corrode more rapidly.
At least in the short term, mill scale is somewhat unpredictable in its effect on
the performance of coatings, although if it remains tightly adhered, intact mill scale
may not have to be removed at all for steel exposed to a mild atmospheric exposure.
However, if the steel surface is to be coated with primers with low wetting properties
or exposed to severe environments such as wetness or immersion in fresh or salt water,
then removal of mill scale by blast cleaning or power tool cleaning is necessary.
Soluble salts are deposited from the atmosphere onto surfaces. If they are permitted
to remain on the surface after cleaning and are coated over, they can attract moisture
that can delaminate the coating and cause blisters. Salts, particularly chlorides, may
also accelerate the corrosion reaction and underfilm corrosion. Methods for measuring
the amount of salt on the surface are described in SSPC-Guide 15, Field Methods for
Retrieval and Analysis of Soluble Salts on Steel or Other Nonporous Substrates (SSPC
2005). In some circumstances, it is desirable to remove soluble salts by power washing
or other methods employing wet methods before power tool or abrasive blast clean-
ing. In other circumstances, salt removal is more efficient after the member has initially
been subjected to abrasive blast cleaning. The extra effort required to remove this non-
visible surface contaminant will help immeasurably in improving coating durability.

303

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


The sun’s ultraviolet light causes exposed organic coating to chalk to some extent.
Chalk is the residue left after deterioration of the coating’s organic binder on exposed
surfaces. All loose chalk must be removed before coating in order to avoid intercoat
adhesion problems.
Sharp edges, such as those which at times may occur on some rolled structural
members or plates, as well as those resulting from flame cutting, welding, grinding,
and shearing, could influence coating performance and may need to be addressed.
Additional guidance on the subject of material anomalies and sharp edges can be found
in Steel Bridge Collaboration Specification S-8.1, Guide Specification for Application
of Coating Systems with Zinc-Rich Primers to Steel Bridges (AASHTO/NSBA 2006).
6.4.2.5.2
Coatings (Paint) Inspection
The importance of coating inspection during surface preparation and coating applica-
tion cannot be ignored or underemphasized.
As a general rule, it is not possible to visually examine a coated surface and know
whether the surface preparation and coating application were done in accordance with
the applicable specification and good painting practice. Determining whether the coat-
ing material was properly mixed; whether a component was substituted, adulterated,
or left out completely; or whether an entire coat of paint was simply skipped, can be
a tedious, costly process. Once a surface has been painted, it is usually not possible to
determine whether each painter in a crew has complied consistently with the specifica-
tion. It is important to recognize the value of adequate inspection.
Unless trained inspectors monitor the entire operation from start to finish, there is
no way to know for sure about the level of specification compliance actually achieved;
coating systems can only perform if they are properly installed.
Certification programs for bridge paint inspectors are offered by the Society for
Protective Coatings. The SSPC Bridge Coatings Inspector (BCI) training course was
developed by a committee of more than a dozen DOT representatives. The course is
appropriate for any level of worker in the coatings industry, including apprentices,
blasters, painters, foremen, superintendents, engineers, and inspectors. Details about
the BCI course can be found at www.sspc.org.
Attendance at the BCI class is designed to instruct the attendees to be able to do
the following:

1. Define the varying professional roles of the inspector, bridge owner, and painting
contractor and their relationship to each other at the project site.
2. Identify what preparation the inspector must make before the start of work in
order to conduct effective inspections.
3. Recognize common coating inspection and related terms.
4. Identify and properly adjust and operate commonly used coating inspection in-
struments and test equipment.
5. Identify fundamental surface preparation and coating application processes.

304

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


6. Identify key documents (e.g., specification, product data sheets, and technical bul-
letins; industry technical standards; and references) required to perform competent
inspections.
7. Identify inspection check or hold points.
8. Create inspection documentation, including a basic inspection plan.
9. Identify processes normally inspected and documented.
10. Identify common coating application defects.

A variety of other inspection training classes are offered by private organizations


and nonprofit societies. Although all these classes have strengths, SSPC-trained and
-certified BCIs have received training prepared specifically for the coating of steel
bridges. The SSPC-trained BCI is considered by many to be the most experienced
and best-trained BCI in the bridge coatings inspection field, and many of SSPC’s coat-
ings inspection training courses include the preparation of inspection plans in the
curriculum.
To provide guidance to those responsible for creating quality control plans and
a training document for course participants, SSPC developed a Guide for Planning
Coatings Inspection in 2008. The planning guide describes the importance of qual-
ity monitoring on a project to reduce the risk of coating failure and the challenges
associated with trying to assess quality after the project is complete. It also stresses the
importance of planning the inspection to increase the likelihood that inspections are
performed and the results properly documented. The intended purpose of the planning
guide is to assist coating inspectors, quality control personnel, and owners with a tool
to help ensure the coating or lining installation is the best it can be.
6.4.2.5.3
Worst First Versus Engineered Maintenance Painting
Many agencies are perennially short of maintenance painting funds. As a result, the
bridges that receive coating attention are those that appear to be in the most distress
(worst first). By the time a structure appears to need the most attention, it is probably
well past the point at which the spot-on zone cleaning can be effectively employed.
SSPC Technology Update TU-3, Overcoating (2004), offers guidance when consider-
ing overcoating. Subsection A.2.1 (Bridge Painting Using Risk Tables) of Appendix A
deals with the percentage of the surface at which the cost of the repairs approaches
that of full removal. The percentage identified as the critical percentage beyond which
the surfaces are rusted or distressed such that surface preparation is necessary is 16%
per ASTM D610, Rust Grade 2 (Rust Grade 10 = essentially no rust, and Rust Grade
0 ≥50% rust).
SSPC reports in this document that according to the Guide for Painting Steel
Structures (AASHTO 1994), “whenever the surface preparation area exceeds 15% to
20% of the surface area, the economics are such that a total removal of lead paint is
the most viable option.”

305

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


From a cost perspective, the difference in cost between spot or zone cleaning and
painting versus full removal is dramatic. According to industry sources, a typical lead-
paint removal project in the northeastern part of the United States (in 2011) averages
~ $13/ft2. Of that amount, about half ($6.50/ ft2) is attributed to surface preparation
(access cost, containment, equipment, abrasive, and labor).
If spot or zone cleaning were possible, costs on a comparative basis would be
about $3.90/ft2 (30%). If cleaning were performed when the surface affected was >3%
(Rust Grade 4) but <10%, costs would be lower still.
It is apparent that a timely touch-up would extend the service life of the paint
project and lower costs. The reasons for being forced into a worst-first mode vary,
but the practice is common. If an engineered approach to maintenance recoating were
employed, overall costs could be reduced by a large percentage, and the condition of
the coatings on the bridge inventory in a given city, district, or state would, in time,
improve.
It is recognized that bridge maintenance activities are driven by many factors, not
all of which are corrosion related. When the matter of maintenance painting is consid-
ered, an engineering approach will help to counter the worst-first approach. If the de
facto use of the worst-first practice is unchanged, coatings costs will be at their highest.
If every bridge that was repainted, even if completely redone, were placed in a mas-
ter schedule of planned paint touch-up in, say 20 years, eventually the worst bridges
would be repainted and those structures that were able to be recoated to extend their
service lives would emerge as the norm. There are further enhancements to the plan-
ning that can be employed. For example, the deterioration of one structure may be
faster than another. In those cases, the examination and evaluation of a structure can
be scheduled in a different cycle.
Early intervention saves money, can stretch the budget to cover additional proj-
ects, and can eventually improve the condition of the steel structures across an area,
district, or even an entire state.
6.4.2.6
Bridge Maintenance Owner’s Manual
Every new bridge and every rehabilitated structure should be delivered with an owner’s
manual containing a lifetime maintenance plan that outlines, much like an automobile
owner’s manual, when designated corrosion-mitigation activities are to be undertaken.
For example, it might be recommended that coating touch-up of minor nicks and
scratches be undertaken every 5 years and that every 25 or 30 years, whenever certain
conditions exist, the structure be touched up and overcoated. Following this mainte-
nance plan based on known, needed activities would allow the owner to achieve the
service life inherent in the structure.
6.5
Strategy Selection Process
Specifiers can choose among a wide variety of coating systems. Several of these systems
are recognized by the coatings industry as having a track record of successful perfor-
mance in a given service environment. These systems are assembled by the coating

306

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


manufacturer according to product. In many cases, a given system can be used in a
multitude of locations and service environments. For example, a system comprising
a zinc-rich primer, an epoxy intermediate coat, and an acrylic polyurethane topcoat
can be used to protect bridge steel in most areas and has a +40-year service history.
A coating system is selected on the basis of the prevailing service environment, the
intended life of the structure, the level or degree of surface preparation possible, the
intended service life of the coating, access, and any other constraints.
Table 6.3 lists common generic coating systems, common service environments
within a given structure, and coating systems candidates for each. This chart repre-
sents a cross section of coating systems.

Table 6.3. Coating Systems for Highway Bridges (New Construction and Maintenance)
Highway Bridges Highway Bridges Highway Bridges
Coating System (New) (Maintenance-1) (Maintenance-2)
Inorganic zinc-rich primer–polyamide epoxy– X
acrylic polyurethane
Polysiloxane X
Organic zinc-rich primer–polyamide epoxy–acrylic X X
polyurethane
Organic zinc-rich primer–polyamide X
epoxy–polysiloxane
Organic zinc-rich primer–polyamide X X
epoxy–fluoropolymer
Organic zinc-rich primer–polyurea X X
Moisture-cure urethane zinc-rich primer–moisture- X X
cure urethane–moisture-cure urethane
Moisture-cure urethane zinc-rich primer–moisture- X X
cure urethane–acrylic polyurethane
Inorganic zinc-rich primer–waterborne acrylic X
Organic zinc-rich primer–waterborne acrylic X
Thermal spray coating–sealer X X
Epoxy sealer–epoxy mastic–acrylic polyurethane X
Epoxy mastic–acrylic polyurethane X
Epoxy mastic–waterborne acrylic X
Moisture-cure urethane sealer–moisture-cure X
urethane–moisture cure-urethane
Moisture-cure urethane–moisture-cure urethane– X
acrylic polyurethane
Alkyd–silicone alkyd X
Calcium sulfonate alkyd (two coats) X

Source: KTA-Tator, Inc.


Note: Maintenance-1 = recoating (total removal and replacement of existing system); Maintenance-2 = overcoating
(spot or zone repair, spot or zone repriming and overcoat); X = coating system is applicable to the types of bridges
indicated.
307

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


Maintenance overcoating is a process in which new coating is applied over exist-
ing coating. Based on industry knowledge and DOT survey information obtained for
this project, the coating systems currently in use for this purpose include acrylic, cal-
cium sulfonate, epoxy sealer–epoxy–urethane, epoxy sealer–urethane, polyester, and
polyaspartic.
Maintenance recoating is a process in which a new coating system is applied over
a surface from which all old coating has been removed. According to the DOT survey
information described in this section, the most commonly used systems consist either
of an organic or inorganic zinc-rich primer with an epoxy midcoat and a urethane
topcoat.
These zinc-rich primer–based systems have proven themselves in the field on thou-
sands of structures for over 40 years. In the bridge industry, few materials have this
proven track record. No doubt the demonstrated longevity of the systems has contrib-
uted to their continued use.
6.5.1
Characteristics by Coating
Table 6.4 lists common coating types and their inherent properties and characteristics.
Although zinc–epoxy–urethane systems are widely used on bridges, special circum-
stances may dictate the use of systems tailored for a specific application. A description
of the more commonly used coatings is included in this section.
An explanation of Table 6.4 and an example of how it can be used to select a coat-
ing material on the basis of the desired performance characteristics follows.
The left column of the chart contains industrial coating types. Note that within a
coating type category, there can be subcategories that are not shown. For example, the
category of organic zinc-rich primer includes an epoxy zinc, urethane zinc, vinyl zinc,
and so forth. This list is not exhaustive, but rather contains some of the more common
coating types. The top row on the chart contains 17 common characteristics.
Once the service environment is identified and the intended use of the coating is
determined, the specifier can review which generic categories of coatings are available.
For example, if the specifier is considering overcoating, five coatings can be consid-
ered for this application (alkyd, calcium sulfonate alkyd, epoxy mastic, moisture-
cure urethane, and waterborne acrylic). However, if the overcoat material must also
demonstrate abrasion resistance, then only two candidates remain, epoxy mastic and
moisture-cure urethane, as the other three do not possess abrasion-resistant proper-
ties. If single-pack paint is desirable (i.e., a product that has all ingredients in a single
container), then the specifier can select a moisture-cure urethane from these two, as the
epoxy mastic is a two-pack product that requires mixing prior to application.
6.5.1.1
Acrylic
Acrylic coatings can be formulated as thermoplastic, solvent-deposited coatings, cross-
linked thermoset coatings, and water-based emulsion coatings.
The acrylic resins, with suitable pigmentation, provide excellent film-forming coat-
ings characterized by excellent light fastness, gloss, and ultraviolet stability. Chemi-
cal resistance to weathering environments is generally excellent, as is resistance to
308

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 6.4. Coating Characteristics Chart

Coating Type

Color and Gloss Retention


Surface Tolerant
Flexible
Easy to Apply
Low Cost
Can Be Modified
Acid Resistant
Caustic Resistant
Abrasion Resistant
Solvent Resistant
Fast Dry
Single Pack
Low VOC Available
Overcoat Material
Chemical Resistance
Immersion
Typical Maximum Service
Temperaturea

Alkydb X X X X X X X X X 250oF
Silicone alkyd X X X X X X X 250oF
Calcium sulfonate alkyd X X X X X X X 250oF
Epoxyb X X X X X X X X 250oF
Epoxy mastic X X X X X X X X X X 250oF
Urethaneb X X X X X X X X X 250oF
Moisture-cure urethane X X X X X X X X X X X X X 250oF
b
Inorganic zinc rich X X X X 750oF
Binder
Organic zinc richb X X Mcuz X X
dependent
Waterborne acrylicb X X X X X X X X 250oF
Polyurea X X X X X X X X X X 350oF
200oF–
Polysiloxane X X X X X X X X X
1,400oF
Wire
Thermal spray coating X WD WD X X X X X WD X
dependent

Source: Reprinted with permission of SSPC: The Society for Protective Coatings.
Note: x = characteristic is applicable to coating type. Mcuz = moisture-cured urethane zinc; WD = wire dependent.
a
Consult coating material manufacturer’s product data sheet.
b
Most common types of coatings used on steel bridges.

309

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


moisture. Most acrylic coatings are not suitable for immersion service or strong chemi-
cal environments. In some state DOTs water-based coatings are required as topcoats
because of local environmental rules banning VOCs or local preferences.
6.5.1.2
Cross-Linked Thermoset Coatings
Chemically cured coatings refer to coatings that harden or cure and attain their final
resistance properties by virtue of a chemical reaction either with a copolymer or by
­reaction with moisture. Coatings that chemically cross link by copolymerization include
the epoxy family of coatings, including urethanes. Chemically cured coatings that react
with water are moisture-cured polyurethanes and all of the inorganic zinc-rich coatings.
Coatings based on chemically cured binders can be formulated to have excellent
resistance to acid, alkalis, and moisture and to resist abrasion, ultraviolet degradation,
and thermal degradation. Chemical and moisture resistance increases as the cross-link-
ing density increases within the larger macromolecule. The rate at which the molecule
cross links is dependent not only on the reactants, but also on the cross-linking mech-
anism. Most importantly, external factors such as temperature and, with moisture
reactions, atmospheric humidity, affect the rate and extent of cross linking. Thus, for
chemically altered converted coatings, after application the coating must set through
solvent or water volatilization and then harden and attain its final cured properties via
the cross-linking reaction, which is temperature and/or moisture dependent. A reac-
tion that is too fast may lead to an overcured, hard, impervious coating that cannot
be recoated or topcoated with a properly adherent subsequent coat. This is always
a problem in maintenance repainting when a renewal coat is applied to the original
cross-linked coating system after an extended period of time. As a general rule, curing
of most chemically cross-linked systems should proceed for approximately 7 days at
75ºF before the coating system is exposed to severely corrosive conditions. In corrosive
environments, the structure may have to be enclosed in a containment device, and the
coating manufacturer’s guidance on these issues should be sought.
Following is a discussion of the more common chemically cross-linked binders or
resins used for coatings.
6.5.1.2.1
Epoxies
Chemically curing epoxies usually come in two packages: one consists of the e­ poxy
resin, pigments, and some solvent, and the other is the curing agent. For bridge coat-
ings, the two packages are mixed immediately before application and, on curing,
­develop the large macromolecule structure that provides a tough, water-resistant,
­durable film. However, the film is subject to chalking when exposed to sunlight and is
normally topcoated.
6.5.1.2.2
Urethane Coatings
Urethane coatings have chemical- and moisture-resistance properties similar to the
­epoxies, but they can also be formulated in a variety of light-stable colors and hues
that maintain their gloss and wet look after prolonged outdoor exposure.

310

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Acrylic urethanes are perhaps the most widely used corrosion-protection urethanes
for atmospheric service on bridges. These coatings, when properly formulated, have
excellent weatherability, gloss, and color retention and good chemical and moisture
resistance. They can be readily tinted and pigmented to provide a variety of deep and
pastel colors at a lower cost per gallon than the next most popular class, the polyester
urethane. They have excellent weathering properties.
6.5.1.3
Zinc-Rich Coatings
Zinc-rich coatings are a unique class of coating materials that provide galvanic pro-
tection to a ferrous substrate. As the name implies, the binder is highly loaded with a
metallic zinc dust pigment. After the coating is applied to a thoroughly cleaned sub-
strate, the binder holds the metallic zinc particles in contact with the steel and with
each other. Thus, metal-to-metal contact of two dissimilar metals is made, resulting in
a galvanic cell. In this metallic couple, zinc becomes the anode and sacrifices itself to
protect the underlying (cathodic) steel.
The major advantage of corrosion protection using zinc-rich coatings is that pit-
ting corrosion is eliminated, even at voids, pinholes, scratches, and abrasions in the
coating system. This cannot be said of any nonzinc type of protective coating, and it
is this protective capability that makes zinc-rich coating so unique and invaluable on
bridges.
This advantage, however, comes with certain disadvantages. The underlying steel
substrate must be cleaned of all mill scale, rust, old paint, and other contaminants
that may interfere with metal-to-metal contact. Thus, the degree of surface prepara-
tion must be quite thorough: blast cleaning should, at a minimum, be an SSPC-SP
6 (Commercial) blast. For more aggressive, immersion-like exposures, SSPC-SP 10,
Near-White Blast Cleaning, or SSPC- SP 5, White Metal Blast Cleaning, is necessary.
No material is perfect. Because of the high reactivity of the zinc dust pigment,
zinc-rich coatings are not suitable outside a pH range of approximately 5.5 to 11, and
most bridges are located in environments that are within this range. (pH ranges from
1 to 14, with pH 7 being neutral.) Strong acids and strong alkalis will attack the zinc
dust pigment, and even if topcoated, penetration of the chemicals may occur through
pinholes, scratches, voids, or discontinuities within the topcoat, leading to aggressive
attack of the zinc-rich primer.
However, despite these disadvantages, the advantages that accrue by eliminating
pitting corrosion make zinc-rich coatings, either topcoated or untopcoated, one of the
most widely used corrosion-prevention coatings for painting steel bridges.
Zinc-rich coatings can be used either as primers with topcoats or as complete one-
coat systems. Both organic and inorganic zinc-rich coatings are used extensively for
steel protection on bridges and highway structures, and any area where fresh or salt-
water corrosion, mild fume, and high humidity and resultant corrosion are a problem.

311

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


6.5.1.3.1
Types of Zinc-Rich Coatings
Zinc-rich coatings can be subcategorized into two types: those with organic or with
­inorganic binders. The organic types are similar in many ways to the epoxy and ure-
thane coating systems previously discussed, except that sufficient zinc dust pigment is
added to provide galvanic protection. The inorganic zinc-rich coatings use a different
binder chemistry and are quite unlike the organic zinc-rich coatings.
6.5.1.3.1a
Organic Zinc-Rich Coatings
Organic zinc-rich coatings are most commonly formulated from epoxy polyamide,
urethane, vinyl, and chlorinated rubber binders.
The drying, hardening, and ultimate curing of the organic zinc-rich coating is
predicated on the type of binder used. The organic nature of organic zinc-rich primers
makes them more tolerant of deficient surface preparation, as they more readily wet
and seal poorly prepared surfaces where residues of rust or old paint may remain. Sim-
ilarly, topcoating with the same generic type of topcoat is more readily accomplished
because organic zinc-rich coatings of all types generally have a less porous surface and
are more akin to conventional non-zinc-rich coatings than are the inorganic zinc-rich
coatings.
6.5.1.3.1b
Inorganic Zinc-Rich Coatings
SSPC has categorized inorganic zinc-rich coatings for use in the bridge industry in two
major groups: self-cured water-based alkali metal silicates and self-cured solvent-based
alkali silicates.
Although the binder in both cases is an inorganic silicate, essentially the same
material as glass or sand, the curing of the binder is different.
6.5.1.3.1c
Self-Curing Water-Based Alkali Silicates
The most common of these silicate binders is based on potassium and lithium sili-
cates or combinations of the two. Lithium hydroxide–colloidal silica and quaternary
­ammonium silicate binders are also included in this category. Self-curing alkali silicate
zinc-rich coatings become hard within minutes and are considered generally resistant
to precipitation within half an hour after application.
When final curing is ultimately attained, most water-based zinc-rich coatings expe-
rience a color change, often from a reddish-gray or light gray color to a darker bluish-
gray color.
6.5.1.3.1d
Solvent-Reducible, Self-Cured Inorganic Zinc-Rich Coatings
The binders for this class of coatings are essentially modifications of partially hydro-
lyzed alkyl silicates. Of these, the ethyl silicate type is most commonly used.

312

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


During the condensation phase of the reaction, the partially polymerized silicate
combines with atmospheric moisture to eliminate alcohol, which vaporizes. After
complete hydrolysis, the cross-linked network forms a matrix to hold the pigment
particles.
6.5.1.3.2
Durability of Zinc-Rich Primer–Coated Structures
Thousands of zinc-rich paint–coated structures have been built since the late 1960s
that are in good condition. In these cases, a third of the desired service life of 100 years
has already been met.
The ability of the coating to last 100 years or more appears to be achievable.
Improved coating systems that have been extensively tested by the National Trans-
portation Product Evaluation Program (see Section 6.5.2) can be expected to perform
for 100 years. Naturally, periodic systematic planned maintenance painting must be
performed. Simply put, “painting it now and coming back in 100 years” expecting to
see a coating system in good condition is not believed to be achievable at this time.
6.5.1.4
Hot-Dip Galvanizing
Galvanizing is the process in which steel pieces or parts are immersed in a kettle or vat
filled with molten zinc, resulting in a metallurgically bonded alloy coating that protects
the steel from corrosion.
When galvanizing is exposed to the natural wet and dry cycles of the atmosphere,
it develops a zinc by-product layer on the surface. This layer is stable and nonreactive
unless exposed to aggressive chlorides or sulfides. The protective layer is a key compo-
nent in the longevity of the HDG coating in the atmosphere.
The American Galvanizers Association projects that HDG items will last 75 to
100 years. Figure 6.19 shows the relationship between time to first maintenance and
zinc thickness for various types of environments. The various environments shown in
the key are listed in the order that the lines are shown in the graph, from top to bottom.
Although there are some important differences between zinc-rich coatings and gal-
vanizing, the extensive field performance history of zinc-rich coatings in combination
with the American Galvanizers Association data strongly suggests that steel that has
been properly coated with a zinc coating and has additional coating layers to extend
the service life of the zinc coating beneath can last 100 years or more.
6.5.1.5
Thermal Spray Metalizing
Zinc in wire form may be applied to clean steel by feeding it into a heated gun where
it is heated, melted, and spray applied using combustion gases or auxiliary compressed
air to provide ample velocity. Metalizing may be used on any size steel object, so
limitations due to vat size and awkward shapes are eliminated. Applying a consistent
coating in recesses, hollows, and cavities adds a measure of complexity. Pure zinc can
be used, but often zinc is alloyed with 15% aluminum to provide a smoother film.

313

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


Figure 6.19.  Service life chart for HDG coatings in various environments.
Source: Photo courtesy of the American ­Galvanizers Association.
Figure 6.19. Service life chart for HDG coatings in various environments.

Metalizing spray application generates a smaller spray pattern, and application is


normally slower than spray painting; accordingly, zinc thermal spray is generally more
costly.
FHWA funded two studies relative to zinc thermal spray that are reported in Publi-
cations FHWA-RD-91-060 (Kogler and Mott 1992) and FHWA-RD-96-058 (Kogler et
al. 1997). The 1992 report stated that “metalized test systems . . . performed extremely
well in the battery of accelerated tests.” The later study reported on the testing of 13
coating systems for 5- to 6.5-year periods at three seaside sites. Three of the systems
were 100% zinc, 100% aluminum, and 85% zinc–15% aluminum metalizing. In most
cases, there was virtually no rusting or undercut in the metalized coating, and they
were reported to have “near perfect corrosion performance.” In this study, Kogler et
al. (1997) noted that the metalizing did not perform any better with or without the
epoxy topcoat.
In summary, metalized coatings outperform conventional liquid-applied zinc-rich
coating systems, but they are less easily applied and they are less aesthetically pleasing
without having high gloss topcoats.
6.5.1.6
Composite Protection
It is believed that in a given exposure, galvanizing will outperform metalizing because
of the alloying effect of molten zinc and steel (iron). Metalizing has been shown to
produce impressive corrosion protection, and generally, zinc-metalized surfaces will
outperform zinc-rich paint–coated surfaces. Zinc-rich paint–coated members have a
314

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


proven field history. When difficult conditions are foreseen, all three methods can be
combined to provide protection from corrosion. For example, steel bearings can per-
haps be hot-dip galvanized, and girder ends or cross frames could be metalized. In
addi­tion, it may be possible and desirable to coat a longer section on the girder end
than the traditional 1.5 times the girder depth. The best set of practices should be set
forth for the myriad of service conditions encountered, depending on the exposure
conditions encountered on both a macro- and microenvironment basis.
6.5.2
Performance Evaluation of Individual
Protective Coating Systems Products
Independent verification of coating system performance based on laboratory testing
and/or field exposure is a critical component to selecting a coating system. A given
coating system may have multiple manufacturers. It is not safe to assume that all
coating systems within a given generic category are created equal. Therefore, careful
evaluation of coating system performance is needed before full-scale field application
to determine which of the candidate systems will perform the best.
AASHTO oversees a materials testing branch—the National Transportation Prod-
uct Evaluation Program (NTPEP)—comprising highway safety and construction mate-
rials project panels. These panels are made up of state highway agency personnel whose
objective is providing quality and responsive engineering for the testing and evaluation
of products, materials, and devices that are commonly used by the ­AASHTO member
DOTs. In 1997, the Structural Steel Coatings (SSC) Panel was created to develop a
standard specification, a corresponding project work plan, and a reporting system for
testing industrial coating systems for use on bridge and highway structures. Data are
generated by prequalified independent testing laboratories and uploaded to a central
database known as Datamine for DOT access. All testing is paid for by the coating
manufacturer.
The advantage of this type of performance evaluation is that many agencies within
a given industry can access performance data with little or no associated costs. Limi-
tations include the ability to keep the database current as new coating systems come
to market, the time associated with generating the performance data (SSC requires
approximately 10 months), and the application of the same performance criteria to the
data for a coating that may be used on a bridge structure in northern Minnesota and
on a bridge in Phoenix, Arizona, two very different service environments.
Some facility owners and agencies may choose to employ a combination of perfor-
mance evaluation methods. For example, a bridge owner may subscribe to Datamine
(industry-specific performance evaluation) and may also suspend or mount racks of
test panels containing candidate coating systems for a bridge structure.

315

Chapter 6. CORROSION PREVENTION OF STEEL BRIDGES


7
Fatigue and Fracture
of Steel Structures

7.1
Introduction
This chapter introduces basic principles related to fatigue and fracture in steel bridges
and discusses factors that cause fatigue and fracture. Various available options for re-
pairing observed cracking in steel bridges are also presented. These options are adapted
from the Manual for Repair and Retrofit of Fatigue Cracks in Steel Bridges (Dexter
and Ocel 2013) and are proposed as a guide for the detailing of repairs and retrofits for
fatigue cracks. This chapter only contains summarized information from this manual
and thus should not be the only means used to develop specifications needed for the
repair and retrofit of fatigue-damaged details. Refer to the referenced manual (Dexter
and Ocel 2013) for additional detailed descriptions of the topics, procedures, and ex-
amples presented in this chapter. Further, this chapter should be used in combination
with other existing codes, specifications, and engineering judgment.
7.2
Background
Cracks found in steel elements of bridges can usually be attributed to fatigue. Fatigue
in metals is described as the process by which cracks initiate and grow under repeated
loads. These fatigue cracks can lead to failure if the remaining uncracked section can
no longer carry the loads experienced by the structure. In the case of bridge struc-
tures, fatigue failure usually occurs as a result of the crack growth that initiates from
existing discontinuities. In fatigue, these existing discontinuities are treated as existing
cracks. All fabricated steel elements contain discontinuities, and most contain high
stress concentrations at weld toes. The stress levels causing the failure due to fatigue
are usually considerably lower than those that can cause failure under static loading

316
conditions. Fatigue cracks usually form under large numbers of load cycles and worsen
with higher stress ranges (Fisher et al. 1998).
Cracks and discontinuities are expected in steel structures and do not n ­ ecessarily
mean that the member will fail as long as the proper precautions are taken. Most
modern structures are redundant and allow for the excess stresses in the cracked mem-
bers to be redistributed, thus keeping the fatigue crack from propagating any further
without intervention. However, it is important to assess tension elements that contain
cracks to determine the potential for fracture. Bridges that do not possess redundancy
for the redistribution of stresses face failure of the entire structure if one of the mem-
bers were to fail; these members are known as fracture critical members. These struc-
tures call for more careful attention, as a fatigue crack can be detrimental to the life of
the bridge (Fisher et al. 1998).
Fatigue failure often occurs very suddenly, with little warning; however, the pro-
cess begins at the onset of the structure’s usage, implying that fatigue is progressive.
Another important aspect of fatigue is that it is a local phenomenon, occurring in areas
of high stresses and strains due to load transfer, abrupt changes in geometry, residual
stresses, and material imperfections. The damage caused by fatigue is permanent and
is not reversible. Fatigue cracks exist in many structures, but not all of them are criti-
cal; certain criteria must be met before the cracks are detrimental to the structural
element. Fracture (i.e., separation of a component into two or more parts) occurs once
the remaining uncracked portion of the member can no longer handle the stresses and
strains (Stephens et al. 2001).
The entire fatigue process includes the nucleation (formation) of a fatigue crack,
crack propagation (growth), and final fracture (failure). The nucleation of a fatigue
crack takes place at the microscopic level, dealing primarily with the microstructure
of the material. Discontinuities are common sites of crack nucleation and include per-
sistent slip bands, inclusions, pores, second-phase particles, corrosion pits, voids, and
twin and grain boundaries. However, cracks primarily tend to nucleate along slip lines
in the direction of planes of maximum shear (Stephens et al. 2001).
Once a fatigue crack forms and continues to undergo repeated loading, it tends
to coalesce and grow along the plane of maximum tensile stress range. The crack will
grow with each load cycle, even if only by a small amount. As the cracked member is
loaded, the crack will open, causing an increase in stress at the crack tip; this increased
stress consequently drives the crack to grow even larger. Fatigue crack growth is bro-
ken up into two stages, Stage 1 and Stage 2, as seen in Figure 7.1. Stage 1 refers to the
growth in the direction of the principal shear plane, and Stage 2 refers to the growth
along the plane of maximum principal tensile stress. Fatigue cracks tend to grow trans­
crystalline (through grains), but some fatigue cracks can grow intercrystalline (along
grain boundaries). Crack growth mechanisms include striation formation, microvoid
coalescence, and microcleavage. Striations are microscopic “ripples” that are repre-
sentative of the fatigue cycles experienced by the element. Striations can be used to
investigate the rate of crack growth and are very useful in forensic studies. Microvoid
coalescence involves the formation, growth, and joining of microvoids during plastic

317

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


Figure 7.1.  Schematic of Stages 1 and 2 of
fatigue crack growth.
Source: Stephens et al. 2001.

deformation. Microcleavage is a fracture along specific crystallographic planes and


tends to be a brittle fatigue mode (Stephens et al. 2001).
This section presents only a brief and very general summary of the fatigue pro-
cess, but it is important for the practicing engineer to understand the principles of the
fatigue damage process in order to be proficient in fatigue design.
Minimizing fatigue cracks can be achieved by first avoiding the use of known
details that have proven to have low resistance to fatigue. Fatigue cracks can also be
reduced by the use of better fabrication and welding processes that decrease the num-
ber of inherent defects and also allocate for the detection and repair of such cracks
before the bridge is opened to the public. In-service inspection is necessary to discover
new fatigue cracks and monitor existing ones. Once cracks are located through the
inspection process, it is necessary to perform an assessment to determine the risk of
fracture. Furthermore, for the proper repair methods to be implemented, the cause and
type of crack need to be determined (Fisher et al. 1998).
Methods for determining the fatigue resistance of a detail include nominal stress
approach, hot-spot stress approach, and fracture mechanics approach.
7.2.1
Nominal Stress Approach
The nominal stress design approach is a simple way to determine the fatigue resistance
of a detail by using equations for bending and axial loads to compute the nominal
stress near a weld toe. Test data from full-scale fatigue tests are needed in order to use
this design method. Stress ranges (S) versus the number of cycles to failure (N) curves
are developed from the test data. The curves are grouped into categories to aid in
organizing the details according to their fatigue resistance. Figure 7.2 shows the S-N
curves as used in AASHTO specifications. The detail categories reflect and account for
318

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure
Source: 7.2.  S-N
AASHTO curves used in AASHTO, American Institute of Steel Construction (AISC), American Welding
(2012).
Society (AWS), and American Railway Engineering and Maintenance-of-Way Association (AREMA) specifications.
Source: AASHTO 2012.
Figure 7.2. S-N curves used in AASHTO, American Institute of Steel Construction (AISC), American

Welding Society (AWS), and American Railway Engineering and Maintenance-of-Way Association
the variations in the combined geometric and local notch stress concentrations. Each
(AREMA) specifications.
category has a constant-amplitude fatigue limit (CAFL), also referred to as constant-
amplitude fatigue threshold (CAFT). Stress ranges that fall below the CAFL are not
expected to exhibit any fatigue failures during constant-amplitude testing.
Most bridges with a service life of 75 years are designed as having an infinite life,
with no occurrence of fatigue cracking. In the LRFD Bridge Design Specifications
(LRFD specifications) (AASHTO 2012), the fatigue design live load is taken as 0.75
times the HS20 for finite load-induced fatigue life and 1.5 times the HS20 for infi-
nite load-induced fatigue life. This fatigue load is used to calculate the nominal stress
ranges to be used with the S-N curves. If the resulting nominal stress range is less than
half of the CAFL, it is assumed the bridge is designed for infinite life. This ensures
that the fatigue limit–state stress range is below the CAFL. The fatigue limit state is
the stress range in which 0.01% of the test data exceeds the CAFL (AASHTO 2012).

319

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


7.2.2
Hot-Spot Stress Approach
The hot-spot stress approach is similar to the nominal stress approach, except that the
S-N curves are based on the geometric stress ranges, also known as hot-spot stresses.
This process is beneficial in instances when the nominal stress approach breaks down,
such as offshore tubular structures in which the fatigue resistance is heavily dependent
on the geometry of the tubes. Using this design process involves determining the stress
concentration factor by using parametric equations or finite element analysis. Dis­
advantages arise with the variability of different hot-spot definitions, varying baseline
S-N curves, and difficulties with the CAFL.
7.2.3
Fracture Mechanics Approach
In the case of bridge structures, the fracture mechanics approach is the most complex
approach and tends to be difficult to implement during the bridge design process. The
fracture mechanics approach is divided into two main categories: linear elastic fracture
mechanics (LEFM) and plastic fracture mechanics. LEFM is used when remotely ap-
plied stresses are in elastic ranges. It should be noted that at the crack tip, a stress sin-
gularity is always present and stresses tend to approach infinity. In the case of LEFM,
the rate of crack growth is related to stress ranges, but in the case of plastic fracture
mechanics, rate of crack growth needs to be related to a parameter related to energy
dissipation or plastic strain (Azizinamini and Radziminski 1989).
For the case of LEFM, the Paris law, shown in Equation 7.1, can be useful in deter-
mining the crack growth rate for many engineering applications:

da
= C ∗ ∆K m (7.1)
dN
where
a = crack size (mm or in.),
N = number of cycles,
C = material constant,
ΔK = stress intensity factor range  MPa ⋅ m 2 or ksi ⋅ in. 2  , and
1 1

 
m = material constant.

Although fracture mechanics is rarely used in design, it can serve as a qualitative


tool to give the designer a better understanding of structures containing cracks and
discontinuities. Fracture mechanics is best suited when the local behavior of structure
in the vicinity of the crack is of interest, such as how fast the crack will grow. However,
addressing such problems requires a detailed understanding of the parameters that
affect crack performance.

320

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


7.3
Crack Detection Techniques
Cracks are not always obvious to the human eye and can be difficult to locate at times.
Several methods in practice aid in the detection of cracks, two of which are dye pen-
etrant and magnetic particle inspection.
Dye penetrant consists of three parts: cleaning the area of a suspected crack, appli-
cation of a liquid dye, and finally the application of a white developer. Figure 7.3
shows a crack exposed using a red dye penetrant.
Magnetic particle inspection works by inducing a magnetic field by using a hand-
held device around a crack. The magnetic field is disrupted at the crack, and a con-
centration of a magnetic field results. A fine iron powder is sprinkled over the area of
interest and is attracted to the magnetic field, exposing the crack.
7.4
Repair and Retrofit Methods
This section presents several methods for the repair and retrofit of fatigue critical
details. Such techniques can be categorized as surface treatments, repair of through-
thickness cracks, and connection or global structure modification to reduce the causes
of cracking.
7.4.1
Surface Treatments
Surface treatments are usually performed on weld toes to increase the fatigue strength
of uncracked welds. Such treatments, including grinding, gas tungsten arc, and impact
treatments, aim to improve the weld geometry in order to reduce stress concentrations,
remove discontinuities, or reduce residual tensile stresses.

Figure 7.3.  Crack exposed using a red dye


penetrant.

Crack

321

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


7.4.2
Reshaping by Grinding
Grinding can be used as an effective measure for increasing the fatigue life of the
weld toe by removing portions of the weld that contain small cracks. Grinding has
proven more effective on larger welds in structures such as offshore structures with
large ­tubular joints. It should be noted that grinding is ineffective against microcracks
because the process tends to create new microcracks as it removes the existing ones.
Two types of grinding methods are commonly used: disc grinding (Figure 7.4) and burr
grinding (Figure 7.5). Both methods have advantages and disadvantages. Disc grinding
can be more effective at removing the weld material with faster speeds; however, the
operator needs to use caution to avoid gouging the metal or removing too much of the
weld material. Burr grinding is typically easier to operate and works in more confined
spaces than disc grinding (Gregory et al. 1989).
7.4.3
Gas Tungsten Arc or Plasma Remelting
Gas tungsten arc aims to reduce the stress concentrations at the weld toe and also
remove slag intrusions. This process involves melting a small volume of the weld toe
and base material by using tungsten electrodes. For this process to effectively increase
the weld’s fatigue life, the operator needs great skill, which consequently increases the
cost of this process.

Figure 7.4.  Disc grinding.

322

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 7.5.  Burr grinding.

7.4.4
Impact Treatments
Compressive residual stresses can be induced around the weld toe by using impact
treatments. These compressive stresses reduce the effective tensile stress range, extend-
ing the fatigue life of welds. Since impact treatments enhance the weld profile and
residual stresses, the process can only affect stresses transverse to the impacted weld.
Thus, impact treatments are most effective on transversely loaded welds and have no
effect on longitudinally loaded welds. The most common types of impact treatments
include air hammer peening and ultrasonic impact treatment.
Air hammer peening uses an air-powered hammer with a blunt tip that p ­ lastically
deforms the weld toe. This simple method can increase the fatigue resistance by at
least one detail category. For instance, a Category C detail could be improved to a
Category B detail. Air hammer peening reduces the number of slag intrusions, but at
the same time it creates lap-type defects. These lap-type defects can be reduced by light
grinding following the peening (Hausammann et al. 1983).
Ultrasonic impact treatment (UIT), which uses low-amplitude and high-frequency
displacements, has proven to be more effective than hammer peening. However, this
treatment can be more costly as it is still a proprietary method (Tryfyakov et al. 1993;
Roy et al. 2003).

323

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


7.4.5
Hole Drilling
Hole drilling is the most widely used method for the repair of fatigue cracks. The pro-
cess involves drilling a hole at the tip of the crack (propagating end). The larger the
hole, the more effective it is at arresting the fatigue crack from propagating, as long as
the hole is not detrimental to the stiffness of the member. Holes should not be made
smaller than 1 in. in diameter as holes smaller than this tend to be ineffective. Holes
should be sealed against corrosion and plugged to hide the hole from the public. Equa-
tions have been developed to simplify the process of hole size selection for in-plane
fatigue, as shown by Equation 7.2 (Fisher et al. 1980):

∆K
ρ
(
≤ 10.5 σ y for σ y in MPa )
∆K
ρ
(
≤ 4 σ y for σ y in ksi ) (7.2)

∆K = Sr π a

where
ΔK = stress intensity factor,
r = radius of the hole,
sy = material constant, and
Sr = nominal stress range at crack tip.

At the tip of cracks, singularity exists and stresses approach infinity. Drilling elimi-
nates the high stress concentration and prevents further crack growth.
7.4.6
Vee-and-Weld
The vee-and-weld method is best for long, through-thickness cracks. The process in-
cludes removing the material along the crack in the shape of a V and then filling the
groove with weld material. The groove can be made using several methods, the pre-
ferred being air arc gouging. Grinding can also be done, but it tends to smear the crack
path, making it harder to detect the crack and follow its path. Other m ­ ethods need to
be used in addition to vee-and-weld repairs in order to reduce the stress ranges at the
location of the repair. Additional methods are necessary because the vee-and-weld re-
pairs have a fatigue life that is only equal to that of the original uncracked weld (Dexter
et al. 2003).
7.4.7
Adding Doubler Splice Plates
Doubler plates can be added at crack locations to increase the cross-sectional area and
therefore reduce the stress ranges experienced by that section (see Figure 7.6). Dou-
bler plates are designed to restore the section properties of the cracked section to the
uncracked state by using design processes identical to those of field splice connections.

324

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 7.6.  Bolted doubler plate repair.

7.4.8
Posttensioning
Posttensioning methods that are applied to cracked sections can prolong the fatigue
life of the structure. Posttensioning induces forces on the cracked section that put the
effective stress ranges into compression, keeping the crack closed and unable to propa-
gate. Drilling a hole at the crack tip is recommended in addition to using one of the
various types of posttensioning methods.
7.4.9
Detail Modification
Detail modification is used when it is necessary to lower the effective stress range in
order to repair the cracked section. This modification can be achieved in a number of
ways, among them increasing the cross-sectional area, changing connection geometry,
or eliminating sharp corners from details.
7.5
Fatigue Caused by Secondary Stresses
Secondary stresses can arise when a structure is designed as a series of individual
components and the designer does not account for the global system behavior. These
stresses can cause unexpected fatigue cracking. This section discusses these stresses
and the methods used to repair the fatigue cracks that form under secondary stresses.
7.5.1
Out-of-Plane Distortion
Differential displacements between girders and lateral bracing elements introduce
­fatigue to the web-gap regions of the girders. This phenomenon causes highly local-
ized bending of the web gap, as shown in Figure 7.7, causing fatigue cracks. In order

325

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


Figure 7.7.  Web-gap fatigue mechanism
from displacement continuity: (top)
differential girder displacements cause
a force couple to develop within the
diaphragm and (bottom) zoomed view of
web-gap deformation girder displacement.

to properly repair the fatigue cracks the out-of-plane bending needs to be reduced
or eliminated. It is important to note that web-gap fatigue retrofits need to maintain
symmetry.
7.5.1.1
Repair Methods Specific to Out-of-Plane Distortion
7.5.1.1.1
Hole Drilling
Hole drilling can be effective at reducing the crack growth but not eliminating the
cause of the fatigue. See Section 7.4.5 on hole drilling for additional details; however,
note that the hole-sizing equations were developed for in-plane fatigue and may not
have the same effect on out-of-plane distortion–induced fatigue.
7.5.1.1.2
Diaphragm or Crossframe Removal
Diaphragms and crossframes transfer the secondary forces between girders when dif-
ferential displacement of the girders occurs. Removing these members eliminates the
causes of the fatigue-induced cracks in the web gaps. However, several issues of con-
cern have arisen from the removal of such bridge elements. It has been shown that
the removal of the lateral bracing elements can be detrimental to the structure when
they are not properly removed. In negative moment regions, the lateral bracing keeps
the compression flange from buckling. Also, if the diaphragms and crossframes are
removed because the bridge deck needs to be replaced, no lateral bracing will exist to
keep the girders stable when the deck is removed. Some studies show that crossframes

326

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


are effective to some extent in distributing the applied traffic loads (Brakke 2002;
Flemming 2002). However, extensive investigation of crossframes indicated that it is
the stiffness of the deck that is mainly responsible for distribution of truck loads be-
tween girders and that crossframes do not contribute to load distribution (Azizinamini
et al. 1995a, 1995b).
7.5.1.1.3
Diaphragm Repositioning
It has been shown that lowering diaphragms closer to the bottom flange of the girders
(in negative moment regions) and reducing the number of connecting bolts can reduce
the effective stress range. This was seen in a case study of the I-35W Bridge over the
Mississippi River in Minneapolis, Minnesota (Bergson 1998). See Figure 7.8.
7.5.1.1.4
Bolt Loosening
Loosening the connection bolts can reduce the effect of the out-of-plane displace-
ments. The holes are specified to be larger than the size of the bolts, and this extra
space negates the effects of small differential displacements. However, the effectiveness
of loosening the bolts is limited to the extra space provided by the oversized holes and
whether the holes are in bearing due to misaligned connection plates. Additional mea-
sures are needed to ensure that the loosened nut does not fall off the bolt as a result of
structural vibrations (Wipf et al. 1998).
7.5.1.1.5
Web-Gap Stiffening
Permanently attaching the connection plate to the girder flange can reduce or eliminate
the effects of out-of-plane displacements. Several methods exist for the attachment of
the connection plate to the girder flange: all welded, all bolted, welded and bolted,
adhesives, and nails.

Figure 7.8.  Schematic of diaphragm repositioning retrofit specified on Minnesota Bridge.

Figure 7.8. Schematic of diaphragm repositioning retrofit specified on Minnesota Bridge. 327

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


7.5.1.1.5a
Welded Attachments
The all-welded retrofit for connection plates can be difficult to implement. For in-
stance, the welded connection itself can cause fatigue cracks (Keating et al. 1996). In
addition, it can be very difficult to properly weld high-strength steels as well as flanges
that are embedded in concrete. Although AASHTO now requires transverse welds or
bolted connections on both the girder flanges for the positive attachment of the con-
nection plate, the all-welded attachment retrofit has rarely been specified for reasons
previously mentioned (AASHTO 2012). Figure 7.9 shows a fillet welded connection
plate-to-girder detail.
7.5.1.1.5b
Bolted Connections
The connection plate can be bolted to the girder flange by using angles or sections. It
is important to properly size the angles and tee sections and determine the number of
bolts needed in order to provide for the proper stiffness of the section. Bolted tee sec-
tions are preferred over double angles because they provide greater stiffness (Fisher et
al. 1990). See Figure 7.10.
7.5.1.1.5c
Welded and Bolted Connections
Hybrid connections that use both welded and bolted connections may be more benefi-
cial in areas with certain clearance issues.

Figure 7.9.  Connection plate-to-girder fillet weld detailing.


Source: Keating 2001.

328

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 7.10.  Schematic of concrete deck haunch removal to allow for bolt installation.
Source: Keating 2001.

7.5.1.1.5d
Adhesives
Adhesives become attractive when short-term positive attachments are needed. They
can be less expensive than the bolted or welded options because they do not require
the removal of any concrete (Hu 2005). See Figure 7.11.
7.5.1.1.5e
Nails
Powder-actuated fasteners are the newest and perhaps the best alternative for stiffen-
ing web gaps. These fasteners are made of high-strength materials and are propelled
into the girder flanges by using explosive discharges. Concern has been raised on the
possible fatigue issues of these powder-actuated nails, but research has shown that
the fasteners perform adequately with little detriment to the members. Because of
dimensional issues, nails are only used for the flange connection; bolts are used for
attachments to the connection plate, as shown in Figure 7.12. When determining the
number of nails required, it is imperative that the manufacturer’s recommended nail
shear resistance be used (Niessner and Seeger 1999).

329

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


Leave a minimum 6.4 mm (0.25 inch) gap
between angle/tee and longitudinal weld of
girder to prevent a fretting fatigue problem

Figure 7.11.  Work plan for stiffening retrofit of web gaps with adhesives.
Source: Hu 2005.

Leave a minimum 6.4 mm (0.25 inch) gap


between angle/tee and longitudinal weld of
girder to prevent a fretting fatigue problem

Figure 7.12.  Work plan for web-gap retrofit using nails.

330

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


7.5.1.1.6
Web-Gap Softening
Web-gap softening entails the removal of portions of material to make the web gap
more flexible. See Figure 7.13.
A portion of the connection plate can be removed to increase the size of the web
gap and therefore reduce the stresses resulting from out-of-plane distortion. After
flame cutting the connection plate, it is important to grind the portion of the web
smooth and flush where the connection plate was previously attached.
A simpler and faster approach to softening the web gap would be drilling large
holes in the web of the girder close to the web gap, as shown in Figure 7.14. This
process is much like the smaller-hole retrofits discussed earlier; however, the larger-
diameter holes are able to capture several cracks as opposed to the “Swiss cheese”
method of numerous smaller holes.

Figure 7.13.  Work plan for web-gap softening used on Poplar Street Bridges in East Saint Louis, Illinois.
Source: Koob et al. 1985.

331

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


Figure 7.14.  Schematic of typical large-diameter hole retrofit.
Source: Koob and McGormley 1998.

7.5.2
Tie Girder–Floor Beam Connection
Tied arch bridges exhibit a specific type of web-gap fatigue in the connections between
the floor beams and the tie girders. This fatigue arises from the displacement incom-
patibility between the floor beams that are composite with the bridge deck and the
noncomposite tie girder. The mode of deformation is illustrated in Figure 7.15. Several
retrofits have been implemented in the field and subsequently studied. These studies
are presented in the report (Dexter and Ocel 2013) from which this section has been
adapted.
7.5.3
Cantilever Bracket Cracking
Floor beam cantilevered brackets are used on bridges with large deck overhangs and
can be susceptible to secondary stress fatigue. Several retrofit options exist for reduc-
ing the displacement incompatibility of the girder and floor beams. These retrofits deal
mainly with the modification of the tie plates that span over the girders. The main
idea is either to remove any positive attachments between the girders and tie plates or
to add spacer plates to create a gap between the elements. Figures 7.16 through 7.18
show the deformation modes and possible retrofit options.

332

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 7.15.  Schematic of tie girder-to-floor beam cracking driving force: (a) a generic
deck system of an arch bridge using a tie girder, (b) close-up view of members showing
deformation caused by tie girder rotation, and (c) close-up view of web-gap deformation
of floor beam web.
333

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


Figure 7.16.  Typical cross section of a two-girder bridge with cantilever bracket outriggers.

Figure 7.16. Typical cross section of a two-girder bridge with cantilever bracket outriggers.

Figure 7.17.  (left) Zoomed-in view of tie plate detail and (right) and deformation mode that causes
cracking.

334

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 7.18.  Retrofit of tie plate cracking through addition of spacer plates.

7.6
Retrofit Validation of Secondary Stress Fatigue
Because of the unknown nature of retrofitting web-gap fatigue, it is necessary to vali-
date particular retrofits before retrofitting an entire bridge. The simplest plan to vali-
date a particular retrofit is to first instrument an uncracked detail and then perform
the necessary retrofit and validate whether the retrofit adequately lowered the stress
ranges and out-of-plane displacements. Typical instrumentation includes the use of
strain gauges or displacement measurement devices, or both.
Strain gauges are common and effective instruments used to determine the effec-
tive stress ranges of bridge elements. Strain gauges can be either spot welded or glued
to the element of interest. Figure 7.19 shows preferable strain gauge layouts for retrofit
validation.
Measuring the displacements has become the preferable method for validating
retrofits because displacement measurements are quicker and more cost-effective than
strain measurements. That displacement gauges are only usable for stiffening retrofits
should be taken into account. Two common types of displacement measuring devices
are linear variable differential transformers (LVDTs) and dial gauges. Figure 7.20
shows a schematic for the placement of the devices to measure the web-gap displace-
ment. Maximum distortion-induced fatigue strains and stresses do not always cor-
relate with the largest values of differential deflection, and an instrumentation plan
based primarily on displacements may not capture the whole picture.
In the final analysis, however, the choice of instrumentation to validate the retrofit
type is context sensitive. The designer must select an instrumentation type that best
matches the retrofit type.
335

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


Figure 7.19.  Recommended strain gauge placement for retrofit validation for (top) floor beam–tie
girder connection or beam cope cracking and (bottom) out-of-plane distortion.

7.7
Load-Controlled Fatigue Crack Repair
7.7.1
Coverplates
Several methods have been explored in the retrofitting of coverplates, including grind-
ing, air hammer peening, gas tungsten arc, and bolted splice plates. Grinding has
proven ineffective and is not recommended. Bolted splice plates are an effective option
for girders with severed flanges and in instances in which the listed methods are unde-
sirable. Figure 7.21 shows a splice plate retrofit.

336

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 7.20.  Recommended web-gap displacement instrumentation.

Figure 7.21.  Detailing of splice plate retrofit for cracked coverplate details.

7.7.2
Eyebars and Hangers
Eyebars are long slender bars or rods with forged eyes at the ends that are commonly
used as tension members in truss bridges. Hangers are vertically oriented tension mem-
bers that support load. Because eyebars and hangers are often the sole components
supporting the particular tensile load, they are usually classified as fracture critical
members; inherent flaws can lead to fatigue cracks.

337

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


Assessing the integrity of eyebars, especially in old truss bridges, is very difficult.
An example of an eyebar with extensive corrosion is shown in Figure 7.22.
In the case of truss bridges, adding truss members, rather than replacing them, is
recommended. Removing even a single truss member or connection could easily lead
to bridge failure. In retrofitting many existing old truss bridges, extensive corrosion
of eyebars in tension could be addressed by adding additional tension members while
keeping the existing ones in place. Figure 7.23 shows a possible retrofit alternative for
eyebars in tension in existing truss bridges (Azizinamini 2002).

Figure 7.22.  Corrosion of eyebar


connection.

Figure 7.23.  Retrofit option for


eyebar connection.

338

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


7.7.3
Temporary Tack Welds
Tack welds used to temporarily hold members in place during construction can be
sources of concern for creating fatigue cracks. Fatigue cracks tend to form at the ends
of the tack welds and then propagate into the base metal. The most susceptible types
of tack welds are longitudinal and those located at the end of tack-welded members.
Tack welds can be removed using grinding methods.
7.7.4
Connection Angles
Connection angles used to connect diaphragms to girder webs are susceptible to f­ atigue
cracking as a result of differential girder displacements, as shown in Figure 7.24. The
thickness of such angles needs to be reduced in order to reduce the flexural rigidity
of the elements and negate the causes of the differential girder displacements. Equa-
tion 7.3 helps determine the proper angle thickness to prevent cracking (Dexter and
Fisher 1999):

Figure 7.24.  Deformation mode of connection angles as a result of displacement compatibility.

339

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


 g2 
t ≤ 12   (7.3)
 L
where
t = angle thickness,
g = bolt gauge, and
L = distance between girders.

7.7.5
Web Gusset Plates
Another source of fatigue cracks are web gusset plates, which can experience fatigue
damage in response to weld root defects and locations of intersecting welds. Intersect-
ing welds can be retrofitted by coring holes at the intersections; this procedure will
not only remove the intersecting welds, but also reduce the web constraint. See Fig-
ures 7.25 and 7.26.

Figure 7.25.  (top) Typical cross section of a deep girder bridge and (bottom) plan view of
web gusset detail.
340

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 7.26.  Retrofit detail for gusset plates
with intersecting welds (Crack Site #1).

Fatigue cracks have also been found to initiate from the ends of the gusset plates
and propagate into the girder webs. This issue can be solved using impact treatments
or grinding the weld termination, as shown in Figure 7.27.

1. Grind weld termination using a carbide burr grinder according to figure


below. Make sure to transition the weld tangent to the girder web.

2. Inspect ground area near weld to make sure all weld discontinuities had
been removed.

3. If evidence of discontinuities still exist, grind more weld away until clean
weld is found.

4. Polish ground area with sanding wheel to ANSI roughness of 500 or less.

5. Prime and paint ground area as required.

Figure 7.27.  Detailing of Crack Site #2 retrofit.

341

Chapter 7. FATIGUE AND FRACTURE OF STEEL STRUCTURES


7.7.6
Longitudinal Stiffeners
Fatigue cracks can form at the butt welds of longitudinal stiffeners, primarily as a re-
sult of poor workmanship and inherent defects of the welds. Drilling a large-diameter
hole in the longitudinal stiffener located next to the girder has proven to keep the crack
from developing into the girder web. See Figure 7.28.
7.7.7
Coped Beam Ends
Fatigue cracks have also been observed in coped beam ends. These cracks can be at-
tributed to stresses from flame-cut copes not ground smooth or the presence of bend-
ing moments at these copes, which were designed as simply supported. Square-cut
copes have the least fatigue resistance. Holes can be drilled as a retrofit option, or the
cope can be cut back with a radius.

Figure 7.28.  Work plan of longitudinal butt weld retrofit.

342

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


8
JOINTLESS BRIDGES

8.1
Introduction
A jointless bridge has a continuous deck with no expansion joints over the super-
structure, abutments, and piers. Jointless bridges are commonly referred to as integral
bridges. In this type of bridge structure, all movement due to thermal, creep, and
shrinkage strain is accommodated either within the system itself or at the ends of the
approach slabs where the slabs abut the roadway pavement. Because there are no
joints, ride quality is improved, and maintenance can be greatly reduced.
Leaking deck joints have been a major cause of bridge deterioration and reduced
service life, especially where roadway drainage carrying deicing chemicals can spill
onto the bridge elements below. Elimination of bridge-deck expansion joints is there-
fore an important consideration in bridge system selection to provide long-term service
life, as discussed in Chapter 2.
This chapter summarizes the design, construction, and maintenance provisions
related to the use of jointless bridges.
8.2
History of Jointless Bridges
A detailed history of jointless bridges is provided by Burke, Jr. (2009). The following
is a brief summary of that history. The Ohio highway department was the first to rou-
tinely use continuous construction for multispan bridges, beginning in 1930, although
expansion joints were present at the abutments.
The next step—elimination of deck joints at the abutments—was undertaken by
the Ohio Department of Transportation in 1938 with construction of the Teens Run
Bridge near Eureka, Ohio. The five-span, continuously reinforced, concrete slab bridge
was the first integral bridge in the United States.

343
Use of integral bridges continues to increase both in the United States and abroad.
Several countries adopting the practice are Japan (1996) and South Korea (2002), with
the United Kingdom recently using integral bridges in routine applications. A survey
conducted in 1987 indicated 20 of the 30 respondent transportation departments were
using integral construction details for continuous bridges.
One point of contention is the length limit for which integral systems can be
applied. North Dakota, South Dakota, and Tennessee have used continuous integral
bridges with steel main members at spans in the 300-ft range for many years. Similar
bridges with concrete main members have been constructed at lengths of 500 to 800 ft
in Kansas, California, Colorado, and Tennessee.
The Tennessee DOT is in the lead when it comes to span length. The Long Island
Bridge of Kingsport, built in 1980, has 29 continuous spans without deck expan-
sion joints. Additionally, Tennessee recently completed the seven-span Happy Hollow
Creek Bridge, which is a curved, prestressed concrete bridge with a total length of
1,175 ft. Despite its extreme length, the structure falls within the Tennessee DOT’s
policy for integral bridges.
Seamless bridges, another type of jointless bridge introduced by SHRP 2 Project
R19A for practice in the United States, allow elimination of expansion joints even at the
end of the approach slab. The seamless bridge system was first introduced by Bridge et
al. (2000) in Australia for use with continuously reinforced concrete pavement for the
approach roadways. Most commonly used pavements in the United States, however,
are jointed plain concrete (JPCP) and flexible pavements, which require a modified
application. Seamless bridges do not have any joints, even at the ends of an approach
slab (hence seamless). Instead, a pavement transition zone is used to dissipate the
­thermal displacements of the bridge. The transition zones can be rather lengthy rela-
tive to the bridge. The benefit, however, is that movements at the end of the transition
zone are very small.
8.3
Types of Jointless Bridges
Three main types of jointless bridges are described in this chapter: integral and semi-
integral jointless bridges, which are commonly used in practice, and a new class of
jointless bridges referred to as seamless jointless bridges. The main characteristic of the
seamless system is that expansion joints are eliminated altogether, and the bridge deck
is connected to the approach road pavement without a joint.
Figure 8.1 is a rendering of a typical layout of a jointless bridge, shown with the
superstructure cast in an integral abutment.
8.3.1
Integral Bridges
Integral bridges have the superstructure constructed monolithically with the abutments,
encasing the ends of the superstructure within the backwall. The main characteristics
of integral bridges are their jointless construction and flexible abutment foundations.
The system is structurally continuous, and the abutment foundation is flexible longitu-
dinally. The movement of the superstructure is accommodated by the foundation.

344

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 8.1.  Elements of jointless integral bridges.
Figure 8.1. Elements of jointless integral bridges.

Figure 8.1 shows, schematically, the main elements of an integral bridge system,
which consist of bridge deck, girders, integral cast abutments, and approach slabs. The
bridge movement is accommodated at the ends of the approach slabs. Sleeper slabs are
commonly used to provide vertical support for the ends of the approach slab where
the slabs abut the roadway pavement (not shown in Figure 8.1). In addition, jointless
integral bridges can be continuous multispan structures with intermediate piers (also
not shown in Figure 8.1). Various details are described in greater detail in Section 8.7;
see Section 8.7.3 for further discussion of sleeper slabs.
8.3.2
Semi-Integral Bridges
Semi-integral bridges are defined as having an end diaphragm that serves as the abut-
ment backwall and that is cast encasing the superstructure ends. In this system, the
superstructure rests on expansion bearings, and the end diaphragm is not restrained
longitudinally with respect to the pile cap or abutment stem. The deck may be slid-
ing or cast monolithically with the backwall, but it does not have a joint above the
abutment. The foundation is rigid longitudinally, where superstructure movement is
accommodated through bearings.
The main elements of a semi-integral bridge system consist of bridge deck, girders,
abutment stem and bearing seat, integral cast diaphragm backwall, approach slab, and
sleeper slab. The bridge movement is accommodated at the ends of the approach slabs.
Greater detail is provided in Section 8.7.
8.3.3
Seamless Bridges
The seamless bridge system is characterized by eliminating the need for expansion
joints, even at the ends of the approach slabs, while limiting the longitudinal expan-
sion and contraction of the bridge superstructure. Imposing a limitation on longitudi-
nal expansion and contraction of the bridge superstructure results in development of
longitudinal forces that need to be resisted with appropriate design features. Longitu­
dinal expansion and contraction are limited, but not eliminated. This philosophy is

345

Chapter 8. JOINTLESS BRIDGES


used to reduce the level of longitudinal forces that can be developed, while making
the gap between the end of the transition zone and the start of pavement manageable
(to less than about 0.25 in.). The foundation requirements are very similar to those of
integral abutments. A seamless bridge system for jointed pavement types used in the
United States was developed by SHRP 2 Project R19A and is described in Appendix E.
Additional information on this system can be found in Ala (2011) and in Ala and
­Azizinamini (in press a and b).
Figure 8.2 shows, schematically, the main elements of the seamless bridge system,
which consist of the bridge deck, transition zone, and roadway pavement. The bridge
movement is accommodated within the transition zone, and the movement at the end
of the transition zone is relatively small. This eliminates having expansion joints at
the end of transition zone, where the roadway pavement starts. The thickness of the
transition zone and approach slab near the abutment is increased to account for lack
of support from the soil below. The assumption is that the transition zone near the
abutment has no support and resists the imposed loads by flexure.
8.3.4
Advantages of Jointless Bridges
Henry Derthick, former engineer of structures at the Tennessee DOT, once stated,
“The only good joint is no joint.” In keeping with this statement, known advantages
of the jointless bridge systems include

• Lower initial cost;


• Lower maintenance cost;
• Prevention of leakage of moisture to bridge elements below deck, resulting in
­longer service life;
• Improved ride quality;

Approach
Transition Zone
Bridge

JPCP
Abutment

Figure 8.2.  Seamless bridge system.

346

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


• Easier and faster construction;
• Easier inspection;
• Simplified bridge details;
• Elimination of bearings (except for semi-integral bridges);
• Ideally suitable for bridges with skew and curvature or located in high seismic
areas; and
• Enhanced buoyancy resistance of the bridge.

Because of these advantages, many DOTs have started using jointless bridges;
however, the design provisions vary significantly from one state to another.
8.3.5
Cost-Effectiveness of Jointless Bridges
Jointless bridges have a significant cost savings advantage compared with traditional
bridges with expansion joints. Cost savings are realized both during initial construc-
tion and throughout the life of the bridge with reduced maintenance. This is particu-
larly true for bridges with integral abutments.
Most components of typical bridges with joints and jointless bridges are similar in
construction and cost (e.g., deck, beams, cross frames). Thus, a comparison is made
relative to the different components that distinguish each type of construction (e.g., the
costs of the abutments and expansion joints). In addition, because unit pricing of each
item is consistent neither from region to region nor over time, a qualitative comparison
is made using relative costs.
It is recognized that different states and municipalities have different specifications
and construction techniques; however, initial construction of a typical abutment with
an expansion joint will most often include the following:

• Excavation;
• Two rows of piling (in some cases);
• Concrete cap, stem, diaphragm, and backwall with reinforcing (three pours);
• Elastomeric bearings, per beam;
• Precision-cast bridge seats for bearings;
• Expansion joint; and
• Porous backfill.

Similarly, typical construction of an integral bridge includes

• Excavation;
• One row of piling;
• Concrete cap, integral backwall, and diaphragm (two pours);
• Sleeper slab with reinforcement; and
• Porous backfill.

347

Chapter 8. JOINTLESS BRIDGES


The important differences between an integral abutment and a traditional abut-
ment with a jointed deck include a lack of expansion joint, no bearings, reduced num-
ber of required piles, reduced number of concrete pours, and inclusion of a sleeper
slab. Taking these differences into consideration, the cost savings are readily apparent.
The sleeper slab adds a few cubic yards of concrete and an extra detail to the cost, but
removal of the expansion joint, removal of the second row of piling (overall reduced
number of piles), reduction in the required amount of concrete and reinforcing in the
stem–backwall unit, and elimination of beam bearings at the abutment greatly reduce
the cost relative to adding the sleeper slab. The overall reduction in initial construction
cost can be over 40% for each abutment. (The percentage difference was estimated
using Ohio 2010 bid planning costs for a typical 32-ft-wide bridge.)
The life-cycle cost of the two types of abutments differs, as well. A service life of
100 years is used for the comparison, although it is acknowledged that differences in
estimated service life can affect the parameters. For standard jointed bridges, common
armored expansion joints typically require gland replacement on the order of 8 to 12
years, depending on condition severity. Additionally, the entire joint including armor
will need to be replaced along with the deck at least once (based on an estimated deck
life of 30 to 50 years). Again, this estimate depends on the severity of the conditions.
It is also expected that the bearings will need replacing at least once over the life of the
bridge, the cost of which includes jacking of the bridge.
Similar to expansion joints, the sleeper slab joint seal, if used, will require replace-
ment on the same, or at least similar, schedule. Likewise, the deck will need replacing
on a similar schedule. Deck replacement of an integral bridge requires additional con-
sideration of certain construction items, but it does not require a significant increase in
construction cost compared with traditional deck replacement.
For comparative purposes, consider that a typical bridge abutment with expansion
joints will require

• Expansion gland replacement every ~10 years,


• Deck replacement every ~50 years,
• Expansion joint replacement every ~50 years, and
• Bearing replacement every ~50 years.

A typical jointless bridge abutment will require

• Sleeper slab joint seal (if used) replacement every ~10 years and
• Deck replacement every ~50 years.

Cost is similar for deck replacement and gland and seal replacement; however,
there is a significant increase in cost when replacing the expansion joint.
A qualitative cost comparison is presented in Figure 8.3 and Figure 8.4. Both fig-
ures consider the difference in the initial cost at Year Zero and the accumulated cost
difference over the life of the bridge. The figures show the difference in costs; that
is, similar costs have been removed from the equation (i.e., the cost of replacing the
deck itself is removed from the equation because it is similar for both types of bridge).

348

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 8.3.  Lifetime cost analysis
Expansion Joint Replacement, of jointed versus jointless bridge over
50-yr cycle time.
Joint Repair,
10-yr cycle
Cost difference

Jointless
Relative Total Cost

Abutment Construction
With Exp. Joints

0 20 40 60 80 100

Life of Bridge (yr.)

Figure 8.4.  Lifetime cost differential


analysis of jointed versus jointless
Jointless
bridge.
With Exp. Joints

Relative Cost

Abutment Joint Maintenance Total Difference


Construction

Figure 8.3 shows the estimated cost comparison through the 100-year life of the struc-
ture. Figure 8.4 shows the cost comparison differentiating the initial difference in the
construction costs, the lifetime maintenance costs, and the overall total cost difference
over the life of the bridge.
8.4
Factors Affecting Performance of Jointless Bridges
Factors affecting the performance of jointless bridges include curvature, skew, bear-
ing, and connection of superstructure and substructure. Other factors that should be
considered include site conditions, deterioration of piles, and abutment walls.
8.4.1
Curvature
Horizontal curvature changes the internal forces of integral abutment bridges. These
changes are more important for bridges with either of the following conditions when
the length and radius are measured at the centerline of the bridge: length-over-radius
ratio greater than 0.5 or radius of curvature less than 1,000 ft.

349

Chapter 8. JOINTLESS BRIDGES


If a curved bridge does not have either of these conditions, the response of the
curved bridge can be estimated by the response of a straight bridge of the same length
(Doust 2011). This estimation is not valid for the internal forces during construction.
8.4.2
Skew
In skewed integral abutment bridges, the soil passive pressure developed in response
to thermal elongation can prevent the transverse movement of the bridge. Appendix B
provides additional detail on this subject. However, if the friction created by contact
between soil and abutment wall is insufficient, and depending on the transverse stiffness
of the abutment, either significant transverse forces or significant transverse movements
could be generated (Oesterle et al. 2005).
8.4.3
Bearings
In multispan integral bridges with rigid piers, the superstructure is commonly seated
on piers through the use of bearing devices. In curved bridges or wide, straight bridges,
fixed bearings are not recommended except at the points of zero movement. In curved
integral bridges, there may be no point of zero movement throughout the bridge.
Guided bearings are not recommended for curved or wide, straight integral bridges
because the displacements do not happen in just one particular direction. In such cases,
guided bearings behave like fixed bearings, creating large internal forces at the piers.
Multidirectional elastomeric or sliding bearings are the proper types of pier bear-
ings for integral bridges. If such bearings are used, the superstructure movement is
mainly controlled by the integral abutments.
8.4.4
Connection Between Superstructure and Substructure
The choice of how the superstructure is connected to the substructure has a significant
impact on how the bridge will behave. Choosing the abutment type sets the major
design considerations for the bridge with respect to jointless behavior. Methodology
for designing the abutments for the various types of jointless bridges is presented in
Section 8.6.
The consideration for the connection to the piers is equally important. The super-
structure can be made integral with a pier or designed to transfer loads to the pier with
more traditional assumptions. It is important to note in the planning stages how the
pier will react as the bridge expands and contracts. Piers must be sufficiently designed,
whether they are intended to flex with the structure (slender piers) or are designed to
resist the movement (stout piers). The latter case is generally not preferable as it often
leads to overdesigned substructures, because the movement from the continuous deck
superstructure can generally be accommodated by simply using an expansion bearing
to accommodate the movement. The design of different pier types is discussed in more
detail in Section 8.6.2.9.

350

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


8.4.5
Other Considerations
Other factors that can affect the performance of jointless bridges are primarily associ-
ated with foundation conditions.
8.4.5.1
Site Condition
Integral abutments for jointless bridges are usually supported on a single row of piles to
provide flexibility. Also, piles are typically used to minimize settlement of the abutment
and differential settlement within the superstructure. However, when rock is close to
the substructure bearing surface, a different type of foundation may be required. One
solution is to use semi-integral abutments, described in Section 8.3.2, in which the
abutment foundations are supported directly by and keyed into the rock. The end dia-
phragm serving as the abutment backwall and encasing the superstructure ends rests
on bearings supported by the abutment foundations. The deck and approach slabs are
cast monolithically with the backwall. The abutment foundations are rigid, and the
longitudinal movement of the superstructure is accommodated through the bearings.
As an alternative to the semi-integral abutments, spread footings may potentially
be considered for integral abutments when rock is close to the surface, particularly
for single-span bridges, and when the foundation is assumed to slide. However, sig-
nificant friction forces would have to be overcome, and this concept has typically
not been considered. Differential settlement would be another concern for the use of
spread footings on soils to support abutments for multispan continuous bridges, but
Moulton et al. (1985) and Hearn (1995) indicate that the magnitude of settlement for
abutments supported by spread footings is similar to that for abutments supported by
piles. However, there is very little experience with the actual use of spread footings for
integral abutments either on rock or on competent soil near the surface. Hence, it is
recommended that experience be gained by starting with relatively short simple-span
bridges. Use can then progress to longer structures and multispan structures as success-
ful experience is gained.
The following recommendations pertain to abutments supported by shallow
spread footings, in which end movement may be accommodated by sliding:

• For footings founded on rock, a layer of granular fill should be used (on top of a
leveling layer of fill concrete, as needed) between the footing and rock to facilitate
sliding. The footing should not be keyed into rock.
• The abutment wall should be designed for shear and moments resulting from both
expansion and contraction movements. The resistance to contraction should in-
clude friction on the bottom of the footing and passive soil pressure from the berm
soil on the front face of the abutment.
• Sufficient drainage, distance from the face of the slope, and slope protection are
essential to keep soil from washing out below the footing. For footings supported
on a layer of granular soil for sliding on rock, use of geotextile material may be
considered to contain the granular soil. For footings supported on soil, mechanical
stabilization of the soil below the footing may be appropriate.

351

Chapter 8. JOINTLESS BRIDGES


Another possible solution for use in conditions in which rock is close to surface is to
drill large-diameter holes in the rock and use piles, which would consequently allow the
use of typical integral abutment construction. It must be noted that the concepts of inte-
gral abutments supported on spread footings or supported on piles placed in holes drilled
into rock are not common practice. These two concepts are suggested for consideration
when site conditions would otherwise inhibit the use of typical jointless construction.
8.4.5.2
Deterioration of Piling
Accelerated pile deterioration is generally not considered except in specialized cor-
rosive locations. Designers should consult with either a geotechnical engineer or ge-
ologist to mitigate possible impacts for this condition. More commonly, corrosion is
thought of as a minimal concern for piles, but it has been recorded (Beavers and Durr
1998) and more recently evaluated (Decker et al. 2008). Additionally, the state of
Iowa has been investigating deterioration of piles just below the pile cap of integral
abutments (Iowa Department of Transportation 2010)). Initial results note that Iowa
investigators have discovered corrosion immediately below abutment footings of what
would be considered normal conditions.
Piling deterioration is of increased importance for integral abutments because of
the additional strains placed on the substructure from the longitudinal expansion
of the superstructure. The potential for section loss based on soil conditions should
be accounted for as presented by Article 10.7.5 of the LRFD Bridge Design Specifica-
tions (LRFD specifications), which states minimum considerations for the effects of
corrosion and deterioration of piling (AASHTO 2012). Adhering to these guidelines
should provide sufficient protection against advanced corrosion and thus failure of the
integral abutment system.
8.4.5.3
Jointless Bridge Abutments with Mechanically Stabilized Earth Walls
If setting the abutment on top of an embankment slope or reducing the total bridge
length is impractical, full-height abutments with a mechanically stabilized earth (MSE)
retaining wall may be considered in the design of jointless bridges. When MSE walls
are used, steps must be taken to prevent excess pressure on the retaining wall intro-
duced by the movement of the backwall and pile.
For integral abutments, per FHWA Demonstration Project 82 (Elias et al. 1997)
the horizontal force and its distribution with depth may be developed using pile load-­
deflection methods (p-y curves) and added as a supplementary horizontal force to be
resisted by the MSE wall reinforcements. This force will vary depending on the level
of horizontal load, pile diameter, pile spacing, and distance from the pile to the back of
the panels.
Per Demonstration Project 82, the following additional design details have been
used successfully: first, providing a clear horizontal distance of about 1.5 ft (0.5 m)
between the back of the panels and the front edge of the pile; and second, providing
a casing around the piles, through the reinforced fill, where significant negative skin
friction is anticipated.

352

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Where pile locations interfere with the reinforcement, specific methods for field
installation must be developed. Simple cutting of the reinforcement is not permissible.
For integral abutments and for seamless bridges, MSE walls can still be used, but
they must be sufficiently isolated from the soil movement caused by the movement of
the piles. Alternatives suggested by Nicholson et al. (1997) for the use of MSE walls
with jointless bridge abutments are shown in Figure 8.5.
Figure 8.5a illustrates the use of a semi-integral abutment or stub integral abut-
ment on spread footings. In this approach, the MSE reinforcement should be designed
for the sliding forces in the bearings of the semi-integral abutment or the frictional
sliding forces of the spread footings.

Figure 8.5. Alternatives to integral full-height wall abutments using reinforced-soil retaining structure.
Source: Nicholson et al. 1997.

353

Chapter 8. JOINTLESS BRIDGES


Figure 8.5b illustrates use of pile encased in a pressure-relieving sleeve that isolates
the pile movements from the surrounding soil. Hassiotis (2007) has reported tests with
an integral abutment supported on piles encased in corrugated steel sleeves backfilled
with sand. Lui et al. (2005) indicate that the Iowa DOT criteria for use of MSE walls
with integral abutments requires each pile to be encased in a corrugated metal sleeve.
The reinforced soil should include sand up to the bottom of the sleeve, and the remain-
der of the sleeve should be filled with bentonite to the top.
Figure 8.5c illustrates the use of a semi-integral abutment supported on a pier in
front of the MSE wall. No additional considerations are necessary for semi-integral
abutments because the lateral movement is dissipated through the bearings.
8.5
Strategy Selection Process
Each type of jointless construction has a range of parameters that is appropriate for
particular bridges or provides various advantages over another type of system. The fol-
lowing tables provide guidance in selecting bridge type based on limiting parameters.
Table 8.1 assists in the selection of the primary system and provides three options with
regard to foundation types: integral, semi-integral, and seamless. The maximum length
of each system is not set; rather, it is based on design calculations. Typical details
that could be employed with each system are provided, and corresponding sections
are noted. Table 8.1 also briefly describes the major advantages and disadvantages of
each system. Relatively, the semi-integral abutment type provides larger longitudinal
movement capabilities than the integral and seamless systems. The trade-off is the
need to add sliding bearing, which will result in reduction in service life. As indicated
in Table 8.1, the relative maintenance of integral and seamless abutment types is lower
than maintenance costs for the semi-integral abutment type. The main reason for the
difference in cost is the need for incorporating bearing at the abutment. All three
abutment types are applicable to existing bridges when it is desirable to eliminate the
expansion devices at the abutment. In some situations, cost may prevent use of integral
or semi-integral abutment types.
Table 8.2 provides further guidance on the substructure type that is appropriate
for use with each type of jointless bridge. As indicated in Table 8.2, for the integral
abutment type, H-piles or prestressed piles or concrete-filled tube (CFT) piles could
be used. In the case of prestressed piles, the relative lateral displacement movement is
lower than H-piles or CFT piles. In the case of prestressed piles, cracking and corro-
sion is of concern. In the case of H-piles and CFT piles, corrosion is of concern. These
concerns are reflected qualitatively in assessing the potential for each abutment type to
achieve 100 years of service life.
Table 8.3 provides guidance on the types of connections and bearings used at the
piers when used in a jointless bridge. Considering the discussions provided in previ-
ous sections, the four right-hand columns of Table 8.3 provide qualitative rankings
of each option with respect to maximum longitudinal movement capabilities, relative
maintenance cost, applicability to existing bridges, and potential to achieve 100+ year
service life.

354

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 8.1.  Strategy Table for Abutment Type Selection in Jointless Bridges: Straight Bridges
Qualitative
Maximum Longitudinal Qualitative Applicability
Bridge Length Typical Movement Maintenance to Existing
Strategy Used Details Advantages Disadvantages Demand Ranking Bridges
Integral Established Figure 8.31, Eliminates need for Difficult to inspect Low Low Yes
based on design Figure 8.33, bearings. damage to piles due to
provisions stated Figure 8.34, pile movement.
in the Guide Figure 8.35
Semi- Established Figure 8.36, Provides reduced Needs bearings. Medium Medium Yes
integral based on design Figure 8.37, longitudinal force transfer
provisions stated Figure 8.38 to piles.
in the Guide
Seamless Established Figure 8.25, Eliminates need for Possible, initial higher Low Low Yes
based on design Figure 8.26 expansion joints. cost. Difficult to inspect
provisions stated Eliminates concerns when damage to piles due to
in the Guide there is skew or curvature. pile movement; long
Eliminates need for transition zone off bridge.
bearings and sleeper slab.

355

Chapter 8. JOINTLESS BRIDGES


356
Table 8.2.  Strategy Table for Foundation at Abutments in Jointless Bridges: Straight Bridges

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Qualitative Potential for
Longitudinal Achieving
Typical Movement 100+ Years of
Strategy Details Advantages Disadvantages Demand Service Life
Integral abutment H-pile Figure 8.33 Economical for small movements. Relatively low Medium Medium
Easy to construct. strength, ductility and
buckling capacity.
Prestressed Figure 8.32 Very stiff and high axial load capacity. Prone to concrete Low Medium
pile deterioration and
corrosion of strands.
CFT pile Figure 8.32 CFT has high strength and ductility and Higher initial cost. High Medium
higher buckling capacity, which will
accommodate much larger bridge length.
Semi-integral abutment Any foundation type could be used with the semi-integral abutment.
Seamless The strategy for selection of seamless foundations is the same as for integral abutments. The detailing for seamless bridges is
primarily in the transition zones at the ends of the bridge as detailed in Section 8.6.
Table 8.3.  Strategy Table for Connection Between Piers and Superstructure in Jointless Bridges: Straight Bridges
Degree of Potential for
Qualitative Difficulty Achieving
Longitudinal Qualitative to Apply 100+ Years
Detail Movement Maintenance to Existing of Service
Strategy Figure Advantages Disadvantages Demand Ranking Bridges Life
Girders Integral- Figure 8.27a, Eliminates need May cause Low Low Medium High
continuous frame action Figure 8.29 for bearings over transverse
over pier pier. cracking in the
pier.
Fixed bearing Figure 8.27b No longitudinal May cause Low Medium Medium Medium
(rotational movement transverse
movement requirement for cracking in the
allowed) bearing over pier. pier.
Expansion Figure 8.27c Reduced bending Bearing designed High Medium Low Low
bearing of pier column. for both rotation
and longitudinal
movement.
Girders not Simple for Figure 8.27b, Eliminates joints Restraint Varies with Low Low High
continuous dead and Figure 8.27c, and protects moments and bearing type
over pier continuous for Figure 8.28, girder ends; cracking in
live load Figure 8.30 viable option for diaphragms.
seismic retrofit.
Link slab Figure 8.27d, Low cost. May crack and Varies with High Low Medium
Figure 8.31 cause leakage bearing type
over joint.

357

Chapter 8. JOINTLESS BRIDGES


8.6
Design Provisions for Jointless Bridges
Design procedures for integral abutment bridges can range from a simplified method
of analysis to a more detailed approach. This section includes provisions for both. Sec-
tion 8.6.1 provides requirements for using the simplified approach, and Section 8.6.2
describes detailed methods of analysis that should be used if Section 8.6.1 require-
ments are not met.
8.6.1
Simplified Method of Analysis
The simplified analysis method is provided to eliminate many design steps for simple
bridges that do not require detailed analysis. A bridge should meet the following re-
quirements for use of the simplified analysis method (VTrans 2009):

• The skew angle should be less than or equal to 20°;


• The bridge can be straight or curved, but with parallel girders;
• Abutments and piers should be parallel;
• Abutment height should be limited to 13 ft.;
• Heights of the abutment at the bridge ends should be approximately the same
(maximum difference 20%);
• The slope of the bridge in the longitudinal direction should be less than 5%;
• The length of the wing wall, attached to the abutment, should be less than 10 ft;
and
• The length of the pile should be greater than or equal to 16 ft.

The main characteristics of the simplified method of analysis are as follows:

• The internal forces of the superstructure and substructure are obtained using a
two-dimensional (2D) analysis.
• Conservatively, the superstructure may be assumed to be simply supported at the
two abutment ends.
• The design of the pile can be accomplished by separating the pile from other bridge
elements and treating it as an axial member. The moment capacity of the pile
section is affected by the applied axial load. As the pile axial load increases, the
moment that causes the formation of the plastic hinge in the pile will decrease.
Once the plastic hinge is formed, the pile head can be assumed to act as a pin. In
the simplified approach the pile is modeled as an axial element with one end (the
end close to the abutment) subjected to axial load and constant moment equal to
the moment capacity of the pile cross section. The interaction equation in LRFD
specifications Article 6.9.2.2 can be employed to determine the capacity of the pile
section.

358

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


8.6.2
Detailed Method of Analysis
If the requirements for the simplified method of analysis are not met, the bridge must
be analyzed using the detailed analysis approach. In this approach, there is no limita-
tion on total skew angle and so forth. For years, jointless integral abutment bridges
were designed with an imposed maximum limitation on total bridge length, with max-
imum length of steel structures less than that of concrete. Design provisions for the
detailed method of analysis, as outlined in this chapter, do not include bridge length
limitations; rather, one must meet the specified design provisions.
In the detailed method of analysis, the superstructure and substructure are m
­ odeled
as an integral system using three-dimensional (3D) finite element analysis, with girder
webs modeled using shell elements. Use of grid-type analysis should be carried out
with caution and is not recommended, primarily because the torsional stiffness of line
elements in some of the available commercial programs does not include the contribu-
tion of warping torsional stiffness.
8.6.2.1
Loads
8.6.2.1.1
Dead Loads
Dead loads include the weight of all components including superstructure and sub-
structure elements and include all permanent loads in accordance with LRFD speci-
fications Article 3.5. The dead loads are distributed to the foundation through tradi-
tional assumptions or in accordance with the owner’s bridge design provisions.
8.6.2.1.2
Live Loads
Live loads and their associated impact are applied in accordance with LRFD speci-
fications Article 3.6 or in accordance with the owner’s bridge design provisions. For
integral abutments and piers, application of live loads will cause rotation and induce
moments that will need to be considered in the design.
Horizontal live load (braking force and centrifugal force) are subject to distribu-
tion with respect to the stiffness of the integral and semi-integral abutments. In tradi-
tional design, longitudinal forces are distributed to the substructure based on bearing
fixity (expansion versus fixed against horizontal movement) and relative substructure
flexibility. For jointless bridges, the backfill is in full contact with the end diaphragm
(backwall) and provides a significant amount of stiffness relative to the other substruc-
ture components. For integral abutments, in which the bearing condition is fixed, it is
acceptable to assume for bridges with one to three spans that the longitudinal forces
are absorbed by the passive pressure and stiffness provided by the backfill soil. This
should be verified by a geotechnical engineer. As bridges get longer and additional
substructure units are introduced, a relative stiffness analysis should be performed.
However, even with multiple piers with some having fixed-expansion bearings, inte-
gral abutments can be expected to absorb as much as 80% of the longitudinal force.

359

Chapter 8. JOINTLESS BRIDGES


8.6.2.1.3
Soil Load on Abutment
The magnitude of soil pressure behind the abutment wall and the nonlinear distribu-
tion of this pressure depend on wall displacement, soil type, depth, pile stiffness, and
also the direction of the displacement (Faraji et al. 2001). As a wall moves toward
the backfill, passive pressure is engaged, and when it moves away, active pressure and
surcharge pressure may be generated.
Full passive pressure builds up for relatively long bridge lengths. For shorter bridge
lengths, only part of the passive pressure is developed for expansion because thermal
expansion is limited. For all bridges, the maximum passive pressure force is calculated
as shown in Equation 8.1:

Pp = 1 2 Kpγ H 2 (8.1)

where
Pp = passive pressure force,
Kp = passive pressure coefficient,
g = unit weight of soil, and
H = height of soil face.

Kp is not necessarily the maximum Kp associated with full passive pressure. The
value of Kp should be calculated using Figure 8.6 and Figure 8.7 (Clough and D ­ uncan
1991). The extreme values for expansion and contraction are proportional to the
height of the wall. The movement required to reach the maximum passive pressure is
on the order of 10 times the movement required to reach the active soil pressure. The
movement required to reach the extreme pressures are larger for loose soils than that
for dense soils (Figure 8.6 and Figure 8.7, respectively). Table 8.4 highlights the move-
ments required to achieve maximum pressures.
The force-deflection relationship should be based on the design curves (Barker et
al. 1991) shown in Figure 8.6 and Figure 8.7 (Clough and Duncan 1991). The stiffness
of the springs behind the abutment wall is nonlinear and depends on the type of soil.
8.6.2.1.4
Soil Load on Piles
The design of piles should consider the soil–structure interaction by using p-y curves
such as in the procedure recommended by the American Petroleum Institute for off-
shore platform design (API 1993).
Soil–structure interaction analysis of piles can be performed using available soft-
ware: LPILE, COM624P, and FB-MultiPier are several that use this approach. Further
information on this topic is provided in Article 10.7 of the LRFD specifications.

360

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 8.6.  Relationship between wall movement and earth pressure.
Source: Clough and Duncan 1991.

8.6.2.1.5
Thermal Loads
In order to account for the effect of temperature changes in the design of jointless
bridges, two effects should be considered: the effect of uniform temperature change
and the effect of temperature gradient within the structure. These two effects are ex-
plained in the following subsections.
8.6.2.1.5a
Uniform Temperature Change
The calculation of uniform temperature changes should be in accordance with LRFD
specifications Article 3.12.2, in which two procedures are recommended, Procedure
A and Procedure B, as described in this subsection. Either procedure may be used for
concrete deck bridges that have concrete or steel girders. For all other types of bridges,
Procedure A should be employed.
Table 8.5 presents the temperature ranges used in Procedure A to calculate the
design thermal movements. The difference between these values and the base construc-
tion temperature should be used to calculate thermal movements.

361

Chapter 8. JOINTLESS BRIDGES


Figure 8.7. Relationship between wall movement and earth pressure for a wall with compacted backfill.
Source: Clough
Source: Clough andand Duncan
Duncan 1991.(1991).

Figure 8.7. Relationship between wall movement and earth pressure for a wall with compacted backfill.
Table 8.4.  Approximate Magnitudes of Movements Required to Reach Extreme
Soil Pressure ­Condition
Values of Δ/H
Type of Backfill Active Passive
Dense sand 0.001 0.01
Medium-dense sand 0.002 0.02
Loose sand 0.004 0.04
Compacted silt 0.002 0.02
Compacted lean clay a
0.01 0.05
Compacted fat claya 0.01 0.05

Source: Clough and Duncan 1991.


Note: Δ = movement of top of the wall required to reach extreme soil pressure, by tilting or
lateral translation; H = height of the wall.
a
Under stress conditions close to the minimum active or maximum passive pressures, cohesive
soils creep continually. The movement shown would produce temporary passive pressures. If
pressures remain constant with time, the movements shown will increase. If movement remains
constant, active pressures will increase, while passive pressures will decrease.
362

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 8.5.  Procedure A: Temperature Changes
Climate Steel or Aluminum Concrete Wood
Moderate (°F) 0 to 120 10 to 80 10 to 75
Cold (°F) –30 to 120 0 to 80 0 to 75

Source: LRFD specifications Table 3.12.2.1-1.

Procedure B considers the range of temperature change, which is the difference


between maximum design temperature and minimum design temperature. The maxi-
mum design temperature for concrete girder bridges with concrete deck is provided by
Figure 8.8, and the minimum design temperature is given in Figure 8.9. The maximum
and minimum design temperatures for steel girder bridges are given in Figure 8.10 and
Figure 8.11, respectively.
8.6.2.1.5b
Temperature Gradient
The effect of temperature gradient may typically be ignored; however, if the designer
decides to consider the effect of temperature gradient, the following provisions, taken
from LRFD specifications Article 3.12.3, are recommended. The profile of the tem-
perature in steel and concrete girder bridges may be taken as shown in Figure 8.12, in
which t is the thickness of the concrete deck. Dimension A in Figure 8.12 should be
taken as follows:

• 12.0 in. for concrete superstructures deeper than 16 in.;


• Depth of superstructure minus 4.0 in. for concrete superstructures shallower than
16 in.; and
• 12.0 in. for steel superstructures.

Figure 8.8.  Maximum design temperature for concrete girder bridges.


Source: LRFD specifications Figure 3.12.2.2-1.
Source: LRFD Specifications Figure 3.12.2.2-1.
363
Figure 8.8. Maximum design temperature for concrete girder bridges.
Chapter 8. JOINTLESS BRIDGES
Figure 8.9.  Minimum design temperature for concrete girder bridges.
Source:
Source: LRFDLRFD Specifications
specifications Figure 3.12.2.2-2.
Figure 3.12.2.2-2.

Figure 8.9. Minimum design temperature for concrete girder bridges.

Figure 8.10.  Maximum design temperature for steel girder bridges.


Source:
Source: LRFDLRFD Specifications
specifications Figure 3.12.2.2-3.
Figure 3.12.2.2-3.

Figure 8.10. Maximum design temperature for steel girder bridges.


The values for T1 and T2 are given in Table 8.6 and vary by solar radiation zone as
determined from the map shown in Figure 8.13. The values in Table 8.6 are positive
temperature values. The negative temperature values are obtained by multiplying the
values from the same table by –0.3 for plain concrete decks and by –0.2 for decks with
asphalt overlay. The value of T3 should be taken as 0°F, unless a specific field study is
carried out to determine this value, in which case T3 should not exceed 5°F.

364

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 8.11. LRFD
Source: Minimum design temperature
Specifications for steel girder bridges.
Figure 3.12.2.2-4.
Source: LRFD specifications Figure 3.12.2.2-4.
Figure 8.11. Minimum design temperature for steel girder bridges.

Figure 8.12.  Positive vertical temperature gradient in concrete


and steel superstructures.
Source: LRFD specifications Figure 3.12.3-2.

Table 8.6. Basis for Temperature Gradients


Zone T1 (°F) T2 (°F)
1 54 14
2 46 12
3 41 11
4 38 9

Source: LRFD specifications Table 3.12.3-1.

365

Chapter 8. JOINTLESS BRIDGES


Figure 8.13.  Solar radiation zones for the United States.
Source: LRFD specifications Figure 3.12.3-1.

When considering the temperature gradient in the section profile, the analysis
should consider axial extension, flexural deformation, and internal stresses (LRFD
specifications Article 4.6.6). The response of the structure to temperature gradient
can be divided into three parts: axial expansion, flexural deformation, and additional
stresses. These components are discussed in this subsection.
Axial expansion is due to the uniform portion of the temperature gradient and
can be calculated as shown by Equation 8.2 (LRFD specifications Equation C4.6.6-1):

1
TUG =
Ac ∫ ∫ TG dw ⋅ dz (8.2)
where
TG = temperature gradient (Δ°F),
TUG = temperature averaged across the cross section (°F),
Ac = cross-section area transformed for steel beams (in.2),
w = width of element in cross section (in.), and
z = vertical distance from center of gravity of cross section (in.).

The corresponding uniform axial strain (εu) is then taken as Equation 8.3 (LRFD
specifications Equation C4.6.6-2):

ε u = α (TUG + TU ) (8.3)

where α is the coefficient of thermal expansion (in./in./°F), and TU is the uniform speci-
fied temperature (°F).

366

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


The consequence of a temperature gradient is flexural deformation, the develop-
ment of curvature (f) over the cross section, which can be calculated using Equa-
tion 8.4:

α 1
φ=
Ic ∫ ∫ TG z dw. dz = R (8.4)

where Ic is the inertia of the cross section transformed for steel beams (in.4), and R is
the radius of curvature (in.).
Any additional stresses because of curvature, created by thermal gradient, shall be
calculated as

σ E = E [α TG − α TUG − φz ] (8.5)

where E is the modulus of elasticity (ksi).


8.6.2.1.6
Creep
Concrete creep strains should be calculated using LRFD specifications Article 5.4.2.3.2.
Time dependence and changes in concrete strength should be taken into account in
determining the effect of concrete creep. The creep coefficient (ψ) can be determined
using Equations 8.6 through 8.10 (LRFD specifications Equation 5.4.2.3.2-1):

Ψ ( t , ti ) = 1.9ks khc kf ktd ti−0.118 (8.6)

in which

V
ks = 1.45 − 0.13 ≥ 1.0 (8.7)
S

khc = 1.56 − 0.008H (8.8)

5
kf = (8.9)
1 + fci′

 t 
ktd =  (8.10)
 61 − 4 fci′ + t 

where
H = relative humidity (%). In the absence of better information, H may be taken
from Figure 8.14;
ks = factor for the effect of the volume-to-surface ratio of the component;
kf = factor for the effect of concrete strength;
khc = humidity factor for creep;
ktd = time development factor;

367

Chapter 8. JOINTLESS BRIDGES


Figure 8.14.  Percentage annual average
ambient relative humidity.
Source: LRFD specifications Figure 5.4.2.3.3-1.

t = maturity of concrete (day), defined as age of concrete between time of


loading for creep calculations, or end of curing for shrinkage calculations,
and time being considered for analysis of creep or shrinkage effects;
ti = age of concrete at time of load application (day);
V/S = volume-to-surface ratio (in.); and
fci′ = specified compressive strength of concrete at time of prestressing for pre-
tensioned members and at time of initial loading for nonprestressed mem-
bers. If concrete age at time of initial loading is unknown at design time,
fci′ may be taken as 0.80 fc′ (ksi).
8.6.2.1.7
Shrinkage
Concrete shrinkage should be calculated in accordance with the provisions of LRFD
specifications Article 5.4.2.3.3, where appropriate. For concrete elements, shrinkage
strain (εsh) can be calculated using Equations 8.11 and 8.12 (LRFD specifications
Equation 5.4.2.3.3-1):

ε sh = ks khs kf ktd 0.48 × 10−3 (8.11)

in which

khs = ( 2.00 − 0.014H ) (8.12)

where khs is the humidity factor for shrinkage.


368

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


This article states that if the concrete is exposed to drying before 5 days of curing
have elapsed, the shrinkage as determined in Equation 8.11 should be increased by
20%.
8.6.2.1.8
Settlement
Settlement is not a deterrent to the use of jointless bridges if sufficiently accounted for
in the design of the affected components. AASHTO provides guidance on estimating
settlement for structures in LRFD specifications Article 10.7.2.3.
Bridges with simple spans and simple abutment bearings are able to accommodate
shifting and the associated rotation of the end spans with flexibility of the bearings.
With continuous jointless superstructures and integral abutments, vertical or longitu-
dinal movement of the foundation will introduce additional stresses in the superstruc-
ture, deck, or both. In addition, with semi-integral abutments, vertical movement of
the foundation will introduce additional stresses in the superstructure, deck, or both.
Figure 8.15 demonstrates this concept with an exaggerated illustration showing settle-
ment (Δ). In instances in which traditional bearings are used, the superstructure is free
to rotate to accommodate the movement. In contrast, when the superstructure is inte-
gral with the substructure, the superstructure is not permitted to rotate or shift, and
thus forces are introduced from the fixed-end displacement.
The best approach to address settlement, in general, is to increase pile length so
that settlement is not a design consideration. If increasing the pile length is not an
option and design for settlement must be considered, then one of two strategies can be
used to reduce or eliminate the effect of settlement: (1) evaluate the anticipated settle-
ment and account for the resulting forces in the design, or (2) determine the maximum
permissible displacement allowable by design and take measures to ensure that that
settlement limit is not exceeded.

Δ Figure 8.15.  Comparison of settlement


effects on the superstructure (not to scale).

Δ
V M
M V

369

Chapter 8. JOINTLESS BRIDGES


8.6.2.1.9
Wind
Wind load needs to be considered in accordance with LRFD specifications Article 3.8.
As with braking and centrifugal forces, longitudinal and transverse forces resulting
from wind loads should also take into consideration the considerable stiffness of the
integral abutments (see Section 8.6.2.1.2).
8.6.2.1.10
Other Loads
All other loads prescribed by the LRFD specifications, such as collision forces and
water and ice loads, need to be applied to jointless structures in the same manner as
other structures. As with all designs, it is the engineer’s responsibility to determine
and apply the necessary load conditions appropriate for the unique situation of each
jointless bridge.
8.6.2.2
Load Combinations and Limit States
This section summarizes available information in the LRFD specifications and de-
veloped by SHRP 2 Project R19A related to load combinations and limit states to be
considered for jointless bridges.
8.6.2.2.1
Load Combinations
The following loads should be considered for jointless bridges:

DC = dead load of structural components and nonstructural attachments,


DW = dead load of wearing surfaces and utilities,
EH = horizontal earth pressure load,
LL = vehicular live load,
WS = wind load on structure,
WL = wind on live load,
TU = uniform temperature,
CR = creep,
SH = shrinkage,
TG = temperature gradient, and
SE = settlement.

8.6.2.2.2
Load Factors and Combinations
Table 8.7 lists load combinations required in the design of jointless bridges, and Table 8.8
shows load factors for permanent loads.
The LRFD specifications state that the load factor for the temperature gradient
(γTG) should be considered on a project-specific basis or may be taken as follows:

• 0 at the strength limit states;


• 1.0 at the service limit states when live load is not considered; and
• 0.50 at the service limit state when live load is considered.
370

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 8.7.  Load Combinations and Load Factors
Load DC TU
Combination DW CR
Limit State EH LL WS WL SH TG SE
Strength I γp 1.75 — — 0.50/1.20 γTG γSE
Strength II γp 1.35 — — 0.50/1.20 γTG γSE
Strength III γp — 1.40 — 0.50/1.20 γTG γSE
Strength IV γp — — — 0.50/1.20 — —
Strength V γp 1.35 0.40 1.00 0.50/1.20 γTG γSE
Service I 1.00 1.00 0.30 1.00 1.00/1.20 γTG γSE
Service II 1.00 1.30 — — 1.00/1.20 — —
Service III 1.00 0.80 — — 1.00/1.20 γTG γSE
Service IV 1.00 — 0.70 — 1.00/1.20 — 1.00

Note: — = not applicable.


Source: LRFD specifications Table 3.4.1-1.

Table 8.8.  Load Factors for Permanent Loads (γp)


Load Factor
Type of Load and Foundation Type Maximum Minimum
DC: Component and attachments 1.25 0.90
DC: Strength IV only 1.50 0.90
DW: Wearing surface and utilities 1.50 0.65
EH: Horizontal earth pressure
Active 1.50 0.90
At rest 1.35 0.90

Source: LRFD specifications Table 3.4.1-2.

Because the effects of γTG are typically self-limiting and do not significantly affect
strength or ductility at strength limit states for the types of bridge girders typically used
in jointless bridges, γTG can commonly be taken as zero for the design of foundations
in integral and semi-integral abutments.
Similarly, the load factor for settlement (γSE) should be considered on project-­
specific information or may be taken as 1.0. Load combinations that include settle-
ment should also be applied without settlement.
8.6.2.3
Bridge Movement
Three methods are provided for calculating bridge maximum end displacements. The
first approach is applicable to straight bridges, the second approach addresses trans-
verse movement of skewed bridges, and the third approach is a general method to
calculate the movement of curved girder bridges.

371

Chapter 8. JOINTLESS BRIDGES


8.6.2.3.1
Displacement of Straight (Nonskew) Bridges
Bridges expand and contract because of temperature changes and time-dependent vol-
ume changes associated with concrete creep and shrinkage. In jointless bridges, it is
important to estimate the maximum expansion and contraction at each end of a bridge
to determine the longitudinal displacement expected for the abutment piles. It is also
important to predict the movement at each pier and the joint width needed between
the approach slab and the pavement. Another important movement is the maximum
total thermal movement at each end resulting from the total effective temperature
range. The starting point to determine the maximum passive pressure should conser-
vatively be at the maximum contraction (Oesterle et al. 2005). The maximum passive
pressure is related to the end movement, with reexpansion for the full effective tem-
perature range.
Calculation of the length change for a prestressed concrete bridge can be accom-
plished through use of typical design values for the coefficient of thermal expansion
combined with creep and shrinkage strains. However, the overall variability of these
factors adds uncertainty to the calculated end movements. Although a coefficient of
thermal expansion for concrete is typically assumed to be 5.5 × 10–6 to 6.0 × 10–6/°F,
it is known that this value can range from approximately 3.0 × 10–6 to 7.0 × 10–6/°F
(­Kosmatka and Panarese 1988). Also, the variability of creep, shrinkage, and modu-
lus of elasticity of concrete is known to be significant (Bazant and Panula 1980). In
addition, resistance to length change from abutments and piers, combined with the
variability of the restraint (primarily caused by the variability of the soil), leads to
unequal movement at each end of a bridge (even in theoretically symmetrical bridges)
and uncertainty as to the magnitude of the movement at each end. Finally, the effective
setting temperature of the bridge and the age of concrete girders at completion of the
superstructure are typically unknown, making the relative magnitude of expansion
and contraction and the starting point for temperature, creep, and shrinkage calcula-
tions uncertain.
To investigate the effects of the variability of these parameters and to provide guid-
ance in formulating recommendations for design calculations, Monte Carlo studies
were carried out to calculate bridge movements in order to generate a large number
of computer analyses using the statistical variation of material parameters affecting
the movement (Oesterle 2005; Oesterle and Volz 2005). Within each analysis, val-
ues for the coefficient of thermal expansion, temperature at construction, creep and
shrinkage parameters of concrete, modulus of elasticity of concrete, and soil stiffness
were selected based on statistical distributions of the values of these parameters. The
variations in calculated bridge end abutment movements were then used to deter-
mine a 98% confidence interval for the maximum calculated movements. These maxi-
mum values were used to determine magnification factors, referred to as Γ factors, for
modification of calculated values to account for uncertainty in the various parameters
affecting results.

372

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


The procedures presented in the following sections outline how to determine the
maximum end movements of jointless bridges, including use of these Γ factors. In these
calculations, it is assumed that the bridge has unknown construction timing and that
no specific data on material properties are available.
For prestressed concrete bridges, the following steps should be used to estimate the
longitudinal movement:

• Determine the average construction temperature using the procedure described in


Section 8.6.2.1.5a.
• Determine the maximum and minimum effective bridge temperatures based on the
recommendations of Procedure B in Section 8.6.2.1.5a.
• Assume the parameters for concrete presented in Table 8.9.

Table 8.9. Concrete Parameters


Value Type Coefficient of Expansion Modulus of Elasticity
Value (English) 6.0 × 10 /°F
–6
57,000 fc′ (psi)

Value (metric) 10.8 × 10–6/°C


4.700 fc′ (MPa)

Source: Oesterle et al. 2005.

• Determine the point of zero movement of the fixity point of the bridge based
on the stiffness of the piers and the abutments. Use Section 8.6.2.1.3 provisions
(Clough and Duncan 1991) to estimate the backfill passive pressure and the p-y
method to evaluate the nonlinear behavior of the soil surrounding the piles. For
symmetric bridges, the middle of the bridge will be the point of fixity.
• Use Equations 8.13 through 8.16 to calculate the strain values in the bridge:

ε th = α∆T (8.13)

ε sh,deck − ε sh,girder
ε sh = ε sh,girder +
( EA)girder (8.14)
1+
( EA)deck

 
 
 1 
ε cr = ε cr,girder  
1 +
( )girder
EA

(8.15)

 ( EA)deck 

∆λ = Γε total λ (8.16)

373

Chapter 8. JOINTLESS BRIDGES


where
Dl = maximum end movement;
eth = thermal strain;
esh = shrinkage strain;
ecr = creep strain;
a = coefficient of thermal expansion;
E = modulus of elasticity;
A = cross-section area;
l = length from the point of fixity to the end of the bridge. Note that for an
unsymmetrical bridge two values of l are involved; and
G = magnification factor to account for uncertainty listed in Table 8.10, where
etotal = eth – esh – ecr for expansion, and etotal = –eth – esh – ecr for contraction.

• For maximum expansion, which occurs shortly after construction, use the tem-
perature difference between the maximum effective bridge temperature and the
mean construction temperature for the bridge location based on the Federal Con-
struction Council’s Technical Report No. 65 (FCC 1979). For creep and shrinkage
calculations, assume the girders are 90 days old. Based on Monte Carlo simula-
tion, Γ should be 1.6 to account for uncertainties with 98% confidence that the
movement will be less than the calculated value.
• For maximum contraction, which occurs after several years of service, use the
temperature difference between the minimum effective bridge temperature and the
mean construction temperature. For creep and shrinkage, assume ultimate values
with the girder to be 10 days old at the time of casting the deck. Based on Monte
Carlo simulation, Γ should be 1.35 to account for uncertainties with 98% confi-
dence that the movement will be less than the calculated value.
• For maximum thermal reexpansion from a starting point of full contraction,
use the full effective bridge temperature range without any creep and shrinkage
movements. Based on Monte Carlo simulation, Γ should be 1.2 to account for
uncertainties.
• The Γ values in the first two columns of Table 8.10 for maximum expansion and
maximum contraction are relatively large and possibly overconservative because
they are affected by the relatively large uncertainty of the construction or setting
temperature. Further studies to include a more deterministic method to incorpo-
rate the construction temperature for a given bridge may reduce these magnifica-
tion factors for a more efficient design approach.

For reinforced concrete bridges, the same procedure as that used for prestressed
concrete bridges should be used to calculate bridge end movements. In the case of
shortening, movement caused by creep is not a factor. Magnification factors for differ-
ent cases are listed in Table 8.10.

374

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 8.10.  Summary of Recommended Magnification Factors
Maximum Maximum Maximum Thermal
Bridge Type Expansion Contraction Reexpansion
Prestressed concrete G = 1.6 G = 1.35 G = 1.2
bridges creep + shrinkage + creep + shrinkage + thermal
thermal thermal
Reinforced concrete G = 1.6 G = 1.4 G = 1.2
bridges shrinkage + thermal shrinkage + thermal thermal
Composite steel G = 1.7 G = 1.5 G = 1.2
bridges shrinkage + thermal shrinkage + thermal thermal

Source: Oesterle 2005.

For steel girder bridges, the same procedure as that used for prestressed concrete
bridges should be used to calculate bridge end movements, except that the extreme
effective bridge temperatures should be calculated using the recommendations of Sec-
tion 8.6.2.1.5.
Other steel material parameters are provided in Table 8.11.
The effective coefficient of thermal expansion for steel composite bridges can be
estimated as shown by Equation 8.17 (Emanuel and Hulsey 1977):

(α EA)girder + (α EA)deck
αe = (8.17)
( EA)girder + ( EA)deck
Magnification factors for different cases are listed in Table 8.10.
8.6.2.3.2
Displacement of Skewed Bridges
For information on displacement of skewed bridges, refer to Appendix B.
8.6.2.3.3
Displacement of Curved Bridges
A procedure has been developed (Doust 2011) to determine the magnitude and direc-
tion of bridge end displacement in the case of curved integral abutment bridges. The
related material is provided in Appendix F, which explains the assumptions and limita-
tions of the approach.

Table 8.11. Recommended Steel Parameters


Value Type Coefficient of Expansion Modulus of Elasticity
Value (English) 6.5 × 10 /°F
–6
2.9 × 107 (psi)
Value (metric) 11.7 × 10–6/°C 2.0 × 105 (MPa)

Source: Oesterle 2005.

375

Chapter 8. JOINTLESS BRIDGES


8.6.2.4
Design of Pile Foundation
The main steps in design of piles are as follows:

• Based on subsurface explorations, develop a soil profile for the site. Details of
strength profiles, compressibility characteristics, stress history, and geology of the
subsurface materials should be included. Further, identify favorable and unfavor-
able strata in the affected subsurface zones.
• Estimate the loads for the strength and serviceability limit states.
• Determine the water profiles for the site and the expected depth of scour during
100-year and 500-year flood events.
• Select technically feasible pile types and pile lengths based on constructability, and
consider the strength, serviceability and extreme event limit states. Eliminate the
unsatisfactory alternatives.
• Make a general comparison between the technically feasible piles, and then design
with the most cost-effective alternative based on the following steps:

–– Estimate the axial and lateral pile nominal resistance considering soil and struc-
tural capacity.
–– Determine the required number of piles and their spacing.
–– Estimate the resistance of the pile group on the basis of pile group interaction.
If the group resistance is not sufficient, modify the number of piles and/or the
pile spacing.
–– Check the possibility of punching of the pile into any weak stratum that may be
present beneath the bearing stratum.
–– Determine the tolerable deformations of the structure and estimate its verti-
cal and lateral deformations. If the deformations are greater than the tolerable
magnitudes, increase the length of the piles or number of the pile spacing.
–– If the pile group is subject to uplift, check its uplift lateral.
–– Determine the loads on top of pile under design lateral displacements to deter-
mine design forces for interaction with the pile cap.
–– Determine whether pile load tests are needed to verify the design and apply the
appropriate resistance factors.
These requirements are summarized in Table 8.12 and categorized as to whether
the requirement applies to the strength or serviceability limit state. In certain cases the
extreme-event limit state governs the design of piles.

376

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 8.12.  Summary of Strength, Serviceability, and Extreme-Event Limit
States That Must Be Considered in Design of Pile Foundations
Strength Limit Serviceability Extreme-Event
Design Consideration State Limit State Limit State
Structural capacity of single pile X — X
Bearing capacity of single pile X — X
Bearing capacity of pile groups X — X
Punching into lower weak X — X
stratum
Settlement of pile groups — X X
Tensile capacity of piles during X — X
uplift
Uplift capacity of single piles X — X
Structural capacity of piles under X — X
lateral loading
Lateral movement of pile groups — X X
when subjected to lateral loads

Note: X = applicable; — = not applicable.


Source: Adapted from Barker et al. 1991.

8.6.2.4.1
Pile Orientation: Straight and Curved Bridges
Abutment piles of straight bridges should be oriented so that the strong axis of the
piles is perpendicular to the longitudinal direction of the bridge (Doust 2011). This
orientation results in strong-axis bending of the piles in response to longitudinal move-
ment of straight nonskew bridges.
A procedure has been developed to determine the optimum abutment pile orienta-
tion in the case of curved girder integral bridges (Doust 2011). Appendix F provides
the suggested approach and current limitations.
8.6.2.4.2
Pile Design
Design of piles should consider strength; ductility; fatigue; stability; pile group interac-
tion; and minimum penetration length required to satisfy the requirements for uplift,
scour, downdrag, liquefaction, lateral loads, seismic forces, and other extreme-event
loadings.
Figures 8.16 and 8.17 provide the design aids for design of piles for integral abut-
ment systems. These design aids are based on research conducted within SHRP 2
­Project R19A (Azizinamini et al., submitted for publication; Sherafati 2011). A sum-
mary of the steps in developing these design aids, which include consider strength,
fatigue, and local and global stability, is provided in Appendix C. Figures 8.16 and
8.17 provide four charts that allow the determination of maximum lateral movement
capacity of a single pile versus the applied axial load to the pile. The design aids are for

377

Chapter 8. JOINTLESS BRIDGES


(a) (b)

Figure 8.16.  Maximum displacement of compact HP sections in soft clay (cu = 2.9 psi) for (a) HP 10x57 and
(b) HP 12x84.

(a) (b)

Figure 8.17.  Maximum displacement of compact HP sections in medium clay (cu = 5.8 psi) for (a) HP 10x57
and (b) HP 12x84.

two HP piles (HP 10x57 and HP 12x84) and four soil conditions. The yield strength
of the HP piles is assumed to be 50 ksi. The design aids provided in Figures 8.16 and
8.17 make design of piles for integral abutment system an easy process. Knowing the
applied axial load to the pile allows determination of maximum lateral movement that
the pile can accommodate.

378

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


8.6.2.4.2a
Geotechnical Axial Resistance
The axial nominal resistance of a pile is the sum of its tip and friction resistance minus
the weight of the pile, as shown by Equation 8.18:

Qnom = Qs + Qt − W (8.18)

where
Qnom = nominal bearing capacity of a pile,
Qs = pile shaft resistance (Asqs),
Qt = pile tip resistance (Atqt),
W = weight of the pile,
As = surface area of the pile shaft,
qs = unit skin resistance of the pile,
At = area of the pile tip, and
qt = unit tip resistance of the pile.

In most situations (except for large concrete piles in pile bent piers), the weight of
the pile is small compared with the other terms and is usually disregarded.
8.6.2.4.2b
Global Stability
Global stability is referred to as buckling of the pile between end supports as opposed
to local flange or web buckling. In general, global stability is not a governing design
provision unless a significant length of pile is above ground level and is unsupported
against lateral buckling (Sherafati et al. 2012).
8.6.2.4.2c
Lateral Deformation of Pile Groups
Provisions of LRFD specifications Article 10.7.2.4 should be used when the p-y method
of analysis is used to evaluate pile group horizontal movement.
8.6.2.4.2d
Minimum Penetration Length
LRFD specifications Article 10.7.1.5 specifies the provisions for the minimum penetra-
tion length necessary to satisfy the requirements for uplift, scour, settlement, down-
drag, liquefaction, lateral loads, seismic response, and other extreme-event loading
conditions. This guidance is also appropriate for the design of jointless bridges and
should be followed by the designer.
Tensile loading of a foundation (uplift) may be caused by swelling soils, frost
heave, buoyancy, lateral loads, and tensile loading during construction activities. Piles
subjected to uplift should be designed to withstand tensile stresses and pullout from
the subsurface materials.

379

Chapter 8. JOINTLESS BRIDGES


Tensile loading for a foundation design is well covered in the LRFD specifica-
tions, which provide guidance to design against uplift for both single piles and pile
groups. The design of single piles, drilled shafts, and micropiles in groups is addressed
in Articles 10.7.3, 10.8.3, and 10.9, respectively.
Scour around the foundation is an important issue that should be considered in
the design. In geotechnical analysis, it should be assumed that the subsurface materials
above the scour line do not exist to provide bearing or lateral support. Three scour
types should be considered in design (Barker et al. 1991):

• Aggradation and degradation are long-term effects. Aggradation is defined as the


deposit of stream bed material eroded from other portions of a stream. Degrada-
tion is the removal of stream bed material and thus lowering of the bed elevation.
• General scour and contraction scour are distinguished by removal of bed material
across the entire width of the stream as a result of increasing flow velocities.
• Local scour occurs when bed material is removed from a small portion of the
width of the stream. Bridge piers and abutments induce acceleration of the flow
because of obstruction of the flow and cause vortices that wash away the bed
material.

Scour is usually evaluated for a design flood with a return period of 100 years,
with a check flood not to exceed the 500-year event, or from an overtopping flood of
lesser recurrence (AASHTO 2012).
To increase the safety against pile failure caused by scour, a few longer piles should
be used rather than many short piles.
Settlement is not a deterrent to the use of jointless bridges if it is accounted for in
the design of the affected components (see Section 8.6.2.1.8). Minimum penetration
lengths with respect to settlement calculations for the foundation are not an additional
concern for jointless bridges.
8.6.2.4.3
Analysis Tools
This section provides a general discussion of different analysis approaches and avail-
able tools designers can use to analyze jointless systems.
8.6.2.4.3a
Simplified Analysis (p-y Method)
The ability to estimate the response of laterally loaded piles is of great importance in
the design of jointless bridges. This design consideration is similar to a beam-on-elastic
foundation model. If the piles are deep enough, modeling the soil with Winkler springs
is a useful method. In this method the soil is considered as a series of independent
­layers providing resistance (p) to the pile deflection (y). This resistance (p) may be a
highly nonlinear function of the deflection (y). The proper form of a p-y relation is
influenced by many factors, including

• Variation of soil properties with depth;


• Shape of the pile deflection;

380

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


• The state of stress and strain throughout the affected soil zone; and
• The rate sequence and history of load cycles.

8.6.2.4.3b
Finite Element Analysis
Finite element modeling can be used to analyze a jointless bridge. There can be several
levels of finite element analysis for such a structure ranging from a simplified analysis
to a refined analysis.
In a simplified finite element analysis, different elements including composite
­girders, abutment walls, piers, and piles are modeled using frame elements. The model­
ing can be 2D or 3D; however, a 3D analysis is preferred. The soil–structure interac-
tion should be modeled by means of springs. Each spring’s load-deflection curve can
be assumed to be linear for a simplified model. In the case of 2D models, the girder
distribution factors should be calculated using appropriate equations from the LRFD
specifications.
In contrast, a refined finite element analysis is a 3D modeling of a jointless bridge.
In this approach, shell elements can be used to model the bridge elements. The soil–
structure interaction can be modeled using nonlinear springs, which can model the
abutment–soil interaction of the gap created between the abutment wall and soil as a
result of contraction.
Based on the importance and complexity of the bridge, the level of detail included
in the finite element model can vary. Engineering judgment should be exersized in
developing the 3D finite element model.
8.6.2.5
Design of Other Foundation Types
It is recognized that other foundation types may be appropriate for jointless applica-
tions depending on the requirements for the individual bridge. Additional consider-
ations for other foundations are discussed here.
8.6.2.5.1
Drilled Shafts
Drilled shafts should be designed considering the same design requirements as piles.
Note, however, that traditional drilled shaft diameters of 30 in. and larger may prove
to be too stiff for longer bridge lengths. Semi-integral abutments may be designed
using drilled shafts with no additional consideration. Refer to LRFD specifications
Article 10.8 for more information on the design of drilled shafts.
8.6.2.5.2
Spread Footings
Use of spread footings directly over rock with integral abutments is not common prac-
tice and is not recommended.

381

Chapter 8. JOINTLESS BRIDGES


8.6.2.5.3
Micropiles
Micropiles may be a viable option for jointless bridges. Note that micropiles, similar to
regular piling, should only be used in a single row for integral abutments. In addition,
the micropile design must include consideration of the cyclic nature of the bending
load resulting from the integral abutment configuration. Multiple rows of micropiles
should only be used for semi-integral abutments.
8.6.2.6
Design of Pile Cap
Depending on the selected jointless system, the pile caps of jointless bridges may re-
quire special consideration. The pile cap of an integral abutment no longer serves
solely as a transfer for gravity loads. The pile cap must transfer longitudinal move-
ments and other forces introduced by making the abutment integral.
8.6.2.6.1
Integral Pile Cap Design
Pile head elevation design for integral abutments can take one of two forms. The first
option is to fix the pile head against rotation. Alternatively, as recently demonstrated
by SHRP 2 Project R19A (Azizinamini et al., submitted for publication; Sherafati and
Azizinamini 2014), the pile head can be fitted with an elastomer-based collar that al-
lows for limited end rotation and displacements to occur, which alleviates some of the
stresses induced by bending of the piles.
8.6.2.6.1a
Encased Piles (Fixed-Head Condition)
The pile cap for integral abutments must take into account and be able to develop the
moment resulting from the restraint of the embedded pile head. Figure 8.18 illustrates
how the shear restraint develops as a moment over the pile length due to fixity (assum-
ing no soil support) as the force couple develops. In turn, this force is resisted by the
pile cap, as shown in Figure 8.19. Note that the shear (V) and bending resultant (Cm)
will be additive. Wasserman and Walker (1996) indicate that the depth of the resulting
stress block is as shown by Equation 8.19:

 lpe 
ap = 0.85   (8.19)
 2

where ap is the depth of the stress block, and lpe is the pile embedment length within
the cap.
It can be seen intuitively that increasing the embedment length will directly
decrease the bending resultant stresses on the cap, both by increasing the moment arm
and the length of ap.
Taking Mp as the plastic moment of the pile, the force couple balance is repre-
sented by Equation 8.20:

Mp = Cm D′ (8.20)

382

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


or

( )
Mp = ap bfcb2 lpe − ap (8.21)

where b is either the pile section depth (weak axis bending) or flange width (strong
axis bending), respective to pile orientation, or pipe diameter; and fcb2 is the bending
resultant stress. Cm and D′ are given in Figure 8.19.

Figure 8.18.  Moment transfer from pile


to cap.

Figure 8.19. Moment
transfer from pile to cap:
internal force balance.
(Not to scale.)

383

Chapter 8. JOINTLESS BRIDGES


It follows then that the maximum stress on the concrete cap is the combined resul-
tant of the bending and shear stresses (fcb1), as shown in Equation 8.22:

fcb1 = fcb2 + V ap b (8.22)

8.6.2.6.1b
Pin Head Piles (Flexible Condition)
By providing rotational capacity at the pile head, the stiffness of the piling system is
reduced, and the moments developed in the pile as a result of lateral movement are
decreased, because the pile will deform in a single curvature rather than double cur-
vature shape. Because the major criterion limiting the application of jointless bridges
is the capacity of the piles to accommodate lateral movement, the proposed detail can
allow the application of jointless construction to longer bridge lengths (Azizinamini
submitted for publication).
The proposed pile cap detail consists of an elastomeric casing at the pile head. To
alle­viate the stress concentration at the top of the pile caused by rotation, steel plates
that slide by each other are key to the design detail. One of these plates is welded to the
end of the pile, and the other is embedded in the concrete with shear studs. Figure 8.20
shows the pin head detail for the case of a concrete-filled tube (CFT) pile. However, the
suggested detail can be adopted for H-piles, as well.
This system offers several advantages; it

• Can provide longer service life (by allowing integral construction for longer
bridges);
• Can effectively be used for jointless skewed or curved bridges;
• Allows construction of longer jointless bridges;
• Develops smaller forces in the abutment and superstructure, since the lateral stiff-
ness of the pile is reduced; and
• Reduces construction cost and time.

Design considerations include the material around the pile head, which is intended
to have a very low elastic modulus to provide rotational capacity. Since the material
for the detail experiences large strains, it must be able to accommodate these strains
when subjected to the applied cyclic rotations. Elastomeric material, regularly used as
bearings for girder bridges, is recommended for the detail (Sherafati and Azizinamini
2014). Sufficient thickness needs to be provided to ensure the efficiency of the detail.
Preliminary results indicate that the minimum thickness of elastomer should be 4 in.
8.6.2.6.2
Semi-Integral Pile Cap Design
The pile cap in semi-integral bridges is mainly subjected to axial load and possibly the
moment created by axial load eccentricity applied to the pile cap. Figure 8.21 shows a
prefabricated pile cap.

384

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 8.20.  Proposed pin head detail for a concrete-filled tube (CFT).

(a) (b)
Figure 8.21.  Prefabricated pile cap: (a) lowering cap into place; (b) final position.

385

Chapter 8. JOINTLESS BRIDGES


8.6.2.6.3
Seamless Details
The design of the pile cap for a seamless bridge should follow the same procedure as
for integral abutment bridges.
8.6.2.7
Design of End Diaphragm (Backwall)
In addition to being integral with the superstructure, the end diaphragm acts as a
backwall for integral and semi-integral jointless bridge systems. The end diaphragm is
designed to resist forces resulting from soil loads; in the remainder of this discussion it
is referred to as the backwall. The soil loads include the passive pressure force that is
created by superstructure thermal expansion. The calculation of this passive pressure
(Pp) is shown in Sections 8.6.2.1.3 and 8.6.2.1.4. Modeling, as shown in Figure 8.22,
can be used to design the backwall.
The following subsections discuss additional backwall design considerations that
vary for each jointless bridge type.
8.6.2.7.1
Integral
The backwall for an integral bridge abutment must be designed to adequately transfer
forces across the construction joint and into the foundation cap for each direction in
which the pile bends. This transfer of forces is illustrated through an example of a
strut-and-tie model in Figure 8.23. In this figure, Section AA shows the local section
recommended for a local region (dp + b) over which the forces can be transferred and
a suggested reinforcing pattern. The length dp is the distance from the forward face of
the pile cap to the face of the pile.

Figure 8.22. Lateral pressure restraint


by superstructure.
Source: Oesterle et al. 2005.

S Pp

Beam (typ.)

386

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 8.23.  Lateral pressure restraint by superstructure.
Source: Oesterle et al. 2005.

8.6.2.7.2
Semi-Integral
Semi-integral backwalls do not require additional considerations above those outlined
in Section 8.6.2.7.1. The one item of note, however, is that if removable forms are not
used to form the bottom of the backwall over the foundation cap, the joint-fill material
used should be sufficiently stiff to support the concrete weight, yet flexible enough not
to interfere with the movement permitted by the bearings. This has been successfully
accomplished with expanded polystyrene filler.
8.6.2.7.3
Seamless
The design of backwalls for seamless bridges should be the same as for integral bridges.
8.6.2.8
Design of Approach Slab
Jointless bridges require approach slabs for two main reasons: (1) the slab needs to be
positively attached to the deck or substructure, or both, to eliminate the joint over the
abutment; and (2) the slab must span the area behind the abutment where the potential

387

Chapter 8. JOINTLESS BRIDGES


for backfill settlement exists. Backfill settlement will occur and introduce voids regard-
less of the degree of compaction and must be considered in design (Schaefer and Koch
1992).
8.6.2.8.1
Integral and Semi-Integral
For both integral and semi-integral abutments, the length of the approach slab is de-
termined by the extent of the backfill. Gangarao and Thippeswamy (1996) determined
that the rate of backfill settlement decreased significantly beyond 20 ft from the back
face of the backwall. This is a typical standard approach slab dimension shown in
several state standards. The study by Schaefer and Koch (1992) demonstrated that
backfill movements occur within a 1.5 horizontal–to–1.0 vertical line from the bot-
tom of the abutment for integral abutments. A general recommendation for the design
length of the approach slab is to conservatively set at a 2.0 horizontal–to–1.0 vertical
slope from the bottom of the abutment. A 20-ft minimum should be considered for
both integral and semi-integral abutments, as shown in Figure 8.24.
In addition, experience from several states has found that the approach slab should
be positively attached to the backwall by at least No. 8 reinforcing bars anchored with
a hook, as shown in Figure 8.24. The condition shown in the figure allows for a sepa-
rate pour of the approach slab designed as a simple span. Creating a moment connec-
tion between the approach slab and the deck slab is not recommended. The connection
should be detailed to act as a pin with tension steel transferred across the approach
span into the backwall for integral and semi-integral abutments. If a moment connec-
tion is desired, it is recommended to use a seamless deck transition for the design (see
Section 8.6.2.8.2).
A final consideration for the approach slab is the development of compression
forces. Sufficient allowance for expansion of the superstructure must be accommo-
dated in the sleeper slab. (See Section 8.7.3 for sleeper slab details.) Otherwise, com-
pression can be introduced into the slab resulting from closing the expansion gap and
then activating the passive pressure behind the sleeper slab or contact with the adja-
cent roadway pavement, which is a major issue for spalling and buckling of adjacent
pavement.
8.6.2.8.2
Seamless Deck Transition Zone
Details of seamless systems developed by SHRP 2 Project R19A are provided in Ap-
pendix E. The system is shown in Figure 8.25 and Figure 8.26.
Figure 8.25 shows that beyond the abutment a “Transition Zone” is required that
replaces the approach slab. The proposed transition introduces simplicity and ease of
construction (Jung et al. 2007). The concept slowly transitions from a heavily rein-
forced region to a plain jointed condition over an extended transition length. Within
the heavily reinforced region, crack spacing is quite small. As the level of reinforce-
ment is reduced, the crack spacing increases. These cracks may be allowed to occur
naturally or may be forced by shallow saw cuts in the pavement.

388

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 8.24.  Determination of approach slab length.

Figure 8.25.  Seamless paving over bridge transitioning to jointed pavement.

Immediately adjacent to the bridge is a thickened and reinforced approach zone.


The approach zone behaves similarly to a reinforced concrete slab bridge and is
intended to carry flexural forces that may arise as a result of settlement.
The design of the approach zone is similar to the design of an approach slab for
an integral or semi-integral bridge. The transition zone is not designed, per se, but the
reinforcing spacing is reduced in specified stages. This transition zone reinforcement is
shown in Figure 8.26.

389

Chapter 8. JOINTLESS BRIDGES


12' 12' 15'

Reinforcing Steel Optional Dowel Dowel

Profile View

Saw Cuts or Induced Design Crack CRC Pavement


JC Pavement

CRCP
Longitudinal
Steel

100% Steel Zone 60% Steel Zone 30% Steel Zone


Transition Transition

Plan View

Figure 8.26.  Continuously reinforced transition zone to jointed pavement.


Source: Jung et al. 2007.

8.6.2.9
Design of Superstructure–Pier Connection
By definition, bridge decks in jointless bridges are continuous, including the region
over the piers. The connection between the piers and the bridge deck could be integral,
pinned, or expansion types, or they could be connected with a link slab. Figure 8.27
shows these different configurations conceptually.
In integral-type connections (Figure 8.27a), the pier and superstructure are mono-
lithic with frame action developed between the superstructure and substructure. The
advantage of this type of connection is the elimination of bearings. Further, the system
provides higher levels of redundancy, especially in highly seismic areas. The longitudi-
nal movement of the bridge superstructure is not affected by making the piers integral
with superstructure. However, the longitudinal expansion of the deck must be consid-
ered in the design of the pier columns, pier foundations, and their connection to the
superstructure.
In pinned connections (Figure 8.27b), bearings are used to restrict longitudinal
movement. Rotation at the bearing is allowed. Although designated as a pin-type
connection, typical bridge terminology in which a bearing is not permitted to move
390

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 8.27.  (a) Integral, (b) pinned,
and (c) and (d) expansion type bearings
for jointless bridges.

longitudinally is designated as a fixed bearing, commonly denoted as “F” (Fixed) in


traditional design plans. For this connection, longitudinal movement between super-
structure and pier is not permitted. Similar to integral-type connections, the longitudi-
nal expansion of the deck must be considered in the design of the pier.
In expansion connections (Figure 8.27c), bearings are necessary and are required
to accommodate both rotation and longitudinal movements. This detail uses tradi-
tional expansion bearings as determined by design requirements.
For the first three connection types shown in Figure 8.27 (integral, pinned, and
expansion), the superstructure is made continuous over the pier. This structural con-
tinuity can be accomplished in one of two ways: (1) the superstructure splices can be
positioned such that they are made at or near the dead load inflection points for the
continuous bridge, or (2) a continuity splice can be used over the pier. This second
option is commonly referred to as simple for dead load, continuous for live load.
This construction method is shown conceptually in Figure 8.28 for an integral pier.
The beams are placed as simply supported over the pier; the beams are either spliced
mechanically or additional reinforcing is provided for the diaphragm; and finally, a
closure pour is made.
The last construction option is the expansion condition using a link slab (Fig-
ure 8.27d). A linkage slab is used when the beams are not positively connected, as
is the case with the other details shown in Figure 8.27. For this condition, the super­
structure is designed and constructed with traditional bearing considerations.
The design considerations for each of these pier cap connections are provided in
the following sections.
8.6.2.9.1
Integral Pier Cap
When making the superstructure truly integral with the pier cap, it must be recognized
that both positive and negative moments will be introduced to the cap from live loads
and other transient loads. Sufficient strength needs to be provided through the deck,

391

Chapter 8. JOINTLESS BRIDGES


(a) (b) (c)

Figure 8.28.  (a) and (b) Simple for dead and continuous for live pier detail after the placing of
girders and (c) after the closure pour.
Source: Azizinamini et al. 2008.

integral diaphragm, and pier cap. Although this type of connection eliminates the need
for bearings at the piers and can increase clearance, it introduces more complex forces
in the superstructure and the piers (see Figure 8.29).
The longitudinal deflection movement of the foundation and the pier accommo-
dates the total longitudinal movement expected at the top of the integral piers. (See
Section 8.6.2.10 for additional information.) When designing the integral cap, the
connection between the beams, continuity diaphragm, pier cap, and pier column must
be sufficient to transfer the moments resulting from this deflection. Resolution of these
forces should be computed by an analytical method or structural model with the abil-
ity to properly capture the behavior of the whole bridge system.

Figure 8.29.  Integral cap as completed.

392

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


8.6.2.9.2
Fixed (Pinned) and Expansion Pier Caps
Similar to the integral pier cap, for both fixed (pinned) and expansion pier caps the
superstructure is made continuous over the pier. The difference is that it is not made
integral with the pier caps. The connection between the superstructure and the pier is
treated with traditional bearings.
Various details are used over the interior supports of multispan bridges to elimi-
nate joints. One of the concepts implemented by various owner agencies for concrete
bridges has been the use of simple spans for dead load made continuous for live load
(Freyermuth 1969; Oesterle et al. 1989). The girders are simply supported for dead
load, but continuity is achieved with deck steel as negative moment reinforcement over
the piers. In addition, the girders are made integral with the interior pier diaphragms,
and commonly positive moment reinforcement is included, as shown in Figure 8.30.
Badie et al. (2001) discussed the alternative use of an interior steel pier diaphragm
with prestressed girders to speed construction and achieve better overall design econ-
omy. The concept of a simple span made continuous has also been applied to eliminate
interior joints and improve the construction speed and design economy for short- and
medium-span steel girder bridges (Azizinamini et al. 2008).
As discussed in Section 8.6.2.4, attention must be paid to the effects of provid-
ing positive moment restraint at the diaphragms. Some simple-made continuous pre-
stressed concrete girder bridges have experienced severe cracking in the girders near
the interior diaphragms. One extensively studied example is the Francis Case Memo-
rial Bridge spanning the Washington Channel of the Potomac River in the District
of Columbia (Telang and Mehrabi 2003). The prime cause of this distress was the
restraint of upward camber of the prestressed girders under the influence of prestress-
ing and temperature gradient. According to Telang and Mehrabi (2003), “By provid-
ing a large amount of positive moment reinforcement at the diaphragms, designers
inadvertently make the diaphragm area stronger than the adjacent girder sections,
thereby forcing the cracking to occur in far more critical but weaker areas of the girder
span.” Their article continues, “In closing, it is important to note that this seemingly
simple transformation of simple-span prestressed girders to continuous spans should
be attempted with caution, and significant attention must be paid during analysis and
design to include loading conditions that can cause counterintuitive behavior such as

Deck Reinforcement

Positive Moment
Reinforcement

Figure 8.30.  Precast, prestressed girders connected with live load continuity.

393

Chapter 8. JOINTLESS BRIDGES


secondary positive moments at the piers. Most importantly, positive moment rein-
forcement should be designed and detailed such that any cracking, if it occurs, should
be limited to the relatively less critical diaphragm region of this type of structural sys-
tem.” Further discussion of this problem and solutions to avoid it have been published
by Oesterle et al. (2004a) and Arockiasamy and Sivakumar (2005) and are discussed
in Section 8.6.2.4.
8.6.2.9.3
Link Slab Expansion Pier Cap
A link slab is a type of detail used in conjunction with existing or new bridges hav-
ing girders that act as simple beams for both dead and live loads. In this type of deck
detail, the slab spans continuously over the longitudinal gap between the adjacent
span girders, and the girders are kept as simple spans (see Figure 8.31). The length of
the deck connecting the two adjacent simple-span girders is called a link slab (Caner
and Zia 1998). Link slabs generally require less deck reinforcement, but they have
more girder positive moment demands than simple-made-continuous designs. Limited
analysis and laboratory experiments were carried out and design recommendations are
provided in Caner and Zia (1998). The use of this detail has been very limited due to
field-observed cracking. In fact, link slabs are not common in Snow Belt states. A crack
is invariably formed due to deck slab rotation as the bridge is loaded with live loads.
SHRP 2 Project R19A research studies on the link slab indicated that it offered
negligible rotational end restraint to the bridge girders and that the link slab can be
analyzed as a beam subjected to the same end rotations as the adjacent girders. The
researchers found that under service-load conditions, the link slab would crack pri-
marily due to bending. Prior methods developed by Gastal (1986) and El-Safty (1994)
were capable of predicting the forces, stresses, and crack widths in the link slab due to
thermal and shrinkage effects and creep. Caner (1996) modified the procedures devel-
oped by Gastal and El-Safty to properly capture the link slab actions. All these solu-
tions were based on beam theory. The reinforcing bar stresses compared reasonably
well with the data measured from the experimental tests. The predicted crack widths

Figure 8.31.  Conceptual detail for link slab.

394

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


were somewhat larger than the measured crack widths. The researchers concluded that
bending and cracking under live load plus impact are the governing factors that must
be considered in the design of the link slab.
Caner and Zia (1998) suggested design of the link slab using only one layer of
rebar placed near the top of the deck, but they suggested that two layers could be used
to improve performance in bridges having horizontal restraints.
8.6.2.10
Design of Integral Piers
As discussed in the section on design of integral pier caps, the total longitudinal move-
ment expected at the top of integral piers is accommodated by two modes of defor-
mation: longitudinal movement via rotation of the foundation system and flexural
­deflection of the pier. Pier deflection can be both elastic and inelastic in response.
8.6.2.10.1
Foundation Rotation
For spread footings, Zederbaum (1969) provides an equation, shown here as Equa-
tion 8.23, to estimate the rotational stiffness of the soil or rock responding to an
­applied moment:
3Es I f
Kθ = (8.23)
b
where
Kq = rotational stiffness of the foundation,
b = one-third of the spread footing width,
Es = compression modulus of the soil or rock, and
If = the moment of inertia of the footing base.

For pile-supported and drilled shaft–supported foundations, the rotational stiff-


ness is estimated from the elastic stiffness of the pile or shaft group. Rotation of the
foundation can be attributed to the elastic shortening and elongation of the piles or
shafts for multiple rows. Note that the elongation and shortening add additional uplift
and downward forces, respectively, that must be accounted for in the foundation
design. In a single row of piles or drilled shafts, the rotational stiffness is based on the
cantilever response of the single row. The length of the cantilever is based on the soil–
structure interaction at the foundation and can be based on the assumed or calculated
point of fixity for the pile or shaft.
8.6.2.10.2
Pier Displacement
The differential between the pier displacement at the integral cap and the rotation of
the foundation is the deflection of the pier column. The resulting design moment can
be estimated by the following steps. First, the expected movement of the superstruc-
ture at the pier cap should be calculated, as outlined in Section 8.6.2.3. Alternatively,
this can be sufficiently approximated by determining the point of zero movement on
the bridge and multiplying the end displacement by the ratio of the distance from the

395

Chapter 8. JOINTLESS BRIDGES


fixed point to the pier to the distance from fixity to the end support. Second, assume
that 30% of the expected lateral deflection is accommodated by the foundation rota-
tion. Thirty percent is based on a parametric study that demonstrated foundation
­rotation can vary from 30% to 80% with an average close to 45%. Third, the antici-
pated bending moment should be calculated using Equation 8.24:

6EI e ∆b
M= (8.24)
H2
where
E = concrete modulus,
Ie = effective section modulus,
Db = lateral deflection at the pier cap, and
H = height of the pier.

Note that the value of the effective section modulus may depend on the applied
loading and thus a simultaneous or iterative solution may be required. For fixed
(pinned) continuous piers, divide the result of Equation 8.24 by 2 (fixed-end moment
for a fixed-guided beam is one-half that of a fixed-fixed beam). The value of the effec-
tive section modulus for reinforced concrete piers can be obtained from Equation 8.25:

 Mcr 
3   M 3 
Ie =   I g + 1 −  cr   Icr (8.25)
 Ma    Ma  
where
Mcr = cracking moment,
Ma = applied moment,
Ig = moment of inertia of gross section, and
Icr = moment of inertia of cracked section.

8.6.2.11
Design of Wingwalls
The design of wingwalls depends on their orientation relative to the abutment stem,
their method of support, and the abutment skew. There are various possible configura-
tions for wingwalls, but the traditional configurations include U-shaped, straight, or
flared, the last being some degree of angle between the first two.
Oesterle et al. (2005) indicate that the U-shape configuration is preferable for
wingwalls in that this configuration inherently reduces the passive pressure introduced
by the longitudinal movement of the abutment end diaphragms. Additionally, they
note that the U-shape configuration conveniently contains the soil behind the abut-
ment and decreases bulging of the embankment soil.
Use of both straight and flared walls leads to the development of passive pressure on
the wingwalls as the jointless abutment moves. Oesterle et al. (2005) note that this pres-
sure can be expected to decrease as the distance from the abutment increases, but that
the degradation cannot be effectively predicted. Thus, the wingwalls need to be designed
for the same passive pressure as that of the abutment end diaphragm across the length.

396

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


For integral and seamless bridges, additional considerations for wingwalls include
the loading effect they have on the bridge structure. When cantilevered from the abut-
ment stem, the weight of the wingwalls will create additional torsion and/or bending
along the length of the abutment. These forces are resisted by a counteracting negative
moment at the end of the external beam or girder.
If wingwalls for integral abutments are placed on supports, such as piles or a spread
foundation, the support must also be able to accommodate the movements of the joint-
less bridge. For this condition, Oesterle et al. (2005) note that the shear and moment
developed in the wingwall foundation must be transferred through the wingwall struc-
ture to the abutment and superstructure. They also note that U-shaped wingwalls on
piles create significant resistance to abutment rotation, which creates partial fixity for
beam end moments on the exterior beams or girders. These additional moments need
to be included in the design of the connections of the exterior beams to the integral
abutment.
8.7
Details
The introduction of different mechanisms for transferring force to the foundations
requires that additional details be considered when designing jointless bridges. The
following sections present specific details for each jointless bridge type. The term back-
wall is used to describe the end diaphragm that resists soil loads.
8.7.1
General Abutment Details for Jointless Bridges
In this section, details that have been used successfully in the past by some states are
presented along with general concepts. The figures represent recent research efforts and
the accumulated experience of several states that have used jointless bridge technology.
8.7.1.1
Integral Abutments
Figure 8.32 shows the overall concept for an integral bridge abutment, including the
typical layout with the beam, end diaphragm (backwall), and pile cap, all integral.
Although it is not necessary in all cases, the beam shown in this figure is sitting on a
temporary pedestal to achieve proper alignment before being cast integral with the
rest of the abutment. Alternatively, the cap can be stepped to accommodate elevations
before pouring the backwall. For proper alignment and to allow for rotations that
occur when placing the beam, a small elastomeric pad should be placed at the girder
bearing even though each beam will eventually be cast composite with the abutment.
Note that the need to design the pads and for what capacity has not been thoroughly
studied. The pads need not be designed to meet the criteria for rotational capacity,
which is now addressed as a shear strain component. The maximum rotation of the
pad is realized during placement of the beam. A reasonable assumption is to design the
pad to accommodate only noncomposite bearing pressure.
Drainage is also important to avoid ice expansion and removal of the backfill by
washout. A drain pipe should be placed at the appropriate location to properly remove
any water that might otherwise accumulate behind the backwall.

397

Chapter 8. JOINTLESS BRIDGES


Figure 8.32.  General integral abutment concept.

Additional end diaphragm details are presented in Figure 8.33. Note that an H-pile
foundation is shown in the figure; however, each of the foundation types noted in the
strategy table in Section 8.5 can be interchanged. The minimum embedment length
of 2 ft, within the pile cap, should be maintained for H-piles, prestressed piles, and
CFT piles, as shown in Figure 8.34. Also shown in the figure is an approximate cap
height of 5 ft, typical of cold-weather regions, which allows for embedment below
the frost depth and to provide 2 ft between the finished grade and the bottom of the
beam. A depth of 3 to 3.5 ft is more common where frost depth need not be consid-
ered. Another alternative to the holes through the beam shown in Figure 8.34 is to
use threaded inserts, which are preferred by some precast concrete companies to ease
securing them in the forms.
398

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 8.33.  Integral abutment details.

Figure 8.34.  Integral abutment rotation


detail.
Source: Ohio DOT.

399

Chapter 8. JOINTLESS BRIDGES


Figure 8.34 is an adaptation from an Ohio DOT standard drawing showing a pre-
stressed concrete beam. A standard detail for most prestressed girders includes provid-
ing holes through the beam for reinforcing. This reinforcing provides continuity though
the backwall for bending and limits the differential deflection between the superstruc-
ture and backwall where tension forces develop in the top portion of the web.
In contrast to the design recommendations in Section 8.6.2.7, the state of Ohio
allows for rotation in the backwall across the construction joint instead of designing
rebar to transfer the forces through the stem. The configuration is used to accommo-
date the rotation of the superstructure, as shown in Figure 8.34. At the centerline of
bearing, reinforcing is crossed at the bearing pivot location and expansion-joint mate-
rial is placed to permit a limited amount of rotation. Note that Ohio limits the length
of its bridges with integral abutments to 250 ft, so consideration of this limit should
be made before adopting this detail for other bridges.
Figure 8.35 presents another standard integral detail drawing from the New York
State Thruway Authority that shows a steel beam connection. This detail is more
typical of DOT design standards in that the reinforcing is continuous across the con-
struction joint. When comparing this detail with Figure 8.33, although both details
have had repeated success, there are two obvious differences: (1) Figure 8.33 shows

Figure 8.35.  Integral abutment details.


Source: New York DOT.

400

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


a bent hook bar connecting the approach slab, but Figure 8.35 shows that continuity
is maintained by a straight bar connecting the approach slab to the deck; and (2) the
Figure 8.33 detail uses a shear key, but the Figure 8.35 detail relies solely on the con-
tinuity of the reinforcing across the construction joint. Each detail has demonstrated
success in application, and the designer should consider which option may be more
appropriate for each bridge’s unique situation.
For more information on backwall detailing, see Section 8.7.2.
8.7.1.2
Semi-Integral Abutments
Figure 8.36 shows the overall concept for a semi-integral bridge abutment. It includes
the typical layout with the beam and end diaphragm (backwall) cast integral.
Drainage and porous backfill are necessary for the same reasons as for integral
abutments: formation of ice and integrity of the backfill. In semi-integral abutments,
two bearing strategies have been used successfully. In the first method, the pile cap

(a)

Expanded
1” PFJ Polystyrene
Elastomeric Bearing

(b)
Figure 8.36.  General semi-integral abutment concept: (a) section (with pedestal) and
(b) elevation (with no pedestal).

401

Chapter 8. JOINTLESS BRIDGES


may be cast level and the superstructure superelevation can be accommodated through
the use of bearing pedestals. The second method is to step the pile cap. In the latter
case, the polystyrene filler must be used on both the top of the cap and on the sides of
the step to allow for movement. Due to the nature of the superstructure movement, it
is recommended that the first case with pedestals be used for locations of high skew
(larger than 20°) and bridges on a curve. If it is desired to inspect the bearings during
the life of the bridge, removable filler material should be placed in front of the bearings.
Figure 8.37 shows the successful detailing strategies that have been used in vari-
ous states. The foundation shown is for a drilled shaft, but other foundation types are
equally applicable. Similar to integral abutments, dowel holes are placed through the
beam or girder. Unlike integral abutments, bearings are used to accommodate move-
ment between the superstructure and the foundation. Efforts must be made to seal
the gap between the cap and backwall yet still accommodate movement. This seal
has traditionally been a preformed filler surrounding the bearing area and a layer of
waterproofing applied to the rear face of the seam before placing the backfill. For more
information on backwall detailing see Section 8.7.2.
Other than the backwall and treatment of the bearing area, detailing for the rest
of a semi-integral abutment is the same as for a traditional design.
Figure 8.38 shows an alternative detail used in instances in which the diaphragm
is extended and a lip is dropped down over the pile cap. This detail replaces the neo-
prene sheeting that provided the barrier between the porous backfill and the expanded

Figure 8.37.  Semi-integral details.

402

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 8.38.  Semi-integral details with
extended diaphragm.

polystyrene filler surrounding the bearings. Preformed elastomeric material is placed


between the extended diaphragm and the abutment stem.
8.7.1.3
Seamless Abutments
Detail recommendations for the transition zone are not well established, and no stan-
dard details are available for reference. However, the recommendations for the abut-
ment cap and backwall are the same as those presented in Section 8.7.1.1.
8.7.2
Pile Cap and Backwall
Oesterle et al. (2005) recommend that vertical reinforcement for the moment from
the soil load be distributed with 75% of the bars within 25% of the beam spacing on
either side of the beam. Furthermore, for crack control they recommend the center-
to-center spacing (s) of the flexural reinforcement not to exceed (in inches) the value
calculated by Equation 8.26:
540
s≤ − 2.5cc or
fs
 36  (8.26)
s ≤ 12  
 fs 

where cc is clear cover from the nearest surface in tension, and fs is calculated stress
(ksi) at service load, or alternately, as 0.60Fy.
This limitation is taken from ACI 318-05 Section 10.6.4 rules for the distribution
of flexural reinforcing to control cracking in one-way slabs. Further commentary on
this requirement can be found in that section.

403

Chapter 8. JOINTLESS BRIDGES


8.7.3
Sleeper Slab
A sleeper slab is appropriate for all integral or semi-integral bridges and is placed at
the roadway end of the approach slab. The intent of this slab is to provide a relatively
solid foundation for the far end of the approach slab and to provide a location for
limited expansion and contraction (see Figure 8.39). Although no formal design is sug-
gested, a typical suggested detail has been provided by Wasserman and Walker (1996).
A potential problem with the design shown in Figure 8.39 is that it presents a
potential for cracking in the approach pavement where it suddenly transitions to the
thin piece above the sleeper slab. Although this type of transition might ease final grad-
ing, it is preferable to have the stem of the inverted “T” of the sleeper slab extend to
final grade and thus avoid any sharp transitions.
The state of New York has adopted this sleeper slab detail and modified it to
marry the adjoining pavement design based on the type of surfacing used. Figure 8.40
and Figure 8.41 show the sleeper slab for concrete and asphaltic pavements, respec-
tively. In these figures, note how the state formed the joint such that both the pave-
ment and approach slab are graded at full depth up to the sleeper slab that provides
the transition.
The location of the sleeper slab should be placed so that the entirety of the slab is
outside the failure plane, as discussed in Section 8.6.2.8.1.

Figure 8.39.  Suggested helper (sleeper) slab details.


Source: Wasserman and Walker 1996.

404

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 8.40.  Sleeper slab with concrete pavement approach.
Source: New York DOT.

Figure 8.41.  Sleeper slab with asphalt pavement approach.


Source: New York DOT.

405

Chapter 8. JOINTLESS BRIDGES


8.7.4
Details for Skewed and Curved Bridges
Transverse movements of integral abutments associated with large skews or horizontal
curves should be accommodated by the details for barrier walls, drainage structures,
and the ends of the approach slabs. In addition, the foundation and pier structure stiff-
ness will likely be significant for movement parallel to the pier cap. It is recommended
that the connection between the bottoms of the girders and diaphragms and the pier
caps be flexible in this direction. This approach, however, may not be appropriate for
seismic design, in which case the design of the diaphragms should consider the interior
pier restraint of the rigid body rotations that result from passive abutment restraint of
longitudinal thermal expansion.
8.8
Construction
8.8.1
Construction Stability
In response to concerns about the repetitive bending stresses on the pile, it is recom-
mended that no seam (weld) be placed at the top 30 ft of the pile. This placement of the
seam will ensure proper ductility and eliminate the possibility of having a poor fatigue
detail near the region of higher bending response. In addition, it will better ensure
proper alignment of the pile at the cap.
The order of construction is also important, as described in Section 8.8.4.
8.8.2
Utilities
Nonflexible utilities should not be permitted to pass through integral and semi-integral
abutments. Multiple DOTs report experiencing problems in which the flexibility of the
integral cap creates issues with rigid utilities. Only utilities that are able to sufficiently
flex with the movement of the integral abutment should be permitted, but it is prefer-
able to locate all utilities adjacent to the bridge structure.
8.8.3
Cracking Control
Vertical cracks have often been found at the bottom of diaphragms between precast
beams over the piers in the positive moment connection region near the external (­ fascia)
girders. On the interior girders encased in the diaphragm, spalling of the diaphragm
has been observed near the bottom flange. This spalling results from the bottom flange
slipping outward (away from the diaphragm) due to the end rotation of the girder
asso­ciated with creep and thermal changes. These vertical cracks in the diaphragm and
end rotation of the girders serve to relieve tensile stresses resulting from creep, shrink-
age, and thermal movement and are not detrimental to the integrity of the structure.
Attempts to control this cracking through overreinforcing may result in cracks in less
desirable locations.

406

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Horizontal cracks and efflorescence have been found on the forward face of inte-
gral abutments at the construction joint on top of the pile cap. These alterations can
be alleviated by placing adequate sealing from water behind the stem across the con-
struction joint.
Settlement of the approach slab is common and can cause cracking and further
damage to the barrier rail. Rails that are attached to both the deck and approach slab
should be jointed to accommodate the differential settlement.
8.8.4
Construction Sequencing
Guidelines for concrete bridge-deck materials and construction to control transverse
cracking in concrete bridge decks are presented in NCHRP Report 380 (Krauss 1996).
Among the issues that affect deck cracking are weather, time of placement, curing,
vibration, finishing, loads, and placement sequencing. Certain current practices are
presented here for jointless bridges.
For jointless bridges, the construction sequence should generally be as follows:

1. Embankment should be completed before pile driving and should allow for con-
solidation (if required).
2. Piling should be placed, and predrilled holes filled and forms constructed (if used).
3. Abutments and wingwalls should be constructed to the elevation of the bearing
seat.
4. Semi-integral elastomeric bearings should be set; or integral beam pads should be
set allowing for rotation from beam and deck dead load.
5. Beams should be set.
6. The deck slab and the integral backwall should be cast. The ends of the slab should
be poured last to minimize locked-in stresses at the supports.
7. Drainage and backfill should be placed behind the abutments after the deck has
achieved the appropriate strength. It is important that the backfill be placed simul-
taneously behind each abutment so the bridge is not inadvertently shifted in the
unsupported direction.
8. The approach slab should be cast, ideally with the bridge in the thermally con-
tracted position (i.e., early morning). This avoids putting the slab into tension until
the concrete has gained sufficient strength.

It should be emphasized that simultaneous placement of the abutment backfill (Step 7)


is particularly important for semi-integral abutments because in semi-integral bridges
the superstructure sits on flexible bearings rather than being positively attached to the
abutment, and it is more likely to move in response to the pressure from the compact-
ing procedures.

407

Chapter 8. JOINTLESS BRIDGES


8.8.5
Fill Compaction
Construction can follow normal compaction procedures as specified by the owner
agency except as noted in the section on construction sequencing. Fill compaction has
been modified and adjusted using several variables, including the use of specialized
material. However, general experience has indicated that properly compacted normal
fill material is sufficient for jointless bridge construction, and proper drainage behind
the backwall is more important.
8.9
Maintenance and Repair
8.9.1
Problems with Jointless Construction
Although integral-type bridges will eliminate some of the more troublesome prob-
lems associated with jointed bridges and yield significantly more durable structures,
they will not eliminate endemic highway construction problems that are somewhat
related to accelerated construction, all-weather construction, marginal construction
super­vision, and other construction issues identified in Chapter 3 on materials and
Chapter 4 on bridge decks.
Transverse and diagonal deck slab cracks, stage construction issues, lateral rota-
tion of superstructure, erosion of embankments, marginal quality of structure move-
ment systems, and other problems have appeared to trouble design, construction, and
maintenance engineers. Except for early-age deck slab cracking, these problems are
generally the result of failure to anticipate and apply typical design and construction
provisions to achieve trouble-free construction and more durable structures.
8.9.1.1
Deck Cracking
Diagonal deck slab cracks located at acute corners of integral-type bridges are occa-
sionally reported. When constructing integral-type bridges, stationary abutments and
moving superstructure must be joined together by cast-in-place continuity connections.
Consequently, these connections could be stressed and cracked if a substantial tem-
perature drop were to occur during initial concrete setting, or if concrete placement
sequences were not suitably controlled. To address this problem, one or more of the fol-
lowing procedures should be used: place continuity connections at sunrise, place deck
slab and continuity connections at sunrise, place continuity connections after deck slab
placement, or use crack sealers.
8.9.1.2
Lateral Rotation of Semi-Integral Bridges
One of the primary aspects of semi-integral bridges that must be considered and ad-
dressed is the design and proper orientation of guided bearings for the superstructure
of skewed bridges. Unfortunately, many of the retention devices currently being used
are not fully functional because of friction and binding and consequently, the long-term
stability of some abutments, especially those not supported by rigid foundations, may
not have been provided for effectively. However, it appears inevitable that this aspect

408

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


of the semi-integral bridge concept will be improved when bearing manu­facturers and
bridge design engineers combine their expertise to design and manufacture more func-
tional structure movement systems for these applications.
8.9.1.3
Approach Slabs
Shortly after the state of Ohio adapted the integral concept to continuous steel beam
bridges in the early 1960s, slab distress was experienced. Where the bridges in ques-
tion were constructed adjacent to compressible asphalt concrete approach pave-
ments, approach slab seats at the ends of bridge superstructures were found to be
fractured, approach slabs had settled, and the vertical discontinuity in the roadway
surface at the approach slab–superstructure interface was hindering movement of
vehicular traffic.
8.9.1.4
Drainage
Washout has been noted on several existing structures in which drainage was not
properly designed or maintained, including some where the piles became exposed. It is
imperative that proper drainage material including geotextiles and perforated piping
be placed behind the abutments. The preferred alternative is to direct water away from
the bridge approach, but it is acknowledged that this can be difficult to accomplish in
many cases.
In addition, improper drainage can lead to washout at one end of the bridge and
not the other. For semi-integral abutments, this leads to an unbalanced soil pressure,
which can lead to additional maintenance issues at the bearing locations.
Drainage can also affect settlement of the sleeper slab and create settlement of the
approach slab. It is recommended that runoff be intercepted or diverted so that it does
not reach the end of the approach slab.
In regions that experience freezing temperatures, proper drainage is also important
to minimize the potential for frozen soil behind the abutment. The magnitude of the
potential restraining force is unknown for frozen soil, but it will be minimized with
proper drainage (Briaud et al. 1997).
8.9.1.5
Cycle-Control Joints
Probably the most significant unresolved problem regarding integral and semi-integral
bridges is the availability of cost-effective functional and durable cycle-control joints,
which are the moveable transverse joints used between approach slabs of integral-type
bridges and approach pavements. The usual pavement movement joints, composed
of preformed fillers, are currently being used for the shortest bridges. For the longest
bridges, finger-plate joints with easily maintainable curb inlets and drainage troughs
have been successfully employed. However, for intermediate-length bridges, develop-
ment of a suitable cycle-control joint is still in the evolutionary stages.

409

Chapter 8. JOINTLESS BRIDGES


8.9.2
Deck Replacement
Figure 8.42 shows what can happen when the proper procedures and sequences for
deck replacement and integral abutment backfilling are not followed. It should be
anticipated that large compressive forces are acting on the whole structure as a result
of soil pressure on the abutments and restrained expansion of the girders. To ensure
the global stability of the structure, one of two procedures must be followed. The first
procedure, which should always be used for whole deck replacement, is to use proper
construction sequencing as follows:

1. Remove the approach slab.


2. Remove backfill to the bottom of the stem for integral abutments or to the bottom
of the end diaphragm for semi-integral abutments. Excavation should be done
simultaneously behind both backwalls.
3. Remove the deck.
4. Replace the deck according to the guidance provided in Section 8.8.4.

The second option is to calculate the stress applied by the passive pressure of the
abutment backwalls. This force can then be applied to the superstructure with por-
tions of the deck removed to check the stability of the system and each structural item
that might be affected by the removal of the deck. This procedure includes checking
both local and global buckling stability. It is recommended that this be used only for
partial-width deck repair.
8.9.3
Bearing Replacement for Semi-Integral Jointless Bridges
An additional factor when detailing semi-integral jointless bridges, should bearing re-
pair or replacement be required, is to incorporate appropriate features at the initial
design stage to facilitate superstructure jacking. In general, since superstructure and
abutments in semi-integral bridges are separated by elastomeric bearings, it would
be easy to place flat jacks between the abutment and the end diaphragm to raise the
super­structure and approach slab and allow replacement of the bearing.

Figure 8.42.  Buckled girder flanges resulting from improper


deck sequencing.

410

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


8.10
Retrofits
A large percentage of existing bridges are simple-span bridges that rely on expansion
joints at piers and abutments to accommodate longitudinal movements. Most of the
deficient bridges in the United States include these jointed structures, which require
upgrade and repair. Retrofitting existing jointed bridges to jointless ones is highly
recommended.
The following considerations are required in integral conversion (Leathers 1990):

1. The existing structural elements should be able to properly function without the
expansion joint.
2. Movement calculation should be based on the LRFD specifications.
3. Continuity can be achieved by making either the deck or the girders continuous.
4. All obsolete and deteriorated bearings should be replaced with elastomeric bearing
devices.
5. If the abutment is unrestrained, a fixed integral condition can be developed for
many of the shorter bridges. Abutments that are free to rotate, such as a stub abut-
ment on one row of piles or an abutment hinged at the footing, are considered un-
restrained. A semi-integral condition can be developed for restrained abutments.

8.10.1
Details over the Pier
Two practical options that can be used with or without integral abutments are avail-
able for retrofitting existing jointed bridges into jointless bridges:

1. Provide beam continuity for live load only. In this case, the negative moment con-
tinuity is provided over the piers, with or without positive moment continuity at
these locations; or
2. Provide deck slab continuity only. In this option, although the deck is continuous,
beams are technically, simply supported. This method involves removing some
length of slab at the ends of the adjacent beams, splicing the existing reinforcement
and adding new bars, then recasting that part of the deck.

When retrofit of an existing open joint is considered, the following approach may
be used, as shown in Figure 8.43 (note for this detail only the deck is made continuous):

1. Remove concrete as necessary to eliminate existing armoring.


2. Add negative moment steel at the level of existing top-deck steel sufficient to resist
transverse cracking.
3. Reconstruct with regular concrete to original grade.

Because the deck slab will be exposed to longitudinal flexure as a result of the rota-
tion of beam ends responding to the movement of vehicular traffic, cracks will occur
over the link slab. However, for short- and medium-span bridges, the deck cracking

411

Chapter 8. JOINTLESS BRIDGES


Figure 8.43.  Integral conversion at piers.
Source: Leathers 1990.

associated with such behavior is preferred over the long-term consequences associated
with open movable deck joints or poorly executed joint seals.
8.10.2
Details over the Abutment
For existing stub abutments with a single row of piles, the following procedure (shown
in Figure 8.44) should be used (integral abutment retrofit):

1. Check the capacity of piles and the pile cap connection for the expected movement.
2. Remove the approach slab, and excavate backfill to the elevation of the existing
ground on the front face.
3. Demolish the existing backwall to the top of bridge seat. Cast reinforced concrete
around beam ends.
4. Provide drainage, backfill, and new approach slab behind the new abutment.

For existing stub abutments with rigid foundation or existing full-height wall abut-
ments, the following procedure (shown in Figure 8.45) should be used (semi-integral
abutment retrofit):

1. Remove the approach slab and excavate the backfill to the elevation of the existing
ground on the front face.
2. Remove the existing abutment backwall to the top of the bridge seat.
3. Provide a sliding surface between the pile cap and the abutment stem, which is cast
integrally with the beam ends and approach slab.
4. Provide details for both horizontal and vertical sliding joints by using lateral guide
bearings, sheet seals, and drainage and backfill.

8.10.3
Converting Jointed Bridges to Jointless Bridges
General experience has shown that most common bridge types can be converted to
jointless bridges to enhance their performance with the same goal as new construc-
tion (i.e., joint elimination). Examples of candidates that have already been converted
are pin-and-hanger bridges and multispan and simple-span bridges for both steel and
concrete superstructures.
Several states have had success converting old pin-and-hanger expansion joints to
a bolted full-moment connection, thus eliminating the expansion joints.

412

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


(a) (b)

Figure 8.44.  Conversion of a bridge with moveable deck joints at the superstructure–
abutment interface with integral abutment (a) before and (b) after conversion.

(a) (b)

Figure 8.45.  Conversion of a very-short-span bridge with moveable deck joints at


the superstructure–abutment interface with integral abutment (a) before and (b) after
conversion.

413

Chapter 8. JOINTLESS BRIDGES


The state of New Mexico has presented several case studies (Maberry et al. 2005).
In one of them, simple-span concrete girders were converted by incorporating a link
slab. The project demonstrated that attention must be paid to the bearings. The greatly
increased expansion that would transfer to the outer bearing locations was overlooked
by the retrofit assessment. The resulting expansion loads were absorbed by the pile
caps, which quickly deteriorated.
The key to any conversion is the ability of the bridge to withstand the new conti-
nuity loading and expansion demands introduced by the changed load path. Because
of the complex nature of the converted structure, it is recommended that conversions
be treated with the same level of analysis as required for a new design.

414

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


9
Expansion Devices

9.1
Introduction
Bridge elements are subjected to various loads, including traffic and environmental
loads that result in movement of the bridge elements. One of the key factors affecting
the service life of bridges is how to address thermal expansion and contraction of the
bridge elements. This design issue is handled in two distinct ways. One option is to
provide expansion joints at designated locations within the superstructure. By doing
so, the designer forces the entire thermal deformation to take place at these discrete
locations. The other option is to make the superstructure and deck continuous and
assume that the thermal movement will be accommodated by the flexibility of the sub-
structure, such as with the use of integral abutments. In such cases, the joint is usually
moved away from the bridge abutment and placed near the end of the approach slab.
This chapter provides essential information for enhancing the service life of bridge
expansion joints. The process begins with developing an understanding of the types
of bridge joints that may be considered for a particular project. The viable expansion
joint types are further evaluated for factors that adversely affect their service life, and
individual strategies are developed to mitigate these adverse effects. The overall strat-
egy selection is then developed, blending these sometimes-conflicting individual strate-
gies. The components of an overall strategy should include

• Identification of appropriate design methodologies,


• Selection of durable expansion device types considering life-cycle costs,
• Specification of best practices for construction, and
• Development of an effective maintenance plan.

415
Section 9.2 describes the different types of expansion devices used in practice, as
well as their various advantages and disadvantages. Section 9.3 discusses reported
potential factors that could affect the service life of the different expansion devices,
and Section 9.4 provides strategies for enhancing expansion device service life.
9.2
Descriptions of Various Expansion Joint Devices
Many types of expansion joints exist to meet the needs of a wide array of bridge
types. However, it is important for the designing engineer to select the proper expan-
sion joint, as this is the most crucial step in enhancing the service life of an expansion
­device. When selecting the proper expansion device it is important to consider the
total anticipated displacements, bridge skew, and other special considerations, such as
proprietary system requirements and accommodations for deicing agents.
An ideal expansion joint (Lee 1994) is able to

• Accommodate thermal expansion and contraction,


• Accommodate movement caused by traffic-induced loads,
• Provide a smooth ride,
• Prevent the creation of hazards and safety issues,
• Accommodate needs during snow removal,
• Prevent leaking of moisture and other chemicals to elements below the superstructure,
• Have a long service life,
• Be maintenance free or require minimal maintenance, and
• Be cost-effective.

There are many different joint types, and often the terminology used to describe
them differs from agency to agency or manufacturer to manufacturer. One approach
to categorizing expansion joint systems is to divide the various joints into three catego-
ries on the basis of the maximum longitudinal movements that they can accommodate,
as follows:

1. Expansion joints capable of accommodating small longitudinal movements (less


than about 3 in.),
2. Expansion joints capable of accommodating medium longitudinal movements (be-
tween 3 and 5 in.), and
3. Expansion joints capable of accommodating large longitudinal movements (in
­excess of 5 in.).

The three types of expansion joint are described in the remainder of this section.

416

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


9.2.1
Expansion Joints with Small Movement Capabilities
9.2.1.1
Compression Seal Joints
In an expansion joint that uses a compression seal, the opening throughout the entire
bridge-deck width is filled with a neoprene elastomeric section that forms a waterproof
joint, as shown in Figure 9.1 (Purvis 2003). The neoprene elastomeric is installed by
squeezing and inserting the seal into a preformed joint opening (Chang and Lee 2002).
Steel or special concrete materials are sometimes used to strengthen the joint face. To
facilitate the movement of the joint, the seals are usually made semihollow with inter-
nal diagonals and vertical neoprene webs (like a truss structure). However, some types
are made of closed section foam. The squeezing of the neoprene elastomeric section is
meant to ensure that it will be in compression, thereby sealing the joint. Figure 9.1b
shows an armored version of a compression seal joint. A recent trend within transpor-
tation agencies has been to eliminate the horizontal leg of the angle used and to attach
the vertical leg to the deck with shear studs. This expansion joint type allows total
movement of up to 3 in.
9.2.1.1.1
Features

• Movement from 0.25 to 3.0 in. is allowed.


• The seal needs to be made the right size for the needed range of movement.
• The semihollow cross section with the internal diagonal and vertical neoprene
webbings forms a truss action that allows the section to compress toward the sides
of the joints and become watertight.
• Sometimes foams are used instead of semihollow cross sections; these foams form
a closed section, but the movement range is the same.

(a) (a) (b) (b)

Figure 9.1.  (a) Plain and (b) armored compression seal.


Source: Purvis 2003.

417

Chapter 9. EXPANSION DEVICES


• The compression seal fits into the sides of the joint by using a lubricant material
that also functions as an adhesive that bonds the seal to its place.
• Splices should be avoided in this type of seal.

9.2.1.1.2
Advantages

• Some agencies report minimal required maintenance and a good life span for these
seals.
• This type of joint is recommended mostly for areas with moderate temperature
extremes.

9.2.1.1.3
Disadvantages

• It is reported that large sustained compressive movement forces air from the seal
material, which may not recover when the joint opening expands.
• Some agencies do not report good performance for this type of seal, citing damage
from debris, traffic, and snowplows.
• Leakage has been reported shortly after installation.
• It has been reported to dislodge and lose compression over time.
• This type of joint is not recommended for extreme temperature ranges.
• The seal needs to be installed at relatively low temperatures, or it is more difficult
to install and prone to damage.
• According to some reports, the compression seal loses its capability to retain initial
compression recovery due to loss of resilience, particularly if the movement range
is large.

9.2.1.2
Poured Sealants
The poured sealant systems used today mainly consist of two parts: a polyethylene
foam backer rod and a pourable silicone sealer (Purvis 2003), as shown in Figure 9.2.
The backer rod keeps the sealant from spilling through the opening. During installa-
tion, it is very important to keep the joint edges clean so that the silicone sealant bonds
tightly. The performance of this type of joint will improve if it is poured when the
ambient temperature is at the middle of the historical range.
Traditionally, this type of joint could handle movements up to 3/16 in., but newer
systems can accommodate up to 3 in. of movement (Purvis 2003; Chang and Lee
2001).

418

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 9.2.  Poured sealant joint type.

9.2.1.2.1
Features

• Traditional systems were good for shorter spans on which the movement need was
3
/16 in. (5 mm) or less. However, newer systems accommodate larger movements
depending on the sealant material.
• The most common sealant used today is silicone.

9.2.1.2.2
Advantages

• This type of joint is maintaining popularity because, according to reports, newer


systems are serving well if installed properly.
• The silicone polymer used is self-leveling and rapid curing.
• The performance of this type of seal is not affected by joint walls that are not
perfectly made.
• They are easy to repair.
• It is easy to remove a portion of the sealant, clean the sides of the joint, and refill
the joint.
• The quick maintenance process makes traffic disruption very short.

9.2.1.2.3
Disadvantages

• The first poured seal materials were heated asphalt or coal tar products that did
not perform satisfactorily for many transportation agencies.
• Polymer materials used to have problems such as debonding, splitting, and dam-
age from noncompressible debris.
• Earlier, damage to the deck edge also caused the joint sealant to fail.

419

Chapter 9. EXPANSION DEVICES


9.2.1.2.4
Requirements

• The thickness of the silicone at the center should be no more than half the width
of the joint.
• It is important that the bottom of the silicone does not bond to the material below.
• Poured sealants perform best if the seal is poured when the ambient temperature
(which must be above 40ºF) is at the middle of the historical range or the joint
opening is at the midpoint.

9.2.1.3
Asphaltic Plug Joints
Figure 9.3 shows a typical detail for an asphaltic plug joint. It is a simple detail that
can be used with concrete deck overlays. In this system, a center opening around the
joint is prepared (about 20 in. wide and 2 in. deep), and a steel plate is placed in the
opening as shown in Figure 9.3. The asphaltic material is then placed over the steel
plate to seal the joint. The maximum movement allowed by this system is about 2 in.
(Pemmaraju et al. 2006; Mogawer and Austerman 2004).
9.2.1.3.1
Features

• Movements less than 2 in. (50 mm) are allowed.


• The system is suitable for concrete decks, especially when used with an overlay layer.
• “A popular application is on decks where a waterproof membrane, topped by
bituminous (asphalt) concrete overlay, has been added” (Purvis 2003).
• This system may work better in restricted climate conditions.

9.2.1.3.2
Advantages

• Installation and repair are easy.


• Installation and repair are low cost.
• There is a low susceptibility to snowplow damage.
• The system can be cold milled.

9.2.1.3.3
Disadvantages

• “This seal was developed almost exclusively for bridge-deck joints without curbs,
barriers, parapets, etc., and does not provide an effective method of sealing joint
upturns, especially for longer decks and skewed deck joints where joint movement
will degrade the system, resulting in early system failure” (Purvis 2003).
• Some problems have also been reported, among them are softening in hot weather,
debonding of the joint–pavement interface, and cracking in very cold weather.

420

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


(a)

(b)
Figure 9.3.  Asphaltic plug joints: (a) typical detail and (b) commercially available details.
Source: (a) Mogawer and Austerman (2004) and (b) courtesy D.S. Brown.
Sources: (a) Mogawer and Austerman (2004) and (b) courtesy D.S. Brown.

• In areas of heavy traffic volumes, this seal may rut or delaminate over time.
• Rapid temperature changes can damage this type of seal.

9.2.1.4
Sheet Seal Joints
As shown in Figure 9.4, the sheet seal joint consists of a sheet of fiber-reinforced elas-
tomeric membrane tied to both sides of the joints. This joint type can accommodate
up to 4 in. of longitudinal movement. The kink in the membrane allows the expansion
and contraction of the joint, while keeping it watertight. This system could be used for
rehabilitation of medium-span bridges. The membrane used in this system is provided
in a variety of shapes, configurations, and sizes (Chang and Lee 2001).

421

Chapter 9. EXPANSION DEVICES


Figure 9.4.  Sheet seal expansion joint.
Source: Yuen 2005.

9.2.1.5
Sliding Plate Joints
Figure 9.5 shows a typical sliding plate joint. According to some literature (Purvis
2003), this type of joint could also be classified as an open joint, as it does allow some
level of leakage while preventing most debris from passing though the opening. This
system consists of a steel plate spanning an open joint and embedded in adjoining deck
slabs. It can accommodate movements up to 3 in. Currently, these joint types can even
handle larger movements, depending on the thickness of the plate used, and they are
not used without a trough below.
9.2.1.5.1
Advantages

• Sliding plate joints effectively prevent debris from passing through the opening.

9.2.1.5.2
Disadvantages

• The system is noisy under traffic as a result of loosening over time.


• It is prone to safety hazards due to detaching.
• Anchorage problems can occur between the plate and the concrete.
• Anchors can be corroded over time and are prone to fatigue damage from traffic
loads.

Figure 9.5.  Sliding plate joint system.


Source: Yuen 2005.

422

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


• Damage accumulated at the unsupported edge of the sliding plate causes a weak
spot for traffic-impact loads.
• Snowplow blades can damage both the sliding plate and the anchors.
• The impact on the plate from traffic loads can be increased due to deterioration of
the roadway surface close to the joint.
• Sliding joints are not recommended for highways with considerable truck traffic
loads.
• It is not considered a watertight joint.

9.2.1.6
Open Joints
Figure 9.6 shows an example of an open joint with trough and armoring angles. The
early versions of open joints were constructed without trough and armoring angles;
however, most departments of transportation (DOTs) are no longer using open joints
without a trough. These joint types are best suited for small movements (less than
1 in.). Armoring provides protection against traffic impact, and the trough protects
bridge elements below the joints from moisture and debris that cause deterioration.
9.2.2
Expansion Joints with Medium Movement Capabilities
9.2.2.1
Strip Seals
Figure 9.7 shows a typical configuration for a strip seal joint. The main elements are a
V-shaped membrane made using elastomeric material mechanically attached to an ex-
truded metal piece anchored to the joint edges with studs (Purvis 2003). The maximum

Figure 9.6.  An example of an open joint with


trough and armoring angles.

423

Chapter 9. EXPANSION DEVICES


Figure 9.7.  Strip seal expansion joint.

movement that can be accommodated by this joint system is about 5 in. (Chang and
Lee 2002). Strip seal joints have gained popularity in recent years, and most DOTs are
relatively more satisfied with strip seal joints than with other joint types.
9.2.2.1.1
Advantages

• Strip seals are the most positively evaluated seals by agencies surveyed during the
study reported in NCHRP Synthesis 319 (Purvis 2003).
• The seal is watertight if properly installed.
• Under desirable conditions, strip seals have a longer service life than any other
kind of seal.

9.2.2.1.2
Disadvantages

• Replacement is difficult.
• Strip seal splices should be avoided.
• Snowplow blades can damage the seal.
• Problem areas for this type of seal usually occur at locations of rapid cross-section
change, both in the vertical or horizontal direction of the deck, like the gutter line.
• During expansion, noncompressible material at tiny cracks could create membrane
tears. This material can also cause ruptures, resulting in loss of watertightness.
• Occasionally, the seal comes loose from the extruded metal holding piece.

9.2.3
Expansion Joints with Large Movement Capabilities
9.2.3.1
Modular Expansion Joints
Modular joints were developed to accommodate large longitudinal expansion and
contraction movements by combining multiple neoprene strip elements. Depending
on the number of combined strips, modular joints can accommodate movement from
6 to 28 in. Modular expansion joints are mainly used for bridges with more complex

424

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


movements. They consist of several metallic pieces, which makes them prone to fa-
tigue. The modular expansion joint system consists of sealer, separator beams for seal-
ers, and support bars for separator beams. Sealers can be strip, compression, or sheet
seal type. Separator beams are usually extruded or rolled metal shapes to join the seals
in a series. The separator beams are supported on support beams at frequent intervals
(Chang and Lee 2001). Figure 9.8 shows a schematic of one commercially available
modular joint.
9.2.3.1.1
Features

• The system consists of three elements: sealers, separator beams, and support bars.
• Modular joints are considered a closed joint system, so they protect the compo-
nents below from damage due to water, salt, and other problems associated with
deck runoff.
• They are used for all types of long spans. It is the best alternative for large
movements.

9.2.3.1.2
Advantages

• They can accommodate large movements over 4 in. (100 mm); the typical move-
ment is between 6 and 24 in.

Figure 9.8.  Modular expansion joint system.


Source: Courtesy D.S. Brown.

425

Chapter 9. EXPANSION DEVICES


• They can be sized according to the magnitude of movement and have been de-
signed to accommodate movements of more than 7 ft (very long span).
• Joint seals have been continuously improved.
• The system is improving with experience and research. New design provisions
in the LRFD Bridge Design Specifications (AASHTO 2012) incorporate recent
fatigue-related research.

9.2.3.1.3
Disadvantages

• Problems have included fatigue cracking of welds (older designs); damage to the
equalizing spring, the neoprene sealer material, and the support; and damage from
snowplows.
• The system is complex and must be capable of expanding and contracting to
­accommodate very large movements and must remain watertight while being sup-
ported by a movable framing system.
• The high initial and maintenance costs have caused many DOTs to return to u
­ sing
finger joints for large movements and placing additional emphasis on proper
drainage.
• There are only two primary suppliers, which limits competition.
• Mixed performance results have been reported.

9.2.3.1.4
Requirements

• Some agencies use modular expansion joints exclusively to accommodate large


movements.
• Watertightness and fatigue resistance are important considerations, and some states
have developed or are developing specifications to ensure these considerations.

9.2.3.2
Finger-Plate Joints
Figure 9.9 shows a typical finger-plate joint system. Two pieces of steel are anchored
either to the deck or the steel superstructure and cover the joint opening. Grooves
in each of the plates fit loosely together like fingers. The majority of the finger-plate
joint systems in use have a trough, as shown in Figure 9.9, to catch water and debris
and carry them away from the superstructure to prevent deterioration (Purvis 2003;
Chang and Lee 2001). A variation of this concept is to offset the location of the finger
assembly to one side of the deck opening so that the fingers from one side slide over
an embedded plate in the deck on the other side. The advantage of this design is that
the deck opening is covered, which limits the amount of debris that can fall through.

426

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 9.9.  Examples of a finger-plate joint system.
Figure 9.9. Examples of a finger-plate joint system.

9.2.3.2.1
Features

• The system consists of finger-like plates attached to both sides of the super­structure
and usually includes a drainage trough.
• It is considered an open-joint system.

9.2.3.2.2
Advantages

• These joints can accommodate large movements (over 3 in.).


• The system experiences fewer durability problems in comparison with modular
joints and tends to have fewer problems than many other joints.
• DOTs have opted to use finger joints on longer spans.

9.2.3.2.3
Disadvantages

• The upward bending of the ends of the fingers results in increased noise, a rough
riding surface, and occasionally broken fingers.
• The most common problem is concrete deterioration around the joint.
• The only corrosion protection for the components below is the drainage trough.
• Problems are usually caused by poor drainage trough design.
• Cleaning and flushing are often difficult and are rarely performed.

427

Chapter 9. EXPANSION DEVICES


9.2.3.2.4
Requirements

• Watertightness and good drainage are important considerations for finger-plate


expansion joints. Use of appropriate slope (minimum 1%) can easily address this
requirement.
• When using finger-plate expansion joints, following the installation requirements
specified by the manufacture is important.
• Maintenance of the trough and routinely cleaning the debris are essential when
using finger-plate expansion joints.

9.3
Factors and Considerations Influencing Expansion Joint
Service Life
Joint performance is perhaps the most important factor affecting the deterioration
of bridge elements. A leaky joint allows salt and other chemicals to penetrate below
the deck and causes many maintenance and deterioration problems. A FHWA study
(Fincher 1983) reported that over a 5-year period, more than 60% of joints evaluated
were leaking water, and the other 40% had problems that were actually decreasing
their service lives. Wallbank (1989) evaluated 200 concrete bridges and found leaking
expansion joints to be the major cause of bridge element deterioration.
The reduced service life of bridge expansion devices is primarily related to deficien-
cies, which may be load induced; caused by man-made or natural hazards; or result
from production defects in construction processes and/or design details or operational
procedures. These deficiencies are illustrated in the fault tree shown in Figure 9.10.
9.3.1
Load-Induced Expansion Joint Considerations
Load-induced bridge-deck deterioration can be attributed to either loads induced by
the traffic that the bridge-deck expansion device carries or characteristics dependent
on the overall bridge system. These load-induced factors are introduced in the fault
tree in Figure 9.11.
9.3.1.1
Traffic-Induced Load Considerations
Traffic-induced loads include the effects of truck and other vehicular traffic on the rid-
ing surface of the bridge. Bridge-deck loading, and thus the loading on the expansion
devices, has a degree of uncertainty that must be addressed during the design of the
bridge, including vehicle weights and frequency of application; vehicle axle and tire
spacing; vehicle location on the deck; and variability in vehicle suspension systems,
which can affect impact assumptions. The assumptions in loading must be carefully
reviewed when establishing the criteria to be used and checked for bridge expansion
device design. Typically the service life of bridge-deck expansion devices will be af-
fected by fatigue response to repetitive loading, overload, and wear and abrasion.

428

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Reduced Service Life of
Bridge Deck

Caused by
Caused by Deficiency
Obsolescence

Natural or Man-Made Production/


Load Induced
Hazards Operation Defects

Figure 9.10.  Expansion joint reduced service life fault tree.

9.3.1.1.1
Fatigue
Fatigue is caused by the repetition of applied loads that result in a degradation of the
strength resistance of the components used to resist tensile stresses that occur in expan-
sion devices (especially in modular expansion joints and other large-movement joints).
9.3.1.1.2
Overload
Despite weight limit regulations in most states that define load limits for permit and
legal truck configurations, overloads exceeding these limits are not specifically con-
trolled. There are few weigh stations for checking these loads, and they are usually
only found on major roadway facilities, such as Interstate highways. Enforcement of
the laws against these overloads on other facilities is limited to spot checks by code
enforcement officials.

429

Chapter 9. EXPANSION DEVICES


Load Induced

Traffic-Induced System-
Loads Dependent Loads

System-
Snow
Fatigue Overload Wear Thermal Framing
Plow
Restraint

Figure 9.11.  Load-induced deficiency fault tree.


Figure 9.11. Load-induced deficiency fault tree.

Overloads result in additional flexural stresses in bridge decks and expansion


devices that can cause excessive cracking and movements not accommodated by the
original design. Heavier tire loads may also affect the wear and abrasion on the struc-
ture. Multiple applications of these loads can also affect the fatigue behavior of the
deck expansion devices.
9.3.1.1.3
Wear and Abrasion
Wear and abrasion is typically affected by high traffic volumes, high tire loads, and the
types of tires used by the vehicles. In cold climates, tires may have enhanced features
to aid in traction, such as deep grooves, studs, and chains. These added tire features,
although they aid traction, can abrade the surface of the bridge expansion joints.
9.3.1.1.4
Snow Plow
Impacts from snow plows are special considerations that need to be addressed in the
selection of expansion joints. Repeated impacts from snow plows can cause expansion
devices to function incorrectly and even make driving over them unsafe.

430

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


9.3.1.2
System-Dependent Loads
System-induced loads include the effects of the bridge system configuration on the
behavior of the bridge-deck expansion devices. These effects are accentuated by re-
straints provided through the bridge system and bridge-deck boundary conditions and
can result in significant locked-in stresses.
9.3.1.2.1
System-Framing Restraint
Boundary conditions for the longitudinal and transverse restraint of bridges can result
in tension that may lead to cracking in bridge decks under cold-weather conditions.
These boundary condition restraints are introduced through design by the choices
made relating to bearing types and their fixity and sliding assumptions. Fixed bear-
ings provide an anchor that is intended to restrict “walking” of the superstructure,
which results from shrinkage and cycling of expansion and contraction; they are usu-
ally located longitudinally near the point of zero movement of a supported multispan
superstructure unit. These fixed bearings, used in combination with bearings designed
to allow the superstructure to either move or slide over the top of the substructure,
reduce restraining forces that would otherwise be required to resist the movement.
9.3.1.2.2
Thermal
Temperature changes can result in changes in bridge movement, which can affect the
service life of expansion joints.
9.3.2
Natural or Man-Made Hazard Considerations
The environment to which the bridge deck is subjected can have a significant influence
on the service life of bridge decks and consequently the expansion devices.
These environmental influences include hazards from both natural and man-made
sources and include effects from areas with adverse thermal climates, coastal climates,
and chemical climates, as well as from chemical properties of the materials and outside
agents, such as fire. These natural and man-made hazards are introduced in the fault
tree in Figure 9.12.
9.3.2.1
Thermal Climate
Thermal climate influences on bridge-deck expansion device service life performance
are primarily due to cold-weather climates. These influences are both man-made, from
the application of deicing salts, and natural, in the case of freeze–thaw.
9.3.2.1.1
Application of Deicing Salts
Agencies in cold-weather climates dealing with ice and snow on roadways and bridges
have traditionally applied deicing salts to melt the ice and snow to facilitate tire trac-
tion on their facilities. These chloride-laden compounds can corrode expansion joint
steel components.

431

Chapter 9. EXPANSION DEVICES


Natural or Man-
Made Hazards

Thermal Climate Coastal Climate Chemical Climate

Corrosion-
Deicing Salts Salt Spray Inducing
Corrosion Chemicals

Freeze–Thaw

Figure 9.12.  Natural or man-made hazards fault tree.


Figure 9.12. Natural or man-made hazards fault tree.

The cross slope built into bridge decks for drainage purposes causes salt to wash
toward the bridge gutter adjacent to the traffic railing barriers bounding the bridge.
Removal equipment that scrapes snow from the bridge deck will also deposit residual
snow laden with deicing salts at this location, resulting in a very high concentration
of chlorides.
9.3.2.1.2
Freeze–Thaw
Spalling and cracking of the concrete deck as a result of freeze–thaw at the sites of ex-
pansion joints could have adverse effects on the joint-to-deck connections and possibly
on the performance of the device, as well.

432

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


9.3.2.2
Coastal Climate
Coastal climate influences on bridge expansion joint service life performance are pri-
marily due to chlorides introduced through salt spray and effects from high humidity.
These influences both occur naturally.
Coastal regions are subjected to a chloride-laden saltwater environment and a
combination of wind and wave action that causes these chlorides to become airborne
as salt spray. The susceptibility of the bridge expansion joints to these environmental
influences depends on the height of the bridge deck above the water elevation. The salt
spray wets the surfaces, leaving a chloride residual that can cause steel components of
expansion devices to corrode.
9.3.2.3
Chemical Climate
Chemical climate influences on service life performance can be attributed to corrosion-
inducing chemicals or other chemicals that can have detrimental effects on exposed
neoprene elements. These influences can occur naturally or can be man-made.
Corrosion-inducing chemicals from adjacent industries and residuals from pollu-
tion can be introduced to the bridge deck and joints and attribute to reduction in bridge
expansion joint service life. As an example, oil- and coal-burning facilities release sul-
fur dioxide and nitrogen oxide into the air, which cause acid rain consisting of sulfuric
and nitric acids. These acids can dissolve cement compounds in the cement paste and
calcareous aggregates and leave crystallized salts on concrete surfaces that can lead to
spalling and the corrosion of reinforcing bars at the site of expansion devices, as well
as components of the joints themselves.
9.3.3
Production and Operation Bridge-Deck Considerations
Decisions made for the production of bridge expansion devices and activities during
the bridge’s operation can have a significant influence on the service life of bridge ex-
pansion joints. These production and operation influences, which are introduced in the
fault tree in Figure 9.13, include decisions made during the design and detailing of the
bridge expansion joints and decisions regarding the quality of construction, the level of
inspection and testing performed during operations, and maintenance implementation.
9.3.3.1
Design and Detailing Bridge-Deck Considerations
Decisions made during the design and detailing phase of a bridge project can signifi-
cantly affect the service life of the bridge. It is incumbent on designers to understand
the implications of these decisions in order to make rational choices that will improve
the service life of bridge expansion devices.
9.3.3.2
Construction
Attention to good practices during construction and installation is crucial to the long-
term durability of expansion devices. Well-qualified and well-trained workers and
well-executed workmanship increase productivity, reduce material waste, and provide

433

Chapter 9. EXPANSION DEVICES


Production/
Operation Defects

Design/Detailing Construction Inspection Maintenance

Visual
Expansion
Installation
Joints

Nondestructive
Handling
Testing

Figure 9.13.  Production–operation and design–detailing defects fault tree.


Figure 9.13. Production–operation and design–detailing defects fault tree.

expected service life. Proper use of test methods is needed to ensure that quality expan-
sion joints are achieved.
9.3.3.3
Maintenance Plan: Monitoring of Condition
An effective maintenance plan must include provisions to adequately monitor the
structure in order to identify deficiencies at an early age so that appropriate repairs can
be performed before the deficiency propagates into an irreparable condition. The plan
should address proper inspection, collection of data, and prompt action on reported
deficiencies.
9.3.3.4
Inspection Requirements and Intervals
Inspection is a valuable tool for increasing the service life of bridges. Identification of
deterioration at an early stage is essential to providing corrective actions in a timely
manner in order to maximize the owner’s investment in the structure. The scope and
434

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


interval of these inspections should be calibrated on the basis of bridge expansion
­device performance history for the applicable deck hazard exposures. Inspection per-
sonnel should be properly trained to identify this deterioration.
9.4
Overall Strategies for Enhanced Bridge Expansion Joints
Service Life
Providing bridge expansion joints with enhanced service life requires a full under-
standing of the potential deterioration mechanisms. These mechanisms are associated
with load-induced conditions, local environmental hazards, production-created defi-
ciencies, and lack of effective operational procedures. This process is presented in this
section.
The intention of this section is to provide guidelines for selecting the most appro-
priate individual strategy to achieve the desired service life.
9.4.1
Design Methodology
With limited funds available for bridge construction, first-cost economy is often an
overriding factor in critical bridge joint selection decisions. However, to take advantage
of the long-term advantages of durable expansion joints, service life enhancement strat-
egies must apply a cascading series of economy, design, construction, and maintenance
measures. The success of the strategy selection process is dependent on the ­ability to
predict service life and the incorporation of best practices to enhance service life.
9.4.2
Expansion Joint System Selection
In general, selection of bridge expansion joints is historically based on locally accepted
construction practices and procedures, which result in an economical deployment.
Proposed enhancements to these systems would need to be assessed on the specific
long-term performance of these systems. As noted in Section 9.3, there are numerous
deficiencies that result in a reduction of service life, and these conditions may not nec-
essarily occur at the specific project site under consideration. It is the design engineer’s
responsibility to understand local practice and its performance history and to perform
sufficient investigation of the project site, its environment, and other design conditions
before selecting an expansion joint and its potential enhancements.
9.4.3
Life-Cycle Cost Analysis
Once a series of strategies has been developed, the evaluation of these strategies is
performed through a life-cycle cost analysis, which assesses the overall long-term cost
of a strategy throughout the target service life of the structure. Life-cycle cost analysis
is discussed in detail in Chapter 11 of the Guide.
The analysis should consider all costs associated with the construction required,
implementation trials and testing, and all future maintenance actions required through
the life of the structure.

435

Chapter 9. EXPANSION DEVICES


9.4.4
Construction Practice Specifications
Once the bridge expansion joint system is selected, a proper set of specifications must
be developed to ensure the appropriate standard of care is used during construction. It
should be noted that many agencies and manufacturers have specific requirements for
temperature setting for expansion joints to allow maximum range of movements; these
requirements need to be considered during design and construction.
9.4.5
Maintenance Plan
An effective maintenance plan must be developed to ensure the assumptions regarding
maintenance upkeep made in the bridge expansion joint selection process are prop-
erly identified for staff and budget requirements. If the bridge owner cannot com-
mit to such a program, then strategies for low-maintenance life-cycle costs should be
recommended.
9.4.6
Retrofit Practices for Expansion Devices
Expansion joints in bridges are not typically retrofitted as it is common practice to
replace the joints when they reach the end of their service life. Due to the perishable
nature of these expansion joints, it is important that they be properly maintained in
order to maximize the life of such devices.
9.4.7
Technology Tables for Expansion Joint Devices
Table 9.1 is a series of technology tables that summarizes service life and durability is-
sues related to expansion joints and aids the designing engineer in selecting the proper
expansion device. The table provides service life issues related to most expansion joints
used in practice. The information could be used in several ways at the design or main-
tenance level. The purpose of the technology table is to identify the most common
types of service life problems encountered in commonly used expansion joints; this
information can subsequently be used to provide possible solutions to service life prob-
lems encountered with expansion joints.
A review of the technology tables indicates that the following modes of failure are
observed when various expansion joints are used:

• Damage to various expansion joint components, such as steel armors, as a result


of snowplow usage;
• Inadequate design, installation, and maintenance;
• Accumulation of debris, leading to limited movement;
• Lack of timely maintenance;
• Limited service life; and
• Leakage.

The ultimate consequence of these service life problems is that joints start leaking
and causing damage to bridge elements below the deck.

436

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 9.1. Bridge Joints Technology Tables

Field-Molded (or Field-Formed) Joints


Range of Movement: Less than 1 in.
Expected Range of Service Life: 8 to 9 years
System
Preservation
Service Life Problem Solutions Advantages Disadvantages Requirements
Spalling and cracking of Deeper notch. Field-molded It requires It requires regular
adjacent concrete Filling with silicone. joints are maintenance 3-year inspection
inexpensive and and has a short and replacement
Beveling edges of concrete.
easy to install. service life. In in most cases.
Placement of correct They are best most cases, it
silicone thickness. suited for single- needs complete
Correctly mixing silicone span bridges replacement
material. with a maximum every 2 to 3
Installation problems Training installation length of years.
technicians. ~100 ft. This detail
is an economical
Providing detailed
solution for
installation plans by the
simple-span
manufacturer or contractor.
bridges with
Having a representative of
spans of less than
the joint supplier on site
~100 ft, especially
during installation.
in low-traffic
Snowplow damage Not allowing inferior quality areas where a
of bonding agents. few hours of
interruption to
Installing slightly below the traffic is not a
top of the deck elevation. major problem.
Debris accumulation Inspecting and cleaning
regularly.
Water leakage Draining water away from
the joint.
Insulation.
Hot-weather damage Using high-quality silicone
sealer.

(continued)

437

Chapter 9. EXPANSION DEVICES


Table 9.1. Bridge Joints Technology Tables (continued)

Plug Seal Joints


Range of Movement: Less than 2 in.
Expected Range of Service Life: 1.5 to 20 years
Service Life Problem Solution Advantages Disadvantages
Seal material does not Reapplication of the joint. Low instances of snowplow Damage due to sudden
always fill the joint damage. changes in temperature.
completely Low cost of installation and
Early joint failures in Avoid the system in skewed maintenance.
skewed and/or long and/or long decks.
decks

Softening in hot Proper selection of materials


weather considering climate
conditions.
Debonding of the joint– Reapplication of the sealant.
pavement interface
Cracking in very cold Proper selection of materials
weather considering the climate
conditions.
Rutting or delamination Reapplication of the joint.
over time due to heavy
traffic volumes
(continued)

438

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 9.1. Bridge Joints Technology Tables (continued)

Compression Seal Joints


Range of Movement: 0.25 to 2.5 in.
Expected Range of Service Life: 2 to 20 years
Service Life Problem Solution Advantages Disadvantages
High deformation due Use only at moderate Ease of installation. Poor workmanship, leading
to extreme temperature temperature extremes. Not very prone to snow to service life problems, is
variation plow damage. reported. In some instances,
Damage due to Use better consolidation of Have received very good poor workmanship
snowplow concrete around the steel performance ratings from has resulted in joints
armors. various agencies. leaking immediately after
installation.
Leakage after Use high-solids urethane
installation adhesive prior to insertion
of the seal.
Prevent twisting during
installation.
Dislodging over time Use better adhesive
materials to anchor the seal
to its place.
Installing at low
temperature.
Loss of compression Use material with lower
over time creep property.
Size using a working range
of 40% to 85% of its
uncompressed width.
Ozone sensitive Use material with high
resistance property for
ozone side effects.
(continued)

439

Chapter 9. EXPANSION DEVICES


Table 9.1. Bridge Joints Technology Tables (continued)

Sliding Plate Joints


Range of Movement: Less than 4 in.
Expected Range of Service Life: Up to 30 years
Service Life Problem Solution Advantages Disadvantages
Corrosion of steel Determine selection of Comparable to armored Movement of the sliding
material used corrosion-resistant material butt joints, except that plate can easily be hindered
(anchor). the sliding plate prevents, by debris that accumulates
Bending and fatigue of Use thicker plates. for the most part, the in the slot.
the sliding plate movement of debris below
Better design of surround
the joint.
material.
Use high-tensile-strength
steel.
Drainage (clogging) Trough with steep slope.
Conduct regular
maintenance.

Strip Seal Joints


Range of Movement: Less than 5 in.
Expected Range of Service Life: Up to 20 years
Service Life Problem Solution Advantages Disadvantages
Tearing of neoprene Replace neoprene Good performance record. If the membrane is set low,
membranes membrane. High degree of water the debris can accumulate
Corrosion Determine selection of tightness. faster than it can be
corrosion-resistant material maintained. If it is set high,
(armoring). it will be prone to snow
plow damage.
Debris accumulation Perform regular
High initial cost and
maintenance.
replacement.
Damage due to Set the neoprene
snowplow membranes lower.
(continued)

440

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 9.1. Bridge Joints Technology Tables (continued)

Finger-Plate Joints
Range of Movement: More than 4 in.
Expected Range of Service Life: 15 to 50 years
Service Life Problem Solution Advantages Disadvantages
Clogging of the joint Trough with steep slope. Suitable for large High initial cost and prone
Perform regular movements. to snowplow damage if
maintenance. Allows flushing of debris not designed correctly.
and prevents leakage and Horizontal misalignment
Anchorage failure Construction QC.
overflow. during construction can
Design. Good performance record. cause the fingers to jam.
Use stronger materials. Vertical misalignment can
Bent fingers (load Use thick plates. result in poor ride quality
related) and noise.
Use replaceable fingers.
Misaligned fingers Ensure proper installation
and avoid horizontal or
vertical misalignments.
Corrosion Design (corrosion-resistant
material).
(continued)

441

Chapter 9. EXPANSION DEVICES


Table 9.1. Bridge Joints Technology Tables (continued)

Modular Expansion Joints


Range of Movement: Greater than 5 in.
Expected Range of Service Life: 10 million cycles
Service Life Problem Solution Advantages Disadvantages
Fatigue cracking of High-purity materials, new Can accommodate very Incorporates many
welds fatigue design and details. large movement. components and is high
Water leakage at seal Curbs to guide drainage maintenance; therefore,
splice away. many agencies are reverting
to the use of finger-plate
Debris accumulation in Perform maintenance crew
joints.
seals clean out once per year.
Run joint through opening
in parapet with a catch
basin.
Direct drainage away from
joints in order to wash
debris away.
Reflective cracking in Use a thicker top plate to
concrete above support support box.
box Use an increased concrete
cover over support box.
Use a #3 reinforcing bar
over anchorage detail to
minimize cracks.
Drainage build-up Curbs or upturns.

442

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


9.4.8
Strategy Table for Expansion Joints
Selecting the appropriate expansion device for the required movement range is the
most critical step in enhancing the service the life of these joints. Table 9.2 summarizes
the information presented in this chapter and proposes recommended strategies when
expansion devices are to be used. The table is intended to serve as a guide to the de-
signer in selecting bridge expansion devices.

443

Chapter 9. EXPANSION DEVICES


444
Table 9.2.  Strategy Table for Expansion Joints

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Maximum Life-Cycle Cost Difficulty Expected
Longitudinal Potential Options to Associated Service
Movements Deterioration Improve with Performance Life
(in.) Strategy Mode Service Life Initial Maintenance Replacement Record (years)
Less than 1 in. Field-molded or other See Table 9.1 Provide second Low High Low Good 1 to 3
equivalent joint types layer protection
Between 1 and Compression seal See Table 9.1 Provide second Medium High High Very good 3 to 30
4 in. joint (3 in. max.) layer protection
Strip seal joints See Table 9.1 Provide second Medium High High Very good 3 to 30
layer protection
Larger than Finger-plate joints See Table 9.1 See Table 9.1 High High High Very good 10 to 50
4 in. Modular expansion See Table 9.1 See Table 9.1 Very high Very high Very high Good 10 to 50
joints
10
Bridge Bearings

10.1
Introduction
Bearings are a critical element within overall bridge systems. Although they represent
only a small part of the overall structure cost, they can potentially cause significant
problems if they function improperly or if possible maintenance, retrofit, or replace-
ment strategies are not envisioned and well planned at the design stage. Bridge super-
structures experience translational movements and end rotations caused by traffic load-
ing, thermal effects, creep and shrinkage, wind and seismic forces, initial construction
tolerances, and other factors. Bridge bearings are designed and built to accommodate
these movements and rotations while supporting required gravity loads, transmitting
those loads to the substructure, and providing necessary restraint for the superstruc-
ture. Proper functioning of bridge bearings is assumed in the analysis and design of
overall bridge systems. Bearing failure or improper behavior can lead to significant
changes in load distribution and overall structural behavior that are not accounted for
in the design and can significantly affect the superstructure–substructure interaction.
This chapter describes various bearing types and provides information concern-
ing factors affecting and increasing their service life. Methods for design for service
life are discussed along with needs for future inspection, maintenance, and possible
replacement.
10.2
Bearing Types
Many bearing types have been developed, primarily to provide efficient, economical
ways to accommodate various levels of load and movement. Each type has certain ad-
vantages and potential disadvantages. Table 10.1 identifies the commonly used bridge
bearing types discussed in this chapter.

445
Table 10.1. Bearing Types
General Category Bearing Type
Elastomeric bearings Plain elastomeric pads
Steel-reinforced elastomeric pads
Cotton duck pads
Sliding bearings Polytetrafluorethylene
Alternative sliding materials
High-load multirotational bearings Pot bearings
Disc bearings
Spherical bearings (cylindrical for unidirectional)
Fabricated steel mechanical bearings Fixed pin
Rocker or roller expansion

10.2.1
Elastomeric Bearings: Plain and Reinforced
Elastomeric bearings have become the most common type of bearing in recent years
because of their desirable performance characteristics, durability, low maintenance
requirements, and relative economy. Elastomeric bearings have no moveable parts.
They accommodate movement and rotation by deformation of an elastomeric pad,
which can be neoprene or natural rubber. Lateral and longitudinal movements are
accommodated by the pad’s ability to deform in shear. These bearings can accommo-
date combined movements in both longitudinal and transverse directions, and circular
elastomeric bearings have been used to accommodate multirotation requirements. Ex-
isting bridges using elastomeric bearings with more than 50 years of very good service
performance are reported.
Plain, unreinforced elastomeric pads are used for short spans on which loads and
movements can be accommodated by a single layer of elastomer.
As vertical load and movement requirements increase, thin reinforcing plates are
combined with multiple layers of elastomer to form a laminated reinforced elastomeric
assembly (see Figure 10.1). Steel and fiberglass reinforcement layers have been used;
however, fiberglass is weaker, more flexible, and does not bond as well to the elastomer
as does steel reinforcement. As a result, the use of thin steel-plate reinforcement has
become more common.
Neoprene is the most widely used elastomer, but some states also use natural
rubber (Stanton et al. 2004), particularly in colder climates, to meet AASHTO low-­
temperature requirements. Natural rubber generally stiffens less than neoprene at low
temperatures. Neoprene has greater resistance to ozone and a wide range of chemicals
than natural rubber, making it more suitable for some harsh chemical environments.
The LRFD Bridge Design Specifications (LRFD specifications) (AASHTO 2012)
currently provide two design methods for the design of elastomeric bearings: Method A,
which is the simpler method and has fewer testing requirements; and Method B, which
requires greater design effort and more extensive testing. Method A leads to viable

446

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 10.1.  Laminated elastomeric pad.
Source: Courtesy (left) D.S. Brown and (right) National Steel Bridge Alliance.

Source: Courtesy (left) D.S. Brown and (right) National Steel Bridge Alliance.

Figure 10.1.designs
Laminated elastomeric
for bridges pad.
up to 150 ft (Stanton et al. 2004), and Method B is generally used
when a reasonable bearing cannot be designed using Method A. The majority of states
use Method A (Stanton et al. 2004).
10.2.2
Cotton Duck Pads
Cotton duck bearing pads are another type of elastomeric bearings that are occasion-
ally used in some states, typically for precast concrete I-girder bridges with span lengths
up to the 150- to 180-ft range. Cotton duck pads (CDPs) are preformed elastomeric
pads consisting of very thin layers of elastomer (less than 0.4 mm [1/60 in.]) interlaid
with cotton or polyester fabric. They are stiff and strong in compression, giving them
much larger compressive load capacities than plain elastomeric pads; however, CDP
shear deflection capability is very limited. The CDP bearings provide a high stiffness in
the direction of applied compressive force and are helpful in limiting problems encoun-
tered during construction of heavy girders because of rotational instability, generally
observed with other elastomeric bearing types. For large shear strain, CDPs may split
and crack or result in girder slip on the CDP. The limited shear deflection capacity is
frequently overcome by the addition of a polytetrafluorethylene (PTFE) sliding surface
to accommodate large movement. When PTFE surfaces are used, they are often com-
bined with stainless steel sliding surfaces, similar to that shown in Figure 10.2. The
overall capacities depend on the stiffness and deformation capacity of the CDP and
vary from manufacturer to manufacturer. To assure adequate performance from CDP,
quality control (QC) testing measures and design recommendations have been devel-
oped and incorporated into the LRFD specifications (Lehman et al. 2003).

447

Chapter 10. BRIDGE BEARINGS


Figure 10.2.  Elastomeric bearing with
PTFE sliding surface.
Source: Courtesy D.S. Brown.

10.2.3
Sliding Bearings
10.2.3.1
Polytetrafluorethylene
When horizontal movements become too large for elastomeric bearings to reason-
ably accommodate in shear, PTFE sliding surfaces can be used to provide additional
­capacity (see Figure 10.2). They are commonly used to provide movement capability
with CDPs, and they are also used to provide for horizontal movement in combina-
tion with other bearing systems that internally provide for compression and rotation,
such as high-load multirotation (HLMR) pot and disc bearings (Figures 10.3 and 10.4,
­respectively). They are also used to accommodate large translations and rotations
when combined with spherical or cylindrical bearings.
PTFE has low frictional characteristics, chemical inertness, and resistance to
weathering and water absorption, making it an attractive material for bridge bearing
applications.
The sliding movement is typically provided by a very smooth stainless steel-plate
sliding on a PTFE surface. The stainless steel surface is larger than the PTFE surface
so that the full movement can be achieved without exposing the PTFE. The stainless
steel is typically placed on top of the PTFE to prevent contamination with dirt or
debris. PTFE sliding bearings may be guided, allowing movement in only one direc-
tion, or nonguided, allowing multidirectional movement. When PTFE sliding surfaces
are combined with elastomeric pads, the elastomeric pad must be designed to accom-
modate the shear force that is needed to overcome the PTFE friction resistance.
448

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Sliding surfaces develop a frictional force that acts on the superstructure, substruc-
ture, and bearing. As a result, friction is an important design consideration, and the
low frictional resistance of PTFE makes it very useful for this application. The coef-
ficient of friction of PTFE increases with decreasing temperature and with decreasing
contact pressure. It also increases if the mating surface is rough or contaminated with
dust or dirt. Proper design, fabrication, and field installation are all essential for proper
performance.
Plain, unfilled PTFE is the most common material used for sliding bearings. Filled
PTFE, with the addition of glass fibers, carbon fibers, or other chemically inert filler
reinforcement, is sometimes used. Filled PTFE has significantly greater resistance to
wear and creep, but it also has a higher friction coefficient by as much as 25% to 30%.
Unfilled PTFE in the form of a woven fabric is occasionally used to provide higher
bearing strength, longer wear, and increased creep resistance.
Lubrication significantly reduces the coefficient of friction, and dimpling of the
PTFE surface has been used as a means to facilitate lubrication. Dimples are spherical
indentations (0.32 in. maximum diameter by 0.08 in. minimum depth and covering
20% to 30% of the surface area) that are machined into the PTFE surface to act as
reservoirs for storing lubrication. Silicone greases are specified because they are effec-
tive at low temperatures and do not attack the sliding material. Dimpled and lubri-
cated PTFE has been used in Europe, but in the United States, it has been used only
in special cases on large spherical bearings in which a very low coefficient of friction
requirement is needed to reduce friction loads on substructures. Dimpled and lubri-
cated PTFE demands a routine maintenance plan, as the coefficient of friction will
significantly increase as the lubrication material is depleted. This increase in coefficient
of friction can have an adverse effect on the service performance of other parts of the
bridge system.
10.2.3.2
Alternatives to Plain PTFE
Maurer sliding material (MSM) is an alternative sliding material developed in ­Germany
as a better-performing substitute for current PTFE-based sliding material, mainly for
high-speed rail applications (Maurer Söhne 2003). The new material is an ultra-high-
molecular weight polyethylene that has performed well in recent field applications
and experimental testing in Europe, where it is one of the most popular sliding sur-
faces in use.
MSM was primarily developed to accommodate bridge movements and related
wear caused by high-speed trains, which induce high rates of movement due to girder
end rotations that result in large accumulated movement over time. Initial specifica-
tions required the bearing material to accommodate a rate of movement up to 15 mm/s
and provide 80 years of service life.
Experimental testing in Europe with dimpled and lubricated specimens subjected
to high loading rates has shown MSM to outperform PTFE in regard to compressive
strength, coefficient of friction, and rate of wear. But because this material is relatively
new, there are no long-term data available. More recently, research conducted under
SHRP 2 Project R19A compared coefficient of friction and wear between lubricated

449

Chapter 10. BRIDGE BEARINGS


and unlubricated MSM and plain PTFE specimens at high movement rates. Unlubri-
cated specimen tests showed MSM to have significantly greater wear resistance than
plain PTFE, but with a greater coefficient of friction. Project R19A testing also com-
pared coefficient of friction and wear of a glass-reinforced PTFE, Fluorogold, with
plain PTFE and MSM. Like MSM, the Fluorogold material had significantly greater
wear resistance, but with a smaller increase in the coefficient of friction.
10.2.3.3
Service Life Design Method for Sliding Surfaces
Appendix G provides further information regarding a potential service life design
method for sliding surfaces that considers a pressure–velocity factor in determining an
effective wear rate for the surface material. The method requires test data to establish
material wear characteristics; therefore, its application as a design method will be
subject to the availability of sufficient existing test data to establish reliable wear rate
curves for different sliding materials. The proposed design provisions are based on
research conducted by SHRP 2 Project R19A (Ala et al., submitted for publication).
10.2.4
High-Load Multirotation Bearings
When design loads and rotations exceed the reasonable limits for elastomeric bearings,
HLMR bearings have typically been considered. HLMR situations often occur with
longer spans, with curved or highly skewed bridges, or with complex framing, such as
with straddle bents. In these cases the axis of rotation or the direction of movement,
or both, are either not fixed or may be difficult to determine.
HLMR bearings include pot, disc, and spherical bearings, each of which is unique
in how it accommodates large loads and rotations. All are fabricated in fixed and
expansion versions. The expansion versions accommodate translational movement by
means of PTFE sliding elements. Expansion versions may be guided, allowing move-
ment in only one direction, or nonguided, allowing multidirectional movement. The
following sections describe and compare each HLMR bearing type.
10.2.4.1
Pot Bearings
The pot bearing was first developed in Germany in the early 1960s, and its use began
in the United States in the early 1970s (Fyfe et al. 2006). The main elements of these
bearings include a shallow steel cylinder, or pot, which contains a tight-fitting elasto-
meric disc that is thinner than the depth of the cylinder. A machined steel piston fits
inside the cylinder and bears directly on the elastomeric disc. Brass rings are used to
seal the elastomer between the piston and pot components (see Figure 10.3).
Vertical load is carried through the piston of the bearing and is resisted by com-
pressive stress in the elastomeric pad. The pad is deformable but almost incompressible
in its confined condition and is often idealized as behaving hydrostatically. In practice,
the elastomer has some shear stiffness and so this idealization is not completely jus-
tified. Rotation can occur about any axis and is accommodated by deformation of
the elastomeric pad. Horizontal loads on a pot bearing are resisted by direct contact
between the pot wall and the piston.

450

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 10.3.  Pot bearing components.
Source: Courtesy D.S. Brown.

To achieve satisfactory performance, pot bearings require a high degree of QC in


the fabrication and field installation process and an accurate determination of design
loads and displacements. Through the years, they have been the most economical and
most common HLMR bearing. They have been implemented on bridges throughout
the country.
10.2.4.2
Disc Bearings
The disc bearing was developed and put into service in Canada in 1970 (Fyfe et al.
2006) and was a proprietary, patented device until recent times. It consists of a hard
polyether urethane disc between upper and lower steel plates with a center shear pin
device to resist horizontal load (Figures 10.4 and 10.5). The discs are stiff enough to
support compressive load, yet can deform to permit rotation. However, rotational
stiffness for a disc bearing is several times that of a pot bearing.

451

Chapter 10. BRIDGE BEARINGS


Figure 10.4.  Disc bearing components.
Source: Courtesy R.J. Watson.

Figure 10.5.  Typical disc bearings.


Source: Courtesy R.J. Watson.

Disc bearings are reasonably economical, but widespread use has been limited
because of their originally patented and proprietary status, which made them available
only from a single source. Now that there are additional bearing manufacturers that
can supply disc bearings, their use has increased.

452

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


10.2.4.3
Spherical and Cylindrical Bearings
Spherical bearings are used primarily for accommodating large rotations about mul-
tiple or unknown axes. Sometimes referred to as curved sliding bearings, spherical
bearings permit rotation about any axis; cylindrical bearings permit rotation about
one axis. In these bearing types, rotation is developed by sliding a convex metal surface
(lower element) against a concave PTFE surface (upper element) (see Figure 10.6). The
rotation occurs about the center of the radius of the curved surface, and the maximum
rotation is limited by the geometry and clearances of the bearing. Translational move-
ment is accomplished by incorporating a flat PTFE sliding surface. Horizontal loads
may be partially resisted by the curved geometry of the spherical head; however, large
horizontal loads may require additional external restraint.
Spherical bearings require highly machined fabrication and are more sensitive to
the quality of the initial manufacture and installation than other HLMR bearings.
Although they are generally the most expensive HLMR type, their advantage is their
ability to accommodate higher gravity loads and rotations.
10.2.5
Fabricated Steel Bearings
Fabricated steel mechanical bearings have been used for both fixed and expansion
conditions (see Figure 10.7) and are the longest-used of any other bearing type. Many
existing bridges have these types of bearings, and some states still use them for new
construction. When functioning properly, mechanical steel bearings generally provide
the closest representation of assumed structural end conditions of all bearing types
and transmit loads through direct metal-to-metal contact. Most fixed bearings rely on
a pin or knuckle to allow rotation while restricting translational movement. R­ ockers,
rollers, and sliding types are commonly used expansion bearings. Typically, steel bear-
ings are expensive to fabricate, install, and maintain, which in part accounts for the
popularity of elastomeric bearings. Further, steel bearings typically provide only uni-
directional movement. These types of bearings are fully designed by the engineer to
accommodate loads, movements, and rotations and can be developed to accommodate
large requirements.

Figure 10.6.  Typical spherical bearing.


Source: Courtesy D.S. Brown.

453

Chapter 10. BRIDGE BEARINGS


Figure 10.7.  Fabricated steel bearings with rocker
expansion and fixed conditions.
Source: Courtesy HDR Engineering, Inc.

Bronze lubricated plate bearings have been used in conjunction with steel bearings
to accommodate smaller amounts of movement at expansion ends, but they are not
used much today. PTFE sliding surfaces have replaced bronze sliding plates because of
a much lower coefficient of friction and lower cost.
10.3
Factors Influencing Service Life of Bearings
This section discusses various factors influencing the service life of bearings by using
a fault tree analysis approach that first identifies service life issues that generally per-
tain to all bearing types. This analysis is followed by specific discussions of service life
­issues pertaining to individual bearing types.
10.3.1
Factors Affecting Service Life of All Bearing Types: Fault Tree
Analysis
A general description of the fault tree analysis approach for identifying factors af-
fecting service life is given in Chapter 1 on general framework. Chapter 2, on system
selection, further applies fault tree analysis to identify factors affecting the service life
of the overall bridge system, which includes deck, superstructure, and substructure
components. This section applies specific parts of the fault tree analysis to identify
service life factors that apply to bearings.
Figure 10.8 shows an overall fault tree diagram that identifies factors affecting
service life for bridge bearings. The diagram identifies factors at descending levels that
cause or contribute to service life reduction.

454

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Reduced Service Life of
Bearings

Caused by Deficiency

Of Bridge System– Of Bearing Element–


Related Items Related Items

Due to Natural or Due to Production/


Due to Loads
Human-Caused Hazards Operation Defects

Figure 10.8.  Fault tree analysis for factors affecting service life of bearings.
Figure 10.8. Fault tree analysis for factors affecting service life of bearings.

As discussed in Chapter 2, the first level affecting the service life of a bridge sys-
tem is either obsolescence (relating to function or operation) or deficiency (relating to
deterioration or damage). In relation to bearings, however, service life issues are typi-
cally caused by deterioration or damage, not by obsolescence, so the fault tree moves
directly to deficiency.
Deficiencies can be either system related (i.e., related to other items within the
bridge system or to the layout of the system) or bearing element related (i.e., related
directly to bearing element performance). Deficiencies can then be subcategorized
to those caused by loads, natural or man-made hazards, or production or operation
defects.
10.3.1.1
Deficiency of System-Related Items
System-related items whose deficiencies can directly affect bearings are primarily at-
tributed to system production or operation defects, as illustrated in Figure 10.9.
These factors typically relate to deficiencies in elements, details, and the general
layout of the overall bridge system that can adversely affect bearing performance.
Other types of production and operation defects are discussed in Section 10.3.1.2.3 as
they relate to individual bearing elements.

455

Chapter 10. BRIDGE BEARINGS


Figure 10.9.  Bridge system–related
deficiencies. Deficiency of Bridge
System–Related Items

Due to Production/
Operation Defects

Leaking Inadequate
Deck Joints Inspectability

Improper
Bearing Inadequate
Orientation Replaceability

10.3.1.1.1
Leaking Deck Joints
Of all system-related deficiencies, leaking deck expansion joints can have the greatest
negative impact on bridge bearings. This applies to both open and sealed joints.
Open joints, such as finger dams or sliding plate dams, typically allow drainage
to pass through and be collected in troughs and below-joint drainage systems. Failure
or clogging of these drainage systems allows deck drainage and debris to spill on all
bridge elements below, including bearings.
Sealed joints, such as compression joints, strip seals, or large modular joints, are
intended to prevent deck drainage from spilling through. However, failure or damage
to these types of joints can also allow deck drainage to leak through to the bridge ele-
ments below.
Bearings located below leaking deck joints—particularly those in northern wet
­climates—are subject to deck drainage, deicing chemicals, and other deck debris,
which is a leading cause of deterioration and reduced service life. Drainage and deicing
chemicals cause corrosion of exposed steel elements, and debris buildup affects proper
rotation and expansion movement.
10.3.1.1.2
Improper Bearing Orientation
In skewed, curved, and wide bridges, bearings are subjected to multidirectional move-
ments or rotations, or both. Improper bearing orientation and/or inadequate multi­
directional movement capacity can lead to higher stresses, wear, and reduced service life.

456

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Bridges wider than three lanes can experience significant transverse thermal move-
ment. Guides and keeper assemblies should be limited to the interior portions of the
bridge that do not experience large transverse movements. Bearing details for outer
portions on wide bridges should be designed to accommodate transverse movement.
In the case of skewed steel bridges, a phenomenon commonly referred to as lay-
overs exists during construction. The layovers subject the bearings to rotation of the
steel girder about the longitudinal axis of the girder (twisting of the section). This
action subjects the bearing to a rotation that is not generally considered in design and
subjects the bearing to multirotation.
10.3.1.1.3
Inadequate Inspectability
Proper inspection of bearings during their service life is critical for an accurate evalu-
ation of performance, wear, or deterioration. Early detection of problems can allow
maintenance or repair before more serious conditions can develop. Shallower bearing
types can be difficult to properly inspect, particularly when limited headroom prevents
close access. Consideration should be given in overall bridge system design to allow
access for proper inspection of bearings.
10.3.1.1.4
Inadequate Replaceability
Regardless of expected service life, bearings are subjected to severe service conditions
and have a high potential for unintended consequences related to improper design,
manufacturing, installation, and maintenance that can lead to shorter service lives
than other bridge elements. Consideration should be given in the overall bridge sys-
tem design to allow for easy replacement of bearings with minimal traffic disrup-
tion. A­ ASHTO and the National Steel Bridge Alliance provide recommended bearing
­details that facilitate replacement (AASHTO/NSBA 2004).
10.3.1.2
Deficiency of Bearing-Related Items
Reduced service life of bearings is often caused by deficiencies of individual bearing
elements themselves. As illustrated in Figure 10.8, bearing deficiency can be caused by
loads, natural or man-made hazards, or production and operation defects.
10.3.1.2.1
Bearing Deficiency Due to Loads
Figure 10.10 illustrates load-related factors that affect bearing service life. Loads can
be traffic loads (primarily truck loads) or system-­dependent loads (primarily due to
thermal activity). Each of these load types can result in element damage due to wear
and fatigue or overload.
Truck traffic applies direct vertical load to bearings, as well as superstructure end
rotation and accompanying horizontal translation. Thermal activity applies horizontal
and transverse translations and girder end rotation to bearings; however, depending on
intended or unintended levels of restraint, thermal activity can also apply a significant
horizontal load to bearings.

457

Chapter 10. BRIDGE BEARINGS


Figure 10.10.  Load-related factors
affecting service life. Due to Loads

Traffic-Induced System-
Loads Dependent Loads

Element
Element
Wear/
Overload
Fatigue

Truck traffic produces high-frequency, low-amplitude cyclic movement at expan-


sion bearings combined with vertical load. Cyclic movement from girder end rotation
results in a total cumulative movement that is significantly greater than total cumula-
tive movement due to thermal activity. This behavior primarily affects wear of sliding
surfaces.
Other loads, such as those due to wind, longitudinal braking, or earthquakes, can
also apply various vertical and horizontal loads to bearings that must be resisted.
All of these loads can ultimately result in bearing element fatigue, wear, or over-
load at various levels depending on the specific bearing type and makeup. For exam-
ple, elastomeric bearings are subject to element fatigue, and PTFE sliding surfaces are
subject to wear.
Overload, either from heavy truck loads or large thermal movement, in which the
bearings experience greater loads than assumed in design, can also lead to reduced ser-
vice life. Overload can result in various forms of damage depending on bearing type.
Incorrect assumptions during the design process can also significantly affect the
service life of bearings because of system restraint. For example, as described in Chap-
ter 8 on jointless bridges, in the case of curved girder bridges, there might not be a
point within the structure that can be designated as a point of zero movement. Such
an assumption can lead to the use of a fixed bearing type without any capability or
allowance for transverse or longitudinal movements. The end result of such a wrong
assumption is that the bearing is subjected to actions that cannot be accommodated by
the bearing and the development of significant damage to the bearing in the process.
Service life applications to specific bearing types are discussed later in this chapter.

458

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


10.3.1.2.2
Bearing Deficiency Due to Natural or Man-Made Hazards
Figure 10.11 illustrates factors affecting service life due to natural or man-made haz-
ards. For bearings, hazard-related deficiencies typically pertain to thermal climates,
coastal climates, or chemical environments.
Thermal climates pertain to cold and wet climates accompanied by snow and ice
that result in high usage of deicing chemicals on roadways and bridge decks. Bridge
bearings are affected by roadway drainage and salts leaking through expansion joints,
or by salt spray rising up from crossed roadways.
Coastal climates are climates near the ocean or other saltwater bodies where
bridge bearing elements can be affected by airborne salt spray.
Chemical environments can include environments near chemical or industrial
facilities where corrosive airborne chemicals can affect exposed bearing elements.
Ultimately, these climates or environments can result in exposed steel element cor-
rosion or degradation of other bearing materials at various levels, depending on the
specific bearing type and make up. More specific environmental factors are discussed
as they relate to individual bearing types later in this chapter.

Figure 10.11.  Factors affecting service life


Due to Natural or Man- due to natural or man-made hazards.
Made Hazards

Chemical/
Thermal Coastal
Atmospheric
Climate Climate
Conditions

Steel Bearing
Element Element
Corrosion Degradation

459

Chapter 10. BRIDGE BEARINGS


Due to Production/
Operation Defects

Design/ Fabrication/
Construction Maintenance
Detail Manufacturing

Element
Improper
Flaws Element Debris
Design
Damage Buildup
Parameters

Improper Corrosion
Improper
Design of Steel
Placement
Clearances Elements

Figure 10.12.  Production and operation defects affecting service life.

10.3.1.2.3
Bearing Deficiency Due to Production or Operation Defects
Figure 10.12 illustrates factors affecting the service life of bearings that are related to
production or operation defects. For bearings, these defects can be in any one of four
general subcategories:

• Design and detail;


• Fabrication and manufacturing;
• Construction; or
• Maintenance.

10.3.1.2.3a
Design and Detail
Improper bearing design and accompanying details can result in significantly reduced
service life. Major base factors include improper design parameters and improper
­design clearances.
Incorrect or improperly computed design values can affect vertical load, move-
ment, and rotation, or combinations of these. All bearings must be designed for proper
superstructure loads, movements, and rotations. Improper calculation and application
460

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


of these parameters at the design stage can result in bearings that are subject to exces-
sive rotation, higher stresses, greater wear, and, ultimately, reduced service life.
Clearances within bearing details should permit proper movement and/or rota-
tion. Details with inadequate clearances can cause binding that limits proper move-
ment or rotation, which results in higher stress, damage, wear, and reduced service life.
10.3.1.2.3b
Fabrication and Manufacturing
Material or fabrication flaws can lead to performance failure and reduced service
life. Proper quality assurance (QA) and QC procedures in accordance with the cur-
rent LRFD Bridge Construction Specifications (LRFD construction specifications)
(­AASHTO 2010a) must be implemented in the fabrication or manufacturing of bridge
bearings to ensure that the completed bearings meet specifications and provide re-
quired levels of performance. See Section 10.4.1.2.3b for additional discussion.
10.3.1.2.3c
Construction
Production and operation defects at the construction stage can be caused by either
element damage or improper placement. Protective care must be taken during field
construction to prevent damage or contamination to sensitive bearing parts such as
sliding surfaces or elastomeric pads or discs. Bearings must be set in the field at proper
positions to accommodate installation temperatures and rotations.
10.3.1.2.3d
Maintenance
Lack of or inadequate bearing maintenance can also lead to reduced service life. Issues
typically relate to lack of cleaning, which results in debris buildup below deck expan-
sion joints, and steel element corrosion. Other types of maintenance-related issues will
be discussed in the following subsections for specific bearing types. Dirt and debris
buildup prevents proper bearing movement and rotation. Debris buildup also retains
and holds moisture and salt against exposed steel elements, which leads to corrosion
if the elements are not cleaned.
10.3.2
Factors Affecting Service Life Unique to Each Bearing Type
This section looks at specific service life issues pertaining to the bearing types described
in Section 10.2 and addressed within the general fault tree categories described in Sec-
tion 10.3.1.
10.3.2.1
Steel-Reinforced Elastomeric Bearings
Steel-reinforced elastomeric (SRE) bearings have been in service in the United States
for over 50 years, and longer elsewhere in the world, with very good results. They
are typically very robust, and of all bearing types they likely have the best chance for
achieving a service life greater than 100 years. When properly designed, manufactured,
and installed, there is very little that can go wrong, and long-term maintenance require-
ments are minimal. In rare instances, certain problems have been observed, most often
associated with production or operation defects relating to design and manufacturing.
461

Chapter 10. BRIDGE BEARINGS


10.3.2.1.1
Improper Design
In rare occurrences, improper design has led to overloaded pads or pads subjected to
excessive lateral movement, causing excessive bulging, splitting, or delamination. The
following paragraphs describe various potential failure modes and other issues with
elastomeric bearings and how they are addressed within current AASHTO design pro-
visions. A significant amount of research has been performed on SRE bearings, and
designs that follow recent provisions in the LRFD specifications should adequately
avoid these issues as they would affect service life.
10.3.2.1.1a
Shear Deformation
Elastomeric bearings accommodate longitudinal and transverse expansion and con-
traction by shear deformation within the elastomer itself (see Figure 10.13). If the
shear displacements of the bearing are large, they may cause some rollover at the acute
ends of the layer, which leads to cracking in the elastomer at the end of the top and
bottom reinforcement plates. This condition is exacerbated by cyclic loading. Fatigue
tests simulating temperature movement, which is low-cycle, high-amplitude move-
ment, indicate that keeping the shear strain below 0.5 will prevent this cracking. This
is not conservative, however, if the deformation is due to high-cycle loading caused by
braking forces or end rotation. It was found that high-cycle fatigue was more damag-
ing, and in those cases the maximum shear strain should be limited to 0.10 (Roeder
et al. 1990).
10.3.2.1.1b
Plate Delamination and Debonding
Shear delamination between layers of elastomer and steel reinforcing plates is the most
significant potential mode of failure. Large shear strains occur in elastomeric bearings
as a result of combined axial load, rotation, and shear, as illustrated in Figure 10.14
(Stanton et al. 2008). Each of these shear strains reaches its maximum value at the
same location, namely the very edge of the layer at which the elastomer is bonded to
the reinforcing plate. Under severe loading, this condition leads to local detachment
and debonding of the elastomer from the plate. Once debonding occurs, the elastomer
starts to extrude from the bearing, which in turn causes significant vertical deflec-
tion. Further, cyclic shear strain causes additional debonding damage than static shear
strain of the same magnitude. Section 14.7.5.3.3 of the LRFD specifications provides

Figure 10.13.  Shear deformation in


elastomeric pad.
Source: Stanton et al. 2008.

462

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 10.14.  Deformations of a laminated elastomeric bearing layer.
Source: Stanton et al. 2008.

updated (2010b) requirements for considering shear strain caused by combined axial
load, rotation, and shear displacement and considers an amplification factor of 1.75
for combined shear strains due to cyclic loading caused by traffic.
10.3.2.1.1c
Instability
Large lateral movement combined with large end rotations result in thick bearing
designs. If the bearing becomes thick enough, instability can affect its performance
­(Stanton et al. 2008). The layered construction and the very low shear modulus of
the rubber combine to cause this potential problem. Quite often the bearing is made
as wide as possible (transverse to the girder axis), and then only a short length is
needed to provide sufficient bearing area for supporting the axial load. However,
too short a length would again risk instability. In such bearings, the axial stress may
therefore be significantly lower than the limit because of the indirect influence of sta-
bility require­ments. Section 14.7.5.3.4 of the LRFD specifications provides updated
(2010) ­requirements for stability and limits the average compression stress to half the
predicted buckling stress.
10.3.2.1.1d
Plate Fracture
Lateral expansion of the elastomeric layers causes tension in the steel reinforcing plates
(Stanton et al. 2008). At extreme loads the plates could fracture, typically splitting
along the longitudinal axis of the bearing. However, in plates of the thickness cur-
rently used, this behavior does not occur until the load has reached 5 to 10 times its
design value, so plate fracture seldom controls design. Plates could actually be thinner,
but they are typically sized in practice to keep them from bending during the molding
463

Chapter 10. BRIDGE BEARINGS


process. Section 14.7.5.3.5 of the LRFD specifications provides requirements for the
thickness of steel reinforcement and considers both strength and fatigue.
10.3.2.1.1e
Compressive Deflection
Instantaneous compressive deflection is limited by the LRFD specifications to avoid
damage to deck joints and seals and to avoid additional impact when traffic passes
from one girder to the other across a joint. A maximum relative live load deflection
across a joint of 1/8 in. is recommended. Section 14.7.5.3.6 of the LRFD specifications
provides requirements for compressive deflection.
10.3.2.1.2
Improper Fabrication, Manufacturing, and Installation
In addition to proper design, it is imperative that proper manufacturing and installa-
tion also be performed, and effective QC during these processes is essential for suc-
cessful performance.
One of the most commonly reported field problems associated with elastomeric
bearings, albeit in only a few instances, has been walking out, or slipping from the
original position under the girder. Slippage has occurred primarily in natural rubber
elastomeric pads that had paraffin added during the manufacturing process in order to
meet ozone degradation requirements established by AASHTO. Neoprene pads have
inherently greater ozone resistance and do not need wax additives. Over time, these
waxes will bleed to the bearing surfaces and drastically reduce the coefficient of fric-
tion between the bearing and its contact surface, which leads to slipping (Chen and
Yura 1995). The use of positive attachment to the substructure or superstructure, such
as bonding, is recommended to prevent this occurrence. However, caution needs to
be exercised in bonding elastomeric pads and contact surfaces in highly skewed steel
bridges because of layovers.
10.3.2.2
Cotton Duck Pads
CDPs are used by only a few states so their performance history is limited. Like
SRE bearings, potential service life issues for CDPs are most often associated with
production and operation defects, and CDPs require proper design, manufacturing,
­installation, and maintenance.
For SRE bearings, longitudinal and transverse bridge movements are accommo-
dated by shear deformation of the elastomer, and the reinforcement has little influ-
ence on the shear stiffness of the bearing. CDPs behave differently. The fabric layers
of CDPs are many and closely spaced and result in significantly larger shear stiffness
and smaller shear deformation capacity than in other elastomeric bearing types. As a
consequence, CDPs tolerate large compressive stresses but limited shear deformation,
because interlayer splitting occurs at much smaller shear strains. Also, the large shear
stiffness of CDPs can result in slip of the girder on the CDP, which may result in abra-
sion and deterioration of the CDP. As a consequence, translational shear strain is lim-
ited to only about 10%. That is, the shear deflection is limited to only 1/10 of the total

464

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


CDP thickness. Larger movement requirements with CDPs must be accommodated by
the addition of a sliding surface such as a PTFE sliding surface (Lehman et al. 2003).
The stiffness and deformation capacity of CDP varies from manufacturer to manu-
facturer. Proper QC testing is necessary to assure that the bearing pad provides ade-
quate performance.
Delamination of elastomer layers or secretion of oil and wax from the CDP are
the common serviceability limit states for CDPs. To control delamination, compressive
stress limits of 3,000 psi for total dead plus live load and 2,000 psi for live load are
recommended. Dynamic or cyclic rotation, which induces uplift or partial separation
of the pad from the load surface, may cause delamination and reduced service life.
Uplift damage depends on the maximum total rotation, as well as the rotation range
caused by the live load variation, and separate rotation limits are provided. The LRFD
specifications now include design provisions to ensure the serviceability and durability
of CDPs.
10.3.2.3
Primary Sliding Surfaces
Factors affecting the service life of bearing sliding surfaces (primary sliding surfaces,
not guide bars) most often relate to deficiencies caused by loads, including both traf-
fic load from trucks and system-induced (thermal movement) loads. These loads can
result in sliding element wear (from cyclic movement) and creep or cold flow (from
compressive overload).
Deficiencies caused by production or operation defects can also occur, primarily
during manufacturing and construction. They typically relate to surface scratching and
damage due to inadequate protection during shipping or installation. Surface damage
leads to increased wear. These problems can be mitigated by proper protection and
inspection during shipping and installation.
Inadequate maintenance, particularly below open or leaking deck expansion
joints, can allow buildup of dirt and other debris that can cause contamination of slid-
ing surfaces and increased wear. Periodic maintenance and cleaning in these areas can
mitigate these problems.
10.3.2.3.1
Wear of Sliding Surfaces
Plain PTFE is most commonly used for bearing sliding surfaces. It wears under service
conditions and may require replacement after a period of time. Low temperatures, fast
sliding speeds, high contact pressures, rough mating surfaces, and contamination of
the sliding interface increase the wear rate. However, fast sliding speed has been shown
to be the more dominant parameter (Stanton et al. 1999). Movement due to tempera-
ture change is low-cycle, high-amplitude movement, with a slow movement rate, and
produces the least amount of wear. However, movement due to truck load and asso­
ciated dynamic effects is high-cycle, low-amplitude movement; its sliding speed can
be much faster, by as much as a factor of 10. Wear rates associated with high sliding
speeds can be as much as 150+ times greater than wear rates at lower sliding speeds.

465

Chapter 10. BRIDGE BEARINGS


Thus, plain PTFE should not be used as a sliding surface for bearings subject to rela-
tively high sliding speeds and low temperatures.
Relatively thin layers of PTFE, from 1/16 to 3/16 in., are commonly used in the
United States, but engineers in other countries often use thicker PTFE layers to accom-
modate wear.
Woven or glass-filled PTFE surfaces provide much higher overall wear resistance,
especially at higher sliding speeds, but these surfaces have higher friction coefficients
that must be taken into account in the bridge system design.
Dimpled and lubricated PTFE also provides exceptional wear resistance and low
friction, but the long-term effectiveness of lubrication is questionable (Stanton et al.
1999).
MSM has been shown to provide exceptional wear resistance, but when used in
a dry condition (without lubrication), it has a much higher friction coefficient than
plain PTFE. When used in a dimpled and lubricated condition, however, its friction
­coefficient reduces considerably and is more comparable to lubricated PTFE.
10.3.2.3.2
Creep or Cold Flow
PTFE may creep (or cold flow) laterally when subjected to high compressive stress and
shorten the life of the bearing. The reduction in PTFE thickness may also allow hard
contact between metal components. Thus, although the compressive stress should be
high to reduce friction, it must also be limited to control creep. PTFE is frequently
recessed for one-half its thickness and bonded to control creep and permit larger com-
pressive stress.
Filled PTFE, which is reinforced with fiberglass or carbon fibers, has significantly
greater resistance to creep and is sometimes used to resist creep or cold flow.
10.3.2.4
High-Load Multirotation Pot Bearings
In past years, pot bearings have had service life problems most often associated with
production or operation defects relating to design and manufacturing. These issues
have ultimately resulted in broken internal seals, leakage or extrusion of the e­ lastomer,
abraded elastomeric pads, and metal-to-metal contact, which have led some state
depart­ments of transportation to recommend avoiding pot bearings altogether. How-
ever, improved design specifications and tighter manufacturing tolerances developed
in the late 1990s have greatly improved overall performance. Current LRFD specifica-
tions (AASHTO 2012) now incorporate research findings reported in NCHRP Report
432: High-Load Multi-Rotational Bridge Bearings (Stanton et al. 1999) and address
specific past service life issues. The following items have been recommended to miti-
gate certain past deficiencies and are now part of standard design and manufacturing
practice:

• The rotational capacity of pot bearings is limited by the clearance between various
elements of the pot, piston, sliding surface, guides, and restraints (Stanton et al.
1999). Inadequate clearances can cause binding between metal components (see
Figure 10.15). Clearance requirements are now included in the specifications.

466

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure 10.15.  Critical clearances affecting rotational capacity.
Source: Courtesy D.S. Brown.

• The thickness of the elastomeric pad also affects rotational capacity. It is currently
controlled by a 15% strain limit on the pad-edge deflection under rotation. This
strain value is based more on past practice than research results, but it is believed
to be a reasonable value.
• Pot bearings are able to sustain many cycles of small rotation better than a smaller
number of large rotation cycles. This is believed to be true because smaller ro-
tations cause deformation of the elastomer but little slip. Slip caused by larger
­rotations abrades the surface of the elastomer. In an attempt to accommodate
this behavior, pot bearings should be designed for larger minimum rotations that
reduce the potential for overrotation. This minimum rotation should reflect in-
creased rotation caused by construction tolerances expected in practice. Further,
greater emphasis should be placed on calculation of rotations caused by service
loads, construction loads, and environmental conditions.
• Rotational resistance, wear, and abrasion are significantly reduced with a smooth
surface finish inside the pot and on the piston. Metalizing of these interior surfaces
for corrosion protection causes a rougher surface that leads to increased dam-
age under cyclic rotation; such metalized surfaces should not be used unless they
are buffed to a smooth finish. In extremely corrosive environments, stainless steel
could be considered.
• The failure of sealing rings causing escape of elastomer has been one of the major
service life issues with pot bearings in past years. Solid circular cross-section brass
rings and multiple flat brass rings have both been used, and each has advantages
and limitations. Circular cross-section rings provide a tight seal, but they are sus-
ceptible to wear under cyclic rotation. Flat rings appear to be more susceptible to
leakage and ring fracture, but they experience less wear. Heavier flat brass rings
have been suggested as a means of improving their performance. The performance
of circular rings could also be improved if the friction and wear were reduced.

467

Chapter 10. BRIDGE BEARINGS


Currently, both circular cross-section and flat brass rings are permitted in the
specifications. Current design and manufacturing in accordance with the LRFD
specifications have greatly resolved past issues with sealing rings and leakage of
elastomer.
• Silicon grease lubrication appears to reduce the wear noted on rings, pot walls,
and pistons and is recommended for use. This lubrication does not reduce the
ultimate wear of the elastomeric disc.
• A relatively small lateral load (5% of gravity load) combined with cyclic rota-
tion can dramatically increase the rotational resistance, wear, and leakage of pot
bearings. This damage is caused by the piston rim dragging against the pot wall
­during rotation. Alternative methods of external restraint are recommended for
large ­lateral loads to mitigate this damage potential.
• Dirt or contamination in the pot increases wear and abrasion of the rings, pot,
and elastomeric disc and increases the potential of elastomer leakage. To mitigate
this potential, pot bearings need to be sealed and protected during shipping and
installation.

Although newer design and manufacturing criteria have improved the overall per-
formance of pot bearings in recent years, it is still recognized that this bearing type has
internal moving parts subject to wear and abrasion that can lead to reduced service life.
Newer requirements for long-term deterioration testing in accordance with the current
LRFD construction specifications provide greater assurance of proper performance.
10.3.2.5
High-Load Multirotation Disc Bearings
Since their first use in the early 1970s, HLMR bearings have had good performance
and few reported field problems. Potential (albeit few) service life problems would
most likely be associated with production or operation defects relating to design and
manufacturing. Current specifications are minimal regarding design and are generally
performance related. Therefore, performance testing and manufacturing QC measures
are necessary to verify compliance.
Previous research testing of disc bearings for combined load and rotation (Stanton
et al. 1999) made several performance conclusions:

• Tests showed that rotation of disc bearings is partly accommodated by uplift of


the steel plates from the urethane disc, especially if the compressive load is light.
This uplift should not result in any problems with fixed bearings, but it could be
a concern with sliding bearings, because uplift of the disc produces edge loading
on the PTFE sliding surface. To mitigate this potential problem, LRFD specifica-
tions limit the edge contact stress on PTFE surfaces, and the LRFD construction
specifications require proof load testing.
• Tests showed the urethane disc to be somewhat deformed and abraded by cyclic
rotations, but the damage was not severe enough to affect the performance of the
bearing. LRFD construction specifications require cyclic deterioration testing to
confirm long-term performance.
468

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


10.3.2.6
High-Load Multirotation Spherical Bearings
Although a very robust system, HLMR spherical bearings have had service life prob-
lems, most often associated with production or operation defects relating to design
and manufacturing. The difficulties appear to be attributable to faulty fabrication by
manufacturers that were not necessarily first-tier suppliers. These bearing types re-
quire very precise manufacturing tolerances to assure proper fit of the curved mating
surfaces.
Rotational capacity can be set at almost any desired level provided adequate clear-
ances are provided.
PTFE surfaces may also eventually wear out. Variations in friction with different
types of PTFE and under different temperature and load conditions cause variations in
behavior that can lead to performance issues. Woven PTFE has often been used with
spherical bearings in the United States; dimpled and lubricated PTFE is often used in
Canada and Europe.
These types of bearings are typically larger than other types and generally require
additional space. Spherical bearings are traditionally considered to be the most expen-
sive HLMR bearing type, but they are also traditionally considered the most reliable.
10.3.2.7
Fabricated Mechanical Steel Bearings
Fabricated mechanical steel bearings have been used for the longest time of any bear-
ing type and have the potential for extended service life if properly protected and
maintained. Factors affecting service life relate to several categories, including loads,
primarily overload, which results in binding or overrotation of rocker bearings; natu-
ral or man-made hazards, which result in steel element corrosion; and production or
operation defects, specifically due to lack of maintenance.
10.3.2.7.1
Loads
Rocker expansion bearings can be designed for wide ranges of movement, but they can
have limited potential for accommodating overload. Excessive translation can lead to
rockers tipping over, which has been reported in a few instances.
10.3.2.7.2
Natural or Man-Made Hazards
Corrosion of steel bearings, particularly those located below open or leaking deck
joints in thermal environments, is the highest cause of reduced service life for these
bearing types. In these locations, steel bearings are highly susceptible to corrosion
from roadway drainage with deicing salts and other dirt accumulation. Coastal cli-
mates and other chemical environments are also catalysts for steel element corrosion.
Pin, roller, and rocker bearings have direct metal-to-metal contact, which creates
an environment of high stress concentration and accumulation of moisture between
surfaces. In the areas of contact, any protective coating on the steel is inevitably dam-
aged by the relative movement. All of these conditions contribute to corrosion.

469

Chapter 10. BRIDGE BEARINGS


Corroded expansion bearings can lock up or freeze, subjecting beams and the
substructure elements below to additional load and potential damage. Fyfe et al.
(2006) reported that steel rocker and roller bearings, along with older type metal slid-
ing plates, are the most susceptible to freezing in position. When bearings are frozen,
bridges develop their own provisions for contraction and expansion, such as pier cap
cracking or rocking of piers or abutments.
Dirt and debris accumulation has caused some rocker-type expansion bearings to
move beyond their limits and actually roll over, causing the superstructure to drop sev-
eral inches or more. Figure 10.16 illustrates a ratcheting effect reported by Modjeski
and Masters (2008) in which rocker bearings at one pier on the Birmingham Bridge in
Pittsburgh, Pennsylvania, tipped over because of corrosion and debris accumulation.
The rockers had an initial lean, and a leaking expansion joint above caused corro-
sion and increasing accumulation of debris on the bearings and under one side of the
­rockers. A ratcheting effect followed in which additional debris and corrosion material
kept accumulating under the rockers, causing additional and increasing tilt. This pro-
cess continued until a lateral force or kick developed against the pier cap that caused
the pier to move just enough to initiate the final tipping of the rocker bearings.
Steel rocker bearings have also performed poorly in seismic events and have been
replaced as part of seismic retrofit in many instances.

Figure 10.16.  Ratcheting effect caused by debris accumulation and corrosion.


Source: Modjeski and Masters 2008.

470

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


10.3.2.7.3
Production and Operation Defects
Lack of maintenance has been a contributing factor to reduced service life for steel
bearings in many instances. Coating loss, surface corrosion, and debris buildup, all
of which lead to reduced service life, can be effectively mitigated by periodic mainte-
nance. In locations below open deck expansion joints, particularly in thermal environ-
ments in which bearings are highly susceptible to roadway drainage, salt, and debris,
periodic cleaning to prevent buildup of deleterious materials is essential.
10.4
Options for Enhancing Service Life of Bearings
This section presents available options for mitigating the various bearing service life
­issues that were identified in Section 10.3. A procedure to select the optimal bear-
ing type for given loads, movements, and environmental conditions is presented
in Section 10.5.
The options for enhancing the service life of bearings address the issues discussed
in the fault tree analysis and the specific categories and subcategories of deficiencies
that were shown to cause damage or deterioration. Bearing deficiencies can be either
related to bridge system deficiencies or related directly to bearing element deficiencies.
Deficiencies can then be attributed to those caused by loads, natural or man-made
hazards, or production and operation defects.
In the discussion of available options for enhancing the service life of individual
bearing types, the following general solution categories are included and are addressed
as applicable:

• Avoidance eliminates or bypasses a particular service life issue. This is a primary


consideration and should be incorporated when possible.
• Mitigation improves performance through enhanced materials, greater protection,
or effective maintenance. When avoidance is not practicable, mitigation techniques
should be implemented to improve service life.
• Acceptance considers that a particular bearing or bearing component may not be
able to achieve a service life equal to the bridge system design service life, even
after mitigation to improve its service life, and may eventually need to be replaced.

In many instances, the need to provide capabilities and details for easy bearing
replacement may still be a necessary consideration because of uncertainties regarding
loads, hazards, or production and operation defects that can cause premature service
life issues.
10.4.1
General Options for All Bearing Types
10.4.1.1
Solutions for Deficiencies of System-Related Items
Table 10.2 summarizes solutions for various service life issues identified in Sec-
tion 10.3.1.1 and briefly identifies what each solution provides and what other con-
siderations may still be needed. System-related issues are primarily due to production
471

Chapter 10. BRIDGE BEARINGS


Table 10.2.  Solutions for Service Life Problems Caused by Bridge System–Related Deficiencies:
All Bearing Types
Service Life Problem Solution Advantages Disadvantages
Leaking deck joints Avoid by using integral or Fully integral abutment Requires an understanding of
(Bridge systems with semi-integral construction at eliminates joints and integral abutment behavior
deck joints can have abutments. bearings. and limitations.
deterioration of bridge Semi-integral eliminates See Chapter 8 on jointless
elements below deck joints. bridges for detail design
caused by drainage Prevents deterioration due provisions.
through leaking deck to drainage and chlorides to
joints carrying deicing elements below deck.
salts and other debris.) Avoid by using integral Eliminates deck joints, Requires system design
construction between bearings, and the need for approach to accommodate
superstructure and future bearing maintenance. bridge movements.
substructure at piers. At piers, can also avoid sharp Some systems require
skews, or longer spans, or posttensioning for integral
higher profiles. cap design and construction.
May be more costly than
isolated construction with
bearings.
Avoid by using continuous Eliminates deck joints and None.
superstructure over piers. protects bearings below.
Mitigate by protecting Coatings on steel surfaces Requires continual
bearings with coatings and protect against corrosion. maintenance.
maintenance. Maintenance improves
coating life and prevents
debris buildup.
Mitigate by repair or Cost-effective solution for Requires continual
replacing leaking joints. existing conditions. maintenance.
Misorientation of Align bearings along girder Proper orientation of guided Requires understanding of
expansion bearings chords for curved bridges, expansion bearings prevents bridge system behavior.
on complex in keeping with current wear and binding against
alignments recommendations. restraint mechanisms.
Align bearings along long
diagonal line on deck for
skewed bridges.
See Chapter 8 on jointless
bridges for additional
detail discussions and
recommendations.
Limited access for Provide details and Early detection of problems Possible additional initial
inspection clearances that facilitate avoids major deterioration or cost.
bearing inspection and distress.
maintenance. Reduces risk.
Difficulty in Provide details that facilitate Minimizes impact to traveling Additional initial cost is
replacing jacking and replacement public during rehabilitation. higher but saves greater
deteriorated under traffic. Facilitates speed of future cost when replacing
bearings replacement. bearings.

472

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


or operational defects and, more specifically, due to design and details of components
within the overall bridge system. Further discussion relating to each issue follows this
table.
10.4.1.1.1
Leaking Deck Joints
The following are possible solutions for avoiding or mitigating bearing damage or
deterioration caused by leaking deck joints. In considering options for improving ser-
vice life, avoiding deck expansion joints by using continuous systems and/or integral
systems offers the best opportunity. When possible, integral systems that eliminate
bearings entirely should also be considered, but in most cases this will not be feasible
or practical, and the use of bearings cannot be avoided.

• Integral abutment systems that avoid deck joints and bearings. Refer to Chapter 8
on jointless bridges for general design guidelines. This has become a popular ap-
proach by many states for improving bridge service life. Fully integral abutment
construction eliminates both deck joints and bearings at abutments and is also
lower in initial cost. Semi-integral construction eliminates deck joints.
• Proper use of integral abutments requires an understanding of pile and cap be-
havior and limitations regarding bridge length and geometrics such as curvature
and skew.
• Integral pier systems that avoid deck joints and bearings. See Chapter 2 on system
selection for a more detailed discussion of integral pier options.
• Bridge systems using integral girder–pier cap construction have been used occa-
sionally by some states to accommodate vertical clearance issues, to avoid sharp
skews, or to develop frame action for seismic design. But they can also serve to
eliminate joints, bearings, and associated future maintenance. Although there is a
higher initial structure cost, savings are realized with integral pier caps in lower
approach fills and lower long-term maintenance.
• Cast-in-place, posttensioned integral bent caps have been used for many years,
but they are not widely employed. The Tennessee Department of Transportation
constructed its first application in 1978, with several others since, and all have per-
formed well. The concept allows main longitudinal girders to pass directly through
the pier’s cap rather than over the top in the traditional manner.
• Integral caps can also maximize column efficiency. Frame action in the longitudinal
direction reduces column design moments at column bases compared with conven-
tional cantilever columns and can also enhance seismic performance. However, it
must also be considered that integral pier columns may need to resist additional
longitudinal moments due to live load, and thus a system analysis is required.
• When multiple piers are made integral with the superstructure, expansion must
be accommodated by flexure of the pier columns. Tall, slender columns are best
suited for this type of construction because their greater flexibility can accommo-
date temperature movement with less force developed.

473

Chapter 10. BRIDGE BEARINGS


• Continuous superstructure systems that avoid or minimize the number of deck
joints include fully continuous, continuous for live load, or continuous deck slabs
(link slabs). These continuous systems can be combined with either integral abut-
ments or conventional abutments. Bearings are still used at piers, but they are
protected by continuous superstructure.
• Mitigate for bearings below deck joints. When bearings must be located below
deck expansion joints, mitigation procedures should be implemented. Depending
on the environmental severity, all exposed steel bearing parts, which include sole
plates, masonry plates, guides, and anchor bolts, should be stainless steel, galva-
nized, or metalized. These areas should also be cleaned periodically as part of a
regular maintenance program to prevent salt and debris buildup.
• Mitigate by repairing or replacing leaking joints. In existing bridges with deck
joints, a cost-effective strategy is to repair or replace joints when they start to leak,
which would prevent any deterioration below the joint. This proactive approach
to preventative maintenance requires periodic inspection with immediate repair or
replacement when necessary.

10.4.1.1.2
Improper Expansion-Bearing Orientation
Proper orientation of bearings in curved and skewed bridges, which are subjected
to multidirectional movements and/or rotations, should be provided. Improper bear-
ing orientation and/or multidirectional movement demand can lead to higher stresses,
damage, wear, and reduced service life.

• Curved girder bridges do not expand and contract along girder lines. A typical
approach is to assume movement to occur along chord lines from the fixed point
to expansion points (see Figure 10.17). See Chapter 8 on jointless bridges for ad-
ditional discussion on determining the point of zero movement or fixed point in
integral abutment bridges.

Figure 10.17.  Recommended bearing orientation on a horizontally curved alignment.


Source: Courtesy National Steel Bridge Alliance.

474

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


• For large skew bridges, one approximate solution is to consider the major axis
of thermal movement along the diagonal from the acute deck corners caused by
thermal movement of the bridge deck. Bearing alignment should be parallel to that
axis (see Figure 10.18). It is necessary for expansion bearings to have multidirec-
tional capabilities.
• Bridges wider than three lanes can experience significant transverse thermal move-
ment. Guides and keeper assemblies should be limited to the interior portions of
the bridge that do not experience large transverse movements. Bearing details for
outer portions on wide bridges should be designed to accommodate transverse
movement combined with longitudinal movement.

10.4.1.1.3
Proper Access for Bearing Inspection
Details, particularly sufficient room and access at the tops of piers and abutments to
permit proper bearing inspection, should be provided. Early detection of bearing-re-
lated issues can allow maintenance or repair before major service life–reducing events
can develop. When possible, shallow bearings should be placed on pedestals at the
tops of piers and abutments in order to provide greater clearance for inspection. Use
of thicker sole plates can also help provide greater clearance.
10.4.1.1.4
Proper Capability for Bearing Replacement
Details that will accommodate the potential need for replacement of all or part of the
bearing system should be provided with the following considerations:

• Jacking locations should be provided at every girder. An alternative is to provide


for jacking under a diaphragm that lifts adjacent girders simultaneously.
• Bearing attachment details should allow ease of replacement. If the bearing is un-
attached, it can easily be pulled from its position when the load is removed. Welds

Figure 10.18.  Recommended bearing orientation for large skew bridges.


Source: Courtesy National Steel Bridge Alliance.

475

Chapter 10. BRIDGE BEARINGS


can be cut, but doing so requires equipment that may be cumbersome in the space
available. Grinding may also be needed to produce a surface flat enough for in-
stalling the new bearing. Any anchor bolts should be placed so that they do not
impede the removal of the bearing.
• Bearings should be detailed so they can be replaced with only ¼ in. of jacking to
avoid causing a bump at the top of the deck that would affect traffic. The details
also need to provide adequate vertical clearance for jacking. Jacking points on
the structure should be designed to accommodate both dead load and live load
in o
­ rder to be able to maintain traffic during bearing replacement. (Actual jack-
ing loads may have to include a factor applied to the design load to break loose
the component before lifting.) Jacking for elastomeric bearings needs to consider
the effect of compressive dead-load strain in determining the amount of lift.

10.4.1.2
Solutions for Bearing-Related Deficiencies
10.4.1.2.1
Deficiencies Due to Traffic- and System-Dependent Loads
Service life issues can be avoided by determining truck traffic–induced cyclic move-
ment for consideration in particular bearing type designs or by determining proper
levels and combinations of load, movement, and rotation.
10.4.1.2.2
Deficiencies Due to Natural or Man-Made Hazards
The potential for steel element corrosion should be avoided by using stainless steel or
mitigated by high-performance protective coatings (galvanizing or metalizing).
10.4.1.2.3
Deficiencies Due to Production or Operation Defects

10.4.1.2.3a
Design- and Detail-Related Deficiencies
Design should take into account proper levels and possible combinations of load,
movement, and rotation in order to avoid problems due to excessive stress, transla-
tion, or overrotation. A considerable number of bearing failures are attributed to im-
proper allowance for displacements.
Construction tolerances and possible construction loadings and how they might
affect bearing loads and movements should also be considered. It is recommended that
the worst possible combination of loads and displacements due to construction toler-
ances be considered. In other words, design displacements should include the largest
sum of displacements caused by the worst out-of-level placement of piers and abut-
ments, the largest camber and deflection, and the most adverse tolerances permissible
in the construction of the bridge and bearing.
In Steel Bridge Bearing Design and Detailing Guidelines (AASHTO/NSBA 2004),
Appendix A provides recommendations for calculating beam rotations for dead load
and live load conditions. There is great variation in the methods used in the industry

476

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


for calculating live load rotations, and the Guide was developed according to the
methods used in several states. A realistic approach for computing beam rotations is
presented in the NSBA guidelines.
Adequate clearances should be provided for horizontal movement and rotation
to prevent binding, wear, or damage to restraining devices, anchor bolts, or internal
bearing elements. Adequate widths and clearances at pier and abutment bridge seats
should be provided to allow for required movement and rotation.
10.4.1.2.3b
Fabrication- and Manufacturing-Related Deficiencies
Proper QA-QC procedures in the fabrication or manufacturing of bridge bearings are
essential to assure that the completed bearings meet specifications and provide the
required levels of performance. The current LRFD construction specifications provide
minimum requirements for packaging, handling, and storage; fabrication tolerances;
materials; and general performance. Testing requirements include material certification
tests, material friction tests, dimension checks, clearance tests, bearing friction tests,
long-term deterioration tests, proof load tests, and horizontal force capacity tests. The
specifications require that manufacturers provide certification that each bearing satis-
fies the requirements of the contract drawings and construction specifications.
10.4.1.2.3c
Construction-Related Deficiencies
Care and protection during field construction should be provided to prevent damage
to protective coatings or damage to or contamination of sensitive bearing parts, such
as sliding surfaces or elastomeric pads or discs.
Proper QA-QC procedures should be provided during construction to assure that
bearings are initially set with specified position, clearances, and rotation.
10.4.1.2.3d
Maintenance-Related Deficiencies
Periodic maintenance cleaning should be provided depending on the location within
the bridge system and environmental hazard conditions to avoid debris buildup that
could affect movement and rotation performance.
Steel surface cleaning and coating maintenance should be provided to prevent cor-
rosion of exposed steel elements.
10.4.2
Options Related to Specific Bearing Types
10.4.2.1
Elastomeric Bearings: Plain and Steel Reinforced
Table 10.3 summarizes solutions for various service life issues identified in Sec-
tion 10.3.2.1 for elastomeric bearings and briefly identifies the advantages and dis­
advantages of each solution. Problems with elastomeric bearings, albeit rare, have
been asso­ciated with production or operational defects relating to design, manufactur-
ing, and installation. Following this table is further discussion relating to each issue.

477

Chapter 10. BRIDGE BEARINGS


Table 10.3.  Solutions for Service Life Problems: Plain and Steel-Reinforced
Elastomeric Bearings
Service Life Problem Solution Advantages Disadvantages
Improper element Avoid by following current Very robust bearing with None.
design resulting AASHTO LRFD design criteria minimal problems and
in excessive shear for Methods A or B. need for maintenance
deformation and when properly designed,
excessive bulging, manufactured, and installed.
splitting, or AASHTO design criteria have
delamination incorporated recent research
to avoid failure modes,
including combined axial
load, shear, and rotation.
Need for greater Combine with high- Can provide unlimited Additional initial cost.
longitudinal performing sliding surfaces. movement capability and
movement capacity longer service life.
Need to Use circular bearing pads. Accommodates May require wider bridge
accommodate multidirectional movement seat.
multidirectional and rotation. Not as efficient as rectangular
complex movement Well suited for cases in which pad with proper orientation.
(skewed or curved transverse movements are
bridges) uncertain.
Slipping or walking Control the use of wax Prevents reduction in Less resistance to ozone
out additives in elastomer. coefficient of friction degradation, but should not
(This has been more of between pad and bearing be an issue with neoprene.
a problem with plain surface.
pads.) Provide positive attachment Restrains pad from walking Small additional initial cost.
to sole plates. out.
Improper Provide effective QA-QC. Assures adherence to None.
fabrication or quality and performance
manufacturing requirements.

10.4.2.1.1
Improper Element Design Resulting in Excessive Shear Deformation and Excessive
Bulging, Splitting, or Delamination
Excessive bulging, splitting, or delamination can be avoided by following recent pro-
visions (2010) in the LRFD specifications for either Method A or B. As discussed,
improper design has, in some instances, resulted in deficiencies. Extensive research has
identified various failure modes when pads are subjected to excessive shear deforma-
tion or overload. See Section 10.3.2.1.1 for further discussion of failure modes.
10.4.2.1.2
Need for Greater Longitudinal Movement Capacity
When there is a need to accommodate large displacements along with smaller axial
loads, such as might be found at the end spans of a long continuous bridge, the use of
a low-profile elastomeric bearing combined with a PTFE or higher-performing sliding

478

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


surface offers a practical solution. In these cases, cyclic end movement due to girder ro-
tation caused by truck load can cause PTFE surface wear. This wear can be avoided by
designing the elastomeric pad to accommodate rotation and the smaller cyclic move-
ment due to truck loads and by designing the sliding surface to accommodate the
larger longitudinal movement due to thermal load.
10.4.2.1.3
Need to Accommodate Multidirectional Complex Movement
When movement direction is not readily determined, such as with curved or skewed
bridges, circular pads can more easily accommodate multidirectional requirements.
Circular pads typically offer an advantage by requiring narrower bridge seats than
rectangular pads when placed on a skew; however, circular pads are not as efficient
from a design standpoint as properly oriented rectangular pads.
10.4.2.1.4
Slipping or Walking Out
Slipping can be avoided in two ways. First, limiting the use of wax additives in the manu-
facturing of the elastomer prevents or reduces future bleeding of the wax to the contact
surface, which results in a much lower coefficient of friction and bearing slipping. Slip-
ping can also be avoided by providing positive attachment of the elastomeric pad to the
bearing sole plates.
10.4.2.1.5
Improper Fabrication or Manufacturing
Effective QC during manufacture is essential for successful performance. Proper clean-
liness and care during fabrication helps prevent debonding of reinforcing plates with
elastomer.
10.4.2.2
Cotton Duck Pads
Table 10.4 summarizes solutions for various service life issues identified in Sec-
tion 10.3.2.2 for CDPs. These bearing types have performed well, but they have had
limited usage. Research testing has identified potential failure modes that could result
from improper design and/or manufacturing and installation. Further discussion relat-
ing to each issue follows.
10.4.2.2.1
Improper Pad Design Resulting in Interlayer Splitting and Delamination
Splitting and delamination damage is caused by excessive horizontal movement, axial
load, and rotation and can be avoided by following the current LRFD specifications.
Key criteria include limiting shear strain to 10% of total pad thickness and limiting
compressive stress to 3,000 psi for total load and 2,000 psi for live load.
10.4.2.2.2
Improper Pad Design Resulting in Overrotation
Overrotation can be avoided by implementing the rotation limits provided in the cur-
rent LRFD specifications. The use of narrow pads in the direction of movement and
rotation can improve expansion and rotational capacity.

479

Chapter 10. BRIDGE BEARINGS


Table 10.4.  Solutions for Service Life Problems: CDPs
Service Life Problem Solution Advantages Disadvantages
Improper pad Avoid with proper design, CDPs are economical and High stiffness reduces
design resulting in manufacturing, and very stiff in supporting capacity for rotation and
interlayer splitting installation. vertical load. horizontal translation.
and delamination Limit shear strain to 10% of Provide greater stability Service life experience and
total pad thickness. during construction. use is limited.

Limit compressive stress to


3,000 psi for total load and
2,000 psi for live load.
Improper pad Use narrow pads in direction Improves expansion and
design resulting in of movement. rotation capacity.
over rotation
Need for greater Combine with high- Can provide unlimited Use of PTFE surfaces can limit
longitudinal performing sliding surfaces. movement capability. service life due to wear.
movement capacity Additional initial cost.
Improper Provide effective QA-QC. Assures adherence to None.
fabrication or quality and performance
manufacturing requirements.

10.4.2.2.3
Need for Greater Longitudinal Movement Capacity
The high shear stiffness of these bearing types reduces their capacity for accommodat-
ing horizontal translation. This lower capacity can be accommodated by combining
CDPs with PTFE sliding surfaces. PTFE can be subject to wear, which can reduce ser-
vice life. See further discussion of solutions for sliding surface wear in Section 10.4.2.3.
10.4.2.2.4
Improper Fabrication or Manufacturing
Proper QA-QC and testing during manufacturing is necessary to confirm required
performance. Stiffness and other performance characteristics can vary between manu-
facturers, so QC testing measures have been developed.
10.4.2.3
Sliding Surface Bearings
Table 10.5 summarizes solutions for various service life issues identified in Sec-
tion 10.3.2.3 for sliding surfaces. PTFE is most commonly used for sliding surfaces. Its
service life is affected by deficiencies caused by loads, which result in wear and creep
or cold flow; and by production or operation defects, which result in damage due to
contamination. It is recommended to use higher-performing sliding surfaces when pos-
sible. The higher initial cost of these surfaces is a very small fraction of total bridge
cost, and cost by itself is not a justifiable reason for not using higher-performing sliding
surfaces in place of PTFE, which is used exclusively in United States. Further discus-
sion relating to each issue is provided following the table.

480

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 10.5.  Solutions for Service Life Problems: PTFE Sliding Surfaces
Service Life Problem Solution Advantages Disadvantages
Wear caused by Possible thicker PTFE surface. Increases time for material to Thickness may be limited by
cyclic truck load wear out. other factors.
Lower service life than
higher-performing sliding
surfaces.
Dimpled and lubricated PTFE. Improves wear resistance and Long-term effectiveness of
reduces coefficient of friction. lubrication is uncertain.
PTFE has much lower service
life than higher-performing
sliding surfaces.
Woven or glass-reinforced Greatly improves wear Has slightly increased
PTFE. resistance. coefficient of friction.
Improved sliding material, Greatly improves wear Proprietary product.
MSM. resistance. Possibly higher initial cost.
Has significantly increased
coefficient of friction without
lubrication.
Combined performance Elastomeric pad can be Requires special design
of high-performing sliding designed to accommodate considerations.
surfaces with elastomeric cyclic truck load.
pad.
Creep or cold flow Compressive overload Recess and bonding holds Has a possibly higher initial
caused by high prevention. PTFE in place. cost but prevents higher
compressive load PTFE material set in recess, future maintenance cost.
and epoxy bond.
Surface damage Protection during shipment Prevents damage to sliding Maintenance needs to be
caused by and installation. surface that increases scheduled and actually
contamination Periodic maintenance coefficient of friction and performed.
cleaning. wear.

10.4.2.3.1
PTFE Wear Due to Cyclic Truck Load
As mentioned throughout this section, the use of higher-wear-resistant sliding surface
materials is recommended, when possible, when surfaces are subject to the fast sliding
speeds associated with cyclic truck load. The reference to PTFE in this section reflects
its widespread use in practice and providing solutions where they are needed. How-
ever, this approach is not meant to recommend the use of PTFE as a material of choice
for sliding surfaces, except in instances in which wear is not a factor.
PTFE use can be improved with the following solutions:

• Thicker PTFE than that used in the United States is used in other countries, but
plain PTFE wears very rapidly, and very thick surfaces would be required to
achieve long service life when subjected to cyclic truck load.

481

Chapter 10. BRIDGE BEARINGS


• Dimpled and lubricated PTFE has been shown to provide both improved wear
resistance and reduced coefficient of friction; however, the long-term effectiveness
of lubrication is uncertain.
• Woven or glass-filled PTFE greatly improves wear resistance, but it has a higher
coefficient of friction.
• MSM is an improved sliding material that has greatly improved wear resistance
and has a reduced coefficient of friction when used in the dimpled and lubricated
condition. The coefficient of friction is greatly increased when unlubricated. This
material is a proprietary product.
• The design of combined elastomeric–PTFE bearings considers the combined per-
formance of both materials. The elastomeric pad can be designed to accommo-
date rotation along with the low-amplitude, high-cycle movement caused by truck
load, and the PTFE can be designed to accommodate the high-amplitude, low-
cycle movement caused by temperature load, which produces the least amount of
wear. These accommodations require some additional design considerations.

10.4.2.3.2
Creep or Cold Flow Due to High Compressive Stress
Creep or cold flow can be avoided by limiting compressive stress, which prevents cold
flow, but lower compressive stress results in a higher coefficient of friction. These
problems can also be avoided by recessing the PTFE for one-half its thickness, which
provides the better solution because it prevents cold flow and also allows higher com-
pressive stress for a lower coefficient of friction.
10.4.2.3.3
PTFE Wear Due to Surface Contamination
Surface contamination can be avoided by protection during shipment and installation.
Bearing assemblies should be shipped together and protected to avoid PTFE surface
damage. In addition, periodic maintenance and cleaning can mitigate the potential for
sliding surface damage by preventing buildup of dirt and other debris.
10.4.2.4
High-Load Multirotation Pot Bearings
Table 10.6 summarizes solutions for various service life issues identified in Section
10.3.2.4 for HLMR pot bearings. Past issues have most often been related to produc-
tion or operation defects relating to design and manufacturing. Expansion pot bear-
ings use PTFE sliding surfaces, which are affected by loads resulting in wear. External
steel surfaces are exposed to environmental hazards, which can cause corrosion; how-
ever, corrosion of external surfaces typically does not affect operation. Current design
and manufacturing procedures have greatly reduced past issues. Further discussion
relating to each issue follows.

482

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 10.6.  Solutions for Service Life Problems: HLMR Pot Bearings
Service Life Problem Solution Advantages Disadvantages
Improper design Proper clearances between This has been the most This has internal moving
and manufacturing pot elements. common and economical parts and requires a
resulting in leakage Proper pad rotational HLMR bearing for many high degree of QC in
of elastomer, capacity. years. manufacturing.
broken sealing Recent sealing ring and pot Internal sealing-ring wear
Proper calculation of
rings, abraded manufacturing improvements could always be a concern.
movements and rotations.
elastomeric pads, have eliminated most of the This should provide for
and internal metal- Smooth surfaces on pot rim earlier failures. bearing replacement in
to-metal contact and inside piston. Long-term deterioration design details.
Improved sealing ring testing assures greater
design and manufacturing performance.
tolerances.
Silicon grease lubrication.
Alternate methods of external
restraint for lateral loads.
Sealing and protecting pot
bearings during shipping and
installation.
Long-term deterioration
testing per current AASHTO
procedures.
Load-induced PTFE See Table 10.5 on PTFE sliding surfaces.
sliding surface wear

10.4.2.4.1
Improper Design and Manufacturing Causing Various Deficiencies
Various recommendations by Stanton et al. (1999) to mitigate or avoid previous defi­
ciencies have been incorporated into the current LRFD specifications. These recom-
mendations include

• Providing proper clearances between various elements of the pot, piston, sliding
surface, guides, and restraints to avoid binding between metal components.
• Providing proper rotational capacity with proper elastomeric pad thickness design,
controlled by a 15% strain limit on the pad-edge deflection under rotation.
• Protecting against overrotation by designing for larger rotations that include pos-
sible rotation due to construction tolerances and by placing greater emphasis on
calculation of rotations due to service loads, construction loads, and environmen-
tal conditions.
• Providing smooth surfaces on the piston and inside the pot to reduce rotational
resistance, wear, and abrasion. Metalizing of these interior surfaces for corrosion
protection should be avoided because it produces a rough surface that leads to

483

Chapter 10. BRIDGE BEARINGS


increased damage under cyclic rotation. In highly corrosive environments, stain-
less steel should be considered, but this affects cost greatly.
• Providing tight control on design and manufacturing of sealing rings to avoid
escape of elastomer. Circular cross-section rings provide a tight seal but are sus-
ceptible to wear under cyclic rotation. Flat rings appear to be more susceptible to
leakage and ring fracture, but they experience less wear. Heavier flat brass rings
have been suggested as a means of improving performance. The performance of
circular rings could also be improved if the internal friction and wear were re-
duced. Multiple flat brass sealing rings have been the most frequently used system
since the mid to late 1990s and have had good results.
• Using silicon grease lubrication to reduce potential wear on rings, pot walls, and
pistons.
• Providing alternate methods of external restraint for lateral loads to avoid having
these loads resisted by the piston rim bearing against the pot wall. Relatively small
lateral load (5% of gravity load) when combined with cyclic rotation can dramati-
cally increase the rotational resistance caused by the piston rim dragging against
the pot wall during rotation and cause wear.
• Sealing and protecting pot bearings during shipping and installation to prevent
dirt or contamination from getting inside the pot, which can lead to increased
wear and abrasion of the rings, pot, and elastomeric disc.
• Providing long-term deterioration testing per current LRFD construction specifica-
tions to assure required performance.

Even with recent improvements in serviceability, this bearing type still has internal
moving parts that are subject to wear and abrasion. This behavior can still lead to
reduced element service life and the potential need for bearing replacement before the
service life of the bridge system is realized.
10.4.2.4.2
Load-Induced Sliding Surface Wear
Expansion bearings using PTFE sliding surfaces are subject to wear from truck loads
or thermal loads. See further discussion of solutions for sliding surface wear in Sec-
tion 10.4.2.3.
10.4.2.5
High-Load Multirotation Disc Bearings
Table 10.7 summarizes solutions for various service life issues identified in Sec-
tion 10.3.2.5 for HLMR disc bearings. Research testing has identified potential issues
that would most often be related to production or operation defects relating to design
and manufacturing. In addition, expansion disc bearings use PTFE sliding surfaces,
which are subject to load-induced wear. External steel surfaces are exposed to environ-
mental hazards, which can cause corrosion. Further discussion relating to each issue
is provided.

484

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 10.7.  Solutions for Service Life Problems: HLMR Disc Bearings
Service Life Problem Solution Advantages Disadvantages
Improper design Proper design, This is a simple concept They have not been
and manufacturing manufacturing, and using a polyether urethane extensively used because of
causing fatigue installation. disc, which provides proprietary status until recent
deformation AASHTO construction multidirectional rotation years.
and abrasion of specifications require capability. Service life experience and
urethane disc cyclic testing to confirm Good performance has been fatigue testing experience are
performance. experienced over the last 40 limited.
Improper Proper analysis and design to years without any known AASHTO design specifications
design causing identify rotations. field problems. are limited.
overrotation and AASHTO design specifications There is concern that
high edge pressure limit edge contact stress on rotational stiffness can cause
and damage to PTFE surfaces. high stresses on sliding
PTFE surfaces on surfaces.
expansion bearings Additional research and
(Overrotation can experience are required to
cause binding on determine if 100-year service
center pin.) life is achievable.

Load-induced PTFE See Table 10.5 for PTFE sliding surfaces.


sliding surface wear

10.4.2.5.1
Improper Design and Manufacturing Causing Fatigue, Abrasion, and Overrotation
Various performance conclusions and recommendations were made by Stanton et al.
(1999) on the basis of tests of disc bearings subjected to combined axial load and rota-
tion. These include the following:

• The LRFD construction specifications requirements for cyclic load testing to con-
firm long-term performance should be followed. Cyclic rotation tests have shown
slight disc deformation and abrasion, but these changes did not affect performance.
• The LRFD specifications requirements limiting edge contact stress on PTFE sur-
faces in sliding expansion bearings should be followed. Tests showed that rotation
of disc bearings is partly accompanied by uplift, which can produce high edge
loading on sliding surfaces that affects service life.
• Overrotation also causes binding on the center shear pin that can be mitigated by
limiting rotation or providing adequate clearance.

10.4.2.5.2
Load-Induced Sliding Surface Wear
Expansion bearings using PTFE sliding surfaces are subject to wear from truck loads
or thermal loads. See further discussion of solutions for sliding surface wear in Sec-
tion 10.4.2.3.

485

Chapter 10. BRIDGE BEARINGS


10.4.2.6
High-Load Multirotation Spherical and Cylindrical Bearings
Table 10.8 summarizes solutions for various service life issues identified in Sec-
tion 10.3.2.6 for HLMR spherical and cylindrical bearings. Issues have most often
been related to production or operation defects relating to design and manufacturing.
Further discussion relating to each issue is provided following the table.
10.4.2.6.1
Improper Design and Manufacturing Causing Binding of Steel Components
Adequate clearances between moving bearing components should be provided to ac-
commodate full rotation demand without premature binding.
10.4.2.6.2
Improper Design Causing Overload of Spherical Surface from High Lateral Loads
An additional external restraint system should be provided to accommodate large lat-
eral loads to avoid generating excessive localized bearing stresses that can damage
PTFE surfaces.
10.4.2.6.3
Improper Manufacturing Causing Surfaces to Not Mate Properly
Tight manufacturing tolerances on curved mating surfaces should be maintained to
avoid excessive localized stresses that can damage PTFE surfaces.

Table 10.8.  Solutions for Service Life Problems: HLMR Spherical and Cylindrical Bearings
Service Life Problem Solution Advantages Disadvantages
Improper design Ensure proper clearances. This is a robust bearing This is the most expensive of
and manufacturing system that is traditionally the HLMR types.
causing binding of considered to be the most It requires a high degree of
steel components reliable HLMR type. manufacturing QC.
and reducing It can be designed to
rotational capacity accommodate large loads
Improper design Provide external restraint and rotations.
causing overload system. Behavior is close to what is
of spherical surface assumed in design.
from high lateral
loads
Improper Ensure proper manufacturing
manufacturing tolerances.
causing surfaces to
not mate properly,
causing excessive
localized stress
Load-induced PTFE See Table 10.5 for PTFE sliding surfaces.
sliding surface wear

486

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


10.4.2.6.4
Load-Induced Sliding Surface Wear
PTFE sliding surfaces are subject to wear from truck loads or thermal loads. See fur-
ther discussion of solutions for sliding surface wear in Section 10.4.2.3.
10.4.2.7
Fabricated Mechanical Steel Bearings
Table 10.9 summarizes solutions for various service life issues identified in Sec-
tion 10.3.2.7 for fabricated mechanical steel bearings. As discussed, factors affecting
service life relate to several categories including loads, primarily overload, which results
in binding or overrotation of rocker bearings; natural or man-made hazards, which re-
sult in steel element corrosion; and production or operation defects, specifically caused
by lack of maintenance. Further discussion relating to each issue is provided.
10.4.2.7.1
Overload
Special emphasis should be placed on determining load, rotation, and movement de-
mands to avoid overrotation, which can lead to excessive tilting and possible tipping.
Adequate clearances between moving components should be provided to avoid prema-
ture binding. Set rocker bearings for proper temperature alignment.

Table 10.9.  Solutions for Service Life Problems: Fabricated Mechanical Steel Bearings
Service Life Problem Solution Advantages Disadvantages
Overload causing Use proper demand This has been the longest- It is expensive to fabricate
binding or excessive parameters, proper used bearing type. and install.
rotation or tipping clearances, and proper initial This bearing type has In comparison with other
setting. potential for 100+ years of bearings, it requires
Corrosive Use stainless steel in extreme service life if maintained and additional maintenance.
environment environments. protected from corrosion,
causing steel Use galvanizing or metalizing freezing, and overrotation.
surface corrosion on bearing assembly
components (fixed or rocker)
and accompanying plates
and anchor bolts.
Improper Establish and follow
maintenance maintenance procedures to
causing debris prevent debris buildup.
buildup, affecting
movement capacity
Improper Establish and follow
maintenance procedures to prevent or
causing freezing, inhibit corrosion of steel
affecting movement surfaces.
and/or rotation
capacity

487

Chapter 10. BRIDGE BEARINGS


10.4.2.7.2
Corrosive Environment
For bearings in corrosive environments, the use of stainless steel can avoid the poten-
tial of corrosion on exposed surfaces and on contact surfaces. Galvanizing, metalizing,
or high-performance paint systems can serve to mitigate the potential for surface cor-
rosion, but a maintenance plan for the longer term will still be required.
10.4.2.7.3
Improper Maintenance Causing Debris Buildup
Proper periodic maintenance is required for bearings located below deck expansion
joints to clean bearing areas and prevent excessive buildup of debris, which can affect
horizontal movement and rotation. Dirt and debris buildup against steel surfaces can
also hold moisture and accelerate coating deterioration and steel corrosion.
10.4.2.7.4
Improper Maintenance Causing Freezing
Proper maintenance is also required in corrosive environments to prevent or inhibit
corrosion of steel surfaces, which can cause freezing and reduce rotation and move-
ment capacity. Proper field cleaning and recoating of existing steel bearings with zinc-
rich paint systems, followed by maintenance touch-up as required, will provide ex-
tended service life. Lubricating pins and knuckles that have metal-to-metal contact can
also help prevent freezing.
10.5
Strategies for Bearing Selection and Design
10.5.1
Available Service Life Design Philosophies
Currently there are no deterioration models for predicting the service life of bearings.
Although service life cannot be accurately predicted, preventive measures should be
taken to avoid deterioration and extend service life.
Previous bearing research has studied the behavior of various bearing types under
static and cyclic loading and has identified potential damage and deterioration modes
that have been addressed by improved AASHTO design and construction specifica-
tions. This research has been performed primarily to address observed field problems
or to improve understanding and design methodology. However, these studies have
focused primarily on developing criteria that will avoid observed problems or improve
performance; models have not been developed for determining how various bearing
types will deteriorate over time under given loading and environmental conditions.
Research on sliding surfaces has determined the potential for a deterioration model for
wear based on various factors, including pressure and sliding velocity, but this model
has not yet been fully developed.
With the lack of deterioration models, experience and expert opinion are the
only methods for predicting service life. However, previous experience with certain
bearing types may not be a good indicator of newer design performance, particularly
with recently updated and improved design and construction requirements. Future

488

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


long-term data collection regarding bearing performance will be necessary for devel-
oping more accurate service life predictions.
10.5.2
Bearing Selection and Design for Service Life
Selecting the proper bearing should always be done as part of the overall bridge system
development and should consider the expected bridge system behavior and optimal
superstructure–substructure interaction. Bearings should be selected and designed as
an integral part of the overall system and should not be designed as an afterthought,
which increases chances for problems. It is recommended to use elastomeric bearing
and higher-performing sliding surfaces when possible. A combination of rectangular
or circular elastomeric bearing pad and higher-performing sliding surfaces can meet
the demands of most bridges and result in a very long service life.
10.5.2.1
Selection Process
The process for selecting the proper bearing type involves four main steps:

Step 1. Determine demand by identifying operational and service life requirements


that the bearing must accommodate.
Step 2. Determine suitable options by identifying bearing types that have the potential
to accommodate demand requirements, and perform preliminary design(s) to
confirm.
Step 3. Evaluate service life mitigation and replacement requirements and evaluate
life-cycle costs.
Step 4. Select optimal bearing type considering all factors.

When considering service life, Steps 1 (demand) and 2 (options) need to address
additional issues beyond loads and movements, and consider all hazards that can have
adverse effects. Figure 10.19 illustrates the overall selection process. When considering
multiple options, the selection process should take into account the various levels of
bearing performance, the initial cost and maintenance requirements, and the reliability
of the bearings and their potential to achieve a long service life.
10.5.2.2
Detailed Steps in Selection Process
Step 1. Identify Demand Requirements
The demand step identifies the desired service life and what requirements the bearings
must accommodate from an operational and environmental standpoint throughout their
service life. Operational requirements have typically included proper determination of
gravity loads, rotations, and translational movements. However, for service life, this
step needs to further consider cyclic movements and cumulative movement due to truck
load, which in some cases can have a severe effect on service life. Environmental demand
has typically involved determination of thermal climate and corresponding temperature
ranges. But for service life, this determination also needs to identify specific local environ-
mental hazards and their consequences that need to be avoided or mitigated in the design.

489

Chapter 10. BRIDGE BEARINGS


Identify DEMAND requirements.

Compare with SUPPLY parameters for


bearing alternatives.

Identify potential bearing alternative(s)


that accommodate loads and
movements.
No

Perform preliminary design to confirm Is bearing


suitability of potential bearing type. suitable?

Yes
Identify factors affecting service life
and mitigation requirements.

Evaluate potential service life of


Is bearing No
alternatives and determine if bearing
has potential for achieving desired life. suitable?

Yes

Compare life-cycle costs considering Determine expected


initial maintenance and replacement frequency of bearing
costs. replacement.

Perform final bearing selection.

Figure 10.19.  Process flowchart for bearing selection and design for service life.

490

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 10.10 provides a format for identifying various demand requirements, and
Steps 1.1 to 1.4 summarize the process.
Step 1.1. Determine operational and service life requirements:

• Review the targeted bridge design service life, which should have been identified at
the bridge system selection stage. See Chapter 2 on bridge system selection.
• Identify total traffic and truck volumes.
• Identify local environmental factors that can affect bearing service life or performance:
–– Thermal movement: design temperature ranges, and
–– Environmental hazards, including
¤¤ Severe corrosive environment—location below deck expansion joints in
northern wet climates;
¤¤ Coastal environment—potential for salt spray; and
¤¤ Chemical environment—potential for other deleterious atmospheric or cor-
rosive activity.

Step 1.2. Determine general bearing requirements on the basis of bridge system
alternatives and superstructure–substructure interaction:

• Consider integral system options that eliminate bearings at abutments and/or at


piers.
• Consider continuous system options that eliminate deck joints at interior piers.
• Determine optimal fixity and expansion options at piers and abutments on the ba-
sis of bridge system evaluation. Consider flexibility of piers in determining options
for multiple pier fixity.
• For curved or skewed bridge systems, determine proper direction of movement;
for curved systems, determine point of fixity.

Step 1.3. Determine superstructure loads and movements for the given bridge
system(s) being considered:

• Gravity loads from dead and live loads;


• End rotations due to all sources, including construction tolerances;
• Longitudinal movements
–– Maximum movement due to temperature change,
–– Cyclic movement due to truck load, and
–– Movement due to posttensioning, creep, and shrinkage;
• Requirements for multidirectional movement; and
• Longitudinal and transverse loads to be resisted by bearings.

Step 1.4. Ensure that loads are distributed to bearings in accordance with system
analysis.

491

Chapter 10. BRIDGE BEARINGS


492
Table 10.10. Demand Requirements for Bearing Selection

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Qualitative Value: Quantitative Value: How Values Are
Demand Description Low, Medium, High If Applicable Used
General Bridge importance and system design service life Example: High Example: 100 years To identify bearing
Requirements design service life and
replacement needs
Traffic Example: High Indicate total traffic To determine cyclic
volumes and truck translation due to
volumes truck load
Environmental Temperature range for thermal climate and Moderate or cold Indicate design To compute
factors superstructure type temperature range superstructure
thermal movement
Environmental Cold climate with deck Example: Severe na To determine severity
hazards deicing chemicals and of environmental
subject to open deck joints hazards to which
Coastal climate subject to Example: High na bearings will be
salt spray subjected, and to
determine mitigation
Other chemically corrosive Example: na Identify type if
needs to achieve
environment applicable
optimal service life
Other atmospheric Example: na Identify type if
environmental factors applicable

(continued)
Table 10.10. Demand Requirements for Bearing Selection (continued)
Quantitative Value: How Values Are
Demand Description If Applicable Used
Bridge Design loads Vertical Maximum For fixed and
System (kip) Permanent expansion bearing
Loads and design
Minimum
Movements
Transverse
Longitudinal
Rotation (rad) Longitudinal Permanent
Cyclic—Maximum
Cyclic—Minimum
Transverse Permanent
Cyclic
Translation (In) Longitudinal Cyclic thermal For expansion
Cyclic truck bearing design

Irreversible
Transverse Cyclic thermal
Cyclic truck
Irreversible

Note: na = not applicable.

493

Chapter 10. BRIDGE BEARINGS


Step 2. Identify Suitable Bearing Options
The options step involves comparing demand requirements with supply parameters for
various bearing types and determining which types are suitable. Often, there may be
more than one option that can meet load and movement performance requirements.
For shorter spans with lighter loads, which represent the largest population of bridges,
plain or reinforced elastomeric pads will typically provide the best overall service.
CDPs with greater load capacity can be a suitable alternative for plain elastomeric
pads when movement demand is light; however, when movement demand increases,
CDPs will require sliding surfaces, which affect service life.
When loads and rotational demands increase, SRE bearings will have to be eval-
uated against HLMR pot or disc bearings for accommodating the required loads.
As long as an SRE bearing can be designed for the combined load, movement, and
rotation, it will also have the greatest potential for achieving the desired service life.
Greater movement capacity can be provided with SRE bearings by combining them
with sliding surfaces, but this combination typically would be considered only when
movement demand is large and vertical loads are relatively light, such as at the ends of
a long, multispan continuous unit.
When load demand is beyond the capacity of SRE bearings, other bearings such as
HLMR pot, disc, or spherical must be considered. Fabricated steel bearings can also be
considered, but cost and service life mitigation issues have to be weighed.
The preliminary design step involves performing preliminary design in accordance
with AASHTO specifications to determine if potential bearing type(s) can actually pro-
vide the required capacities depending on actual bridge layout and to evaluate further
size and geometric requirements.
The following steps summarize the process for identifying viable options:
Step 2.1. Using Table 10.11, other applicable agency or industry guidelines, or
professional experience, identify potential bearing alternatives that accommodate load
and movement requirements.
Step 2.2. Perform preliminary design on potential bearing types following
­AASHTO design requirements. Determine whether selected alternatives can actually
accommodate load and movement demand. Determine bearing size requirements.

Step 3. Evaluate Service Life Factors and Mitigation Requirements


After suitable options are identified in Step 2 based on satisfying vertical load and
movement requirements, Step 3 evaluates the bearing’s ability to resist various factors
that affect service life and considers what mitigation requirements will be necessary.
This step includes an evaluation of whether the bearing type or types have the poten-
tial for achieving the desired service life and identifies possible replacement needs and
maintenance requirements.
Table 10.11 summarizes various supply parameters for individual bearing types
related to service life. It lists relative durability factors for each bearing type and also
identifies key avoidance or mitigation requirements. In addition, relative qualitative
initial costs and maintenance requirements as part of a qualitative life-cycle cost com-
parison are listed. The table also indicates relative qualitative service life potential.

494

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 10.11.  Supply Parameters for Bearing Selection
Bearing Load and Movement Durability Factors Life-Cycle Costs
Type Performance Values

Load (kips)
Rotation (radians)
Movement (in.)
Multidirectional
Rotation/Movement
Capability
Relative Ability to
Accommodate Cyclic
Truck Movement
Resistance
to Corrosive
Environment
Resistance to
Production/
Operation Defects
Avoidance or Mitigation
Requirements
Relative Initial Cost
A = Lowest
Relative
Maintenance
Service Life Potential

Plain Low Low Low Yes High High — High Avoid pad A Low High —
elastomeric pad not splitting by potential for
pad 0 to 100 0.01 0.5 affected proper design 100+ years
Steel- Low to Medium Low to Yes— High High — High — Avoid pad B Low High —
reinforced medium medium circular SRE SRE pad splitting by potential for
elastomeric 0.02 pads pad not splitting or proper design 100+ years
pad 50 to 4 affected delamination
750
Elastomeric Low to Medium High Yes High High High Mitigate PTFE B Low to Low to
pad with medium with high wear with moderate moderate
sliding 0.02 No limit performing improved (moderate
surface 50 to sliding sliding surface with improved
750 surface. sliding
Low with surface)
PTFE.
CDP Low to Low Low No Low High — High — Avoid pad B Low Uncertain due
medium pad not CDP pad splitting by to lack of data
0.005 0.25 affected splitting or proper design
0 to 300 maximum delamination

(continued)

495

Chapter 10. BRIDGE BEARINGS


496
Table 10.11.  Supply Parameters for Bearing Selection (continued)
Bearing Load and Movement Durability Factors Life-Cycle Costs
Type Performance Values

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Load (kips)
Rotation (radians)
Movement (in.)
Multidirectional
Rotation/Movement
Capability
Relative Ability to
Accommodate Cyclic
Truck Movement
Resistance
to Corrosive
Environment
Resistance to
Production/
Operation Defects
Avoidance or Mitigation
Requirements
Relative Initial Cost
A = Lowest
Relative
Maintenance
Service Life Potential

CDP with Low to Low High Yes Low due High — High Mitigate PTFE B Low to Low
sliding medium to wear pad not wear with moderate (moderate
surface 0.005 No limit of sliding affected improved with improved
0 to 300 maximum surface sliding surface sliding
surface)
HLMR High High High Yes Low due Moderate Moderate— Ensure proper E Moderate Moderate
pot to wear for internal design and
250 to 0.02 to No limit of sliding exposed sealing ring construction.
(except 2,500+ 0.04 with surface sides of wear and Improved
with sliding slider pot and elastomer sliding
surface) piston leakage surfaces.

Metalized
surfaces.
HLMR High High High Yes Low due High Moderate Proper E Moderate Moderate
disc to wear disc not design and
250 to 0.02 to No limit of sliding affected PTFE wear construction.
(except 2,500+ 0.03 with surface
with sliding slider Improved
surface) sliding
surfaces.

(continued)
Table 10.11.  Supply Parameters for Bearing Selection (continued)
Bearing Load and Movement Durability Factors Life-Cycle Costs
Type Performance Values

Load (kips)
Rotation (radians)
Movement (in.)
Multidirectional
Rotation/Movement
Capability
Relative Ability to
Accommodate Cyclic
Truck Movement
Resistance
to Corrosive
Environment
Resistance to
Production/
Operation Defects
Avoidance or Mitigation
Requirements
Relative Initial Cost
A = Lowest
Relative
Maintenance
Service Life Potential

HLMR High High High Yes Low due Moderate Moderate. Proper F Moderate Moderate
spherical to wear for design and
No limit No limit No limit of sliding exposed PTFE wear. construction
(curved with surface sides Improved
sliding slider of steel Surfaces sliding
surface; elements not mating surfaces.
except with properly.
sliding Metalized
surface) surfaces.
Fabricated Low to High High No High Low High Use stainless D Moderate High with
steel medium for all steel or mitigation
No limit No limit elements mitigate with Low with for corrosion
(pin fixed, 50 to galvanizing or SS potential
rocker 750+ metalizing
or roller
expansion)
Steel sole na na na na na Low High Use stainless na Moderate High with
plates, base steel or mitigation
plates, and mitigate with Low with for corrosion
anchor bolts galvanizing or SS potential
metalizing

Source: Stanton et al. 2008.


Note: Performance condition limits are approximate and may not occur simultaneously. SS = stainless steel; na = not applicable.

497

Chapter 10. BRIDGE BEARINGS


The following steps summarize the process for evaluating service life requirements:
Step 3.1. For potential bearing alternatives, evaluate factors affecting service life
and identify required avoidance, mitigation, or acceptance options. Consider primary
service life–reduction categories, including reductions

• Due to loads (primarily cyclic truck load),


• Due to environmental hazards (primarily corrosive or deleterious environment),
and
• Due to production or operational defects (primarily element damage or wear).

Step 3.2. Evaluate potential service life of identified bearing alternatives with con-
sideration of any required mitigation and maintenance.

• Use deterioration models (currently not available for bearings)


–– Potential deterioration model for sliding surface resistance to wear.
• Use experience or expert opinion
–– Bearing system and/or material resistance to wear or other deterioration, and
–– Steel element resistance to and protection from corrosion.

Step 3.3. Relate potential service life of identified bearing alternatives to the target
design life of bridge system. If service life of bearing alternative is less than target bridge
system design life, consider the need to replace bearing after service life is exhausted.
Step 3.4. Evaluate life-cycle cost of bearing options considering initial cost, long-
term maintenance cost, and potential replacement cost. This evaluation can be done
qualitatively at this stage.

Step 4. Select Optimal Bearing Type


After determining suitable options according to load and movement requirements,
evaluating service life mitigation and replacement requirements, and considering and
comparing all parameters, the optimal bearing type for the given application can be
determined.
The selection process should summarize final mitigation, maintenance, and
replacement requirements that will need to be incorporated into the final design.
The final bearing design process should fine-tune the preliminary bearing design
after the final bridge system analysis and final load distribution. Final design details
would then be developed to accommodate required clearances and to include any ser-
vice life mitigation measures or details for required bearing replacement.

10.6
Bridge Management Related to Bearings
This section provides guidance related to inspection and maintenance of bearings and
needs for future data collection.

498

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


It is recommended that bridge inspections and inspection data collection for bear-
ings be expanded to identify bearing types, specific conditions, and other relevant data.
These recommendations for more detailed bearing data collection can be used within
the FHWA Long-Term Bridge Performance program, which is intended to study the
deterioration and durability of bridges and the impacts of maintenance and repair.
These recommendations can also be used to supplement the types of data c­ ollected for
use within bridge management systems such as Pontis. Improved data collection
for bearings can be useful in determining and scheduling required maintenance and for
developing more accurate deterioration models.
10.6.1
Bridge Owner’s Manual
Chapter 1 provided a detailed description of a bridge Owner’s Manual, which must
be provided when requested by the owner or if service life design of unique bridges is
involved.
The following data regarding bearing design should be included in the bridge
Owner’s Manual. This information will be helpful in evaluating future bearing
­
performance.

• Bearing type(s);
• Design movements considering temperature and truck loads;
• Design rotations considering dead load, and expected construction rotation and
live load;
• Expected bearing service life and expected replacement schedule;
• Summary of major factors considered that could affect bearing service life;
• Summary of mitigations included in design for factors affecting service life;
• Types and frequency of recommended maintenance;
• Basis for designed-in details and capabilities for future bearing replacement; and
• Special features that should be monitored with future National Bridge Inspection
Standards inspections (see Section 10.6.2).

10.6.2
Recommendations for Inspection
The following list provides inspection recommendations that can provide early indi-
cations of bearing problems for various bearing types. Early indication can facilitate
maintenance scheduling, which can prevent more serious problems leading to bearing
replacement.

• All bearings should be checked for misalignment. All guided sliding bearings
should be checked for binding against the guides. All expansion bearings should be
checked for movement position at respective temperatures and compared against
initial settings as identified in the Owner’s Manual.

499

Chapter 10. BRIDGE BEARINGS


• Elastomeric bearings should be checked for
–– Overrotation;
–– Excessive shear deformation;
–– Splitting or tearing;
–– Excessive bulging; and
–– Sliding or walking out.
• PTFE and stainless steel surfaces should be checked for
–– PTFE fragments indicating wear;
–– Migration of PTFE surface;
–– Scratching, paint, or other contamination (exposed stainless steel surfaces); and
–– Proper position of stainless steel surface on PTFE surface.
• Pot bearings should be checked for
–– Leakage of elastomer from pot;
–– PTFE and stainless steel surfaces for expansion bearings;
–– Steel surface corrosion on exposed surfaces of pot, piston and plates; and
–– Adequate rotational clearances or binding of pot elements due to rotation.
• Disc bearings should be checked for
–– Splitting, cracking, or bulging of urethane disc;
–– PTFE and stainless steel surfaces for expansion bearings; and
–– Steel surface corrosion on plates.
• Mechanical steel bearings should be checked for
–– Surface corrosion;
–– Debris buildup that could prevent movement and rotation;
–– Overrotation; and
–– Freezing.
• Miscellaneous items to check for
–– Bent or misaligned anchor bolts;
–– Loose or missing nuts on anchor bolts;
–– Voids under bearing plates; and
–– General debris buildup.

10.6.3
Future Data Needs
There is a need to study the performance of bridge bearings in actual service conditions
and to accumulate data that would be helpful in determining more accurate life predic-
tions for various bearing types given those service conditions. For example, measuring

500

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


actual longitudinal bearing movements caused by girder end rotations under traffic
load could be useful in understanding and predicting the service life of sliding surfaces.
This type of data should be collected by the FHWA Long-Term Bridge Performance
program and other research programs for enhancing the service life prediction capabil-
ity and development of deterioration models for various bearing types.

501

Chapter 10. BRIDGE BEARINGS


11
LIFE-CYCLE
COST ANALYSIS

11.1
Introduction
This chapter introduces life-cycle cost analysis (LCCA) and its use in the decision-
making process for selecting optimum cost-effective bridge systems, subsystems, and
elements that can achieve long-term service life. It includes general guidelines and best
practices for application of LCCA, outlines the steps in the process, and briefly dis-
cusses the economic principles involved. References to available models are made
without recommending the use of a specific software or model.
LCCA, in lieu of purely initial construction cost evaluation, is essential in evaluat-
ing the long-term benefit of many strategies that can achieve extended service life (in
excess of 100 years) yet require additional and sometimes considerable initial invest-
ment. LCCA can assist agencies with investment decisions by considering the initial
costs and relevant future costs associated with required inspection, maintenance, reha-
bilitation, and possible component replacement, including associated demolition, dis-
posal, and user costs.
In its broadest form, LCCA can be used in the evaluation of alternative bridge
systems. In its more simplified forms, it aids in evaluating alternatives for bridge com-
ponents, such as decks, superstructures, and substructures; or more specialized bridge
element applications, such as comparing alternatives for deck joints or bearings.
This chapter is intended only as a brief discussion of the benefits, principles, and
methodologies involved in LCCA. Additional detail on this entire process and its
application are provided in the attached references, primarily in the Life-Cycle Cost
Analysis Primer (FHWA 2002a) and in NCHRP Report 483: Bridge Life-Cycle
Cost Analysis (Hawk 2003).

502
11.2
LCCA Defined
LCCA is an analysis methodology that assists in comparing and choosing alternative
strategies for achieving long-term service life for bridge systems, subsystems, or ele-
ments. It considers not only the initial construction cost, but also all of the costs that
are expected to occur over the entire service life of the bridge, typically maintenance,
major rehabilitation, component or element replacement (including relevant demoli-
tion and disposal costs), and user costs. Economic methods are used to convert antici-
pated future costs to present dollar values so that lifetime costs of various alternatives
can be directly compared.
11.2.1
Steps in LCCA
The five basic steps in the LCCA process are described in the following sections (FHWA
2002b).

Step 1. Establish Design Alternatives


Step 1 involves establishing the elements of initial design and identifying the associated
activities that will be required throughout the structure’s service life for maintenance,
rehabilitation, or element replacement within a system or subsystem for each alterna-
tive being considered.

Step 2. Determine Activity Timing


The timing of associated activities throughout the period of comparison must be
­determined as part of the identification process. Estimating when and how often cer-
tain activities must be performed is important in making realistic comparisons. This
process might involve identifying certain required maintenance on a yearly basis, or
certain levels of potential rehabilitation due to expected wear after a specified period
of time, or when individual components or elements such as decks or bearings may
have to be replaced. Agency data are important in establishing when various levels of
maintenance, rehabilitation, or replacement may be required. In the absence of data,
expert opinion can be used.

Step 3. Estimate Costs


This step involves estimating the initial construction cost associated with each design
alternative and the costs associated with the various identified future maintenance,
rehabilitation, and replacement activities. Costs are computed on the basis of current
cost data. It is recommended as a best practice that costs include both agency costs and
user costs. This practice is discussed further in Section 11.2.2.

Step 4. Compute Life-Cycle Costs


Step 4 involves computing the present value of all costs identified for a given alterna-
tive. The concepts of present value are further discussed in Section 11.3.1. As part of
this computational process, two approaches, either deterministic or stochastic (prob-
abilistic), address the variability and uncertainty associated with input factors. The

503

Chapter 11. LIFE-CYCLE COST ANALYSIS


deterministic approach (most used) uses fixed discreet values. The stochastic approach
defines input variables by a probability distribution. These computational approaches
are further discussed in Section 11.3.4.

Step 5. Analyze Results


The final step involves comparing the initial and life-cycle costs associated with the
various alternatives and determining the optimal cost-effective solution. If the alter-
natives provide different levels of service, then the alternatives that provide the best
overall long-term benefit can also be compared.
11.2.2
LCCA Cost Types
11.2.2.1
Agency Costs
Agency costs include all the costs borne by the agency or owner of the bridge (includ-
ing design, initial construction, inspection, maintenance, rehabilitation, and element
replacement) and have been the primary elements for consideration in LCCA. Some
costs, such as initial construction, rehabilitation, and replacement, are more easily
estimated on the basis of current industry cost data. Other costs, such as maintenance,
are more difficult and rely on the existence and accuracy of agency historical cost data.
In the absence of these data, expert opinion can be used.
11.2.2.2
User Costs
User costs are primarily associated with reduced traffic capacity in work zones and
involve costs to the users because of delays, vehicle operating costs, and accidents.
Estimating these costs may be the greatest challenge to LCCA implementation. Many
agencies have been reluctant to incorporate user costs into LCCA because of the dif-
ficulty and uncertainty in assigning value to user delay time, or because user costs are
not factored into agency budgets. These issues have led many agencies to give lesser
credence to user costs in evaluating overall lowest-cost solutions; however, minimizing
user impacts and associated costs is a major concern today, especially with implementa-
tion of accelerated bridge construction. A recent pooled-fund study led by the ­Oregon
Department of Transportation (DOT) (Doolen et al. 2011) developed a set of decision-
making tools to determine if accelerated bridge construction techniques are more ef-
fective than traditional construction for a given bridge replacement or rehabili­tation
project. These tools incorporate quantified user costs as part of an LCCA evaluation.
User costs can play an important factor in evaluating various options for long-
term service life. Section 11.3.3 further discusses the approach for estimating user
costs based on traffic volumes and user delays.
11.2.2.3
Vulnerability Costs
Vulnerability costs are associated with extraordinary circumstances and risks, such
as overload, collision, blast, fire, floods, scour, or earthquake, and typically would
not be included in an LCCA for comparison of service life strategies. They are useful,

504

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


however, in evaluating vulnerability of existing bridges that might have a high prob-
ability for one or more of these extreme events.
11.2.3
LCCA Versus Benefit–Cost Analysis
LCCA is a subset of benefit–cost analysis, which compares benefits, as well as costs, in
selecting optimal alternatives. Benefit–cost analysis is useful in comparing alternatives
that do not achieve the same level of service or benefit. For example, benefit–cost anal-
ysis can be useful in comparing bridge replacement options that provide either three
or four lanes of traffic with corresponding different levels of service. Clearly there are
both cost and benefit variations between each option. LCCA is typically considered
alone for service life evaluation of various alternatives that can ultimately provide the
same level of service.
11.3
Elements of LCCA
This section discusses methodologies for determining net present value (NPV) and
discount rates. It further discusses other elements considered in LCCA such as activity
timing and service life, user costs, computational approaches, and analysis tools.
11.3.1
Net Present Value
The NPV concept in LCCA is an economic method for combining initial costs and
present dollar values of future expected costs so that lifetime costs for various alter-
natives can be directly compared. Dollars spent at different times within a structures
life have different present values, so the projected activity costs for an alternative can-
not simply be added together to calculate the total life-cycle cost for that alternative.
Accord­ing to TRB’s Transportation Economic Committee (2013),

A dollar today is worth more than a dollar five years from now, even if there
is no inflation, because today’s dollar can be used productively in the ensuing
five years, yielding a value greater than the initial dollar. Future benefits and
costs are discounted to reflect this fact. The purpose of discounting is to put all
present and future costs into a common metric, their net present value [NPV].

The formula to convert the sum of the initial cost and the present value of future
repair and renewal costs into NPV is given by Equation 11.1:

N  1 
NPV = initial cost + ∑ rehab costk   (11.1)
 (1 + r ) k 
n
k=1

where
r = real discount rate,
k = order number of a rehabilitation activity undertaken in the future,
N = total number of rehabilitation activities, and
nk = year in the future when the cost will be incurred.

505

Chapter 11. LIFE-CYCLE COST ANALYSIS


 1 
The term   is called the discount factor.
 (1 + r ) k 
n

Input factors in computing NPV are the initial cost (usually the cost of planning,
design and construction of a new structure), the cost and timing of rehabilitation activ-
ities, and the real discount rate.
Discounting is a method of considering the opportunity cost of money as it applies
to current versus future funds. It can be thought of in terms of the alternative economic
return that could be gained on funds such as earning interest. The computed potential
amount of interest is based on what is referred to as the discount rate, which is gener-
ally described as having three components:

1. The real opportunity cost of capital to account for productive value of funds;
2. The premium to account for financial risk (i.e., that the loan will not be repaid);
and
3. The anticipated rate of inflation.

According to TRB’s Transportation Economic Committee,

Life-cycle analyses typically ignore inflation because the prediction of future


prices introduces unnecessary uncertainty into the analysis. Therefore, dis-
count rates are typically based on interest rates for government borrowing,
which have little risk, with the inflation component removed, yielding the
“real” interest rate. This rate is typically calculated by subtracting the rate of
inflation (consumer price index) from the interest rate of an investment such
as a 10-year U.S. Treasury bill. For example, if the interest on a 10-year Trea-
sury bill is 5.5% and the inflation rate is 3%, then the discount rate would be
2.5%.

Circular No. A94 (OMB 1992) provides general guidance for conducting cost-
effectiveness analyses and provides specific guidance on the discount rates to be used
in evaluating programs whose benefits and costs are distributed over time. It also
provides standard criteria for deciding whether programs can be justified on economic
principles. The OMB publishes real interest rates for NPV analyses on its website.
As Figure 11.1 shows, the discount rate can have a significant impact on the analy-
sis. A low discount rate favors projects with long-term benefits and near-term costs.
When evaluating alternative projects, a sensitivity analysis using a range of discount
rates can be used to determine the importance or impact of the discount rate in the
relative project performance. Even with a low discount rate, values far in the future
have a relatively low present value.

506

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Increasing Rate

Figure 11.1. Effect of discount rates on present value.

Figure 11.1. Effect of discount rates on present value.


11.3.2
Activity Timing, Service Life, and Life Cycle
11.3.2.1
Deterioration Models
Deterioration models describe the relationship between the condition of the bridge (or
its element) and time, showing how the bridge deteriorates. The model assumes that
no replacements or major repairs are made, but it usually implies that scheduled main-
tenance actions are performed as planned. The basic model applies either to a bridge
as a whole, or to any of its elements (e.g., deck, substructure, bearings, columns). The
shape of a deterioration curve depends on the type of the element and the definitions
of condition states.
Figure 11.2 shows a deterioration curve. If the bridge is placed in service at time
T0, its condition gradually declines, and the deterioration curve represents its condi-
tion over time. Initially the condition is good, but after a period of wear and aging, it
eventually (at time Tf) reaches an unacceptably low condition (Cf). The period between
T0 and Tf is called the service life of the bridge.
In practice, however, realistic deterioration models that are based on actual physi-
cal and chemical deterioration processes are generally not available to accurately pre-
dict service life. The most acceptable deterioration model is in the form of the solution
to Fick’s second law, used to predict the rate of chloride ingress through concrete
cover. This model, including its limitations, is described in Chapter 5 of the Guide. It
is expected that with time additional deterioration models will become available that

507

Chapter 11. LIFE-CYCLE COST ANALYSIS


Bridge Deterioration
Figure 11.2. Bridge
deterioration curve.

Condition
Service Life
Cf

Time of Use
T0 Tf

will greatly enhance quantification of the service life of bridge elements, components,
subsystems, and systems.
Developing deterioration models is a data-intensive procedure that is complicated
by the lack of knowledge of the underlying processes that foster deterioration, as well
as by data availability. In lieu of deterioration models based on actual physical and
chemical deterioration processes, other approximate methods must be used. Bridge
management software programs such as Pontis and BRIDGIT, which are used in nearly
all 50 states, have deterioration models contained within them that are typically based
on expert opinion and analysis of available historical data.
Recent studies by the New York State DOT (Agrawal et al. 2009) and the Florida
DOT (Sobanjo 2011) have further attempted to develop bridge element deteriora-
tion models on the basis of state DOT bridge inspection databases along with expert
opinion. The New York State DOT study applied computerized statistical methods
to develop deterioration curves using inspection data going back to 1981. The study
included the influence of various factors such as average daily truck traffic and climate.
The New York State DOT study implemented a stochastic approach to account
for the uncertainty and randomness of factors affecting the deterioration process. In
the stochastic approach, the ratings of bridge elements (reflective of their condition
at a particular time) and the durations that elements will stay at a particular rating
were assumed to be random variables and were modeled by probability distributions.
The study developed and compared deterioration curves using both Markov chain–
and Weibull distribution–based stochastic models. Markov chain, the most commonly
used model for developing deterioration rates for infrastructure facilities, is used in
advanced bridge management systems such as Pontis and BRIDGIT. It models the
deterioration process by considering the probability of transition from one condition
state to another in a discrete time, and it accounts for the current element condition in
predicting the future condition. A Weibull-based model considers the probability of
how long a bridge element will remain at a particular state, and it also considers past
conditions. The New York State DOT study found that the Weibull-based models gen-
erally provided the best overall fit with historical bridge inspection data.
508

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Methods that use historical data to develop long-term bridge element deteriora-
tion models have certain limitations. Aside from not considering the actual physical
deterioration process, older historical data do not consider more recent improvements
in construction materials or methods, which can greatly affect future service life. Fur-
ther, there is some uncertainty about extrapolating the data beyond the duration for
which the data were collected.
11.3.2.2
Expenditure Stream
As shown in Figure 11.3, a bridge will deteriorate over the period of its service life if
left unattended. However, in most cases a bridge is not left to follow the basic dete-
rioration path and reach an unacceptable condition without interruption. The agency
responsible for the bridge will, from time to time, undertake repairs, rehabilitations,
and renewals that return conditions to higher levels and extend its service life.
The sequence of events and actions that determine the bridge condition through-
out its life cycle is called a life-cycle activity profile. Actions are usually associated with
expenditures that have to be incurred when repair, rehabilitation, and renewal activi-
ties occur. These expenditures may be plotted on a separate diagram that represents
the stream of expenditures associated with construction and repair activities. Such
a diagram is sometimes called a cash flow diagram. An example of such diagram is
shown in Figure 11.4.
In cash flow diagrams, all resource flows are usually attributed either to the begin-
ning or the end of the time period in which they actually occur. In cases in which
a resource flow is extended over several periods, the expenses are represented as a
series of lines. Cash flow diagram expenses are shown as lines graphed in the positive
domain; revenues and other returns (e.g., the terminal value of the bridge) are shown
as negative values. For simplicity, only agency costs have been plotted on the example
cash flow diagram shown in Figure 11.4. When other types of costs (e.g., user costs
and vulnerability costs) are included, the cash flow diagram serves as a graphic repre-
sentation of the NPV computation.
Bridge Condition
Figure 11.3. Bridge
Service Life ­condition life cycle.

Condition

Cf

T0 Time Tf
of use

509

Chapter 11. LIFE-CYCLE COST ANALYSIS


Figure 11.4.  Cash flow diagram.

11.3.3
Estimating User Costs
Work zone user costs are the increased vehicle operating costs, delay, and crash costs
incurred by highway users as a result of construction, maintenance, or rehabilitation
work zones. User costs may represent the greatest data challenge for consideration in
an LCCA. When calculated, user costs are often so large that they may substantially
exceed agency costs, particularly for transportation investments being considered for
high-traffic areas. Congestion statistics and cost can be obtained from the Annual
Urban Mobility Report prepared by the Texas Transportation Institute (Schrank et al.
2011).
FHWA’s Life-Cycle Cost Analysis in Pavement Design (Walls and Smith 1998)
includes a rational step-by-step procedure for determining user costs associated with
work zones:

Work zone is defined in the Highway Capacity Manual (2010) as an area of


a highway where maintenance and construction operations impinge on the
number of lanes available to traffic or affect the operational characteristics of
traffic flowing through the area. . . . In order to analyze work zone user costs,
work zone characteristics associated with alternative designs and supporting
maintenance and rehabilitation strategies must be defined as part of the devel-
opment of alternative [designs].

510

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Work zone characteristics of concern include such factors as work zone
length, number and capacity of lanes open, duration of lane closures, timing
(e.g., hours of the day, days of the week, season of the year) of lane closures,
posted speed, and the availability of and the physical and traffic characteristics
of alternative routes. The strategy for maintaining traffic should include any
anticipated restrictions on the contractor’s or maintenance force’s hours of
operation or ability to establish lane closures.

Specific details in an LCCA should include

1. Projected year work zones occur (Years 5, 8, 12, and so forth);

2. Number of days the work zone will be in place (construction period);

3. Specific hours of each day, as well as the days of the week the work zone
will be in place; and

4. Work zone length and posted speed.

The duration of a work zone (the overall length of time a facility or por-
tion of a facility is out of service or traffic is restricted) can range from sporadic
daily lane closures for maintenance to several months for bridge-deck replace-
ments. [In many cases,] the differential routine maintenance cost between [al-
ternatives] tends to be insignificant when compared with initial construction
and rehabilitation costs. To a great extent, the same is true of user costs result-
ing from routine reactive-type maintenance activities. Routine maintenance
work zones tend to be relatively infrequent, of short duration, and outside
of peak traffic flow periods. As such, analysts should focus attention on user
costs associated with major work zones.

User costs are directly dependent on the volume and operating character-
istics of the traffic on the facility. Each construction, maintenance, and reha-
bilitation activity generally involves some temporary effect on traffic using the
facility. The effect can vary from insignificant for minor work zone restrictions
on low-volume facilities to highly significant for major lane closures on high-
volume facilities.

The major traffic characteristics of interest for each year a work zone will
be established include

1. The overall projected average annual daily traffic volumes on both the
facility and possible alternate routes;

2. The associated 24-hour directional hourly demand distributions; and

3. The vehicle classification distribution of the projected traffic streams.

511

Chapter 11. LIFE-CYCLE COST ANALYSIS


On high-volume routes, distinctions between weekday and weekend
­traffic demand and hourly distributions become important. Seasonal average
­annual daily traffic distribution also becomes important when work zones are
proposed on recreational routes during seasonal peak periods. . . .

Once the individual work zones have been identified, each is evaluated
separately. This is the point at which individual user cost components are
quantified and converted to dollar cost values.

A detailed example of the user cost calculation is given in Life-Cycle Cost Analysis
in Pavement Design (Walls and Smith 1998).
As mentioned in Section 11.2.2.2, the recent pooled-fund study led by the Oregon
DOT (Doolen et al. 2011) developed a set of decision-making tools to evaluate the cost-
effectiveness of using accelerated bridge construction techniques versus conventional
construction. This study incorporated user costs as part of the LCCA comparison.
The tools also incorporated an analytical hierarchy process, which is a technique that
aids decision makers in prioritizing multiple criteria by using a multilevel hierarchical
structure of objectives, criteria, and alternatives. This process considers both quantita-
tive and qualitative criteria and quantifies the qualitative trade-offs and relationships
between criteria by using a hierarchy of criteria. This analysis of trade-offs was impor-
tant for accelerated bridge construction because it quantified various qualitative fac-
tors contributing to user costs, such as user delay from a long detour, and could show
the economic benefit resulting from reduced construction duration.
11.3.4
Computational Approaches
The two approaches used in preparing an LCCA differ dramatically in how they ad-
dress the variability and uncertainty associated with various input factors and with
the risk associated with the various uncertainties. Often there is some level of possible
variability and uncertainty in regard to the values identified for each input parameter.
This possible variation can often have significant effects on the LCCA outcome.
11.3.4.1
Deterministic Approach
Traditionally in the deterministic approach, input variables are treated as fixed values,
as if those values were certain. This approach assigns each LCCA input variable with
a fixed (base case) value based on statistics and nonlinear regression of actually occur-
ring data or professional judgment.
This method does not specifically address the degree of variability or uncertainty
with input values. In order to incorporate uncertainty about input values into the
analysis, a sensitivity analysis can be performed to see the effect of variation on any
one parameter. However, the deterministic approach combined with the sensitivity
analysis has two drawbacks. First, it can only be applied to input variables one by one,
when the real question of interest is how the variation in several variables simultane-
ously can affect the result. Even more importantly, sensitivity analysis alone does not
provide any information on the relative likelihood of different outcomes. For example,
the sensitivity analysis may suggest that if the initial construction cost is 10% higher
512

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


than is assumed in the base case, the corresponding NPV of all costs would be 7%
higher than in the base case. However, it will provide no information on whether this
scenario is likely to occur. In order to characterize the relative likelihood of various
potential outcomes, a stochastic approach should be adopted.
11.3.4.2
Stochastic Approach
A stochastic approach (sometimes referred to as a probabilistic approach or risk analy-
sis approach) defines the value of input variables by a frequency (probability) distri-
bution. For a given project alternative, the uncertain input parameters are identified.
Then, for each uncertain parameter, a sampling distribution of possible values is de-
veloped. Simulation programming randomly draws values from the stochastic descrip-
tion of each input variable and uses these values to compute a forecasted NPV. This
sampling process is repeated through thousands of iterations, and through the process,
an entire probability distribution of NPVs is generated for the project alternative along
with the mean or average NPV for that alternative. The resulting NPV distribution can
then be compared with the projected NPVs for alternatives, and the most economical
option for implementing the project can be determined for any given risk level.
The concept of risk arises from the uncertainty associated with future events and
the inability to know what outcomes will result from particular actions taken today.
Risk can be objective or subjective. Subjective risk is a person’s perception of the likeli-
hood of a particular event; this perception of risk may include the ability to avoid the
risk and an assessment of the consequences of a particular outcome. Objective risk,
on the other hand, incorporates theory, experiments, observation, and other unbiased
information. Ultimately, decision makers who are characterized by different degrees
of risk tolerance and whose perception of risk is intrinsically subjective make the deci-
sions. However, the goal of risk analysis is to provide the best-unbiased risk estimates
to arm the decision makers with the most accurate representation of objective risk.
Figure 11.5 depicts the stochastic approach as a whole. Stated succinctly, it reflects
uncertainty in input factors (e.g., construction and repair costs and the timing of those
activities) in the probability distribution of the results. A description of the specific
steps involved in the stochastic approach is provided.
11.3.4.2.1
Develop Structure and Layout of Problem
The first step in conducting risk analysis consists of reducing the problem to its most
basic elements and describing it in the form of an analytical model. The models for
LCCA problems typically include NPV computation, definition of cost categories,
and determination of other functional relationships, such as bridge condition curves.
­Project alternatives are also identified and described. Cash flow diagrams and life-cycle
curves are convenient tools to clearly present the project details and the features of
alternative implementations.
Once the structure of the problem is fully determined, a list of inputs can be devel-
oped. An example of input variables for an LCCA project is presented in Table 11.1.
For each of the input variables, the general basis used to determine their values is

513

Chapter 11. LIFE-CYCLE COST ANALYSIS


Uncertainty
in Timing
Uncertainty
Uncertainty
in
in Repair
Construction
Costs
Costs

Probability
Distribution
of NPV

Figure 11.5.  Stochastic approach to LCCA.

established, and a subset of input variables for which a stochastic distribution will be
used is specified.

11.3.4.2.2
Develop Input Data
The second step in conducting an LCCA is developing probability distributions for the
uncertain variables identified in the next step. A probability distribution describes the
complete range of values that a variable may assume and weighs the likelihood of its
occurrence. Figure 11.6 illustrates some of the most common probability distributions
shown in a histogram format: uniform, triangular, and normal distributions. The hori-
zontal axis provides a range of all possible values that the variable assumes, and the
vertical axis shows the relative frequency weighting of the occurrence of any particular
value. For the distributions in Figure 11.6, the probability of a range of values is equal
to the area under the curve, and the total area under the curve is equal to one.
The choice of a particular distribution depends on the type of input and the amount
of data available. A triangular distribution (see Figure 11.6) is the most common dis-
tribution used to represent various variables using expert elicitation. Expert opinions
are used to determine the minimum, the maximum, and the most likely value, and the
triangle is constructed using those three points. This method is most appropriate for
modeling such input variables as service life, discount rate, work zone delay, and so
forth. Normal distribution is the most common continuous distribution used to repre-
sent random variables symmetrically distributed around the mean value. It is usually
a good candidate to represent cost-like variables, such as construction cost and main-
tenance cost. A normal distribution can either be used to represent the information
514

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Table 11.1.  LCCA Input Variables
LCCA Component Input Variable Source
Initial and future agency costs Preliminary engineering Estimate
Construction management Estimate
Construction Estimate
Maintenance Assumption
Timing of costs Bridge deterioration Projection
User costs Current traffic Estimate
Future traffic Projection
Hourly demand Estimate
Vehicle distributions Estimate
Value of delay time Assumption
Work zone configuration Assumption
Work zone hours of operation Assumption
Work zone duration Assumption
Work zone activity years Projection
Crash rates Estimate
Crash cost rates Assumption
Vulnerability costs Flood probability Estimate
Flood damage distribution Estimate
Earthquake probability Estimate
Earthquake damage Estimate
Load distribution probability Estimate
Load-related structural damage Estimate
Other parameters Discount rate Assumption

Figure 11.6.  Example probability distributions.

Figure 11.6. Example probability distributions.


515

Chapter 11. LIFE-CYCLE COST ANALYSIS


obtained from the expert elicitation or to use data collected from other sources when
the amount of data is sufficiently large. If little information about the input variable
is available, a uniform distribution might be used as a rough approximation. Uniform
distribution assumes that outside of a certain range, the probability of outcomes drops
to zero. Within the range, however, there is no information as to which outcomes are
more likely, so it is assumed uniform over that range.
Although normal and uniform distributions are symmetric around their mean
­values, the triangular distribution can be either symmetric or asymmetric. Other types
of common asymmetric distributions are exponential and lognormal. When a large
amount of hard data is available, the best distribution can be determined by using sta-
tistical techniques to establish goodness of fit for each of them. However, when data
availability is limited, a triangular distribution can be used as a rough estimate of the
distribution shape.
Figure 11.7 shows a cumulative probability distribution in ascending order that
portrays a cumulative probability of a group of possible events. (Sometimes cumula-
tive distributions are shown in descending order, and then the points on the curve
show the probability of exceeding a particular value.) For example, there is an 80%
probability that the project cost will not exceed $4 million. Each probability distribu-
tion has a corresponding cumulative distribution. Cumulative distributions present the
data in a form that is easy to interpret for the purposes of risk analysis.
11.3.4.2.3
Develop Probability Distributions
When existing data are available, the standard method is to use the data to choose a
functional form of a probability distribution that best fits the available data. Many
statistical methods and software packages can be used to compare common distribu-
tion types with the available data and to determine goodness of fit, that is, to indicate
how well a particular probability distribution fits the data.

Figure 11.7.  Ascending cumulative probability distribution.

516

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


However, when sufficient relevant data are not readily available, group inter-
views are often used to develop probabilities of uncertain variables. Expert panels are
convened to establish the boundaries and general shape of input distributions. The
process of eliciting information from experts may include group meetings, individual
interviews, and formal surveys and questionnaires. The goal of expert elicitation is to
establish the general shape of the input distribution and to determine if there are any
interdependencies among input variables. The process of expert elicitation is shown
on Figure 11.8.
11.3.4.2.4
Perform Simulations
The next step in the LCCA process involving risk analysis is to run a computer simu-
lation of the model in order to obtain results. The process of using random numbers
to sample from a probability distribution is known as Monte Carlo sampling. In the
Monte Carlo simulation process, a series of random numbers is generated by the com-
puter along the cumulative probability scale of input distribution (see Figure 11.7).
Values corresponding to each random number are sampled along the x-scale. The sam-
pled value for one input is then combined with sampled values for all other inputs to
compute the single result. This process is repeated hundreds or thousands of times to
generate a cumulative distribution of the outcomes. The stopping rule involves either
a prespecified number of iterations or a convergence rule, which reflects the situation
in which additional iterations do not significantly affect the distribution of the results.
Monte Carlo simulations require a large number of iterations to assure that values
with low probabilities are sufficiently sampled and represented in the results. Such
sampling is especially important when the input distributions are highly unsymmetric.
When the number of iterations is insufficient, the low-probability outcomes may be
underepresented and not adequately accounted for in simulations. This is especially
significant when a low-probability outcome can have a particularly strong effect on
the results. To avoid such problems, different sampling methods can be employed. For
example, Latin hypercube sampling uses special techniques to generate samples from
all probability ranges with a relatively low total number of iterations.
Frequency

Expert Opinion
Elicitation

Model Parameter

Figure 11.8.  Using expert opinion to develop a probability distribution.


Figure 11.8. Using expert opinion to develop probability distributions.
517

Chapter 11. LIFE-CYCLE COST ANALYSIS


11.3.4.2.5
Interpret Results
The final step in the risk analysis is the interpretation of the results. If the analysis
uses the traditional deterministic approach, the only data available to decision ­makers
would be the means of the output distributions for the alternatives investigated. Based
on the comparisons of the means (50% probability), the difference between Alterna-
tive 1 and Alternative 2 in Figure 11.9 is small and may be assumed to be negligible.
The risk analysis allows one to evaluate a much more nuanced picture. When inter-
preting the risk profile in Figure 11.9, it is important to distinguish the upside risk
from the downside risk. Downside risk for project costs implies a cost overrun, a
chance that the costs will be much higher than anticipated. In contrast, the upside risk
presents an opportunity for low cost, a cost underrun. From Figure 11.9, it can be
seen that Alternative 1 has a greater upside risk than Alternative 2; in other words, it
has a better chance that the cost will be very low. At the same time, Alternative 1 is a
preferred choice because it reduces the downside risk of cost overruns. Another con-
sideration is to compare the two alternatives at the 80% cumulative probability level.
Alternative 1 is about $100,000, and Alternative 2 is $130,000. Although the means
have a negligible risk difference, Alternative 1 exhibits a far smaller financial risk.
As a part of the risk assessment, a sensitivity analysis of simulation results can
be performed to identify the key input variables that have the most influence on the
output distributions. Typically, this analysis is done by computing the degree of cor-
relation between inputs and outputs: the higher the degree of correlation, the more
significant a particular input variable is for determining the results.

Alternative 1 Alternative 2

Figure 11.9.  Cumulative risk profile of NPV for Alternatives 1 and 2.

518

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


From the perspective of most transportation agencies, the application of stochastic
LCCA is relatively new. Stochastic LCCA has become more practical as a result of the
dramatic increases in computer data-processing capabilities. Simulating and account-
ing for simultaneous changes in LCCA input parameters can now be accomplished
easily and quickly and provides invaluable information for making informed decisions.
11.3.5
LCCA Analysis Tools
11.3.5.1
Simplified LCCA Applications
Various tools and software are available for determining LCCA. All steps of a simpli-
fied deterministic LCCA can be performed using generic spreadsheet software such as
Microsoft Excel.
Several commercial programs are available for conducting a stochastic or risk-
based LCCA. These microcomputer-based risk and analysis software programs are
either spreadsheet based or work as add-ons to other generic spreadsheet software.
These tools incorporate Monte Carlo simulation capabilities for stochastic analyses.
Two common applications are presented in Table 11.2.
In using simple generic spreadsheets, the economic life-cycle analysis model has to
be programmed into the spreadsheet, and other considerations such as incorporating
user costs also must be computed separately. However, the user has much more control
over the process and how data are used and presented. For simple applications, generic
spreadsheet solutions are common and easily applied.
11.3.5.2
Comprehensive LCCA Applications
Specialized comprehensive software tools for performing LCCAs have recently been
developed by federal agencies and are available from government public websites. The
available software tools described are capable of calculating comprehensive life-cycle
costs including agency, user, operations and maintenance, disposal, and remaining ser-
vice life costs. They are also capable of performing both deterministic and stochastic
analyses.

Table 11.2. Risk-Based LCCA Software


Software Name Producer
@Risk Palisade Corporation
www.palisade.com
Oracle Crystal Ball Decisioneering Corporation
www.decisioneering.com

519

Chapter 11. LIFE-CYCLE COST ANALYSIS


11.3.5.2.1
BridgeLCC Software from the National Institute of Standards and Technology
BridgeLCC 2.0 is comprehensive LCCA software developed by the National Insti-
tute of Standards and Technology (NIST) to help bridge designers determine the cost-­
effectiveness of alternative bridge designs, construction and repair strategies, and
­construction materials (Ehlen 2003). The software uses a life-cycle costing methodol-
ogy based on the ASTM E917 standard practice for life-cycle costing and a cost classi-
fication scheme developed by NIST (Ehlen 2003). This software is specifically tailored
to highway bridges.
BridgeLCC 2.0 can segregate costs by bearer (agency, user, and third party), by
timing (initial construction; operations, maintenance, and repair; and disposal), and
by component (deck, superstructure, substructure, other, nonelemental, and new
technology introduction). The program also includes advanced features such as cal-
culation of user delay and cost, has capabilities for both deterministic and stochastic
analyses, and includes sensitivity analysis and risk analysis using Monte Carlo simula-
tion (Ehlen 2003).
11.3.5.2.2
RealCost LCCA Software
RealCost was developed by FHWA to support the application of LCCA in the pave-
ment project-level decision-making process (FHWA 2004). RealCost automates
­FHWA’s LCCA methodology as it applies to pavements. Work is being considered
to make it applicable to bridges, but currently, bridges have not been included in this
document.
The software calculates life-cycle values for both agency and user costs associated
with construction and rehabilitation. The software can perform both deterministic
and stochastic modeling (Monte Carlo simulation) and can also include user costs.
Outputs are provided in tabular and graphic format.
RealCost is an add-in for Microsoft Excel providing a graphical user interface that
facilitates the creation of an Excel Workbook containing the input data and results of
the LCCA. RealCost can only evaluate two alternatives at a time.

520

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


REFERENCES

AASHTO. 1983. Standard Specifications for Seismic Design of Highway Bridges. American
Association of State Highway and Transportation Officials, Washington, D.C.
AASHTO. 1994. Guide for Painting Steel Structures. Association of State Highway and
Transportation Officials, Washington, D.C.
AASHTO. 2007. LRFD Bridge Design Specifications, 4th ed. American Association of State
Highway and Transportation Officials, Washington, D.C.
AASHTO. 2008. Guide Specifications for Bridges Vulnerable to Coastal Storms. American
Association of State Highway and Transportation Officials, Washington, D.C.
AASHTO. 2010a. LRFD Bridge Construction Specifications, 3rd ed. American Association
of State Highway and Transportation Officials, Washington, D.C.
AASHTO. 2010b. LRFD Bridge Design Specifications. American Association of State High-
way and Transportation Officials, Washington D.C.
AASHTO. 2011. Guide Specifications for LRFD Seismic Bridge Design, 2nd ed. American
Association of State Highway and Transportation Officials, Washington, D.C.
AASHTO. 2012. LRFD Bridge Design Specifications, 6th ed. American Association of State
Highway and Transportation Officials, Washington, D.C.
AASHTO/NSBA Steel Bridge Collaboration. 2004. Steel Bridge Bearing Design and Detail-
ing Guidelines. NSBA G9.1-2004 and AASHTO SBB-1. National Steel Bridge Alliance and
American Association of State Highway and Transportation Officials, Washington, D.C.
AASHTO/NSBA Steel Bridge Collaboration. 2006. Guide Specification for Application of
Coating Systems with Zinc-Rich Primers to Steel Bridges. NBSA S8.1-2006. National Steel
Bridge Alliance and American Association of State Highway and Transportation Officials,
Washington, D.C.
Aboutaha, R. S. 2001. Investigation of Durability of Wearing Surfaces for FRP Bridge Decks.
Project C-01-50. Syracuse University, Syracuse, N.Y., and Cornell University, Ithaca, N.Y.

521
ACI. 1989. Reapproved 1994. ACI Committee 216, Guide for Determining the Fire Endur-
ance of Concrete Elements. ACI 216R-89. American Concrete Institute, Farmington Hills,
Mich.
ACI. 1996. State-of-the-Art Report on Fiber Reinforced Concrete. ACI 544.1R-96. Ameri-
can Concrete Institute, Farmington Hills, Mich.
ACI. 1997. Cracking of Concrete Members in Direct Tension. ACI 224.2R-92. Committee
224, American Concrete Institute, Farmington Hills, Mich.
ACI. 1998. Report on Alkali-Aggregate Reactivity. ACI 221.1R-98. American Concrete
Institute, Farmington Hills, Mich.
ACI. 2000. Reinforcement for Concrete: Materials and Applications. ACI Education Bul-
letin E2-00. American Concrete Institute, Farmington Hills, Mich.
ACI. 2001a. Control of Cracking in Concrete Structures. ACI 224R-01. American Concrete
Institute, Farmington Hills, Mich.
ACI. 2001b. Protection of Metals in Concrete Against Corrosion. ACI 222R-01. American
Concrete Institute, Farmington Hills, Mich.
ACI. 2003a. Design and Construction Practices to Mitigate Corrosion of Reinforcement in
Concrete Structures. ACI 222.3R-03. American Concrete Institute, Farmington Hills, Mich.
ACI. 2003b. Guide for Structural Lightweight-Aggregate Concrete. ACI 213R-03.
­American Concrete Institute, Farmington Hills, Mich.
ACI. 2004. Concrete Repair Guide. ACI 546R-04. American Concrete Institute,
­Farmington Hills, Mich.
ACI. 2005a. Building Code Requirements for Structural Concrete and Commentary. ACI
318-11, Farmington Hills, Mich.
ACI. 2005b. Guide to Shotcrete. ACI 506R-05. American Concrete Institute, Farmington
Hills, Mich.
ACI. 2006. Guide for the Use of Silica Fume in Concrete. ACI 234R-06. American Con-
crete Institute, Farmington Hills, Mich.
ACI. 2007a. Report on Fiber-Reinforced Polymer (FRP) Reinforcement for Concrete Struc-
tures. ACI 440R-07. American Concrete Institute, Farmington Hills, Mich.
ACI. 2007b. Self-Consolidating Concrete. ACI 237R-07. American Concrete Institute
Farmington Hills, Mich.
ACI. 2008a. Guide to Durable Concrete. ACI 201.2R-08. American Concrete Institute,
Farmington Hills, Mich.
ACI. 2008b. Guide to Fiber-Reinforced Shotcrete. ACI 506.1R-08. American Concrete
Institute, Farmington Hills, Mich.
ACI. 2010. Guide to Cold Weather Concreting. ACI 306R-10. American Concrete Institute,
Farmington Hills, Mich.
ACI. 2011. Building Code Requirements for Structural Concrete and Commentary. ACI
318-11, Farmington Hills, Mich.
ACPA. 1998. Guidelines for Partial Depth Spall Repair. American Concrete Pavement As-
sociation, Skokie, Ill.

522

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Adams, M., J. Nicks, T. Stabile, J. Wu, W. Schlatter, and J. Hartmann. 2011. Geosynthetic
Reinforced Soil Integrated Bridge System Synthesis Report. FHWA-HRT-11-027. Federal
Highway Administration, U.S. Department of Transportation.
AFGC. 2002. Ultra High Performance Fibre-Reinforced Concretes—Interim Recommenda-
tions. Association Française de Génie Civil, Paris.
AGA. 2006. Hot-dip Galvanizing for Corrosion Protection: A Specifier’s Guide. American
Galvanizers Association, Centennial, Colo. Updated here: http://www.galvanizeit.org/­
uploads/publications/Galvanized_Steel_Specifiers_Guide.pdf.
Agrawal, A. K., and Z. Yi. 2008. Blast Effects on Highway Bridges. University Transporta-
tion Research Center Report, City College of New York, New York.
Agrawal, A. K., A. Kawaguchi, and Z. Chen. 2009. Bridge Element Deterioration Rates.
Report No. C-01-51. Transportation Infrastructure Research Consortium, New York State
Department of Transportation, Albany.
AISC. 1994. Manual of Steel Construction: Load and Resistance Factor Design, 2nd ed.
American Institute of Steel Construction, Chicago, Ill.
Ala, N. 2011. Seamless Bridge and Roadway System for the U.S. Practice. PhD dissertation.
University of Nebraska–Lincoln.
Ala, N., and A. Azizinamini. In press a. Analytical Study and Design Provisions for Seam-
less Bridge System for the U.S. Practice. ASCE Bridge Engineering Journal.
Ala, N., and A. Azizinamini. In press b. Experimental Study of New Seamless Bridge Sys-
tem. ASCE Bridge Engineering Journal.
Ala, N., E. Power, and A. Azizinamini. Quantitative Prediction of Service Life of Conven-
tional and High-Performing Sliding Surfaces in Bearing Devices. ASCE Bridge Engineering
Journal, submitted for publication.
Alabama DOT. 1994. Cracks in Precast Prestressed Bulb Tee Girders on Structure Nos.
I-565-45-11.5 A. & B. on I-565 in Huntsville, Alabama. Alabama Department of Transpor-
tation, Montgomery.
Albers, W. F., O. Hag-Elsafi, and S. Alampalli. 2007. Dynamic Analysis of the Bentley
Creek Bridge with FRP Deck. Special Report 150. Transportation Research and Develop-
ment Bureau, New York State Department of Transportation, Albany.
Alblas, B. P., and A. M. van Londen. 1997. The Effect of Chloride Contamination on the
Corrosion of Steel Surfaces: A Literature Review. Journal of Protective Coatings & Linings,
European Edition (PCE), February, pp. 16–25.
America’s Climate Choices: Panel on Advancing the Science of Climate Change, National
Research Council. 2011. Advancing the Science of Climate Change. National Academies
Press, Washington, D.C.
America’s Climate Choices: Panel on Advancing the Science of Climate Change, National
Research Council. 2010. Adapting to the Science of Climate Change. National Academies
Press, Washington, D.C.
Anderson, P., and K. Brabant. 2010. Increased Use of MSE Abutments. Presented at Inter-
national Bridge Conference, Pittsburgh, Pa.
API. 1993. Recommended Practice for Planning, Designing and Constructing Fixed Off-
shore Platforms. American Petroleum Institute, Washington, D.C.

523

REFERENCES
Appleman, B. R., R. Weaver, and J. Bruno. 1995. Maintenance Coating of Weathering Steel:
Field Evaluation and Guidelines. FHWA Report FHWA-RD-92-055. Office of Engineering
and Highway Operations R&D, Federal Highway Administration, McLean, Va.
Arneson, L. A., L. W. Zevenbergen, P. F. Lagasse, and P. E. Clopper. 2012. Hydraulic Engi-
neering Circular No. 18: Evaluating Scour at Bridges, 5th ed. Federal Highway Administra-
tion, U.S. Department of Transportation.
Arockiasamy, M., and M. Sivakumar. 2005. Effects of Restraint Moments in Integral
Abutment Bridges. Integral Abutment and Jointless Bridges: The 2005-FHWA Conference,
Baltimore, Md., pp. 185–198.
Arsoy, S., R. M. Barker, and J. M. Duncan. 2002. Experimental and Analytical Investiga-
tions of Piles and Abutments of Integral Bridges. Research report. Virginia Polytechnic
Institute and State University, Charlottesville.
ASM. 1998. Metals Handbook. ASM Handbook Committee and American Society for
­Metals International, Metals Park, Ohio.
ASTM. 2010. Standard Practice for Measuring Life-Cycle Costs of Buildings and Building
Systems. E 917-05, 2010. ASTM International, West Conshohocken, Pa.
ASTM. 2011. Standard Practices for Cycle Counting in Fatigue Analysis. ASTM E1049-
85(2011)e1. ASTM International, West Conshohocken, Pa.
Azizinamini, A. 2002. Full-Scale Testing of Old Steel Truss Bridge. Journal of Construc­
tional Steel Research, Vol. 58, No. 5–8, pp. 843–858.
Azizinamini, A. 2009. A New Era for Short-Span Bridges. Modern Steel Construction. Na-
tional Steel Bridge Alliance, American Institute of Steel Construction, Chicago, Ill.
Azizinamini, A. 2014. Simple for Dead Load and Continuous for Live Load Steel Bridge
Systems. AISC Engineering Journal, Vol. 51, No. 2.
Azizinamini, A., and J. B. Radziminski. 1989. Static and Cyclic Performance of Semi-Rigid
Steel Beam-to-Column Connections. ASCE Journal of Structural Engineering, Vol. 115, No.
12, pp. 2979–2999.
Azizinamini, A., and S. K. Ghosh. 1997. Steel Reinforced Concrete Structures in 1995
HyogokenNanbu Earthquake. ASCE Journal of Structural Engineering, Vol. 123, No. 8,
pp. 986–992.
Azizinamini, A., T. E. Boothby, Y. Shekar, and G. Barnhill. 1994a. Old Concrete Slab
Bridges I: Experimental Investigation. ASCE Journal of Structural Engineering, Vol. 120,
No. 11, pp. 3284–3304.
Azizinamini, A., Y. Shekar, T. E. Boothby, and G. Barnhill. 1994b. Old Concrete Slab B
­ ridges
II: Analysis. ASCE Journal of Structural Engineering, Vol. 120, No. 11, pp. 3305–3319.
Azizinamini, A., S. Kathol, and M. Beacham. 1995a. Effect of Cross Frames on Behavior of
Steel Girder Bridges. Transportation Research Board, 4th International Bridge Engineering
Conference, San Francisco, Calif., Vol. 1, pp. 117–124.
Azizinamini, A., S. Kathol, and M. Beacham. 1995b. Influence of Cross Frames on Load-
Resisting Capacity of Steel Girder Bridges. AISC Engineering Journal, Vol. 32, No. 3, pp.
107–116.

524

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Azizinamini, A., N. J. Lampe, and A. J. Yakel. 2003a. Toward Development of a Steel
Bridge System—Simple for Dead Load and Continuous for Live Load. Final report.
Nebraska Department of Roads, National Bridge Research Organization, University of
Nebraska–Lincoln.
Azizinamini, A., A. J. Yakel, and J. P. Swendroski. 2003b. Development of a Design
Guide for Phase Construction of Steel Girder Bridges. NDOR Research Project Number
SPRPL1 (038) P530. National Bridge Research Organization, Department of Civil Engi-
neering, University of Nebraska–Lincoln.
Azizinamini, A., A. J. Yakel, and M. Farimani. 2005a. Development of a Steel Bridge
System—Simple for Dead Load and Continuous for Live Load: Vol. 1: Analysis and Rec-
ommendations. Final report. Nebraska Department of Roads, National Bridge Research
Organization, University of Nebraska–Lincoln.
Azizinamini, A., A. J. Yakel, N. J. Lampe, N. Mossahebi, and M. Otte. 2005b. Develop-
ment of a Steel Bridge System Simple for Dead Load and Continuous for Live Load, Vol. 2:
Experimental Results. NDOR P542. Nebraska Department of Roads, Lincoln, Neb.
Azizinamini, A., A. J. Yakel, and S. J. Niroumand. 2008. Use of Steel Bridge System with
Continuity for Live Load Only in Conjunction with Conventional and Accelerated Bridge
Construction Methodologies. Proc., 2008 Accelerated Bridge Construction—Highway for
Life Construction, Federal Highway Administration, Baltimore, Md., pp. 107–114.
Azizinamini, A., A. J. Yakel, A. Sherafati, R. Taghinezhad, and J. H. Gull. Flexible Pile
Head in Jointless Bridges: Design Provisions for H-Piles in Cohesive Soils. ASCE Bridge
Engineering Journal, submitted for publication.
Babaei, K., and A. M. Fouladgar. 1997. Solutions to Concrete Bridge Deck Cracking. Con-
crete International, Vol. 19, No. 7, pp. 34–37.
Badie, S. S., M. K. Tadros, and K. E. Pedersen. 2001. Re-Examination of I-Girder/Pier Con-
nection in Jointless Bridges. PCI Journal, March–April, pp. 62–74.
Ball, C., and D. W. Whitmore. 2005. Innovative corrosion mitigation solutions for existing
concrete structures. International Journal of Materials and Product Technology (IJMPT),
Vol. 23, No. 3/4, 2005
Barker, R. M., J. M. Duncan, K. B. Rojiani, P. S. K. Ooi, C. K. Tan, and S. G. Kim. 1991.
NCHRP Report 343: Manuals for the Design of Bridge Foundations. TRB, National
­Research Council, Washington, D.C.
Barson, J. 1994. Chapter 3: Properties of Bridge Steels. In AISC (American Institute of Steel
Construction). Highway Structures Design Handbook, Volume I. AISC, Chicago, IL.
Baun, M. D. 1993. Steel Fiber-Reinforced Concrete Bridge Deck Overlays: Experimental
Use by Ohio Department of Transportation. In Transportation Research Record 1392,
TRB, National Research Council, Washington, D.C.
Bazant, Z. P., and L. Panula. 1980. Creep and Shrinkage Characterization for Analyzing
Prestressed Concrete Structures. Journal of the Prestressed Concrete Institute, Vol. 25,
pp. 86–122.
Beavers, J. A., and C. L. Durr. 1998. NCHRP Report No. 408: Corrosion of Steel Piling in
Nonmarine Applications. TRB, National Research Council, Washington, D.C.
Bennet, J., and T. J. Schue. 1993. Evaluation of NORCURE Process for Electro­chemical
Chloride Removal from Steel-Reinforced Concrete Bridge Components. SHRP-C-620.
­Strategic Highway Research Program, National Research Council, Washington, D.C.

525

REFERENCES
Bergson, P. 1998. Letter to Donald Flemming, Minnesota state bridge engineer; Re: Bridge
No. 9340—Load Test Results, Dec. 21.
Berke, N. S., and A. Rosenberg. 1989. Technical Review of Calcium Nitrite Corrosion
Inhibitor in Concrete. In Transportation Research Record 1211, TRB, National Research
Council, Washington, D.C.
Bertolini, L., F. Bolzoni, A. Cigada, T. Pastore, and P. Pedeferri. 1993. Cathodic Pro-
tection of New and Old Reinforced Concrete Structures. Corrosion Science, Vol. 35,
pp. 1633–1639.
Bertolini, L., S. W. Yu, and C. L. Page. 1996. Effects of Electrochemical Chloride Extraction
on Chemical and Mechanical Properties of Hydrated Cement Paste. Advances in Cement
Research, Vol. 8, No. 31, pp. 93–100.
Bilow, D. N., and M. E. Kamara. 2008. Fire and Concrete Structures. Presented at Struc-
tures Congress, Structural Engineering Institute, American Society of Civil Engineers,
Reston, Va.
Blunt, J. D., and C. P. Ostertag. 2009. Deflection Hardening and Workability of Hybrid
Fiber Composites. ACI Materials Journal, Vol. 106, No. 3, pp. 265–272.
Bourgin, C., E. Chauveau, and B. Demelin. 2006. Stainless Steel Rebar: The Choice of Ser-
vice Life. Revue de Métallurgie, No. 2, February.
Brakke, B. 2002. Opinions expressed at Fatigue Workshop. Lehigh University, Bethlehem, Pa.
Brandt, T., A. Varma, B. Rankin, S. Marcu, R. Connor, and K. Harries. 2011. Effects of
Fire Damage on the Structural Properties of Steel Bridge Elements. Report FHWA-PA-
2011-009-PIT011. Pennsylvania Department of Transportation, Harrisburg, and University
of Pittsburgh, Pittsburgh, Pa.
Briaud, J. L., R. W. James, and S. B. Hoffman. 1997. NCHRP Synthesis 234: Settlement of
Bridge Approaches (The Bump at the End of the Bridge). TRB, National Research Council,
Washington, D.C.
Bridge, R., S. Griffiths, and G. Bowmaker. 2005. The Concept of a Seamless Concrete
Pavement and Bridge Deck. In Australian Structural Engineering Conference 2005 (M.G.
Stewart and B. Dockrill, eds.) Sydney, New South Wales, Australia, pp. 289–298.
Bruneau, M., J. W. Wilson, and R. Tremblay. 1996. Performance of Steel Bridge During the
1995 Hyogoken-Nanbu (Kobe, Japan) Earthquake. Canadian Journal of Civil Engineering,
Vol. 23, No. 23, pp. 678–713.
BSI. 2012. Cathodic Protection of Steel in Concrete. EN 12696. British Standards Institu-
tion, London.
Buckle, I., I. Friedland, J. Mander, G. Martin, R. Nutt, and M. Power. 2006. Seismic Retro-
fitting Manual for Highway Structures: Part 1—Bridges. Publication FHWA-HRT-06-032.
Office of Infrastructure Research and Development, Federal Highway Administration, U.S.
Department of Transportation.
Burke, M. P., Jr. 2009. Integral and Semi-Integral Bridges. Wiley-Blackwell, Oxford, United
Kingdom.
Caltrans. 2008. Isolated Partial Depth Concrete Repair. In MTAG Volume II: Rigid Pave-
ment Preservation, 2nd ed. Caltrans Division of Maintenance, California Department of
Transportation, Sacramento.

526

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Campbell, T. I., and W. L. Kong. 1987. TFE Sliding Surfaces in Bridge Bearings. Report
ME-87-06. Ontario Ministry of Transportation and Communications, Downsview, Ontario,
Canada.
Caner, A. 1996. Analysis and Design of Jointless Bridge Decks Supported by Simple-Span
Girders. Ph.D. Dissertation. North Carolina State University, Raleigh.
Caner, A., and P. Zia, P. 1998. Behavior and Design of Link Slabs for Jointless Bridge Decks.
PCI Journal, Vol. 43, pp. 68–80.
Chandra, V. 1995. Draft Report on Precast Prestressed Concrete Integral Bridges, State-of-
the-Art. Precast/Prestressed Concrete Institute, Chicago, Ill., p. 113.
Chang, L. M., and Y. J. Lee. 2001. Evaluation and Policy for Bridge Deck Expansion
Joints. FHWA/IN/JTRP-2000/1. Joint Transportation Research Program, Indiana Depart-
ment of Transportation, Indianapolis, and Purdue University, West Lafayette, Ind.
Chang, L. M., and Y. J. Lee. 2002. Evaluation of Performance of Bridge Deck Expansion
Joints. Journal of Performance of Constructed Facility, Vol. 16, No. 1, pp. 3–9.
Chen, R., and J. A. Yura. 1995. Wax Build-Up on the Surfaces of Natural Rubber Bridge
Bearings. Research Report 1304-4. Center for Transportation Research, University of
Texas, Austin.
Chung, R. (ed.). 1996. NIST Special Publication 901: The January 17, 1995 Hyogoken-
Nanbu (Kobe) Earthquake Performance of Structures, Lifelines, and Fire Protection Sys-
tems. Building and Fire Research Laboratory, National Institute of Standards and Technol-
ogy, Gaithersburg, Md.
Clark, W. 2001. Research Needs Statements. Maintenance and Design of Steel Abut-
ment Piles in Iowa Bridges. Transportation Research Board. http://rns.trb.org/dproject.
asp?n=27091. Accessed Oct 16, 2013.
Clemeña, G. G. 2003. Report on the Investigation of the Resistance of Several New
­Metallic Reinforcing Bars to Chloride-Induced Corrosion in Concrete. VTRC 04-R7.
­Virginia Transportation Research Council, Charlottesville.
Clemeña, C. G., and Jackson, D. R. 2000. Trial Application of Electrochemical Chloride
Extraction on Concrete Bridge Components in Virginia, VTRC 00-R18, Virginia Transpor-
tation Research Council.
Clifton, J. R., H. F. Beehgly, and R. G. Mathey. 1974. Non-Metallic Coatings for Concrete
Reinforcing Bars. Report FHWA-RD-74-18. Federal Highway Administration, U.S. Depart-
ment of Transportation.
Clough, G. W., and J. M. Duncan. 1991. Earth Pressures. In Foundation Engineering Hand-
book, Van Nostrand Reinhold, New York, pp. 223–235.
Connor, R. J., R. Dexter, and H. Mahmoud. 2005. NCHRP Synthesis 354: Inspection and
Management of Bridges with Fracture-Critical Details. Transportation Research Board of
the National Academies, Washington, D.C.
Corbett, W. D. 2011. Preparing an Inspection Plan for Bridge Maintenance Painting. Proc.,
2011 SSPC Conference, Las Vegas, Nev. http://www.kta.com/Corbett_Preparing_an_Inspec-
tion_Plan.pdf.
Corven, J., and A. Moreton. 2004. Post-Tensioning Tendon Installation and Grouting
Manual. Corven Engineering, Tallahassee, Fla., and Federal Highway Administration, U.S.
Department of Transportation.

527

REFERENCES
Culmo, M., and R. Seraderian. 2010. Development of the Northeast Extreme Tee (NEXT)
Beam for Accelerated Bridge Construction. PCI Journal, Vol. 55, No. 3, pp. 86–101.
Darwin, D., C. E. Locke, Jr., J. Balma, and J. T. Kahrs. 1999. Evaluation of Stainless Steel
Clad Reinforcing Bars. SL Report No. 99-3. University of Kansas Center for Research, Inc.,
Lawrence.
Darwin, D., J. P. Browning, T. V. Ngyuen, and C. E. Locke. 2002. Mechanical and Cor-
rosion Properties of a High-Strength, High-Chromium Reinforcing Steel for Concrete.
SD2001-05-F. University of Kansas Center for Research, Inc., Lawrence.
Darwin, D., J. Browning, W. Lindquist, H. A. K. McLeod, J. Yuan, M. Toledo, and D.
Reynolds. 2010. Low-Cracking, High-Performance Concrete Bridge Decks: Case S­ tudies
over First 6 Years. In Transportation Research Record: Journal of the Transportation
Research Board, No. 2202, Transportation Research Board of the National Academies,
Washington, D.C., pp. 61–69.
da Silva, M. Ferreira. 2011. Development of Self-Stressing System for Bridge Applica-
tion with Emphasis on Precast Panel Deck System. PhD dissertation. University of
Nebraska–Lincoln.
Davis, R. T. 2008. Creep and Shrinkage of Structural Lightweight Concretes. HPC Bridge
Views, Vol. 49, May–June.
Davison, N., G. K. Glass, A. C. Roberts, and J. M. Taylor. 2003. The Protective Effects of
Electrochemical Treatment in Reinforced Concrete. Corrosion 2003, National Association
of Corrosion Engineers, San Diego, Calif.
Day, K. W. 2006. Concrete Mix Design, Quality Control and Specification, 3rd ed. Taylor
and Francis, New York.
Decker, J. B., K. M. Rollins, and J. C. Ellsworth. 2008. Corrosion Rate Evaluation and
Prediction for Piles Based on Long-Term Field Performance. Journal of Geotechnical and
Geoenvironmental Engineering, Vol. 134, pp. 341–351.
Deshpande, P. G., J. D. Seddelmeyer, H. G. Wheat, D. W. Fowler, and J. O. Jirsa. 2000.
Corrosion Performance of Polymercoated, Metal-Clad and Other Rebars as Reinforcement
in Concrete. Research Report 4904-2. Center for Transportation Research, University of
Texas at Austin.
Dexter, R. J., and Fisher, J. W. 1999. “Fatigue and Fracture.” In Handbook of Bridge Engi-
neering (W. F. Chen, ed.), CRC Press, New York.
Dexter, R., R. Fitzpatrick, and D. St. Peter. 2003. Fatigue Strength and Adequacy of Fatigue
Crack Repairs. Final report. Report No. SSC-425. Ship Structure Committee, U.S. Coast
Guard, Washington, D.C.
Dexter, R. J., and J. M. Ocel. 2013. Manual for Repair and Retrofit of Fatigue Cracks in
Steel Bridges. Draft report (originally completed 2005). Office of Research and Technology
Service, Federal Highway Administration, McLean, Va.
Dicleli, M., and S. M. Albhaisi. 2004. Effect of Cyclic Thermal Loading on the Performance
of Steel H-Piles in Integral Bridges with Stub-Abutments. Journal of Constructional Steel
Research, Vol. 60 (Compendex), pp. 161–182.
Doolen, T., A. Saeedi, and S. Emami. 2011. Accelerated Bridge Construction (ABC) Deci-
sion Making and Economic Modeling Tool. Report No. FHWA-OR-TPF-12-06. Federal
Highway Administration, U.S. Department of Transportation.

528

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Doust, S. E. 2011. Extending Integral Concepts to Curved Bridge Systems. PhD dissertation.
University of Nebraska–Lincoln.
Duwadi, S., and E. Munley. 2011. Hazard Mitigation R&D Series: Article 5—Securing the
Nation’s Bridges. Public Roads, Vol. 74, No. 6.
Edvardsen, C. A. 2008. Design for 100 Plus Years of Service Life: European Approach.
In Proceedings: 2008 Accelerated Bridge Construction—Highway for Life Construction,
Federal Highway Administration, U.S. Department of Transportation, Washington, D.C.,
pp. 17–28.
Eglinton, M. S. 1975. Review of Concrete Behavior in Acidic Soils and Groundwaters.
Technical Note 69. CIRIA, London.
Ehlen, M. A. 2003. BridgeLCC 2.0 Users Manual: Life-Cycle Costing Software for the Pre-
liminary Design of Bridges. NIST GCR 03-85. National Institute of Standards and Technol-
ogy, Technology Administration, U.S. Department of Commerce, Gaithersburg, Md.
Elias, V., P. E. Barry, and R. Christopher. 1997. Mechanically Stabilized Earth Walls and
Reinforced Soil Slopes Design and Construction Guidelines. Demonstration Project 82,
Publication No. FHWA-NHI-00-043. National Highway Institute, Federal Highway
Admin­istration, U.S. Department of Transportation.
Ellor, J. M., W. Young, and J. Repp. 2004. NCHRP Report 528: Thermally Sprayed Metal
Coatings to Protect Steel Pilings: Final Report and Guide. Transportation Research Board
of the National Academies, Washington, D.C.
El-Safty, A. K. 1994. Behavior of Jointless Bridge Decks. PhD dissertation. North Carolina
State University, Raleigh.
Emanuel, J. H., and J. L. Hulsey. 1977. Prediction of the Thermal Coefficient of Expansion
of Concrete. Journal of the American Concrete Institute, Vol. 74, pp. 149–155.
ENR. 1994. Temperature Gradient Cracks Viaduct Girders. Engineering News-Record, Vol.
232, No. 22, p. 14.
Errera, S.J. 1964. Chapter 2: Materials. In Structural Steel Design (Tall, L., Beedle, L.S., and
Galambos, T.V., eds.), Lehigh University, The Ronald Press Company, New York.
Fang, K. I. 1985. Behavior of Ontario-Type Bridge Deck on Steel Girders. PhD dissertation.
University of Texas, Austin.
Faraji, S., J. M. Ting, D. S. Crovo, and H. Ernst. 2001. Nonlinear Analysis of Integral
Bridges: Finite Element Model. Journal of Geotechnical and Geoenvironmental Engineer-
ing, Vol. 127, No. 5, pp. 454–461.
Farimani, R., S. Javidi, D. Kowalski, and A. Azizinamini. 2014. Numerical Analysis and
Design Provision Development of Simple for Dead–Continuous for Live Bridge System.
AISC Engineering Journal, Vol. 51, No. 3.
FCC. 1979. Expansion Joints in Buildings. Technical Report No. 65. Federal Construction
Council, Washington, D.C.
FDOT. 2009. C09-08 Wave Forces on Bridges: Implementation of AASHTO Guide Speci-
fications for Bridges Vulnerable to Coastal Storms. Temporary design bulletin. Florida
Department of Transportation, Tallahassee.
FDOT. 2012. Index 20210 Series Florida U Beams, rev. January 2012. Instructions to
­Design Standards. Florida Department of Transportation, Tallahassee.

529

REFERENCES
FHWA. 1989. Uncoated Weathering Steel in Structures, updated 2011. T-5140.22. Federal
Highway Administration, U.S. Department of Transportation.
FHWA. 1997 (updated 2012). Bridge Coatings Technical Note – Metalized Steel Bridge
Coatings. Federal Highway Administration, U.S. Department of Transportation.
FHWA. 2002a. Life-Cycle Cost Analysis Primer. Federal Highway Administration, U.S.
Department of Transportation.
FHWA. 2002b. High Performance Steel Designers’ Guide, 2nd ed. Federal Highway
Admin­istration, U.S. Department of Transportation.
FHWA. 2004. RealCost Life-Cycle Cost Analysis User Manual. Federal Highway Adminis-
tration, U.S. Department of Transportation.
FHWA. 2011a. Bridge Deck Construction Process Review, updated April 4, 2011. FHWA
Construction. Accessed April 19, 2012.
FHWA. 2011b. Bridge Preservation Guide. Publication FHWA-HIF-11042. Federal High-
way Administration, U.S. Department of Transportation.
FHWA. 2011c. Performance Evaluation of One-Coat Systems for New Steel Bridges. Re-
port No. FHWA-HRT-11-046. Office of Infrastructure Research and Development, Federal
Highway Administration, McLean, Va.
FHWA. 2012. Action: Clarification of Requirements for Fracture Critical Members.
Memorandum June 20, 2012. Federal Highway Administration, U.S. Department of
Transportation.
FIB. 2006. Model Code for Service Life Design. fib Bulletin 34. Lausanne, Switzerland,
p. 116.
Fincher, H. E. 1983. Evaluation of Rubber Expansion Joints for Bridges. Report No.
FHWA/IN/RTC-83/1, FHWA. U.S. Department of Transportation, pp. 15–16.
Fisher, J. W., B. M. Barthelemy, D. R. Mertz, and J. A. Edinger. 1980. NCHRP Report 227:
Fatigue Behavior of Full-Scale Welded Bridge Attachments. TRB, National Research Coun-
cil, Washington, D.C.
Fisher, J. W., J. Jian, D. C. Wagner, and B. T. Yen. 1990. NCHRP Report 336: Distortion-
Induced Fatigue Cracking in Steel Bridges. TRB, National Research Council, Washington,
D.C.
Fisher, J. W., G. L. Kulak, and I. F. C. Smith. 1998. A Fatigue Primer for Structural Engi-
neers. National Steel Bridge Alliance.
Flemming, D. 2002. Opinions expressed at Fatigue Workshop. Lehigh University,
­Bethlehem, Pa.
Fletcher, F. B., Townsend, H. E., and Wilson, A. D. 2003. Corrosion Performance of
Improved Weathering Steels for Bridges. Presented at the World Steel Bridge Symposium,
National Steel Bridge Alliance, Orlando, Fla., November 19–20.
Fletcher, F., Wilson, A., Strasky, J., Kilpatrick, J., Mlcoch, T., and Wrysinski, J. 2005. Stain-
less Steel for Accelerated Bridge Construction. Presented at the FHWA Accelerated Bridge
Construction Conference, San Diego, Calif.
Freyermuth, C. L. 1969. Design of Continuous Highway Bridges with Precast, Prestressed
Concrete Girders. Journal of the Prestressed Concrete Institute, Vol. 14, No. 2, pp. 14–39.

530

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Fyfe, E., P. Milligan, and S. Watson. 2006. A Discussion of Failure Modes on High Load
Bridge Bearings Which Have Resulted in the Current Design Methods, Philosophy, and
Guidelines for Disc Bearings. 6th World Congress on Joints, Bearings and Seismic Systems
for Concrete Structures, Halifax, Nova Scotia, Canada.
Gajda, J. 2007. Mass Concrete for Buildings and Bridges. Portland Cement Association,
Skokie, Ill.
Gangarao, H. V. S., and H. K. Thippeswamy. 1996. Study of Jointless Bridge Behavior and
Development of Design Procedures. Report No. FHWA-WV89. West Virginia Department
of Transportation, Charleston.
Garcia, H. M. 2007. Analysis of an Ultra-High Performance Concrete Two-Way Ribbed
Bridge Deck Slab. FHWA-HRT-07-055. Federal Highway Administration, U.S. Department
of Transportation.
Gastal, F. P. S. L. 1986. Instantaneous and Time-Dependent Response and Strength of Joint-
less Bridge Beams. PhD dissertation. North Carolina State University, Raleigh.
Gere, J. M., and B. J. Goodno. 2012. Mechanics of Materials. Cengage Learning, Stamford,
Conn.
Geren, K., and M. Tadros. 1994. The NU Precast/Prestressed Concrete Bridge I-Girder
Series. PCI Journal, Vol. 39, No. 3, pp. 26–39.
Good-Mojab, C. A., A. J. Patel, and A. R. Romine. 1993. Innovative Materials Development
and Testing: Volume 5: Partial Depth Spall Repair in Jointed Concrete Pavements. SHRP-
H-356. Strategic Highway Research Program, National Research Council, W ­ ashington, D.C.
Goodspeed, C. H., S. Vanikar, and R. A. Cook. 1996. High-Performance Concrete Defined
for Highway Structures. Concrete International, Vol. 18, No. 2, pp. 62–67.
Graybeal, B. 2006. Material Property Characterization of Ultra-High Performance Con-
crete. Report No. FHWA-HRT-06-103. Federal Highway Administration, U.S. Department
of Transportation.
Graybeal, B., and J. Tanesi. 2007. Durability of Ultrahigh-Performance Concrete. ASCE
Journal of Materials in Civil Engineering, Vol. 19, No. 10, pp. 848–854.
Gregory, E., G. Slater, and C. Woodley. 1989. NCHRP 321: Welded Repair of Cracks in
Steel Bridge Members. TRB, National Research Council, Washington, D.C.
Hanna, K. E., G. Morcous, and M. K. Tadros. 2009. Transverse Post-Tensioning Design
and Detailing of Precast, Prestressed Concrete Adjacent Box-Girder Bridges. PCI Journal,
Vol. 54, No. 4, pp. 160–170.
Hanna, K., G. Morcous, and M. Tadros. 2010. Design Aids of NU I-Girder Bridges. Final
report. University of Nebraska–Lincoln and Nebraska Department of Roads, Lincoln.
Hanna, K. E., G. Morcous, and M. K. Tadros. 2011. Adjacent Box Girders Without Inter-
nal Diaphragms or Post-Tensioned Joints. PCI Journal, Vol. 56, No. 4.
Hansen, J., K. E. Hanna, and M. K. Tadros. 2012. Simplified Transverse Post-Tensioning
Construction and Maintenance of Adjacent Box Girders. PCI Journal, Vol. 57, No. 2, pp.
64–79.
Hansink, J. D. 1994. Maintenance Tips. Journal of Protective Coatings and Linings, Vol.
11, No. 3, p. 66.

531

REFERENCES
Hartt, W. H., R. G. Powers, F. P. Marino, M. Paredes, R. Simmons, H. Yu, R. Himiob, and
Y. P. Virmani. 2009. Corrosion Resistant Alloys for Reinforced Concrete. HRT-09-020.
Federal Highway Administration, U.S. Department of Transportation.
Hassiotis, A. 2007. Data Gathering and Design Details of an Integral Abutment Bridge.
18th ASCE Engineering Mechanics Division Conference, Blacksburg, Va.
Hausammann, H., J. W. Fisher, and B. T. Yen. 1983. Effect of Peening on Fatigue Life
of Welded Details. Proc., W. H. Munse Symposium on Behavior of Metal Structures,
­Philadelphia, Pa., pp. 70–83.
Hawk, H. 2003. NCHRP Report 483: Bridge Life-Cycle Cost Analysis. Transportation
Research Board of the National Academies, Washington, D.C.
Hearn, G. 1995. Faulted Pavements at Bridge Abutments. Colorado Transportation Insti-
tute Synthesis, University of Colorado at Boulder, p. 181.
Heffron, R. 2007. Innovative Techniques for Extending the Service Life of Deteriorating
Concrete Piers and Wharves. Proc., 11th Triennial International Conference on Ports, Ports
2007: 30 Years of Sharing Ideas: 1977–2007, San Diego, Calif.
Heimann, J., and G. Schuler. 2010. The Implementation of Full Depth UHPC Waffle Bridge
Deck Panels: Phase 1 Final Report. DTFH61-09-G-00006. Highways for LIFE Technology
Partnerships Program, Federal Highway Administration, Washington, D.C.
Highway Capacity Manual 2010. 2010.Transportation Research Board of the National
Academies, Washington, D.C.
Holm, T. A., and J. P. Ries. 2006. Lightweight Concrete and Aggregates. In Significance of
Tests and Properties of Concrete and Concrete Making Materials. ASTM STP 169D. West
Conshohocken, Pa., pp. 548–560.
Holowka, M., R. A. Dorton, and P. F. Csagoly. 1980. Punching Shear of Restrained Circular
Slabs. Ministry of Transportation and Communications, Downsview, Ontario, Canada.
Horton, R., E. Power, K. Van Ooyen, A. Azizinamini, and G. Krupicka. 2002. High Perfor-
mance Steel Cost Comparison Study—Phase II, Steel Bridges: Emerging Technologies with
Emphasis on High Performance Steel. Salt Lake City, Utah.
Hu, Y. 2005. Use of Adhesives to Retrofit Out-of-Plane Distortion Induced Fatigue Cracks.
PhD dissertation. University of Minnesota, Minneapolis.
ICRI. 2008. Guide for Surface Preparation for the Repair of Deteriorated Concrete Result-
ing from Corrosion. Guideline No. 310.1R-2008. International Concrete Repair Institute,
Rosemont, Ill.
Iowa Department of Transportation. Maintenance and Design of Steel Abutment Piles in
Iowa Bridges. Research needs statement. TRB, National Research Council, Washington,
D.C. 2010. http://rns.trb.org/dproject.asp?n=27091. Accessed Oct 16, 2013.
Javidi, S., A. J. Yakel, and A. Azizinamini. 2014. Experimental Investigation, Application
and Monitoring of Simple-Made-Continuous Bridge Connection for Modular Bridge Con-
struction Method. AISC Engineering Journal, Vol. 51, No. 3.
Jung, W., Y. Yoon, and Y. Sohn. 2003. Predicting the Remaining Service Life of Land Con-
crete by Steel Corrosion. Cement and Concrete Research, Vol. 33, pp. 663–677.
Jung, Y. S., D. G. Zollinger, and S. D. Tayabji. 2007. Best Practices of Concrete Pavement
Transition Design and Construction. Report No. FHWA/TX-07/0-5320-1. Texas Depart-
ment of Transportation, Austin.

532

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Kahn, L. F., and M. Lopez. 2005. Prestress Losses in High Performance Lightweight Con-
crete Pretensioned Bridge Girder. PCI Journal, September–October, pp. 84–94.
Karalar, M., and M. Dicleli. 2010. Development of a New Cycle Counting Method for
Cyclic Thermal Strains in Integral Bridge Piles. In Bridge Maintenance, Safety, Management
and Life-Cycle Optimization (R. Sause, D. Frangopol, and C. Kusko, eds.), CRC Press,
p. 638.
Kavazanjian, E., J. Wang, G. Martin, A. Shamsabadi, I. Lam, S. Dickenson, and C. Hung.
2011. LRFD Seismic Analysis and Design of Transportation Geotechnical Features and
Structural Foundations. Report FHWA-NHI-11-032. Federal Highway Administration, U.S.
Department of Transportation.
Keating, P. B. 2001. Steel Fatigue Repair. Short course materials from Effective Bridge
­Rehabilitation, University of Wisconsin–Madison.
Keating, P., S. Wilson, and T. Kohutek. 1996. Evaluation of Repair Procedures for Web
Gap Fatigue Damage. Research Report 1360-1. Texas Transportation Institute, Texas
A&M University, College Station.
Kepler, J. L., D. Darwin, and C. E. Locke. 2000. Evaluation of Corrosion Protection
­Methods for Reinforced Concrete Highway Structures. SM Report No. 58. University of
Kansas Center for Research, Lawrence.
Kiger, S. A., H. A. Salim, and A. Ibrahim. 2010. Bridge Vulnerability Assessment and Miti-
gation Against Explosions. Final report MTC Project 2007–06. Midwest Transportation
Consortium, Institute for Transportation, Iowa State University, Ames.
Knott, M. A., and O. D. Larsen. 1990. Guide Specifications and Commentary for Vessel
Collision Design of Highway Bridges. Report No. 89. Federal Highway Administration,
U.S. Department of Transportation.
Kodur, V., L. Gu, and M. Garlock. 2010. Review and Assessment of Fire Hazard in Bridges.
In Transportation Research Record: Journal of the Transportation Research Board, No.
2172, Transportation Research Board of the National Academies, Washington, D.C.,
pp. 23–29.
Kogler, R., J. Ault, and C. Farschon. 1997. Environmentally Acceptable Materials for the
Corrosion Protection of Steel Bridges. FHWA Publication No. FHWA-RD-96-058. Federal
Highway Administration, U.S. Department of Transportation.
Kogler, R., and W. Mott. 1992. Environmentally Acceptable Materials for the Corrosion
Protection of Steel Bridges. FHWA Publication No. FHWA-RD-91-060. Federal Highway
Administration, U.S. Department of Transportation.
Koob, M. J., P. D. Frey, and J. M. Hanson. 1985. Evaluation of Web Cracking at Floor
Beam to Stiffener Connections of the Poplar Street Approaches, FAI Route 70, East St.
Louis, St. Claire County, Illinois. Illinois Department of Transportation, Springfield.
Koob, M. J., and J. C. McGormley. 1998. Bridge Condition Report I-80 Spring Street
Bridge over the Missouri River, Council Bluffs, Iowa. Iowa Department of Transportation,
Ames.
Kosmatka, S. H., and W. C. Panarese. 1988. Design and Control of Concrete Mixtures.
Portland Cement Association, Skokie, Ill., pp. 151–162.
Kosmatka, S. H., and M. L. Wilson. 2011. Design and Control of Concrete Mixtures, 15th
ed. Portland Cement Association, Skokie, Ill.

533

REFERENCES
Krauss, P. D. 1996. NCHRP Report 380: Transverse Cracking in Newly Constructed
Bridge Decks. TRB, National Research Council, Washington, D.C.
Lagasse, P. F., P. E. Clopper, J. E. Pagán-Ortiz, L. W. Zevenbergen, L. A. Arneson, J. D.
Schall, and L. G. Girard. 2009. Hydraulic Engineering Circular No. 23: Bridge Scour and
Stream Instability Countermeasures: Experience, Selection, and Design Guidance, 3rd ed.
Federal Highway Administration, U.S. Department of Transportation.
Lampe, N. J., N. Mossahebi, A. J. Yakel, R. Farimani, and A. Azizinamini. 2014. Develop-
ment and Experimental Testing of Simple for Dead–Continuous for Live Bridge System.
AISC Engineering Journal, Vol. 51, No. 2.
Lawler, J. S., D. Zampini, and S. P. Shah. 2002. Permeability of Cracked Hybrid Fiber-­
Reinforced Mortar Under Load. ACI Materials Journal, Vol. 99, No. 4, pp. 379–385.
Leathers, R. C. 1990. Bridge Deck Joint Rehabilitation. FHWA Technical Advisory
T5140.16. Federal Highway Administration, U.S. Department of Transportation.
Lee, D. J. 1994. Bridge Bearing and Expansion Joints, 2nd ed. E & FN Spon, London,
pp. 52–53.
Lee, H., R. D. Cody, A. M. Cody, and P. G. Spry. 2000. Effects of Various Deicing Chemi-
cals on Pavement Concrete Deterioration. Proc., Mid-Continent Transportation Sympo-
sium, Center for Transportation Research and Education, Iowa State University, Ames.
Lee, H., and K. Neville. 1967. Handbook of Epoxy Resins. McGraw-Hill, New York.
Lehman, D. E., C. W. Roeder, R. Larson, and K. Curtin. 2003. Cotton Duck Bearing Pads:
Engineering Evaluation and Design Recommendations. Research Report WA-RD 569.1.
Washington State Transportation Center, University of Washington, Seattle.
Li, V. C. 2002. Advances in ECC Research. ACI Special Publication on Concrete: Material
Science to Applications, SP 206-23, pp. 373–400.
Li, V. C. 2003. On Engineered Cementitious Composites (ECC): A Review of the M ­ aterial
and Its Applications. Journal of Advanced Concrete Technology, Vol. 1, No. 3, pp. 215–230.
Liles, P. 2010. High Strength Lightweight Concrete for Use in Precast, Prestressed Concrete
Bridge Girders in Georgia. HPC Bridge Views, Vol. 61, May–June.
Lopez, W., J. A. Gonzales, and C. Andrade. 1993. Influence of Temperature on the Service
Life of Rebars. Cement and Concrete Research, Vol. 23, pp. 1130–1140.
Lui, D., R. A. Magliola, and K. F. Dunker. 2005. Integral Abutment Bridges: Iowa and
­Colorado Experience. Proc., FHWA Conference on Integral Abutment and Jointless
Bridges, Baltimore, Md.
Maberry, S., J. D. Camp, and J. Bowser. 2005. New Mexico’s Practice and Experience in
Using Continuous Spans for Jointless Bridges. Proc., FHWA Conference on Integral Abut-
ment and Jointless Bridges, Baltimore, Md.
Macdonald, S. (ed.) 2003. Concrete: Building Pathology, Blackwell Science, Inc., Malden,
Mass.
Malone, P. G. 1999. Use of Permeable Formwork in Placing and Curing Concrete. Engineer
Research and Development Center, Army Engineer Waterways Experiment Station, Vicks-
burg, Miss.
Manning, D. M. 1995. NCHRP Synthesis of Highway Practice 220: Waterproofing Mem-
branes for Concrete Bridge Decks. TRB, National Research Council, Washington, D.C.

534

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Markeset, G., S. Rostam, and O. Klinghoffer. 2006. Guide for the Use of Stainless Steel
­Reinforcement in Concrete Structures. Project Report 405. Nordic Innovation Centre
­Project 04118, Norwegian Building Research Institute, Oslo, Norway.
Mather, B. 1990. How to Make Concrete That Will Be Immune to the Effects of Freezing
and Thawing. In Paul Klieger Symposium of Performance of Concrete (D. Whiting, ed.),
ACI SP 126, American Concrete Institute, Farmington Hills, Mich., pp. 1–10.
Maurer Söhne. 2003. Maurer MSM Sliding Bearings. Maurer Report No. 39a-gb. Munich,
Germany.
McDonald, D. B., D. W. Pfeifer, and M. R. Sherman. 1998. Corrosion Evaluation of Epoxy-
Coated, Metallic-Clad and Solid Metallic Reinforcing Bars in Concrete. FHWA-RD-98-153.
Federal Highway Administration, U.S. Department of Transportation.
McDonald, D. B., M. R. Sherman, D. W. Pfeifer, and Y. P. Virmani. 1995. Stainless Steel
Reinforcing As Corrosion Protection. Concrete International, Vol. 17, No. 5, pp. 65–70.
McDonald, D. B., Y. P. Virmani, and D. F. Pfeifer. 1996. Testing the Performance of Copper-
Clad Reinforcing Bars. Concrete International, Vol. 18, No. 11, pp. 39–43.
McEleney, B. 2005. A Primer on Weathering Steel. Structure Magazine, a joint publica-
tion of the National Council of Structural Engineers Associations (NCSEA), Council of
American Structural Engineers (CASE), and ASCE’s Structural Engineering Institute (SEI).
October. Schumann Printers, Fall River, Wisc. http://www.spiweb.com.
Memon, A. H., and A. A. Mufti. 2004. Fatigue Behaviour of Second Generation of Steel-
Free Concrete Bridge Deck Slab. Proc., 2nd International Conference on FRP Composites
in Civil Engineering—CICE 2004, Adelaide, Australia, pp. 765–772.
Mertz, D. R., M. J. Chajes, J. W. Gillespie, D. S. Kukich, S. A. Sabol, N. M. Hawkins,
W. Aquino, and T. B. Deen. 2003. NCHRP Report 503: Application of Fiber-Reinforced
Polymer Composites to the Highway Infrastructure. Transportation Research Board of the
National Academies, Washington, D.C.
Miller, J. R. 1995. The Use of an Inorganic Copolymer Liquid Admixture to Produce
Water­tight Concrete. Proc., 2nd Regional Concrete Conference—Concrete Durability in the
Arabian Gulf, Bahrain.
Miller, R. A., R. Castrodale, A. Mirmiran, and M. Hastak, M. 2004. NCHRP Report 519:
Connection of Simple-Span Precast Concrete Girders for Continuity. Transportation
­Research Board of the National Academies, Washington, D.C.
Mirmiran, A., S. Kulkarni, R. Castrodale, R. Miller, and M. Hastak. 2001. Nonlinear
Continuity Analysis of Precast, Prestressed Concrete Girders with Cast-in-Place Decks and
Diaphragms. PCI Journal, Vol. 46, pp. 60–80.
Modjeski and Masters, Inc. 2008. Birmingham Bridge Forensic Inspection. Final report
summary. Pennsylvania Department of Transportation, Engineering District 11, Bridgeville.
Moehle, J. P., and M. O. Eberhard. 2000. Earthquake Damage to Bridges. In Bridge Engi-
neering Handbook (W.-F. Chen and L. Duan, eds.), CRC Press, Boca Raton, Fla.
Mogawer, W. S., and A. J. Austerman. 2004. Evaluation of Asphaltic Expansion Joints. The
New England Transportation Consortium, Storrs, Conn.
Moore, B., and D. Bierwagen. 2006. Ultra High Performance Concrete Highway Bridge.
Proc., 2006 Concrete Bridge Conference: HPC: Build Fast, Build to Last, Reno, Nev.

535

REFERENCES
Moulton, L. K., H. V. S. Gangarao, and G. T. Halvorsen. 1985. Tolerable Movement Crite-
ria for Highway Bridges. Report No. FHWA/RD-85/107. Federal Highway Administration,
U.S. Department of Transportation.
Naaman, A. E., and J. R. Homrich. 1989. Tensile Stress-Strain Properties of SIFCON. ACI
Materials Journal, Vol. 86, No. 3, pp. 244–251.
NACE. 2000. Impressed Current Cathodic Protection of Reinforcing Steel in Atmospheri-
cally Exposed Concrete Structures. Standard RP0290-2000. National Association of Corro-
sion Engineers International, Houston, Tex.
Naito, C., R. Sause, I. Hodgson, S. Pessiki, C. Desai. 2006. Forensic Evaluation of Pre-
stressed Box Beams from the Lake View Drive Bridge over I-70. Final report. ATLSS
Report No. 06-13. ATLSS Center, Lehigh University. Bethlehem, Pa.
NCHRP. 1977. Rapid Setting Materials for Patching of Concrete, NCHRP Synthesis of
Highway Practice No. 45. National Cooperative Highway Research Program, Transporta-
tion Research Board of the National Academies, Washington, D.C.
NCHRP. 2009. NCHRP Synthesis 393: Adjacent Precast Concrete Box Beam B ­ ridges:
Connection Details. Transportation Research Board of the National Academies,
­Washington, D.C.
Neville, A. M. 1995. Properties of Concrete, 4th ed. Pearson Prentice Hall.
Newhook, J. P., and A. A. Mufti. 1996. A Reinforcing Steel-Free Concrete Deck Slab for the
Salmon River Bridge. Concrete International, June 1996, pp. 30–34.
Nicholson, B. A., J. M. Barr, R. S. Cooke, R. P. Hickman, C. J. F. P. Jones, and H. P. J.
Taylor. 1997. Integral Bridges: Report of a Study Tour to North America. Concrete Bridge
Development Group, Blackwater, Camberley, Surrey, United Kingdom.
Niessner, M., and T. Seeger. 1999. Fatigue Strength of Structural Steel with Powder-­
Actuated Fasteners According to Eurocode 3. Stahlbau, Vol. 68, No. 11, pp. 941–948.
Nmai, C. K., S. A. Farrington, and G. S. Bobrowski. 1992. Organic-Based Corrosion-
Inhibiting Admixture for Reinforced Concrete. Concrete International, Vol. 14, No. 4,
pp. 45–51.
Nmai, C. K., and P. D. Krauss. 1994. Comparative Evaluation of Corrosion-Inhibiting
Chemical Admixtures for Reinforced Concrete. Proc., 3rd International Conference on
Concrete Durability, Nice, France, pp. 245–262.
NRC. 1977. NCHRP Synthesis of Highway Practice No. 45: Rapid-Setting Materials for
Patching of Concrete. TRB, National Research Council, Washington, D.C.
NSBA. 2008. Selecting the Right Bridge Type. In Steel Bridge Design Handbook, National
Steel Bridge Alliance.
NTPEP (National Transportation Product Evaluation Program). 2013. National Transpor-
tation Product Evaluation Program (NTPEP) Information and Operations Guide, Revised.
American Association of State Highway and Transportation Officials, Washington, DC.
http://www.ntpep.org.
Obla, K. H., R. L. Hill, M. D. A. Thomas, S. G. Shashiprakash, and O. Perebatova. 2003.
Properties of Concrete Containing Ultra-Fine Fly Ash. ACI Materials Journal, Vol. 100, No.
5, p. 426.
Oesterle, R. G. 2005. Jointless Bridges, Volume II, Analytical Research and Proposed
­Design Procedures. Final report. Federal Highway Administration, Washington, D.C.

536

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Oesterle, R. G., and H. R. Lotfi. 2005. Transverse Movement in Skewed Integral Abutment
Bridges. Proc., FHWA Conference on Integral Abutment and Jointless Bridges, Baltimore,
Md.
Oesterle, R. G., and J. S. Volz. 2005. Effective Temperature and Longitudinal Movement in
Integral Abutment Bridges. Proc., FHWA Conference on Integral Abutment and Jointless
Bridges, Baltimore, Md.
Oesterle, R. G., J. D. Glikin, and S. C. Larson. 1989. NCHRP Report 322: Design of
Precast Prestressed Bridge Girders Made Continuous. TRB, National Research Council,
Washington, D.C.
Oesterle, R. G., A. B. Mehrabi, H. Tabatabai, A. Scanlon, and C. A. Ligozio. 2004a.
­Continuity Considerations in Prestressed Concrete Jointless Bridges. Proc., 2004
­Structures—Building on the Past: Securing the Future, Structures Congress, Nashville,
Tenn., pp. 227–234.
Oesterle, R. G., A. B. Mehrabi, H. Tabatabai, A. Scanlon, and C. A. Ligozio. 2004b. Evalu-
ation of Continuity in Prestressed Concrete Jointless Bridges. Proc., 2004 Concrete Bridge
Conference, Charlotte, N.C.
Oesterle, R. G., H. Tabatabai, T. J. Lawson, T. M. Refai, J. S. Volz, and A. Scanlon. 2005.
Jointless Bridges, Volume III, Summary Report. Final report. Federal Highway Administra-
tion, Washington, D.C.
Okamura, H., and M. Ouchi. 1999. Self-Compacting Concrete: Development, Present Use
and Future. In Self-Compacting Concrete: Proceedings of the First International RILEM
Symposium (A. Skarendahl and O. Petersson, eds.), RILEM Publications, Cachan Cedex,
France.
OMB. 1992. Circular No. A-94. Revised October 29, 1992. U.S. Office of Management
and Budget, Washington, D.C.
Ozol, M. A. 2006. Alkali Carbonate Rock Reaction. Significance of Tests and Properties
of Concrete and Concrete Making Materials. ASTM STP 169D. West Conshohocken, Pa.,
pp. 410–424.
Ozyildirim, C. 2005. High-Performance Fiber-Reinforced Concrete in a Bridge Deck.
­Virginia Transportation Research Council, Charlottesville.
Ozyildirim, C. 2007. Durable Concrete for Bridges. Aspire, the Concrete Bridge Magazine,
Summer 2007, pp. 15–19.
Ozyildirim, C. 2008. Bulb-T Beams with Self-Consolidating Concrete on the Route 33
Bridge over the Pamunkey River in Virginia. Virginia Transportation Research Council,
Charlottesville.
Ozyildirim, C. 2011. Evaluation of Ultra-High-Performance Fiber-Reinforced Concrete.
Virginia Center for Transportation Innovation and Research, Charlottesville.
Padgett, J., R. DesRoches, B. Nielson, M. Yashinsky, O.-S. Kwon, N. Burdette, and E.
Tavera. 2008. Bridge Damage and Repair Costs from Hurricane Katrina. Journal of Bridge
Engineering, Vol. 13, No. 1, pp. 7–14.
Parsekian, G. A., N. G. Shrive, T. G. Brown, J. Kroman, P. J. Seibert, V. H. Perry, and A.
Boucher. 2008. Innovative Ultra-High Performance Concrete Structures. In Tailor Made
Concrete Structures, Taylor & Francis Group, London, pp. 325–330.

537

REFERENCES
Parsons, Brinckerhoff, Quade & Douglas. 2006. NCHRP Report 525, Volume 12: Making
Transportation Tunnels Safe and Secure. Transportation Research Board of the National
Academies, Washington, D.C.
PCI Bridges Committee. 2008. Reflective Cracking in Adjacent Box Girder Bridge Super-
structures. Subcommittee on Adjacent Box Beam Bridges, Precast/Prestressed Concrete
Institute, Chicago, Ill.
Pedeferri, P. 1996. Cathodic Protection and Cathodic Prevention. Construction and Build-
ing Materials, Vol. 10, No. 5, pp. 391–402.
Pemmaraju Venkata, H. P., G. Tsiatas, and K. W. Lee. 2006. The Buried Joint Approach for
Expansion Joint Retrofit: A Case Study. Presented at 85th Annual Meeting of the Transpor-
tation Research Board, Washington, D.C.
Perry, V. H., and M. Royce. 2010. Innovative Field-Cast UHPC Joints for Precast Bridge
Decks (Side-by-Side Deck Bulb-Tees), Village of Lyons, New York: Design, Prototyping,
Testing and Construction. Third fib International Congress, Washington, D.C.
Pike, R. G. 1973. Nonmetallic Coatings for Concrete Reinforcing Bars. Public Roads, Vol.
37, No. 5, pp. 185–187.
Polder, R. 1994. Electrochemical Chloride Removal of Reinforced Concrete Prisms
Containing Chloride from Sea Water Exposure. UK Corrosion and EUROCORR ’94,
­Bournemouth, United Kingdom, pp. 239–248.
Purvis, R. 2003. NCHRP Synthesis 319: Bridge Deck Joint Performance. Transportation
Research Board of the National Academies, Washington, D.C.
Qian, S., D. Cusson, R. Glazer, and T. Hoogeveen. 2002. The Performance of Corrosion-­
Inhibiting Systems in Concrete Bridge Barrier Walls: 5 Years of Field Data. National
­Research Council, Ottawa, Ontario, Canada.
Ralls, M., L. Ybanez, and J. Panak. 1993. The New Texas U-Beam Bridges: An Aesthetic
and Economical Design Solution. PCI Journal, September–October, pp. 20–29.
Ramniceanu, A., R. Weyers, M. Brown, and M. Sprinkel. 2008. Parameters Governing the
Corrosion Protection Efficiency of Fusion-Bonded Epoxy Coatings on Reinforcing Steel.
Final report. Virginia Transportation Research Council, Charlottesville.
Roeder, C. W., J. F. Stanton, and A. W. Taylor. 1990. Fatigue of Steel-Reinforced E
­ lastomeric
Bearings. Journal of Structural Engineering, Vol. 116, No. 2, pp. 407–426.
Roy, S., J. Fisher, and B. Yen. 2003. Fatigue Resistance of Welded Details Enhanced by Ultra­
sonic Impact Treatment (UIT). International Journal of Fatigue, Vol. 25, pp. 1239–1247.
Russell, H. G. 1978. Performance of Shrinkage-Compensating Concretes in Slabs.
RD057.01D. Portland Cement Association, Skokie, Ill.
Russell, H. 2004. NCHRP Synthesis 333: Concrete Bridge Deck Performance. Transporta-
tion Research Board of the National Academies, Washington, D.C.
Russell, H. G. 2011. Adjacent Precast Concrete Box-Beam Bridges: State of the Practice.
PCI Journal, Vol. 56, No. 1.
Russell, H. 2012. NCHRP Synthesis 425 Waterproofing Membranes for Concrete Bridge
Decks. Transportation Research Board of the National Academies, Washington, D.C.
Russell, H., and H. C. Ozyildirim. 2006. Revising High-Performance Concrete Classifica-
tions. ACI Concrete International, Vol. 28, No. 8, 2006, pp. 43–49.

538

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Sahmaran, M., and V. C. Li. 2010. Engineered Cementitious Composites: Can It Be
­Accepted As a Crack-Free Concrete? In Transportation Research Record: Journal of the
Transportation Research Board, No. 2164, Transportation Research Board of the National
Academies, Washington, D.C., pp. 1–8.
Schaefer, V. R., and J. C. Koch. 1992. Void Development Under Bridge Approaches. Report
No. SD90-03. South Dakota State University, Brookings.
Schindler, A. K., and D. A. Brown. 2006. Evaluation of Self-Consolidating Concrete for
Drilled Shaft Applications at the Lumber River Bridge Project, South Carolina. Research
Report No. 2. Highway Research Center, Auburn University, Auburn, Ala.
Schrank, D., T. Lomax, and B. Eisele. 2011. TTI’s 2011 Urban Mobility Report: Powered
by INRIX Traffic Data. Texas Transportation Institute, College Station.
Scully, J. R., and M. F. Hurley. 2007. Investigation of the Corrosion Propagation Character-
istics of New Metallic Reinforcing Bars. Virginia Transportation Research Council, Virginia
Center for Transportation Innovation and Research, Charlottesville.
Sergi, G. 2009. Ten Year Results of Galvanic Sacrificial Anodes in Steel Reinforced
­Concrete. EUROCORR, Nice, France.
Sherafati, A. 2011. Expanding the Length of Jointless Bridges by Providing Rotational
Capacity over the Pile Head. PhD dissertation. University of Nebraska–Lincoln.
Sherafati, A., M. Sotudeh Chafi, and A. Azizinamini. 2012. Buckling of Piles in Cohesive
Soil Supporting Jointless Bridges. Bridge Structures: Assessment, Design and Construction,
Vol. 8, No. 1, pp. 15–24.
Sherafati, A., and A. Azizinamini. 2014. Flexible Pile Head in Jointless Bridges: Experi­
mental Investigation. Journal of Bridge Engineering.
Sherman, M. R., R. L. Carrasquillo, and D. W. Fowler. 1993. Field Evaluation of Bridge
Corrosion Protection Measures. Texas Department of Transportation, Austin.
Shinozuka, M., D. Ballantyne, R. Borcherdt, I. Buckle, T. O’Rourke, and A. Schiff.
1995. The Hanshin-Awaji Earthquake of January 17, 1995: Performance of Lifelines.
­Technical Report NCEER-95-0015. National Center for Earthquake Engineering Research,
Buffalo, N.Y.
Shutt, C. A. 2006. Precast’s Durability Helps Bridges Resist Fire Damage. ASCENT
­Magazine, Spring.
Smith, L. M. 2001. Applicator Training Bulletin: Cleaning and Painting Galvanized Steel.
Journal of Protective Coatings and Linings, Technology Publishing/PaintSquare, Pittsburgh
Pa., April.
Smith, F. N., and M. Tullmin. 1999. Using Stainless Steels As Long-Lasting Rebar Material.
Materials Performance, Vol. 38, No. 5, pp. 72–76.
Smith, K. L., Peshkin, D.G., Rmeili, E.H., Dam, T.V., Smith, K.D., and Darter, M.I. 1991.
Innovative Materials and Equipment for Pavement Surface Repair – Final Report, Volume
1, Strategic Highway Research Report No. SHRP-M/UFR-91-504. National Research
Council, Washington, D.C.
Sobanjo, J. 2011. Determining Deterioration Models Using Inspection Data in Florida. Pre-
sented at 2100 Southeast Bridge Preservation Partnership Meeting, Raleigh, N.C.
Song, H. W., and V. Saraswathy. 2007. Corrosion Monitoring of Reinforced Concrete
Structures: A Review. International Journal of Electrochemical Science, Vol. 2, pp. 1– 28.

539

REFERENCES
Sprinkel, M. M. 1998. Very-Early-Strength Latex-Modified Concrete Overlay. Virginia
Transportation Research Council, Charlottesville.
Sprinkel, M. M., and C. Ozyildirim. 2000. Evaluation of High-Performance Concrete
Overlays Placed on Route 60 over Lynnhaven Inlet in Virginia. Virginia Transportation
Research Council, Charlottesville.
SSPC. 2000. Standard Method of Evaluating Degree of Rusting on Painted Steel Surfaces.
SSPC-VIS 2. Society for Protective Coatings, Pittsburgh, Pa.
SSPC. 2003. Specification for the Application of Thermal Spray Coatings (Metallizing) of
Aluminum, Zinc, and Their Alloys and Composites for the Corrosion Protection of Steel.
Joint Standard SSPC-CS 23.00/AWS C2.23M/NACE No. 12. Society for Protective Coat-
ings, Pittsburgh, Pa.
SSPC. 2004a. Surface Preparation Commentary for Steel and Concrete Substrates. SSPC-SP
COM. Society for Protective Coatings, Pittsburgh, Pa.
SSPC. 2004b. Guide for Repair of Imperfections in Galvanized or Inorganic Zinc-Rich
Coated Steel Using Organic Zinc-Rich Coating. SSPC-Guide 14. Society for Protective
Coatings, Pittsburgh, Pa.
SSPC. 2004c. Guide to Maintenance Repainting with Oil Base or Alkyd Painting Systems.
SSPC-PA Guide 4. Society for Protective Coatings, Pittsburgh, Pa.
SSPC. 2004d. SSPC Technology Update TU-3, Overcoating. Society for Protective Coat-
ings, Pittsburgh, Pa.
SSPC. 2005. Field Methods for Retrieval and Analysis of Soluble Salts on Steel or Other
Nonporous Substrates. SSPC-Guide 15. Society for Protective Coatings, Pittsburgh, Pa.
SSPC. 2006. Surface Preparation Commentary for Steel and Concrete Surfaces. Society for
Protective Coatings, Pittsburgh, Pa.
SSPC. 2008. Guide for Planning Coatings Inspection. Society for Protective Coatings, Pitts-
burgh, Pa. http://www.sspc.org/.../sspc-guide-for-planning-coatings-inspection.html/.
Stanton, J. F., C. W. Roeder, and T. I. Campbell. 1999. NCHRP Report 432: High-Load
Multi-Rotational Bridge Bearings. TRB, National Research Council, Washington, D.C.
Stanton, J. F., C. W. Roeder, and P. M. Helnwein. 2004. NCHRP 12-68: Rotation Limits
for Elastomeric Bearings: Appendix B. Final report. Department of Civil and Environ­
mental Engineering, University of Washington, Seattle.
Stanton, J. F., C. W. Roeder, P. M. Helnwein, C. White, C. Kuester, and B. Craig. 2008.
NCHRP Report 596: Rotation Limits for Elastomeric Bearings. Transportation Research
Board of the National Academies, Washington, D.C.
Stephens, R. I., A. Fatemi, R. R. Stephens, and H. O. Fuchs. 2001. Metal Fatigue in Engi-
neering. John Wiley and Son, Inc., New York.
Stoddard, R. 2002. Inspection and Repair of a Fire-Damaged Prestressed Girder Bridge.
IBC-04-17. Washington Department of Transportation, Olympia.
Suprenant, B. A. 1996. Evaluating Fire-Damaged Concrete. Publication No. R970020.
Aberdeen Group, Addison, Ill.
Tabatabai, H., R. G. Oesterle, and T. J. Lawson. 2005. Jointless Bridges, Volume I, Part 3,
Experimental Research and Field Studies. Final report. Federal Highway Administration,
U.S. Department of Transportation.

540

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Telang, N. M., and A. M. Mehrabi. 2003. Cracked Girders. Public Roads (FHWA), Vol. 67,
No. 3.
Thomas Telford Service, Ltd. 1993. CEB-FIP Model Code. Comité Euro-International du
Béton, Lausanne, Switzerland.
TRB. 2006. Transportation Research Circular E-C107: Control of Cracking in Concrete:
State of the Art. Transportation Research Board of the National Academies, Washington,
D.C.
TRB, Transportation Economic Committee. 2013. Transportation Benefit-Cost ­Analysis,
Discounting. http://bca.transportationeconomics.org/calculation-issues/discounting.
­Accessed November 11, 2013.
Tripler, A. B., E. L. White, F. H. Haynie, and W. K. Boyd. 1966. NCHRP Report 23:
Methods of Reducing Corrosion of Reinforcing Steel. HRB, National Research Council,
­Washington, D.C.
Tryfyakov, V., P. Mikheev, Y. Kudryavtsev, and D. Reznik. 1993. Ultrsonic Impact Peening
Treatment of Welds and Its Effect on Fatigue Resistance in Air and Seawater. 25th Annual
Offshore Technology Conference, Houston, Tex.
U.S. Global Change Research Program. 2009. Global Climate Change Impacts in the Unit-
ed States: A State of Knowledge Report. Cambridge University Press, Cambridge, United
Kingdom, and New York.
Valum, R., and J. E. Nilsskog. 1999. Production and Quality Control of High Performance
Lightweight Concrete for the Raftsundet Bridge. 5th International Symposium on Utiliza-
tion of High Strength/High Performance Concrete, Sandefjord, Norway.
Velivasakis, E. E., S. K. Henriksen, and D. W. Whitmore. 1997. Halting Corrosion by Chlo-
ride Extraction and Realkalization. Concrete International, Vol. 19, No. 12, pp. 39–45.
Vermeer, M., and S. Rahmstorf. 2009. Global Sea Level Linked to Global Temperature.
Proceedings of the National Academy of Sciences, Vol. 106, No. 51, pp. 21527–21532.
Virmani, Y. P., K. C. Clear, and T. J. Pasko, Jr. 1983. Time-to-Corrosion of Reinforcing Steel
in Concrete Slabs: Vol. 5, Calcium Nitrite Admixture or Epoxy-Coated Reinforcing Bars
As Corrosion Protection Systems. Federal Highway Administration, U.S. Department of
Transportation.
Virmani, Y. P., and G. G. Clemeña. 1998. Corrosion Protection: Concrete Bridges. Report
RD-98-088. Federal Highway Administration, U.S. Department of Transportation.
VTrans. 2009. Integral Abutment Bridge Design Guidelines. Vermont Agency of Transpor-
tation, Montpelier.
Wallbank, E. J. 1989. The Performance of Concrete in Bridges: A Survey of 200 Highway
Bridges. Department of Transport, HMSO, London.
Walls, J., III, and M. R. Smith. 1998. Life-Cycle Cost Analysis in Pavement Design: Interim
Technical Bulletin. FHWA-SA-98-079. Federal Highway Administration, U.S. Department
of Transportation.
Wang, K., D. C. Jansen, S. Shah, and A. F. Karr. 1997. Permeability Study of Cracked Con-
crete. Cement and Concrete Research, Vol. 27, No. 3, pp. 381–393.
Wardhana, K., and F. Hadipriono. 2003. Analysis of Recent Bridge Failures in the United
States. Journal of Performance of Constructed Facilities, Vol. 17, No. 3, pp. 144–150.

541

REFERENCES
Wasserman, E. P., and J. H. Walker. 1996. Integral Abutments for Steel Bridges, American
Iron and Steel Institute, Chicago, Ill.
Weyers, R. E., M. M. Sprinkel, and M. C. Brown. 2006. Summary Report on the Perfor-
mance of Epoxy-Coated Reinforcing Steel in Virginia. Virginia Center for Transportation
Innovation and Research, Charlottesville.
Whitmore, D. 2002. Impressed Current and Galvanic Discrete Anode Cathodic Protection
for Corrosion Protection of Concrete Structures. Conference paper. NACE International,
Houston, Texas.
Williamson, E. B., O. Bayrak, G. D. Williams, C. E. Davis, K. A. Marchand, A. E. McKay,
J. Kulicki, and W. Wassef. 2010. NCHRP Report 645: Blast-Resistant Highway Bridges:
Design and Detailing Guidelines. Transportation Research Board of the National Acad-
emies, Washington, D.C.
Wilson, A. D. 1999. Properties of Recent Production of A709 HPS 70W Bridge Steels.
International Symposium on Steel for Fabricated Structures, ASM International, Cincinnati,
Ohio.
Wilson, A. D. 2005. Improvements to High Performance Steels. Presented at the 2005
World Steel Bridge Symposium, American Institute of Steel Construction, Chicago, Ill.
Wipf, T. J., L. F. Greimann, A. H. Khalil, and D. Wood. 1998. Preventing Cracking at
­Diaphragm/Plate Girder Connections in Steel Bridges. Iowa DOT Project HR-393. Center
for Transportation Research and Education, Iowa State University, Ames.
Wolchuk, R. 1963. Design Manual for Orthotropic Steel Plate Deck Bridges. American
Institute of Steel Construction, Chicago, Ill.
Xi, Y., N. Abu-Hejleh, A. Asiz, and A. Suwito. 2004. Performance Evaluation of Various
Corrosion Protection Systems of Bridges in Colorado. Final report. CDOT-DTD-R2004-1.
Colorado Department of Transportation, Denver.
Yakel, A. J., and A. Azizinamini. 2014. Field Application Case Studies and Long Term
Monitoring of Bridges Utilizing the Simple for Dead–Continuous for Live Bridge System.
AISC Engineering Journal, Vol. 51, No. 3.
Yeomans, S. R. 2004. Galvanized Steel Reinforcement in Concrete. Elsevier, New York.
Yuen, L.H. 2005. Performance of Concrete Bridge Deck Joints. MS thesis. Department of
Civil and Environmental Engineering, Brigham Young University, Provo, Utah.
Zederbaum, J. 1969. Factors Influencing the Longitudinal Movement of Concrete Bridge
Systems with Special Reference to Deck Contraction. In Concrete Bridge Design, ACI Pub-
lication SP 23, American Concrete Institute, Detroit, Mich., pp. 75–95.
Zemajtis, J., and R. E. Weyers. 1996. Concrete Bridge Service Life Extension Using Sealers
in Chloride-Laden Environments. In Transportation Research Record 1561, TRB, National
Research Council, Washington, D.C., pp. 1–5.

542

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


A
Design Provisions for Self-
Stressing System for Bridge
Application with Emphasis on
Precast Panel Deck System

Steel girder bridges often use continuity over the interior supports to reduce interior
forces on the spans. In continuous structures with composite concrete decks, the loca-
tion of maximum negative bending moment is over the interior supports. This moment
produces tensile stresses in the concrete deck and compressive stress in the bottom
flanges of the girders. The tensile stress in the deck leads to cracking, which allows in-
trusion of moisture and road salt, causing corrosion of the reinforcement and support-
ing girders. Continued maintenance is required to forestall the deterioration; however,
replacement of the deck is eventually required.
To help alleviate this problem, a self-stressing system was developed as part of
SHRP 2 Project R19A. Additional details of the development can be found in the
forthcoming final report. The method induces a compressive force in the deck that is
accomplished by raising the interior supports above their final elevation while the deck
is cast (cast-in-place construction) or placed (precast construction). Once the concrete
has cured, the supports are lowered to their final elevation. Continuity of the steel
member and the composite action with the deck produce a compressive stress in the
concrete slab, which is balanced by tensile stresses in the bottom of the steel member.
As a result, the cracking over the interior support is reduced, increasing durability.
In addition, the need for girder splices may be eliminated, making the overall bridge
design more efficient and less expensive than a conventional design.
This appendix describes the construction procedure, design considerations, and
implementation details for using the self-stressing method. A flowchart is provided
to aid in implementation. Simplified formulas applicable to two-span bridges, which
represent the most likely use of the method, are also included.

543
A.1.
Construction Procedure Overview
This section briefly describes the major steps in the construction procedure in order to
establish a frame of reference and to introduce vocabulary used throughout the appen-
dix. Table A.1 illustrates the major steps required for constructing a bridge using the
self-stressing method. These steps will be used as points of reference in the remaining
discussion.
The first stage consists of simply placing the girder on the level supports. The
resulting moments and deflections are those obtained from a continuous beam analysis.
During the second stage, the interior support is raised. The bare steel girder
responds as a simply supported beam subjected to an upward directed point load at
the location of the interior support. Note that the supports could be in the raised posi-
tion before placing the girder. However, due to superposition, the analysis would be
the same as described.
Stage
StSt Stage Next, the concrete Loading
Structure
Structure deck is cast, or precast
Loading panels are placed
Moment
Moment and grouted. The
Deflection
Deflection
StStStStage
Stage
Stage Structureresponds likeLoading
Structure
Structure
structure continuous bare Moment
Loading
a Loading beam, just as itDeflection
Moment
Moment
steel Deflection
Deflection
would be for conven-
ee
Ste St
St e St Stage
eStageStage
Stage tional construction. Loading
Structure
Structure
Structure
Structure Loading
Loading
Loading Moment
Moment
Moment
Moment Deflection
Deflection
Deflection
Deflection
St p pStage Structure
During the fourth Loading
stage, the interior Moment
support is lowered toDeflection
its final position. Just
eep ep epStageStage
Stage Structure
Structure Loading
Loading Moment
Moment Deflection
Deflection
e StStSt
St Stage
Place
Place Structure
asStructure
in Loading
Loading
Stage 2, the response Moment
supported at theDeflection
Moment
is that of a beam Deflection
exterior supports only
ppe ep p
ep
St Stagee Place
Place
Place and subjected to a point load. However, the structure is now composite and the point
girderononStructure
girder Loading Moment Deflection
p1 p Place
PlacePlace
Place
girder
girder
1Stage on
girderon Structure
pep
St
1StSt Place
1 1St Stage on
Stage
Stage
level
level Structure
Structure
Structure Loading
Loading
Loading
Loading Moment
Moment
Moment
Moment Deflection
Deflection
Deflection
Deflection
epe e ePlace girder
girder
Place
girder
level
Place
girder
levelon
on
level
Place onon
TableSt 1 1girder
11A.1. Stage on Structure
supports
supports
  Self-Stressing Method MajorLoading
Steps Moment Deflection
p1pp pgirder level
levellevel
supports
girder
level
girder
supports
Place
girder on
onon
supportson
Step e Stage level Structure Loading Moment Deflection
111 1 Raise Raise
1 p Place2 2Place
level
girder supports
supports
supports
level
Placesupports
level
level
RaiseRaise
Raise
Place
girder
supports on
Place
12on 2 2level interior
interior
2 2girder
Raise
Raise
supports
level
Place Raise
Raise
interior onon
supports
supports
interior
interior
supports
girder girder
girder onon
1221supports
1 2 21Raise support
support
2
2 Raise level interior
interior
support
supports
RaiseRaise
level
girder interior
Raise
Raise
interior
support
support
level
level
interior on
22 2interior
2212support CastCast
supports
supportsupport
support support
interior
interior
2 3 23interior
Raise
level
Cast
support Castsupports
interior
supports
supports
Cast
232St3 3 Stage concrete Structure
concrete Loading Moment Deflection
Raise
support
interior
supports
Raise support
support
supportRaise
Raise
Cast
Cast
3 22e2Cast Cast
Cast
concrete
concrete
concrete
concrete
33 2 3 3 2Cast
3 interior support
Raise
concrete Lower
Lower
interior
interior
interior
concrete
concrete
Cast
concrete Cast
22p22 2Cast Cast
Lower
concreteLower
Lower
3334 34support interior
interior
support
interior support
support
concrete
concrete
4 4 4concrete
234support
Lower Castconcrete
Lower
Lower Lower
Lower
interior
interior
interior
Placeinterior
Lower support
support
4 4 Cast
support
concrete
Cast
CastCast
interior
interior
4 4 interior
support
interior
Lower
3433 3interior on Lower
Lower
support
Lowersupport
girder
1 5 5concrete Relaxatio
Relaxatio
concrete
5a 44Relaxation Cast
4 4interior concrete
supportconcrete
support
support
Lower support
interior
interior
interior
level
5
3 5 5support Relaxatio
Relaxatio
Relaxatio
a asupport n
n support
455a 5a 5aLower
concrete
support
interior
nsupport
Lower Lower
supports
Lower
Relaxatio
n Relaxatio
Relaxatio
Relaxatio
n
5 Relaxatio
5b 4 interior
support
545a a554Relaxatio
5aaa54Restoring Raise
nLower interior
interior
interior
n nRelaxatio
Relaxatio
nRelaxatio
2 n Restoring
Restoring
5 5 5support
force Restoring
Restoring
Restoring
support
a45aab abnRelaxatio support
nsupport
ninterior
interior nforce
force
2
55 5b 5bsupport Restoring
RestoringRestoring
55ab
Restoring
forceforce
force
nsupport
55 5Relaxatio
Restoring Relaxatio
Relaxatio
Relaxatio
5 bb55b b 5Restoring
force
force force
force
Restoring
Restoring
Restoring
544 ab 5aa anforce nn n
Cast
Relaxatio
b53bb bforce force
force
Restoring
force
a nconcrete
5bBRIDGES
DESIGN GUIDE FOR 55 5Restoring
force Restoring
Restoring
Restoring
FOR SERVICE LIFE

b5bb bforce Lower


Restoring
force force
force
4 interior
load is directed downward. This action places the concrete deck over the supports into
compression.
Over time, creep and shrinkage occur in the concrete deck. This may be accounted
for in two stages. First, the creep and shrinkage are seen as an applied curvature on
the structure. If the beam were simply supported by the exterior supports, this applied
curvature would result in additional deflection without inducing additional load.
However, because of the continuity, a restoring force is generated that prevents the
displacement and results in additional stresses.
A.2
Design Considerations
This section discusses the design issues specific to the use of the self-stressing method.
Design of bridges using the self-stressing method should follow the provisions for
I-section and box-section flexural members contained in Sections 6.10 and 6.11, respec-
tively, of the LRFD Bridge Design Specifications (LRFD specifications) (­AASHTO
2012), except as modified here.
A.2.1
General
The use of the self-stressing method is limited to straight I- and box-section steel g­ irders
and is applicable only to continuous multispan structures with a composite deck. Prac-
tical limitations dictate that the method is most likely to be used in two-span struc-
tures. Simplified design aids are provided in Section A.5 for structures with two spans.
A.2.2
Analysis
Two options provided for the analysis of the structure are described in the following
section. Note that the analysis methods should only be used when analyzing the con-
struction steps associated with the self-stressing method and not the overall analysis
procedures covered in Chapter 4 of the LRFD specifications.
A.2.2.1
Simplified Analysis
The simplified analysis method relies on first-order techniques that disregard time
­effects in the concrete. These effects are accounted for using conservative correction
factors presented in the implementation details portion of the provisions (Section A.3).
The correction factors account for the effects of creep and shrinkage in the evaluation
of stresses and deflections. As an alternative, advanced methods of analysis may be
used that directly evaluate these effects.
A.2.2.2
Advanced Analysis
Advanced methods of analysis explicitly consider the effects of creep and shrinkage to
evaluate stresses and deflections. Several examples of advanced methods are the effec-
tive modulus method, adjusted effective modulus method, step-by-step method, and
the rate of creep method.

545

Appendix A. DESIGN PROVISIONS FOR SELF-STRESSING SYSTEM FOR BRIDGE APPLICATION WITH EMPHASIS ON PRECAST PANEL DECK SYSTEM
When the creep and shrinkage strains are known, or otherwise assumed, then
LRFD specifications Section C4.6.6 can be used for calculating the resulting stresses
and deformations.
A.2.3
Forces
The forces and stresses in all components that arise as a result of the self-stressing con-
struction procedure should be considered in evaluating the load effects during design.
For the purpose of design, the locked-in prestressing force shall be considered a dead
load force applied to the composite long-term section (DC2).
LRFD specifications Section 3.4.1 states that when prestressed components are
used in conjunction with steel girders, the force effect should be considered as locked-
in construction loads. However, in this situation the prestressing forces are being
developed by gravity effects rather than applied by prestressing devices. As such, the
variability in the resulting stresses will be of the same magnitude as the variability of
the dead load effects, which leads to the decision of considering the prestress stress as
DC2 loading.
Note that the self-stressing procedure will generate tensile stresses in the bottom
of the steel girders that will serve to offset some of the compressive dead and live load
stresses. The stresses resulting from the self-stressing procedure should be kept sepa-
rate from other dead load stress sources, and the minimum load factor for dead load
(0.9) should be used.
A.2.4
Deflections
The final deflected shape is necessary for determining the camber requirements of the
girders and can be obtained by summing the deflections from the various construction
stages.
A.3
Design Procedure and Implementation Details
This section provides a step-by-step procedure for designing a bridge incorporating the
self-stressing method. All grout and/or adhesives must be adequately cured before low-
ering the interior support. The creep and shrinkage properties of the materials must be
compatible with the intended use and properties assumed during analysis.

Step 1. Determine Required Amount of Prestress


The self-stressing method is a way to introduce compressive stresses in the concrete
deck of a multispan continuous beam. The compressive stresses are generally located
near the interior supports and therefore work to counter the tensile stresses that arise
in this vicinity due to gravity and live loading. The result is a reduction in cracking and
an accompanying increase in service life. The magnitude of the prestress that must be
applied to achieve the desired effects has been determined based on past experience
with decks that have been prestressed using traditional mechanical methods.

546

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Minimum Final Prestress
The recommended minimum level of prestress at the top fiber of the concrete deck over
an interior support, after all losses, is 750 psi.
The simplified (Bernoulli assumption) analysis methods predict a linear variation
of stresses through the thickness of the deck, which produces a maximum stress value
at the face of the concrete. In practice, creep effects quickly blunt this maximum stress
value, resulting in a more uniform stress profile through the depth of the concrete. The
prescribed minimum prestress value at the face of the slab is intended to provide a final
uniform value over the top half of the slab of 250 psi, which is the value recommended
in LRFD specifications Section 9.7.5.3 for longitudinal prestressing of concrete slabs.
Figure A.1 shows the initial stress distribution in the concrete deck and the distri-
bution that develops after some period of time has elapsed.

250
Top of Slab

Final

Initial

Bottom of Slab
Figure A.1.  Stress distribution
σ in concrete deck.

Maximum Initial Prestress


The maximum initial prestress to be applied shall be no greater than 60% of the con-
crete compressive strength.
There is no upper limit recommendation in the literature because the material
maximum strength is a natural upper bound. However, to maintain a safe margin,
the upper limit shall not be greater than 60% of the concrete compressive strength
(0.6 fc′ ), which is the compressive stress limit recommended in LRFD specifications
Section 5.9.4.1.1 for pretensioned and posttensioned concrete components, including
segmentally constructed bridges.

Prestress Adjusted for Losses


In lieu of an exact analysis, the prestress loss may be conservatively estimated as 20%
when the initial prestress value is less than 40% of the concrete compressive strength,
and 30% when the initial prestress value is greater than 40% of the concrete compres-
sive strength.

547

Appendix A. DESIGN PROVISIONS FOR SELF-STRESSING SYSTEM FOR BRIDGE APPLICATION WITH EMPHASIS ON PRECAST PANEL DECK SYSTEM
Equation A.1 gives the initial prestress at the top fiber that is to be applied:
σ pf
σ pi = (A.1)
(1 − rs )

where

spf = final prestress stress,


spi = initial prestress stress, and
rs = loss due to creep and shrinkage.

Step 2. Calculate Amount of Deflection to Obtain Desired Prestress


Determine the height to which the interior support must be raised so that on release it
will provide the desired amount of prestress.
The problem at hand is essentially that of support settlement. How far must the
interior support settle so that the stress in the top of the deck is the value chosen in the
previous design step?
For the following steps (a to d), the structure to be considered is a composite struc-
ture being supported at the exterior supports only, as shown in Figure A.2.

a. Determine the stress at the top fiber of the deck due to point loading applied
at the interior support location.
b. Use the result from (a) to solve for the magnitude of the forces required to
produce the desired prestress determined in Step 1.
c. Calculate the stiffness with respect to point load applied at the interior sup-
port location.
d. Use the stiffness from (c) to solve for the displacement required to produce the
necessary force.

P
Composite

L1 L2

Figure A.2.  Equivalent structure used for calculating stresses during lowering of support.

548

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


For the structure shown in Figure A.2, this displacement (δ) is given by Equation A.2:

σ ts L1L2
δ= (A.2)
3Econc cts
where
sts = initial prestress stress,
L1 = length of Span 1,
L2 = length of Span 2,
Econc = modulus of elasticity of concrete, and
cts = distance from neutral axis to top fiber of slab.

Step 3. Determine Forces Due to Lifting Bare Steel Beam


The results obtained from this step are used to complete the constructability check of
the structure. For the following steps, the structure to be considered is the bare steel
beam being supported at the exterior supports only, as shown in Figure A.3.

L1 L2

Bare Steel

P
Figure A.3.  Equivalent structure used for calculating stresses during the raising of support.

a. Calculate the stiffness with respect to point loads applied at the interior sup-
port locations.
b. Use the stiffness from (a) to calculate the force required to lift the interior sup-
ports to the height determined in the previous design step. For the structure
shown in Figure A.3, this force is given by Equation A.3:

3Esteel Isteel ( L1 + L2 )
P= δ (A.3)
L21L22

where
P = reaction at support due to deflection of support,
d = deflection of support,
L1 = length of Span 1,
L2 = length of Span 2,
Esteel = modulus of elasticity of steel, and
Isteel = moment of inertia of bare steel girder.

549

Appendix A. DESIGN PROVISIONS FOR SELF-STRESSING SYSTEM FOR BRIDGE APPLICATION WITH EMPHASIS ON PRECAST PANEL DECK SYSTEM
c. Using the force given by (b), the reactions, moments, and stresses can be
calculated as needed for design.

The steel girders and any temporary or permanent support structures must be
designed for the concentrated forces that are developed as a result of lifting the girders.

End Anchorages
The calculated vertical displacement may require a lifting force that is greater than the
self-weight of the steel girder, such that the girder would lift off the end supports. In
this situation, the exterior ends of the girder may be anchored to prevent uplift. Once
the concrete deck is in place, the weight of the deck will replace this anchorage force.
Note that loading within the spans can affect uplift at the end supports. Consider
the structure shown in Figure A.4. Loading in the first span will create uplift at the
end support of the opposite span. Therefore, the progression of deck casting or pre-
cast panel placement may affect the need for end anchorages. This possibility must be
properly accounted for through either design or the specification of explicit procedures
to avoid the condition described.

L1 L2

Bare Steel
R1 R2 R3

Figure A.4.  Loading in Span 1 producing uplift at Support R3.

Equation A.4 gives the reaction at the end of Span 2 (unloaded span) caused by the
following combination of loading:

• Self-weight of the steel girder (wsteel),


• An upward displacement of the interior support (δ), and
• Uniform load within Span 1 due to deck placement (wdeck).

Equation A.4 will aid in evaluating the need and magnitude of end anchorages. The
critical condition occurs when Span 1, the loaded span, is longer than Span 2. When the
spans are different lengths, the deck within the shorter span should be cast first.


(
wsteel 3L22 + L1L2 − L21 ) − 3E
steel Isteel
δ−
wdeck L31
(A.4)
8L2 L1L22 8L2 ( L1 + L2 )

550

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


where
wsteel = uniform load due to self-weight of steel,
wdeck = uniform load due to deck placement,
d = deflection of support (positive upward)
L1 = length of Span 1,
L2 = length of Span 2,
Esteel = modulus of elasticity of steel, and
Isteel = moment of inertia of bare steel girder.

For the case of two equal spans (L1 = L2 = L), Equation A.4 can be simplified to
Equation A.5:

3Lwsteel 3Esteel Isteel Lwdeck (A.5)


− 3
δ−
8 L 16
where L is the length of Spans 1 and 2 (equal).

End anchorages, when necessary, must be designed to withstand the concentrated


force that is to be applied.

Step 4. Determine Forces and Stresses Caused by Lowering


Composite Bridge
The forces and stresses imparted on the structure caused by lowering the composite
bridge are obtained from a similar analysis to that performed when the amount of
deflection was originally calculated in Step 3.
For the following steps, the structure to be considered is the composite structure
being supported at the exterior supports only, as shown in Figure A.5.

PReduced
Composite

L1 L2

Figure A.5.  Equivalent structure used for calculating stresses during lowering of support.

a. Calculate the stiffness with respect to a point load applied at the interior sup-
port location.
b. Use the stiffness from (a) to calculate the equivalent point force caused by the
lowering of the support.
c. Reduce the force calculated in (b) to account for the prestress loss due to creep
and shrinkage, as determined in Step 2.
551

Appendix A. DESIGN PROVISIONS FOR SELF-STRESSING SYSTEM FOR BRIDGE APPLICATION WITH EMPHASIS ON PRECAST PANEL DECK SYSTEM
d. Using the reduced force applied to the composite structure supported at the ex-
terior supports, calculate the internal forces and stresses necessary for design.

The resulting forces and stresses from this step should be considered dead load
forces applied to the composite structure for the purpose of design.

Step 5. Determine Deflected Shape


The final deflected shape is necessary for determining the camber requirements of the
girders. The final deflection is the summation of deflections from the various construc-
tion stages.

Bare Steel Deflection


Sources of deflection of the bare steel girder are

• Self-weight of steel;
• Initial lift of interior supports; and
• Casting of wet concrete.

The deflection due to the self-weight of the steel and casting of the wet concrete is
calculated in a conventional manner by using the continuous bare steel structure, as
shown in Figure A.6. Equations for calculating the deformation along the length of the
beam can be found in Section A.5.

L1 L2

Bare Steel
R1 R2 R3

Figure A.6.  Structure for calculating bare steel deflections.

Calculation of the deflection caused by the initial lift of the interior support is
determined considering the bare steel girder supported at the exterior supports only, as
shown in Figure A.7. The structure is subjected to point forces applied at the interior
supports as determined in Step 3. Equations for calculating the deformation along the
length of the beam can be found in Section A.5.

552

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


L1 L2

Bare Steel

P
Figure A.7.  Structure for calculating composite deflections caused by lowering the support.

Composite Deflection
Calculation of the deflection caused by the lowering of the interior support is deter-
mined by considering the composite bridge girder supported at the exterior supports
only, as shown in Figure A.8. The structure is subjected to point forces applied at the
interior supports as determined in Step 3 without the reduction in load meant to ac-
count for creep and shrinkage. Creep and shrinkage have the opposite effect, resulting
in an increase of the total deflection. This effect is discussed in the following section.
Equations for calculating the deformation along the length of the beam can be found
in Step 2.

P
Composite

L1 L2

Figure A.8.  Structure for calculating bare steel deflections caused by initial lifting of support.

Relaxation Deflection
Additional deflections arise due to curvature induced along the beam due to the effects
of creep and shrinkage. The resulting loading can be seen in Figure A.9.

553

Appendix A. DESIGN PROVISIONS FOR SELF-STRESSING SYSTEM FOR BRIDGE APPLICATION WITH EMPHASIS ON PRECAST PANEL DECK SYSTEM
cr

sh

Composite

L1 L2

Figure A.9.  Curvature applied to continuous structure due to creep and shrinkage.

The steps for calculating the deflected shape can be performed using the following
steps, considering the structure supported at the exterior supports only, as shown in
Figure A.10.

cr

sh
Composite

L1 L2

Figure A.10.  Structure for determining restoring force.

a. Calculate the stiffness with respect to a point load applied at the interior sup-
port location.
b. Determine the curvature along the length of the beam. The curvature at a
section can be obtained from Equation A.6. LRFD specifications Section
5.4.2.3.1 provides methods for determining the values of esh and ecr.

554

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


1
ϕ=
Ic ∫ (ε sh + ε cr ) z dz (A.6)
where
j = curvature of section,
Ic = composite moment of inertia,
esh = strain due to shrinkage,
ecr = strain due to creep, and
z = distance from neutral axis.

c. Calculate the displaced shape of the structure due to the applied curvature, as
shown in Figure A.11. The displacement can be calculated using the integra-
tion given in Equation A.7.

cr

sh
Composite

L1 L2

Figure A.11.  Structure for determination of deflection due to curvature.

x x
δ ( x) = ∫0 ∫0 ϕ ( x) dx dx (A.7)
where j(x) is the curvature along the length of the beam.

d. Using the stiffness from (a), determine the force required to offset the displace-
ment at the support location calculated in (c).
e. The resulting deflection due to the relaxation is the sum of the deflections ob-
tained from the applied curvature (Equation A.6) and the application of the
point load determined in (d) on the structure shown in Figure A.10.

Step 6. Carry Out Remainder of Design


The remainder of the design proceeds as it would for a conventional steel girder bridge
with concrete deck.

555

Appendix A. DESIGN PROVISIONS FOR SELF-STRESSING SYSTEM FOR BRIDGE APPLICATION WITH EMPHASIS ON PRECAST PANEL DECK SYSTEM
A.4
Design Flowchart
Figure A.12 provides a design criteria flowchart for the conventional design method
and the self-stressing design method.

Design Criteria

Determine bridge
geometry and
dimensions.

Conventional Self-stressing
design method design method

Calculate
Choose level of
loads/stress due to
compressive stress.
girder weight.

Calculate Determine
load/stress due to amount of
deck weight. displacement.

Check Determine force Determine force


constructability. to anchor needed to raise
(LRFD specifications) girder ends. bridge.

Calculate load/stress Calculate


due to time- load/stress due to
dependent effect. lifting.

Calculate
Calculate load/stress
load/stress due to
due to live load.
lowering.

Service Strength
limit state limit state
(LRFD specifications) (LRFD specifications)

Figure A.12.  Design criteria flowchart.

556
Figure A.12. Design flowchart.
DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE
A.5
Design Aids for Two-Span Bridges
The following figures offer design aids for continuous beam two-span bridges. Figures
A.13 and A.14 are for bridges with equal spans, and Figures A.15 and A.16 are for
bridges with unequal spans.

7
R1 = V1 ............................................. = wl
16
5
R2 = V2 + V3 ...................................... = w l
8
1
R3 = V3 ............................................. = wl
16
9
V2 .................................. ................... = wl
16
7 49
M max at x = l ............................ = wl2
16 512
1
M 1 (at support R2 ) ............................ = w l 2
16
wl 2
M x (when x < l ) ............................... = (7 l 8 x )
16
Figure A.13.  Continuous beam—two equal spans—uniform load on one span.

3 wl
R1 = V1 = R3 = V3 ....................................... =
8
10 wl
R2 ............................................................. =
8
5 wl
V2 = Vmax .................................................... =
8
wl 2
M 1............................................................ =
8
3l 9 wl2
M 2 at ................................................ =
8 128
wl 4
max (at 0.4215 l , approx.from R1 and R3 ).. =
185 EI

Figure A.14.  Continuous beam—two equal spans—uniformly distributed load.

557

Appendix A. DESIGN PROVISIONS FOR SELF-STRESSING SYSTEM FOR BRIDGE APPLICATION WITH EMPHASIS ON PRECAST PANEL DECK SYSTEM
wl1 M1
R1 = V1................................. =
2 l1
R2 ........................................ = w l1 R1 R3
M1
R3 ........................................ =
l2
V2 ......................................... = w l1 R1
V3 ......................................... = R3
wl23
M 1....................................... =
8(l1 + l2 )
R1 w x2
M max when x = ........... = R1 x
w 2

Figure A.15.  Continuous beam—two unequal spans—uniformly distributed load on one span.

M 1 wl1
R1 = V1................................. = +
l1 2
R2 ........................................ = w l1 + w l2 R1 R3
M 1 wl2
R3 = V4 ................................. = +
l2 2
V2 ......................................... = w l1 R1
V3 ......................................... = w l2 R3
wl + wl13
3
M 1....................................... = 2

8(l1 + l2 )
R1 w x12
M x1 when x1 = ............ = R1 x1
w 2
R3 w x22
M x2 when x2 = ............ = R3 x2
w 2

Figure A.16.  Continuous beam—two unequal spans—uniformly distributed load.

558

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


B
Displacement of
Skewed Bridge

This appendix summarizes the work for addressing the effect of skew on lateral move-
ment of the bridge at the abutment.
B.1
Background
A skewed bridge is a bridge with the longitudinal axis at an angle other than 90° with
the piers and abutments. The skew angle (q) is shown in Figure B.1. With skewed inte-
gral abutment bridges, the soil passive pressure developed in response to thermal elon-
gation has a component in the transverse direction, as illustrated in Figure B.1. Within
certain limits of the skew angle, soil friction on the abutment will resist the transverse
component of passive pressure. However, if the soil friction is insufficient, then, de-
pending on the transverse stiffness of the abutment, either significant transverse forces
or significant transverse movements could be generated.
Figure B.2 shows a two-span bridge with a skew angle of 45° (Nicholson et al.
1997). This bridge was constructed in 1969 with semi-integral abutments. The semi-
integral construction included an integral end diaphragm that was designed to move
with the superstructure, which slides longitudinally and is guided transversely by rela-
tively stiff abutments.
Figure B.3 shows cracking in the abutment wall near an acute corner of the super-
structure, presumably caused by transverse forces related to soil pressures.
Figure B.4 shows distress in an asphalt overlay at the skewed end of an approach
slab as a result of transverse movement (Tabatabai et al. 2005).
Figure B.5 shows a closer view of the barrier wall joint from Figure B.4 at the end
of the approach slab. The expansion joint in the barrier wall was made perpendicular
to the longitudinal direction and could not accommodate the transverse movement.

559
Figure B.1.  Components of abutment soil passive pressure
response to thermal elongation in skewed integral abutment
bridges.

Figure B.2.  Two-span semi-integral abutment bridge with an overall length of 89 m
(259 ft), width of 11.6 m (38 ft), and a skew angle of 45°.
Source: Nicholson et al. 1997.

560

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure B.3.  Cracking in the abutment wall near an
acute corner of the superstructure for the bridge shown
in Figure B.2.
Source: Nicholson et al. 1997.

Figure B.4.  Asphalt overlay distress (west end).


Source: Tabatabai et al. 2005.

561

Appendix B. DISPLACEMENT OF SKEWED BRIDGE


Figure B.5.  Barrier distress at
west abutment.
Source: Tabatabai et al. 2005.

Because of potential problems and uncertainty related to the response of skewed


integral abutments, many state departments of transportation limit the skew angle.
A typical limit for maximum skew angle for integral abutment bridges used by many
states is 30°. However, maximum skew angle limits in various states range from 0°
to no limit (Chandra 1995). In response to the potential problems of skewed inte-
gral abutments, studies on jointless and integral abutment bridges were conducted by
FHWA (Oesterle and Lotfi 2005) to

1. Develop a relationship between skew angle and abutment soil friction for limiting
skew.
2. Develop a relationship for the magnitude of forces required to restrain transverse
movement in integral abutment bridges with large skew angles.
3. Develop a relationship between skew angle and expected transverse movement
for a typical integral stub abutment with no special design features to restrain this
movement.
4. Compare analytical results with field data for a skewed bridge that was monitored
as part of the experimental portion of this project.
5. Perform a sensitivity study to demonstrate the relationship between transverse
movement and longitudinal expansion for various skew angles and ratios of bridge
length to width.

This work was accomplished by developing equilibrium and compatibility equa-


tions for end abutment forces and, for the case of a typical stub abutment, solving
these relationships for various skew angles and bridge length-to-width ratios.

562

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


B.2
Analyses for Transverse Response to Thermal Expansion
B.2.1
Skew Angle Limit for Limiting Transverse Effects
Figure B.6 shows the passive soil pressure response (Pp) due to thermal expansion and
soil–abutment interface friction (Faf), assuming no rotation in the plane of the bridge
superstructure. Equation B.1 shows Faf for rotational equilibrium:

Faf ( L cos θ ) = Pp ( L sin θ ) (B.1)

and Equation B.2 shows Faf from interface behavior:

Faf = Pp tan δ (B.2)

where tan d is the friction coefficient for the interface of formed concrete and soil.

Equation B.3 is obtained by substituting Equation B.1 into Equation B.2:

sin θ
tan δ = = tan θ or (B.3)
cos θ

δ =θ

Therefore, the bridge superstructure can be held in rotational equilibrium until


the skew angle exceeds the angle of interface friction. Integral abutments are typically
backfilled with granular material. NCHRP Report 343 lists a friction angle of 22° to
26° for formed concrete against clean gravel, gravel sand mixtures, and well-graded
rock fill (Arsoy et al. 2002). Based on these data, the angle of θ = 20° represents a
reasonably conservative skew angle limit below which special considerations for trans-
verse forces or transverse movement are not needed.

Figure B.6.  Soil pressure load (Pp ) and


soil–abutment interface friction (Faf ).
563

Appendix B. DISPLACEMENT OF SKEWED BRIDGE


With larger skew angles, the integral abutment either can be designed to resist
the transverse force generated by the soil passive pressure in an attempt to guide the
abutment movement to be predominantly longitudinal, or it can be detailed to accom-
modate the transverse movement.
B.2.2
Forces Required to Resist Transverse Movement
Adding lateral resistance of the abutment (Fa) to wall–soil interface friction (Faf) in
Figure B.6, rotational equilibrium is found by Equation B.4:

(F
a + Faf ) (L cos θ) = P (L sin θ) (B.4)
p

Equation B.5 is obtained by substituting from Equation B.2 into Equation B.4:

Fa = Pp ( tan θ − tan δ ) (B.5)

Fa is the summation of abutment lateral resistance from pile and passive pressure on
the substructure surface perpendicular to the abutment.
Figure B.7 shows the relationship between Fa and Pp, assuming the interface fric-
tion angle (d) to be 20º. As shown in Figure B.7, the force required to resist transverse
movement is a significant portion of the soil passive pressure (Pp). Pp is not necessarily
full passive pressure, but can be determined for the end movement by using relation-
ships calculated by Clough and Duncan (1991; also Barker et al. 1991) shown in
Figure B.6. The end movement to consider in calculating passive pressure is the end
movement normal to the abutment (Dln).

Figure B.7.  Relationship between force required for


abutment lateral resistance (Fa ) and passive pressure
response (Pp ) to restrain lateral movement.

564

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


As illustrated in Figure B.8, this end movement is given by Equation B.6:

∆ln = ∆l cos θ (B.6)

where Δl is the maximum expected end movement for thermal reexpansion from the
starting point of full contraction for the full range of effective bridge temperatures, as
discussed in Section 8.6.2.3.1.
From Figure B.8, it can be seen that Δln is reduced with respect to Δl as the skew
angle (θ) increases. This relationship helps offset the increase in Fa /Pp with increasing
θ. However, Fa will still be a sizable portion of Pp.

Figure B.8.  Relationship between end normal movement (Dln) and end thermal
expansion (Dl).

For relatively short bridges or bridges in locations with small effective temperature
ranges, it may be feasible to design the abutment substructure to resist Fa . However, it
should be understood that for whatever means used to develop Fa (battered pile and/
or lateral passive soil resistance), lateral movements are required to develop the resis-
tance. Therefore, details anticipating some transverse movement should be used. The
expected movements are a function of the relative stiffnesses of response for Pp and
Fa. Adding battered piles to an integral abutment for lateral loading will also increase
the stiffness in the longitudinal direction, which induces more demand on the super­
structure and connections between the girders and abutments.

565

Appendix B. DISPLACEMENT OF SKEWED BRIDGE


B.3
Expected Transverse Movement with Typical Integral
Abutment
B.3.1
Method of Analyses
To investigate the relationship between skew angle and expected transverse movement
for a typical integral stub abutment, a set of relationships was derived based on equi-
librium and the compatibility of end abutment forces in the plane of the bridge super-
structure. For this analysis, the superstructure is assumed to act as a rigid body with
rotation b about the center of the deck (for a longitudinally symmetrical bridge). The
rotation occurs to accommodate the thermal end movement (Δl). Forces considered
in response to this movement include soil pressure on the abutment and wingwalls,
wall–soil interface friction on the abutment, and pile forces normal to and in line with
the abutment and wingwalls. Details of the forces, stiffness, and equations of compat-
ibility and equilibrium are provided in the final report on FHWA’s analytical work on
jointless bridges (Oesterle 2005).
A spreadsheet program was used for solving the rotational equilibrium in the plane
of the deck. For a given end thermal movement (Δl), the equilibrium position can be
found using an iterative analysis by progressively increasing the rotation angle (b) until
the sum of the in-plane moments is zero.
B.3.2
Results of Analyses for Instrumented Bridge
As part of the experimental program for studying jointless bridges (Tabatabai et al.
2005), a heavily skewed bridge in Tennessee was instrumented and monitored for
1 year. This bridge carries U.S. Interstate 40 over Ramp 2B in Knox County, ­Tennessee.
It has a three-span steel-plate girder superstructure with an overall length of 415.92 ft
and integral abutments. This structure is sharply skewed, with a skew angle of 59.09º.
The three span lengths are 139.83, 208, and 68.08 ft. The bridge was instrumented to
monitor the longitudinal and transverse movements of the east abutment and obtain
an indication of restraint to the longitudinal expansion.
The east abutment was analyzed using the spreadsheet program developed to solve
for rotational equilibrium (Oesterle 2005; Oesterle and Lotfi 2005). On the basis of
the experimental data, an end movement of Δl = 0.781 in. was used in the analysis. A
measured superstructure rotation angle of β = 0.000224 radians corresponded with
Δl = 0.781 in. Using the spreadsheet to determine rotational equilibrium, an angle of
β = 0.000226 radians was calculated. The calculated value indicated very good agree-
ment with the measured data.

566

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


B.3.3
Sensitivity Analyses for the Effects of Skew Angle on Transverse
Movement and Longitudinal Restraint
To demonstrate the effects of skew angle on expected transverse movement and lon-
gitudinal restraint forces, further analyses were carried out using the spreadsheet pro-
gram. Variables included skew angle and the length-to-width ratio for the bridge. The
abutment for the instrumented bridge is a relatively typical type of stub abutment used
in Tennessee and was used as the baseline abutment for the analyses (Oesterle 2005;
Oesterle and Lotfi 2005).
The instrumented bridge is relatively wide compared with the length. The ratio
of length to width (L/W) for this bridge is 3.15. To demonstrate the sensitivity to the
bridge L/W ratio, the analyses were repeated for abutments reduced to 2/3W and 1/3W.
The length of the wingwalls at each skew angle was kept constant.
Results of these analyses for the ratio of transverse movement to longitudinal
movement (Δt1/Δl) are shown in Figure B.9 for Δl = 1 in. The transverse movement (Δt1)
is the transverse movement of the acute corner of the bridge deck. This is the corner
that experiences the greatest transverse movement because of the skew angle.
The results in Figure B.9 demonstrate the increase in the transverse movement
with increasing skew angle. The data in Figure B.9 also demonstrate the increase
in transverse movement with decreasing L/W. The change in L/W was accomplished

Figure B.9.  Relationship between transverse movement at the acute corner (Dt1) and ­thermal
expansion (Dl) for an expansion of 1 in. with constant length bridge (L = 415.92 ft) and
varying L/W.
567

Appendix B. DISPLACEMENT OF SKEWED BRIDGE


in the analyses by decreasing the width and keeping the length constant. The length
of the wingwall at each skew angle was constant; therefore, the results in Figure B.9
demonstrate the effects of increasing the ratio of the length of the wingwalls to the
length of the abutment wall. The data in Figure B.9 show that increasing the wingwall
length relative to the abutment wall length (which includes increasing the number of
wingwall piles relative to the number of abutment wall piles) can significantly decrease
the transverse movement. However, the wingwalls and abutment must be designed to
transmit the wingwall forces into the superstructure.
Figure B.10 shows the resulting total longitudinal restraint force for these analyses
and demonstrates the decrease in longitudinal restraint with increasing skew angle.
For the full-width bridge with L/W = 3.15, the longitudinal restraint at a skew angle
of θ = 60° is approximately 60% of the longitudinal restraint at θ = 25°. For the larger
L/W = 9.45, the ratio of longitudinal restraint at θ = 60° is approximately 70% of the
restraint at θ = 25°. This change demonstrates the increase in restraint resulting from
the increase in resistance to lateral moment because of the larger ratio of wingwall
length to abutment length.

Figure B.10.  Relationship between resultant longitudinal restraint force and skew angle for thermal
expansion (Dl) of 1 in. with constant length bridge (L = 415.92 ft) and varying L/W.

568

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


B.3.4
Design Recommendations
Because the baseline abutment used in these analyses is a relatively typical stub abut-
ment (but also relatively deep, with an abutment height of 13.0 ft and with strong
axis pile bending for movement normal to the abutment versus weak axis bending
for movement parallel to the abutment), the data in Figure B.9 represent a reasonably
large estimate for the transverse movement of skewed abutments. Although there is
significant uncertainty for actual soil and pile stiffness, the maximum expected end
movement (Δl) discussed in Section 8.6.2.3.1 includes a multiplier to account for un-
certainty. Therefore, it is suggested that the data in Figure B.9 can be used by designers
to determine an approximate estimate for expected transverse movement in skewed
integral abutments resulting from the restraint of longitudinal thermal expansion. In
addition, the relationships between longitudinal restraint force and skew angle shown
in Figure B.10 can be used to estimate the relative decrease in restraint forces in a
skewed bridge. The transverse movements can be used to estimate the transverse forces
on the wingwall resulting from passive soil load and pile and to estimate longitudinal
and transverse movements for the abutment pile for biaxial bending considerations.
All the other components of movement and forces can be determined from Δt1 and Δl
by using equations presented in the full analytical report (Oesterle 2005).

569

Appendix B. DISPLACEMENT OF SKEWED BRIDGE


C
Design of Piles for
Fatigue and Stability

This appendix provides steps that could lead to development of design aid for piles
subjected to axial load and lateral movement. The principal steps are explained and
are customized for development of design aids for 50 ksi steel H-piles.
C.1
Estimation of Maximum Allowable Strain
Seasonal and daily temperature fluctuations subject steel H-piles in jointless bridges to
cyclic loading that can result in fatigue failures of the H-piles. This possibility is espe-
cially important as the magnitude of cyclic strain exceeds elastic limits. The seasonal
and daily temperature fluctuations subject H-piles to one large annual cycle (due to
seasonal temperature change) and a number of smaller load cycles (due to daily tem-
perature loading) (Dicleli and Albhaisi 2004; Karalar and Dicleli 2010). Figure C.1
shows typical H-pile cyclic strain (Dicleli and Albhaisi 2004).
This strain is the maximum longitudinal strain in steel H-piles, typically located at
the point of fixity below the pile head.
The following steps outline one alternative for predicting the fatigue life of steel
elements subjected to variable amplitude cyclic loading. The steps involve concepts of
cycle counting and use of damage models for keeping track of accumulated damages
due to cycling loading (Gere and Goodno 2012).

1. Obtain the loading history to which the steel element is subjected.


2. Develop an S-N type curve for the material under consideration. In general, in
the low-cycle regime (yearly seasonal cycle when the steel element is subjected
to strain exceeding elastic limits), the data should be presented in terms of strain
versus number of cycles to failure.

570
Figure C.1.  Pile strain as a function of
time.
Source: Dicleli and Albhaisi 2004.

3. Use a cycle counting method, such as the rain flow method (ASTM 1049-85), to
convert the variable amplitude loading into equivalent constant amplitude loading.
4. Use a damage model, such as Miner’s rule, to determine the time to failure.

Dicleli and Albhaisi (2004) suggest using Equation C.1 for relating strain ampli-
tude to fatigue life:

( )
m
ε a = M 2N f (C.1)

where
ea = constant strain amplitude,
Nf = fatigue life (number of cycles to failure) corresponding to ea,
M = factor determined from experimental testing (0.0795), and
m = exponent determined from experimental testing (–0.448).

Dicleli and Albhaisi (2004) suggest using Miner’s rule as a damage model for steel
H-piles. Equation C.2 expresses Miner’s rule:
n
n
∑ Ni ≤ 1 (C.2)
i =1 i

where ni is the number of cycles associated with the loading number i, and Ni is the
number of cycles to failure for the same case.
Dicleli and Albhaisi (2004) assume that steel H-piles are subjected to two constant
amplitude loadings, one corresponding to seasonal temperature changes and another
representing daily temperature changes. Therefore, Miner’s rule can be expanded as
shown by Equation C.3:

ns n
+ l = 1 (C.3)
N fs N fl

571

Appendix C. DESIGN OF PILES FOR FATIGUE AND STABILITY


In Equation C.3, ns and ni are the number of small and large amplitude strain cycles
due to temperature changes during the service life of the bridge, respectively, and Nfs and
Nfl are the total number of cycles to failure for the corresponding small and large ampli-
tude strain cycles, respectively. According to Dicleli and Albhaisi (2004), for 100 years of
service life, the number of small amplitude cycles is ns = 14,800, and the number of large
amplitude cycles is nl = 100. These values were obtained by studying the field perfor-
mance of several jointless bridges and developing the types of data shown in Figure C.1.
For small and large amplitude loading, Equation C.1 can be customized, as shown
by Equations C.4 and C.5, respectively (Dicleli and Albhaisi 2004):

( )
m
ε as = M 2N fs (C.4)

( )
m
ε al = M 2N fl (C.5)

To facilitate development of an “allowable” strain to be used in selecting a steel


pile capable of meeting the fatigue requirement, the small strain amplitude (eas) is
related to the large strain amplitude (eal) by using β, a proportionality constant, result-
ing in the relationship shown by Equation C.6:

ε as = βε al (C.6)

In Equation C.6, β is estimated to be 0.25 (Karalar and Dicleli 2010). By substitut-


ing Equation C.6 into Equation C.4 and solving for constant amplitude life to failure,
Equation C.4 and C.5 could then be used to determine Equations C.7 and C.8 (Dicleli
and Albhaisi 2004):
1
1  βε  m
N fs =  al  (C.7)
2 M 
1
1ε m
N fl =  al  (C.8)
2 M 
By substituting Equations C.7 and C.8 into Equation C.3 and solving for eal , the
maximum large amplitude strains that the pile can sustain without fatigue failure can
be obtained as shown by Equation C.9 (Dicleli and Albhaisi 2004):

 
 
 2n s 2nl 
ε al = + 1  (C.9)
1
 β m  1 m 
     
M M
Substituting the previously stated values for the parameters in Equation C.9, which
are ns = 14,800, nl = 100, b = 0.25, M = 0.0795, and m = –0.448, eal is then determined
to be 0.002967.

572

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Based on the calculated maximum large strain amplitude of eal = 0.002967, the
maximum cyclic curvature amplitude Yf at fatigue failure of the pile is expressed by
Equation C.10:
2ε al
ψf = (C.10)
dp

where dp is the width of the pile in the direction of the cyclic displacement.
Knowing the cross section of the steel pile to be used, complete nonlinear moment
curvature characteristics of the pile can be developed. From this relationship, the max-
imum moment that a steel pile can sustain without failure can be estimated using the
maximum “allowable” curvature, as obtained from Equation C.10. The maximum
moment that a steel pile can sustain can then be used to obtain the maximum lat-
eral displacement that the steel pile can accommodate without fatigue failure. The
maximum lateral displacement is obtained through a nonlinear pushover analysis as
described in the next section.
C.2
Pushover Analysis Example
The development of the design aids required conducting a static pushover analysis.
Static nonlinear pushover analysis using the finite element software SAP2000 can be
used to estimate the maximum lateral displacement capacity of steel H-piles based on
fatigue consideration. Only two sections (HP10x57 and HP12x84) meet the compact
ductility requirements for A36 and A50 steels, as described in Section 8.6.2.4.2. These
two cross sections were used for pushover analysis.
C.2.1
Soil–Pile Interaction Model
For the purpose of a pushover analysis, the p-y curve for piles driven in clay can be
simplified as a bilinear curve, as shown in Figure C.2.

Figure C.2.  Actual and modeled p-y curves for clay.


Source: Dicleli and Albhaisi 2004.
Load per unit length, (P)

573

Appendix C. DESIGN OF PILES FOR FATIGUE AND STABILITY


In this figure, the ultimate response Pu is estimated as shown by Equation C.11:

Pu = 9Cu d p (C.11)

where Cu is the undrained shear strength of the clay, and dp is the pile width.

The elastic modulus of the clay soil can be estimated as shown by Equation C.12:

9Cu
Es = (C.12)
5ε 50
where ε50 is the soil strain at 50% of ultimate soil resistance.

Table C.1 lists the corresponding values of Cu and e50 for different consistencies of
clay soil.

Table C.1. Representative Values of Cu and e50


Consistency of Clay Cu (psi) e50

Soft 2.9 0.020


Medium 5.8 0.010
Stiff 17.4 0.005

C.2.2
Description of the Model
To conduct a pushover analysis, the pile was modeled using SAP2000 and divided into
small beam elements, each 1 ft in length. For the purpose of the analysis, a 40-ft length
of pile was modeled for soft- and medium-density clays. The models show that this
length is sufficient to provide a relative fixed condition in the lower portion of the pile.
The pile tip is restrained from movements in all directions.
The soil response to lateral deflection was modeled using nonlinear link elements
placed every foot. The load deflection properties of the link elements were defined
based on the p-y curve described in Chapter 8.
Material properties were assumed to be 36-ksi steel for the pile section. Nonlinear
beam elements with the capability of developing hinges at both ends were used in the
pushover analysis. The properties of these hinges were defined on the basis of the ori-
entation and the level of axial load on the pile.
For a given axial load in the pile, soil condition, and steel section, a pushover anal-
ysis was then performed to obtain the maximum lateral displacement capable of meet-
ing the fatigue limit. On the basis of the assumptions made, the maximum moment that
a pile can sustain without fatigue failure was established. This maximum moment is
used in pushover analysis to establish the maximum lateral displacement. Results of
the pushover analyses for various axial loads are shown in Figure C.3 and Figure C.4.

574

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


C.2.3
Results of the Analyses
Using the described method and by performing pushover analyses, the maximum dis-
placement that steel H-piles with a specified minimum yield strength of 50 ksi can ac-
commodate has been estimated and is shown in the Figures C.3 and C.4.
These figures can be used to determine the maximum lateral displacement that a
pile can sustain, based on fatigue considerations.
An interesting aspect of the data shown in these figures is that piles oriented with
bending about the strong axis provide larger displacement (up to four times). Many
departments of transportation orient the steel piles about the weak axis of bending,
based on the logic that it could provide larger lateral displacement. The results shown
in the figures contradict this belief.

(a) (b)
Figure C.3.  Lateral displacement capacity of compact HP sections in soft clay (c = 2.9 ksi)
(a) HP10x57 (b) HP12x84.

(a) (b)
Figure C.4.  Lateral displacement capacity of compact HP sections in medium clay (c = 5.8 ksi)
(a) HP10x57 (b) HP12x84.

575

Appendix C. DESIGN OF PILES FOR FATIGUE AND STABILITY


D
Restraint Moments

This appendix provides methods of estimating the restraining moment developed in


prestressed girders when girders are made continuous over supports, as well as meth-
ods to mitigate the problem.
D.1
Background
In simple-span noncomposite bridges, time-dependent deformations result in little or
no change in the distribution of forces and moments within the structure. However,
continuous multispan composite bridges are statically indeterminate. As a result, in-
elastic deformations that occur after construction will generally induce statically inde-
terminate forces and restraining moments in the girders.
Sources of inelastic deformation include concrete creep and shrinkage and temper-
ature gradients. For example, a common type of jointless bridge construction consists
of precast, prestressed girders connected with a continuous cast-in-place deck slab, as
illustrated in Figure D.1. The girders are simply supported for dead load but may be
considered continuous for live load. Continuity is established with deck steel as nega-
tive moment reinforcement over the piers. Commonly, a positive moment connection
is also provided in the diaphragms.
It has long been recognized that positive secondary moments develop in the con-
nection at piers of continuous prestressed concrete bridges when the deck is cast at
a relatively young girder age (Freyermuth 1969). Creep of the girder concrete under
the net effects of prestressing and self-weight will tend to produce additional upward
camber with time. The piers prevent this upward movement. When girders are made
continuous at a relatively young age, it is possible that positive moments will develop
at the supports over time, as shown in Figure D.2.

576
Deck Reinforcement M M = Positive Restraint
Moment

Positive Reinforcement

Figure D.1.  A typical precast prestressed bridge simply supported for dead load and made
continuous for live load.

M M

Figure D.2.  Restraint against upward movement, positive secondary moment.

Conversely, differential shrinkage, with the newer deck slab concrete shrinking
more than the girder concrete, causes the continuous structure to bow download. Dif-
ferential shrinkage has a tendency to reduce the positive moment due to creep or result
in negative secondary moments at the supports.
In addition to creep and shrinkage of concrete, temperature gradients can play a
major role if the girders are made continuous. Solar heating of the top deck will tend
to produce upward camber, adding to the positive restraint moment caused by creep.
Large restraining positive moment can cause cracking in the bottom flange near the
pier locations. Heat of hydration in the cast-in-place deck concrete can have a miti-
gating effect on the development of positive restraint moment. The cast-in-place deck
may be heated to a temperature that is higher than the supporting girder temperature
by heat of hydration during the initial hydration when the concrete is still plastic.
Contraction of the deck concrete with subsequent cooling after the concrete has hard-
ened results in a downward deflection, thereby reducing the positive restraint moment
caused by creep and solar heating.
NCHRP and FHWA funded an experimental and analytical research program
on the behavior of continuous and jointless integral abutment prestressed concrete
bridges with cast-in-place deck slab (Oesterle et al. 1989, 2004a, 2004b). Results of
the analytical studies (Oesterle et al. 2004a) showed that the age of the girder when
the deck was cast was the most significant factor in determining whether positive or
negative restraint moments occurred at the interior transverse joints over the piers in
response to the interaction of creep and shrinkage. Results of analytical and experimen-
tal research (Oesterle et al. 1989, 2004a, 2004b) indicated that the live load continuity

577

Appendix D. RESTRAINT MOMENTS


of the bridge may be reduced significantly with long-term and time-dependent loading
effects and with thermal effects.
In the experimental part of the jointless bridge research (Oesterle et al. 2004a,
2004b), testing of materials, bridge components, and a full-scale girder indicated that

1. Expected shrinkage of the deck concrete did not occur in the concrete in the out-
door environment of Skokie, Illinois. Thus, the effects of deck shrinkage to miti-
gate the effects of girder creep did not occur.
2. Heat of hydration effects in the cast-in-place deck concrete can have a mitigating
effect on the development of positive restraint moment.
3. Daily temperature effects of heating and cooling of the deck with respect to the
girder have a significant effect on restraint moments. Solar heating of the deck
causes positive restraint moments of the same order of magnitude as the moments
caused by girder creep and are additive to the moments caused by creep.
4. Tests on a full-scale girder that was monitored and loaded periodically with simu-
lated live load on sunny days and cloudy days during different seasons over an
18-month time frame demonstrated that positive restraint moment and the re-
sulting cracking at the transverse connection significantly reduced continuity for
live load. Using change in beam reactions under application of live load to assess
continuity, the lowest measured percentage of full live load continuity was 48%
measured on a cloudy day in summer.
5. Continuity induces restraint moments, and effective continuity requires assessment
considering all loads. Effective continuity in the test girder was assessed using the
distribution of total reactions supporting the test girder, which included the ef-
fects of dead load, live load, and restraint moments. Effective continuity is defined
as 100% if the distribution of total reactions corresponds to the combination of
simply supported dead load reactions plus fully continuous live load reactions.
Effective continuity is 0% if the distribution of total reactions corresponds to the
combination of simply supported dead load reactions and simply supported live
load reactions. The measured effective continuity in two of the live load tests in the
jointless bridge study was negative (i.e., less than 0%). That is, the total midspan
positive moment in the tested “continuous” girder was slightly higher than the
anticipated positive moment if the girder were a simply supported girder for both
dead load and live load.
6. The positive moment due to combined creep and temperature effects in the test
girder resulted in stresses in the positive moment reinforcement in the connection
over the pier that reached or exceeded yield stress.

The results of this research indicated that use of a positive moment connection in
the diaphragms is not beneficial in determining the net resultant midspan service-level
stresses under dead, live, and restraint loads. Without a positive moment connection
at the supports, effects that would tend to produce a positive restraint moment (creep
in the prestressed girders and solar heating of the deck) will likely cause a crack to
form at the bottom of the diaphragm concrete between the ends of the girders. With

578

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


application of live load that would tend to produce a negative moment at the sup-
port, the crack at the bottom of the diaphragm concrete has to close before full nega-
tive moment develops. The net effect is the loss of some live load continuity, which,
depending on the parameters, can range from 0% to 100% of live load continuity. If
effects that would tend to produce a positive restraint moment are large enough, the
crack at the bottom of the diaphragm can remain open under live load, and the girder
acts as if it is simply supported.
If a positive moment connection is provided, a crack will still likely form at the
bottom of the diaphragm concrete from effects that tend to cause positive restraint
moment. The positive moment connection will decrease the crack width, but a positive
restraint moment will develop. The positive restraint moment superimposed on the
live load negative moment will negate, at least in part, the beneficial effects of the nega-
tive moment continuity connection over the piers (for service load stresses). S­ tudies
(Oesterle et al. 1989, 2004a, 2004b; Mirmiran et al. 2001) have shown that the effect
of the crack at the bottom of the diaphragm that would form without the positive
moment connection is essentially equivalent to superposition of a positive restraint
moment that would form if a positive moment connection were provided (assuming
the amount of positive moment reinforced provided was not excessive).
If effects that tend to cause negative restraint moments in the connection over
the supports predominate, positive moment reinforcement is not needed. Therefore,
these studies indicated that there is no net benefit, in terms of service-level stresses in
the prestressed girder, by providing positive moment reinforcement in the transverse
connections. It is understood, however, that there may be benefit in terms of structural
integrity for providing the positive moment reinforcement.
The results of recently completed NCHRP Project 12-53 are included in NCHRP
Report 519 (Miller et al. 2004). This project was carried out to further examine the
behavior of simple-span precast, prestressed girders made continuous by connections
at the transverse joints over the piers. The focus was on the effectiveness of the posi-
tive moment connection and on design criteria for this connection. Results of ana-
lytical studies (Mirmiran et al. 2001) were similar to those reported in the previous
NCHRP study (Oesterle et al. 1989). That is, if positive restraint moments develop,
these restraint moments must be added to the moments caused by dead and live load,
and the net positive moment at the midspan is essentially independent of the amount
of positive moment reinforcement provided in the transverse connection (assuming
the amount of positive moment reinforcement provided is not excessive). In addition,
analytical studies indicated that cracking in the transverse joint decreases live load
continuity.
NCHRP Project 12-53 also included experimental studies. Live load testing indi-
cated that, contrary to analyses results, the continuity with application of live load
was near 100% unless the positive moment crack at the connection became very large.
The full-scale testing result in the NCHRP 12-53 study, with essentially no live load
continuity lost due to positive moment cracking, differed from the analytical results
in the NCHRP studies (Oesterle et al. 1989; Mirmiran et al. 2001) and the result of
full-scale testing in the jointless bridge study (Oesterle et al. 2004a, 2004b). However,

579

Appendix D. RESTRAINT MOMENTS


live load continuity in the NCHRP 12‑53 study was assessed using change in reactions
with application of live load. It is not clear how restraint moment present in the test
specimen connection was considered.
A reason provided in NCHRP Report 519 for the difference between the analytical
studies and the experimental studies was that the observed positive moment cracks did
not extend into the top flange until the crack was very large, but that in the analytical
model, the crack extends into the top flange as soon as it forms. In the NCHRP 12-53
experimental beams, however, the effects of concrete creep were simulated by apply-
ing posttensioning near the bottom flanges after the diaphragm concrete was cast.
Posttensioning rods were dead-headed at the ends of the girders on each side of the
diaphragm and used to apply a relatively concentrated load near the bottom flanges at
the end of the girders. The additional compressive strain due to the posttensioning was
intended to simulate the creep strain in the girders due to the pretensioned prestress
and produce simulated positive moment cracks in the bottom of the diaphragm con-
crete. Applying the posttensioning forces concentrated near the bottom at the ends of
the girders, however, may have distorted the plane of the ends of the girders so that the
change in crack width over girder depth did not simulate an expected positive moment
crack in an actual bridge. Experimental tests in the jointless bridge study (Oesterle et
al. 2004a, 2004b) were carried out with full-scale girders with positive moment cracks
in the diaphragm that were primarily the result of actual long-term creep in the girders
due to the original pretensioned prestress combined with temperature gradient caused
by actual solar heating.
Several results from the NCHRP 12-53 full-scale tests were similar to those
observed in the jointless bridge study, including

1. The shrinkage strains in the deck concrete were significantly less than expected.
2. The effects of heat of hydration in the deck concrete were significant.
3. Daily thermal effects were significant.

On the basis of the analyses and testing, recommendations for the positive moment
connection in NCHRP Report 519 included

1. The positive moment connection should be provided and designed for the cal-
culated moment due to dead, live, and restraint moment. At least minimum re-
inforcement should be provided for a moment equal to 0.6 Mcr, where Mcr is the
cracking moment of the connection. Also, the design moment should not exceed
1.2 Mcr because providing additional reinforcement is not effective. If the design
moment exceeds 1.2 Mcr, the design parameters should be changed. The easiest
change to reduce the positive moment is to specify a minimum age of the girder at
the time of making the continuity connection.
2. If the contract documents specify that the girders are a minimum age of 90 days
when continuity is established, the restraint moment does not have to be calcu-
lated. This is based on the observation from surveys and analytical work that if the
­girders are more than 90 days old when continuity is formed, it is unlikely that time-­
dependent positive restraint moments from concrete creep and shrinkage will form.

580

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


3. The transverse connection can be considered fully effective if “the calculated stress
at the bottom of the continuity diaphragm for the combination of superimposed
permanent loads, settlement, creep, shrinkage, 50% live load and temperature
gradient, if applicable, is compressive.”

Results presented in NCHRP Report 519 were used to provide extensive and
comprehensive revisions and additions to Article 5.14.1.4 (Bridges Composed of
Simple Span Precast Girders Made Continuous) in the fourth edition of the LRFD
Bridge Design Specifications (LRFD specifications) (AASHTO 2007). Based on Article
5.14.1.4.1, the connections between girders should be designed for all effects that
cause moments at the connections, including restraint moments from time-dependent
effects. Note that although restraint moment due to thermal gradient is not specifi-
cally mentioned in Article 5.14.1.4.1, it should be included. However, Article 5.14.1.4
includes the following two exceptions regarding the need to design for the restraint
moments:

1. Per Article 5.14.1.4.1, multispan bridges composed of precast girders with conti-
nuity diaphragms at interior supports that are designed as a series of simple spans
are not required to satisfy Article 5.14.1.4.
2. Per Article 5.14.1.4.4, if contract documents require a minimum girder age of at
least 90 days when continuity is established, then
a. Positive restraint moments caused by girder creep and shrinkage and deck slab
shrinkage may be taken as zero,
b. Computation of restraint moments shall not be required, and
c. A positive moment connection shall be provided as specified in Article
5.14.1.4.9.

D.2
Design Recommendations
This section provides various alternatives for handling the positive moment developed
in continuous prestress girders.
D.2.1
Restraint Moments in Prestressed Concrete Girders
In general, it is recommended that LRFD specifications Article 5.14.1.4 should be
followed in the design of jointless bridges constructed with precast prestressed girders
made continuous for live load. However, the further considerations discussed in this
section should be taken into account.
D.2.1.1
Thermal Effects
Calculated thermal gradient stress caused by the combined internal restraint and sec-
ondary continuity moments can be very high, particularly when combined with other
secondary effects (Oesterle et al. 2004a, 2004b). NCHRP Report 519 states that daily
thermal effects were significant and mentions that they should be considered in de-
sign. However, results of analyses and example calculations included in the report
581

Appendix D. RESTRAINT MOMENTS


to demonstrate that restraint moment is near zero if the girder age is at least 90 days
when continuity is established did not include the effects of thermal gradient. Also,
­although the commentary to LRFD specifications Article 5.14.1.4.2 mentions temper-
ature variation as a cause of restraint moments, Article 5.14.1.4 does not specifically
address design considerations for thermal effects. It is commonly considered that ther-
mal effects are self-limiting for strength limit states and can generally be disregarded.
However, prestressed girders also have to be designed for service-level and thermal
stresses in continuous prestressed concrete bridges.
D.2.1.2
Differential Shrinkage Effects
The results of the FHWA jointless bridge project indicated that expected shrinkage
based on theoretical shrinkage models and on laboratory shrinkage tests did not occur
in the outside environment. NCHRP Report 519 (Miller et al. 2004) included a similar
observation; however, analyses and example calculations included in NCHRP Report
519 to demonstrate that restraint moment is near zero if the girder age is at least
90 days when continuity is established did include the effects of differential shrinkage
as determined from a theoretical shrinkage model. Results of the analyses presented in
the report show that early negative moment due to differential shrinkage between the
deck and the girder essentially offset the longer-term positive moment that developed
due to creep in the prestressed girder.
D.2.1.3
Combined Creep, Shrinkage, and Thermal Effects
The effects of creep in the prestressed girders and solar heating of the deck are additive
with respect to inducing positive moment at the connection over the supports. When
creep and solar heating are combined with an absence of differential shrinkage, it is
not clear, even in bridges constructed with 90-day-old girders, that positive moments
will not be significant.
D.2.1.4
Potential Negative Moment
Limiting construction to the use of girders with a minimum age of 90 days will in-
crease the potential that factors that induce negative restraint moments over the sup-
ports may predominate. Increasing the potential for negative moment increases the
risk of cracking in the deck over the support regions. Deck cracking over the support
regions may have a more detrimental effect on long-term durability of a bridge than
positive moment cracking in the diaphragm.
D.2.1.5
Uncertainties in Determining Restraint Moments
In addition to concrete creep, shrinkage, and solar heating of the deck, various other
effects can contribute significantly to restraint moments. These effects include differen-
tial settlement of supports; heat of hydration of the deck concrete during construction;
variation of the coefficient of thermal expansion between the girder and the deck; and
seasonal moisture changes in the concrete that cause shrinkage reversals. In addition,
in jointless bridges with integral abutments, additional forces may be imparted on the

582

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


positive moment connection by the restraint of the abutment to longitudinal tempera-
ture movements. All of these factors contribute to restraint forces within a continuous
jointless bridge structure. In some instances, these factors are additive, but in others,
they oppose one another. The magnitudes of these effects to be considered in design
and the critical combinations are uncertain. Although methods are available to esti-
mate restraint moments due to all of these effects, the moments that actually occur
may be significantly different from the estimated values.
D.2.1.6
Effects of Excessive Positive Moment Reinforcement
In spite of all the uncertainties regarding magnitudes and combinations of restraint
moments, there have been few cases of distress related to these secondary stresses.
In general, concrete cracking and reinforcement yielding will diminish the stresses
caused by the secondary effects. However, an overly strong connection combined with
the effects of creep and thermal gradient may result in excessive positive restraint
­moment (ENR 1994; Alabama DOT 1994; Telang and Mehrabi 2003). A strong posi-
tive ­moment connection increases the positive moment along the span and in some
cases may result in cracking in the beams. Figure D.3 shows an example bridge (Telang
and ­Mehrabi 2003) with significant flexural cracking of this type. The flexural crack
­occurred at the end of the embedment of the positive moment connection bars near the
ends of the prestressed girders with a large quantity of positive moment reinforcement.
In contrast, Figure D.4 shows the end region of another girder in the same example
bridge where cracking and spalling occurred within the diaphragm. The diaphragm
cracking and spalling were associated with positive moment connection bars bent out
of place during erection (because of constructability issues) for several girders in the
bridge such that the connection bars became ineffective. However, no flexural cracking
occurred within the span of these girders.

Figure D.3.  Cracks near girder supports.

583

Appendix D. RESTRAINT MOMENTS


Figure D.4.  Crack and spall at diaphragm over pier support.

Because of the uncertainty associated with calculations of positive continuity


moments resulting from the variability of the creep and shrinkage effects, tempera-
ture gradient, differential coefficient of expansion effects, locked-in heat of hydration
effects, settlement, and cracking, calculations to determine restraint moments are com-
plex and probably unreliable. To eliminate the need to attempt to calculate restraint
moments and to simplify the design, the following recommendations (Options 1 and
2 below) for positive moment connections were developed on the basis of the work in
the NCHRP projects (Oesterle et at. 1989; Mirmiran et al. 2001; Miller et al. 2004),
the FHWA jointless bridge project (Oesterle et al. 2004a, 2004b), and the LRFD speci-
fications (AASHTO 2007).
D.2.1.6.1
Option 1 for Positive Moment Connections
Positive moment connection reinforcement at the piers should not be provided. This
approach prevents the development of significant positive restraint moments in the
pier diaphragms (and eliminates constructability issues with the overlapping reinforce-
ment). The girders should be analyzed as simply supported for dead plus live loads
at service levels. This practice is allowed by LRFD specifications Article 5.14.1.4.1;
it eliminates the requirement to calculate restraint moments (without the need to age
girders prior to construction); and, as stated in the commentary of the LRFD specifi-
cations, it has been used successfully by several state departments of transportation.
D.2.1.6.2
Option 2 for Positive Moment Connections
If positive moment connections are used to improve structural integrity and to pro-
vide some crack control, as recommended in the commentary of LRFD specifications
Article 5.14.1.4.1, it is suggested that the positive moment capacity (fMn) be limited
to the minimum moment of 0.6 Mcr recommended in the LRFD specifications. Note

584

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


that Mcr should be determined using the properties of the diaphragm concrete. If addi­
tional reinforcement is used to increase crack control, the upper limit recommended
by the LRFD specifications of fMn  = 1.2 Mcr should not be exceeded. To eliminate the
need for calculation of restraint moments, the girders should be analyzed as simply
supported for dead plus live loads at service levels as allowed by LRFD specifica-
tions Article 5.14.1.4.1. However, positive restraint moments are likely to occur. In
spite of this, additional stresses in the girders due to positive restraint moment can be
minimized by limiting the capacity of the connection fMn so that the connection acts
like a fuse that will yield before the development of detrimental stresses. Therefore,
the girder service load stresses should be checked along the length of the girder under
simple supported dead and live loads plus fMn of the positive moment connections
superimposed on the spans, such that the allowable tensile stress in the bottom of the
beam of 0.19 fc′ ksi (6 fc′ psi) is not exceeded. Particular attention should be paid
to the region of termination of the positive moment steel if mild reinforcement is used
for the connection.
For both Options 1 and 2, the girder–diaphragm interface should consider details
to allow relative movement between the bottom of the girder and diaphragm concrete
for girders partially embedded in the diaphragm concrete. For the exterior surface of
fascia girders, providing a sealed crack control joint at the beam–diaphragm interface
should be considered.
Negative moment reinforcement should be provided over the supports, and dia-
phragm concrete should be provided between the ends of the girder bottom flanges.
Negative restraint moments may develop, for example when the deck and diaphragms
are cast when the concrete girders are older. However, parametric studies carried out
in the FHWA jointless bridge project indicate that, with high restraint moments, crack-
ing occurs in the deck, and sufficient moment redistribution occurs to prevent the
deck reinforcement from becoming overstressed. Therefore, restraint moments do not
have to be calculated. Negative moment reinforcement in the deck can be designed
for applied dead and live load moments calculated on the basis of uncracked section
properties. It can be assumed that the girder is simply supported for dead load and
fully continuous for live and superimposed dead loads because of the parapets, barrier
walls, wear surface, and so forth. Because the deck in the negative moment region is
considered reinforced concrete, the negative moment connection is only designed for
strength limit states.
D.2.2
Restraint Moments in Composite Steel Bridge Girders
Temperature gradients and differential coefficients of thermal expansion in continu-
ous composite steel beams produce both positive and negative restraint moments, but
the shrinkage of deck concrete and the heat of hydration locked-in strains produce
negative restraint moments. Deck slab cracking partially relieves negative restraint
moments.

585

Appendix D. RESTRAINT MOMENTS


The parametric studies in the FHWA jointless bridge project indicate that stresses
in both the concrete deck slab and steel beams are not excessive under the combination
of dead and live load forces combined with positive restraint moments. Consequently,
explicit calculations considering positive restraint moments are not necessary.
The analyses for effects of negative restraint moments in composite steel beams
indicated that, in general, if negative moments are high, deck cracking results in
re­
distribution, and calculated stresses are not excessive. However, analyses also
included the effects of a negative temperature gradient, which produces negative
restraint moments, combined with dead and live load and restraint of longitudinal
expansion provided by passive pressure in backfill and the lateral force in the piles of
integral abutments. These analyses indicated that, under certain circumstances, calcu-
lated compressive stresses in the bottom flange of the steel beams near interior sup-
ports may be excessive, even after allowance for redistribution of the stresses because
of deck cracking. On the basis of the parametric studies, the combination of loads
described above may become critical for larger beam spacing. Calculations indicate
that, for stringer spacing equal to or greater than 7 ft for A36 beams and 9 ft for A572,
Grade 50 beams, an explicit check of the effects of the combined load effects of dead
and live loads, negative temperature gradient, and restraint of longitudinal expansion
may be required to check for lateral torsional buckling of the bottom flange near inte-
rior supports.

586

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


E
Design Steps for
Seamless Bridge
System Developed by
SHRP 2 Project R19A

Expansion joints are one of the main causes for high maintenance costs in bridges. A
new seamless bridge system was envisioned within SHRP 2 Project R19A that should
result in bridges with long service lives by eliminating the joints over the entire length
of the bridge, approach slab, and a segment of the roadway (Ala and Azizinamini
in press a and b). The system is similar to a system developed in Australia for use with
continuously reinforced concrete pavements (CRCP) (Bridge et al. 2005). Proposed
modifications have been made to the Australian system to adapt it to U.S. practice, in
which most pavements are either jointed plain concrete pavement (JPCP) or flexible
pavement (Ala 2011). Although pavement within a particular roadway may be jointed
or flexible, the segment of roadway containing the bridge and the proposed seamless
transition is similar in nature to CRCP. Therefore, transition details would be similar
to those used when transitioning from CRCP to jointed or flexible pavements.
The key factor is establishing an effective longitudinal force transfer mechanism
from the transition slab to the base soil that minimizes the length of the transition. The
goal is achieving limited end movements, a predictable and controlled crack pattern,
and controlled axial forces in the system.
The system developed to meet these needs is shown in Figure E.1. The transition
slab is connected to a secondary slab that is embedded below. The two slabs are con-
nected by a series of small piles. The secondary slab increases the stiffness of the transi-
tion region, resulting in the desired short transition length. A similar system without
the transition slab may lose its effectiveness after multiple cycles as a result of the
compaction of the soil surrounding the small piles.
A special reinforcement reduction detail is used over the length of the transition
zone to achieve a controlled crack pattern when the bridge system is in tension. The
system behavior in tension (temperature reduction–bridge contraction) is an important
factor because the crack pattern plays a major role in design life and maintenance

587
costs. Figure E.2 shows a transition in which the reinforcement detail helps to main-
tain the desirable crack pattern (Jung et al. 2007). The reinforcement is reduced over
the length of the transition region as the force is reduced.

Bridge Road pavement


Transition
Approach Zone
Transition Zone

Soil-nails

Embeded
Slab

Transition Zone
Bridge Approach

JPCP
Abutment Small Piles

Secondary Slab

Figure E.1.  Schematic and rendering of the recommended practice for bridge–roadway interface.

Saw Cuts or Induced Design Crack CRC Pavement


JC Pavement

100% Steel Zone 60% Steel Zone 30% Steel Zone


Transition Transition
Figure E.2.  Gradual transition continuously reinforced to jointed pavement.
Source: Jung et al. 2007.

588

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Analysis, design, and construction of a seamless bridge and approach slab system
are similar to other bridge structures. However, some new components are involved
in the system that are not typically seen in other bridge systems. The new compo-
nents include a transition slab, secondary slab, small piles, and the connection of the
4.0"
small piles to the
5/8"transition concrete slabs. Figure PL16x16x1.0"
Both Flange
E.3 shows an example from the
6.0" 4.0"

2.0"
4.0" CL

3.0"
4.0"
5/8" PL16x16x1.0" 6.0" 4.0"

6.0"
Both Flange

16"
2.5" W10x49 2.5"

2.0"
4.0" CL

3.0"
3/8"

6.0"

16"
2.5" W10x49 2.5"
16"
58.0"
3/8"

16"
58.0"

Figure E.3.  Small piles to be used in the test to connect the upper and lower slabs.

589

Appendix E. DESIGN STEPS FOR SEAMLESS BRIDGE SYSTEM DEVELOPED BY SHRP 2 PROJECT R19A
developmental phase of the concept of the small piles used to connect the transition
slab to the secondary slab.
The initial system design is an iterative process in which the length of the transition
and secondary slabs; the shape, size, and spacing of small piles; and the embedment
depth of the secondary slab are determined via structural analyses of various sys-
tem configurations. Demand in all components is determined during the initial design
phase. The various parts of the system may then be designed according to the applica-
ble LRFD Bridge Design Specifications (LRFD specifications) (AASHTO 2012). The
reduced cracked stiffness of the system in tension may be neglected in the initial design.
Relevant pavement design loadings are the longitudinal strains (thermal effects,
creep, and shrinkage) and the out-of‑plane effects caused by traffic wheel loads,
­settlement of approach embankments, and rotational effects transferred from the
bridge deck.
E.1
Structural Analysis
Until further research is completed to develop a simplified analysis approach, the seam-
less bridge should be analyzed as a holistic system with all components incorporated
into the analysis. To account for the effect of temperature changes in design of the
transition region of the seamless bridge system, only the effect of uniform temperature
change needs to be considered. The calculation of uniform temperature change should
be in accordance with LRFD specifications Article 3.12.2. The interaction between
the soil and the small piles can be modeled in the structural analysis by using springs.
The spring stiffness around the small piles highly depends on the relative density of the
compacted soil material (geomaterial) surrounding the small piles and the confinement
pressure. Because the soil material is manually compacted, the relative density of the
compacted soil needs to be measured during the compaction process, and this compac-
tion should be related to the soil stiffness. The connection of the small piles to the slabs
can be assumed rigid for analysis purposes.
The structural analysis should take into account the effects of longitudinal stiffness
reduction due to cracking of the transition slab in tension (temperature reduction–
bridge contraction). Iterative structural analyses of the seamless bridge and roadway
system in conjunction with cracked section analyses are required. For the first itera-
tion, the tensile forces in the structure are assumed equal to the compressive forces due
to thermal expansion. Cracked section analyses are carried out for various segments
of the transition slab, and the axial stiffness of the slab segments are modified. The
structure is analyzed with the modified in‑plane stiffness to determine the in‑plane ten-
sile axial forces in the system. This process is repeated until convergence of the axial
forces is achieved.
E.2
Design of System Components
Once the initial system design has been completed, design of the individual system
components can be performed.

590

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


E.2.1
Approach Slab and Bridge Deck
Extra reinforcement may be required in the approach slab for crack control under
the tensile in‑plane forces resulting from thermal expansion. The bridge deck also has
to be checked for cracking. The approach slab should be checked for compressive
­thermal stresses to avoid concrete crushing. The approach slab should be designed for
the differential settlement of the bridge abutment and the transition system. Embank-
ment settlement is another important criterion to check for the approach slab. The
bridge approach embankments should be designed to achieve a long‑term settlement
of less than 3/4 in. to minimize traffic comfort issues on the motorway pavement. To
account for the probable geotechnical and construction variations, however, a more
conservative approach embankment settlement of 1.5 in. should be assumed for the
seamless pavement design (Thomas Telford Service Ltd. 1993).
E.2.2
Transition Slab
The main objective of providing a transition slab is to provide a means for controlling
the movement of the system to the point at which no expansion devices are needed
where the transition slab meets the pavement. The design items for the transition slab
include designing against compressive force created by thermal expansion; achieving a
uniform cracking pattern in the transition zone during contraction, preventing punch-
ing shear failure at the pile-to-slab connection area; and ensuring adequate flexural
capacity at these locations. The thickness of the transition slab should be determined
on the basis of (1) the punching shear requirements, (2) the connection requirements
for developing the moment introduced from the small piles, and (3) the in‑plane hori-
zontal stiffness of the system to reduce the movement of the end joint. Reinforce-
ment of the transition slab should be determined from cracked section analysis ­under
­tensile in‑plane forces. The transition slab should be checked for the maximum bend-
ing ­moments between the rows of small piles. Stirrups (tie bars) may be required for
the connection to the slab around the ends of the small piles. The transition slab
should also be designed for the design truck-axle load exerted at the midspan of the
slab between the small piles. Both slabs should be designed for punching shear and
one‑way shear. Detailed design provisions are provided in VTrans (2009) and also Ala
and Azizinamini (in press a).
E.2.3
Secondary Slab
The length of the secondary slab should be greater than or equal to the length of the
transition slab. Likewise, the secondary slab thickness is designed for punching shear
and requirements to develop the moment introduced from the small piles into the slab
(the secondary slab thickness will most likely be equal to the transition slab). The sec-
ondary slab should also be designed for the bending moment due to the soil pressure
underneath. This slab should be designed for punching and one‑way shear.

591

Appendix E. DESIGN STEPS FOR SEAMLESS BRIDGE SYSTEM DEVELOPED BY SHRP 2 PROJECT R19A
E.2.4
Small Piles
The stiffness, number, and arrangement of small piles connecting the transition and
secondary slabs should be determined to control the longitudinal movement of the
transition slab at the end where it meets the pavement. This limit eliminates expansion
joint devices at these locations. Increasing the stiffness of the small piles will reduce the
longitudinal movement of the transition slab at the end of transition zone. However,
piles with high flexural stiffness will also create high stresses (tension or compression)
in the transition slab and bridge deck, in addition to the secondary slab. Therefore, the
design of small piles should consider a balance between longitudinal movement at the
end of transition slab and the maximum longitudinal force that can be accommodated
in the transition slab and bridge deck. Small piles with high stiffness will demand more
sophisticated connection details to the transition and secondary slabs. The maximum
longitudinal movement at the end of the transition slab, where it meets the pavement,
should be limited to about 0.25 in.
E.2.5
Connection of Small Piles to Slabs
The connection design for attaching the small piles to the top (transition) and bottom
(secondary) slabs should use high factors of safety and ensure that they stay elastic
when the weak element of the entire system fails. This design concept is similar to the
philosophy used in seismic design in which some of the bridge elements are protected
and remain elastic while plastic hinges form in other parts of the structure. The con-
nection should be designed for cyclic loading, as the system will be subjected to daily
and seasonal temperature fluctuation. Figure E.4 shows one possible connection detail
that was used during the experimental phase. The experimental results indicated that
the area around the connection could have a larger thickness or, alternatively, could
use advanced materials such as ultrahigh-performance concrete. Research is needed to
develop more economical connection details.

Stirrups Concrete Slab


Bars

Studs Baseplate
Figure E.4.  Recommended small
pile–concrete slab connection.

592

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


E.2.6
Geomaterial
Design of the geomaterial consists of the selection of the geomaterial type and the com-
paction requirement. The required compaction depends on the stiffness requirement
around the small piles. It is very important to achieve the required compaction (level
and consistency) around the small piles, so stringent quality control is required during
soil compaction. It is highly recommended to use granular material in this region for
ease of compaction, which will result in smaller long-term settlement and smaller gap
development around the piles caused by pile movements.
Moisture density relation (compaction) tests, maximum and minimum density
(relative density) tests, and in‑place moisture content and density determinations dur-
ing placement of the backfill (using a nuclear moisture density meter) are the recom-
mended soil mechanics tests.
E.3
Cracked Section Analysis
Methods of determining the maximum probable crack width and stiffness reduction
for an axially tensioned concrete member are explained in ACI Report 224.2R-92
(ACI 1997). The maximum probable crack width (Wmax) in a fully cracked member
can be determined from Equation E.1:

Wmax = 0.10 × 10−3 f s 3 dc A (E.1)

where
dc = distance from center of bar to extreme tension fiber (in.);
fs = service stress in the reinforcement (ksi); and
A = effective tension area of concrete surrounding the tension reinforcement,
having the same centroid as the reinforcement, divided by the number of
bars (in.2).

Figure E.5 demonstrates the calculation of A. S is the bar spacing, and H is the
total thickness of the slab.

Figure E.5.  Determination of effective tension area of concrete surrounding the tension reinforcement for
an axially tensioned concrete member.

593

Appendix E. DESIGN STEPS FOR SEAMLESS BRIDGE SYSTEM DEVELOPED BY SHRP 2 PROJECT R19A
As shown in Figure E.5, the parameter 3 dc A can be determined to be dc 3 2S dc
for both one and two layers of reinforcement.
The crack width allowed is inserted in Equation E.1 to obtain the service stress and
strain in the reinforcement.
The LRFD specifications define an exposure factor (γe) that is 1.00 for a Class 1
exposure condition and 0.75 for a Class 2 exposure condition. The crack width associ-
ated with Class 1 and 2 exposure conditions are 0.017 and 0.012, respectively.
The amount of required reinforcing steel can be determined from the axial tensile
force (P) in the member (determined from the structural analysis), as shown by Equa-
tion E.2:

P
P = f s .As ⇒ As = (E.2)
fs
where As is the reinforcement steel area.
The axial force in various segments of the transition slab (P) is determined from
structural analysis. The transition slab may crack when it is in tension, which will result
in reduction of axial stiffness. The reduced axial stiffness of the cracked transition slab
should be used in structural analysis, requiring an iterative cracked section analysis.
In this iterative analysis the section axial stiffness is modified on the basis of the axial
force determined from the previous analysis. Next, the structure is analyzed using the
modified axial stiffness. This process is repeated until convergence. For the first itera-
tion, the slab can be assumed uncracked (the tensile force can be taken the same as the
compressive force developed in the slab due to temperature increase).
The following equations describe the method for determining the reduced axial
stiffness of the concrete member in tension.
ACI Report 224.2R-92 (1997) suggests using Equation E.3 for determining the
direct tensile strength of the concrete ( ft′ ):
1
ft′ = 0.33[γ c ⋅ fc′ ]2 (E.3)

where: From the LRFD specifications, for a given fc′ , the unit weight can be determined
from gc = 0.14 + 0.0001 fc′ , and Ec can be determined from Ec = 33000K1γ c1.5 fc′ .
( )
The stress in the reinforcing bars after the crack occurs f s′,cr is determined from
ACI Report 224.2R-92 (1997), as shown by Equation E.4:
1 
f s′,cr = ft′  − 1 + n  (E.4)
p 

where ρ is the reinforcing ratio (As/Ag), and n is the modular ratio of steel to concrete.
The axial load that causes first cracking in the axially tensioned member is shown
by Equation E.5:

Pcr = f s′,cr ⋅ As (E.5)

594

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


During the cracked section analysis, if the force in a segment of the slab is smaller
than Pcr, the slab will not crack, and no modification will be required in the structural
analysis. Otherwise, if the force in a segment of the transition slab exceeds the above
Pcr, the segment will crack, and the modified axial stiffness of the segment should be
determined and used for the next iteration.
For a cracked section, the average strain in the tensile member can be calculated
from the CEB-FIP Model Code (Thomas Telford Service Ltd. 1993), using Equa-
tion E.6 to determine the modified axial stiffness:
  f s,cr  
2
ε m = ε s 1 − k   (E.6)
  f s  

where es = fs/Es, k = 1.0 for the first loading and 0.5 for repeated or sustained loading.
For the seamless system, k = 0.5 should be used.
Equation E.7 provides the effective modulus of elasticity of steel bars:

Es
Esm = (E.7)
 f  
2
1 − k  c ,cr  
  fs  

The effective axial cross‑sectional stiffness of the tensile concrete member can be
written as (EA)eff = EsmAs. The ratio term (EA)eff/(EA) is the section modification factor
that should be used in the structural analysis to modify the axial stiffness of the mem-
ber in tension. Figure E.6 shows the general flowchart for the cracked section analysis.

 
Figure E.6.  Cracked section analysis flowchart.

595

Appendix E. DESIGN STEPS FOR SEAMLESS BRIDGE SYSTEM DEVELOPED BY SHRP 2 PROJECT R19A
F
CURVED GIRDER BRIDGES

F.1
Background
This section contains a procedure developed to extend the application of jointless
bridges to curved steel I-girder bridges (Doust 2011). Several limitations must be ob-
served when using the suggested approach that reflect the range of parameters consid-
ered in its development. The study considered several bridge configurations for which
detailed finite element analyses were conducted. These analyses were then used to
(1) comprehend the performance of jointless curved girder bridges and (2) develop
­approximate solutions that are in reasonable agreement with the results of detailed
­finite element analysis. Various assumptions and cases were considered during the
devel­opment of the suggested approach:

1. Steel I-girder superstructure made composite with concrete deck;


2. Concrete integral abutments at the bridge ends supported on steel H-piles;
3. One or more intermediate piers isolated from the bridge superstructure by elasto-
meric bearings;
4. Concrete parapets integrally connected to the concrete deck;
5. Superstructure superelevation ranging between 0% and 6%;
6. Abutment wall height ranging between 9 and 13 ft;
7. Wingwalls separated from the abutment wall by means of joints;
8. Approach slab connected to abutment wall using a pinned connection detail;
9. Bridge plan symmetric with respect to the midlength of the bridge;
10. Radial piers and abutments (i.e., the lines of all abutments and piers intersect at
the bridge center of curvature);
596
11. Bridge arc length-over-width ratio larger than 3.0;
12. Ratio of the lengths of end spans to interior spans approximately equal to 0.8; and
13. All intermediate spans of approximately equal length.

The following two sections present step-by-step procedures to calculate the magni-
tude and direction of bridge end displacements and determine the optimum abutment
pile orientation.
F.2
Calculating Magnitude and Direction of End Displacement
For curved integral abutment bridges meeting the limitations described earlier, the
following procedure can be employed to calculate the magnitude and direction of end
displacements:

1. Determine the point of zero movement for the bridge and consequently the bridge
length along the centerline of the bridge (L0) that should be used in calculating the
end displacement. For symmetric bridges supported on a substructure with rela-
tively symmetrical stiffness, it can be assumed that L0 is equal to half the bridge
total arc length. Otherwise, a more detailed approach that takes into account the
relative stiffnesses of the supports should be used to calculate the point of zero
movement.
2. Determine the effective coefficient of thermal expansion by using Equation F.1:

( EAα )deck + ( EAα )girder


α equivalent = (F.1)
( EA)deck + ( EA)girder
3. Calculate the bridge shortening due to contraction by using Equation F.2:

∆ construction = α equivalent ⋅ ∆T ⋅ L0 (F.2)

4. Find the modification factor for bridge shortening due to contraction by using the
information provided in Figure F.1, which provides the relationship between the
radius of curvature and the modification factor used in Equation F.5.
5. Determine the equivalent shrinkage strain by using Equation F.3:

( EA)deck
(
ε sh,equivalent = ε sh,girder + ε sh,deck − ε sh,girder ) ( EA) + ( EA)girder (F.3)
deck

6. Calculate the bridge shortening due to shrinkage by using Equation F.4:

∆ shrinkage = ε sh,equivalent ⋅ L0 (F.4)

7. Find the modification factor for bridge shortening due to shrinkage by using
Figure F.2.

597

Appendix F. CURVED GIRDER BRIDGES


1.15

Modification Factor, gTUc


1.1

Outer
1.05 Arc

1 Inner
Arc

0.95

0.9
100 1000 10000
Radius (ft)
Figure F.1.  Modification factor for bridge contraction.

1.45

1.4
Modification Factor, g Sh

1.35

1.3
Outer
1.25 Arc

Inner
1.2 Arc

1.15

1.1
100 1000 10000
Radius (ft)
Figure F.2.  Modification factor for bridge shrinkage.

598

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


8. Calculate the total factored bridge shortening by using Equation F.5:

( )
∆ total = 1.3 γ TUc ∆ thermal + γ Sh ∆ shrinkage (F.5)

9. Calculate the bridge width effect factor with the following equations. These fac-
tors are calculated for the inner (Equation F.6) and outer (Equation F.7) corners of
the bridge separately. The purpose of these factors is to determine the direction of
end displacement.

W
kin = 1 + 0.84 (F.6)
LC

W
kout = 1 − 0.84 (F.7)
LC
10. Find the direction of the bridge corner displacements by using Equations F.8 and
F.9 for inner and outer displacements, respectively:

  L 
α in = kin 90 − 11   in degrees (F.8)
  R 

  L 
α out = kout 90 − 11   in degrees (F.9)
  R 

11. Knowing the total bridge shortening found in Step 8 and the direction found in
Step 10, solve Equations F.10 through F.16 to find the new location of the bridge
corner. The corner of the bridge is assumed to be originally located at the coor-
dinates xA = RA and yA = 0, in which RA is the radius of the bridge at that specific
corner.

(
xA′ = − ab + a 2b2 − b2 − R′ 2 1 + a2 ( )( ) ) / (1 + a ) (F.10)
2

yA′ = axA′ + b (F.11)

where
a = − tan α (F.12)

b = R tan α (F.13)
y 
γ = tan−1  A′  (F.14)
 xA ′ 

L′ = 2R′ ( β − γ ) (F.15)

in which

599

Appendix F. CURVED GIRDER BRIDGES


L
β= (F.16)
2R
12. Using the new coordinates of the bridge corner xA′ and yA′, the components of
bridge corner displacement are found as shown by Equations F.17 and F.18:

∆ x = xA′ − RA (F.17)

∆ y = yA′ (F.18)

F.3
Optimum Pile Orientation
In curved bridges, the optimum orientation of the piles depends mainly on the bridge
geometry. In contrast to straight bridges, the optimum direction is not the same for all
curved bridges. This section presents a method to find the optimum pile orientation
in a curved bridge that is based on finite element simulation of several curved integral
steel I-girder bridges (Doust 2011). The same concept employed for straight bridges
is also used for curved girder bridges; namely, the piles should be oriented so that
the strong axis of their sections is perpendicular to the direction of bridge maximum
displacement.
The following steps should be used to obtain the optimal abutment pile orientation:

1. The critical load combination for the design of the piles should be determined to
be either expansion based or contraction based. Figure F.3 may help for determining
the controlling load combination.

Bridge Length (ft)


100 400 700 1000
100

Expansion
Control
Radius (ft)

1000
Contraction
Control

10000

Figure F.3.  Controlling type of load combination.

600

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


2. The direction of bridge maximum end displacement, as defined in Figure F.4,
should be determined using the curves presented in Figure F.5.
3. The strong axis of the abutment piles perpendicular to the displacement direction
found in Step 2 should be oriented. If the type of critical load combination can-
not be distinguished for a specific bridge, the bridge should be analyzed for both
expansion-control and contraction-control pile orientations from Step 2, and then
the optimum orientation should be chosen.

Figure F.4.  Direction of bridge


maximum and displacement.

A'

y
u uy
x A ux

150
125
100
Angle of Direction (deg.)

75
50
25 Contraction-Control
0 Expansion-Control
-25 0.00 1.00 2.00 3.00

-50
-75
-100
L/R

Figure F.5.  Angle of direction of bridge end displacement.

601

Appendix F. CURVED GIRDER BRIDGES


G
Design Provision for
Sliding Surfaces Used
in Bearing Devices
for Service Life

G.1
Introduction
This appendix provides a procedure for designing sliding surfaces for service life that
is applicable to various bearing devices that allow rotation and use sliding surfaces to
allow horizontal movements. A key factor in the process is being able to predict the
service life of sliding surfaces, which is based on the following essential parameters:

1. Wear rate of the sliding material, which can be obtained through experimental
work;
2. Total accumulated movements, which can be approximated from loading demand
(traffic and thermal loads) and analysis; and
3. The speed or velocity of movement, which can be determined from analysis de-
pending on movement due to truck load or temperature change.

The target service life of the bridge system is established by the owner. The designer
must ensure that the bearing device incorporating a sliding surface can provide a ser-
vice life exceeding the bridge system. If the service life of the sliding surface is less than
the service life of the bridge system, steps must be taken to accommodate replacement
of the sliding surface or the entire bearing.
The following sections provide detailed descriptions of the parameters listed and
the design steps.

602
G.2
Elements of Design Provisions
G.2.1
Wear Rate
Tests have shown that plain polytetrafluorethylene (PTFE) will wear over time, causing
reduction in thickness, which ultimately affects service life. If the right type of slid-
ing material is selected along with the right thickness, there is greater probability of
achieving the desired service life.
The rate of wear, which can be identified as the anticipated thickness reduction per
length traveled, can be used to approximately predict service life. The rate of wear is
affected by contact pressure, travel speed, temperature, and lubrication. Considering
these factors, Equation G.1 can be used to estimate the wear rate for a sliding surface:

wear rate = base wear rate (material, P, V) × CT × CL (G.1)

where wear rate is defined in terms of mil thickness per mile of travel distance; base
wear rate (material, P, V) is defined as a function of material type, contact pressure,
and velocity, based on experimental tests; and

= modification factor for the effects of low temperature (function of material


CT
type);
= modification factor for the effects of lubrication (function of material type);
CL
P = contact pressure acting normal to the sliding surface; and
V = travel speed of the sliding bearing (see Section G.2.3).

The base wear rate defined in this procedure is the wear rate determined from tests
conducted at various combinations of speed and contact pressure at room temperature,
without lubrication. Stanton et al. (1999) showed that low temperature and lubrica-
tion also contributed to wear rate. Low temperatures increased wear, but lubrication
significantly reduced wear. The effects of these parameters can be seen in Table G.1.
To account for these effects, the factors CT and CL are added to Equation G.1. These
factors are a function of material type and must be determined from tests. At this time,
there is insufficient data to develop these factors accurately for final service life design,
but estimates can be drawn from Table G.1.
Research by Campbell and Kong (1987) on wear of PTFE sliding surfaces indi-
cated that the value of pressure times velocity, referred to as the PV factor, could be
used as a base parameter to predict the corresponding rate of wear. Their research indi-
cated that there was a PV threshold below which there would be a low-wear regime,
and above which there would be a high-wear regime.
Limited proof-of-concept testing in SHRP 2 Project R19A resulted in preliminary
development of PV curves for two types of PTFE sliding materials and an alternate
non-PTFE sliding material. These studies confirmed the concept of a PV factor affect-
ing wear rate for PTFE-based materials, and further confirmed the concept of PV
threshold. However, because of the limited amount of data, additional tests need to be

603

Appendix G. DESIGN PROVISION FOR SLIDING SURFACES USED IN BEARING DEVICES FOR SERVICE LIFE
Table G.1.  PTFE Wear Rates
V T PV Wear Rate
Material Lubrication (in./min) (°F) (lb/in.2 ft/min) (mil/mi)
Unfilled Dimpled, 2.5 68 625 0.3
PTFE lubricated 25 68 6,250 0.5
Flat, 2.5 68 625 0.7
unlubricated 25 68 6,250 189
2.5 –13 625 10
25 –13 6,250 259
Woven Flat, 2.5 68 625 0.3
PTFE unlubricated 25 68 6,250 17
2.5 –13 625 27
25 –13 6,250 24
15% Glass Flat, 2.5 68 625 –1
filled unlubricated 25 68 6,250 –0.5
2.5 –13 625 No result
25 –13 6,250 6
25% Glass Flat, 2.5 68 625 –0.3
filled unlubricated 25 68 6,250 2
2.5 –13 625 4
25 –13 6,250 46

Source: Stanton et al. 1999.


Note: Pressure is 3,000 psi.

carried out to develop final PV versus wear rate curves that can be reliably used for
actual service life design.
Figure G.1 shows wear rate versus PV data for plain PTFE sliding surfaces. It com-
bines data from the SHRP 2 R19A study with data from NCHRP Report 432 (Stanton
et al. 1999) and shows the relative low-wear and high-wear regimes. Data shown in
red are from NCHRP Report 432. As stated, further testing is required to develop
more accurate curves within each of these regions. Figure G.2 shows a similar curve
for glass-reinforced PTFE (Fluorogold) from SHRP 2 R19A tests. Rates of wear for
reinforced PTFE are significantly reduced from plain PTFE and could be considered as
an alternative for plain PTFE in conditions of high PV.
Table G.1 presents wear data from NCHRP Report 432 (Stanton et al. 1999)
and shows wear rates for various PTFE-based materials at constant pressure but with
variations in sliding speed, temperature, and lubrication. These data can be helpful in
providing input for parameters in Equation G.1, but as mentioned, further testing is
required to establish final values.

604

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Figure G.1.  Wear rate versus PV factor
for plain PTFE.

NCHRP 432 Tests


SHRP 2 Tests Dimpled unlubricated
SHRP 2 Tests Plain

Figure G.2.  Wear rate versus PV factor


Wear Rate – Reinforced PTFE
for glass-reinforced PTFE.

♦ NCHRP 432 Tests Glass Filled


SHRP 2 Tests Fluorogold

605

Appendix G. DESIGN PROVISION FOR SLIDING SURFACES USED IN BEARING DEVICES FOR SERVICE LIFE
G.2.2
Estimating Total Accumulated Movements
The total travel demand is the total accumulated distance that the sliding surface will
be traveling throughout the service life of the bridge system. This total travel demand
can be estimated using (1) specified bridge system service life, (2) traffic and thermal
loading demands, and (3) calculated horizontal movements related to applied traffic
and thermal loadings.
Sliding surfaces are a means to accommodate horizontal movements associated
with traffic load and daily and seasonal bridge superstructure expansion and contrac-
tion. The total bridge movement at the bearing [(TD)Demand] in miles is produced by the
following three mechanisms:

1. Traffic-induced horizontal movement, (TD)Tr. The total accumulated travel with


this type of movement can be considerably greater than that associated with
Mechanisms 2 and 3.
2. Daily temperature–induced horizontal movement, (TD)DT.
3. Seasonal temperature-induced horizontal movement, (TD)ST.

The total movement due to temperature is the combination of daily and seasonal
movements.
G.2.2.1
Traffic-Induced Horizontal Movement (TD)Tr
Equation G.2 estimates total horizontal movement of the sliding surface, in miles due
to traffic movement, for the designed service life [(SL)B] in years:

365
(TD)Tr = 2 × A × θ × D1 × n × 1.33 × (ADTT)SL × ( SL)B × (G.2)
63,360
where
(TD)Tr = traffic-induced horizontal movement (mi);
A = 1, if each end of the girder is free to move in horizontal direction, or
2, if all horizontal movements are accommodated at one end, with
the other end pinned against horizontal movement;
(ADTT)SL = single-lane ADTT (average daily truck traffic);
θ = rotation of girder end with sliding bearing (rad);
D1 = depth of neutral axis measured from the bottom flange (in.);
(SL)B = design service life (years);
1.33 = impact factor for truck load; and
n = number of equivalent full-amplitude horizontal movement cycles per
truck passage (due to free vibration) initially taken equal to 1.0 for
this procedure.

606

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


In Equation G.2, A is a parameter that accounts for boundary conditions at both
ends of the span. The term θ × D1 is the horizontal movement due to girder end rota-
tion. The factor 2 accounts for the full cycle of movement, which includes deflection
and rebound.
When a truck passes over a span, the girders deflect to a maximum amount as
the truck approaches midspan and then recover as the truck moves toward the end of
the span. However, because of dynamic behavior, the girders may continue to vibrate
until the girder deflection is damped out. The cycles produced after truck passage have
successively smaller amplitudes, and the decay is dependent on the damping ratios.
This characteristic is represented by the term n, which is the equivalent number of
cycles with full amplitude that corresponds to the total number of cycles with decreas-
ingly smaller amplitude. For the purposes of this procedure, however, this term can
be neglected (using n = 1). Although it is recognized that this behavior occurs, its
true magnitude as it applies to bearing movement requires further study and field
verification.
In Equation G.2, ADTT is the average daily truck traffic, and (SL)B is the owner-
specified service life of the bridge system. The constant terms in Equation G.2 are
conversion factors.
G.2.2.2
Daily Temperature–Induced Horizontal Movement, (TD)DT
Equation G.3 estimates total horizontal movement (in miles) of the sliding surface due
to daily temperature fluctuation over the designed service life of the bridge (in years):

(TD)DT = DLDaily × (SL)B × 365/5,280 (G.3)

where
DLDaily = 2aLDTDaily;
a = coefficient of thermal expansion;
L = maximum span length (ft) or length contributing to expansion in the
case of multiple spans; and
DTDaily = maximum daily temperature fluctuation.

G.2.2.3
Seasonal Temperature–Induced Horizontal Movement, (TD)ST
Equation G.4 estimates the total horizontal movement (in miles) of the sliding surface
due to yearly temperature fluctuation over the designed service life (in years):

1
(TD)ST = ∆LAnnual × ( SL )B × (G.4)
5,280
where DLAnnual is 2aLDTAnnual, and DTAnnual is the maximum annual temperature fluctuation.

607

Appendix G. DESIGN PROVISION FOR SLIDING SURFACES USED IN BEARING DEVICES FOR SERVICE LIFE
G.2.2.4
Total Induced Horizontal Movement, (TD)Demand
Finally, Equation G.5 estimates the total bridge movement at the bearing:

(TD)Demand = (TD)Tr + (TD)DT + (TD)ST (G.5)

G.2.3
Estimating Speed of Sliding Surface Movement
The service life calculation described previously involves the use of PV curves that are
specific for the material used for the sliding surface. The term V is the speed at which
the sliding surface moves, which depends on whether the movement is caused by truck
load or temperature load. This section provides a procedure for calculating V in the
PV expression.
G.2.3.1
Speed of Movement per Truck Passage
The speed of travel (V) for the sliding bearing for movement caused by truck passage
can be determined from the general equation shown in Equation G.6:

total horizontal movement per truck passage


average travel speed = (G.6)
travel time
The total horizontal movement of the sliding surface per truck passage can be
determined from Equation G.7:

total horizontal movement per truck passage = 2 × 1.33 × A × θ × D1 × n  (G.7)

where
A = 1, if each end of the girder is free to move in the horizontal direction, or
2, if all horizontal movements are accommodated at one end with the
other end pinned against horizontal movement;
θ = rotation of the girder end with sliding surface (rad);
D1 = depth of neutral axis measured from the bottom flange (in.);
1.33 = impact factor for truck load; and
n = number of equivalent full-amplitude horizontal movement cycles per
truck passage (due to free vibration) initially taken equal to 1.0 for this
procedure.

The total travel time is the time that it will take for the accumulated horizontal
movement caused by the passage of one truck to occur. It is the time for the first cycle
and for all succeeding dynamic vibration cycles to take place. If the component of the
time due to dynamic vibration cycles as described in Section G.2.2 is neglected, the
resulting time for the movement can be determined from Equation G.8:

bridge span length


t= (G.8)
truck speed

608

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


G.2.3.2
Speed of Movement per Temperature Variation
The speed of travel (V) for the sliding bearing for movement caused by daily and
seasonal temperature change is a much slower velocity that can be estimated from
Equation G.9:

total horizontal movement per temperature change


average travel speed = (G.9)
travel time
The total horizontal movement of the sliding surface due to temperature move-
ment is estimated by determining the total yearly temperature movement caused
by daily temperature change and seasonal temperature change, as shown in Equa-
tions G.10 to G.12:

total movement due to daily temperature change = ΔLDaily × 365 (G.10)

total movement due to seasonal temperature change = ΔLAnnual (G.11)

total temperature movement = (ΔLDaily × 365) + ΔLAnnual (G.12)

The total travel time for the total temperature movement as defined in Equation
G.12 is 365 days, which can be converted into consistent units.
G.3
Design Process for Sliding Surfaces
G.3.1
Steps in Design Process
The following steps could be used to select the type of sliding material and its required
thickness to meet service life requirements:

Step 1. Calculate the total travel distance demand [(TD)Demand, in miles] using the
procedures in Section G.2.2.
Step 2. Determine the velocity of movement on the basis of traffic load or tempera-
ture movement using the procedures in Section G.2.3.
Step 3. Select a trial sliding surface type and determine the corresponding wear
rate, based on PV curves for the type of material (in inches per mile), using the proce-
dures in Section G.2.1.
Step 4. Calculate the thickness demand, which is the total predicted wear or reduc-
tion in thickness for the sliding surface, by using Equations G.13 and G.14:

(thickness)Demand = (TD)Demand × wear rate × α (G.13)

gross thickness = (thickness)Demand + thickness of recess (G.14)

609

Appendix G. DESIGN PROVISION FOR SLIDING SURFACES USED IN BEARING DEVICES FOR SERVICE LIFE
where

(TD)Demand = total induced horizontal movement (see Equation G.4), and


α = factor to assure that the thickness will not be zero at the end of the
service life to prevent undesirable metal-to-metal contact (>1.0).

The (thickness)Demand is the thickness that is subject to wear. Accordingly, the gross
thickness is the thickness subject to wear plus the recessed thickness that is used to
positively connect the sliding surface to the backing plate.
Step 5. Establish the gross thickness of the material to be specified in the design
plan. The thickness of commercially available sliding surfaces must be larger than the
gross thickness calculated in Step 4.
Step 6. If the commercially available thicknesses are less than the required gross
thickness, then there are two available approaches: (1) select another material, such
as a reinforced or braided PTFE or other sliding material type that could meet the
demand by repeating Steps 3 through 5; or (2) calculate the service life of the commer-
cially available thicknesses and develop a replacement strategy accordingly.
G.3.2
Design Process Application
The design process has application to all bearing types that use sliding surfaces to per-
mit horizontal movement and where horizontal movement is caused by truck load or
temperature load. It can also be applied to evaluate the service life of sliding surfaces
that are used in combination with elastomeric pads. In these cases, the elastomeric pad
is designed to accommodate the high-cycle, low-amplitude horizontal movement due
to truck load, and the sliding surface is designed to accommodate the larger-amplitude,
low-cycle movement due to temperature. This approach has advantages for the design
of expansion bearings at the end of a series of continuous spans where the temperature
movement is large and the superstructure reactions are low. Combining e­ lastomeric
pads with sliding surfaces reduces the required thickness of the elastomeric pads and
permits the use of more durable elastomeric bearings in cases in which high-load
multi­rotation (HLMR) types would have been required because of the excessive height
required for the elastomeric pads. Further advantages of the reduced elastomeric pad
thickness include better stability during construction and operation and reduced in-
stantaneous and long-term compressive deflection.

610

DESIGN GUIDE FOR BRIDGES FOR SERVICE LIFE


Related SHRP 2 Research

Geotechnical Solutions for Soil Improvement, Rapid Embankment Construction, and


Stabilization of the Pavement Working Platform (R02)

Innovative Bridge Designs for Rapid Renewal (R04)

Nondestructive Testing to Identify Concrete Bridge Deck Deterioration (R06A)

Mapping Voids, Debonding, Delaminations, Moisture, and Other Defects Behind or


Within Tunnel Linings (R06G)

Performance Specifications for Rapid Highway Renewal (R07)

Process of Managing Risk on Rapid Renewal Projects (R09)

Project Management Strategies for Complex Projects (R10)

Bridges for Service Life Beyond 100 Years: Service Limit State Design (R19B)
TRB Oversight Committee for the
Strategic Highway Research Program 2*

Chair: Kirk T. Steudle, Director, Michigan Department of Transportation

MEMBERS
H. Norman Abramson, Executive Vice President (retired), Southwest Research Institute
Alan C. Clark, MPO Director, Houston–Galveston Area Council
Frank L. Danchetz, Vice President, ARCADIS-US, Inc.
Malcolm Dougherty, Director, California Department of Transportation
Stanley Gee, Executive Deputy Commissioner, New York State Department of Transportation
Mary L. Klein, President and CEO, NatureServe
Michael P. Lewis, Director, Rhode Island Department of Transportation
John R. Njord, Executive Director (retired), Utah Department of Transportation
Charles F. Potts, Chief Executive Officer, Heritage Construction and Materials
Ananth K. Prasad, Secretary, Florida Department of Transportation
Gerald M. Ross, Chief Engineer (retired), Georgia Department of Transportation
George E. Schoener, Executive Director, I-95 Corridor Coalition
Kumares C. Sinha, Olson Distinguished Professor of Civil Engineering, Purdue University
Paul Trombino III, Director, Iowa Department of Transportation

EX OFFICIO MEMBERS
Victor M. Mendez, Administrator, Federal Highway Administration
David L. Strickland, Administrator, National Highway Transportation Safety Administration
Frederick “Bud” Wright, Executive Director, American Association of State Highway and Transportation Officials

LIAISONS
Ken Jacoby, Communications and Outreach Team Director, Office of Corporate Research, Technology, and Innovation Management,
Federal Highway Administration
Tony Kane, Director, Engineering and Technical Services, American Association of State Highway and Transportation Officials
Jeffrey F. Paniati, Executive Director, Federal Highway Administration
John Pearson, Program Director, Council of Deputy Ministers Responsible for Transportation and Highway Safety, Canada
Michael F. Trentacoste, Associate Administrator, Research, Development, and Technology, Federal Highway Administration

* Membership as of March 2014.

RENEWAL TECHNICAL COORDINATING COMMITTEE*

Chair: Daniel D’Angelo, Recovery Acting Manager, Director and Deputy Chief Engineer, Office of Design, New York State Department
of Transportation

MEMBERS
Rachel Arulraj, President, InfoInnovation
Michael E. Ayers, Consultant, Technology Services, American Concrete Pavement Association
Thomas E. Baker, State Materials Engineer, Washington State Department of Transportation
John E. Breen, Al-Rashid Chair in Civil Engineering Emeritus, University of Texas at Austin
Steven D. DeWitt, Chief Engineer (retired), North Carolina Turnpike Authority
Tom W. Donovan, Senior Right of Way Agent (retired), California Department of Transportation
Alan D. Fisher, Manager, Construction Structures Group, Cianbro Corporation
Michael Hemmingsen, Davison Transportation Service Center Manager (retired), Michigan Department of Transportation
Bruce Johnson, State Bridge Engineer, Oregon Department of Transportation, Bridge Engineering Section
Leonnie Kavanagh, PhD Candidate, Seasonal Lecturer, Civil Engineering Department, University of Manitoba
Cathy Nelson, Technical Services Manager/Chief Engineer (retired), Oregon Department of Transportation
John J. Robinson, Jr., Assistant Chief Counsel, Pennsylvania Department of Transportation, Governor’s Office of General Counsel
Ted M. Scott II, Director, Engineering, American Trucking Associations, Inc.
Gary D. Taylor, Professional Engineer
Gary C. Whited, Program Manager, Construction and Materials Support Center, University of Wisconsin–Madison

AASHTO LIAISON
James T. McDonnell, Program Director for Engineering, American Association of State Highway and Transportation Officials

FHWA LIAISONS
Steve Gaj, Leader, System Management and Monitoring Team, Office of Asset Management, Federal Highway Administration
Cheryl Allen Richter, Assistant Director, Pavement Research and Development, Office of Infrastructure Research and Development,
Federal Highway Administration
J. B. “Butch” Wlaschin, Director, Office of Asset Management, Federal Highway Administration

CANADA LIAISON
Lance Vigfusson, Assistant Deputy Minister of Engineering & Operations, Manitoba Infrastructure and Transportation

*Membership as of March 2014.

You might also like