0% found this document useful (0 votes)
187 views232 pages

Oceanography and Oceans GEOLOGY BASICS

The document titled 'Oceans and Oceanography' provides a comprehensive overview of Earth's oceans, covering their physical properties, circulation patterns, and the impact they have on climate and human activities. It discusses the composition of seawater, the features of ocean basins, and the historical development of oceanography as a science. Additionally, it highlights the significance of ocean currents, such as the Gulf Stream and El Niño, in influencing global weather patterns and marine ecosystems.
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)
187 views232 pages

Oceanography and Oceans GEOLOGY BASICS

The document titled 'Oceans and Oceanography' provides a comprehensive overview of Earth's oceans, covering their physical properties, circulation patterns, and the impact they have on climate and human activities. It discusses the composition of seawater, the features of ocean basins, and the historical development of oceanography as a science. Additionally, it highlights the significance of ocean currents, such as the Gulf Stream and El Niño, in influencing global weather patterns and marine ecosystems.
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

OCEANS AND OCEANOGRAPHY

THE LIVING EARTH

OCEANS AND OCEANOGRAPHY


EDITED BY JOHN P. RAFFERTY, ASSOCIATE EDITOR, EARTH AND LIFE SCIENCES
Published in 2011 by Britannica Educational Publishing
(a trademark of Encyclopædia Britannica, Inc.)
in association with Rosen Educational Services, LLC
29 East 21st Street, New York, NY 10010.

Copyright © 2011 Encyclopædia Britannica, Inc. Britannica, Encyclopædia


Britannica, and the Thistle logo are registered trademarks of Encyclopædia
Britannica, Inc. All rights reserved.

Rosen Educational Services materials copyright © 2011 Rosen Educational


Services, LLC.
All rights reserved.

Distributed exclusively by Rosen Educational Services.


For a listing of additional Britannica Educational Publishing titles, call toll
free (800) 237-9932.

First Edition

Britannica Educational Publishing


Michael I. Levy: Executive Editor
J.E. Luebering: Senior Manager
Marilyn L. Barton: Senior Coordinator, Production Control
Steven Bosco: Director, Editorial Technologies
Lisa S. Braucher: Senior Producer and Data Editor
Yvette Charboneau: Senior Copy Editor
Kathy Nakamura: Manager, Media Acquisition
John P. Rafferty: Associate Editor, Earth and Life Sciences

Rosen Educational Services


Jeanne Nagle: Senior Editor
Nelson Sá: Art Director
Cindy Reiman: Photography Manager
Matthew Cauli: Designer, Cover Design
Introduction by Therese Shea Harasymiw

Library of Congress Cataloging-in-Publication Data

Oceans and oceanography / edited by John P. Rafferty.—1st ed.


p. cm.—(The living earth)
“In association with Britannica Educational Publishing, Rosen Educational
Services.”
Includes bibliographical references and index.
ISBN 978-1-61530406-6 (eBook)
1. Oceanography. I. Rafferty, John P.
GC11.2.O258 2011
551.46—dc22

2010032151
On the cover: A clown fish swims amid the anemones along the ocean floor.
Shutterstock.com

On page x: A diver examines the many diverse marine life forms located
within the Verde sea passage, south of Manila, Philippines. Jay
Directo/AFP/Getty Images

On pages v, 1, 49, 104, 139, 180, 226, 249, 250, 252, 260: A cresting wave.
Jean-Paul Nacivet/Photographer’s Choice/Getty Images
CONTENTS

Introduction

Chapter 1: The Ocean


Relative Distribution of the Oceans
Hypsometry
Major Subdivisions of the Oceans
Origin of Ocean Waters
Composition of Seawater
Dissolved Inorganic Substances

Dissolved Organic Substances


Effects of Human Activities
The Physical Properties of Seawater
Salinity Distribution
Horse Latitudes
Temperature Distribution
Thermal Properties
Density of Seawater and Pressure
Optical Properties
Acoustic Properties
Chemical Evolution of the Oceans
The Early Oceans
The Transition Stage
The Modern Oceans

Chapter 2: Ocean Basins

Sea Level
Exploration of the Ocean Basins
Seasat
Deep-Sea Sediments
Sediment Types
Sedimentation Patterns
Evolution of the Ocean Basins Through Plate Movements
Island Arcs
Ocean Basin Features
Abyssal Hills
Continental Margins
Continental Rises
Continental Shelves
Continental Slopes
Deep-Sea Trenches
Oceanic Crust
Oceanic Plateaus

Oceanic Ridges

Chapter 3: Circulation and Waves


The Circulation of the Ocean Waters
Distribution of Ocean Currents
Causes of Ocean Currents
Pressure Gradients
Coriolis Effect
Frictional Forces
Geostrophic Currents
Density Currents in the Oceans
Ekman Layer
Wind-Driven Circulation
Equatorial Currents
The Subtropical Gyres
The Subpolar Gyres
Antarctic Circumpolar Current
Thermohaline Circulation
Waves of the Sea
Surface Gravity Waves
Tsunamis

Seiche Waves and Internal Waves


Ocean Tides
Tide-Generating Forces
Tidal Bores

Chapter 4: The Climatic and Economic Impacts of the Oceans


Seasonal and Interannual Ocean-Atmosphere Interactions
The Link Between Ocean Surface Temperature and Climate Anomalies
Formation of Tropical Cyclones

Effects of Tropical Cyclones on Ocean Waters


Influence on Atmospheric Circulation and Rainfall
The Gulf Stream and Kuroshio Systems
The Gulf Stream
The Kuroshio Current
The Poleward Transfer of Heat
El Niño/Southern Oscillation and Climatic Change
The El Niño Phenomenon

The Southern Oscillation


Economic Aspects of the Oceans
Medium for Transportation and Communications
Source of Food and Water
Energy Resources
Source of Minerals and Other Raw Materials
Deepwater Disaster
Waste Disposal and Other Related Actions

Chapter 5: Oceanography

Ancient Oceanography
The Discovery of the Hydrologic Cycle
The Initial Studies of the Tides
Early Oceanography
The Rise of Subterranean Water
Evaporation from the Sea
Oceanography in the 19th Century
Oceanography in the 20th Century

Ocean Circulation, Currents, and Waves


Ocean Bathymetry
Bathymetry
Water Resources and Seawater Chemistry
Desalinization, Tidal Power, and Minerals from the Sea
Marine Geology
Undersea Exploration
Basic Elements of Undersea Exploration
Dead Reckoning
Methodology and Instrumentation
Nansen Bottle
Sonar
Conclusion

Chapter 6: Marine Explorers and Oceanographers


Alexander Agassiz
Prince Albert I of Monaco

Robert Ballard
Jacques-Yves Cousteau
Robert S. Dietz
Vagn Walfrid Ekman
Richard H. Fleming
Bjørn Helland-Hansen
Columbus O’Donnell Iselin
Matthew Fontaine Maury
Walter Munk
Sir John Murray
Fridtjof Nansen

Jacques Piccard
Francis P. Shepard
Henry Melson Stommel
Georg Wüst

Appendix
Glossary
Bibliography
Index
INTRODUCTION

Of all the planets in the solar system, Earth is singled out as the “Blue Planet.”
This colorful label is indicative of the presence of water on the planet’s surface.
When viewed from space, Earth’s oceans, which cover nearly three-fourths of
the total sphere, appear blue. The sheer preponderance of water makes it easy to
comprehend how the oceans influence nearly every life process on Earth. This
volume takes a comprehensive look at the world’s oceans, examining them from
the tip of their wave crests to the depths of their basins. Topics covered include
currents, sea life, the ocean’s effect on climate, and the science of oceanography.
Although they are, in truth, one continuous body of water, the world’s oceans
are commonly referred to in the plural. Earth has three major oceans: the Pacific,
the Atlantic, and the Indian Ocean. The Pacific is the largest in area and volume,
and the Indian is the smallest. The Arctic Ocean is technically considered part of
the northern Atlantic. While the waters surrounding Antarctica are sometimes
called the Southern Ocean, they are considered by many to be part of the
southern Pacific, Atlantic, and Indian oceans.
The origin of oceanic waters remains something of an unknown, though a
few theories offer an explanation. One states that as the planet was forming, its
developing crust reacted with water vapor and other gases in the atmosphere to
produce liquid water. Of the two kinds of crust that solidified over Earth’s
mantle, the thicker oceanic crust formed a basin for the oceans and served as a
cradle for Earth’s early living organisms. The earliest fossils of algae and
bacteria date from 3.3 billion years ago.
Topographical studies of the ocean floor have revealed that Earth’s ocean
basins possess several terrain features that would be familiar to land dwellers.
Similar to its terrestrial counterpart, the ocean floor is covered with hills,
mountains, featureless plains, and deep gorges. The majority of the ocean floor,
however, lies at depths of between 4,000 metres (13,000 feet) and 5,000 metres
(16,500 feet). The Mariana Trench, located in the western part of the North
Pacific Ocean, plunges to over 11,000 metres (36,200 feet). Little was known
about ocean basins until the development of sonar in the early 1900s. Sound
waves emitted by sonar allow marine researchers to detect objects thousands of
feet under water. Other tools are also used. Satellites, global positioning systems,
radar, and echo-sounder systems are among the most important remote-sensing
tools. These tools have been used to map areas several parts of the ocean, the
Mid-Atlantic Ridge being one of the most prominent mapped features.
Core samples of sediment from deep areas of the ocean floor have been a
boon to the theory that revolves around plate tectonics. The top layer of Earth’s
crust is made up of tectonic plates in constant, albeit slow, movement. Beginning
200 million years ago, the supercontinent Pangea—a land-mass made up of all
of Earth’s continents—began to break apart, eventually fragmenting into the
present-day continents separated by the oceans. Even today, spreading plates at
the bottom of Earth’s oceans spew molten rock from the mantle, creating new
floor. The theory of plate tectonics explains the relative “newness” of the ocean
floor compared to the age of the planet; the floor is constantly recycling itself.
The composition of oceanic water was established by about 1.5 billion years
ago. The primary mixture is 96.5 percent water, 2.5 percent salts, and a small
percentage of substances that include dissolved inorganic and organic materials,
other particles, and atmospheric gases. The amount of salt lost through oceanic
processes is equal to returns from continental drainage.
The composition of ocean water is ideal for the countless organisms that,
over hundreds of millions of years, have adapted to the mix of water and salt.
Inorganic chemicals, such as phosphorus, nitrogen, and zinc, provide further
nourishment, as does sediment from continental shelves. In return, marine life
breaks down much of the material that finds its way into the ocean.
Photosynthesis, the essential process that brings energy to marine food chains,
occurs just below the water’s surface, where the ingredients of solar energy,
carbon dioxide, and nutrient salts are all available. The byproduct of
photosynthesis is oxygen, and Earth’s waters are large producers of oxygen.
Complex circulation patterns influence the chemical composition of oceans.
Not surprisingly, wind is a major factor in circulating ocean waters. As wind
blows across the water, the ocean responds with surface waves. Vertical
movements of water in the oceans, referred to as upwelling, cycle waters up into
the surface layer. Downwelling occurs when water is carried vertically down
from the surface. These motions exchange the cold, deeper layers of the ocean,
which are rich in nutrients and carbon dioxide, with the warmer surface waters,
rich in oxygen.
Several other forces govern the direction and formation of oceanic currents.
The most significant are horizontal gradients of pressure, Coriolis forces,
friction, and gravity. Horizontal pressure causes water molecules to move
horizontally from regions of high pressure to regions of low pressure. Surface
waters warm as they absorb solar radiation and become less dense. The Coriolis
forces refer to the effects of Earth’s rotation on the movement of oceanic waters.
Spinning out of vast circular systems called gyres, currents move clockwise in
the Northern Hemisphere and counterclockwise in the Southern.
Once water is set in motion, it won’t stop until it meets a stronger, opposing
force. It can be slowed, however, by friction from contact with slower currents
and waves, as well as with essentially nonmoving volumes of water near the
ocean floor. Currents also decelerate as gravity pulls them down. The region of
the ocean affected by wind is the Ekman layer, which extends about 100 metres
(330 feet) below the surface of the ocean. Below this layer, the circulation of
currents is much slower.
Two properties of ocean water contribute enable circulation at great depths:
temperature and salinity. When water cools, it becomes denser and sinks,
displacing the water below. The more saline the water, the denser it is. Together,
these properties create a process called thermohaline circulation, first discovered
in 1960. Thermohaline circulation is constant in polar regions, where water at
the surface cools, sinks, and is replaced with more water. The current is created
and travels at depth until it finally upwells again in a return current to polar
regions. Thermohaline circulation is responsible for slowly moving a huge
volume of water all over the world. Carrying nutrients obtained from the ocean
floor, the upwelling process brings these nutrients to the surface, replenishing the
needs of ocean plant and animal life.
Oceanic water also interacts with the atmosphere, exchanging enormous
quantities of oxygen and carbon dioxide. The ocean absorbs much of the sun’s
radiation as heat is delivered downward into the ocean’s depths. Warm-water
currents carry heat. When this heat is transferred to the atmosphere, air rises,
creating regions of low pressure. In contrast, cold-water currents cause high
pressure systems. Both warm and cold currents affect the climate, but the reverse
is also true. Some scientists believe that global warming could shut down or
weaken parts of the thermohaline ocean current system. Some scientists
speculate that an influx of freshwater from ice sheets and glaciers running into
the North Atlantic could “freshen,” and thus disrupt the sections of the
thermohaline circulation. Since freshwater is less dense than salt water,
significant amounts of freshwater entering the North Atlantic may lower the
density of surface waters and stop the sinking motion that drives thermohaline
circulation.
Several well-documented current systems greatly impact climates. The Gulf
Stream is a warm current flowing northeastward in the North Atlantic off the
North American coast. It is part of a clockwise-rotating gyre that begins with the
westward-moving North Equatorial Current. Some consider the Florida current
sweeping warm water up the Florida and Carolina coasts part of this system as
well. The Gulf Stream tends to change over time, at times even seeming to
disappear and then reappear. Its winds carry warm, moist air to northwestern
Europe. In winter, the air over the North Atlantic west of Norway is more than
22 °C (40 °F) warmer than the average for that latitude. The Gulf Stream is one
of several western boundary currents. The Kuroshio in the northwestern Pacific
is an example of another.
Although these currents bring moderate and warm weather to the coasts,
occasionally irregular events cause shifts in the currents and dramatic changes in
weather. El Niño is the name for unusually warm weather conditions that occur
periodically along the Pacific coast of South America, near the Equator. El Niño
occurs when trade winds (persistent winds blowing from the west) weaken,
which reduces the upwelling of cool, nutrient-rich water, replacing warm surface
water. The warming of the water kills plankton, a major food source for fish, and
results in a temporary but major disruption of the region’s marine ecosystem. El
Niño also causes an atypical increase in precipitation in many areas, some
thousands of miles away from the Equator. It is unclear why the trade winds
weaken, so El Niño is hard to predict. However, the ecological and climatic
effects may be felt for about a year.
These and other facts about the world’s seas are known are mainly due to
advances in oceanography, which is the study of all things related to the ocean.
The science of oceanography is divided into four main components: marine
geology, chemical oceanography, physical oceanography, and marine ecology.
Marine geology focuses on the features and evolution of the oceans. Physical
oceanography, chemical oceanography, and marine ecology are all closely
related fields with overlapping concerns. In addition to tangible properties such
as temperature, physical oceanography studies include the movement of ocean
waters and their interactions with Earth’s atmosphere. In contrast, chemical
oceanography focuses on the composition of seawater and its chemical
relationships with marine organisms, while marine ecology, which is also called
biological oceanography, is the study of those organisms.
Modern oceanic exploration is conducted from platforms above water: ships,
buoys, aircraft, or satellites. Modern investigations began during the Age of Sail,
but newer vessels designed to withstand the actions of weather and waves are
more accurate in their underwater measurements. Perhaps tomorrow’s vessels
will be stationed on the ocean floor—a location that is increasingly being viewed
as an area to harvest. Ocean basins are rich in minerals such as iron and
manganese. Amazingly, the ocean contains very nearly every element found on
Earth. Deep-submergence research vehicles have located mineral deposits, as
well as never-before-seen biological communities, at seafloor spreading centres.
The ocean is a trove of significant resources. It is a source of food for
humans and many organisms that live in marine ecosystems. After
desalinization, ocean water is made useful for agricultural purposes, irrigating
fields in places with little access to freshwater. Additionally, the oceans figure
prominently in trade and transportation. Companies pump oil from hundreds of
offshore wells in their quest to supply a world full of consumers hungry for
petroleum. Offshore drilling is not without its danger to marine ecosystems, as
evidenced by 2010’s Deepwater Horizon disaster in the Gulf of Mexico.
Thankfully, there also is clean energy to be had from the sea, as tides and waves
are converted to electricity.
However, human activities are exerting an ever-increasing influence over the
world’s oceans. Atmospheric carbon dioxide, which is often transferred to the
oceans, continues to increase. Phytoplankton and other organisms in the oceans,
along with chemical processes in the water, absorb tremendous amounts of
carbon dioxide. As carbon emissions from human activities increase, there is a
question of whether the ocean will be able to absorb more in the future.
Additionally, man-made pollutants alter oceanic processes by blocking light
from reaching photosynthetic organisms and those that rely on vision, change the
chemical composition of seawater in coastal and runoff areas, and introduce
potentially harmful compounds to marine life.
Despite centuries of research, the ocean remains one of the few places on
Earth that has not been completely explored and described. Oceanographers
continue to provide insight into formerly unknown aspects of oceanography.
Perhaps no oceanographer was so famous as Jacques-Yves Cousteau, whose
name has become synonymous with deep-sea exploration. In addition, Robert
Ballard’s discovery of the Titanic in 1985 sparked a curiosity in deep-sea
archaeology. Additional knowledge of the oceans would enable scientists to
accurately predict phenomena such as the timing and intensity of El Niño and
the behaviour of earthquake-predicated tsunamis. Having a greater
understanding of marine ecology would enable humans to harvest the resources
they need from the oceans in sustainable ways.
The oceans have played a crucial roles in the origin and development of life
on Earth, and they continue to be essential for maintaining life. Whether they
continue to provide these critical services, remaining a crucial resource for the
future of this big “Blue Planet,” is largely up to the actions of human beings in
the coming decades.
CHAPTER 1
THE OCEAN

The ocean is a continuous body of salt water that is contained in enormous


basins on Earth’s surface. When viewed from space, the predominance of the
oceans on Earth is readily apparent. The oceans and their marginal seas cover
nearly 71 percent of Earth’s surface, with an average depth of 3,795 metres
(12,450 feet). The exposed land occupies the remaining 29 percent of the
planetary surface and has a mean elevation of only 840 metres (2,756 feet).
Actually, all the elevated land could be hidden under the oceans and Earth
reduced to a smooth sphere that would be completely covered by a continuous
layer of seawater 2,686 metres (8,812 feet) deep. This is known as the sphere
depth of the oceans and serves to underscore the abundance of water on Earth’s
surface.
Earth is unique in the solar system because of its distance from the Sun and
its period of rotation. These combine to subject Earth to a solar radiation level
that maintains the planet at a mean surface temperature of 17 °C (62.6 °F), which
varies little over annual and night-day cycles. This mean temperature allows
water to exist on Earth in all three of its phases—solid, liquid, and gaseous. No
other planet in the solar system has this feature. The liquid phase predominates
on Earth. By volume, 97.957 percent of the water on the planet exists as oceanic
water and associated sea ice. The gaseous phase and droplet water in the
atmosphere constitute 0.001 percent. Fresh water in lakes and streams makes up
0.036 percent, while groundwater is 10 times more abundant at 0.365 percent.
Glaciers and ice caps constitute 1.641 percent of Earth’s total water volume.
Each of the above is considered to be a reservoir of water. Water
continuously circulates between these reservoirs in what is called the hydrologic
cycle, which is driven by energy from the Sun. Evaporation, precipitation,
movement of the atmosphere, and the downhill flow of river water, glaciers, and
groundwater keep water in motion between the reservoirs and maintain the
hydrologic cycle.
The large range of volumes in these reservoirs and the rates at which water
cycles between them combine to create important conditions on Earth. If small
changes occur in the rate at which water is cycled into or out of a reservoir, the
volume of a reservoir changes. These volume changes may be relatively large
and rapid in a small reservoir or small and slow in a large reservoir. A small
percentage change in the volume of the oceans may produce a large proportional
change in the land-ice reservoir, thereby promoting glacial and interglacial
stages. The rate at which water enters or leaves a reservoir divided into the
reservoir volume determines the residence time of water in the reservoir. The
residence time of water in a reservoir, in turn, governs many of the properties of
that reservoir.

RELATIVE DISTRIBUTION OF THE OCEANS


Those conducting oceanic research generally recognize the existence of three
major oceans, the Pacific, Atlantic, and Indian. (The Arctic Ocean is considered
an extension of the Atlantic.) Arbitrary boundaries separate these three bodies of
water in the Southern Hemisphere. One boundary extends southward to
Antarctica from the Cape of Good Hope, while another stretches southward from
Cape Horn. The last one passes through Malaysia and Indonesia to Australia,
and then on to Antarctica. Many subdivisions can be made to distinguish the
limits of seas and gulfs that have historical, political, and sometimes ecological
significance. However, water properties, ocean currents, and biological
populations do not necessarily recognize these boundaries. Indeed, many
researchers do not either. The oceanic area surrounding the Antarctic is
considered by some to be the Southern Ocean.

Boundaries of the world’s oceans and seas. Encyclopædia Britannica,


Inc.

If area-volume analyses of the oceans are to be made, then boundaries must


be established to separate individual regions. In 1921 Erwin Kossina, a German
geographer, published tables giving the distribution of oceanic water with depth
for the oceans and adjacent seas. This work was updated in 1966 by H.W.
Menard and S.M. Smith. The latter only slightly changed the numbers derived
by Kossina. This was remarkable, since the original effort relied entirely on the
sparse depth measurements accumulated by individual wire soundings, while the
more recent work had the benefit of acoustic depth soundings collected since the
1920s. This type of analysis, called hypsometry, allows quantification of the
surface area distribution of the oceans and their marginal seas with depth.
The distribution of oceanic surface area with 5° increments of latitude shows
that the distribution of land and water on Earth’s surface is markedly different in
the Northern and Southern hemispheres. The Southern Hemisphere may be
called the water hemisphere, while the Northern Hemisphere is the land
hemisphere. This is especially true in the temperate latitudes.
This asymmetry of land and water distribution between the Northern and
Southern hemispheres makes the two hemispheres behave very differently in
response to the annual variation in solar radiation received by Earth. The
Southern Hemisphere shows only a small change in surface temperature from
summer to winter at temperate latitudes. This variation is controlled primarily by
the ocean’s response to seasonal changes in heating and cooling. The Northern
Hemisphere has one change in surface temperature controlled by its oceanic area
and another controlled by its land area. In the temperate latitudes of the Northern
Hemisphere, the land is much warmer than the oceanic area in summer and
much colder in winter. This situation creates large-scale seasonal changes in
atmospheric circulation and climate in the Northern Hemisphere that are not
found in the Southern Hemisphere.

HYPSOMETRY

Hypsometry is the science of measuring the elevation and depth of features


on Earth’s surface with respect to sea level. Data collected using
hypsometers, wire sounders, echo sounders, and satellite-based altimeters is
used to quantify the distribution of land at different elevations across a
given area and the surface-area distribution of the oceans and their marginal
seas with depth. Scientists can show how the areas of oceans, marginal seas,
and terrestrial basins change with elevation and depth using a special curve
known as a hypsometric, or hypsographic, curve.

MAJOR SUBDIVISIONS OF THE OCEANS


If the volume of an ocean is divided by its surface area, the mean depth is
obtained. With or without marginal seas, the Pacific is the largest ocean in both
surface area and volume, the Atlantic is next, and the Indian is the smallest. The
Atlantic exhibits the largest change in surface area and volume when its marginal
seas are subtracted. This indicates that the Atlantic has the greatest area of
bordering seas, many of which are shallow.
Hypsometry can show how the area of each ocean or marginal sea changes as
depth changes. A special curve known as a hypsometric, or hypsographic, curve
can be drawn that portrays how the surface area of Earth is distributed with
elevation and depth. This curve has been drawn to represent the total Earth and
all of its oceans; likewise, curves can be constructed for each individual ocean
and sea. The average depth of the world’s oceans, 3,795 metres (12,451 feet),
and the average elevation of the land, 840 metres (2,756 feet), are indicated. The
highest point on land, Mount Everest (8,850 metres [29,035 feet]), and the
deepest point in the ocean, located in the Mariana Trench (11,034 metres [36,201
feet]), mark the upper and lower limits of the curve, respectively. Since this
curve is drawn on a grid of elevation versus Earth’s area, the area under the
curve covering the 29.2 percent of Earth’s surface that is above sea level is the
volume of land above sea level. Similarly, the area between sea level and the
curve depicting the remaining 70.8 percent of Earth’s surface below sea level
represents the volume of water contained in the oceans.
Hypsographic curve showing how the surface area of Earth is
distributed with elevation and depth. Copyright Encyclopædia
Britannica, Inc.; rendering for this edition by Rosen Educational
Services

Portions of this curve describe the area of Earth’s surface that exists between
elevation or depth increments. On land, little of Earth’s total area—only about 4
percent—is at elevations above 2,000 metres (6,562 feet). Most of the land, 25.3
percent of the total Earth, is between 0 and 2,000 metres. About 13.6 percent of
the total land area is at higher elevations, with 86.4 percent between 0 and 2,000
metres when the areas are determined relative to land area only. In the oceans,
the percentages of the area devoted to depth increments yield information about
the typical structure and shape of the oceanic basins. The small depth increment
of 0–200 metres (0–660 feet) occupies about 5.4 percent of Earth’s total area or
7.6 percent of the oceans’ area. This approximates the world’s area of continental
shelves, the shallow flat borderlands of the continents that have been alternately
covered by the oceans during interglacial stages and uncovered during glacial
periods.
At depths between 200 and 1,000 metres (about 660 and 3,300 feet) and
between 1,000 and 2,000 metres (about 3,300 and 6,600 feet), an area only
slightly larger—6.02 percent of Earth’s total area or 8.5 percent of the oceans’
area—is found. These depths are related to the regions of the oceans that have
very steep slopes where depth increases rapidly. These are the continental slope
regions that mark the true edge of the continental landmasses. Marginal seas of
moderate depths and the tops of seamounts, however, add their area to these
depth zones when all the oceans are considered. The majority of the oceanic area
lies between 4,000 and 5,000 metres (about 13,100 and 16,400 feet).
The continental shelf region varies immensely from place to place. The
seaward boundary of the continental shelf historically is determined by the 100-
fathom, or 200-metre, depth contour. However, 85 fathoms, or 170 metres (about
560 feet), is a closer approximation. The true boundary at any given location is
marked by a rapid change in slope of the seafloor known as the shelf break. This
change in slope may be nearly at the coastline in areas where crustal plates
converge, as along the west coast of North and South America, or it may be
located more than 1,000 km (620 miles) seaward of the coast, as off the north
coast of Siberia. The average width of the shelf is about 75 km (47 miles), and
the shelf has an average slope of about 0.01°, a slope that is barely discernible to
the human eye. Seaward of the shelf break, the continental slope is inclined by
about 4°.

ORIGIN OF OCEAN WATERS


The huge volume of water contained in the oceans (and seas), 137 × 107 cubic
km (about 33 × 107 cubic miles), has been produced during the geologic history
of Earth. There is little information on the early history of Earth’s waters.
However, fossils dated from the Precambrian some 3.3 billion years ago show
that bacteria and cyanobacteria (blue-green algae) existed, indicating the
presence of water during this period. Carbonate sedimentary rocks, obviously
laid down in an aquatic environment, have been dated to 1 billion years ago.
Also, there is fossil evidence of primitive marine algae and invertebrates from
the outset of the Cambrian Period some 542 million years ago.
The presence of water on Earth at even earlier times is not documented by
physical evidence. It has been suggested, however, that the early hydrosphere
formed in response to condensation from the early atmosphere. The ratios of
certain elements on Earth indicate that the planet formed by the accumulation of
cosmic dust and was slowly warmed by radioactive and compressional heating.
This heating led to the gradual separation and migration of materials to form
Earth’s core, mantle, and crust. The early atmosphere is thought to have been
highly reducing and rich in gases, notably in hydrogen, and to include water
vapour.
Earth’s surface temperature and the partial pressures of the individual gases
in the early atmosphere affected the atmosphere’s equilibration with the
terrestrial surface. As time progressed and the planetary interior continued to
warm, the composition of the gases escaping from within Earth gradually
changed the properties of its atmosphere, producing a gaseous mixture rich in
carbon dioxide (CO2), carbon monoxide (CO), and molecular nitrogen (N2).
Photodissociation (i.e., separation due to the energy of light) of water vapour
into molecular hydrogen (H2) and molecular oxygen (O2) in the upper
atmosphere allowed the hydrogen to escape and led to a progressive increase of
the partial pressure of oxygen at Earth’s surface. The reaction of this oxygen
with the materials of the surface gradually caused the vapour pressure of water
vapour to increase to a level at which liquid water could form. This water in
liquid form accumulated in isolated depressions of Earth’s surface, forming the
nascent oceans.
The high carbon dioxide content of the atmosphere at this time would have
allowed a buildup of dissolved carbon dioxide in the water and made these early
oceans acidic and capable of dissolving surface rocks that would add to the
water’s salt content. Water must have evaporated and condensed rapidly and
accumulated slowly at first. The required buildup of atmospheric oxygen was
slow because much of this gas was used to oxidize methane, ammonia, and
exposed rocks high in iron. Gradually, the partial pressure of the oxygen gas in
the atmosphere rose as photosynthesis by bacteria and photodissociation
continued to supply oxygen. Biological processes involving algae increased, and
they gradually decreased the carbon dioxide content and increased the oxygen
content of the atmosphere until the oxygen produced by biological processes
outweighed that produced by photodissociation. This, in turn, accelerated the
formation of surface water and the development of the oceans.

COMPOSITION OF SEAWATER
The chemical composition of seawater is influenced by a wide variety of
chemical transport mechanisms. Rivers add dissolved and particulate chemicals
to the oceanic margins. Wind-borne particulates are carried to mid-ocean regions
thousands of kilometres from their continental source areas. Hydrothermal
solutions that have circulated through crustal materials beneath the seafloor add
both dissolved and particulate materials to the deep ocean. Organisms in the
upper ocean convert dissolved materials to solids, which eventually settle to
greater oceanic depths. Particulates in transit to the seafloor, as well as materials
both on and within the seafloor, undergo chemical exchange with surrounding
solutions. Through these local and regional chemical input and removal
mechanisms, each element in the oceans tends to exhibit spatial and temporal
concentration variations. Physical mixing in the oceans (thermohaline and wind-
driven circulation) tends to homogenize the chemical composition of seawater.
The opposing influences of physical mixing and of biogeochemical input and
removal mechanisms result in a substantial variety of chemical distributions in
the oceans.

DISSOLVED INORGANIC SUBSTANCES


In contrast to the behaviour of most oceanic substances, the concentrations of the
principal inorganic constituents of the oceans are remarkably constant.
Calculations indicate that, for the main constituents of seawater, the time
required for thorough oceanic mixing is quite short compared with the time that
would be required for input or removal processes to significantly change a
constituent’s concentration. The concentrations of the principal constituents of
the oceans vary primarily in response to a comparatively rapid exchange of
water (precipitation and evaporation), with relative concentrations remaining
nearly constant.
Salinity is used by oceanographers as a measure of the total salt content of
seawater. Practical salinity, symbol S, is determined through measurements of a
ratio between the electrical conductivity of seawater and the electrical
conductivity of a standard solution. Practical salinity can be used to calculate
precisely the density of seawater samples. Because of the constant relative
proportions of the principal constituents, salinity can also be used to directly
calculate the concentrations of the major ions in seawater. The measure of
practical salinity was originally developed to provide an approximate measure of
the total mass of salt in 1 kg (2.2 pounds) of seawater. Seawater with S equal to
35 contains approximately 35 grams (1.2 ounces) of salt and 965 grams (34
ounces) of water.

Many other constituents are of great importance to the biogeochemistry of


the oceans. Such chemicals as inorganic phosphorus (HPO2−/4 and PO3−/4) and
inorganic nitrogen (NO−/3, NO−/2, and NH+/4) are essential to the growth of
marine organisms. Nitrogen and phosphorus are incorporated into the tissues of
marine organisms in approximately a 16:1 ratio and are eventually returned to
solution in approximately the same proportion. As a consequence, in much of the
oceanic waters dissolved inorganic phosphorus and nitrogen exhibit a close
covariance. Dissolved inorganic phosphorus distributions in the Pacific Ocean
strongly bear the imprint of phosphorus incorporation by organisms in the
surface waters of the ocean and of the return of the phosphorus to solution via a
rain of biological debris remineralized in the deep ocean. Inorganic phosphate
concentrations in the western Pacific range from somewhat less than 0.1
micromole per kg (1 × 10−7 mole per kg) at the surface to approximately 3
micromoles/kg (3 × 10−6 mole/kg) at depth. Inorganic nitrogen ranges between
somewhat less than 1 micromole/kg and 45 micromoles/kg along the same
section of ocean and exhibits a striking covariance with phosphate.
A variety of elements essential to the growth of marine organisms, as well as
some elements that have no known biological function, exhibit nutrient-like
behaviour broadly similar to nitrate and phosphate. Silicate is incorporated into
the hard structural parts of certain types of marine organisms (diatoms and
radiolarians) that are abundant in the upper ocean. Dissolved silicate
concentrations range between less than 1 micromole/kg (1 × 10−6 mole/kg) in
surface waters to approximately 180 micromoles/kg (1.8 × 10−4 mole/kg) in the
deep North Pacific. The concentration of zinc, a metal essential to a variety of
biological functions, ranges between approximately 0.05 nanomole/kg (5 × 10−11
mole/kg) in the surface ocean to as much as 6 nanomoles/kg (6 × 10−9 mole/kg)
in the deep Pacific. The distribution of zinc in the oceans is observed to
generally parallel silicate distributions. Cadmium, though having no known
biological function, generally exhibits distributions that are covariant with
phosphate and concentrations that are even lower than those of zinc.
Many elements, including the essential trace metals iron, cobalt, and copper,
show surface depletions but in general exhibit behaviour more complex than that
of phosphate, nitrate, and silicate. Some of the complexities observed in
elemental oceanic distributions are attributable to the adsorption of elements on
the surface of sinking particles. Adsorptive processes, either exclusive of or in
addition to biological uptake, serve to remove elements from the upper ocean
and deliver them to greater depths. The distribution patterns of a number of trace
elements are complicated by their participation in oxidation-reduction (electron-
exchange) reactions. In general, electron-exchange reactions lead to profound
changes in the solubility and reactivity of trace metals in seawater. Such
reactions are important to the oceanic behaviour of a variety of elements,
including iron, manganese, copper, cobalt, chromium, and cerium.
The processes that deliver dissolved, particulate, and gaseous materials to the
oceans ensure that they contain, at some concentration, very nearly every
element that is found in Earth’s crust and atmosphere. The principal components
of the atmosphere, nitrogen (78.1 percent), oxygen (21.0 percent), argon (0.93
percent), and carbon dioxide (0.035 percent), occur in seawater in variable
proportions, depending on their solubilities and oceanic chemical reactions. In
equilibrium with the atmosphere, the concentrations of the unreactive gases,
nitrogen and argon, in seawater (0 °C [32 °F], salinity 35) are 616
micromoles/kg and 17 micromoles/kg, respectively. For seawater at 35 °C (95
°F), these concentrations would decrease by approximately a factor of two. The
solubility behaviours of argon and oxygen are quite similar. For seawater in
equilibrium with the atmosphere, the ratio of oxygen and argon concentrations is
approximately 20.45. Since oxygen is a reactive gas essential to life, oxygen
concentrations in seawater that are not in direct equilibrium with the atmosphere
are quite variable. Although oxygen is produced by photosynthetic organisms at
shallow, sunlit ocean depths, oxygen concentrations in near-surface waters are
established primarily by exchange with the atmosphere. Oxygen concentrations
in the oceans generally exhibit minimum values at intermediate depths and
relatively high values in deep waters. This distribution pattern results from a
combination of biological oxygen utilization and physical mixing of the ocean
waters. Estimates of the extent of oxygen utilization in the oceans can be
obtained by comparing concentrations of oxygen with those of argon, since the
latter are only influenced by physical processes. The physical processes that
influence oxygen distributions include, in particular, the large-scale
replenishment of oceanic bottom waters with cold, dense, oxygen-rich waters
sinking toward the bottom from high latitudes. Due to the release of nutrients
that accompanies the consumption of oxygen by biological debris, dissolved
oxygen concentrations generally appear as a mirror image of dissolved nutrient
concentrations.
While the atmosphere is a vast repository of oxygen compared with the
oceans, the total carbon dioxide content of the oceans is very large compared
with that of the atmosphere. Carbon dioxide reacts with water in seawater to
form carbonic acid (H2CO3), bicarbonate ions (HCO−/3), and carbonate ions
(CO2-/3). Approximately 90 percent of the total organic carbon in seawater is
present as bicarbonate ions. The formation of bicarbonate and carbonate ions
from carbon dioxide is accompanied by the liberation of hydrogen ions (H+).
Reactions between hydrogen ions and the various forms of inorganic carbon
buffer the acidity of seawater. The relatively high concentrations of both total
inorganic carbon and boron—as B(OH)3 and B(OH)−/4—in seawater are
sufficient to maintain the pH of seawater between 7.4 and 8.3. (The term pH is
defined as the negative logarithm of the hydrogen ion concentration in moles per
kg. Thus, a pH equal to 8 is equivalent to 1 × 10−8 mole of H+ ions per kg of
seawater.) This is quite important because the extent and rate of many reactions
in seawater are highly pH-dependent. Carbon dioxide produced by the
combination of oxygen and organic carbon generally produces an acidity
maximum (pH minimum) near the depth of the oxygen minimum in seawater. In
addition to exchange with the atmosphere and, through respiration, with the
biosphere, dissolved inorganic carbon concentrations in seawater are influenced
by the formation and dissolution of the calcareous shells (CaCO3) of organisms
(foraminiferans, coccolithophores, and pteropods) abundant in the upper ocean.

DISSOLVED ORGANIC SUBSTANCES


Processes involving dissolved and particulate organic carbon are of central
importance in shaping the chemical character of seawater. Marine organic carbon
principally originates in the uppermost 100 metres (330 feet) of the oceans
where dissolved inorganic carbon is photosynthetically converted to organic
materials. The “rain” of organic-rich particulate materials, resulting directly and
indirectly from photosynthetic production, is a principal factor behind the
distributions of many organic and inorganic substances in the oceans. A large
fraction of the vertical flux of materials in the uppermost waters is converted to
dissolved substances within the upper 400 metres (about 1,300 feet) of the
oceans. Dissolved organic carbon (DOC) accounts for at least 90 percent of the
total organic carbon in the oceans. Estimates of DOC appropriate to the surface
of the open ocean range between roughly 100 and 500 micromoles of carbon per
kg of seawater. DOC concentrations in the deep ocean are 5 to 10 times lower
than surface values.
DOC occurs in an extraordinary variety of forms, and, in general, its
composition is controversial and poorly understood. Conventional techniques
have indicated that, in surface waters, about 15 percent of DOC can be identified
as carbohydrates and combined amino acids. At least 1–2 percent of DOC in
surface waters occurs as lipids and 20–25 percent as relatively unreactive humic
substances. The relative abundances of reactive organic substances, such as
amino acids and carbohydrates, are considerably reduced in deep ocean waters.
Dissolved and particulate organic carbon in the surface ocean participates in diel
cycles (i.e., those of a 24-hour period) related to photosynthetic production and
photochemical transformations. The influence of dissolved organic matter on
ocean chemistry is often out of proportion to its oceanic abundance.
Photochemical reactions involving DOC can influence the chemistry of vital
trace nutrients such as iron, and, even at dissolved concentrations on the order of
one nanomole/kg (1 × 10−9 mole/kg), dissolved organic substances in the upper
ocean waters are capable of greatly altering the bioavailability of essential trace
nutrients, as, for example, copper and zinc.

A drainpipe spews untreated sewage near Gaza, releasing harmful


chemicals into the Mediterranean Sea. Human activity is a prime
source of such toxins. Warrick Page/Getty Images

EFFECTS OF HUMAN ACTIVITIES


Although the oceans constitute an enormous reservoir, human activities have
begun to influence their composition on both a local and a global scale. The
addition of nutrients (through the discharge of untreated sewage or the seepage
of soluble mineral fertilizers, for example) to coastal waters results in increased
phytoplankton growth, high levels of dissolved and particulate organic materials,
decreased penetration of light through seawater, and alteration of the community
structure of bottom-dwelling organisms. Through industrial and automotive
emissions, lead concentrations in the surface ocean have increased dramatically
on a global scale compared with preindustrial levels. Certain toxic organic
compounds, such as polychlorinated biphenyls (PCBs), are found in seawater
and marine organisms and are attributable solely to the activities of humankind.
Although most radioactivity in seawater is natural (approximately 90 percent
as potassium-40 and less than 1 percent each as rubidium-87 and uranium-238),
strontium-90 and certain other artificial radioisotopes have unique environmental
pathways and potential for bioaccumulation. Among the most dramatic
influences of human activities on a global scale is the remarkable increase of
carbon dioxide levels in the atmosphere. Atmospheric carbon dioxide levels are
expected to double by the middle of the 21st century, with potentially profound
consequences for global climate and agricultural patterns. It is thought that the
oceans, as a great reservoir of carbon dioxide, will ameliorate this consequence
of human activities to some degree.

THE PHYSICAL PROPERTIES OF SEAWATER


Water is a unique substance. Not only is water the most abundant substance at
Earth’s surface, but it also has the most naturally occurring physical states of any
Earth material or substance (solid, liquid, and gas) and the greatest capacity to
do things without being altered significantly. It is essential for sustaining life on
Earth and affects the physical environment in a myriad of ways, as evidenced by
the sculpting of landscape features by moving water, the maintaining of Earth’s
radiation balance by atmospheric water vapour transfer, and the transporting of
inorganic and organic materials about the planet’s surface by the oceans. The
addition of salt to water changes the behaviour of water only slightly.

SALINITY DISTRIBUTION
A discussion of salinity, the salt content of the oceans, requires an understanding
of two important concepts: (1) the present-day oceans are considered to be in
steady state, receiving as much salt as they lose, and (2) the oceans have been
mixed over such a long time period that the composition of sea salt is
everywhere the same in the open ocean. This uniformity of salt content results in
oceans in which the salinity varies little over space or time.
The range of salinity observed in the open ocean is from 33 to 37 grams (1.2
to 1.3 ounces) of salt per kg of seawater or parts per thousand (0/00). For the most
part, the observed departure from a mean value of approximately 350/00 is caused
by processes at Earth’s surface that locally add or remove fresh water. Regions
of high evaporation have elevated surface salinities, while regions of high
precipitation have depressed surface salinities. In nearshore regions close to
large freshwater sources, the salinity may be lowered by dilution. This is
especially true in areas where the region of the ocean receiving the fresh water is
isolated from the open ocean by the geography of the land.
Areas of the Baltic Sea may have salinity values depressed to 100/00 or less.
Increased salinity by evaporation is accentuated where isolation of the water
occurs. This effect is found in the Red Sea, where the surface salinity rises to
410/00. Coastal lagoon salinities in areas of high evaporation may be much higher.
The removal of fresh water by evaporation or the addition of fresh water by
precipitation does not affect the constancy of composition of the sea salt in the
open sea. A river draining a particular soil type, however, may bring to the
oceans only certain salts that will locally alter the salt composition. In areas of
high evaporation where the salinity is driven to very high values, precipitation of
particular salts may alter the composition too. At high latitudes where sea ice
forms seasonally, the salinity of the seawater is elevated during ice formation
and reduced when the ice melts.
At depth in the oceans, salinity may be altered as seawater percolates into
fissures associated with deep-ocean ridges and crustal rifts involving volcanism.
This water then returns to the ocean as superheated water carrying dissolved
salts from the magmatic material within the crust. It may lose much of its
dissolved load to precipitates on the seafloor and gradually blend in with the
surrounding seawater, sharing its remaining dissolved substances.
Salt concentrations as high as 2560/00 have been found in hot but dense pools
of brine trapped in depressions at the bottom of the Red Sea. The composition of
the salts in these pools is not the same as the sea salt of the open oceans.
The salinities of the open oceans found at the greater depths are quite
uniform in both time and space with average values of 34.5 to 350/00. These
salinities are determined by surface processes such as those described above
when the water, now at depth, was last in contact with the surface.
The intertropical convergence, with its high precipitation centred about 5° N,
supports the tropical rainforests of the world and leaves its imprint on the oceans
as a latitudinal depression of surface salinity. At approximately 30°–35° N and
30°–35° S, the subtropical zones called the horse latitudes are belts of high
evaporation that produce major deserts and grasslands on the continents and
cause the surface salinity to rise. At 50°–60° N and 50°–60° S, precipitation
again increases.
HORSE LATITUDES

The horse latitudes are two subtropical atmospheric high-pressure belts that
encircle Earth around latitudes 30°–35° N and 30°–35° S and that generate
light winds and clear skies. Because they contain dry, subsiding air, they
produce arid climates in the areas below them. The Sahara, for example, is
situated in a horse latitude. The Southern Hemisphere, which has more
water area than the Northern, has the more continuous belt of subsiding air.
The belts contain several separate high-pressure centres and shift a few
degrees away from the equator in summer.
The horse latitudes were named by the crews of sailing ships, who
sometimes threw horses overboard to conserve water when their ships were
becalmed in the high-pressure belts.

TEMPERATURE DISTRIBUTION
Mid-ocean surface temperatures vary with latitude in response to the balance
between incoming solar radiation and outgoing long-wave radiation. There is an
excess of incoming solar radiation at latitudes less than approximately 45° and
an excess of radiation loss at latitudes higher than approximately 45°.
Superimposed on this radiation balance are seasonal changes in the intensity of
solar radiation and the duration of daylight hours due to the tilt of Earth’s axis to
the plane of the ecliptic and the rotation of the planet about this axis. The
combined effect of these variables is that average ocean surface temperatures are
higher at low latitudes than at high latitudes. Because the Sun, with respect to
Earth, migrates annually between the Tropic of Cancer and the Tropic of
Capricorn, the yearly change in heating of Earth’s surface is small at low
latitudes and large at mid-and higher latitudes.
Water has an extremely high heat capacity, and heat is mixed downward
during summer surface-heating conditions and upward during winter surface
cooling. This heat transfer reduces the actual change in ocean surface
temperatures over the annual cycle. In the tropics the ocean surface is warm
year-round, varying seasonally about 1 to 2 °C (1.8 to 3.6 °F). At mid-latitudes
the mid-ocean temperatures vary about 8 °C (14.4 °F) over the year. At the polar
latitudes the surface temperature remains near the ice point of seawater—about
−1.9 °C (28.6 °F).
Land temperatures have a large annual range at high latitudes because of the
low heat capacity of the land surface. Proximity to land, isolation of water from
the open ocean, and processes that control stability of the surface water combine
to increase the annual range of nearshore ocean surface temperature.
In winter, prevailing winds carry cold air masses off the continents in
temperate and subarctic latitudes, cooling the adjacent surface seawater below
that of the mid-ocean level. In summer, the opposite effect occurs, as warm
continental air masses move out over the adjacent sea. This creates a greater
annual range in sea surface temperatures at mid-latitudes on the western sides of
the oceans of the Northern Hemisphere but has only a small effect in the
Southern Hemisphere as there is little land present. Instead, the oceans of the
Southern Hemisphere act to control the air temperature, which in turn influences
the land temperatures of the temperate zone and reduces the annual temperature
range over the land.
Currents carry water having the characteristics of one latitudinal zone to
another zone. The northward displacement of warm water to higher latitudes by
the Gulf Stream of the North Atlantic and the Kuroshio (Japan Current) of the
North Pacific creates sharp changes in temperature along the current boundaries
or thermal fronts, where these northward-moving flows meet colder water
flowing southward from higher latitudes. Cold water currents flowing from
higher to lower latitudes also displace surface isotherms from near constant
latitudinal positions. At low latitudes the trade winds act to move water away
from the lee coasts of the landmasses to produce areas of coastal upwelling of
water from depth and reduce surface temperatures.
Temperatures in the oceans decrease with increasing depth. There are no
seasonal changes at the greater depths. The temperature range extends from 30
°C (86 °F) at the sea surface to −1 °C (30.2 °F) at the seabed. Like salinity, the
temperature at depth is determined by the conditions that the water encountered
when it was last at the surface. In the low latitudes the temperature change from
top to bottom in the oceans is large. In high temperate and Arctic regions, the
formation of dense water at the surface that sinks to depth produces nearly
isothermal conditions with depth.
Areas of the oceans that experience an annual change in surface heating have
a shallow wind-mixed layer of elevated temperature in the summer. Below this
nearly isothermal layer 10 to 20 metres (about 33 to 66 feet) thick, the
temperature decreases rapidly with depth, forming a shallow seasonal
thermocline (i.e., layer of sharp vertical temperature change). During winter
cooling and increased wind mixing at the ocean surface, convective overturning
and mixing erase this shallow thermocline and deepen the isothermal layer. The
seasonal thermocline re-forms when summer returns. At greater depths, a weaker
nonseasonal thermocline is found separating water from temperate and subpolar
sources.
Below this permanent thermocline, temperatures decrease slowly. In the very
deep ocean basins, the temperature may be observed to increase slightly with
depth. This occurs when the deepest parts of the oceans are filled by water with a
single temperature from a common source. This water experiences an adiabatic
temperature rise as it sinks. Such a temperature rise does not make the water
column unstable because the increased temperature is caused by compression,
which increases the density of the water. For example, surface seawater of 2 °C
(35.6 °F) sinking to a depth of 10,000 metres (32,800 feet) increases its
temperature by about 1.3 °C (2.3 °F). When measuring deep-sea temperatures,
the adiabatic temperature rise, which is a function of salinity, initial temperature,
and pressure change, is calculated and subtracted from the observed temperature
to obtain the potential temperature. Potential temperatures are used to identify a
common type of water and to trace this water back to its source.

THERMAL PROPERTIES
The unit of heat called the gram calorie is defined as the amount of heat required
to raise the temperature of 1 gram (0.04 ounce) of water 1 °C (1.8 °F). The
kilocalorie, or food calorie, is the amount of heat required to raise 1 kg (2.2
pounds) of water 1 °C. Heat capacity is the amount of heat required to raise 1
gram of material 1 °C under constant pressure. In the International System of
Units (SI), the heat capacity of water is 1 kilocalorie per kg per degree Celsius.
Water has the highest heat capacity of all common Earth materials; therefore,
water on Earth acts as a thermal buffer, resisting temperature change as it gains
or loses heat energy.
The heat capacity of any material can be divided by the heat capacity of
water to give a ratio known as the specific heat of the material. Specific heat is
numerically equal to heat capacity but has no units. In other words, it is a ratio
without units. When salt is present, the heat capacity of water decreases slightly.
Seawater of 350/00 has a specific heat of 0.932 compared to 1.000 for pure water.
Pure water freezes at 0 °C (32 °F) and boils at 100 °C (212 °F) under normal
pressure conditions. When salt is added, the freezing point is lowered and the
boiling point is raised. The addition of salt also lowers the temperature of
maximum density below that of pure water (4 °C [39.2 °F]). The temperature of
maximum density decreases faster than the freezing point as salt is added.
At 300/00 salinity, the temperature of maximum density is lower than the
initial freezing point of saltwater. Therefore, a maximum density is never
achieved, as seawater of this salinity is cooled because freezing occurs first. At
24.700/00 salinity, the freezing point and the temperature of maximum density
coincide at −1.332 °C (29.6 °F). At salinities typical of the open oceans, which
are greater than 24.70/00, the freezing point is always higher than the temperature
of maximum density.
When water changes its state, hydrogen bonds between molecules are either
formed or broken. Energy is required to break the hydrogen bonds, which allows
water to pass from a solid to a liquid state or from a liquid to a gaseous state.
When hydrogen bonds are formed, permitting water to change from a liquid to a
solid or from a gas to a liquid, energy is liberated. The heat energy input required
to change water from a solid at 0 °C to a liquid at 0 °C is the latent heat of fusion
and is 80 calories per gram of ice. Water’s latent heat of fusion is the highest of
all common materials. Because of this, heat is released when ice forms and is
absorbed during melting, which tends to buffer air temperatures as land and sea
ice form and melt seasonally.
When water converts from a liquid to a gas, a quantity of heat energy known
as the latent heat of vaporization is required to break the hydrogen bonds. At 100
°C, 540 calories per gram of water are needed to convert 1 gram of liquid water
to 1 gram of water vapour under normal pressure. Water can evaporate at
temperatures below the boiling point, and ice can evaporate into a gas without
first melting in a process called sublimation. Evaporation below 100 °C and
sublimation require more energy per gram than 540 calories. At 20 °C (68 °F)
about 585 calories are required to vaporize 1 gram of water. When water vapour
condenses back to liquid water, the latent heat of vaporization is liberated. The
evaporation of water from the surface of Earth and its condensation in the
atmosphere constitute the single most important way that heat from Earth’s
surface is transferred to the atmosphere. This process is the source of the power
that drives hurricanes and a principal mechanism for cooling the surface of the
oceans. The latent heat of vaporization of water is the highest of all common
substances.

DENSITY OF SEAWATER AND PRESSURE


The density of a material is given in units of mass per unit volume and expressed
in kilograms per cubic metre in the SI system of units. In oceanography the
density of seawater has been expressed historically in grams per cubic
centimetre. The density of seawater is a function of temperature, salinity, and
pressure. Because oceanographers require density measurements to be accurate
to the fifth decimal place, manipulation of the data requires writing many
numbers to record each measurement. Also, the pressure effect can be neglected
in many instances by using potential temperature. These two factors led
oceanographers to adopt a density unit called sigma-t (σt). This value is obtained
by subtracting 1.0 from the density and multiplying the remainder by 1,000. The
σt has no units and is an abbreviated density of seawater controlled by salinity
and temperature only. The σt of seawater increases with increasing salinity and
decreasing temperature.

The relationship between pressure and density is demonstrated by observing


the effect of pressure on the density of seawater at 350/00 and 0 °C (32 °F).
Because a 1-metre (3.3-foot) column of seawater produces a pressure of about 1
decibar (0.1 atmosphere), the pressure in decibars is approximately equal to the
depth in metres. (One decibar is one-tenth of a bar, which in turn is equal to 105
newtons per square metre.)
Increasing density values demonstrate the compressibility of seawater under
the tremendous pressures present in the deep ocean. If the average pressure over
4,000 metres (13,100 feet: the approximate mean depth of the ocean) is
calculated, it is found to be approximated by that at 2,000 metres (about 6,600
feet). The average volume change due to pressure for each gram of water in the
entire water column is (1/1.02813–1/1.03747) cm3/g, or 0.00876 cm3/g. Because
the number of grams of water in a column of seawater 4 × 105 cm in length is
equal to the number of centimetres times the average density of the water,
1.03747 g/cm3, the expansion of the entire water column is about 4 × 105 cm ×
0.00876 cm3/g × 1.03747 g/cm3, or an average sea level rise of about 36 metres
(118 feet) if the area of the oceans is considered constant.

The temperature of maximum density and the freezing point of water


decrease as salt is added to water, and the temperature of maximum density
decreases more rapidly than the freezing point. At salinities less than 24.70/00 the
density maximum is reached before the ice point, while at the higher salinities
more typical of the open oceans the maximum density is never achieved
naturally. This ability of low-salinity water and, of course, fresh water to pass
through a density maximum makes them both behave differently from marine
systems when water is cooled at the surface and density-driven overturn occurs.
During the fall a lake is cooled at its surface, the surface water sinks, and
convective overturn proceeds as the density of the surface water increases with
the decreasing temperature. By the time the surface water reaches 4 °C (39.2 °F),
the temperature of maximum density for fresh water, the density-driven
convective overturn has reached the bottom of the lake, and overturn ceases.
Further cooling of the surface produces less dense water, and the lake becomes
stably stratified with regard to temperature-controlled density. Only a relatively
shallow surface layer is cooled below 4 °C. When this surface layer is cooled to
the ice point, 0 °C, ice is formed as the latent heat of fusion is extracted. In a
deep lake the temperature at depth remains at 4 °C. In the spring the surface
water warms up and the ice melts. A shallow convective overturn resumes until
the lake is once more isothermal at 4 °C. Continued warming of the surface
produces a stable water column.
In seawater in which the salinity exceeds 24.70/00, convective overturn also
occurs during the cooling cycle and penetrates to a depth determined by the
salinity and temperature-controlled density of the cooled water. Since no density
maximum is passed, the thermally driven convective overturn is continuous until
the ice point is reached where sea ice forms with the extraction of the latent heat
of fusion. Since salt is largely excluded from the ice in most cases, the salinity of
the water beneath the ice increases slightly and a convective overturn that is both
salt-and temperature-driven continues as sea ice forms.
The continuing overturn requires that a large volume of water be cooled to a
new ice point dictated by the salinity increase before additional ice forms. In this
manner, very dense seawater that is both cold and of elevated salinity is formed.
Such areas as the Weddell Sea in Antarctica produce the densest water of the
oceans. This water, known as Antarctic Bottom Water, sinks to the deepest
depths of the oceans. The continuing overturn slows the rate at which the sea ice
forms, limiting the seasonal thickness of the ice. Other factors that control the
thickness of ice are the rate at which heat is conducted through the ice layer and
the insulation provided by snow on the ice. Seasonal sea ice seldom exceeds
about 2 metres (6.6 feet) in thickness. During the warmer season, melting sea ice
supplies a freshwater layer to the sea surface and thereby stabilizes the water
column.
Surface processes that alter the temperature and salinity of seawater drive the
vertical circulation of the oceans. Known as thermohaline circulation, it
continually replaces seawater at depth with water from the surface and slowly
replaces surface water elsewhere with water rising from deeper depths.

OPTICAL PROPERTIES
Water is transparent to the wavelengths of electromagnetic radiation that fall
within the visible spectrum and is opaque to wavelengths above and below this
band. However, once in the water, visible light is subject to both refraction and
attenuation.
Light rays that enter the water at any angle other than a right angle are
refracted (i.e., bent) because the light waves travel at a slower speed in water
than they do in air. The amount of refraction, referred to as the refractive index,
is affected by both the salinity and temperature of the water. The refractive index
increases with increasing salinity and decreasing temperature. This relationship
allows the refractive index of a sample of seawater at a constant temperature to
be used to determine the salinity of the sample.
Some of the Sun’s radiant energy is reflected at the ocean surface and does
not enter the ocean. That which penetrates the water’s surface is attenuated by
absorption and conversion to other forms of energy, such as heat that warms or
evaporates water, or is used by plants to fuel photosynthesis. Sunlight that is not
absorbed can be scattered by molecules and particulates suspended in the water.
Scattered light is deflected into new directional paths and may wander randomly
to eventually be either absorbed or directed upward and out of the water. It is this
upward scattered light and the light reflected from particles that determine the
colour of the oceans, as seen from above.
Water molecules, dissolved salts, organic substances, and suspended
particulates combine to cause the intensity of available solar radiation to
decrease with depth. Observations of light attenuation in ocean waters indicate
that not only does the intensity of solar radiation decrease with depth but also the
wavelengths present in the solar spectrum are not attenuated at the same rates.
Both short wavelengths (ultraviolet) and long wavelengths (infrared) are
absorbed rapidly and are not available for scattering. Only blue-green
wavelengths penetrate to any depth, and because the blue-green light is most
available for scattering, the oceans appear blue to the human eye. Changes in the
colour of the ocean waters are caused either by the colour of the particulates in
suspension and dissolved substances or by the changing quality of the solar
radiation at the ocean surface as determined by the angle of the Sun and
atmospheric conditions. In the clearest ocean waters only about 1 percent of the
surface radiation remains at a depth of 150 metres (about 500 feet). No sunlight
penetrates below 1,000 metres (3,300 feet).
There are many ways of measuring light attenuation in the oceans. A
common method involves the use of a Secchi disk, a weighted round white disk
about 30 cm (11.8 inches) in diameter. The Secchi disk is lowered into the ocean
to the depth where it disappears from view; its reflectance equals the intensity of
light backscattered from the water. This depth in metres divided into 1.7 yields
an attenuation, or extinction, coefficient for available light as averaged over the
Secchi disk depth. The light extinction coefficient, x, may then be used in a form
of Beer’s law, Iz = I0exz, to estimate Iz, the intensity of light at depth z from I0,
the intensity of light at the ocean surface. This method gives no indication of the
attenuation change with depth or the attenuation of specific wavelengths of light.

A photocell may be lowered into the ocean to measure light intensity at


discrete depths and to determine light reduction from the surface value or from
the previous depth value. The photocell may sense all available wavelengths or
may be equipped with filters that pass only certain wavelengths of light. Since Iz
and I0 are known, changing light intensity values may be used in Beer’s law to
determine how the attenuation coefficient changes with depth and quality of
light. Measurements of this type are used to determine the level of
photosynthesis as a function of radiant energy level with depth and to measure
changes in the turbidity of the water caused by particulate distribution with
depth.
Different areas of the oceans tend to have different optical properties. Near
rivers, silt increases the suspended particle effect. Where nutrients and sunlight
are abundant, phytoplankton (unicellular plants) increase the opacity of the water
and lend it their colour. Organic substances from excretion and decomposition
also have colour and absorb light.
Solar radiation received at the ocean surface is constantly changing in time
and space. Cloud cover, atmospheric dust, atmospheric gas composition,
roughness of the ocean surface, and elevation angle of the Sun combine to
change both the quality and quantity of light that enters the ocean. When the
Sun’s rays are perpendicular to a smooth ocean surface, reflectance is low. When
the solar rays are oblique to the ocean surface, reflectance is increased. If the
ocean is rough with waves, reflectance is increased when the Sun is at high
elevation and decreased when it is at low elevation. Since latitude plays a role in
the elevation of the Sun above the horizon, light penetration is always less at the
higher latitudes. Cloud cover, density layering, fog, and dust cause refraction and
atmospheric scattering of sunlight. When strongly scattered, the Sun’s rays are
not unidirectional and there are no shadows. Light enters the ocean from all
angles under this condition, and the elevation angle of the Sun loses its
importance in controlling surface reflectance. The solar energy available to
penetrate the ocean is 100 percent minus the tabulated reflectance value.

These data indicate that water is a good absorber of solar radiation.

ACOUSTIC PROPERTIES
Water is an excellent conductor of sound, considerably better than air. The
attenuation of sound by absorption and conversion to other energy forms is a
function of sound frequency and the properties of water.
The attenuation coefficient, x, in Beer’s law, as applied to sound, where Iz
and I0 are now sound intensity values, is dependent on the viscosity of water and
inversely proportional to the frequency of the sound and the density of the water.
High-pitched sounds are absorbed and converted to heat faster than low-pitched
sounds. Sound velocity in water is determined by the square root of elasticity
divided by the water’s density. Because water is only slightly compressible, it
has a large value of elasticity and therefore conducts sound rapidly. Since both
the elasticity and density of seawater change with temperature, salinity, and
pressure, so does the velocity of sound.
In the oceans the speed of sound varies between 1,450 and 1,570 metres
(4,757 and 5,151 feet) per second. It increases about 4.5 metres (14.8 feet) per
second per each degree C (1.8 °F) increase and 1.3 metres (4.3 feet) per second
per each 10/00 increase in salinity. Increasing pressure also increases the speed of
sound at the rate of about 1.7 metres (5.6 feet) per second for an increase in
pressure of 100 metres (about 330 feet) in depth, which is equal to
approximately 10 bars, or 10 atmospheres.
The greatest changes in temperature and salinity with depth that affect the
speed of sound are found near the surface. Changes of sound speed in the
horizontal are usually slight except in areas where abrupt boundaries exist
between waters of different properties. The effects of salinity and temperature on
sound speed are more important than the effect of pressure in the upper layers.
Deeper in the ocean, salinity and temperature change less with depth, and
pressure becomes the important controlling factor.
In regions of surface dilution, salinity increases with depth near the surface,
while in areas of high evaporation salinity decreases with depth. Temperature
usually decreases with depth and normally exerts a greater influence on sound
speed than does the salinity in the surface layer of the open oceans. In the case of
surface dilution, salinity and temperature effects on the speed of sound oppose
each other, while in the case of evaporation they reinforce each other, causing
the speed of sound to decrease with depth. Beneath the upper oceanic layers the
speed of sound increases with depth.
If a sound wave (sonic pulse) travels at a right angle to these layers, as in
depth sounding, no refraction occurs; however, the speed changes continuously
with depth, and an average sound speed for the entire water column must be
used to determine the depth of water. Variations in the speed of sound cause
sound waves to refract when they travel obliquely through layers of water that
have different properties of salinity and temperature. Sound waves traveling
downward and moving obliquely to the water layers will bend upward when the
speed of sound increases with depth and downward when the speed decreases
with depth. This refraction of the sound is important in the sonar detection of
submarines because the actual path of a sound wave must be known to determine
a submarine’s position relative to the transmitter of the sound. Refraction also
produces shadow zones that sound waves do not penetrate because of their
curvature.
At depths of approximately 1,000 metres (about 3,300 feet), pressure
becomes the important factor: it combines with temperature and salinity to
produce a zone of minimum sound speed. This zone has been named the SOFAR
(sound fixing and ranging) channel. If a sound is generated by a point source in
the SOFAR zone, it becomes trapped by refraction. Dispersed horizontally rather
than in three directions, the sound is able to travel for great distances.
Hydrophones lowered to this depth many kilometres from the origin of the sound
are able to detect the sound pulse. The difference in arrival time of the pulse at
separate listening posts may be used to triangulate the position of the pulse
source.
Hearing is an important sensory mechanism for marine animals because
seawater is more transparent to sound than to light. Animals communicate with
each other over long distances and also locate objects by sending directional
sound signals that reflect from targets and are received as echoes. Information
about the size of a target is gained by varying the frequency of the sound; high-
frequency (or short-wavelength) sound waves reflect better from small targets
than low-frequency sound waves. The intensity and quality of the returning
signal also provide information about the properties of the reflecting target.

CHEMICAL EVOLUTION OF THE OCEANS


The chemical history of the oceans has been divided into three stages. The first
is an early stage in which the Earth’s crust was cooling and reacting with volatile
or highly reactive gases of an acidic, reducing nature to produce the oceans and
an initial sedimentary rock mass. This stage lasted until about 3.5 billion years
ago. The second stage was a period of transition from the initial to essentially
modern conditions, and it is estimated to have ended 2 to 1.5 billion years ago.
Since that time it is likely that there has been little change in seawater
composition.

THE EARLY OCEANS


The initial accretion of Earth by agglomeration of solid particles occurred about
4.6 billion years ago. Heating of this initially cool, unsorted conglomerate by the
decay of radioactive elements and the conversion of kinetic and potential energy
to heat resulted in the development of a liquid iron core and the gross internal
zonation of Earth. It has been concluded that formation of Earth’s core took
about 500 million years. It is likely that core formation resulted in the escape of
an original primitive atmosphere and its replacement by one derived from loss of
volatile substances from Earth’s interior. Whether most of this degassing took
place during core formation or soon afterward or whether there has been
significant degassing of Earth’s interior throughout geologic time is uncertain.
Recent models of Earth formation, however, suggest early differentiation of
Earth into three major zones (core, mantle, and crust) and attendant early loss of
volatile substances from the interior. It is also likely that Earth, after initial cold
agglomeration, reached temperatures such that the whole Earth approached the
molten state. As the initial crust of Earth solidified, volatile gases would be
released to form an atmosphere that would contain water, later to become the
hydrosphere; carbon gases, such as carbon dioxide, methane, and carbon
monoxide; sulfur gases, mostly hydrogen sulfide; and halogen compounds, such
as hydrochloric acid. Nitrogen also may have been present, along with minor
amounts of other gases. Gases of low atomic number, such as hydrogen and
helium, would escape Earth’s gravitational field. Substances degassed from the
planetary interior have been called excess volatiles because their masses cannot
be accounted for simply by rock weathering.
At an initial crustal temperature of about 600 °C (1,100 °F), almost all these
compounds, including water (H2O), would be in the atmosphere. The sequence
of events that occurred as the crust cooled is difficult to construct. Below 100 °C
(212 °F) all the H2O would have condensed, and the acid gases would have
reacted with the original igneous crustal minerals to form sediments and an
initial ocean. There are at least two possible pathways by which these initial
steps could have been accomplished.
One pathway assumes that the 600 °C (1,100 °F) atmosphere contains,
together with other compounds, water (as vapour), carbon dioxide, and
hydrochloric acid in the ratio of 20:3:1 and would cool to the critical temperature
of water. The water vapour therefore would have condensed into an early hot
ocean. At this stage, the hydrochloric acid would be dissolved in the ocean
(about 1 mole per litre), but most of the carbon dioxide would still be in the
atmosphere with about 0.5 mole per litre in the ocean water. This early acid
ocean would react vigorously with crustal minerals, dissolving out silica and
cations and creating a residue that consisted principally of aluminous clay
minerals that would form the sediments of the early ocean basins. This pathway
of reaction assumes that reaction rates are slow relative to cooling. A second
pathway of reaction, which assumes that cooling is slow, is also possible. In this
case, at a temperature of about 400 °C (750 °F) most of the water vapour would
be removed from the atmosphere by hydration reactions with pyroxenes and
olivines. Under these conditions, water vapour would not condense until some
unknown temperature was reached, and Earth might have had at an early stage in
its history an atmosphere rich in carbon dioxide and no ocean: the surface would
have been much like that of present-day Venus.
The pathways described are two of several possibilities for the early surface
environment of Earth. In either case, after Earth’s surface had cooled to 100 °C
(212 °F), it would have taken only a short time geologically for the acid gases to
be used up in reactions involving igneous rock minerals. The presence of
bacteria and possibly algae in the fossil record of rocks older than 3 billion years
attests to the fact that Earth’s surface had cooled to temperatures lower than 100
°C (212 °F) by this time and that the neutralization of the original acid gases had
taken place. If most of the degassing of primary volatile substances from Earth’s
interior occurred early, the chloride released by reaction of hydrochloric acid
with rock minerals would be found in the oceans and seas or in evaporite
deposits, and the oceans would have a salinity and volume comparable to those
that they have today.
This conclusion is based on the assumption that there has been no drastic
change in the ratios of volatiles released through geologic time. The overall
generalized reaction indicative of the chemistry leading to formation of the early
oceans can be written in the form: primary igneous rock minerals + acid volatiles
+ H2O → sedimentary rocks + oceans + atmosphere. Notice from this equation
that if all the acid volatiles and H2O were released early in the history of Earth
and in the proportions found today, then the total original sedimentary rock mass
produced would be equal to that of the present time, and ocean salinity and
volume would be near what they are now. If, on the other hand, degassing were
linear with time, then the sedimentary rock mass would have accumulated at a
linear rate, as would oceanic volume. However, the salinity of the oceans would
remain nearly the same if the ratios of volatiles degassed did not change with
time. The most likely situation is that presented here—namely, that major
degassing occurred early in Earth history, after which minor amounts of volatiles
were released episodically or continuously for the remainder of geologic time.
The salt content of the oceans based on the constant proportions of volatiles
released would depend primarily on the ratio of sodium chloride (NaCl) locked
up in evaporites to that dissolved in the oceans. If all the sodium chloride in
evaporites were added to the oceans today, the salinity would be roughly
doubled. This value gives a sense of the maximum salinity the oceans could have
attained throughout geologic time.
One component missing from the early terrestrial surface was free oxygen
because it would not have been a constituent released from the cooling crust. As
noted earlier, early production of oxygen was by photodissociation of water in
the atmosphere as a result of absorption of ultraviolet light. The reaction is 2H2O
+ hν → O2 + 2H2, in which hν represents a photon of ultraviolet light. The
hydrogen produced would escape into space, and the O2 would react with the
early reduced gases by reactions such as 2H2S + 3O2 → 2SO2 + 2H2O. Oxygen
production by photodissociation gave the early reduced atmosphere a start
toward present-day conditions, but it was not until the appearance of
photosynthetic organisms approximately 3.3 billion years ago that it was
possible for the accumulation of oxygen in the atmosphere to proceed at a rate
sufficient to lead to today’s oxygenated environment. The photosynthetic
reaction leading to oxygen production may be written 6CO2 + 6H2O + hν →
C6H12O6 + 6O2, in which C6H12O6 represents sugar.

THE TRANSITION STAGE


The nature of the rock record from the time of the first sedimentary rocks (about
3.5 billion years ago) to approximately 2 to 1.5 billion years ago suggests that
the amount of oxygen in the atmosphere was significantly lower than today and
that there were continuous chemical trends in the sedimentary rocks formed and,
more subtly, in oceanic composition. The source rocks of sediments during this
time were likely to be more basaltic than would later ones; sedimentary detritus
was formed by the alteration of these rocks in an oxygen-deficient atmosphere
and accumulated primarily under anaerobic marine conditions. The chief
difference between reactions involving mineral-ocean equilibriums at this time
and at the present time was the role played by ferrous iron. The concentration of
dissolved iron in the present-day oceans is low because of the insolubility of
oxidized iron oxides. During the period 3.5 to 1.5 billion years ago, oxygen-
deficient environments were prevalent; these favoured the formation of minerals
containing ferrous iron (reduced state of iron) from the alteration of basaltic
rocks. Indeed, the iron carbonate siderite and the iron silicate greenalite, in close
association with chert and the iron sulfide pyrite, are characteristic minerals that
occur in middle Precambrian iron formations (those about 1.5 to 2.4 billion years
old). The chert originally was deposited as amorphous silica; equilibrium
between amorphous silica, siderite, and greenalite at 25 °C (77 °F) and 1
atmosphere of total pressure requires a carbon dioxide pressure of about 10-2.5
atmosphere, or 10 times the present-day value.
The oceans at this time can be thought of as the solution resulting from an
acid leach of basaltic rocks, and because the neutralization of the volatile acid
gases was not restricted primarily to land areas as it is presently, much of this
alteration may have occurred by submarine processes. The atmosphere at the
time was oxygen-deficient; anaerobic depositional environments with internal
carbon dioxide pressures of about 10-2.5 atmosphere were prevalent, and the
atmosphere itself may have had a carbon dioxide pressure near 10-2.5
atmosphere. If so, the pH of early ocean water was lower than that of modern
seawater, the calcium concentration was higher, and the early ocean water was
probably saturated with respect to amorphous silica (about 120 parts per million
[ppm]).
To simulate what might have occurred, it is helpful to imagine emptying the
Pacific basin, throwing in great masses of broken basaltic material, filling it with
hydrochloric acid so that the acid becomes neutralized, and then carbonating the
solution by bubbling carbon dioxide through it. Oxygen would not be permitted
into the system. The hydrochloric acid would leach the rocks, resulting in the
release and precipitation of silica and the production of a chloride ocean
containing sodium, potassium, calcium, magnesium, aluminum, iron, and
reduced sulfur species in the proportions present in the rocks. As complete
neutralization was approached, aluminum could begin to precipitate as
hydroxides and then combine with precipitated silica to form cation-deficient
aluminosilicates. The aluminosilicates, as the end of the neutralization process
was reached, would combine with more silica and with cations to form minerals
like chlorite, and ferrous iron would combine with silica and sulfur to make
greenalite and pyrite. In the final solution, chlorine would be balanced by
sodium and calcium in roughly equal proportions, with subordinate potassium
and magnesium; aluminum would be quantitatively removed, and silicon would
be at saturation with amorphous silica. If this solution were then carbonated,
calcium would be removed as calcium carbonate, and the chlorine balance would
be maintained by abstraction of more sodium from the primary rock. The
sediments produced in this system would contain chiefly silica, ferrous iron
silicates, chloritic minerals, calcium carbonate, calcium magnesium carbonates,
and minor pyrite.
If the hydrochloric acid added were in excess of the carbon dioxide, the
resultant ocean would have a high content of calcium chloride, but the pH would
still be near neutrality. If the carbon dioxide added were in excess of the
chlorine, calcium would be precipitated as the carbonate until it reached a level
approximately that of the present oceans—namely, a few hundred parts per
million.
If this newly created ocean were left undisturbed for a few hundred million
years, its waters would evaporate and be transported onto the continents (in the
form of precipitation); streams would transport their loads into it. The sediment
created in this ocean would be uplifted and incorporated into the continents.
Gradually, the influence of the continental debris would be felt, and the pH
might shift slightly. Iron would be oxidized out of the ferrous silicates to produce
iron oxides, but the water composition would not vary a great deal.
The primary minerals of igneous rocks are all mildly basic compounds.
When they react in excess with acids such as hydrochloric acid and carbon
dioxide, they produce neutral or mildly alkaline solutions plus a set of altered
aluminosilicate and carbonate reaction products. It is highly unlikely that ocean
water has changed through time from a solution approximately in equilibrium
with these reaction products, which are clay minerals and carbonates.

THE MODERN OCEANS


The oceans probably achieved their modern characteristics 2 to 1.5 billion years
ago. The chemical and mineralogical compositions and the relative proportions
of sedimentary rocks of this age differ little from their Paleozoic counterparts
(those dating from about 570 million to 245 million years ago). The fact that the
acid sulfur gases had been neutralized to sulfate by this time is borne out by
calcium sulfate deposits of late Precambrian age (roughly 570 million to 1.6
billion years old). Chemically precipitated ferric oxides in late Precambrian
sedimentary rocks indicate available free oxygen, whatever its percentage. The
chemistry and mineralogy of middle and late Precambrian shales is similar to
that of Paleozoic shales. Thus, it appears that continuous cycling of sediments
like those of the present time has occurred for 1.5 to 2 billion years and that
these sediments have controlled oceanic composition.
It was once thought that the saltiness of the modern oceans simply represents
the storage of salts derived from rock weathering and transported to the oceans
by fluvial processes. With increasing knowledge of the age of Earth, however, it
was realized that, at the present-day rate of delivery of salts to the oceans or even
at much reduced rates, the total salt content and the mass of individual salts in
the oceans could be attained in geologically short-time intervals compared to
Earth’s age. The total mass of salt in the oceans can be accounted for at present-
day rates of stream delivery in about 12 million years. The mass of dissolved
silica in ocean water can be doubled in only 20,000 years by addition of stream-
derived silica; to double sodium would take 70 million years. It then became
apparent that the oceans were not simply an accumulator of salts, but as water
evaporated from the oceans, along with some salt, the introduced salts must be
removed in the form of minerals. Thus, the concept of the oceans as a chemical
system changed from that of a simple accumulator to that of a steady-state
system in which rates of inflow of materials into the oceans equal rates of
outflow. The steady-state concept permits influx to vary with time, but it would
be matched by nearly simultaneous and equal variation of efflux. Calculations of
rates of addition of elements to the oceanic system and removal from it show
that for at least 100 million years the oceanic system has been in a steady state
with approximately fixed rates of major element inflow and outflow and, thus,
fixed chemical composition.
CHAPTER 2
OCEAN BASINS

Ocean basins constitute any of several vast submarine regions that collectively
cover nearly three-quarters of Earth’s surface. Together they contain the
overwhelming majority of all water on the planet and have an average depth of
almost 4 km (about 2.5 miles). A number of major features of the basins depart
from this average—for example, the mountainous ocean ridges, deep-sea
trenches, and jagged, linear fracture zones. Other significant features of the
ocean floor include aseismic ridges, abyssal hills, and seamounts and guyots.
The basins also contain a variable amount of sedimentary fill that is thinnest on
the ocean ridges and usually thickest near the continental margins.
While the ocean basins lie much lower than sea level, the continents stand
high—about 1 km (0.6 mile) above sea level. The physical explanation for this
condition is that the continental crust is light and thick while the oceanic crust is
dense and thin. Both the continental and oceanic crusts lie over a more uniform
layer called the mantle. As an analogy, one can think of a thick piece of
styrofoam and a thin piece of wood floating in a tub of water. The styrofoam
rises higher out of the water than the wood.

Major features of the ocean basins.

The ocean basins are transient features over geologic time, changing shape
and depth while the process of plate tectonics occurs. The surface layer of Earth,
the lithosphere, consists of a number of rigid plates that are in continual motion.
The boundaries between the lithospheric plates form the principal relief features
of the ocean basins: the crests of oceanic ridges are spreading centres where two
plates move apart from each other at a rate of several centimetres per year.
Molten rock material wells up from the underlying mantle into the gap between
the diverging plates and solidifies into oceanic crust, thereby creating new ocean
floor. At the deep-sea trenches, two plates converge, with one plate sliding down
under the other into the mantle where it is melted. Thus, for each segment of
new ocean floor created at the ridges, an equal amount of old oceanic crust is
destroyed at the trenches, or so-called subduction zones. It is for this reason that
the oldest segment of ocean floor, found in the far western Pacific, is apparently
only about 200 million years old, even though the age of Earth is estimated to be
at least 4.6 billion years.
Three-dimensional diagram showing crustal generation and
destruction according to the theory of plate tectonics; included are the
three kinds of plate boundaries—divergent, convergent (or collision),
and strike-slip (or transform). Encyclopædia Britannica, Inc.

The dominant factors that govern seafloor relief and topography are the
thermal properties of the oceanic plates, tensional forces in the plates, volcanic
activity, and sedimentation. In brief, the oceanic ridges rise about 2 km (1.2
miles) above the seafloor because the plates near these spreading centres are
warm and thermally expanded. In contrast, plates in the subduction zones are
generally cooler. Tensional forces resulting in plate divergence at the spreading
centres also create block-faulted mountains and abyssal hills, which trend
parallel to the oceanic ridges. Seamounts and guyots, as well as abyssal hills and
most aseismic ridges, are produced by volcanism. Continuing sedimentation
throughout the ocean basin serves to blanket and bury many of the faulted
mountains and abyssal hills with time. Erosion plays a relatively minor role in
shaping the face of the deep seafloor, in contrast to the continents. This is
because deep ocean currents are generally slow (they flow at less than 50 cm [20
inches] per second) and lack sufficient power.

SEA LEVEL

Sea level is the position of the air-sea interface on Earth’s surface to which
all terrestrial elevations and submarine depths are referred. The sea level
constantly changes at every locality with the changes in tides, atmospheric
pressure, and wind conditions. Longer-term changes in sea level are
influenced by Earth’s changing climates. Consequently, the level is better
defined as mean sea level, the height of the sea surface averaged over all
stages of the tide over a long period of time.
Global mean sea level rose at an average rate of about 1.2 mm (0.05
inch) per year over much of the 20th century, with shorter terms during
which the rise was significantly faster (5.5 mm [0.2 inches] per year during
the period from 1946 to 1956). This variable rise has been shown to have
occurred for a very long time. The sea level appears to have been very close
to its present position 35,000 years ago. It dropped 130 metres (426 feet) or
more during the interval from 30,000 to 15,000 years ago and has been
rising ever since. Fluctuations of equivalent magnitude probably have
accompanied the alternate growth and melting of continental glaciers
during the Pleistocene Epoch (from 2.6 million to 11,700 years ago)
because the ocean’s waters are the ultimate source of glacial ice. Slower
changes in the shapes and sizes of the ocean basins have less effect.

EXPLORATION OF THE OCEAN BASINS


Mapping the characteristics of the ocean basin has been difficult for several
reasons. First, the oceans are not easy to travel over; second, until recent times
navigation has been extremely crude, so that individual observations have been
only loosely correlated with one another; and, finally, the oceans are opaque to
light—that is, the deep seafloor cannot be seen from the ocean surface. Modern
technology has given rise to customized research vessels, satellite and electronic
navigation, and sophisticated acoustic instruments that mitigate some of these
problems.
The Challenger Expedition, mounted by the British in 1872–76, provided the
first systematic view of a few of the major features of the seafloor. Scientists
aboard the HMS Challenger determined ocean depths by means of wire-line
soundings and discovered the Mid-Atlantic Ridge. Dredges brought up samples
of rocks and sediments off the seafloor. The main advance in mapping, however,
did not occur until sonar was developed in the early 20th century. This system
for detecting the presence of objects underwater by acoustic echo provided
marine researchers with a highly useful tool, since sound can be detected over
several thousands of kilometers in the ocean (visible light, by comparison, can
penetrate only 100 metres [about 330 feet] or so of water).
Modern sonar systems include the Seabeam multibeam echo sounder and the
GLORIA scanning sonar. They operate on the principle that the depth (or
distance) of the seafloor can be determined by multiplying one-half the elapsed
time between a downgoing acoustic pulse and its echo by the speed of sound in
seawater (about 1,500 metres [4,900 feet] per second). Such multifrequency
sonar systems permit the use of different pulse frequencies to meet different
scientific objectives. Acoustic pulses of 12 kilohertz (kHz), for example, are
normally employed to measure ocean depth, while lower frequencies—3.5 kHz
to less than 100 hertz (Hz)—are used to map the thickness of sediments in the
ocean basins. Very high frequencies of 100 kHz or more are employed in side-
scanning sonar to measure the texture of the seafloor. The acoustic pulses are
normally generated by piezoelectric transducers. For determining subbottom
structure, low-frequency acoustic pulses are produced by explosives, compressed
air, or water-jet implosion. Near-bottom sonar systems, such as the Deep Tow of
the Scripps Institution of Oceanography (in La Jolla, Calif., U.S.), produce even
more detailed images of the seafloor and subbottom structure. The Deep Tow
package contains both echo sounders and side-scanning sonars, along with
associated geophysical instruments, and is towed behind a ship at slow speed 10
to 100 metres (33 to 330 feet) above the seafloor. It yields very precise
measurements of even finer-scale features than are resolvable with Seabeam and
other comparable systems.
Another notable instrument system is ANGUS, a deep-towed camera sled
that can take thousands of high-resolution photographs of the seafloor during a
single day. It has been successfully used in the detection of hydrothermal vents
at spreading centres. Overlapping photographic images make it possible to
construct photomosaic strips about 10 to 20 metres (about 33 to 66 feet) wide
that reveal details on the order of centimetres.
Three major navigation systems are in use in modern marine geology. These
include electromagnetic systems such as Loran and Earth-orbiting satellites.
Acoustic transponder arrays of two or more stations placed on the seafloor a few
kilometres apart are used to navigate deeply towed instruments, submersibles,
and occasionally surface research vessels when detailed mapping is conducted in
small areas. These systems measure the distance between the instrument package
and the transponder sites and, using simple geometry, compute fixes accurate to
a few metres. Although the individual transponders can be used to determine
positions relative to the array with great accuracy, the preciseness of the position
of the array itself depends on which system is employed to locate it.
Earth-orbiting satellites such as Seasat and Geosat have uncovered some
significant topographic features of the ocean basins. Seasat, launched in 1978,
carried a radar altimeter into orbit. This device was used to measure the distance
between the satellite path and the surfaces of the ocean and continents to 0.1
metre (0.3 foot). The measurements revealed that the shape of the ocean surface
is warped by seafloor features: massive seamounts cause the surface to bulge
over them because of gravitational attraction. Similarly, the ocean surface
downwarps occur over deep-sea trenches. Using these satellite measurements of
the ocean surface, William F. Haxby computed the gravity field there. The
resulting gravity map provides comprehensive coverage of the ocean surface on
a 5’-by-5’ grid that depicts 5 nautical miles on each side at the Equator.
Coverage as complete as this is not available from echo soundings made from
ships. Because the gravity field at the ocean surface is a highly sensitive
indicator of marine topography, this map reveals various previously uncharted
features, including seamounts, ridges, and fracture zones, while improving the
detail on other known features. In addition, the gravity map shows a linear
pattern of gravity anomalies that cut obliquely across the grain of the
topography. These anomalies are most pronounced in the Pacific basin; they are
apparently about 100 km (about 60 miles) across and some 1,000 km (about 600
miles) long. They have an amplitude of approximately 10 milligals (0.001
percent of Earth’s gravity attraction) and are aligned west-northwest—very close
to the direction in which the Pacific Plate moves over the mantle below.

SEASAT

Seasat is an experimental U.S. ocean surveillance satellite launched June


27, 1978. During its 99 days of operation, Seasat orbited Earth 14 times
daily and monitored nearly 96 percent of its oceanic surface every 36 hours.
Instruments on the unmanned spacecraft, engineered to penetrate cloud
cover, provided data on a wide array of oceanographic conditions and
features, including wave height, water temperature, currents, winds,
icebergs, and coastal characteristics. Although Seasat ceased data
transmission on Oct. 9, 1978, as a result of a power failure, it achieved its
primary purpose: to demonstrate that much useful information about
oceanographic phenomena could be obtained by means of satellite
surveillance. Data transmitted by Seasat was made available to scientists
representing 23 government and academic organizations. The information
was also used to aid the crews of transoceanic vessels and aircraft.

DEEP-SEA SEDIMENTS
The ocean basin floor is everywhere covered by sediments of different types and
origins. The only exception are the crests of the spreading centres where new
ocean floor has not existed long enough to accumulate a sediment cover.
Sediment thickness in the oceans averages about 450 metres (1,500 feet). The
sediment cover in the Pacific basin ranges from 300 to 600 metres (about 1,000
to 2,000 feet) thick, and that in the Atlantic is about 1,000 metres (3,300 feet).
Generally, the thickness of sediment on the oceanic crust increases with the age
of the crust. Oceanic crust adjacent to the continents can be deeply buried by
several kilometres of sediment. Deep-sea sediments can reveal much about the
last 200 million years of Earth history, including seafloor spreading, the history
of ocean life, the behaviour of Earth’s magnetic field, and the changes in the
ocean currents and climate.
The study of ocean sediments has been accomplished by several means.
Bottom samplers, such as dredges and cores up to 30 metres (about 100 feet)
long, have been lowered from ships by wire to retrieve samples of the upper
sediment layers. Deep-sea drilling has retrieved core samples from the entire
sediment layer in several hundred locations in the ocean basins. The seismic
reflection method has been used to map the thickness of sediments in many parts
of the oceans. Besides thickness, seismic reflection data can often reveal
sediment type and the processes of sedimentation.

SEDIMENT TYPES
Deep-sea sediments can be classified as terrigenous, originating from land; as
biogenic, consisting largely of the skeletal debris of microorganisms; or as
authigenic, formed in place on the seafloor. Pelagic sediments, either terrigenous
or biogenic, are those that are deposited very slowly in the open ocean either by
settling through the volume of oceanic water or by precipitation. The sinking
rates of pelagic sediment grains are extremely slow because they ordinarily are
no larger than several micrometres. However, fine particles are normally bundled
into fecal pellets by zooplankton, which allows sinking at a rate of 40 to 400
metres (130 to 1,300 feet) per day.

SEDIMENTATION PATTERNS
The patterns of sedimentation in the ocean basins have not been static over
geologic time. The existing basins, no more than 200 million years old, contain a
highly variable sedimentary record. The major factor behind the variations is
plate movements and related changes in climate and ocean water circulation.
Since about 200 million years ago, a single vast ocean basin has given way to
five or six smaller ones. The Pacific Ocean basin has shrunk, while the North
and South Atlantic basins have been created. The climate has changed from
warm and mild to cool, stormy, and glacial. Plate movements have altered the
course of surface and deep ocean currents and changed the patterns of upwelling,
productivity, and biogenic sedimentation. Seaways have opened and closed. The
Strait of Gibraltar, for example, was closed off about 6 million years ago,
allowing the entire Mediterranean Sea to evaporate and leave thick salt deposits
on its floor. Changes in seafloor spreading rates and glaciations have caused sea
levels to rise and fall, greatly altering the deep-sea sedimentation pattern of both
terrigenous and biogenic sediments. The calcite compensation depth (CCD), or
the depth at which the rate of carbonate accumulation equals the rate of
carbonate dissolution, has fluctuated more than 2,000 metres (about 6,600 feet)
in response to changes in carbonate supply and the corrosive nature of ocean
bottom waters. Bottom currents have changed, becoming erosive or
nondepositional in some regions to produce geological unconformities (that is,
gaps in the geological record) and redistribute enormous volumes of sediment to
other locations. The Pacific Plate has been steadily moving northward, so that
biogenic sediments of the equatorial regions are found in core samples taken in
the barren North Pacific.

EVOLUTION OF THE OCEAN BASINS THROUGH PLATE


MOVEMENTS
Through most of geologic time, probably extending back 2 billion years, the
ocean basins have both grown and been consumed as plate tectonics continued
on Earth. The latest phase of ocean basin growth began just less than 200 million
years ago with the breakup of the supercontinent Pangea, the enormous landmass
composed of nearly all the present-day continents. Since that time the major
developments have included a shrinking of the Pacific basin at the expense of
the growing Atlantic and Arctic basins, the opening of the Tethys seaway
circling the globe in tropical latitudes and its subsequent closing, and the
opening of the Southern Ocean as the southern continents moved north away
from Antarctica.
The oldest known oceanic crust (estimated to be about 200 million years old)
is located in the far western equatorial Pacific, east of the Mariana Island arc.
The Pacific ocean floor at this site was generated during seafloor spreading from
a pattern of ridges and plates that had existed for some unknown period of time.
At least five different seafloor spreading centres were involved. In the Indian
Ocean the oldest segment of seafloor was formed about 165 million to 145
million years ago by the rifting away of Africa and South America from
Gondwana, a supercontinent consisting largely of the present-day continents of
the Southern Hemisphere. At this time Africa was joined to South America,
Eurasia, and North America. Today this old seafloor is found along the east coast
of Africa from the Somali Basin to the east coast of South Africa and adjacent to
Queen Maud Land and Enderby Land in East Antarctica.
Close to 180 million years ago (but before 165 million years ago), North
America and Eurasia, which together made up most of the large northern
continent of Laurasia, began drifting away from Africa and South America,
creating the first seafloor in the central region of the North Atlantic and opening
the Gulf of Mexico. The Tethys seaway also opened during this rifting phase as
Europe pulled away from Africa. Shortly after this time continental fragments,
including possibly Tibet, Myanmar (Burma), and Malaya, rifted away from the
northwest coast of Australia and moved northward, thereby creating the oldest
seafloor in the Timor Sea. During this period spreading continued in the Pacific
basin with the growth of the Pacific Plate and the consumption by subduction of
its bordering plates, including the Izanagi, Farallon, and Phoenix. The Pacific
Plate moved northward during this phase and continues to do so today.
India and Madagascar, as a unit, rifted away from Australia and Antarctica
prior to 130 million years ago and began drifting northward, creating seafloor
adjacent to Western Australia and East Antarctica. Possibly simultaneously or
shortly after this rifting began, South America started to separate from Africa,
initiating the formation of seafloor in the South Atlantic Ocean.
Between 90 million and 80 million years ago, Madagascar and India
separated, and the spreading ridges in the Indian Ocean were reorganized. India
began drifting northward directly toward Asia. During this same period, Europe,
joined to Greenland, began drifting away from North America, which resulted in
the emergence of the seafloor in the Labrador Sea and the northernmost Atlantic
Ocean. This spreading phase affected the passages in the Tethys seaway between
Europe (Iberia) and northwest Africa, intermittently opening and closing it. In
the southwest Pacific, New Zealand, along with the Lord Howe Rise and the
Norfolk Ridge, rifted away from Australia and Antarctica between 80 million
and 60 million years ago, opening the Tasman Sea.
About 60 million years ago a new rift and oceanic ridge formed between
Greenland and Europe, separating them and initiating the formation of oceanic
crust in the Norwegian Sea and the Eurasian basin in the eastern Arctic Ocean.
The Amerasian basin in the western Arctic Ocean had formed during an earlier
spreading phase from about 130 million to 110 million years ago. Between 60
million and 50 million years ago, significant events occurred in the Indian Ocean
and southwest Pacific. Australia began drifting northward, away from East
Antarctica, creating seafloor there. The northward movement of Australia
resulted in the emergence of several subduction zones and island arcs in the
southwest and equatorial Pacific. The Indian subcontinent first touched against
the Asian continent about 53 million years ago, developing structures that
preceded the main Himalayan orogeny (mountain-building event), which began
in earnest some 40 million years ago.

ISLAND ARCS

Island arcs are long, curved chains of oceanic islands associated with
intense volcanic and seismic activity and orogenic (mountain-building)
processes. Prime examples of this form of geologic feature include the
Aleutian-Alaska Arc and the Kuril-Kamchatka Arc.
Most island arcs consist of two parallel, arcuate rows of islands. The
inner row of such a double arc is composed of a string of explosive
volcanoes, while the outer row is made up of nonvolcanic islands. In the
case of single arcs, many of the constituent islands are volcanically active.
An island arc typically has a landmass or a partially enclosed, unusually
shallow sea on its concave side. Along the convex side there almost
invariably exists a long, narrow deep-sea trench. The greatest ocean depths
are found in these depressions of the seafloor, as in the case of the Mariana
and Tonga trenches.
Destructive earthquakes occur frequently at the site of island arcs.
Unlike the shallow earthquakes that are recorded in other areas of the
world, these are deep-focus seismic events emanating from as much as 370
miles (600 km) below the base of an arc. The quakes tend to have foci of
progressively greater depth toward the arc’s concave side.
The majority of island arcs occur along the western margin of the
Pacific Basin. The few exceptions are the East Indian and the West Indian
arcs and the Scotia Arc in the South Atlantic. According to prevailing
theory, island arcs are formed where two lithospheric plates (enormous rigid
slabs that constitute segments of Earth’s surface) converge. Upon colliding,
one of the plates—that bearing heavy, oceanic crust—buckles downward
and is forced into the partially molten lower mantle beneath the second
plate with lighter, continental crust. An island arc is built up from the
surface of the overriding plate by the extrusion of basalts and andesites. The
basalts are thought to be derived from the semimolten mantle, whereas the
andesites are probably generated by the partial melting of the descending
plate and the sediments that have accumulated on its surface.

Less than 30 million years ago, seafloor spreading ceased in the Labrador
Sea. Along the west coast of North America, the Pacific Plate and the North
American Plate converged along what is now California shortly after 30 million
years ago. This resulted in the cessation of a long history of subduction in the
area and the gradual conversion of this continental margin to a transform fault
zone. Continued closure between Africa and Europe, which began about 100
million years ago, caused the isolation of the Mediterranean Sea, so that by 6
million years ago this water body had completely evaporated.
The present-day Mediterranean seafloor was formed during a complex
sequence of rifting between small plates in this region, beginning with the
separation of North America and Europe from Africa about 200 million years
ago. In the eastern Mediterranean the seafloor is no older than about 100 million
years. West of Italy it was created during subsequent spreading between 30
million and 20 million years ago.
The Caribbean Sea and the Gulf of Mexico formed as a result of the relative
movement between North America and South America. The seafloor of the Gulf
of Mexico began forming some 160 million to 150 million years ago. A proto-or
ancient Caribbean seafloor also was formed during this period but was later
subducted. The present Caribbean seafloor consists of a captured piece of the
Farallon Plate (from the Pacific basin) and is estimated to be for the most part of
Cretaceous age (i.e., about 120 million to 85 million years old).
The seafloor in the western portion of the Philippine Sea developed between
60 million and 35 million years ago. In the east it was formed by backarc
spreading from 30 million years ago. The origin of the older crust is not
completely clear. It either was created by spreading in the Pacific basin and
subsequent capture by the formation of the Bonin and Mariana arcs, or it resulted
from backarc spreading behind trenches to the south.

OCEAN BASIN FEATURES


Ocean basins contain a number of physical features. Some of these, such as
abyssal hills and deep-sea trenches, are analogous to topographic features on
land. Although most of the ocean floor is made up of oceanic crust, some
features of ocean basins (such as continental shelves and continental slopes) are
made of continental crust.

ABYSSAL HILLS
Abyssal hills are small, topographically well-defined submarine hills that may
rise from several metres to several hundred metres above the abyssal seafloor.
(Abyssal hills appear some 3,000 to 6,000 metres [10,000 to 20,000 feet] below
the surface of the ocean.) Typical abyssal hills have diameters of several to
several hundred metres. They elongate parallel to spreading centres or to marine
magnetic anomalies and cover the entire flanks and crests of oceanic ridges.
Abyssal hill provinces, areas of abyssal seafloor occupied exclusively by such
hills, characteristically occur seaward of the smooth abyssal plains at the bases
of continental rises. Isolated hills and groups of hills also protrude from abyssal
plain surfaces, and the base of an abyssal plain accumulation of marine
sediment, as revealed by subbottom seismic profiling, generally matches the
undulating topography and relief of abyssal hill provinces.
Apparently, the hills are constructed by two processes: volcanism and block
faulting. The relative contribution of each may depend on the spreading rate. At
slower rates, faulting of the oceanic crust is a dominant factor in forming the
relief, and the relief of the hills is greater as the rate is slower. At the crest of a
spreading centre, volcanism in the neovolcanic zone initiates the construction of
volcanic hills. The zone of active faulting is where they form or are modified by
block faulting. The existence of discrete and separate volcanic hills indicates that
volcanism at a spreading centre is episodic.
Abyssal hills, although generally covered with marine sediments, probably
are identical in composition and origin to the extrusive basaltic prominences on
the upper flanks of mid-ocean ridges and rises. Thus, it is believed that abyssal
hills underlie most of the ocean floor, locally buried by accumulations of abyssal
sediment. In the Atlantic Ocean, long abyssal hill provinces parallel both flanks
of the Mid-Atlantic Ridge along most of its length. The Pacific Ocean has a
smaller supply of continental sediment than the Atlantic Ocean, and numerous
trenches and local rises separate the main ocean floor from the continents,
preventing the seaward transport of sediment. Consequently, between 80 and 85
percent of the Pacific abyssal floor is occupied by abyssal hills.
CONTINENTAL MARGINS
Continental margins make up the submarine edge of the continental crust. They
are distinguished by relatively light and isostatically high-floating material in
comparison with the adjacent oceanic crust. It is the name for the collective area
that encompasses the continental shelf, continental slope, and continental rise.
The characteristics of the various continental margins are shaped by a
number of factors. Chief among these are tectonics, fluctuations of sea level, the
size of the rivers that empty onto a margin as determined by the amount of
sediment they carry, and the energy conditions or strength of the ocean waves
and currents along the margin.

The broad, gentle pitch of the continental shelf gives way to the
relatively steep continental slope. The more gradual transition to the
abyssal plain is a sediment-filled region called the continental rise.
The continental shelf, slope, and rise are collectively called the
continental margin. Depth is exaggerated here for effect.
Encyclopædia Britannica, Inc.

MARGIN TYPES
Continental margins on the leading edges of tectonic plates, like those around the
rim of the Pacific Ocean, are usually narrow and have steep continental slopes
and either poorly developed continental rises or none at all. The continental
slope is often steep and falls away directly into a deep-sea trench. In many cases,
the leading-edge margins are backed by mountain ranges. Continental margins
on the trailing side of tectonic plates, like those around the Atlantic Ocean, are
broad, with gentle continental slopes and well-developed continental rises. The
adjacent land area is commonly a broad coastal plain that, depending on the state
of sea level, may become submerged from time to time and hence part of the
continental margin.
Since continental margins are the shallowest parts of the world’s oceans, they
are most affected by changes in sea level. Worldwide changes in sea level, called
eustatic sea-level changes, have occurred throughout geologic history. The most
common causes of such sea-level changes are global climatic fluctuations that
lead to major glacial advances and retreats—that is, ice ages and interglacial
periods. Other causes that are not as well understood may include major
mountain-building events and isostatic changes in crustal plates. When
continental glaciers advance, as they did several times during the Pleistocene
Epoch (which extended from about 2.6 million to 11,700 years ago), water that
would normally be in the oceans is locked up as ice on land, resulting in a drop
in sea level. As the glaciers retreat, more water is fed to the ocean basins and the
sea level rises. Fluctuations from highstand to lowstand have totaled 250 metres
(about 800 feet) or more during Cenozoic time (roughly the last 65.5 million
years), with concomitant fluctuations in exposure and flooding of the continental
margins. (During a highstand the sea level is above the edge of the continental
shelf, while during a lowstand it is below the shelf edge.)
Rivers bring a variety of sediments to the coast. These are classified by their
mineralogy and by particle size and include sand, silt, and clay. To
sedimentologists, sand is a grain of any composition from 63 to 2,000
micrometres (0.002 to 0.08 inch) in its largest diameter. Silt is 4 to 62
micrometres (0.0002 to 0.002 inch), and clay is any particle less than 4
micrometres. Most of the detrital minerals brought to the continental margins by
rivers in sand and silt sizes are quartz, feldspars, and mica; those of clay size are
a suite of clay minerals that most commonly include smectite, kaolinite, and
illite. (Clay can, in other words, refer either to particle size or to a group of
minerals.) These, then, are the mineral constituents that together with calcium
carbonates produced in the oceans by biogenic activity as shells and the hard
parts of plants and animals, go to make up the sedimentary packages that are
deposited on and constitute a fundamental part of continental margins.
A constant battle is being waged between the rivers that bring sediments
eroded from the land to the sea and the waves and currents of the receiving body
of water. This dynamic struggle goes on year after year, century after century,
sometimes for millions of years. Take, for example, the north coast of the Gulf of
Mexico, into which the Mississippi River flows. The continental margin at this
site is subject to relatively low wave and current energy, and so the river filled
up most of the adjacent continental shelf with a delta and typically dumps over
200 million tons of sediment each year directly at the top of the continental
slope. By contrast, the Columbia River in the Pacific Northwest of the United
States carries 131 million tons to the coast, where the sediments are attacked by
the large waves and currents normal for that margin. As a result, sediments are
widely dispersed, and the shelf is not filled with a large subaerial delta.
The effects of this battle are easily seen where human activities have
interfered with the transport of sediments to the sea by major rivers. For
example, the Nile River delta is retreating rapidly, widening the submerged
portion of the continental margin, because the Aswan High Dam has trapped
much of the sediment normally fed to the delta front. The lower Mississippi
River has been artificially maintained in a channel by high man-made levees.
These have stopped the floods that fed much of the western delta margin.
Because of this, coupled with a slow rise in sea level and the effects of canals
dug in the delta wetlands, the coast has begun to retreat significantly.
When rivers carrying sediment from the interiors of continents reach the sea,
several things happen. Velocity in the river jet decreases rapidly, and the sand
particles drop out to be picked up by the waves and currents along the coast,
where they feed beaches or barrier island systems. If the river has a large enough
discharge, the finer-than-sand-sized materials may be carried for kilometres onto
the margin in a fresh-or brackish-water plume. The surf system then acts as a
wave filter, trapping the sand in the coastal zone but allowing the finer materials
to be carried out onto the margin. When estuaries are the receiving bodies of
water on the coastal boundaries of continental margins, as in the case of the east
coast of North America, virtually all the sediments brought down by the rivers
are trapped within the confines of the estuaries.
In addition to the two primary types of continental margins, there also are
special types that do not readily fit either category. One of the most intensely
studied margins of the world is the Borderland, the continental margin of
southern California and northern Baja California. It consists of a series of
offshore basins and ridges, some of which are exposed as islands. This system of
basins and ridges formed as the result of faulting associated with the movement
of the Pacific Plate past the North American Plate. It remains tectonically active
today and is related to the San Andreas Fault system of California. A second
special type is the marginal plateau. The Blake Plateau off the east coast of
Florida is a good example. Such a plateau constitutes a portion of a continental
margin that has many of the features of a normal system but is found at much
greater depth—1,000 metres (about 3,300 feet) in the case of the Blake Plateau.
Continental margins can be either constructional or erosional over varying
periods of geologic time, depending on the combination of factors discussed
above. When deposition exceeds erosion, the margin grows seaward, a process
of progradation that builds out as well as up. When the erosive forces are
predominant, the margin remains static or actually retreats over time. Some
geologists think that the continental margin of the eastern United States has
retreated as much as 5–30 km (3–19 miles) since the end of the Cretaceous
Period some 65.5 million years ago.

ECONOMIC IMPORTANCE OF CONTINENTAL MARGINS


Continental margins are very significant economically. Most of the major
fisheries of the world are located on them. Of these, sport fisheries and related
tourist industries are becoming increasingly important to the economies of
developed nations. Paradoxically, continental margins also are one of the world’s
biggest dump sites. All kinds of wastes are disposed of along the margins, and
the effects of pollution have become a major global concern.
Continental margins are the only parts of the world’s oceans to be effectively
exploited for mineral resources. Millions of tons of sand are mined by dredges
each year off the U.S. coasts alone for beach renourishment projects. From time
to time placer deposits also have been worked. Examples include tin off
Indonesia, gold off Alaska, and diamonds off Namibia. By far and away the
largest mineral resources to be exploited from continental margins are oil and
natural gas. Exploration of the continental margins by major oil companies has
intensified and is expected to continue for the foreseeable future because the
margins are the most likely sites of giant undiscovered petroleum deposits.
Continental margins are made of thick accumulations of sedimentary rock, the
type of rock in which oil and gas generally occur. In fact, most of the
sedimentary rocks exposed on the continents were originally deposited on
continental margins; thus, even the hydrocarbon deposits found on land were
formed for the most part on ancient continental margins.

CONTINENTAL RISES
Continental rises are major depositional regimes in oceans made up of thick
sequences of continental material that accumulate between the continental slope
and the abyssal plain. Continental rises form as a result of three sedimentary
processes: mass wasting, the deposition from contour currents, and the vertical
settling of clastic and biogenic particles.
The first such process is a downslope movement of sediments by mass
wasting, a set of gravity-deposition events, including submarine landslides,
slumps, debris flows, and high-velocity sediment-laden density flows known as
turbidity currents. Several phenomena may initiate gravity events. In tectonically
active areas, earthquakes are important triggering mechanisms. Even in the
Atlantic they play a significant role. One of the few documented major gravity
events took place on the Grand Banks of Newfoundland in 1929, when an
earthquake triggered a gravity flow that possibly attained velocities of more than
90 km (56 miles) per hour and was traced for hundreds of kilometres as it
successively broke transatlantic cables. Other triggering events may be
oversteepening of deposits on the sharply inclined portions of the continental
slope, breaking internal waves that have been shown to affect the upper slope,
and storm waves and storm-induced currents.
A second process that may be equally important, although its overall
significance is subject to considerable scientific debate, is deposition from
bottom currents that flow parallel to the slope of the continental rise—namely,
contour currents. Resulting sediment accumulations are called contourites. The
major points of contention concerning the efficacy of contour currents are (1)
whether or not they are strong enough—they flow at a speed of about 20 cm (8
inches) per second—to build the huge thicknesses of sediment that make up the
rises and (2) how the sediments get into the contour currents in the first place. It
is probable that most of the mass of rise material is originally brought downslope
by gravity events and then redistributed by contour currents.
Vertical settling through the water column of both clastic and biogenic
particles is the third contributor of slope and rise sediments. These pelagic
sediments are composed of clay minerals and fine-grained particles (chiefly
quartz, mica, and carbonate) swept off the continental shelf, wind-blown dust,
organic detritus, and the tests of plankton. Chief among the last group are the
tests of foraminiferans, pteropods, and coccolithophores that are composed of
calcium carbonate and those of diatoms and radiolarians that are made of silicon
dioxide.

CONTINENTAL SHELVES
Continental shelves are broad, relatively shallow submarine terraces of
continental crust that form the edge of a continental landmass. The geology of
continental shelves is often similar to that of the adjacent exposed portion of the
continent, and most shelves have a gently rolling topography called ridge and
swale. Continental shelves make up about 8 percent of the entire area covered by
oceans.

STRUCTURE
A continental shelf typically extends from the coast to depths of 100–200 metres
(about 330–660 feet). It is gently inclined seaward at an average slope of about
0.1°. In nearly all instances, it ends at its seaward edge with an abrupt drop
called the shelf break. Below this lies the continental slope, a much steeper zone
that usually merges with a section of the ocean floor called the continental rise at
a depth of roughly 4,000 to 5,000 metres (13,000 to 16,500 feet). A few
continental margins—such as those off the Mediterranean coast of France and at
Porcupine Bank, off the western coast of Ireland—do not have a sharply defined
break in slope but rather maintain a generally convex shape to the seafloor.
The average width of continental shelves is about 65 km (40 miles). Almost
everywhere the shelves represent simply a continuation of the continental
landmass beneath the ocean margins. Accordingly, they are narrow, rough, and
steep off mountainous coasts but broad and comparatively level offshore from
plains. The shelf along the mountainous western coast of the United States, for
example, is narrow, measuring only about 32 km (20 miles) wide, whereas that
fringing the eastern coast extends more than 120 km (75 miles) in width.
Exceptionally broad shelves occur off northern Australia and Argentina.
Continental shelves are usually covered with a layer of sand, silts, and silty
muds. Their surfaces exhibit some relief, featuring small hills and ridges that
alternate with shallow depressions and valleylike troughs. In a few cases, steep-
walled V-shaped submarine canyons cut deeply into both the shelf and the slope
below.

ORIGIN
American oceanographer Donald J.P. Swift has called continental shelves
“palimpsests,” parchment writing tablets upon which stories are written after
each previous writing has been erased. Each new stand of sea level “writes” a
new story of sedimentation on the shelf after the previous episode has been
erased by the rise or fall that preceded it, but with some traces of the previous
environment of deposition or last erosional event remaining. The “eraser” is the
surf, a high-energy force that erodes and reworks everything as it passes over,
winnowing out the finer-than-sand-sized sediment and leaving the coarser
material behind. An interpreted seismic line shows the complicated array of
channels (eroded and then filled), old deltaic deposits, ancient erosional surfaces,
and winnowed sand bodies that make up the continental shelf southwest of Cape
San Blas on the panhandle of Florida.
How the above processes affect any particular margin depends on its tectonic
setting and the size of the rivers that drain into it. On continental shelves backed
by high mountain ranges, such as the Pacific coast of North and South America,
the difference between high and low sea-level stands may be difficult to detect,
being one of degree perhaps noticeable only by marginally increased
sedimentation rates during lowstands, or intervals of decreased sea level. In
many ways, continental shelves on tectonically active margins at present sea
levels approximate lowstands on trailing-edge, or passive, margins.
When sea level is lowered on a trailing-edge shelf that has no adjacent high
mountains, such as the Atlantic coast of North America, rivers are rejuvenated.
In other words, their base level is lowered and they begin to erode their beds,
carrying sediment from the continent across the former continental shelf that is
now exposed and depositing it at the new coast. When sea level falls below the
shelf break, the coast lies on the continental slope. As sea level rises again on
tectonically stable or sinking shelves, small and medium-sized river mouths
drown and estuaries form, trapping the sediment within them and starving the
shelves. In these cases, sediment for the shelf is primarily produced by erosion
of the coastline as the surf zone advances landward with rising sea level. Fine-
grained material is winnowed out, to be either deposited back in the estuaries or
carried in steps by advective processes across the shelf to the deeper water
beyond. As a result, continental shelf surfaces on trailing-edge margins into
which no large rivers flow are veneered with a sand sheet lying over a complex
of older deposits, some of which peek through the surface as outcrops—vestiges
of an earlier story written on the palimpsest. Large rivers that drain a large, high
continent, such as the Mississippi, are able to keep pace with rising sea level and
deliver enough sediment to keep an estuary from forming, and, at a high
stillstand like that of the present, even fill their entire shelf area.
For many years after World War II, the period when many of the world’s
continental shelves were first described in detail, it was thought that the sand
deposits on continental shelves were “relict,” deposits left stranded by a higher
sea level from the higher-energy regime of the surf zone that passed over them
perhaps as much as a few thousand years before. Geophysical investigations of
the shelf area since the mid-1970s have revealed the presence of many types of
sand waves and ripple marks in seafloor sediments that show submerged
continental shelf sediments to be constantly undergoing reworking and erosion.
As scientific understanding of the physical processes that affect continental
shelves increased, it was found that currents set up by large winter storms,
monsoons, and hurricanes and typhoons are reworking the bottom by winnowing
out the fine-grained materials and carrying them either back into the estuaries or
beyond the shelf break, where they are lost from the system.
In short, the kind of sediment that covers the surface of a continental shelf is
determined by the interplay among the tectonic setting, the size of the rivers that
empty into it (size based on how much sediment they carry), and the wave
energy that affects it, just as is the case with continental margins in general.
Shelves such as that of western Florida that have been cut off from clastic input
(that is, sediments composed chiefly of quartz and clay minerals derived from
erosion of the continent) may be covered with carbonate sediments. In some
cases, as in the islands of the Bahamas, the carbonate shelf, called a bank, is cut
off from a continental source by deep water. Continental shelves with rivers that
carry sediments from the continents to the shelf and beyond only at lowstands of
sea level and those that drain mountainous areas on high-energy coasts are
dominated by quartz sands. In addition, shelves with rivers that drain large
continental areas and carry enough sediment to keep abreast of sea-level rise or
dominate ambient wave-energy conditions will accumulate muddy sediment
deposits out across their surfaces.
Since the 1970s an increasing number of investigators have sought to explain
the origin of continental shelves and their related structures in terms of plate
tectonics. According to this theory, the shelves of the Pacific Ocean, for
example, formed as the leading edges of continental margins on lithospheric
plates that terminate either at fracture zones (sites where two such plates slide
past each other) or at subduction zones (sites where one of the colliding plates
plunges into the underlying partially molten asthenosphere). Shelves of such
origin tend to be steep, deformed, and covered by a thin layer of erosional
debris. The Atlantic continental shelves, on the other hand, show little or no
tectonic deformation and bear a thick veneer of sedimentary material. They are
thought to be remnants of the trailing edges of the enormous plates that split
apart and receded many millions of years ago to form the Atlantic basin. As the
edges of the plates gradually contracted and subsided, large amounts of sand,
silts, and mud from the continents settled and accumulated along their seaward
side.

CONTINENTAL SLOPES
Continental slopes form the seaward borders of continental shelves. The world’s
combined continental slope has a total length of approximately 300,000 km
(200,000 miles) and descends at an average angle in excess of 4 °From the shelf
break at the edge of the continental shelf to the beginning of the ocean basins at
depths of 100 to 3,200 metres (330 to 10,500 feet).
The gradient of the slope is lowest off stable coasts without major rivers and
highest off coasts with young mountain ranges and narrow continental shelves.
Most Pacific slopes are steeper than Atlantic slopes. Gradients are flattest in the
Indian Ocean. About one-half of all continental slopes descend into deep-sea
trenches or shallower depressions, and most of the remainder terminate in fans
of marine sediment or in continental rises. The transition from continental crust
to oceanic crust usually occurs below the continental slope.
About 8.5 percent of the ocean floor is covered by the continental slope-rise
system. This system is an expression of the edge of the continental crustal block.
Beyond the shelf-slope break, the continental crust thins quickly, and the rise lies
partly on the continental crust and partly on the oceanic crust of the deep sea.
Although the continental slope averages about 4°, it can approach vertical on
carbonate margins, on faulted margins, or on leading-edge, tectonically active
margins. Steep slopes usually have either a very poorly developed continental
rise or none at all and are called escarpments.
Continental slopes are indented by numerous submarine canyons and
mounds. The Blake Plateau off the southeastern United States and the
continental borderland off southern California are examples of continental slopes
separated from continental shelves by plateaus of intermediate depth. Slopes off
mountainous coastlines and narrow shelves often have outcrops of rock.
The predominant sediments of continental slopes are muds; there are smaller
amounts of sediments of sand or gravel. Over geologic time, the continental
slopes are temporary depositional sites for sediments. During lowstands of sea
level, rivers may dump their sedimentary burden directly on them. Sediments
build up until the mass becomes unstable and sloughs off to the lower slope and
the continental rise. During highstands of sea level, these processes slow down
as the coastline retreats landward across the continental shelf, and more of the
sediments delivered to the coast are trapped in estuaries and lagoons. Still the
process continues, albeit slowly, as sediments are brought across the shelf break
by winnowing of the shelf surface and by advection. Slopes are sometimes
scoured by such major ocean currents as the Florida Current that work to erode
their surfaces. Off active major deposition centres, such as the Mississippi delta,
slope sequences may accumulate through progradation, while the active slope
front is continuously shedding sediments downslope by gravity processes.

DEEP-SEA TRENCHES
These structures, which are also called oceanic trenches, are long, narrow, steep-
sided depressions in the ocean bottom. Deep-sea trenches are locations in which
the maximum oceanic depths, approximately 7,300 to more than 11,000 metres
(24,000 to 36,000 feet), occur. They typically form in locations where one
tectonic plate subducts under another. The deepest known depression of this kind
is the Mariana Trench, which lies east of the Mariana Islands in the western
North Pacific Ocean; it reaches 11,034 metres (36,200 feet) at its deepest point.

TYPES
Deep-sea trenches generally lie seaward of and parallel to adjacent island arcs or
mountain ranges of the continental margins. They are closely associated with
and found in subduction zones—that is, locations where a lithospheric plate
bearing oceanic crust slides down into the upper mantle under the force of
gravity. The result is a topographic depression where the oceanic plate comes in
contact with the overriding plate, which may be either oceanic or continental. If
the overriding plate is oceanic, an island arc develops. The trench forms an arc in
plan view, and islands with explosive volcanoes develop on the overriding plate.
If the overriding plate is continental, a marginal trench forms where the
topographic depression appears to follow the outline of the continental margin.
Explosive volcanoes are found there too.
Features of a typical island arc. Copyright Encyclopædia Britannica,
Inc.; rendering for this edition by Rosen Educational Services

Both types of subduction zones are associated with large earthquakes that
originate at a depth of as much as 700 km (435 miles). The deep earthquakes
below subduction zones occur in a plane that dips 30° or more under the
overriding plate. Typical trench depths are 8 to 10 km (5 to 6 miles). The longest
trench is the Peru-Chile Trench, which extends some 5,900 km (about 3,700
miles) along the west coast of South America. Trenches are relatively narrow,
usually less than 100 km (about 60 miles) wide.
Of Earth’s 20 major trenches, 17 are found in the Pacific basin, a vast area
rimmed by trenches of both marginal and island arc varieties. Marginal trenches
bound the west coast of Central and South America from the Gulf of California
to southern Chile. Although they are deeply buried in sediment, trenches are
found along the western North American continental margin from Cape
Mendocino (in northern California) to the Canadian border. The Aleutian Trench
extends from the northernmost point in the Gulf of Alaska west to the
Kamchatka Peninsula in far eastern Russia. It can be classified as a marginal
trench in the east but is more properly termed an island arc west of Alaska.
In the western Pacific the trenches are associated with island arcs. These
include the Kuril, Japan, Bonin, Mariana, Ryukyu, and Philippine trenches that
extend from Kamchatka to near the Equator. A complex pattern of island arcs is
found in Indonesia. The major island arc here is the Java Trench, extending from
northern Australia to the northwestern end of Sumatra in the northeast Indian
Ocean. The region of New Guinea and the Solomon Islands includes the New
Britain and Solomon trenches, the latter of which joins the New Hebrides Trench
directly to the south. East of this area the Tonga and Kermadec trenches extend
south from the Fiji Islands to New Zealand.
Two island arcs occur in the Atlantic Ocean. The South Sandwich Trench is
located west of the Mid-Atlantic Ridge between South America and Antarctica.
The Puerto Rico Trench joins the Lesser Antilles Island arc in the eastern
Caribbean.
A few trenches are partially filled with sediments derived from the bordering
continents. The Aleutian Trench is effectively buried east of Kodiak Island in the
Gulf of Alaska. There the ocean floor is smooth and flat. To the west, farther
from the sediment supply of Alaska, the trench reaches depths of more than 7
km (about 4 miles). The Lesser Antilles trench in the eastern Caribbean is buried
by sediments originating from South America.
STRUCTURE
Deep-sea trenches and their approaches are striking features on the ocean floor.
In general, the cross sections of deep-sea trenches are V-shaped with steeper
landward sides. Typical slopes range between 4° and 16°, although slopes as
steep as 45° have been measured in the Tonga Trench of the equatorial South
Pacific. Narrow, flat abyssal plains of ponded sediment generally occupy trench
axes; however, in most deep-sea trenches the accumulated material is relatively
shallow since the bottom of the trench subducts into Earth’s interior.
Oceanward of trenches the seafloor is usually bulged upward in an outer
ridge or rise of up to 1,000 metres (about 3,300 feet) in relief. This condition is
thought to be the elastic response of the oceanic plate bending down into a
subduction zone. The landward or island-arc slope of the trench is often
interrupted by a submarine ridge, which sometimes breaks the ocean surface, as
in the case of the Java Trench. Such a ridge is constructed from deformed
sediments scraped off the top of the descending oceanic plate and is termed an
accretionary prism. A line of explosive volcanoes, extruding (erupting) a lava
that forms the volcanic rock andesite, is found on the overriding plate usually
100 km (about 60 miles) or so from the trench. In marginal trenches these
volcanoes form mountain chains, such as the Cascades in the Pacific Northwest
or the great volcanoes of the Andes. In island arcs they form active volcanic
island chains, such as the Mariana Islands.
Behind the volcanic line of island arcs are sometimes found young, narrow
ocean basins. These basins are bounded on the opposite side by submarine
ridges. Such interarc, or backarc, basins are sites of seafloor spreading directly
caused by the dynamics of subduction. They originate at the volcanic line, so
that the outer bounding submarine ridge, or third arc, represents an older portion
of the volcanic line that has spread away. These backarc basins bear many of the
features characteristic of oceanic spreading centres. Well-studied examples of
these features are found in the Lau Basin of the Tonga arc and also west of the
Mariana Islands. The Sea of Japan originated from backarc spreading behind the
Japanese arc that began some 30 million years ago. At least two backarc basins
have opened behind the Mariana arc, creating seafloor in two phases from about
30 to 17 million years ago in the western Parece Vela Basin and from 5 million
years ago in the Mariana Trough next to the islands. The Mariana Trough is a
backarc basin occurring due east of the Mariana arc.
It is important to note that some seafloor features bear the name trench and
are deep linear troughs; however, they do not occur in subduction zones. The
Vema Trench on the Mid-Indian Ridge is a fracture zone. The Vityaz Trench
northwest of Fiji is an aseismic (inactive) feature of unknown origin. The
Diamantina trench (Diamantina Fracture Zone) extends westward from the
southwest coast of Australia. It is a rift valley that was formed when Australia
separated from Antarctica between 60 million and 50 million years ago.

ORIGIN OF DEEP-SEA TRENCHES


Geophysical data provide important clues concerning the origin of trenches. No
abnormalities in the flow of internal Earth heat or variations in Earth’s magnetic
field occur at trenches. Precision measurements reveal that the force of gravity
generally is lower than normal, however. These negative gravity anomalies are
interpreted to mean that the segments of the lithosphere (that is, the crust and
upper mantle comprising the rigid, outermost shell of Earth) that underlie
trenches are being forced down against buoyant isostatic forces.
This interpretation of gravity data is substantiated by seismological studies.
All trenches are associated with zones of earthquake foci. Along the periphery of
the Pacific Ocean, earthquakes occur close to and landward of the trenches, at
depths within Earth of 55 km (34 miles) or less. With increased landward
distance away from the trenches, earthquakes occur at greater and greater depths
—500 km (310 miles) or more. Seismic foci thus define tabular zones
approximately 20 km (12 miles) thick that dip landward at about 45° beneath the
continents. Analyses of these seismic zones and of individual earthquakes
suggest that the seismicity results from the descent of a lithospheric plate with its
associated crust into the asthenosphere (that is, the partially molten layer beneath
the lithosphere); oceanic trenches are topographic expressions of this movement.
The sinking of oceanic lithosphere helps to explain the relative scarcity of
sediment that has accumulated within the trenches. Small quantities of brown or
red clay, siliceous organic remains, volcanic ash and lapilli, and coarse, graded
layers that result from turbidity currents and from the slumping of the trench
walls occur at trench axes. Sediments on trench walls shallower than 4,500
metres (about 14,800 feet) are predominantly calcareous foraminiferal oozes.
Large quantities of sediment cannot accumulate because they either are dragged
into Earth’s interior by the plunging oceanic lithosphere or are distorted into
folded masses and molded into new material of the continental periphery.

OCEANIC CRUST
The oceanic crust is the outermost layer of Earth’s lithosphere that is found
under the oceans. The oceanic crust is formed at spreading centres on oceanic
ridges, and it is about 6 km (4 miles) thick. It is composed of several layers, not
including the overlying sediment. The topmost layer, about 500 metres (1,650
feet) thick, includes lavas made of basalt (that is, rock material consisting largely
of plagioclase [feldspar] and pyroxene). Oceanic crust differs from continental
crust in several ways: it is thinner, denser, younger, of different chemical
composition, and formed above the subduction zones.
The lavas are generally of two types: pillow lavas and sheet flows. Pillow
lavas appear to be shaped exactly as the name implies—like large overstuffed
pillows about 1 metre (3 feet) in cross section and 1 to several metres long. They
commonly form small hills tens of metres high at the spreading centres. Sheet
flows have the appearance of wrinkled bed sheets. They commonly are thin
(only about 10 cm [4 inches] thick) and cover a broader area than pillow lavas.
There is evidence that sheet flows are erupted at higher temperatures than those
of the pillow variety. On the East Pacific Rise at 8° S latitude, a series of sheet
flow eruptions (possibly since the mid-1960s) have covered more than 220
square km (85 square miles) of seafloor to an average depth of 70 metres (230
feet).
Below the lava is a layer composed of feeder, or sheeted, dikes that measures
more than 1 km (0.6 mile) thick. Dikes are fractures that serve as the plumbing
system for transporting magmas (molten rock material) to the seafloor to
produce lavas. They are about 1 metre (3 feet) wide, subvertical, and elongate
along the trend of the spreading centre where they formed, and they abut one
another’s sides—hence the term sheeted. These dikes also are of basaltic
composition. There are two layers below the dikes totaling about 4.5 km (3
miles) in thickness. Both of these include gabbros, which are essentially basalts
with coarser mineral grains. These gabbro layers are thought to represent the
magma chambers, or pockets of lava, that ultimately erupt on the seafloor. The
upper gabbro layer is isotropic (uniform) in structure. In some places this layer
includes pods of plagiogranite, a differentiated rock richer in silica than gabbro.
The lower gabbro layer has a stratified structure and evidently represents the
floor or sides of the magma chamber. This layered structure is called cumulate,
meaning that the layers (which measure up to several metres thick) result from
the sedimentation of minerals out of the liquid magma. The layers in the
cumulate gabbro have less silica but are richer in iron and magnesium than the
upper portions of the crust. Olivine, an iron-magnesium silicate, is a common
mineral in the lower gabbro layer.
A cross section of Earth’s outer layers, from the crust through the
lower mantle. Encyclopædia Britannica, Inc.

The oceanic crust lies atop Earth’s mantle, as does the continental crust.
Mantle rock is composed mostly of peridotite, which consists primarily of the
mineral olivine with small amounts of pyroxene and amphibole.

INVESTIGATIONS OF THE OCEANIC CRUST


Knowledge of the structure and composition of the oceanic crust comes from
several sources. Bottom sampling during early exploration brought up all
varieties of the above-mentioned rocks, but the structure of the crust and the
abundance of the constituent rocks were unclear. Simultaneously, seismic
refraction experiments enabled researchers to determine the layered nature of the
oceanic crust. These experiments involved measuring the travel times of seismic
waves generated by explosions (such as dynamite blasts) set off over distances
of several tens of kilometres. The results of early refraction experiments revealed
the existence of two layers beneath the sediment cover. More sophisticated
experiments and analyses led to dividing these layers into two parts, each with a
different seismic wave velocity, which increases with depth. The seismic
velocity is a kind of fingerprint that can be attributed to a limited number of rock
types. Sampled rock data and seismic results were combined to yield a model for
the structure and composition of the crust.

STUDY OF OPHIOLITES
Great strides in understanding the oceanic crust were made by the study of
ophiolites. These are slices of the ocean floor that have been thrust above sea
level by the action of plate tectonics. In various places in the world, the entire
sequence of oceanic crust and upper mantle is exposed. These areas include,
among others, Newfoundland and the Pacific Coast Ranges of California, the
island of Cyprus in the Mediterranean Sea, and the mountains in Oman on the
southeastern tip of the Arabian Peninsula. Ophiolites reveal the structure and
composition of the oceanic crust in astonishing detail. Also, the process of
crustal formation and hydrothermal circulation, as well as the origin of marine
magnetic anomalies, can be studied with comparative clarity. Although it is clear
that ophiolites are of marine origin, there is some controversy as to whether they
represent typical oceanic crust or crust formed in settings other than an oceanic
spreading centre—behind island arcs, for example.
The age of the oceanic crust does not go back farther than about 200 million
years. Such crust is being formed today at oceanic spreading centres. Many
ophiolites are much older than the oldest oceanic crust, demonstrating continuity
of the formation processes over hundreds of millions of years. Methods that may
be used to determine the age of the crustal material include direct dating of rock
samples by radiometric dating (measuring the relative abundances of a particular
radioactive isotope and its daughter isotopes in the samples) or by the analyses
of fossil evidence, marine magnetic anomalies, or ocean depth. Of these,
magnetic anomalies deserve special attention.
A marine magnetic anomaly is a variation in strength of Earth’s magnetic
field caused by magnetism in rocks of the ocean floor. Marine magnetic
anomalies typically represent 1 percent of the total geomagnetic field strength.
They can be stronger (“positive”) or weaker (“negative”) than the average total
field. Also, the magnetic anomalies occur in long bands that run parallel to
spreading centres for hundreds of kilometres and may reach up to a few tens of
kilometres in width.
MARINE MAGNETIC ANOMALIES
Marine magnetic anomalies were first discovered off the coast of the western
United States in the late 1950s and completely baffled scientists. The anomalies
were charted from southern California to northern Washington and out several
hundred kilometres. Russian-born American geophysicist Victor Vacquier
noticed that these linear anomalies ended at the fracture zones mapped in this
area. In addition, he noticed that they had unique shapes, occurred in a
predictable sequence across their trends, and could be correlated across the
fracture zones. Soon thereafter, linear magnetic anomalies were mapped over the
Reykjanes Ridge south of Iceland. They were found to occur on both sides of the
ridge crest and parallel to it. Simultaneously, Alan Cox and several other
American geophysicists documented evidence that Earth’s magnetic field had
reversed in the past: the north magnetic pole had been the south magnetic pole
about 700,000 years ago, and there were reasons to believe older reversals
existed. Also at this time, American geophysicist Robert S. Dietz and American
geologist Harry H. Hess were formulating the theory of seafloor spreading—the
hypothesis that oceanic crust is created at the crests of the oceanic ridges and
consumed in the deep-sea trenches.
It remained for English geologists Frederick J. Vine and Drummond H.
Matthews and Canadian geophysicist Lawrence W. Morley to put these
observations together in a theory that explained marine magnetic anomalies. The
theory rests on three assumptions: (1) that Earth’s magnetic field periodically
reverses polarity, (2) that seafloor spreading occurs, and (3) that the oceanic
crust is permanently magnetized as it forms and cools at spreading centres. The
theory expresses the assumptions—namely, that the oceanic crust records
reversals of Earth’s field as it is formed during seafloor spreading. Positive
anomalies result when the crust is magnetized in a “normal” polarity parallel to
the ambient field of Earth, and negative anomalies result when the crust is
“reversely” magnetized in an opposite sense. As the magnetized crust moves
down the flanks of a ridge away from the spreading centre, it remains
permanently magnetized and “carries” the magnetic anomalies along with it.
A brilliant leap in understanding was now possible. If the age of the field
reversals were known, the age of the ocean crust could be predicted by mapping
the corresponding anomaly. By the mid-1960s, Cox and his colleagues had put
together a schedule of reversals for the last four or five million years by studying
the ages and magnetic polarities of lava flows found on land. Vine and the
Canadian geologist J. Tuzo Wilson applied the time scale to marine magnetic
anomalies mapped over the Juan de Fuca Ridge, a spreading centre off the
northwest United States. They thus dated the crust there and also computed the
first seafloor spreading rate of about 30 mm (1.2 inches) per year. The rate is
computed by dividing the distance of an anomaly from the ridge crest by the age
of the anomaly twice. Thus the oceanic crust at the Juan de Fuca Ridge is
moving at about 15 mm (0.6 inch) per year away from the ridge crest and at
about 60 mm (2.4 inches) per year away from the crustal segment on the
opposite side of the crest.
During the 1960s and ’70s marine magnetic anomalies were mapped over
wide areas of the ocean basins. By using estimates of the ages of oceanic crust
obtained from core samples by deep-sea drilling, a magnetic anomaly time scale
was constructed, and at the same time the spreading history for the ocean basins
covering the last 200 million years or so was proposed.
It is thought that the most important contributor to marine magnetic
anomalies is the layer of lavas in the upper oceanic crust. A secondary
contribution originates in the upper layer of gabbros. The dike layer is
essentially demagnetized by the action of hydrothermal waters at the spreading
centres. The dominant mechanism of permanent magnetization is the
thermoremanent magnetization (or TRM) of iron-titanium oxide minerals. These
minerals lock in a TRM as they cool below 200–300 °C (392–572 °F) in the
presence of Earth’s magnetic field. Although several processes are capable of
altering the TRM, including reheating and oxidation at the seafloor, it is
remarkably robust, as is evidenced by magnetic anomalies as old as 165 million
years in the far western equatorial Pacific.

OCEANIC PLATEAUS
Oceanic plateaus, which are also called submarine plateaus, are large regions of
submarine elevation that rise sharply at least 200 metres (about 660 feet) above
the surrounding deep-sea floor. Oceanic plateaus are characterized principally by
extensive, relatively flat or gently tilted summits. Most oceanic plateaus were
named early in the 20th century prior to the invention of sonic sounding, and
many of these features have been shown by modern bathymetric data to be
portions of the oceanic ridges. Thus, the Albatross Plateau of the eastern
equatorial Pacific now is recognized as belonging to the East Pacific Rise and
has been shown to possess a much more irregular summit than early data
indicated.
Most plateaus are steplike interruptions of the continental slopes and appear
to be downwarped or downfaulted blocks of former continental shelves. These
marginal plateaus are exemplified by the Blake Plateau off the southeastern
United States. This plateau’s flat surface lies between 700 and 1,000 m (2,300
and 3,300 feet) below sea level, is more than 300 km (185 miles) wide, and
covers approximately 130,000 square km (50,000 square miles) of seafloor. The
crust underlying the plateau, although relatively thin and veneered by flat-lying
marine sediments, is otherwise continental in character.
Other plateaus, such as the coral-capped plateaus of the South China Sea,
occur in the ocean well beyond the continental margins. They stand above the
surrounding deep-sea floor as isolated topographic highs and are believed to be
composed of continental rock cores overlain by flat-lying marine sediments.
Presumably, these mid-ocean plateaus are minor fragments of continent that have
been isolated during continental drift and seafloorspreading.
Crew members aboard a drilling ship inspect a rock core during a
scientific expedition that succeeded for the first time in drilling
through the upper oceanic crust. JOI Alliance/IODP

OCEANIC RIDGES
Oceanic ridges are continuous submarine mountain chains that extend
approximately 80,000 km (50,000 miles) through all the world’s oceans.
Individually, ocean ridges are the largest features in ocean basins. Collectively,
the oceanic ridge system is the most prominent feature on Earth’s surface after
the continents and the ocean basins themselves. In the past these features were
referred to as mid-ocean ridges, but, as will be seen, the largest oceanic ridge,
the East Pacific Rise, is far from a mid-ocean location, and the nomenclature is
thus inaccurate. Oceanic ridges are not to be confused with aseismic ridges,
which have an entirely different origin.

Oceanic ridges offset by transform faults and fracture zones. The


arrows show the direction of movement across the transform faults.
Copyright Encyclopædia Britannica, Inc.; rendering for this edition by
Rosen Educational Services

PRINCIPAL CHARACTERISTICS
Oceanic ridges are found in every ocean basin and appear to girdle the planet.
The ridges rise from depths near 5 km (3 miles) to an essentially uniform depth
of about 2.6 km (1.6 miles) and are roughly symmetrical in cross section. They
can be thousands of kilometres wide. In places, the crests of the ridges are offset
across transform faults within fracture zones, and these faults can be followed
down the flanks of the ridges. (Transform faults are those along which lateral
movement occurs.) The flanks are marked by sets of mountains and hills that are
elongate and parallel to the ridge trend.
New oceanic crust (and part of Earth’s upper mantle, which, together with
the crust, makes up the lithosphere) is formed at seafloor spreading centres at
these crests of the oceanic ridges. Because of this, certain unique geologic
features are found there. Fresh basaltic lavas are exposed on the seafloor at the
ridge crests. These lavas are progressively buried by sediments as the seafloor
spreads away from the site. The flow of heat out of the crust is many times
greater at the crests than elsewhere in the world. Earthquakes are common along
the crests and in the transform faults that join the offset ridge segments. Analysis
of earthquakes occurring at the ridge crests indicates that the oceanic crust is
under tension there. A high-amplitude magnetic anomaly is centred over the
crests because fresh lavas at the crests are being magnetized in the direction of
the present geomagnetic field.
The depths over the oceanic ridges are rather precisely correlated with the
age of the ocean crust; specifically, it has been demonstrated that the ocean depth
is proportional to the square root of crustal age. The theory explaining this
relationship holds that the increase in depth with age is due to the thermal
contraction of the oceanic crust and upper mantle as they are carried away from
the seafloor spreading centre in an oceanic plate. Because such a tectonic plate is
ultimately about 100 km (62 miles) thick, contraction of only a few percent
predicts the entire relief of an oceanic ridge. It then follows that the width of a
ridge can be defined as twice the distance from the crest to the point where the
plate has cooled to a steady thermal state. Most of the cooling takes place within
70 or 80 million years, by which time the ocean depth is about 5 to 5.5 km (3.1
to 3.5 miles). Because this cooling is a function of age, slow-spreading ridges,
such as the Mid-Atlantic Ridge, are narrower than faster-spreading ridges, such
as the East Pacific Rise. Further, a correlation has been found between global
spreading rates and the transgression and regression of ocean waters onto the
continents. About 100 million years ago, during the early Cretaceous Period
when global spreading rates were uniformly high, oceanic ridges occupied
comparatively more of the ocean basins, causing the ocean waters to transgress
(spill over) onto the continents, leaving marine sediments in areas now well
away from coastlines.
Besides ridge width, other features appear to be a function of spreading rate.
Global spreading rates range from 10 mm (0.4 inch) per year or less up to 160
mm (6.3 inches) per year. Oceanic ridges can be classified as slow (up to 50 mm
[about 2 inches]) per year, intermediate (up to 90 mm [about 3.5 inches]) per
year, and fast (up to 160 mm per year). Slow-spreading ridges are characterized
by a rift valley at the crest. Such a valley is fault-controlled. It is typically 1.4
km (0.9 mile) deep and 20–40 km (about 12–25 miles) wide. Faster-spreading
ridges lack rift valleys. At intermediate rates, the crest regions are broad highs
with occasional fault-bounded valleys no deeper than 200 metres (about 660
feet). At fast rates, an axial high is present at the crest. The slow-spreading rifted
ridges have rough faulted topography on their flanks, while the faster-spreading
ridges have much smoother flanks.
DISTRIBUTION OF MAJOR RIDGES AND SPREADING CENTRES
Oceanic spreading centres are found in all the ocean basins. In the Arctic Ocean
a slow-rate spreading centre is located near the eastern side in the Eurasian
basin. It can be followed south, offset by transform faults, to Iceland. Iceland has
been created by a hot spot located directly below an oceanic spreading centre.
The ridge leading south from Iceland is named the Reykjanes Ridge, and,
although it spreads at 20 mm (0.8 inch) per year or less, it lacks a rift valley. This
is thought to be the result of the influence of the hot spot.

The Atlantic Ocean


The Mid-Atlantic Ridge extends from south of Iceland to the extreme South
Atlantic Ocean near 60° S latitude. It bisects the Atlantic Ocean basin, which led
to the earlier designation of mid-ocean ridge for features of this type. The Mid-
Atlantic Ridge became known in a rudimentary fashion during the 19th century.
In 1855 Matthew Fontaine Maury of the U.S. Navy prepared a chart of the
Atlantic in which he identified it as a shallow “middle ground.” During the
1950s the American oceanographers Bruce Heezen and Maurice Ewing
proposed that it was a continuous mountain range.
In the North Atlantic the ridge spreads slowly and displays a rift valley and
mountainous flanks. In the South Atlantic spreading rates are between slow and
intermediate, and rift valleys are generally absent, as they occur only near
transform faults.
The Atlantic Ocean, with depth contours and submarine features.

The Indian Ocean


A very slow oceanic ridge, the Southwest Indian Ridge, bisects the ocean
between Africa and Antarctica. It joins the Mid-Indian and Southeast Indian
ridges east of Madagascar. The Carlsberg Ridge is found at the north end of the
Mid-Indian Ridge. It continues north to join spreading centres in the Gulf of
Aden and Red Sea. Spreading is very slow at this point but approaches
intermediate rates on the Carlsberg and Mid-Indian ridges. The Southeast Indian
Ridge spreads at intermediate rates. This ridge continues from the western Indian
Ocean in a southeasterly direction, bisecting the ocean between Australia and
Antarctica. Rifted crests and rugged mountainous flanks are characteristic of the
Southwest Indian Ridge. The Mid-Indian Ridge has fewer features of this kind,
and the Southeast Indian Ridge has generally smoother topography. The latter
also displays distinct asymmetric seafloor spreading south of Australia. Analysis
of magnetic anomalies shows that rates on opposite sides of the spreading centre
have been unequal at many times over the past 50 or 60 million years.

The Indian Ocean, with depth contours and undersea features.


The Indian Ocean, with depth contours and undersea features.
Encyclopaedia Britannica, Inc.

Pacific Ocean
The Pacific-Antarctic Ridge can be followed from a point midway between New
Zealand and Antarctica northeast to where it joins the East Pacific Rise off the
margin of South America. The former spreads at intermediate to fast rates.
The East Pacific Rise extends from this site northward to the Gulf of
California, where it joins the transform zone of the Pacific-North American plate
boundary. Offshore from Chile and Peru, the East Pacific Rise is currently
spreading at fast rates of 159 mm (6.3 inches) per year or more. Rates decrease
to about 60 mm (about 2.4 inches) per year at the mouth of the Gulf of
California. The crest of the ridge displays a low topographic rise along its length
rather than a rift valley. The East Pacific Rise was first detected during the
Challenger Expedition of the 1870s. It was described in its gross form during the
1950s and ’60s by oceanographers, including Heezen, Ewing, and Henry W.
Menard. During the 1980s, Kenneth C. Macdonald, Paul J. Fox, and Peter F.
Lonsdale discovered that the main spreading centre appears to be interrupted and
offset a few kilometres to one side at various places along the crest of the East
Pacific Rise. However, the ends of the offset spreading centres overlap each
other by several kilometres. These were identified as a new type of geologic
feature of oceanic spreading centres and were designated overlapping spreading
centres. Such centres are thought to result from interruptions of the magma
supply to the crest along its length and define a fundamental segmentation of the
ridge on a scale of tens to hundreds of kilometres.

The Pacific Ocean, with depth contours and submarine features.

Many smaller spreading centres branch off the major ones or are found
behind island arcs. In the western Pacific, spreading centres occur on the Fiji
Plateau between the New Hebrides and Fiji Islands and in the Woodlark Basin
between New Guinea and the Solomon Islands. A series of spreading centres and
transform faults lie between the East Pacific Rise and South America near 40° to
50° S latitude. The Scotia Sea between South America and the Antarctic
Peninsula contains a spreading centre. The Galápagos spreading centre trends
east-west between the East Pacific Rise and South America near the Equator.
Three short spreading centres are found a few hundred kilometres off the shore
of the Pacific Northwest. These are the Gorda Ridges off northern California, the
Juan de Fuca Ridge off Oregon and Washington, and the Explorer Ridge off
Vancouver Island.
In a careful study of the seafloor spreading history of the Galapagos and the
Juan de Fuca spreading centres, the American geophysicist Richard N. Hey
developed the idea of the propagating rift. In this phenomenon, one branch of a
spreading centre ending in a transform fault lengthens at the expense of the
spreading centre across the fault. The rift and fault propagate at one to five times
the spreading rate and create chevron patterns in magnetic anomalies and the
grain of the seafloor topography resembling the wake of a boat.

SPREADING CENTRES AND ASSOCIATED PHENOMENA


From the 1970s highly detailed studies of spreading centres using deeply towed
instruments, photography, and manned submersibles have resulted in new
revelations about the processes of seafloor spreading. The most profound
discoveries have been of deep-sea hydrothermal vents and previously unknown
biological communities.

Spreading Centre Zones


Spreading centres are divided into several geologic zones. The neovolcanic zone
is at the very axis. It is 1–2 km (0.6–1.2 miles) wide and is the site of recent and
active volcanism and of the hydrothermal vents. It is marked by chains of small
volcanoes or volcanic ridges. Adjacent to the neovolcanic zone is one marked by
fissures in the seafloor. This may be 1 to 2 km wide. Beyond this point occurs a
zone of active faulting. Here, fissures develop into normal faults with vertical
offsets. This zone may be 10 km (about 6 miles) or more wide. At slow
spreading rates the faults have offsets of hundreds of metres, creating rift valleys
and rift mountains. At faster rates the vertical offsets are 50 metres (about 160
feet) or less. A deep rift valley is not formed because the vertical uplifts are
cancelled out by faults that downdrop uplifted blocks. This results in linear,
fault-bounded abyssal hills and valleys trending parallel to the spreading centre.
Warm springs emanating from the seafloor in the neovolcanic zone were first
found on the Galápagos spreading centre. These waters were measured to have
temperatures about 20 °C (36 °F) above the ambient temperature. In 1979
hydrothermal vents with temperatures near 350 °C (662 °F) were discovered on
the East Pacific Rise off Mexico. Since then similar vents have been found on
the spreading centres off the Pacific Northwest coast of the United States, on the
south end of the northern Mid-Atlantic Ridge, and at many locations on the East
Pacific Rise.

Hydrothermal Vents
Hydrothermal vents are localized discharges of heated seawater. They result
from cold seawater percolating down into the hot oceanic crust through the zone
of fissures and returning to the seafloor in a pipelike flow at the axis of the
neovolcanic zone. The heated waters often carry sulfide minerals of zinc, iron,
and copper leached from the crust. Outflow of these heated waters probably
accounts for 20 percent of Earth’s heat loss. Exotic biological communities exist
around the hydrothermal vents. These ecosystems are totally independent of
energy from the Sun. They are not dependent on photosynthesis but rather on
chemosynthesis by sulfur-fixing bacteria. The sulfide minerals precipitated in the
neovolcanic zone can accumulate in substantial amounts and are sometimes
buried by lava flows at a later time. Such deposits are mined as commercial ores
in ophiolites on Cyprus and in Oman.

Magma Chambers
Magma chambers have been detected beneath the crest of the East Pacific Rise
by seismic experiments. (The principle underlying the experiments is that
partially molten or molten rock slows the travel of seismic waves and also
strongly reflects them.) The depth to the top of the chambers is about 2 km (1.2
miles) below the seafloor. The width is more difficult to ascertain but is probably
1 to 4 km (0.6 to 2.5 miles). Their thickness seems to be about 2 to 6 km (1.2 to
3.7 miles), on the basis of studies of ophiolites. The chambers have been mapped
along the trend of the crest between 9° and 13° N latitude. The top is relatively
continuous, but is apparently interrupted by offsets of transform faults and
overlapping spreading centres.
CHAPTER 3
CIRCULATION AND WAVES

Seawater is in constant motion. Water circulates around the globe through ocean
currents, which are, generally speaking, propelled by winds affecting water at
the ocean’s surface above and by changes in the density of fluids below. Ocean
circulation is also affected by pressure gradients, the Coriolis effect, friction, and
other factors. Ocean waves are manifestations of energy transferred through a
liquid medium. Waves (which occur as gravity waves and tides) are most
apparent at the ocean’s surface and in locations where the oceans border
landmasses.

THE CIRCULATION OF THE OCEAN WATERS


The general circulation of the oceans defines the average movement of seawater,
which, like the atmosphere, follows a specific pattern. Superimposed on this
pattern are oscillations of tides and waves, which are not considered part of the
general circulation. There also are meanders and eddies that represent temporal
variations of the general circulation. The ocean circulation pattern exchanges
water of varying characteristics, such as temperature and salinity, within the
interconnected network of oceans and is an important part of the heat and
freshwater fluxes of the global climate. Horizontal movements are called
currents, which range in magnitude from a few centimetres per second to as
much as 4 metres (13.1 feet) per second. A characteristic surface speed is about 5
to 50 cm (2 to 20 inches) per second. Currents diminish in intensity with
increasing depth. Vertical movements, often referred to as upwelling and
downwelling, exhibit much lower speeds, amounting to only a few metres per
month. As seawater is nearly incompressible, vertical movements are associated
with regions of convergence and divergence in the horizontal flow patterns.
Ocean circulation derives its energy at the sea surface from two sources that
define two circulation types: (1) wind-driven circulation forced by wind stress
on the sea surface, inducing a momentum exchange, and (2) thermohaline
circulation driven by the variations in water density imposed at the sea surface
by exchange of ocean heat and water with the atmosphere, inducing a buoyancy
exchange. These two circulation types are not fully independent, since the sea-
air buoyancy and momentum exchange are dependent on wind speed. The wind-
driven circulation is the more vigorous of the two and is configured as large
gyres that dominate an ocean region. The wind-driven circulation is strongest in
the surface layer. The thermohaline circulation is more sluggish, with a typical
speed of 1 cm (0.4 inch) per second, but this flow extends to the seafloor and
forms circulation patterns that envelop the global ocean.

DISTRIBUTION OF OCEAN CURRENTS


Maps of the general circulation at the sea surface are constructed from a vast
amount of data obtained from inspecting the residual drift of ships after course
direction and speed are accounted for in a process called dead reckoning. This
information is amplified by satellite-tracked drifters at sea. The pattern is nearly
entirely that of wind-driven circulation.
Deep-ocean circulation consists mainly of thermohaline circulation. The
currents are inferred from the distribution of seawater properties, which trace the
spreading of specific water masses. The distribution of density or field of mass is
also used to estimate the deep currents. Direct observations of subsurface
currents are made by deploying current meters from bottom-anchored moorings
and by setting out neutral buoyant instruments whose drift at depth is tracked
acoustically.

Major surface currents of the world’s oceans. Subsurface currents also


move vast amounts of water, but they are not known in such detail.
Merriam-Webster, Inc.

CAUSES OF OCEAN CURRENTS


The general circulation is governed by the equation of motion, one of Sir Isaac
Newton’s fundamental laws of mechanics applied to a continuous volume of
water. This equation states that the product of mass and current acceleration
equals the vector sum of all forces that act on the mass. Besides gravity, the most
important forces that cause and affect ocean currents are horizontal pressure-
gradient forces, Coriolis forces, and frictional forces. Temporal and inertial terms
are generally of secondary importance to the general flow, though they become
important for transient features of meanders and eddies.

PRESSURE GRADIENTS
The hydrostatic pressure, p, at any depth below the sea surface is given by the
equation p = gρz, where g is the acceleration of gravity, ρ is the density of
seawater, which increases with depth, and z is the depth below the sea surface.
This is called the hydrostatic equation, which is a good approximation for the
equation of motion for forces acting along the vertical. Horizontal differences in
density (due to variations of temperature and salinity) measured along a specific
depth cause the hydrostatic pressure to vary along a horizontal plane or
geopotential surface, a surface perpendicular to the direction of the gravity
acceleration. Horizontal gradients of pressure, though much smaller than vertical
changes in pressure, give rise to ocean currents.
In a homogeneous ocean, which would have a constant potential density,
horizontal pressure differences are possible only if the sea surface is tilted. In
this case, surfaces of equal pressure, called isobaric surfaces, are tilted in the
deeper layers by the same amount as the sea surface. This is referred to as the
barotropic field of mass. The unchanged pressure gradient gives rise to a current
speed independent of depth. The oceans of the world, however, are not
homogeneous. Horizontal variations in temperature and salinity cause the
horizontal pressure gradient to vary with depth. This is the baroclinic field of
mass, which leads to currents that vary with depth. The horizontal pressure
gradient in the ocean is a combination of these two mass fields.
The tilt, or topographic relief, of the isobaric surface marking sea surface
(defined as p = 0) can be constructed from a three-dimensional density
distribution using the hydrostatic equation. Since the absolute value of pressure
is not known at any depth in the ocean, the sea surface slope is presented relative
to that of a deep isobaric surface; it is assumed that the deep isobaric surface is
level. Since the wind-driven circulation attenuates with increasing depth, an
associated decrease of isobaric tilt with increasing depth is expected.
Representation of the sea surface relief relative to a deep reference surface is a
good representation of the absolute shape of the sea surface. The total relief of
the sea surface amounts to about 2 metres (6.6 feet), with “hills” in the
subtropics and “valleys” in the polar regions. This pressure head drives the
surface circulation.

CORIOLIS EFFECT
The rotation of Earth about its axis causes moving particles to behave in a way
that can only be understood by adding a rotational dependent force. To an
observer in space, a moving body would continue to move in a straight line
unless the motion were acted upon by some other force. To an Earth-bound
observer, however, this motion cannot be along a straight line because the
reference frame is the rotating Earth. This is similar to the effect that would be
experienced by an observer standing on a large turntable if an object moved over
the turntable in a straight line relative to the “outside” world. An apparent
deflection of the path of the moving object would be seen. If the turntable
rotated counterclockwise, the apparent deflection would be to the right of the
direction of the moving object, relative to the observer fixed on the turntable.
This remarkable effect is evident in the behaviour of ocean currents. It is
called the Coriolis force, named after Gustave-Gaspard Coriolis, a 19th-century
French engineer and mathematician. For Earth, horizontal deflections due to the
rotational induced Coriolis force act on particles moving in any horizontal
direction. There also are apparent vertical forces, but these are of minor
importance to ocean currents. Because Earth rotates from west to east about its
axis, an observer in the Northern Hemisphere would notice a deflection of a
moving body toward the right. In the Southern Hemisphere, this deflection
would be toward the left. At the equator there would be no apparent horizontal
deflection.
It can be shown that the Coriolis force always acts perpendicular to motion.
Its horizontal component, Cf, is proportional to the sine of the geographic
latitude (θ, given as a positive value for the Northern Hemisphere and a negative
value for the Southern Hemisphere) and the speed, c, of the moving body. It is
given by Cf = c (2ω sin θ), where ω = 7.29 × 10−5 radian per second is the
angular velocity of Earth’s rotation.

FRICTIONAL FORCES
Movement of water through the oceans is slowed by friction, with surrounding
fluid moving at a different velocity. A faster-moving fluid layer tends to drag
along a slower-moving layer, and a slower-moving layer will tend to reduce the
speed of a faster-moving layer. This momentum transfer between the layers is
referred to as frictional forces. The momentum transfer is a product of
turbulence that moves kinetic energy to smaller scales until at the centimetre
scale it is dissipated as heat. The wind blowing over the sea surface transfers
momentum to the water. This frictional force at the sea surface (i.e., the wind
stress) produces the wind-driven circulation. Currents moving along the ocean
floor and the sides of the ocean also are subject to the influence of boundary-
layer friction. The motionless ocean floor removes momentum from the
circulation of the ocean waters.

GEOSTROPHIC CURRENTS
For most of the ocean volume away from the boundary layers, which have a
characteristic thickness of 100 metres (about 330 feet), frictional forces are of
minor importance, and the equation of motion for horizontal forces can be
expressed as a simple balance of horizontal pressure gradient and Coriolis force.
This is called geostrophic balance.
On a nonrotating Earth, water would be accelerated by a horizontal pressure
gradient and would flow from high to low pressure. On the rotating Earth,
however, the Coriolis force deflects the motion, and the acceleration ceases only
when the speed, c, of the current is just fast enough to produce a Coriolis force
that can exactly balance the horizontal pressure-gradient force. This geostrophic
balance is given as dp/dn = ρc2ω sin θ, where dp/dn is the horizontal pressure
gradient. From this balance, it follows that the current direction must be
perpendicular to the pressure gradient because the Coriolis force always acts
perpendicular to the motion. In the Northern Hemisphere this direction is such
that the high pressure is to the right when looking in current direction, while in
the Southern Hemisphere it is to the left. This type of current is called a
geostrophic current. The simple equation given above provides the basis for an
indirect method of computing ocean currents. The relief of the sea surface also
defines the streamlines (paths) of the geostrophic current at the surface relative
to the deep reference level. The hills represent high pressure, and the valleys
stand for low pressure. Clockwise rotation in the Northern Hemisphere with
higher pressure in the centre of rotation is called anticyclonic motion.
Counterclockwise rotation with lower pressure in its centre is cyclonic motion.
In the Southern Hemisphere the sense of rotation is the opposite, because the
effect of the Coriolis force has changed its sign of deflection.

DENSITY CURRENTS IN THE OCEANS

Density currents are currents that are kept in motion by the force of gravity
acting on a relatively small density difference caused by variations in
salinity, temperature, or sediment concentration. Salinity and temperature
variations produce stratification in oceans. Below the surface layer, which is
disturbed by waves and is lighter than the deeper waters because it is
warmer or less saline, the oceans are composed of layers of water that have
distinctive chemical and physical characteristics, which move more or less
independently of each other and which do not lose their individuality by
mixing even after they have flowed for hundreds of kilometres from their
point of origin.
An example of this type of density current, or stratified flow, is
provided by the water of the Mediterranean Sea as it flows through the
Strait of Gibraltar out into the Atlantic. Because the Mediterranean is
enclosed in a basin that is relatively small compared with the ocean basins
and because it is located in a relatively arid climate, evaporation exceeds
the supply of fresh water from rivers. The result is that the Mediterranean
contains water that is both warmer and more saline than normal deep-sea
water, the temperature ranging from 12.7 to 14.5 °C (54.9 to 58.1 °F) and
the salinity from 38.4 to 39.0 parts per thousand. Because of these
characteristics, the Mediterranean water is considerably denser than the
water in the upper parts of the North Atlantic, which has a salinity of about
36 parts per thousand and a temperature of about 13 °C (55.4 °F). The
density contrast causes the lighter Atlantic water to flow into the
Mediterranean in the upper part of the Strait of Gibraltar (down to a depth
of about 200 metres [about 660 feet]) and the denser Mediterranean water
to flow out into the Atlantic in the lower part of the strait (from about 200
metres to the top of the sill separating the Mediterranean from the Atlantic
at a depth of 320 metres [1,050 feet]).
Because the strait is only about 20 km (12.4 miles) wide, both inflow
and outflow achieve relatively high speeds. Near the surface the inflow may
have speeds as high as 2 metres (6.6 feet) per second, and the outflow
reaches speeds of more than 1 metre (3.3 feet) per second at a depth of
about 275 metres (about 900 feet). One result of the high current speeds in
the strait is that there is a considerable amount of mixing, which reduces the
salinity of the outflowing Mediterranean water to about 37 parts per
thousand. The outflowing water sinks to a depth of about 1,500 metres
(about 4,920 feet) or more, where it encounters colder, denser Atlantic
water. It then spreads out as a layer of more saline water between two
Atlantic water masses.

EKMAN LAYER
The wind exerts stress on the ocean surface proportional to the square of the
wind speed and in the direction of the wind, setting the surface water in motion.
This motion extends to a depth of about 100 metres (330 feet) in what is called
the Ekman layer, after the Swedish oceanographer V. Walfrid Ekman, who in
1902 deduced these results in a theoretical model constructed to help explain
observations of wind drift in the Arctic. Within the oceanic Ekman layer the
wind stress is balanced by the Coriolis force and frictional forces. The surface
water is directed at an angle of 45° to the wind, to the right in the Northern
Hemisphere and to the left in the Southern Hemisphere. With increasing depth in
the boundary layer, the current speed is reduced, and the direction rotates farther
away from the wind direction following a spiral form, becoming antiparallel to
the surface flow at the base of the layer where the speed is of the surface
speed. This so-called Ekman spiral may be the exception rather than the rule, as
the specific conditions are not often met, though deflection of a wind-driven
surface current at somewhat smaller than 45° is observed when the wind field
blows with a steady force and direction for the better part of a day. The average
water particle within the Ekman layer moves at an angle of 90° to the wind; this
movement is to the right of the wind direction in the Northern Hemisphere and
to its left in the Southern Hemisphere. This phenomenon is called Ekman
transport, and its effects are widely observed in the oceans.
Since the wind varies from place to place, so does the Ekman transport,
forming convergence and divergence zones of surface water. A region of
convergence forces surface water downward in a process called downwelling,
while a region of divergence draws water from below into the surface Ekman
layer in a process known as upwelling. Upwelling and downwelling also occur
where the wind blows parallel to a coastline. The principal upwelling regions of
the world are along the eastern boundary of the subtropical ocean waters, as, for
example, the coastal region of Peru and northwestern Africa. Upwelling in these
regions cools the surface water and brings nutrient-rich subsurface water into the
sunlit layer of the ocean, resulting in a biologically productive region. Upwelling
and high productivity also are found along divergence zones at the equator and
around Antarctica. The primary downwelling regions are in the subtropical
ocean waters—e.g., the Sargasso Sea in the North Atlantic. Such areas are
devoid of nutrients and are poor in marine life.
The vertical movements of ocean waters into or out of the base of the Ekman
layer amount to less than 1 metre (3.3 feet) per day, but they are important since
they extend the wind-driven effects into deeper waters. Within an upwelling
region, the water column below the Ekman layer is drawn upward. This process,
with conservation of angular momentum on the rotating Earth, induces the water
column to drift toward the poles. Conversely, downwelling forces water into the
water column below the Ekman layer, inducing drift toward the equator. An
additional consequence of upwelling and downwelling for stratified waters is to
create a baroclinic field of mass. Surface water is less dense than deeper water.
Ekman convergences have the effect of accumulating less dense surface water.
This water floats above the surrounding water, forming a hill in sea level and
driving an anticyclonic geostrophic current that extends well below the Ekman
layer. Divergences do the opposite; they remove the less dense surface water,
replacing it with denser, deeper water. This induces a depression in sea level
with a cyclonic geostrophic current.
The ocean current pattern produced by the wind-induced Ekman transport is
called the Sverdrup transport, after the Norwegian oceanographer H.U.
Sverdrup, who formulated the basic theory in 1947. Several years later (1950),
the American geophysicist and oceanographer Walter H. Munk and others
expanded Sverdrup’s work, explaining many of the major features of the wind-
driven general circulation by using the mean climatological wind stress
distribution at the sea surface as a driving force.

WIND-DRIVEN CIRCULATION
Wind stress induces a circulation pattern that is similar for each ocean. In each
case, the wind-driven circulation is divided into large gyres that stretch across
the entire ocean: subtropical gyres extend from the equatorial current system to
the maximum westerlies in a wind field near 50° latitude, and subpolar gyres
extend poleward of the maximum westerlies. The depth penetration of the wind-
driven currents depends on the intensity of ocean stratification: for those regions
of strong stratification, such as the tropics, the surface currents extend to a depth
of less than 1,000 metres (about 3,300 feet). Within the low-stratification polar
regions, the wind-driven circulation reaches all the way to the seafloor.

EQUATORIAL CURRENTS
At the equator the currents are for the most part directed toward the west, the
North Equatorial Current in the Northern Hemisphere and the South Equatorial
Current in the Southern Hemisphere. Near the thermal equator, where the
warmest surface water is found, there occurs the eastward-flowing Equatorial
Counter Current. This current is slightly north of the geographic equator,
drawing the northern fringe of the South Equatorial Current to 5° N. The offset
to the Northern Hemisphere matches a similar offset in the wind field. The east-
to-west wind across the tropical ocean waters induces Ekman transport
divergence at the equator, which cools the surface water there.
At the geographic equator a jetlike current is found just below the sea
surface, flowing toward the east counter to the surface current. This is called the
Equatorial Undercurrent. It attains speeds of more than 1 metre (3.3 feet) per
second at a depth of nearly 100 metres (330 feet). It is driven by higher sea level
in the western margins of the tropical ocean, producing a pressure gradient,
which in the absence of a horizontal Coriolis force drives a west-to-east current
along the equator. The wind field reverses the flow within the surface layer,
inducing the Equatorial Undercurrent.
Equatorial circulation undergoes variations following the irregular periods of
roughly three to eight years of the Southern Oscillation (i.e., fluctuations of
atmospheric pressure over the tropical Indo-Pacific region). Weakening of the
east-to-west wind during a phase of the Southern Oscillation allows warm water
in the western margin to slip back to the east by increasing the flow of the
Equatorial Counter Current. Surface water temperatures and sea level decrease in
the west and increase in the east. This event is called El Niño. The combined El
Niño/Southern Oscillation effect has received much attention because it is
associated with global-scale climatic variability. In the tropical Indian Ocean, the
strong seasonal winds of the monsoons induce a similarly strong seasonal
circulation pattern.

THE SUBTROPICAL GYRES


These are anticyclonic circulation features. The Ekman transport within these
gyres forces surface water to sink, giving rise to the subtropical convergence
near 20°–30° latitude. The centre of the subtropical gyre is shifted to the west.
This westward intensification of ocean currents was explained by the American
meteorologist and oceanographer Henry M. Stommel (1948) as resulting from
the fact that the horizontal Coriolis force increases with latitude. This causes the
poleward-flowing western boundary current to be a jetlike current that attains
speeds of 2 to 4 metres (6.6 to 13.1 feet) per second. This current transports the
excess heat of the low latitudes to higher latitudes. The flow within the
equatorward-flowing interior and eastern boundary of the subtropical gyres is
quite different. It is more of a slow drift of cooler water that rarely exceeds 10
cm (3.9 inches) per second. Associated with these currents is coastal upwelling
that results from offshore Ekman transport.
The strongest of the western boundary currents is the Gulf Stream in the
North Atlantic Ocean. It carries about 30 million cubic metres (1.1 billion cubic
feet) of ocean water per second through the Straits of Florida and roughly 80
million cubic metres (2.8 billion cubic feet) per second as it flows past Cape
Hatteras off the coast of North Carolina, U.S. Responding to the large-scale wind
field over the North Atlantic, the Gulf Stream separates from the continental
margin at Cape Hatteras. After separation, it forms waves or meanders that
eventually generate many eddies of warm and cold water. The warm eddies,
composed of thermocline water normally found south of the Gulf Stream, are
injected into the waters of the continental slope off the coast of the northeastern
United States. They drift to the southeast at rates of approximately 5 to 8 cm (2
to 3.1 inches) per second, and after a year they rejoin the Gulf Stream north of
Cape Hatteras. Cold eddies of slope water are injected into the region south of
the Gulf Stream and drift to the southwest. After two years they reenter the Gulf
Stream just north of the Antilles Islands. The path that they follow defines a
clockwise-flowing recirculation gyre seaward of the Gulf Stream.
Among the other western boundary currents, the Kuroshio of the North
Pacific is perhaps the most like the Gulf Stream, having a similar transport and
array of eddies. The Brazil and East Australian currents are relatively weak. The
Agulhas Current has a transport close to that of the Gulf Stream. It remains in
contact with the margin of Africa around the southern rim of the continent. It
then separates from the margin and curls back to the Indian Ocean in what is
called the Agulhas Retroflection. Not all the water carried by the Agulhas
returns to the east; about 10 to 20 percent is injected into the South Atlantic
Ocean as large eddies that slowly migrate across it.

THE SUBPOLAR GYRES


The subpolar gyres are cyclonic circulation features. The Ekman transport within
these features forces upwelling and surface water divergence. In the North
Atlantic the subpolar gyre consists of the North Atlantic Current at its
equatorward side and the Norwegian Current that carries relatively warm water
northward along the coast of Norway. The heat released from the Norwegian
Current into the atmosphere maintains a moderate climate in northern Europe.
Along the east coast of Greenland is the southward-flowing cold East Greenland
Current. It loops around the southern tip of Greenland and continues flowing
into the Labrador Sea. The southward flow that continues off the coast of
Canada is called the Labrador Current. This current separates for the most part
from the coast near Newfoundland to complete the subpolar gyre of the North
Atlantic. Some of the cold water of the Labrador Current, however, extends
farther south.
In the North Pacific the subpolar gyre is composed of the northward-flowing
Alaska Current, the Aleutian Current (also known as the Subarctic Current), and
the southward-flowing cold Oyashio Current. The North Pacific Current forms
the separation between the subpolar and subtropical gyres of the North Pacific.
In the Southern Hemisphere, the subpolar gyres are less defined. Large
cyclonic flowing gyres lie poleward of the Antarctic Circumpolar Current and
can be considered counterparts to the Northern Hemispheric subpolar gyres. The
best-formed is the Weddell Gyre of the South Atlantic sector of the Southern
Ocean. The Antarctic coastal current flows toward the west. The northward-
flowing current off the east coast of the Antarctic Peninsula carries cold
Antarctic coastal water into the circumpolar belt. Another cyclonic gyre occurs
north of the Ross Sea.

ANTARCTIC CIRCUMPOLAR CURRENT


The Southern Ocean links the major oceans by a deep circumpolar belt in the
50°–60° S range. In this belt flows the Antarctic Circumpolar Current from west
to east, encircling the globe at high latitudes. It transports 125 million cubic
metres (4.4 billion cubic feet) of seawater per second over a path of about 24,000
km (14,900 miles) and is the most important factor in diminishing the
differences between oceans. The Antarctic Circumpolar Current is not a well-
defined single-axis current but rather consists of a series of individual filaments
separated by frontal zones. It reaches the seafloor and is guided along its course
by the irregular bottom topography. Large meanders and eddies develop in the
current as it flows. These features induce poleward transfer of heat, which may
be significant in balancing the oceanic heat loss to the atmosphere above the
Antarctic region farther south.

THERMOHALINE CIRCULATION
The general circulation of the oceans consists primarily of the wind-driven
currents. These, however, are superimposed on the much more sluggish
circulation driven by horizontal differences in temperature and salinity—namely,
the thermohaline circulation. The thermohaline circulation reaches down to the
seafloor and is often referred to as the deep, or abyssal, ocean circulation.
Measuring seawater temperature and salinity distribution is the chief method of
studying the deep-flow patterns. Other properties also are examined; for
example, the concentrations of oxygen, carbon-14, and such synthetically
produced compounds as chlorofluorocarbons are measured to obtain resident
times and spreading rates of deep water.
In some areas of the ocean, generally during the winter season, cooling or net
evaporation causes surface water to become dense enough to sink. Convection
penetrates to a level where the density of the sinking water matches that of the
surrounding water. It then spreads slowly into the rest of the ocean. Other water
must replace the surface water that sinks. This sets up the thermohaline
circulation. The basic thermohaline circulation is one of sinking of cold water in
the polar regions, chiefly in the northern North Atlantic and near Antarctica.
These dense water masses spread into the full extent of the ocean and gradually
upwell to feed a slow return flow to the sinking regions. A theory for the
thermohaline circulation pattern was proposed by Stommel and Arnold Arons in
1960.

Thermohaline circulation transports and mixes the water of the


oceans. In the process it transports heat, which influences regional
climate patterns. The density of seawater is determined by the
temperature and salinity of a volume of seawater at a particular
location. The difference in density between one location and another
drives the thermohaline circulation. Encyclopaedia Britannica, Inc.

In the Northern Hemisphere, the primary region of deep water formation is


the North Atlantic; minor amounts of deep water are formed in the Red Sea and
Persian Gulf. A variety of water types contribute to the so-called North Atlantic
Deep Water. Each one of them differs, though they share a common attribute of
being relatively warm (greater than 2 °C [35.6 °F]) and salty (greater than 34.9
parts per thousand) compared with the other major producer of deep and bottom
water, the Southern Ocean (0 °C [32 °F] and 34.7 parts per thousand). North
Atlantic Deep Water is primarily formed in the Greenland and Norwegian seas,
where cooling of the salty water introduced by the Norwegian Current induces
sinking. This water spills over the rim of the ridge that stretches from Greenland
to Scotland, extending to the seafloor to the south as a convective plume. It then
flows southward, pressed against the western edge of the North Atlantic.
Additional deep water is formed in the Labrador Sea. This water, somewhat less
dense than the overflow water from the Greenland and Norwegian seas, has been
observed sinking to a depth of 3,000 metres (about 9,800 feet) within convective
features referred to as chimneys. Vertical velocities as high as 10 cm (3.9 inches)
per second have been observed within these convective features.
A third variety of North Atlantic Deep Water is derived from net evaporation
within the Mediterranean Sea. This draws surface water into the Mediterranean
through the Strait of Gibraltar. The mass of salty water formed within the
Mediterranean exits as a deeper stream. It descends to depths of 1,000 to 2,000
metres (3,300 to 6,600 feet) in the North Atlantic Ocean, forming the uppermost
layer of North Atlantic Deep Water. The outflow in the Strait of Gibraltar
reaches as high as 2 metres (about 7 feet) per second, but its total transport
amounts to only 5 percent of the total North Atlantic Deep Water formed. The
outflow of the Mediterranean plays a significant role in boosting the salinity of
North Atlantic Deep Water.
The blend of North Atlantic Deep Water, with a total formation rate of 15 to
20 × 106 cubic metres (5.3 to 7.1 × 108 cubic feet) per second, quickly ventilates
the Atlantic Ocean, resulting in a residence time of less than 200 years. The deep
water spreads away from its source along the western side of the Atlantic Ocean
and, on reaching the Antarctic Circumpolar Current, spreads into the Indian and
Pacific oceans. The sinking of North Atlantic Deep Water is compensated for by
the slow upwelling of deep water, mainly in the Southern Ocean, to replenish the
upper stratum of water that has descended as North Atlantic Deep Water. North
Atlantic Deep Water exported to the other oceans must be balanced by the inflow
of upper-layer water into the Atlantic. Some water returns as cold, lowsalinity
Pacific water through the Drake Passage in the form of what is known as
Antarctic Intermediate Water, and some returns as warm salty thermocline water
from the Indian Ocean around the southern rim of Africa.
Remnants of North Atlantic Deep Water mix with Southern Ocean water to
spread along the seafloor into the North Pacific Ocean. Here, it upwells to a level
of 2,000–3,000 metres (6,600–9,800 feet) and returns to the south lower in
salinity and oxygen but higher in nutrient concentrations as North Pacific Deep
Water. This North Pacific Deep Water is eventually swept eastward with the
Antarctic Circumpolar Current. Modification of deep water in the North Pacific
is the direct consequence of vertical mixing, which carries into the deep ocean
the low salinity properties of North Pacific Intermediate Water. The latter is
formed in the northwestern Pacific Ocean. Because of the immenseness of the
North Pacific and the extremely long residence time (more than 500 years) of the
water, enormous quantities of North Pacific Deep Water can be produced by
vertical mixing.
Considerable volumes of cold water generally of low salinity are formed in
the Southern Ocean. Such water masses spread into the interior of the global
ocean and to a large extent are responsible for the anomalous cold, lowsalinity
state of the modern oceans. The circumstances leading to this role for the
Southern Ocean are related to the existence of a deep-ocean circumpolar belt
around Antarctica that was established some 25 million years ago by the shifting
lithospheric plates which make up Earth’s surface. This belt establishes the
Antarctic Circumpolar Current, which isolates Antarctica from the warm surface
waters of the subtropics. The Antarctic Circumpolar Current does not completely
sever contact with the lower latitudes. The Southern Ocean does have access to
the waters of the north, but through deep-and bottom-water pathways. The basic
dynamics of the Antarctic Circumpolar Current lifts dense deep water occurring
north of the current to the ocean surface south of it. Once exposed to the cold
Antarctic air masses, the upwelling deep water is converted to the cold Antarctic
Bottom Water and Antarctic Intermediate Water. The southward and upwelling
deep water, which carries heat injected into the deep ocean by processes farther
north, is balanced by the northward spread of cooler, fresher, oxygenated water
masses of the Southern Ocean. It is estimated that the overturning rate of water
south of the Antarctic Circumpolar Current amounts to 35 to 45 million cubic
metres (1.2 to 1.6 billion cubic feet) per second, most of which becomes
Antarctic Bottom Water.
The primary site of Antarctic Bottom Water formation is within the
continental margins of the Weddell Sea, though some is produced in other
coastal regions, such as the Ross Sea. Also, there is evidence of deep convective
overturning farther offshore. Antarctic Bottom Water, formed at a rate of 30
million cubic metres (1.1 billion cubic feet) per second, slips below the Antarctic
Circumpolar Current and spreads to regions well north of the equator. Slowly
upwelling and modified by mixing with less dense water, it returns to the
Southern Ocean as deep water.
The remaining upwelling of deep water spreads near the surface to the north,
where it forms Antarctic Intermediate Water within the Antarctic Circumpolar
Current zone and spreads along the base of the thermoclines farther north. This
water mass forms a sheet of lowsalinity water that demarcates the lower
boundary of the subtropical thermocline. It upwells into the thermocline, partly
compensating for the sinking of North Atlantic Deep Water.

WAVES OF THE SEA


There are many types of ocean waves. Waves differ from each other in size and
in terms of the forces that drive them. Waves represent an oscillatory motion of
seawater at regular time intervals or periods. Some may be running, or
progressive, waves in which the crests propagate, while others are stationary, or
standing, waves. Two of the more common types of waves, gravity waves and
tides, are considered here. For gravity waves, the stabilizing force—i.e., the
force that attempts to restore the crests and troughs of the waves to the average
sea level—is Earth’s gravity. The distance between the crests, or wavelength, of
gravity waves range from a few centimetres to many kilometres. Tiny waves at
the ocean surface with a wavelength of less than 1.7 cm (0.7 inch) are called
capillary waves. Their restoring force is the surface tension of seawater.
Capillary waves are direct products of the wind stress exerted on the sea surface
and tend to feed wind energy into gravity waves, which characteristically have
longer wavelengths.

Types of surface waves and their relative energy levels. Copyright


Encyclopædia Britannica, Inc.; rendering for this edition by Rosen
Educational Services

Tides are essentially gravity waves that have long periods of oscillation.
They may be called forced waves, because they have fixed, prescribed periods
that are strictly determined by astronomical forces induced by the relative
movements of the Moon, Earth, and Sun. Sometimes the term “tidal wave” is
used incorrectly to include such phenomena as surges, which are called storm
tides, or destructive waves known as tsunamis that are induced by undersea
earthquakes. In the following discussion, the use of the words “tide” and “tidal”
is restricted to tides of astronomical origin and the forces and phenomena
associated with them.

SURFACE GRAVITY WAVES


Of the nontidal kinds of running surface waves, three types may be
distinguished: wind waves and swell, wind surges, and sea waves of seismic
origin (tsunamis).

WIND WAVES AND SWELL


Wind waves are the wind-generated gravity waves. After the wind has abated or
shifted or the waves have migrated away from the wind field, such waves
continue to propagate as swell.
The dependence of the sizes of the waves on the wind field is a complicated
one. A general impression of this dependence is given by the descriptions of the
various states of the sea corresponding to the scale of wind strengths known as
the Beaufort scale, named after the British admiral Sir Francis Beaufort, who
drafted it in 1808 using as his yardstick the surface of sail that a fully rigged
warship of those days could carry in the various wind forces. When considering
the descriptions of the sea surface, it must be remembered that the size of the
waves depends not only on the strength of the wind but also on its duration and
its fetch—i.e., the length of its path over the sea.
The theory of waves starts with the concept of simple waves, those forming a
strictly periodic pattern with one wavelength and one wave period and
propagating in one direction. Real waves, however, always have a more irregular
appearance. They may be described as composite waves, in which a whole
spectrum of wavelengths, or periods, is present and which have more or less
diverging directions of propagation. In reporting observed wave heights and
periods (or lengths) or in forecasting them, one height or one period is
mentioned as the height or period, however, and some agreement is needed in
order to guarantee uniformity of meaning. The height of simple waves means the
elevation difference between the top of a crest and the bottom of a trough. The
significant height, a characteristic height of irregular waves, is by convention the
average of the highest one-third of the observed wave heights. Period, or
wavelength, can be determined from the average of a number of observed time
intervals between the passing of successive well-developed wave crests over a
certain point, or of observed distances between them.
Wave period and wavelength are coupled by a simple relationship:
wavelength equals wave period times wave speed, or L = TC, when L is
wavelength, T is wave period, and C is wave speed.
The wave speed of surface gravity waves depends on the depth of water and
on the wavelength, or period; the speed increases with increasing depth and
increasing wavelength, or period. If the water is sufficiently deep, the wave
speed is independent of water depth. This relationship of wave speed to
wavelength and water depth (d) is given by the equations below. With g being
the gravity acceleration (9.8 metres [32 feet] per second squared), C2 = gd, when
the wavelength is 20 times greater than the water depth (waves of this kind are
called long gravity waves or shallow-water waves); and C2 = gL/2π, when the
wavelength is less than two times the water depth (such waves are called short
waves or deep water waves). For waves with lengths between 2 and 20 times the
water depth, the wave speed is governed by a more complicated equation
combining these effects:

where tanh is the hyperbolic tangent.


A few examples are listed below for short waves, giving the period in
seconds, the wavelength in metres, and wave speed in metres per second:

Waves often appear in groups as the result of interference of wave trains of


slightly differing wavelengths. A wave group as a whole has a group speed that
generally is less than the speed of propagation of the individual waves; the two
speeds are equal only for groups composed of long waves. For deepwater waves,
the group velocity (V) is half the wave speed (C). In the physical sense, group
velocity is the velocity of propagation of wave energy. From the dynamics of the
waves, it follows that the wave energy per unit area of the sea surface is
proportional to the square of the wave height, except for the very last stage of
waves running into shallow water, shortly before they become breakers.
The height of wind waves increases with increasing wind speed and with
increasing duration and fetch of the wind (i.e., the distance over which the wind
blows). Together with height, the dominant wavelength also increases. Finally,
however, the waves reach a state of saturation because they attain the maximum
significant height to which the wind can raise them, even if duration and fetch
are unlimited. For instance, winds of 5 metres (16 feet) per second, 15 metres
(about 50 feet) per second, and 25 metres (about 80 feet) per second may raise
waves with significant heights up to 0.5 metre (1.6 feet), 4.5 metres (15 feet),
and 12.5 metres (about 40 feet), respectively, with corresponding wavelengths of
16 metres (about 52 feet), 140 metres (about 460 feet), and 400 metres (1,300
feet), respectively.
After becoming swell, the waves may travel thousands of kilometres over the
ocean, particularly if the swell is from the large storms of moderate and high
latitudes, whence it easily may travel into the subtropical and equatorial zones,
and the swell of the trade winds, which runs into the equatorial calms. In
traveling, the swell waves gradually become lower; energy is lost by internal
friction and air resistance and by energy dissipation because of some divergence
of the directions of propagation (fanning out). With respect to the energy loss,
there is a selective damping of the composite waves, the shorter waves of the
wave mixture suffering a stronger damping over a given distance than the longer
ones. As a consequence, the dominant wavelength of the spectrum shifts toward
the greater wavelengths. Therefore, an old swell must always be a long swell.
When waves run into shallow water, their speed of propagation and
wavelength decrease, but the period remains the same. Eventually, the group
velocity, the velocity of energy propagation, also decreases, and this decrease
causes the height to increase. The latter effect may, however, be affected by
refraction of the waves, a swerving of the wave crests toward the depth lines and
a corresponding deviation of the direction of propagation. Refraction may cause
a convergence or divergence of the energy stream and result in a raising or
lowering of the waves, especially over nearshore elevations or depressions of the
sea bottom.
In the final stage, the shape of the waves changes, and the crests become
narrower and steeper until, finally, the waves become breakers (surf). Generally,
this occurs where the depth is 1.3 times the wave height.

WIND SURGES
Running wind surges are long waves caused by a piling up of the water over a
large area through the action of a traveling wind or pressure field. Examples
include the surge in front of a traveling storm cyclone, particularly the
devastating hurricane surge caused by a tropical cyclone, and the surge
occasionally caused by a wind convergence line, such as a traveling front with a
sharp wind shift.

WAVES OF SEISMIC ORIGIN


A tsunami (Japanese: tsu, “harbour,” and nami, “sea”) is a very long wave of
seismic origin that is caused by a submarine or coastal earthquake, landslide, or
volcanic eruption. Such a wave may have a length of hundreds of kilometres and
a period on the order of a quarter of an hour. It travels across the ocean at a
tremendous speed. (Tsunamis are long waves traveling at the wave speed given
by C2 = gd.) To a depth of 4,000 metres (about 13,100 feet), for instance, the
corresponding wave speed is about 200 metres (about 660 feet) per second, or
800 km (about 500 miles) per hour. In the open ocean the height of tsunamis
may be less than 1 metre (3.3 feet), and they pass unnoticed. As they approach a
continental shelf, however, their speed is reduced and their height increases
dramatically. Tsunamis have caused enormous destruction of life and property,
piling up in coastal waters at places thousands of kilometres away from their
point of origin, particularly in the Pacific Ocean.

After being generated by an undersea earthquake or landslide, a


tsunami may propagate unnoticed over vast reaches of open ocean
before cresting in shallow water and inundating a coastline.
Encyclopædia Britannica, Inc.

TSUNAMIS

Tsunamis, which are also called seismic sea waves, are catastrophic ocean
waves usually caused by submarine earthquakes, underwater or coastal
landslides, or volcanic eruptions. The term tsunami is Japanese for “harbour
wave.” The term tidal wave is frequently used for such a wave, but it is a
misnomer, for the wave has no connection with the tides.
After an earthquake or other generating impulse occurs, a train of
simple, progressive oscillatory waves is propagated great distances over the
ocean surface in ever-widening circles, much like the waves produced by a
pebble falling into a shallow pool. In deep water a tsunami can travel as fast
as 800 km (500 miles) per hour. The wavelengths are enormous, about 100
to 200 km (60 to 120 miles), but the wave amplitudes (heights) are very
small, only about 30 to 60 cm (1 to 2 feet). The waves’ periods (the lengths
of time for successive crests or troughs to pass a single point) are very long,
varying from five minutes to more than an hour. These long periods,
coupled with the extremely low steepness of the waves, enables them to be
completely obscured in deep water by normal wind waves and swell.

The aftermath of the December 2004 tsunami in Aceh, Indon.


Philip A. McDaniel/U.S. Navy
As the waves approach the coast of a continent, however, friction with
the rising sea bottom reduces the velocity of the waves. As the velocity
lessens, the wavelengths become shortened and the wave amplitudes
increase. Coastal waters may rise as high as 30 metres (100 feet) above
normal sea level in 10 to 15 minutes. By a poorly understood process, the
continental shelf waters begin to oscillate after the rise in sea level.
Between three and five major oscillations generate most of the damage,
frequently appearing as powerful “run-ups” of rushing water that uproot
trees, pull buildings off their foundations, carry boats far inshore, and wash
away entire beaches, peninsulas, and other lowlying coastal formations.
Frequently the succeeding outflow of water is just as destructive as the run-
up or even more so. In any case, oscillations may continue for several days
until the ocean surface reaches equilibrium.
The Pacific Tsunami Warning Center, located near Honolulu, Hawaii,
was established in 1949, three years after a tsunami generated by a
submarine earthquake near the Aleutian Islands struck the island of Hawaii
around Hilo, killing 159 people. Following the disaster of December 2004,
UNESCO set a goal of establishing similar systems for the Indian Ocean
and eventually the entire globe.

SEICHE WAVES AND INTERNAL WAVES


A freestanding wave may arise in an enclosed or nearly enclosed basin as a free
swinging or sloshing of the whole water mass. Such a standing wave is also
called a seiche, after the name given to the oscillating movements of the water of
Lake Geneva, Switzerland, where this phenomenon first was studied seriously.
In addition, gravity waves may also occur on internal “surfaces” within oceans.
These surfaces represent strata of rapidly changing water density with increasing
depth, and the associated waves are called internal waves.

STANDING, OR SEICHE, WAVES


In seiche waves, the period of oscillation is independent of the force that first
brought the water mass out of equilibrium (and that is supposed to have ceased
thereafter), but depends only on the dimensions of the enclosing basin and on the
direction in which the water mass is swinging. Assuming a simple rectangular
basin of constant depth and the most simple lengthwise oscillation, the period of
oscillation (T) is equal to two times the length of the basin divided by the wave
speed computed from the shallow-water formula above. This relationship may
be written: T = L/C, in which L equals two times the length of the basin and C is
the wave speed found from the formula, using the known depth of the basin.
Besides this fundamental tone (or response to stimuli), the water mass also may
swing according to an overtone, showing one or more nodal lines across the
basin.
The water in an open bay or marginal sea also may perform such a free
oscillation as a standing wave, the difference being that in an open bay the
greatest horizontal displacements are not in the middle of the bay but at the
mouth. For the fundamental period of oscillation, the formula given above is
used with a wavelength equal to four times the length (from the mouth to the
closed end) of the bay. In practice, of course, it is more difficult than that,
because the form of a bay or marginal sea is irregular and the depth differs from
place to place. The North Sea has a period of lengthwise swinging of about 36
hours. The cause of such free oscillations may be a temporary wind or pressure
field, which brought the sea surface out of its horizontal position and which
afterward ceased to act more or less abruptly, leaving the water mass out of
equilibrium.

INTERNAL WAVES
Internal waves manifest themselves by a regular rising and sinking of the water
layers around which they centre, whereas the height of the sea surface is hardly
affected at all. Because the restoring force, excited by the internal deformation of
the water layers of equal density, is much smaller than in the case of surface
waves, internal waves are much slower than the latter. Given the same
wavelength, the period is much longer (the movements of the water particles
being much more sluggish), and the speed of propagation is much smaller; the
formulas for the speed of surface waves include the acceleration of gravity, g,
but those for internal waves include the gravity factor times the difference
between the densities of the upper and the lower water layer divided by their
sum.
The cause of internal waves may lie in the action of tidal forces (the period
then equaling the tidal period) or in the action of a wind or pressure fluctuation.
Sometimes, a ship may cause internal waves (dead water) if there is a shallow,
brackish upper layer.

OCEAN TIDES
The tides may be regarded as forced waves, partially running waves and partially
standing waves. They are manifested by vertical movements of the sea surface
(the height maximum and minimum are called high water [HW] and low water
[LW]) and in alternating horizontal movements of the water, the tidal currents.
The words ebb and flow are used to designate the falling tide and the rising tide,
respectively.

TIDE-GENERATING FORCES
The forces that cause the tides are called the tide-generating forces. A tide-
generating force is the resultant force of the attracting force of the Moon or the
Sun and the force of inertia (centrifugal force) that results from the orbital
movement of Earth around the common centre of gravity of the Earth-Moon or
Earth-Sun system.
Considering the Earth-Moon system, at any time the tide-generating force is
directed vertically upward at the two places on Earth where the Moon is in the
vertical (on the same and on the opposite side of Earth); it is directed vertically
downward at all places (forming a circle) where the Moon is in the horizon at
that moment. At all other places, the tide-generating force also has a horizontal
component. Because this pattern of forces is coupled to the position of the Moon
with respect to Earth and because for any place on Earth’s surface the relative
position of the Moon with respect to that place has, on the average, a periodicity
of 24 hours 50 minutes, the tide-generating force felt at any place has that same
periodicity. When the Moon is in the plane of the equator, the force runs through
two identical cycles within this time interval because of the symmetry of the
global pattern of forces described above. Consequently, the tidal period is 12
hours 25 minutes in this case; it is the period of the semidiurnal lunar tide. The
fact that the Moon is alternately to the north and to the south of the equator
causes an inequality of the two successive cycles within the time interval of 24
hours 50 minutes. The effect of this inequality is formally described as the
superposition of a partial tide called the diurnal lunar tide, with the period of 24
hours 50 minutes, on the semidiurnal lunar tide.
In the same manner, the Sun causes a semidiurnal solar tide, with a 12-hour
period, and a diurnal solar tide, with a 24-hour period. In a complete description
of the local variations of the tidal forces, still other partial tides play a role
because of further inequalities in the orbital motions of the Moon and Earth.
The interference of the solar-tidal forces with the lunar-tidal forces (the lunar
forces are about 2.2 times as strong) causes the regular variation of the tidal
range between spring tide, when it has its maximum, and neap tide, when it has
its minimum.

Tides are caused by the gravitational pull of the Sun and the Moon on
Earth’s water. When the Sun, Moon, and Earth form a straight line
(top), higher and lower tides than usual are generated. In contrast,
when the lines between the Sun and Earth and the Moon and Earth are
perpendicular to one another (bottom), high tides and low tides are
moderated. Encyclopædia Britannica, Inc.

Although the tide-generating forces are very small in comparison with


Earth’s force of gravity (the lunar tidal force at its maximum being only 1.14 ×
10-7 times the force of gravity), their effects upon the sea are considerable
because of their horizontal component. Since Earth is not surrounded by an
uninterrupted envelope of water but rather shows a very irregular alternation of
sea and land, the mechanism of the response of the oceans and seas to the tidal
forces is extremely complex. A further complication is caused by the deflecting
force of Earth’s rotation.
In enclosures formed by gulfs and bays, the local tide is generated by
interaction with the tides of the adjacent open ocean. Such a tide often takes the
form of a running tidal wave that rotates within the confines of the enclosure. In
some semi-enclosed seas, such as the Mediterranean, Black, and Baltic seas, a
standing wave, or tidal seiche, may be generated by the local tide-raising forces.
In these seas, the tidal range of sea level is only on the order of centimetres.
In the open ocean, it generally is on the order of tens of centimetres. In bays and
adjacent seas, however, the tidal range may be much greater, because the shape
of a bay or adjacent sea may favour the enhancement of the tide inside; in
particular, there may be a resonance of the basin concerned with the tide. The
largest known tides occur in the Bay of Fundy, where spring tidal ranges up to
15 metres (about 50 feet) have been measured.

TIDAL BORES
Tidal bores form on rivers and estuaries near a coast where there is a large tidal
range and the incoming tide is confined to a narrow channel. They consist of a
surge of water moving swiftly upstream headed by a wave or series of waves.
Such bores are quite common. There is a large one, known as the mascaret on
the Seine, which forms on spring tides and reaches as far upriver as Rouen.
There is a well-known bore on the Severn, in England, and another forms on the
Petitcodiac River, which empties into the Bay of Fundy in New Brunswick. The
classic example is the bore on the Qiantang (Ch’ien-t’ang) described by
Commander W. Usborne Moore of the British navy in 1888 and 1892. He
reported heights of 2.5 to 3.5 metres (8.2 to 11.5 feet).
When a tidal bore forms in a river, the direction of flow of the water changes
abruptly as the bore passes. Before it arrives, the water may be still or, more
usually, a small freshwater current flows outward toward the sea. The tide comes
in as a “wall of water” that passes up the river. Behind the bore, the current flows
upriver. At the division between the moving water behind the bore and the still
water in front, there is a wave, the water surface behind being higher than it is in
front. This wave must travel more quickly than the water particles behind it,
because, as the advancing water travels upriver, it collects the still water in front
and sets it in motion. Upriver, the advancing tide will consist not of salt water
from the sea but rather of fresh water that has passed farther down and been
collected and returned in front of the incoming tide. It is therefore necessary to
distinguish between the velocity of the advancing wave and that of the water
particles just behind it.
CHAPTER 4
THE CLIMATIC AND ECONOMIC IMPACTS OF THE
OCEANS

Since the oceans are integral parts of Earth’s climate, they can influence events
deep within the interiors of continents. Major currents, such as the Gulf Stream
and Kuroshio, deliver warm water from the tropics to coastal subarctic regions,
and these subarctic regions are characterized by milder climates than other areas
located at similar latitudes. The oceans are spawning grounds for tropical
cyclones, which often bring damaging high winds and flooding to coastal areas.
The oceans are also tremendous reservoirs of heat, and the movement of large
volumes of warm water can temporarily disrupt normal precipitation and heating
patterns. The movement of warm water eastward across the tropical Pacific
Ocean as part of the El Niño/Southern Oscillation affects the weather in regions
both adjacent to and far from the tropical Pacific.
The oceans are also sites of significant economic activity. Ocean fishes and
other organisms are routinely harvested as food by humans. The process of
desalinization allows the oceans to provide water resources to thirsty human
settlements. Energy resources, such as petroleum and natural gas, and raw
materials are also extracted from the oceans, and these bodies of water also serve
as media for transportation and communication and sites of waste disposal.

SEASONAL AND INTERANNUAL OCEAN-ATMOSPHERE


INTERACTIONS
The notion of a connection between the temperature of the surface layers of the
oceans and the circulation of the lowest layer of the atmosphere, the troposphere,
is a familiar one. The surface mixed layer of the ocean is a huge reservoir of heat
when compared to the overlying atmosphere. The heat capacity of an
atmospheric column of unit area cross-section extending from the ocean surface
to the outermost layers of the atmosphere is equivalent to the heat capacity of a
column of seawater of 2.6-metre (8.5-feet) depth. The surface layer of the oceans
is continuously being stirred by the overlying winds and waves, and thus a
surface mixed layer is formed that has vertically uniform properties in
temperature and salinity. This mixed layer, which is in direct contact with the
atmosphere, has a minimum depth of 20 metres (about 66 feet) in summer and a
maximum depth exceeding 100 metres (about 330 feet) in late winter in the mid-
latitudes. In lower latitudes the seasonal variation in the mixed layer is less
marked than at higher latitudes, except in regions such as the Arabian Sea, where
the onset of the southwestern Indian monsoon may produce large changes in the
depth of the mixed layer. Temperature anomalies (i.e., deviations from the
normal seasonal temperature) in the surface mixed layer have a long residence
time compared with those of the overlying turbulent atmosphere. Hence they
may persist for a number of consecutive seasons and even for years.
Observational studies to investigate the relationship between anomalies in
ocean surface temperature and the tropospheric circulation have been undertaken
primarily in the Pacific and Atlantic. They have identified large-scale ocean
surface temperature anomalies that have similar spatial scales to monthly and
seasonal anomalies in atmospheric circulation. The longevity of the ocean
surface temperature anomalies, as compared with the shorter dynamical and
thermodynamical “memory” of the atmosphere, has suggested that they may be
an important predictor for seasonal and interannual climate anomalies.
THE LINK BETWEEN OCEAN SURFACE TEMPERATURE
AND CLIMATE ANOMALIES
First, it is useful to consider some examples of the association between
anomalies in ocean surface temperature and irregular changes in climate. The
Sahel, a region that borders the southern fringe of the Sahara in Africa,
experienced a number of devastating droughts during the 1970s and ’80s, which
can be compared with a much wetter period during the 1950s. Data was obtained
that showed the difference in ocean surface temperature during the period from
July to September between the “driest” and “wettest” rainfall seasons in the
Sahel after 1950. Of particular note were the higher-than-normal surface
temperatures in the tropical South Atlantic, Indian, and Southeast Pacific oceans
and the lower-than-normal temperatures in the North Atlantic and Pacific
oceans. This example illustrates that climate anomalies in one region of the
world may be linked to ocean surface temperature changes on a global scale.
Global atmospheric modeling studies undertaken during the mid-1980s have
indicated that the positions of the main rainfall zones in the tropics are sensitive
to anomalies in ocean surface temperature.
Shorter-lived climate anomalies, on time scales of months to one or two
years, also have been related to ocean surface temperature anomalies. The
equatorial oceans have the largest influence on these climate anomalies because
of the evaporation of water. A relatively small change in ocean surface
temperature, say, of 1 °C (1.8 °F), may result in a large change in the evaporation
of water into the atmosphere. The increased water vapour in the lower
atmosphere is condensed in regions of upward motion known as convergence
zones. This process liberates latent heat of condensation, which in turn provides
a major fraction of the energy to drive tropical circulation and is one of the
mechanisms responsible for the El Niño/Southern Oscillation (ENSO)
phenomenon.
Given the sensitivity of the tropical atmosphere to variations in tropical sea
surface temperature, there also has been considerable interest in their influence
on extratropical circulation. The sensitivity of the tropospheric circulation to
surface temperature in both the tropical Pacific and Atlantic oceans has been
shown in theoretical and observational studies alike. Figures were prepared to
demonstrate the correlation between the equatorial ocean surface temperature in
the east Pacific (the location of El Niño) and the atmospheric circulation in the
middle troposphere during winter. The atmospheric pattern was a characteristic
circulation type known as the Pacific-North American (PNA) mode. Such
patterns are intrinsic modes of the atmosphere, which may be forced by thermal
anomalies in the tropical atmosphere and which in their turn are forced by
tropical ocean surface temperature anomalies.
As noted earlier, enhanced tropical sea surface temperatures increase
evaporation into the atmosphere. In the 1982–83 El Niño event a pattern of
circulation anomalies occurred throughout the Northern Hemisphere during
winter. These modes of the atmosphere, however, account for much less than 50
percent of the variability of the circulation in mid-latitudes, though in certain
regions (northern Japan, southern Canada, and the southern United States), they
may have sufficient amplitude for them to be used for predicting seasonal
surface temperature perhaps up to two seasons in advance.
The response of the atmosphere to mid-latitude ocean surface anomalies has
been difficult to detect unambiguously because of the complexity of the
turbulent westerly flow between 20° and 60° latitude in both hemispheres. This
flow has many properties of nonlinear chaotic systems and thus exhibits
behaviour that is difficult to predict beyond a couple of weeks. The atmosphere
alone can exhibit large fluctuations on seasonal and longer time scales without
any change in external forcing conditions, such as ocean surface temperature.
Notwithstanding this inherent problem, some effects of ocean surface
temperature anomalies on the atmosphere have been observed and modeled.
The influence of the oceans on the atmosphere in the mid-latitudes is greatest
during autumn and early winter when the ocean mixed layer releases to the
atmosphere the large quantities of heat that it has stored up over the previous
summer. Anomalies in ocean surface temperature are indicative of either a
surplus or a deficiency of heat available to the atmosphere. The response of the
atmosphere to ocean surface temperature, however, is not random
geographically. The circulation over the North Atlantic and northern Europe
during early winter has been found to be sensitive to large ocean surface
temperature anomalies south of Newfoundland. When a warm positive anomaly
exists in this region, an anomalous surface anticyclone occurs in the central
Atlantic at a similar latitude to the temperature anomaly, and an anomalous
cyclonic circulation is located over the North Sea, Scandinavia, and central
Europe. With colder than normal water south of Newfoundland, the circulation
patterns are reversed, producing cyclonic circulation over the central Atlantic
and anticyclonic circulation over Europe. The sensitivity of the atmosphere to
ocean surface temperature anomalies in this particular region is thought to be
related to the position of the overlying storm tracks and jet stream. The region is
the most active in the Northern Hemisphere for the growth of storms associated
with very large heat fluxes from the surface layer of the ocean.
Another example of a similar type of air-sea interaction event has been
documented over the North Pacific Ocean. A statistical seasonal relationship
exists between the summer ocean temperature anomaly in the Gulf of Alaska and
the atmospheric circulation over the Pacific and North America during the
following autumn and winter. The presence of warmer-than-normal ocean
surface temperature in the Gulf of Alaska results in increased cyclone
development during the subsequent autumn and winter. The relationship has
been established by means of monthly sea surface temperature and atmospheric
pressure data collected over 30 years in the North Pacific Ocean.
The air-sea interaction events in both the North Pacific and North Atlantic
oceans raise questions as to how the anomalies in ocean surface temperature in
these areas are initiated, how they are maintained, and whether they yield useful
information for atmospheric prediction beyond the normal time scales of weather
forecasting (i.e., one to two weeks). Statistical analysis of previous case studies
have shown that ocean surface temperature anomalies initially develop in
response to anomalous atmospheric forcing. Once developed, however, the
temperature anomaly of the ocean surface tends to reinforce and thereby
maintain the anomalous atmospheric circulation. The mechanisms thought to be
responsible for this behaviour in the ocean are the surface wind drift, wind
mixing, and the interchange of heat between the ocean and atmosphere. The
question of prediction is therefore difficult to answer, as these events depend on
a synchronous and interconnected behaviour between the atmosphere and the
surface layer of the ocean, which allows for positive feedback between the two
systems.

FORMATION OF TROPICAL CYCLONES


Tropical cyclones represent still another example of sea-air interactions. These
storm systems are known as hurricanes in the North Atlantic and eastern North
Pacific and as typhoons in the western North Pacific. The winds of such systems
revolve around a centre of low pressure in an anticlockwise direction in the
Northern Hemisphere and in a clockwise direction in the Southern Hemisphere.
The winds attain velocities in excess of 115 km per hour, or 65 knots, in most
cases. Tropical cyclones may last from a few hours to as long as two weeks, the
average lifetime being six days.
A satellite image shows the approach of Hurricane Andrew toward the
Florida coast over the course of three days in August of 1992. NASA
Goddard Space Flight Center

The oceans provide the source of energy for tropical cyclones both by direct
heat transfer from their surface (known as sensible heat) and by the evaporation
of water. This water is subsequently condensed within a storm system, thereby
releasing latent heat energy. When a tropical cyclone moves over land, this
energy is severely depleted and the circulation of the winds is consequently
weakened.
Such storms are truly phenomena of the tropical oceans. They originate in
two distinct latitude zones, between 4° and 22° S and between 4° and 35° N.
They are absent in the equatorial zone between 4° S and 4° N. Most tropical
cyclones are spawned on the poleward side of the region known as the
intertropical convergence zone (ITCZ).
More than two-thirds of observed tropical cyclones originate in the Northern
Hemisphere, and roughly the same proportion occur in the Eastern Hemisphere.
The North Pacific has more than one-third of all such storms, while the southeast
Pacific and South Atlantic are normally devoid of them. Most northern
hemispheric tropical cyclones occur between May and November, with peak
periods in August and September. The majority of southern hemispheric
cyclones occur between December and April, with peaks in January and
February.
EFFECTS OF TROPICAL CYCLONES ON OCEAN WATERS
A tropical cyclone can affect the thermal structure and currents in the surface
layer of the ocean waters in its path. Cooling of the surface layer occurs in the
wake of such a storm. Maximum cooling occurs on the right of a hurricane’s
path in the Northern Hemisphere. In the wake of Hurricane Hilda’s passage
through the Gulf of Mexico in 1964 at a translational speed of only 5 knots (5.75
miles per hour), the surface waters were cooled by as much as 6 °C (42.8 °F).
Tropical cyclones that have higher translational velocities cause less cooling of
the surface. The surface cooling is caused primarily by wind-induced upwelling
of cooler water from below the surface layer. The warm surface water is
simultaneously transported toward the periphery of the cyclone, where it
downwells into the deeper ocean layers. Heat loss across the air-sea interface and
the wind-induced mixing of the surface water with those of the cooler subsurface
layers make a significant but smaller contribution to surface cooling.
In addition to surface cooling, tropical cyclones may induce large horizontal
surge currents and vertical displacements of the thermocline. The surge currents
have their largest amplitude at the surface, where they may reach velocities
approaching 1 metre (3.3 feet) per second. The horizontal currents and the
vertical displacement of the thermocline observed in the wake of a tropical
cyclone oscillate close to the inertial period. These oscillations remain for a few
days after the passage of the storm and spread outward from the rear of the
system as an internal wake on the thermocline. The vertical motion may
transport nutrients from the deeper layers into the sunlit surface waters, which in
turn promotes phytoplankton blooms (i.e., the rapid growth of diatoms and other
minute one-celled organisms). The ocean surface temperature normally recovers
to its precyclone value within 10 days of a storm’s passage.

INFLUENCE ON ATMOSPHERIC CIRCULATION AND RAINFALL


Tropical cyclones play an important role in the general circulation of the
atmosphere, accounting for 2 percent of the global annual rainfall and between 4
and 5 percent of the global rainfall in August and September at the height of the
Northern Hemispheric cyclone season. For a local area, the occurrence of a
single tropical cyclone can have a major impact on the region’s annual rainfall.
Furthermore, tropical cyclones contribute approximately 2 percent of the kinetic
energy of the general circulation of the atmosphere, some of which is exported
from the tropics to higher latitudes.
THE GULF STREAM AND KUROSHIO SYSTEMS
The Gulf Stream and the Kuroshio Current are prominent elements of Earth’s
oceanic circulation system. Geographically, both occur off the eastern coasts of
large continents and flow from the tropics toward the poles, bringing tremendous
amounts of warm water northward. Both currents influence weather and climate
in the Northern Hemisphere.

THE GULF STREAM


This major current system, as described earlier, is a western boundary current
that flows poleward along a boundary separating the warm and more saline
waters of the Sargasso Sea to the east from the colder, slightly fresher
continental slope waters to the north and west. The warm, saline Sargasso Sea,
composed of a water mass known as North Atlantic Central Water, has a
temperature that ranges from 8 to 19 °C (46 to 66 °F) and a salinity between
35.10 and 36.70 parts per thousand. This is one of the two dominant water
masses of the North Atlantic Ocean, the other being the North Atlantic Deep
Water, which has a temperature of 2.2 to 3.5 °C (36 to 38.3 °F) and a salinity
between 34.90 and 34.97 parts per thousand, and which occupies the deepest
layers of the ocean (generally below 1,000 metres [about 3,300 feet]). The North
Atlantic Central Water occupies the upper layer of the North Atlantic Ocean
between roughly 20° and 40° N. The “lens” of this water is at its lowest depth of
1,000 metres in the northwest Atlantic and becomes progressively shallower to
the east and south. To the north it shallows abruptly and outcrops at the surface
in winter, and it is at this point that the Gulf Stream is most intense.
The Gulf Stream flows along the rim of the warm North Atlantic Central
Water northward from the Florida Straits along the continental slope of North
America to Cape Hatteras. There, it leaves the continental slope and turns
northeastward as an intense meandering current that extends toward the Grand
Banks of Newfoundland. Its maximum velocity is typically between 1 and 2
metres (about 3.3 to 6.6 feet) per second. At this stage, a part of the current loops
back onto itself, flowing south and east. Another part flows eastward toward
Spain and Portugal, while the remaining water flows northeastward as the North
Atlantic Drift (also called the North Atlantic Current) into the northernmost
regions of the North Atlantic Ocean between Scotland and Iceland.
The southward-flowing currents are generally weaker than the Gulf Stream
and occur in the eastern lens of the North Atlantic Central Water or the
subtropical gyre. The circulation to the south on the southern rim of the
subtropical gyre is completed by the westward-flowing North Equatorial
Current, part of which flows into the Gulf of Mexico; the remaining part flows
northward as the Antilles Current. This subtropical gyre of warm North Atlantic
Central Water is the hub of the energy that drives the North Atlantic circulation.
It is principally forced by the overlying atmospheric circulation, which at these
latitudes is dominated by the clockwise circulation of a subtropical anticyclone.
This circulation is not steady and fluctuates in particular on its poleward side
where extratropical cyclones in the westerlies periodically make incursions into
the region. On the western side, hurricanes (during the period from May to
November) occasionally disturb the atmospheric circulation. Because of the
energy of the subtropical gyre and its associated currents, these short-term
fluctuations have little influence on it, however. The gyre obtains most of its
energy from the climatological wind distribution over periods of one or two
decades. This wind distribution drives a system of surface currents in the
uppermost 100 metres (about 330 feet) of the ocean.
Nonetheless, these currents are not simply a reflection of the surface wind
circulation as they are influenced by the Coriolis force. The wind-driven current
decays with depth, becoming negligible below 100 metres. The water in this
surface layer is transported to the right and perpendicular to the surface wind
stress because of the Coriolis force. Hence an eastward-directed wind on the
poleward side of the subtropical anticyclone would transport the surface layer of
the ocean to the south. On the equatorward side of the anticyclone the trade
winds would cause a contrary drift of the surface layer to the north and west.
Thus surface waters under the subtropical anticyclone are driven toward the mid-
latitudes at about 30° N. These surface waters, which are warmed by solar
heating and have a high salinity by virtue of the predominance of evaporation
over precipitation at these latitudes, then converge and are forced downward into
the deeper ocean.
Over many decades this process forms a deep lens of warm, saline North
Atlantic Central Water. The shape of the lens of water is distorted by other
dynamical effects, the principal one being the change in the vertical component
of the Coriolis force with latitude known as the beta effect. This effect involves
the displacement of the warm water lens toward the west, so that the deepest part
of the lens is situated to the north of the island of Bermuda rather than in the
central Atlantic Ocean. This warm lens of water plays an important role,
establishing as it does a horizontal pressure gradient force in and below the
wind-drift current. The sea level over the deepest part of the lens is about 1
metre (3.3 feet) higher than outside the lens. The Coriolis force in balance with
this horizontal pressure gradient force gives rise to a dynamically induced
geostrophic current, which occurs throughout the upper layer of warm water.
The strength of this geostrophic current is determined by the horizontal pressure
gradient through the slope in sea level. The slope in sea level across the Gulf
Stream has been measured by satellite radar altimeter to be 1 metre over a
horizontal distance of 100 km (about 60 miles), which is sufficient to cause a
surface geostrophic current of 1 metre per second at 43° N.
The large-scale circulation of the Gulf Stream system is, however, only one
aspect of a far more complex and richer structure of circulation. Embedded
within the mean flow is a variety of eddy structures that not only put kinetic
energy into circulation but also carry heat and other important properties, such as
nutrients for biological systems. The best known of these eddies are the Gulf
Stream rings, which develop in meanders of the current east of Cape Hatteras.
Though the eddies were mentioned as early as 1793 by Jonathan Williams, a
grandnephew of Benjamin Franklin, they were not systematically studied until
the early 1930s by the oceanographer Phil E. Church. Intensive research
programs were finally undertaken during the 1970s.
Gulf Stream rings have either warm or cold cores. The warm rings are
typically 100 to 300 km (about 60 to 186 miles) in diameter and have a
clockwise rotation. They consist of waters from the Gulf Stream and Sargasso
Sea and form when the meanders in the Gulf Stream pinch off on its continental
slope side. They move generally westward, flowing at the speed of the slope
waters, and are reabsorbed into the Gulf Stream at Cape Hatteras after a typical
lifetime of about six months. The cold core rings, composed of a mixture of Gulf
Stream and continental slope waters, are formed when the meanders pinch off to
the south of the Gulf Stream. They are a little larger than their warm-core
counterparts, characteristically having diameters of 200 to 300 km (about 124 to
186 miles) and an anticlockwise rotation. They move generally southwestward
into the Sargasso Sea and have lifetimes of one to two years. The cold-core rings
are usually more numerous than warm-core rings, typically 10 each year as
compared with five warm-core rings annually.

THE KUROSHIO CURRENT


The Kuroshio Current (Japanese: “Black Current”), which is also called the
Japan Current, is a strong surface oceanic current of the Pacific Ocean, the
northeasterly flowing continuation of the Pacific North Equatorial Current
between Luzon of the Philippines and the east coast of Japan. The temperature
and salinity of Kuroshio water are relatively high for the region, about 20 °C (68
°F) and 34.5 parts per thousand, respectively. Only about 400 metres (1,300 feet)
deep, the Kuroshio travels at rates ranging between 50 and 300 cm (20 and 120
inches) per second.
Flowing past Taiwan (Formosa) and the Ryukyu Islands, the current skirts
the east coast of Kyushu, where, during the summer, it branches west and then
northeast through the Korea Strait to parallel the west coast of Honshu in the Sea
of Japan as the Tsushima Current. In the vicinity of latitude 35° N (about central
Honshu), the bulk of the Kuroshio turns east to receive the southward-flowing
Oya Current. This flow, known as the Kuroshio Extension, eventually becomes
the North Pacific Current (also known as the North Pacific West Wind Drift).
Much of this current’s force is lost west of the Hawaiian Islands as a great south-
flowing eddy, the Kuroshio countercurrent, joins the Pacific North Equatorial
Current and directs the warm water back to the Philippine Sea. The remainder of
the original flow continues east to split off the coast of Canada and form the
Alaska and California currents. The Kuroshio exhibits distinct seasonal
fluctuations. It is strongest from May to August. Receding some in late summer
and autumn, it begins to increase from January to February only to weaken in
early spring. Similar to the Gulf Stream (Atlantic) in its creation and flow
patterns, the Kuroshio has an important warming effect upon the south and
southeast coastal regions of Japan as far north as Tokyo.
The existence of the Kuroshio was known to European geographers as early
as 1650, as shown by a map drawn by Bernhardus Varenius. It was also noted by
Captain J. King, a member of the British expedition under Captain James Cook
(1776–80). It is called Kuroshio (“Black Current”) because it appears a deeper
blue than does the sea through which it flows.

THE POLEWARD TRANSFER OF HEAT


A significant characteristic of the large-scale North Atlantic circulation is the
poleward transport of heat. Heat is transferred in a northward direction
throughout the North Atlantic. This heat is absorbed by the tropical waters of the
Pacific and Indian oceans, as well as of the Atlantic, and is then transferred to
the high latitudes, where it is finally given up to the atmosphere.
The mechanism for the heat transfer is principally by thermohaline
circulation rather than by wind-driven circulation. Circulation of the
thermohaline type involves a large-scale overturning of the ocean, with warm
and saline water in the upper 1,000 metres (about 330 feet) moving northward
and being cooled in the Labrador, Greenland, and Norwegian seas. The density
of the water in contact with the atmosphere is increased by surface cooling, and
the water subsequently sinks below the surface layer to the lowest depths of the
ocean. This water is mixed with the surrounding water masses by a variety of
processes to form North Atlantic Deep Water. The water moves slowly
southward as the lower limb of the thermohaline circulation. It is this
overturning circulation that is responsible for the warm winter climate of
northwestern Europe (notably the British Isles and Norway) rather than the
horizontal wind-driven circulation discussed above. The North Atlantic Drift,
which is an extension of the Gulf Stream system to the south, provides this
northward flow of warm and saline waters into the polar seas. This feature
makes the circulation of the North Atlantic Ocean uniquely different from that of
the Pacific Ocean, which has a less effective thermohaline circulation. Although
there is a northward transfer of heat in the North Pacific, the subtropical wind-
driven gyre in the upper ocean is mainly responsible for it. Thus the Kuroshio on
the western boundary of the North Pacific gyre is principally driven by the
surface wind circulation of the North Pacific.
Studies of the sediment cores obtained from the ocean floor have indicated
that the ocean surface temperature was as much as 10 °C (18 °F) cooler than
today in the northernmost region of the North Atlantic Ocean during the last
glacial maximum some 18,000 years ago. This difference in surface temperature
would indicate that the warm North Atlantic Drift was much reduced compared
to what it is at present, and hence the thermohaline circulation was considerably
weaker. In contrast, the Gulf Stream was probably more intense than it is today
and exhibited a large shift from its present path to an eastward flow at 40° N.

EL NIÑO/SOUTHERN OSCILLATION AND CLIMATIC


CHANGE
As was explained earlier, the oceans can moderate the climate of certain regions.
Not only do they affect such geographic variations, but they also influence
temporal changes in climate. The time scales of climate variability range from a
few years to millions of years and include the so-called ice age cycles that repeat
every 20,000 to 40,000 years, interrupted by interglacial periods of “optimum”
climate, such as the present. The climatic modulations that occur at shorter
scales include such periods as the Little Ice Age from the early 16th to the mid-
19th centuries, when the global average temperature was approximately 1 °C
(1.8 °F) lower than it is today. Several climate fluctuations on the scale of
decades have occurred in the 20th century, such as warming from 1910 to 1940,
cooling from 1940 to 1970, and the warming trend since 1970.
Although many of the mechanisms of climate change are understood, it is
usually difficult to pinpoint the specific causes. Scientists acknowledge that
climate can be affected by factors external to the land-ocean-atmosphere climate
system, such as variations in solar brightness, the shading effect of aerosols
injected into the atmosphere by volcanic activity, or the increased atmospheric
concentration of “greenhouse” gases (e.g., carbon dioxide, nitrous oxide,
methane, and chlorofluorocarbons) produced by human activities. However,
none of these factors explain the periodic variations observed during the 20th
century, which may simply be manifestations of the natural variability of
climate. The existence of natural variability at many time scales makes the
identification of causative factors such as human-induced warming more
difficult. Whether change is natural or caused, the oceans play a key role and
have a moderating effect on influencing factors.

THE EL NIÑO PHENOMENON


The shortest, or interannual, time scale relates to natural variations that are
perceived as years of unusual weather—e.g., excessive heat, drought, or
storminess. Such changes are so common in many regions that any given year is
about as likely to be considered as exceptional as typical. The best example of
the influence of the oceans on interannual climate anomalies is the occurrence of
El Niño conditions in the eastern Pacific Ocean at irregular intervals of about 3–
10 years. The stronger El Niño episodes of enhanced ocean temperatures (2–8
°C [3.6–14.4 °F] above normal) are typically accompanied by altered weather
patterns around the globe, such as droughts in Australia, northeastern Brazil, and
the highlands of southern Peru, excessive summer rainfall along the coast of
Ecuador and northern Peru, severe winter storminess along the coast of central
Chile, and unusual winter weather along the west coast of North America.
The effects of El Niño have been documented in Peru since the Spanish
conquest in 1525. The Spanish term “la corriente de El Niño” was introduced by
fishermen of the Peruvian port of Paita in the 19th century; it refers to a warm,
southward ocean current that temporarily displaces the normally cool,
northward-flowing Humboldt, or Peru, Current. (The name is a pious reference
to the Christ child, chosen because of the typical appearance of the
countercurrent during the Christmas season.) Under normal conditions, coastal
winds can induce upwelling of the lower-layer nutrients to the surface, where
they support an abundant ecosystem. During an El Niño event, however, the
upper layer thickens so that the upwelled water contains fewer nutrients, thus
contributing to a collapse of marine productivity.
By the end of the 19th century Peruvian geographers recognized that every
few years this countercurrent is more intense than normal, extends farther south,
and is associated with torrential rainfall over the otherwise dry northern desert.
The abnormal countercurrent also was observed to bring tropical debris, as well
as such flora and fauna as bananas and aquatic reptiles, from the coastal region
of Ecuador farther north. Increasingly during the 20th century, El Niño has come
to connote an exceptional year rather than the original annual event.
As Peruvians began to exploit the guano of marine birds for fertilizer in the
early 20th century, they noticed El Niño-related deteriorations in the normally
high marine productivity of the coast of Peru as manifested by large reductions
in the bird populations that depend on anchovies and sardines for sustenance.
The preoccupation with El Niño increased after mid-century, as the Peruvian
fishing industry rapidly expanded to exploit the anchovies directly. (Fish meal
produced from the anchovies was exported to industrialized nations as a feed
supplement for livestock.) By 1971 the Peruvian fishing fleet had become the
largest in its history; it had extracted very nearly 13 million metric tons (about
14.3 million tons) of anchovies in that year alone. Peru was catapulted into first
place among fishing nations, and scientists expressed serious concern that fish
stocks were being depleted beyond self-sustaining levels, even for the extremely
productive marine ecosystem of Peru. The strong El Niño of 1972–73 captured
world attention because of the drastic reduction in anchovy catches to a small
fraction of prior levels. The anchovy catch did not return to previous levels, and
the effects of plummeting fish meal exports reverberated throughout the world
commodity markets.
The upwelling process in the ocean along the coast of Peru. A
thermocline and a nutricline separate the warm, nutrient-deficient
upper layer from the cool, enriched layer below. Encyclopædia
Britannica, Inc.
Fishing for anchovies off the coast of Peru. Robert Harding Picture
Library

El Niño was only a curiosity to the scientific community in the first half of
the 20th century, thought to be geographically limited to the west coast of South
America. There was little data, mainly gathered coincidentally from foreign
oceanographic cruises, and it was generally believed that El Niño occurred when
the normally northward coastal winds off Peru, which cause the upwelling of
cool, nutrient-rich water along the coast, decreased, ceased, or reversed in
direction. When systematic and extensive oceanographic measurements were
made in the Pacific in 1957–58 as part of the International Geophysical Year, it
was found that El Niño had occurred during the same period and was also
associated with extensive warming over most of the Pacific equatorial zone.
Eventually tide-gauge and other measurements made throughout the tropical
Pacific showed that the coastal El Niño was but one manifestation of basinwide
ocean circulation changes that occur in response to a massive weakening of the
westward-blowing trade winds in the western and central equatorial Pacific and
not to localized wind anomalies along the Peru coast.

THE SOUTHERN OSCILLATION


The wind anomalies are a manifestation of an atmospheric counterpart to the
oceanic El Niño. At the turn of the century, the British climatologist Gilbert
Walker set out to determine the connections between the Asian monsoon and
other climatic fluctuations around the globe in an effort to predict unusual
monsoon years that bring drought and famine to the Asian sector. Unaware of
any connection to El Niño, he discovered a coherent interannual fluctuation of
atmospheric pressure over the tropical Indo-Pacific region, which he termed the
Southern Oscillation (SO). During years of reduced rainfall over northern
Australia and Indonesia, the pressure in that region (e.g., at what are now
Darwin and Jakarta) was anomalously high and wind patterns were altered.
Simultaneously, in the eastern South Pacific pressures were unusually low,
negatively correlated with those at Darwin and Jakarta. A Southern Oscillation
Index (SOI), based on pressure differences between the two regions (east minus
west), showed low, negative values at such times, which were termed the “low
phase” of the SO. During more normal “high-phase” years, the pressures were
low over Indonesia and high in the eastern Pacific, with high, positive values of
the SOI. In papers published during the 1920s and ’30s, Walker gave statistical
evidence for widespread climatic anomalies around the globe being associated
with the SO pressure “seesaw.”
In the 1950s, years after Walker’s investigations, it was noted that the low-
phase years of the SOI corresponded with periods of high ocean temperatures
along the Peruvian coast, but no physical connection between the SO and El
Niño was recognized until Jacob Bjerknes, in the early 1960s, tried to understand
the large geographic scale of the anomalies observed during the 1957–58 El
Niño event. Bjerknes, a meteorologist, formulated the first conceptual model of
the large-scale ocean-atmosphere interactions that occur during El Niño
episodes. His model has been refined through intensive research since the early
1970s.
During a year or two prior to an El Niño event (high-phase years of the SO),
the westward trade winds typically blow more intensely along the equator in the
equatorial Pacific, causing warm upper-ocean water to accumulate in a thickened
surface layer in the western Pacific where sea level rises. Meanwhile, the
stronger, upwelling-favourable winds in the eastern Pacific induce colder surface
water and lowered sea levels off South America. Toward the end of the year
preceding an El Niño, the area of intense tropical storm activity over Indonesia
migrates eastward toward the equatorial Pacific west of the International Date
Line (which corresponds in general to the 180th meridian of longitude), bringing
episodes of eastward wind reversals to that region of the ocean. These wind
bursts excite extremely long ocean waves, known as Kelvin waves
(imperceptible to an observer), that propagate eastward toward the coast of
South America, where they cause the upper ocean layer of relatively warm water
to thicken and sea level to rise.
The tropical storms of the western Pacific also occur in other years, though
less frequently, and produce similar Kelvin waves, but an El Niño event does not
result and the waves continue poleward along the coast toward Chile and
California, detectable only in tide-gauge measurements. Something else occurs
prior to an El Niño that is not fully understood: as the Kelvin waves travel
eastward along the equator, an anomalous eastward current carries warm western
Pacific water farther east, and the warm surface layer deepens in the central
equatorial Pacific (east of the international dateline). Additional surface warming
takes place as the upwelling-favourable winds bring warmer subsurface water to
the surface. (The subsurface water is warmer now, rather than cooler, because
the overlying layer of warmer water is now significantly deeper than before.)
The anomalous warming creates conditions favourable for the further migration
of the tropical storm centre toward the east, giving renewed vigour to eastward
winds, more Kelvin waves, and additional warming. Each increment of
anomalies in one medium (e.g., the ocean) induces further anomalies in the other
(the atmosphere) and vice versa, giving rise to an unstable growth of anomalies
through a process of positive feedbacks. During this time, the SO is found in its
low phase.
After several months of these unstable ocean-atmosphere interactions, the
entire equatorial zone becomes considerably warmer (2–5 °C [3.6–9 °F]) than
normal, and a sizable volume of warm upper ocean water is transported from the
western to the eastern Pacific. As a result, sea levels fall by 10–20 cm (4–8
inches) in the west and rise by larger amounts off the coast of South America,
where sea surface temperature anomalies may vary from 2 to 8 °C (3.6 to 14.4
°F) above normal. Anomalous conditions typically persist for 10–14 months
before returning to normal. The warming off South America occurs even though
the upwelling-favourable winds there continue unabated: the upwelled water is
warmer now, rather than cooler as before, and its associated nutrients are less
plentiful, thereby failing to sustain the marine ecosystem at its prior productive
levels.
The current focus of oceanographic research is on understanding the
circumstances leading to the demise of the El Niño event and the onset of
another such event several years later. The most widely held hypothesis is that a
second class of long equatorial ocean waves—Rossby waves with a shallow
surface layer—is generated by the El Niño and that they propagate westward to
the landmasses of Asia. There, the Rossby waves reflect off the Asian coast
eastward along the equator in the form of upwelling Kelvin waves, resulting in a
thinning of the upper ocean warm layer and a cooling of the ocean as the winds
bring deeper, cooler water to the surface. This process is thought to initiate one
to two years of colder-than-average conditions until Rossby waves of a contrary
sense (i.e., with a thickened surface layer) are again generated, functioning as a
switching mechanism, this time to start another El Niño sequence.
Another goal of scientists is to understand climate change on the scale of
centuries or longer and to make projections about the changes that will occur
within the next few generations. Yet, determinations of current climatic trends
from recent data are made difficult by natural variability at shorter time scales,
such as the El Niño phenomenon. Many scientists are attempting to understand
the mechanisms of change during an El Niño event from improved global
measurements so as to determine how the ocean-atmosphere engine operates at
longer time scales. Others are studying prehistoric records preserved in trees,
sediments, and fossil corals in an effort to reconstruct past variations, including
those like the El Niño. Their aim is to remove such short-term variations so as to
be able to make more accurate estimates of long-term trends.

ECONOMIC ASPECTS OF THE OCEANS


The sea is generally accepted by scientists as the place where life began on
Earth. Without the sea, life as it is known today could not exist. Among other
functions, it acts as a great heat reservoir, leveling the temperature extremes that
would otherwise prevail over Earth and expand the desert areas. The oceans
provide the least expensive form of transportation known, and the coasts serve as
a major recreational site. More importantly, the sea is a valuable source of food
and a potentially important source of energy and minerals, all of which are
required in ever-increasing quantities by industrialized and developing nations
alike.

MEDIUM FOR TRANSPORTATION AND COMMUNICATIONS


From the beginning of recorded history, people have used the sea as a means of
transporting themselves and their goods. The bulk of the tonnage of products
transported throughout the world today continues to be moved in ocean vessels.
The size of these vessels ranges from small boats capable of carrying a few tons
to bulk carriers (e.g., supertankers) capable of transporting more than 500,000
tons of oil. The cost of transporting goods on the ocean depends on the product,
the form of shipment, and the type of vessel. As the per capita consumption of
materials increases, the outlook for marine transportation is one of ever-
increasing tonnages and size of carrying vessels.
Since the laying of the transatlantic cable in the 19th century, the oceans
have served as a major means of communication between continents and islands.
Hundreds of seafloor cables connect many large centres of world population.
With the development of satellite communications, seafloor cables as a means of
communication have decreased somewhat in importance, but they will continue
to carry information for many decades to come.
In addition to communications, cable and pipes laid on the seafloor carry
electrical energy, oil, and other commodities in many parts of the world.

SOURCE OF FOOD AND WATER


Many millions of tons of edible fish and shellfish are taken from the oceans each
year. Yet, the food-producing potential of the sea is considerably greater than
this. The prevailing methods by which fish are taken from the oceans are
inefficient. This problem is compounded by the fact that only a handful of
varieties are targeted by commercial fishermen (anchovy, sardine, herring, cod,
mackerel, and pollack constitute more than half of the total annual catch).
Overfishing of certain varieties and areas along the continental shelves also has
resulted in declining numbers and therefore catches.
Several alternatives have been proposed to enhance productivity. One
possibility is to extract protein concentrates from all types of fish, promoting the
use of varieties formerly ignored for foodstuff. It has been estimated that this
procedure could produce a sustained yield of 2 billion tons of food annually for
the world’s populace. Another alternative would be to use protein derived from
algae cultivated on the continental shelf. A more viable course of action would
be to develop the continental shelf for fish-and kelp-farming. The Japanese have
instituted a substantial research program in this area. They have farmed oysters
in their oceanic bays for many years and more recently raised sea bream and
shrimp in protected environments. Commercial shrimp and oyster farms also
have been developed in the United States in the shallow waters of estuaries and
bays. Both commercial and research aquaculture projects have been conducted
to raise abalone, Maine lobster, salmon, and edible forms of seaweed under
controlled conditions. Limited sea farming has been practiced in France and
Italy, both of which raise mussels and certain other shellfish.

Workers clean the harvest at an oyster farm in Point Reyes Station,


California. Seafood production can be greatly enhanced by such
operations. Justin Sullivan/Getty Images

Various marine organisms have been artificially cultivated as sources for


therapeutic drugs. For example, during the early 1980s American researchers
isolated a substance called bryostatin 1 from the moss animal Burgula neritina;
this substance has been shown to slow the development of tumours.
The need for water for domestic and agricultural purposes grew steadily
throughout the latter half of the 20th century. As a result, increasing attention has
been given to desalting ocean waters and brackish waters in inland seas.
Throughout the world, more than 3,500 land-based desalination plants,
producing a total of more than 8 billion litres (in excess of 2 billion gallons) per
day, were in operation by the early 1990s. In general, the desalination plants are
located in areas where the population has outstripped the onshore water supply
and where high-cost desalinated water can be afforded. This situation tends to
arise in coastal-desert areas or on densely populated islands, because the cost of
pumping water through pipelines to interior areas would add prohibitively to the
basic cost at the sites of desalination. The United States operates a little more
than 30 percent of the world’s desalination facilities. Another 20 percent or so
are found in the Middle East, chiefly in Israel, Saudi Arabia, Kuwait, and the
United Arab Emirates. Together, these Middle Eastern plants produce roughly
two-thirds of the world’s desalinated water.
A population usually can afford to pay about 10 times as much for water for
domestic purposes as it does for agricultural water. Large-scale nuclear
desalination facilities promise to lower the cost of desalted water at the
desalination sites to a level that most industries and some agricultural enterprises
can afford. As yet, however, no such plant has been constructed.
At present desalination is accomplished primarily by distillation and
membrane processes. Distillation processes involve some form of evaporation
and subsequent condensation. Many of the largest commercial desalination
plants use multiple-effect distillation (i.e., multistage-flash distillation). Roughly
half of the world’s desalting facilities employ distillation processes and account
for approximately 75 percent of all desalinated water produced annually.
Membrane processes for desalting include reverse osmosis and
electrodialysis. Of the two, reverse osmosis is the more widely used, particularly
for desalting brackish waters from inland seas. In this method the natural process
of osmosis is reversed by applying pressure to brine that is in contact with an
osmotic membrane. The membrane impedes the passage of salt ions while
allowing fresh water to move through. In electrodialysis an electric potential is
used to drive positive and negative ions of the dissolved salts through
membranous filters, thereby sharply reducing the salt content of the water
between the filters.
In the future, it can be expected that the ocean will become an increasingly
important source of fresh water. If production and transportation costs can be
lowered sufficiently, it may be possible to produce fresh water to irrigate large
areas that border the oceans in many parts of the world.

ENERGY RESOURCES
There are several recognized techniques by which energy can be extracted from
the sea. The major problem in taking energy resources from the sea is that they
tend to be diffused over a large lateral area. A point-concentration energy source
is necessary if it is to be exploited economically.

TIDAL POWER GENERATION


Hydraulic turbine generator units are used to extract energy from ocean tides. As
of the late 1980s, operating tidal units included a 240-megawatt plant in France
on the estuary of the Rance River and several smaller installations such as, for
example, a 40,000-kilowatt pilot plant in Russia on the Barents Sea. By the early
21st century some of these technologies had become commercially available,
and a number of other countries, such as Scotland and South Korea, were
expanding their tidal power generation capacity.

OCEAN THERMAL ENERGY CONVERSION


Another more promising technology, known as ocean thermal energy conversion
(OTEC), makes use of the temperature differential between the warm surface
waters of the oceans, heated by solar radiation, and the deeper cold waters to
generate power in a conventional heat engine. The difference in temperature
between the surface and lower water layer can be as large as 50 °C (90 °F) over
vertical distances of as little as 90 metres (295 feet) in some ocean areas. To be
economically practical, the temperature differential should be at least 20 °C (36
°F) in the first 1,000 metres (3,300 feet) below the surface.

The La Rance River tidal power plant near Saint Malo, in western
France. Marchel Mochet/AFP/Getty Images

The OTEC concept was first proposed in the early 1880s by the French
engineer Jacques-Arsìne d’Arsonval. His idea called for a closed-cycle system, a
design that has been adapted for most present-day OTEC pilot plants. Such a
system employs a secondary working fluid (a refrigerant) such as ammonia. Heat
transferred from the warm surface ocean water causes the working fluid to
vaporize through a heat exchanger. The vapour then expands under moderate
pressures, turning a turbine connected to a generator and thereby producing
electricity. Cold seawater pumped up from the ocean depths to a second heat
exchanger provides a surface cool enough to cause the vapour to condense. The
working fluid remains within the closed system, vaporizing and reliquefying
continuously.
Some researchers have centred their attention on an open-cycle OTEC
system that employs water vapour as the working fluid and dispenses with the
use of a refrigerant. In this kind of system warm surface seawater is partially
vaporized as it is injected into a near vacuum. The resultant steam is expanded
through a low-pressure steam turbogenerator to produce electric power. Cold
seawater is used to condense the steam, and a vacuum pump maintains the
proper system pressure.
The prospects for the commercial application of OTEC technology seem
bright, particularly on islands and in developing nations in the tropical regions
where conditions are most favourable for OTEC plant operation. It has been
estimated that the tropical ocean waters absorb solar radiation equivalent in heat
content to that of about 170 billion barrels of oil each day. Removal of this much
heat from the ocean would not significantly alter its temperature, but would
permit the generation of about 10 million megawatts of electricity on a
continuous basis.

SOURCE OF MINERALS AND OTHER RAW MATERIALS


The ocean floor continues to be the site of petroleum and mineral extraction. Oil
and natural gas deposits occur in rock layers beneath the ocean floor, and several
countries and multinational corporations spend huge sums of money to extract
these resources annually. Minerals recovered from the ocean floor and marine
sediments below include phosphorite, from which phosphorus can be extracted,
calcareous oozes, which can be used to make cement, and precious metals.

PETROLEUM
In the mid-1950s the production of oil and gas from oceanic areas was
negligible. By the early 1980s about 14 million barrels per day, or about 25
percent of the world’s production, came from offshore wells, and the amount
continues to grow. More than 700 offshore drilling and production rigs were at
work by the early 21st century at more than 200 offshore locations throughout
the world, drilling, completing, and maintaining offshore oil wells.
It was once thought that only the continental-shelf areas contained potential
petroleum resources, but discoveries of oil deposits in deeper waters of the Gulf
of Mexico (about 3,000 to 4,000 metres [9,800 to 13,100 feet]) have changed
that view. It is now believed that the continental slopes and neighbouring
seafloor areas contain large oil deposits, thus enhancing potential petroleum
reserves of the ocean bottom.
Offshore drilling is not without its drawbacks, however. Not only is it
difficult and expensive to drill on the continental shelf and in deeper water, but
there is also the risk of accidental discharges of oil that can cause serious
damage to the environment and marine organisms. In spite of technological
advances, inadvertent leaks, sometimes on a large scale, continue to occur.

DEEPWATER DISASTER

The Deepwater Horizon oil rig, owned and operated by offshore-oil-drilling


company Transocean and leased by oil company BP, was situated in the
Macondo oil prospect in the Mississippi Canyon, a valley in the continental
shelf in the Gulf of Mexico. On the night of April 20, 2010, a surge of
natural gas blasted through a concrete cap recently installed to seal the well
for later use. The gas traveled up the rig’s riser to the platform, where it
ignited, killing 11 workers and injuring 17. The rig capsized and sank on
the morning of April 22—ironically, on Earth Day that year—rupturing the
riser, through which drilling mud was injected in order to counteract the
upward pressure of oil and natural gas. Without the opposing force, oil
began to discharge into the gulf.
Although BP attempted to activate the rig’s blowout preventer (BOP), a
fail-safe mechanism designed to close the channel through which oil was
drawn, the device malfunctioned. Several attempts to contain the oil were
attempted over the ensuing weeks, including capping the well and drilling
two relief wells, which were channels paralleling and eventually
intersecting the original well. These efforts had varying rates of success.
Meanwhile, the volume of oil escaping the damaged well formed a slick
extending over thousands of square miles of the Gulf of Mexico. To clean
oil from the open water, dispersants—substances that emulsified the oil,
thus allowing for easier metabolism by bacteria—were pumped directly
into the leak and applied aerially to the slick. Booms to corral portions of
the slick were deployed, and the contained oil was then siphoned off or
burned. As oil began to contaminate Louisiana beaches in May, it was
manually removed.
More difficult to clean were the state’s marshes and estuaries, where the
topography was knit together by delicate plant life. By June, oil and tar
balls had made landfall on the beaches of Mississippi, Florida, and
Alabama. Thousands of oil-plastered birds, mammals, and sea turtles were
transported to rehabilitation centres; countless fish and smaller sea life
perished. Various cleanup efforts were coordinated by the National
Response Team, a group of government agencies headed by the U.S. Coast
Guard and the Environmental Protection Agency (EPA).
With many residents of Gulf Coast states dependent on fishing and
tourism to support themselves, economic prospects for the region were dire.
More than a third of the gulf was closed to fishing due to fears of
contamination, and few travelers were willing to face the prospect of
petroleum-sullied beaches. Following demands by U.S. Pres. Barack
Obama, BP, which was already spending millions of dollars for cleanup
efforts daily, created a $20 billion compensation fund for those affected by
the spill.

MINERALS
The rivers of the world dump billions of tons of material into the oceans each
year. Seafloor springs and volcanic eruptions also add many millions of tons of
elements. Even the winds contribute solid materials to the oceans in appreciable
quantities. Most of these sediments rapidly settle to the seafloor in nearshore
areas, in some cases forming potentially valuable placer mineral deposits. The
dissolved load of the rivers, however, mixes with seawater and is gradually
dispersed over Earth’s total oceanic envelope. Because of the nature of the
minerals and their mode of formation, it is convenient to consider the occurrence
of ocean deposits in several environments—marine beaches, seawater,
continental shelves, sub-seafloor consolidated rocks, and the marine sediments
of the deep-sea floor. Minerals are mined from all of these environments except
for the deep-sea floor, which was only recently recognized as a repository for
mineral deposits of unbelievable extent and significant economic value.
Minerals that resist the chemical and mechanical processes of erosion in
nature and that possess a density greater than that of Earth’s common minerals
have a tendency to concentrate in gravity deposits known as placers. During the
Pleistocene glaciations, sea level was appreciably lowered as the ocean water
was transferred to the continental glaciers. Because of the cyclical nature of the
ice ages and the intervening warm periods, a series of beaches were formed in
nearshore areas both above and below present sea level. Also, when sea level
was lowered in past ages, the streams that today flow into the sea coursed much
further seaward, carrying placer minerals to be deposited in channels that are
now submerged. With geophysical exploration techniques, these channels and
beaches can be easily delineated, even though these features are totally covered
by Holocene sediments—i.e., those deposited during the past 10,000 years.
Sand and gravel are mined from a number of offshore locations around the
world, generally with hydraulic dredges. They are used primarily for
construction purposes or for beach replenishment or nearshore fills.
Sulfur, which is taken from salt domes in the Gulf of Mexico, is mined by a
process in which pressurized hot water is pumped into the sulfur-containing cap
of the dome, melting the sulfur and forcing it to the surface. Compressed air is
also used to pump sulfur to the surface; the still-molten sulfur is then conveyed
to the shore through insulated pipelines.
Of considerable interest are the seafloor phosphorite deposits on the coastal
shelves of many nations. The phosphorite off California occurs as nodules that
vary in shape from flat slabs a few metres across to small spherical forms termed
oolites. The nodules commonly are found as a single layer at the surface of
coarse-grained sediments. Phosphorite composition from the California offshore
area is surprisingly uniform and contains potentially economically attractive
amounts of phosphorus. Another type of phosphate deposit has been discovered
off the west coast of Mexico. It contains as much as 40 percent apatite (common
phosphate mineral), and some experts have speculated that up to 20 billion tons
of recoverable phosphate rock exist in the deposit.
Mineral deposits of enormous size and potential economic significance have
been discovered on the deep ocean floor. Minerals formed in the deep sea are
frequently found in high concentrations because there is relatively little clastic
material generated in these areas to dilute the chemical precipitates.
An estimated 1016 tons of calcareous oozes, formed by the deposition of
calcareous shells and skeletons of planktonic organisms, cover some 130 million
square km (about 50 million square miles) of the ocean floor. In a few instances,
these oozes, which occur within a few hundred kilometres of most nations
bordering the sea, are almost pure calcium carbonate; however, they often show
a composition similar to that of the limestones used in the manufacture of
portland cement.
Covering about 39 million square km (about 15 million square miles) of the
ocean floor in great bands across the northern and southern ends of the Pacific
Ocean and across the southern ends of the Indian and Atlantic oceans are other
oozes, consisting of the siliceous shells and skeletons of plankton animals and
plants. Normally these oozes could serve in most of the applications for which
diatomaceous earth is used, for fire and sound insulation, for lightweight
concrete formulations, as filters, and as soil conditioners.
An estimated 1016 tons of red clay covers about 104 million square km
(about 40 million square miles) of the ocean floor. Although compositional
analyses are not particularly exciting, red clay may possess some value as a raw
material in the clay-products industries, or it may serve as a source of metals in
the future. The average assay for alumina is about 15 percent, but red clays from
specific locations have assayed as high as 25 percent alumina; copper contents as
high as 0.20 percent also have been found. A few hundredths of a percent of
such metals as nickel and cobalt and a percent or so of manganese also are
generally present in a micronodular fraction of the clays and in all likelihood can
be separated and concentrated from the other materials by a screening process or
by some other physical method.
Underlying the hot brines in the Red Sea are basins containing metal-rich
sediments that potentially may prove to be of considerable significance. It has
been estimated that the largest of several such pools, the Atlantis II Deep,
contains several billion dollars worth of copper, zinc, silver, and gold in
relatively high grades. These pools lie in about 2,000 metres (about 6,600 feet)
of water midway between the Sudan and the Arabian Peninsula. Because of their
gellike nature, pumping these sediments to the surface may prove relatively
uncomplicated. These deposits are forming today under present geochemical
conditions and are similar in character to certain major ore deposits on land.
The discovery, in 1978, of polymetallic sulfides at the mid-ocean spreading
centres has aroused much interest. These sulfides include sediments enriched in
iron and manganese. Sites of rich deposits have been located at the Galápagos
spreading centre in the Gulf of California and at the East Pacific Rise.
From an economic standpoint, the most interesting oceanic sediments are
manganese nodules—small, black to brown, friable lumps found to be widely
distributed throughout the major oceans in the late 19th century by the
Challenger and Albatross expeditions.
Many theories have been proposed to account for the formation of
manganese nodules, the best probably being that the ocean is saturated at its
present state of acidity-alkalinity in iron and manganese. For this reason, these
elements precipitate as colloidal particles that gradually increase in size and filter
down to the seafloor. Colloids of manganese and iron oxides collect many metals
and tend to agglomerate as nodules at the seafloor rather than settle as particles
in the general sediments. An estimated 1.5 trillion tons of manganese nodules lie
on the Pacific Ocean floor alone. Averaging about 4 cm (1.6 inches) in diameter
and found in concentrations as high as 38,600 tons per square km (100,000 tons
per square mile), these manganese nodules contain as much as 2.5 percent
copper, 2.0 percent nickel, 0.2 percent cobalt, and 35 percent manganese. In
some deposits, the content of cobalt and manganese is as high as 2.5 percent and
50 percent, respectively. Such concentrations would be considered high-grade
ores if found on land, and, because of the large horizontal extent of the deposit,
they are a potential source of many important industrial metals.
Relatively simple mechanical cable bucket or hydraulic dredges with
submerged motors and pumps can effect the mining of the nodules at rates as
high as 10,000 to 15,000 tons per day, from depths as great as 6,000 metres
(19,700 feet). The estimated costs of mining and processing the nodules indicate
that copper, nickel, cobalt, and other metals can be economically produced from
this source.
The prospects of mining the manganese nodules and metal-rich sediments
have brought home the need to resolve long-standing legal problems relating to
the ownership of marine resources. During the 18th century the extent of the
territorial sea (and therefore rights) was established as 5.6 km (3 nautical miles)
from a nation’s shoreline. The area beyond the territorial sea, the so-called high
seas, was regarded as open to all nations. By the mid-1940s technological
advances had extended offshore oil drilling beyond the territorial limit. This
situation, together with the desire of various coastal nations to protect their
fishing grounds, eventually resulted in an attempt to codify international law
concerning territorial waters, ocean resources, and sea lanes. A 1982 treaty that
called for the enactment of the United Nations Law of the Sea Convention was
initially signed by 138 countries; some 30 other states, including the United
States, the United Kingdom, and West Germany, refused to sign, however. By
the end of the first decade of the 21st century, the number of signatories had
grown to 159. The treaty extended the territorial limit of each coastal country to
a distance of 12 nautical miles and granted it sovereign rights over natural
resources—living and nonliving—within an exclusive economic zone (EEZ) of
200 nautical miles. The countries that initially refused to sign the treaty objected
to its provisions governing seabed mining. The treaty declared the minerals on
the seafloor beneath the high seas the “common heritage of mankind” and
stipulated that their exploitation be directed by a global authority. While private
and national mining concerns are allowed to conduct exploration and set up
extraction operations, the question of seabed mineral ownership and mining
rights remains largely unresolved. This situation is viewed in some quarters as
the primary obstacle to full and effective utilization of seabed resources.

WASTE DISPOSAL AND OTHER RELATED ACTIONS


One of the least known but most significant uses of the sea is as an enormous
dump site. In the past, the oceans were able to assimilate the wastes of society
without noticeable adverse effects. However, industrialization and other
concomitant developments, along with sharp increases in global population,
have given rise to quantities and forms of waste that are now taxing the capacity
of the oceans to absorb them. Extensive marginal areas of the oceans have been
heavily polluted by human wastes ranging from the raw sewage of urban centres
to junked appliances and automobiles. Less apparent but more insidious forms of
pollution are toxic chemicals, nuclear wastes, and oily bilges pumped by
practically all vessels using petroleum for power.
Some other human activities are equally harmful to the marine environment.
Massive oil spills from tanker accidents, such as the 1989 mishap involving the
Exxon Valdez in Prince William Sound, Alaska, and the 2010 Deepwater
Horizon offshore drilling incident have not only disfigured innumerable beaches
and estuaries but caused widespread damage to wildlife as well. Large power
plants are generally located along coastlines to reduce the costs involved in
cooling their condensers by water-circulation systems. Although the whole of the
ocean never will be affected by the waste heat dissipated by these plants,
detrimental environmental effects can be caused in the immediate area of the
power-plant outfall. Herbicides and pesticides (especially the organochlorides
still used by some countries) reach the oceans via the wind and rivers and
contaminate marine organisms.
The fringes of the oceans—the beaches, lagoons, and bays—are the most
sensitive to human action, but the continued dumping of wastes, attended by
other abuses, will eventually affect the entire marine environment.
CHAPTER 5
OCEANOGRAPHY

Oceanography is the scientific discipline concerned with all aspects of the


world’s oceans and seas, including their physical and chemical properties, their
origin and geologic framework, and the forms of life that inhabit the marine
environment. Traditionally, oceanography has been divided into four separate
but related branches: physical oceanography, chemical oceanography, marine
geology, and marine ecology. Physical oceanography deals with the properties of
seawater (temperature, density, pressure, and so on), its movement (waves,
currents, and tides), and the interactions between the ocean waters and the
atmosphere. Chemical oceanography has to do with the composition of seawater
and the biogeochemical cycles that affect it. Marine geology focuses on the
structure, features, and evolution of the ocean basins. Marine ecology, also called
biological oceanography, involves the study of the plants and animals of the sea,
including life cycles and food production.
Oceanography is the sum of these several branches. Oceanographic research
entails the sampling of seawater and marine life for close study, the remote
sensing of oceanic processes with aircraft and Earth-orbiting satellites, and the
exploration of the seafloor by means of deep-sea drilling and seismic profiling of
the terrestrial crust below the ocean bottom. Greater knowledge of the world’s
oceans enables scientists to more accurately predict, for example, long-term
weather and climatic changes and also leads to more efficient exploitation of
Earth’s resources. Oceanography also is vital to understanding the effect of
pollutants on ocean waters and to the preservation of the quality of the oceans’
waters in the face of increasing human demands made on them.
Research diver deploying self-contained instrument package. Courtesy
of National Oceanic and Atmospheric Administration

ANCIENT OCEANOGRAPHY
The only substance known to the ancient philosophers in its solid, liquid, and
gaseous states, water is prominently featured in early theories about the origin
and operations of Earth. Thales of Miletus (c. 624–c. 545 BCE) is credited with a
belief that water is the essential substance of the planet, and Anaximander of
Miletus (c. 610–545 BCE) held that water was probably the source of life. In the
system proposed by Empedocles of Agrigentum (c. 490–430 BCE), water shared
the primacy Thales had given it with three other elements: fire, air, and earth.
The doctrine of the four earthly elements was later embodied in the universal
system of Aristotle and thereby influenced Western scientific thought until late
in the 17th century.

THE DISCOVERY OF THE HYDROLOGIC CYCLE


The idea that the waters of Earth undergo cyclical motions, changing from
seawater to vapour to precipitation and then flowing back to the ocean, is
probably older than any of the surviving texts that hint at or frame it explicitly.
The idea of the hydrological cycle developed independently in China as early
as the 4th century BCE and was explicitly stated in the Lüshi chunqiu (“The
Spring and Autumn [Annals] of Mr. Lü”), written in the 3rd century BCE. A
circulatory system of a different kind, involving movements of water on a large
scale within Earth, was envisioned by Plato (c. 428–348/347 BCE). In one of his
two explanations for the origin of rivers and springs, he described Earth as
perforated by passages connecting with Tartarus, a vast subterranean reservoir.
A coherent theory of precipitation is found in the writings of Aristotle.
Moisture on Earth is changed to airy vapour by heat from above. Because it is
the nature of heat to rise, the heat in the vapour carries it aloft. When the heat
begins to leave the vapour, the vapour turns to water. The formation of water
from air produces clouds. Heat remaining in the clouds is further opposed by the
cold inherent in the water and is driven away. The cold presses the particles of
the cloud closer together, restoring in them the true nature of the element water.
Water naturally moves downward, and so it falls from the cloud as raindrops.
Snow falls from clouds that have frozen.
In Aristotle’s system the four earthly elements were not stable but could
change into one another. If air can change to water in the sky, it should also be
able to change into water underground.

THE INITIAL STUDIES OF THE TIDES


The tides of the Mediterranean, being inconspicuous in most places, attracted
little notice from Greek and Roman naturalists. Poseidonius (135–50 BCE) first
correlated variations in the tides with phases of the Moon. By contrast, the tides
along the eastern shores of Asia generally have a considerable range and were
the subject of close observation and much speculation among the Chinese. In
particular, the tidal bore on the Qiantang River near Hangzhou attracted early
attention; with its front ranging up to 3.7 metres (12.1 feet) in height, this bore is
one of the largest in the world. As early as the 2nd century BCE, the Chinese had
recognized a connection between tides and tidal bores and the lunar cycle.

EARLY OCEANOGRAPHY
The groundwork for early ocean science can be attributed to Henry the
Navigator, the 15th-century Portuguese prince whose school of oceanography at
Sagres, Port., provided training for hundreds of seamen and advanced
substantially the fields of ship design, simulation, and instrumentation. In
addition, the idea that there is a circulatory system within Earth, by which
seawater is conveyed to mountaintops and there discharged, persisted until early
in the 18th century.

Areas reached by explorers under the sponsorship of Henry the


Navigator.
Two questions left unresolved by this theory were acknowledged even by its
advocates. How is seawater forced uphill? How is the salt lost in the process?

THE RISE OF SUBTERRANEAN WATER


René Descartes supposed that the seawater diffused through subterranean
channels into large caverns below the tops of mountains. The Jesuit philosopher
Athanasius Kircher in his Mundus subterraneus (1664; “Subterranean World”)
suggested that the tides pump seawater through hidden channels to points of
outlet at springs. To explain the rise of subterranean water beneath mountains,
the chemist Robert Plot appealed to the pressure of air, which forces water up the
insides of mountains. The idea of a great subterranean sea connecting with the
ocean and supplying it with water together with all springs and rivers was
resurrected in 1695 in John Woodward’s Essay Towards a Natural History of the
Earth and Terrestrial Bodies.
The French Huguenot Bernard Palissy maintained, to the contrary, that
rainfall is the sole source of rivers and springs. In his Discours admirables
(1580; Admirable Discourses) he described how rainwater falling on mountains
enters cracks in the ground and flows down along these until, diverted by some
obstruction, it flows out on the surface as springs. Palissy scorned the idea that
seawater courses in veins to the tops of mountains. For this to be true, sea level
would have to be higher than mountaintops—an impossibility. In his Discours
Palissy suggested that water would rise above the level at which it was first
encountered in a well provided the source of the groundwater came from a place
higher than the bottom of the well. This is an early reference to conditions
essential to the occurrence of artesian water, a popular subject among Italian
hydrologists of the 17th and 18th centuries.
In the latter part of the 17th century, Pierre Perrault and Edmé Mariotte
conducted hydrologic investigations in the basin of the Seine River that
established that the local annual precipitation was more than ample to account
for the annual runoff.

EVAPORATION FROM THE SEA


The question remained as to whether the amount of water evaporated from the
sea is sufficient to account for the precipitation that feeds the streams. The
English astronomer-mathematician Edmond Halley measured the rate of
evaporation from pans of water exposed to the air during hot summer days.
Assuming that this same rate would obtain for the Mediterranean, Halley
calculated that some 5.28 billion tons of water are evaporated from this sea
during a summer day. Assuming further that each of the nine major rivers
flowing into the Mediterranean has a daily discharge 10 times that of the
Thames, he calculated that a daily inflow of fresh water back into that sea would
be 1.827 billion tons, only slightly more than a third of the amount lost by
evaporation. Halley went on to explain what happens to the remainder. A part
falls back into the sea as rain before it reaches land. Another part is taken up by
plants.
In the course of the hydrologic cycle, Halley reasoned, the rivers constantly
bring salt into the sea in solution, but the salt is left behind when seawater
evaporates to replenish the streams with rainwater. Thus the sea must be growing
steadily saltier.

OCEANOGRAPHY IN THE 19TH CENTURY


Modern efforts to study systematically the physical and biological properties of
the Atlantic began in earnest during the 1800s and were notable for several
pioneering research expeditions, the results of which form the basis for present-
day scientific understanding of the oceans. While crude sampling and inaccurate
measurement techniques led to numerous misconceptions during this time, the
period also marked the advent of large-scale, multiyear scientific expeditions.
Incremental advances in both oceanographic theory and technique evolved
from early interdisciplinary studies of Atlantic processes. As early as 1770, the
American Benjamin Franklin published the first good map of the Gulf Stream,
based on data collected by Timothy Folger from the logs of transatlantic mail
ships. The work of the American naval officer Matthew Fontaine Maury in the
1840s and ’50s paved the way for generations of future researchers. His
exhaustive calculations of Atlantic winds and currents, as well as his early
seafloor maps, were the beginning of modern oceanography in the United States.
Map of the Gulf Stream drawn by Benjamin Franklin. Library of
Congress, Washington, D.C.

In 1807 Thomas Jefferson ordered the establishment of the U.S. Coast


Survey (later Coast and Geodetic Survey and now the National Ocean Survey).
Modeled after British and French agencies that had grown up in the 1700s, the
agency was charged with the responsibilities of hydrographic and geodetic
surveying, studies of tides, and preparation of charts. Beginning in 1842, the
U.S. Navy undertook expansive oceanographic operations through its office of
charts and instruments. Lieut. Matthew Fontaine Maury promoted international
cooperation in gathering meteorologic and hydrologic data at sea. In 1847 Maury
compiled the first wind and current charts for the North Atlantic and in 1854
issued the first depth map to 4,000 fathoms (7,300 metres). His Physical
Geography of the Sea (1855) is generally considered the first oceanographic
textbook.
The voyage of the Beagle (1831–36) is remembered for Darwin’s biological
and geologic contributions. From his observations in the South Pacific, Darwin
formulated a theory for the origin of coral reefs, which with minor changes has
stood the test of time. He viewed the fringing reefs, barrier reefs, and atolls as
successive stages in a developmental sequence. The volcanic islands around
which the reef-building organisms are attached slowly sink, but at the same time
the shallow-water organisms that form the reefs build their colonies upward so
as to remain in the sunlit layers of water. With submergence of the island, what
began as a fringing reef girdling a landmass at last becomes an atoll enclosing a
lagoon.
The advent of the telegraph and the dream of a transatlantic cable required
improved knowledge of bathymetry (measurement of ocean depth), currents,
topography, and bottom sediments. British and American naval ships were
instrumental in conducting hydrographic surveys in support of the early attempts
to lay a transatlantic cable; the first successful cable was laid in 1866. A
watershed expedition made by HMS Challenger in 1872–76 generated
thousands of observations in the Atlantic and other ocean basins, culminating in
the publication of 50 volumes of data on currents, water depth, temperature,
ocean sediments, and animal and plant species. Other important contributions of
the late 19th and early 20th centuries include those of Albert I of Monaco and of
many Scandinavians, including Bjørn Helland-Hansen and V. Walfrid Ekman.
Prince Albert financed a fleet of oceanographic vessels whose efforts led to
improved understanding of North Atlantic currents and to the discovery of many
new species of mid-depth fishes.
Laying telegraphic cables across the Atlantic called for investigations of the
configuration of the ocean floor, of the currents that sweep the bottom, and of the
benthonic animals that might damage the cables. The explorations of the British
ships Lightning and Porcupine in 1868 and 1869 turned up surprising
oceanographic information. Following closely upon these voyages, the
Challenger was authorized to determine “the conditions of the Deep Sea
throughout the Great Ocean Basins.”
The Challenger left port in December of 1872 and returned in May 1876,
after logging 127,600 km (68,890 nautical miles). Under the direction of Wyville
Thomson, Scottish professor of natural history, it occupied 350 stations scattered
over all oceans except the Arctic. The work involved in analyzing the
information gathered during the expedition was completed by Thomson’s
shipmate Sir John Murray, and the results filled 50 large volumes. Hundreds of
new species of marine organisms were described, including new forms of life
from deep waters. The temperature of water at the bottom of the oceans was
found to be nearly constant below the 2,000-fathom level, averaging about 2.5
°C (36.5 °F) in the North Atlantic and 2 °C (35 °F) in the North Pacific.
Soundings showed wide variations in depths of water, and from the dredgings of
the bottom came new types of sediment—red clay as well as oozes made
predominantly of the minute skeletons of foraminifera, radiolarians, or diatoms.
Improved charts of the principal surface currents were produced, and the precise
location of many oceanic islands was determined for the first time. Seventy-
seven samples of seawater were taken at different stations from depths ranging
downward to about 1.5 km (0.9 mile). The German-born chemist Wilhelm
Dittmar conducted quantitative determinations of the seven major constituents
(other than the hydrogen and oxygen of the water itself)—namely, sodium,
calcium, magnesium, potassium, chloride, bromide, and sulfate. Surprisingly, the
percentages of these components turned out to be nearly the same in all samples.
Efforts to analyze the rise and fall of the tides in mathematical terms
reflecting the relative and constantly changing positions of Earth, Moon, and
Sun, and thus to predict the tides at particular localities, has never been entirely
successful because of local variations in configuration of shore and seafloor.
Nevertheless, harmonic tidal analysis gives essential first approximations that
are essential to tidal prediction. In 1884 a mechanical analog tidal prediction
device was invented by William Ferrel of the U.S. Coast and Geodetic Survey,
and improved models were used until 1965, when the work of the analog
machines was taken over by electronic computers.

OCEANOGRAPHY IN THE 20TH CENTURY


The science of oceanography in the 20th century saw advancements in the
understanding of ocean circulation, the measurement of ocean depth
(bathymetry), the recognition of the limits of Earth’s water resources, and the
continued development of the ocean’s economic resources. New innovations
included the development of desalinization techniques, tidal power, and new
methods to extract mineral, oil, and natural gas resources.
A few critical developments early in the century set the stage for later
oceanographic investigation. The disaster in 1912 of the Titanic catalyzed
research efforts concerned with iceberg flows and current patterns in the North
Atlantic, accelerated the development of both radio and sonar, and led to the
establishment of the International Ice Patrol. In the field of marine
communications, the Italian Guglielmo Marconi was demonstrating his new
invention—wireless radio—in Europe and the United States during this period,
having used it in 1899 to report from sea the results of the America’s Cup yacht
races. In 1925–27 a series of scientific voyages by the research vessel Meteor
established Germany as a leader in marine research. Operating in the waters of
the South Atlantic, the Meteor traversed the basin 14 times, mapping the seafloor
by means of sonar and measuring salinity and temperature distributions at
various depths.

OCEAN CIRCULATION, CURRENTS, AND WAVES


Results of many investigations suggest that the forces that drive the ocean
currents originate at the interface between water and air. The direct transfer of
momentum from the atmosphere to the sea is doubtless the most important
driving force for currents in the upper parts of the ocean. Next in importance are
differential heating, evaporation, and precipitation across the air-sea boundary,
altering the density of seawater and thus initiating movement of water masses
with different densities. Studies of the properties and motion of water at depth
have shown that strong currents also exist in the deep sea and that distinct types
of water travel far from their geographic sources. For example, the highly saline
water of the Mediterranean that flows through the Strait of Gibraltar has been
traced over a large part of the Atlantic, where it forms a deepwater stratum that
is circulated far beyond that ocean in currents around Antarctica.
Improvements in devices for determining the motion of seawater in three
dimensions have led to the discovery of new currents and to the disclosure of
unexpected complexities in the circulation of the oceans generally. In 1951 a
huge countercurrent moving eastward across the Pacific was found below depths
as shallow as 20 metres (about 66 feet), and in the following year an analogous
equatorial undercurrent was discovered in the Atlantic. In 1957 a deep
countercurrent was detected beneath the Gulf Stream with the aid of subsurface
floats emitting acoustic signals.
Since the 1970s Earth-orbiting satellites have yielded much information on
the temperature distribution and thermal energy of ocean currents such as the
Gulf Stream. Chemical analyses from Geosecs makes possible the determination
of circulation paths, speeds, and mixing rates of ocean currents.
Surface waves of the ocean are also exceedingly complex, at most places and
times reflecting the coexistence and interferences of several independent wave
systems. During World War II, interest in forecasting wave characteristics was
stimulated by the need for this critical information in the planning of amphibious
operations. The oceanographers H.U. Sverdrup and Walter Heinrich Munk
combined theory and empirical relationships in developing a method of
forecasting “significant wave height”—the average height of the highest third of
the waves in a wave train. Subsequently this method was improved to permit
wave forecasters to predict optimal routes for mariners.
A scientist reviews data at the Tsunami Early Warning Centre in
Hyderabad, India. Tsunamis are the most destructive of ocean waves.
AFP/Getty Images

Forecasting of the most destructive of all waves, tsunamis caused by


submarine quakes and volcanic eruptions, is another recent development. Soon
after 159 persons were killed in Hawaii by the tsunami of 1946, the U.S. Coast
and Geodetic Survey established a seismic sea-wave warning system. Using a
seismic network to locate epicentres of submarine quakes, the installation
predicts the arrival of tsunamis at points around the Pacific basin often hours
before the arrival of the waves.

OCEAN BATHYMETRY
Modern bathymetric charts show that about 20 percent of the surfaces of the
continents are submerged to form continental shelves. Altogether the shelves
form an area about the size of Africa. Continental slopes, which slant down from
the outer edges of the shelves to the abyssal plains of the seafloor, are nearly
everywhere furrowed by submarine canyons. The depths to which these canyons
have been cut below sea level seem to rule out the possibility that they are
drowned valleys cut by ordinary streams. More likely, the canyons were eroded
by turbidity currents, dense mixtures of mud and water that originate as
mudslides in the heads of the canyons and pour down their bottoms.
Profiling of the Pacific basin prior to and during World War II resulted in the
discovery of hundreds of isolated eminences rising 1,000 or more metres (about
3,300 feet) above the floor. Of particular interest were seamounts in the shape of
truncated cones, whose flat tops rise to between 1.6 km (1 mile) and a few
hundred metres below the surface. Harry H. Hess interpreted the flat-topped
seamounts (guyots) as volcanic mountains planed off by action of waves before
they subsided to their present depths. Subsequent drilling in guyots west of
Hawaii confirmed this view; samples of rocks from the tops contained fossils of
Cretaceous age representing reef-building organisms of the kind that inhabit
shallow water.

BATHYMETRY

Bathymetry is the measurement of ocean depth. The earliest technique


involved lowering a heavy rope or cable of known length over the side of a
ship, then measuring the amount needed to reach the bottom. Tedious and
frequently inaccurate, this method yielded the depth at only a single point
rather than a continuous measurement; inaccuracies arose because the rope
did not necessarily travel straight to the bottom but instead might be
deflected by subsurface currents or movements of the vessel.
A more satisfactory approach, though not without problems, is echo
sounding, widely used today, in which a sound pulse travels from the vessel
to the ocean floor, is reflected, and returns. By calculations involving the
time elapsed between generation of the pulse and its return and the speed of
sound in water, a continuous record of seafloor topography can be made.
Most echo sounders perform these calculations mechanically, producing a
graphic record in the form of a paper chart. Misleading reflections caused
by the presence of undersea canyons or mountains plus variations in the
speed of sound through water caused by differences in temperature, depth,
and salinity limit the accuracy of echo sounding, though these problems can
be met somewhat by crossing and recrossing the same area. Sonar has also
been employed in bathymetric studies, as have underwater cameras.

WATER RESOURCES AND SEAWATER CHEMISTRY


Quantitative studies of the distribution of water have revealed that an
astonishingly small part of Earth’s water is contained in lakes and rivers. Ninety-
seven percent of all the water is in the oceans, and, of the fresh water
constituting the remainder, three-fourths is locked up in glacial ice and most of
the rest is in the ground. Approximate figures are also now available for the
amounts of water involved in the different stages of the hydrologic cycle. Of the
859 mm (34 inches) of annual global precipitation, 23 percent falls on the lands;
but only about a third of the precipitation on the lands runs directly back to the
sea, the remainder being recycled through the atmosphere by evaporation and
transpiration.
Subsurface groundwater accumulates by infiltration of rainwater into soil
and bedrock. Some may run off into rivers and lakes, and some may reemerge as
springs or aquifers. Advanced techniques are used extensively in groundwater
studies nowadays. The rate of groundwater flow, for example, can be calculated
from the breakdown of radioactive carbon-14 by measuring the time it takes for
rainwater to pass through the ground, while numerical modeling is used to study
heat and mass transfer in groundwater. High-precision equipment is used for
measuring down-hole temperature, pressure, flow rate, and water level.
Groundwater hydrology is important in studies of fractured reservoirs,
subsidence resulting from fluid withdrawal, geothermal resource exploration,
radioactive waste disposal, and aquifer thermal-energy storage.
Chemical analyses of trace elements and isotopes of seawater are conducted
as part of the Geochemical Ocean Sections (Geosecs) program. Of the 92
naturally occurring elements, nearly 80 have been detected in seawater or in the
organisms that inhabit it, and it is thought to be only a matter of time until traces
of the others are detected. Contrary to the idea widely circulated in the older
literature of oceanography, that the relative proportions of the oceans’ dissolved
constituents are constant, investigations since 1962 have revealed statistically
significant variations in the ratios of calcium and strontium to chlorinity. The
role of organisms as influences on the composition of seawater has become
better understood with advances in marine biology. It is now known that plants
and animals may collect certain elements to concentrations as much as 100,000
times their normal amounts in seawater. Abnormally high concentrations of
beryllium, scandium, chromium, and iodine have been found in algae; of copper
and arsenic in both the soft and skeletal parts of invertebrate animals; and of
zirconium and cerium in plankton.

DESALINIZATION, TIDAL POWER, AND MINERALS FROM THE SEA


For ages a source of food and common salt, the sea is increasingly becoming a
source of water, chemicals, and energy. In 1967 Key West, Fla., became the first
U.S. city to be supplied solely by water from the sea, drawing its supplies from a
plant that produces more than 7,570,000 litres (about 2 million gallons) of
refined water daily.
Many ambitious schemes for using tidal power have been devised. The first
major hydrographic project of this kind was not completed until 1967, when a
dam and electrical generating equipment were installed across the Rance River
in Brittany.
The seafloor and the strata below the continental shelves are also sources of
mineral wealth. Magnesia was extracted from the Mediterranean in the late 19th
century; at present nearly all the magnesium metal used in the United States is
mined from the sea at Freeport, Texas. Concretions of manganese oxide,
evidently formed in the process of subaqueous weathering of volcanic rocks,
have been found in dense concentrations with a total abundance of 1011 tons. In
addition to the manganese, these concretions contain copper, nickel, cobalt, zinc,
and molybdenum. To date, oil and gas have been the most valuable products to
be produced from beneath the sea.

MARINE GEOLOGY

Marine geology, which is also called geologic oceanography, is the


scientific discipline that is concerned with all geological aspects of the
continental shelves and slopes and the ocean basins. In practice, the
principal focus of marine geology has been on marine sedimentation and on
the interpretation of the many bottom samples that have been obtained
through the years. The advent of the concept of seafloor spreading in the
1960s, however, broadened the scope of marine geology considerably.
Many investigations of midoceanic ridges, remanent magnetism of rocks on
the seafloor, geochemical analyses of deep brine pools, and of seafloor
spreading and continental drift may be considered within the general realm
of marine geology.

UNDERSEA EXPLORATION
Undersea exploration is the investigation and description of the ocean waters and
the seafloor and of layers of rock beneath. The study of the physical and
chemical properties of seawater, the various forms of life in the sea, and the
geological and geophysical features of Earth’s crust are topics typically included
within the scope of this endeavour. Researchers in the field define and measure
such properties; prepare maps in order to identify patterns; and utilize these
maps, measurements, and theoretical models to achieve a better grasp of how
Earth works as a whole. This knowledge enables scientists to predict, for
example, long-term weather and climatic changes and leads to more efficient
exploration and exploitation of Earth’s resources, which in turn result in better
management of the environment in general.
The multidisciplinary expedition of the British ship Challenger in 1872–76
was the first major undersea survey. Although its main goal was to search for
deep-sea life by means of net tows and dredging, the findings of its physical and
chemical studies expanded scientific knowledge of temperature and salinity
distribution of the open seas. Moreover, depth measurements by wire soundings
were carried out all over the globe during the expedition.
Since the time of the Challenger voyage, scientists have learned much about
the mechanics of the ocean, what it contains, and what lies below its surface.
Investigators have produced global maps showing the distribution of surface
winds as well as of heat and rainfall, which all work together to drive the ocean
in its unceasing motion. They have discovered that storms at the surface can
penetrate deep into the ocean and, in fact, cause deep-sea sediments to be rippled
and moved. Recent studies also have revealed that storms called eddies occur
within the ocean itself and that such a climatic anomaly as El Niño is caused by
an interaction of the ocean and the atmosphere.
Other investigations have shown that the ocean absorbs large amounts of
carbon dioxide and hence plays a major role in delaying its buildup in the
atmosphere. Without the moderating effect of the ocean, the steadily increasing
input of carbon dioxide into the atmosphere (due to the extensive burning of
coal, oil, and natural gas) would result in the rapid onset of the so-called
greenhouse effect—i.e., a warming of Earth caused by the absorption and
reradiating of infrared energy to the terrestrial surface by carbon dioxide and
water vapour in the air.
The field of marine biology has benefitted from the development of new
sampling methods. Among these, broad ranging acoustical techniques have
revealed diverse fish populations and their distribution, while direct, close up
observation made possible by deep-sea submersibles has resulted in the
discovery of unusual (and unexpected) species and phenomena.
In the area of geology, undersea exploration of the topography of the seafloor
and its gravitational and magnetic properties has led to the recognition of global
patterns of continental plate motion. These patterns form the basis of the concept
of plate tectonics, which synthesized earlier hypotheses of continental drift and
seafloor spreading. As noted earlier, this concept not only revolutionized
scientific understanding of Earth’s dynamic features (e.g., seismic activity,
mountain-building, and volcanism) but also yielded discoveries of economic and
political impact. Earth scientists found that the midocean centres of seafloor
spreading also are sites of important metal deposits. The hydrothermal
circulations associated with these centres produce sizable accumulations of
metals important to the world economy, including zinc, copper, lead, silver, and
gold. Rich deposits of manganese, cobalt, nickel, and other commercially
valuable metals have been found in nodules distributed over the entire ocean
floor. The latter discovery proved to be a major factor in the establishment of the
Convention of the Law of the Sea (1982), which calls for the sharing of these
resources among developed and developing nations alike. Exploitation of these
findings awaits only the introduction of commercially viable techniques for
deep-sea mining and transportation.

BASIC ELEMENTS OF UNDERSEA EXPLORATION


Undersea exploration relies on accurate navigation and the use of scientific
platforms that operate far from land. Although aircraft, ships, and other vehicles
may be used to transport researchers to and from study sites, they can also
deploy sensors, tools, and other types of scientific equipment. Some aircraft and
ships can even serve as mobile laboratories. It should be noted that exploration
of any kind is useful only when the location of the discoveries can be noted
precisely. Thus, navigation has always been a key to undersea exploration.

PLATFORMS
Undersea exploration of any kind must be conducted from platforms, in most
cases, ships, buoys, aircraft, or satellites. Typical oceanographic vessels capable
of carrying out a full complement of underwater exploratory activities range in
size from about 50 to 150 metres (about 160 to 500 feet). They support scientific
crews of 16 to 50 persons and generally permit a full spectrum of
interdisciplinary studies. One example of a research vessel of this kind is the
Melville, operated by the Scripps Institution of Oceanography. It has a
displacement of 2,075 tons and can carry 25 scientists in addition to 25 crew
members. It is powered by a dual cycloidal propulsion system, which provides
remarkable manoeuvrability.
The JOIDES Resolution, operated by Texas A & M University for the Joint
Oceanographic Institutions for Deep Earth Sampling, represents a major advance
in research vessels. A converted commercial drill ship, it measures 145 metres
(475 feet) in length, has a displacement of 18,600 tons, and is equipped with a
derrick that extends 62 metres (about 200 feet) above the waterline. A computer-
controlled dynamic positioning system enables the ship to remain over a specific
location while drilling in water to depths as great as 8,300 metres (about 27,200
feet). The drilling system of the ship is designed to collect cores from below the
ocean floor; it can handle 9,200 metres (30,200 feet) of drill pipe. The vessel
thus can sample most of the ocean floor, including the bottoms of deep ocean
basins and trenches. The JOIDES Resolution has other notable capabilities. It
can operate in waves as high as 8 metres (about 26 feet), winds up to 23 metres
(75 feet) per second, and currents as strong as 1.3 metres (4.3 feet) per second. It
has been outfitted for use in ice so that it can conduct drilling operations in high
latitudes. The ship can accommodate 50 scientists as well as the crew and
drilling team, and its geophysical laboratories total nearly 930 square metres
(about 10,000 square feet).

JOIDES Resolution, a deep-sea drilling vessel that uses a computer-


controlled, acoustic dynamic positioning system to maintain location
over the drilling site. The derrick is visible amidships. Courtesy of
Ocean Drilling Program, Texas A & M University

Other specialized vessels include the deep submergence research vehicle


known as Alvin, which can carry a pilot and two scientific observers to a depth
of 4,000 metres (about 13,100 feet). The manoeuvrability of the Alvin was
pivotal to the discoveries of the mineral deposits at the midocean seafloor
spreading centres and of previously unknown biological communities living at
those sites. Another versatile vessel is the Floating Instrument Platform (FLIP).
It is a long narrow platform that is towed in a horizontal position to a research
site. Once on location, the ballast tanks are flooded to flip the ship to a vertical
position. Only 17 metres (56 feet) of the ship extend above the waterline, with
the remaining 92 metres (about 300 feet) completely submerged. The rise and
fall of the waves cause a very small change in the displacement, resulting in a
high degree of stability.
New ship designs that promise even greater stability and ease of use include
that of the Small Waterplane Area Twin Hull (SWATH) variety. This design type
requires the use of twin submerged, streamlined hulls to support a structure that
rides above the water surface. The deck shape is entirely unconstrained by the
hull shape, as is the case for conventional surface vessels. Ship motion is greatly
reduced because of the depth of the submerged hulls. For a given displacement, a
SWATH-type vessel can provide twice the amount of deck space that a single-
hull ship can, with only 10 percent of the motion of the single-hull design type.
In addition, a large centre opening, or well, can be used to display and recover
instruments.

NAVIGATION
There are various ways by which the position of a vessel at sea can be
determined. In cases where external references such as stars or radio and satellite
beacons are unavailable or undetectable, inertial navigation, which relies on a
stable gyroscope for determining position, is commonly employed. It is far more
accurate than the long-used technique of dead reckoning, which is dependent on
a knowledge of the ship’s original position and the effects of the winds and
ocean currents on the vessel.
Another modern position-fixing method is all-weather, long-range radio
navigation. It was introduced during World War II as Loran (long-range
navigation) A, a system that determines position by measuring the difference in
the time of reception of synchronized pulses from widely spaced transmitting
stations. The latest version of this system, Loran C, uses low-frequency
transmissions and derives its high degree of accuracy from precise time-
difference measurements of the pulsed signals and the inherent stability of signal
propagation. Users of Loran C are able to identify a position with an accuracy of
0.4 km (0.25 mile) and a repeatability of 15 metres (about 50 feet) at a distance
of up to about 2,220 km (about 1,400 miles) from the reference stations. The
Loran C system covers heavily travelled regions in the North Pacific and North
Atlantic oceans, parts of the Indian Ocean, and the Mediterranean Sea.

DEAD RECKONING

Dead reckoning is the determination without the aid of celestial navigation


of the position of a ship or aircraft from the record of the courses sailed or
flown, the distance made (which can be estimated from velocity), the
known starting point, and the known or estimated drift.
Some marine navigators differentiate between the dead-reckoning
position, for which they use the course steered and their estimated speed
through the water, and the estimated position, which is the dead-reckoning
position corrected for effects of current, wind, and other factors. Because
the uncertainty of dead reckoning increases over time and maybe over
distance, celestial observations are taken intermittently to determine a more
reliable position (called a fix), from which a new dead reckoning is begun.
Dead reckoning is also embedded in Kalman filtering techniques, which
mathematically combine a sequence of navigation solutions to obtain the
best estimate of the navigator’s current position, velocity, attitude angles,
and so forth.
A number of devices used for the determination of dead reckoning—
such as a plotter (a protractor attached to a straightedge) and computing
charts, now chiefly used by operators of smaller vehicles—have been
replaced in most larger aircraft and military vessels by one or more dead-
reckoning computers, which input direction and speed (wind velocity can
be manually inserted). Some of these computers include an inertial
guidance system or have a unit that measures Doppler effects, and some can
be programmed to pick up signals from electronic or optical sensing units.
The use of more than one such device tends to increase reliability.

Satellite navigation has proved to be the most accurate method of locating


geographical position. A polar-orbiting satellite system called Transit was
established in the early 1960s by the United States to provide global coverage
for ships at sea. In this system, a vessel pinpoints its position relative to a set of
satellites whose orbits are known by measuring the Doppler shift of a received
signal—i.e., the change in the frequency of the received signal from that of the
transmitted signal. The Transit system suffers from one major drawback.
Because of the limited number of system satellites, the frequency with which
position determinations can be made each day is relatively low, particularly in
the tropics. The system is being improved to provide nearly continuous
positioning capability at sea. This expanded version, the Global Positioning
System (GPS), is to have 18 satellites, six in each of three orbital planes spaced
120° apart. The GPS is designed to provide fixes anywhere on Earth to an
accuracy of 20 metres (about 66 feet) and a relative accuracy 10 times greater.

METHODOLOGY AND INSTRUMENTATION


The physical, chemical, and biological study of the oceans requires the
development of sampling techniques and equipment tailored to the demands of
marine environments. Water samplers and chemical analyzers are typically used
to study the chemistry of seawater, and acoustic and other types of remote
sensors are often used for bathymetry. Although remote-sensing equipment is
also used to study marine life and the geologic formations of ocean basins,
sometimes researchers must collect samples with nets, dredges, and other tools
in order to bring these objects of study to the surface for closer inspection.

WATER SAMPLING FOR TEMPERATURE AND SALINITY


The temperature, chemical environment, and movement and mixing of seawater
are fundamental to understanding the physical, chemical, and biological features
of the ocean and the geology of the ocean floor. Traditionally, oceanographers
have collected seawater by means of specially adapted water-sampling bottles.
The most universal water sampler used today, the Nansen bottle, is a
modification of a type developed in the latter part of the 19th century by the
Norwegian Arctic explorer and oceanographer Fridtjof Nansen. It is a metal
sampler equipped with special closing valves that are actuated when the bottle,
attached by one end to a wire that carries it to the desired depth, rotates about
that end. A mercury thermometer fastened to the bottle records the temperature
at the specified depth. The design of the device is such that, when it is inverted,
its mercury column breaks. The amount of mercury remaining in the graduated
capillary portion of the thermometer indicates the temperature at the point of
inversion. This type of reversing thermometer and the Nansen bottle are
extensively used by oceanographers because of their accuracy and dependability
in a harsh environment.
The temperature and salinity of the ocean have been mapped with data
gathered by many ships over many years. This information is used for tracing
heat and water movement and mixing, as well as for making density
measurements, which are employed in calculating ocean currents. It was noted
as early as the Challenger Expedition that the salt dissolved in seawater has
remarkably constant major constituents. As a consequence, it is possible to map
water density patterns within the sea with measurements of only the water
temperature and one major property of the sea salt (e.g., the chloride ion content
or the electrical conductivity) to arrive at an accurate estimate of the density of a
given sample.
Standard laboratory techniques such as titration are routinely used at sea for
determining chlorinity. Chlorinity can be briefly defined as the number of grams
of chlorine, bromine, and iodine contained in 1 kg (2.2 pounds) of seawater,
assuming that the bromine and iodine are replaced by chlorine. Salinity is the
total weight of dissolved solids, in grams, found in 1 kg of seawater and may be
determined from the concentration of chlorinity because of the constancy of
major constituents. In the traditional technique, a solution of silver nitrate of a
known strength is added to a sample of seawater to produce the same reaction as
with “standard” seawater. The difference in the amounts added gives the degree
of chlorinity. To ensure worldwide uniformity in chlorinity and salinity
determinations, the International Council for the Exploration of the Sea prepared
a universal reference, Eau de Mer Normale (“Standard Seawater”), in 1902. A
new primary standard, prepared in 1937 and having a chlorinity of 19.381 parts
per 1,000, is used to determine the chlorinities of all batches of standard
seawater. It also is utilized to calibrate electrical conductivity measurements.
Accurate and continuous measurements of temperature as it changes with
depth are required for understanding how the ocean moves and mixes heat. To
provide the necessary detail, temperature profilers had to be developed; then,
with the introduction of reliable conductivity sensors, salinity profilers were
added. An instrument called the bathythermograph (BT), which has been used
since the early 1940s to obtain a graphic record of water temperature at various
depths, can be lowered from a ship while it is moving at reduced speed. In this
instrument a depth element (pressure-operated bellows) drives a slide of smoked
glass or metal at right angles to a stylus. Actuated by a thermal element (liquid-
filled bourdon tube) that expands and contracts in response to changes in
temperature, the stylus scribes a continuous record of temperature and depth.
An expendable bathythermograph (XBT) was developed during the 1970s
and has come into increasingly wider use. Unlike the BT, this instrument
requires an electrical system aboard the research platform. It detects temperature
variations by means of a thermistor (an electrical resistance element made of a
semiconductor material) and depends on a known fall rate for depth
determination. The sensor unit of the XBT is connected to the research platform
by a leak-proof, insulated two-conductor cable. This cable is wound around a
pair of large spools in an arrangement resembling that of a fisherman’s spinning
reel. In operation, the cable is unwound from each of the spools in a direction
that is parallel to the axis of the respective spool. As a result, the cable unwinds
from both the platform—either a ship or an airplane—and the sensor unit
simultaneously but independently. Because of this double-spool arrangement, the
sensor unit can free-fall from wherever it hits the sea surface and is completely
unaffected by the direction or speed of the craft from which it was deployed.
One of the principal reasons why the XBT has proved so useful is that it can
provide a record of considerable depth even when it is deployed from a ship
moving at full speed.
Until the late 1950s, salinity was universally determined by titration. Since
then, shipboard electrical conductivity systems have become widely used.
Salinity-Temperature-Depth (STD) and the more recent Conductivity-
Temperature-Depth (CTD) systems have greatly improved on-site hydrographic
sampling methods. They have enabled oceanographers to learn much about
small-scale temperature and salinity distributions.
The most recent version of the CTD systems features rapid-response
conductivity and temperature sensors. The conductivity sensor consists of a tiny
cell with four platinum electrodes. This type of conductivity cell virtually
eliminates errors resulting from the polarization that occurs where the electrodes
come in contact with seawater. The temperature sensor combines a tiny
thermistor with a platinum-resistance thermometer. Its operations are carried out
in such a way as to fully exploit the fast response of the thermistor and the high
accuracy of the platinum thermometer. In addition, the system uses a strain
gauge as a pressure sensor, the gauge being adjusted to reduce temperature
effects to a minimum. This CTD system is extremely reliable. While its
temperature precision is greater than 0.001 °C (0.002 °F) over a range of −3 to
+32 °C (26.6 to 89.6 °F), its conductivity precision is on the order of one part per
million.
Electrical conductivity measurement of seawater salinity has been so
effective that it has given rise to a new practical salinity scale, one that is defined
on the basis of conductivity ratio. This scale has proved to be a more reliable
way of determining density (i.e., the weight of any given volume of seawater at a
specified temperature) than the chlorinity scale traditionally used. Such is the
case because chlorinity is ion specific while conductivity is sensitive to changes
in any ion. Investigators have found that measurements of conductivity ratio
make it possible to predict density with a precision almost one order of
magnitude greater than was permitted by the chlorinity measurements of the
past.

WATER SAMPLING FOR CHEMICAL CONSTITUENTS


Nutrient concentration (e.g., phosphate, nitrate, silicate), the pH (acidity), and
the proportion of dissolved gases are used by the ocean chemist to determine the
age, origin, and movement of water masses and their effect on marine life.
Analysis of dissolved gases, for example, is useful in tracing ocean mixing, in
studying gas production in the ocean, and in elucidating the natural cycles of
atmospheric pollutants. Many such measurements are conducted aboard ship by
autoanalyzers, devices that continually monitor a flow of seawater by spectral
techniques. Those analyses that cannot be accomplished by an autoanalyzer are
carried out with discrete samples in shipboard or shore-based laboratories.

NANSEN BOTTLE

The Nansen bottle is an ocean-water sampler devised late in the 19th


century by the Norwegian oceanographer Fridtjof Nansen and subsequently
modified by various workers. The standard Nansen bottle is made of metal
and has a capacity of 1.25 litres. It is equipped with plug valves at either
end. The bottle is affixed to a winch wire with its valves open, and the
winch wire is paid out until the bottle is approximately at its desired
sampling depth. A weight, or “messenger,” is then allowed to slide down
the cable. The upper attachment of the Nansen bottle is disengaged from the
cable by the impact of the messenger; and the bottle is reversed end over
end, its valves closing in the process to trap the water sample.
Thermometers usually are attached to the Nansen bottle to record the
temperature and pressure of the sample site. Several Nansen bottles are
employed during a single hydrographic cast, each bottle releasing another
messenger when tripped, in order to trigger the deeper bottles in turn.
A schematic drawing of the Nansen bottle, one of the first
oceanographic water-sampling devices. Oceanographic Museum
of Monaco/National Oceanic and Atmospheric
Administration/Department of Commerce

The Niskin bottle, created by American inventor Shale Niskin in 1966,


is more widely used than the Nansen bottle in modern ocean-water
sampling activities. Although it is similar to the Nansen bottle in most
respects, the Niskin bottle is viewed as an improvement over Nansen’s
design because of its plastic construction and because it does not require
end-over-end movement to collect samples.

Radioactive chemical tracers are of special interest. Radioisotopes serve as


time clocks, thus offering a means of determining the age of water masses, the
absolute rates of oceanic mixing, and the generation and destruction of plant
tissue. The distribution of these time clocks is controlled by the interaction of
physical and biological processes, and so these influences must be disentangled
before the clocks can be read. A notable example is the use of carbon-14 (14C).
Today, a number of oceanographic laboratories make carbon-14 measurements
of oceanic dissolved carbon for the study of mixing and transport processes in
the deep ocean. Until recently large samples of water—about 200 litres (one litre
= 0.264 gallon)—were required for analysis. New techniques use a linear
accelerator (a device that greatly increases the velocity of electrically charged
atomic and subatomic particles) as a sophisticated mass spectrometer to directly
determine abundancy ratios of carbon-14/carbon-13/carbon-12 atoms. The
advantage of the newer methodology is that only very small sample amounts—
about 250 millilitres (one millilitre = 0.034 fluid ounce)—are required for high
accuracy measurements.

MEASUREMENTS OF OCEAN CURRENTS


Ocean currents can be measured indirectly through data on density and directly
with current meters. In the indirect technique, water density is computed from
temperature and salinity observations, and pressure is then calculated from
density. The resulting highs and lows of ocean pressure can be used to estimate
ocean currents. The indirect technique establishes currents relative to a particular
pressure surface; it is best for large-scale, low-frequency currents.
Direct measurement of currents is used to establish absolute currents and to
monitor rapidly varying changes. In order to measure currents directly, a current
meter must accurately record the speed and direction of flow, and the platform or
mooring has to be reliable, readily deployable, and extremely sturdy.
Researchers are able to make continuous measurements of currents at levels
below the surface layer for periods of more than a year.
A typical system for the direct measure of ocean currents has three principal
components: a surface or near-surface float; a line consisting of segments of wire
and nylon that holds the current meters; and a release mechanism, signalled
acoustically, which will drop an anchor when the system is ready to be brought
back. A current meter typically employs a rotor equipped with a small direction
vane that moves freely in line with the meter.
One of the most important advances in modern instrument design has been
the introduction of low-power, solid-state microelectronics. The accuracy of the
Vector Averaging Current Meter (VACM), for example, has been improved
appreciably by the use of integrated circuits, as has its data-handling capability.
Because of the latter, the VACM can sample the direction and speed of currents
roughly eight times during each revolution of the rotor. It then computes the
north and east components of speed and stores this data, together with direction
and time measurements, on a compact cassette recorder. The VACM is capable
of making accurate measurements in wave fields as well as from moorings at the
ocean surface because of its direct vector-averaging feature.
Currents also can be measured by drifting floats, either at the surface or at a
given depth. Tracking the location of the floats is critical. Surface floats can be
followed by satellite, but subsurface drifters must be tracked acoustically. A
drifter of this sort acts as an acoustical source and transmits signals that can be
followed by a ship with hydrophones suspended into the sea. For such tracking,
a low sound frequency is crucial because the higher the frequency of sound, the
more rapidly is its energy absorbed by the sea. The longest range floats available
during the mid-1980s operated at a frequency as low as about 250 hertz. Long-
range floats usually drift along channels known as sound fixing and ranging
(SOFAR) channels, which occur in various areas of the ocean where a particular
combination of temperature and pressure conditions affect the speed of sound. In
a sense, the SOFAR channel acts as a type of acoustic waveguide that focusses
sound; as a consequence, several watts of sound can be detected as far away as
2,000 km (about 1,200 miles) or so.
Measuring vertical velocity in the ocean posed a major problem for years
because of the difficulty of devising a platform that does not move vertically.
During the 1960s oceanographers finally came up with a solution: they
employed a neutrally buoyant float for measuring vertical velocities. This form
of vertical-current meter consists of a cylindrical float on which fins are mounted
at an angle. When water moves past the float, it causes the float to turn on its
axis. Measurement of the rotation in relation to a compass yields the amount of
vertical water movement.
An extension of the neutrally buoyant float is the self-propelled, guided float.
One such system, called a Self-Propelled Underwater Research Vehicle
(SPURV), manoeuvres below the surface of the sea in response to acoustic
signals from the research vessel. It can be used to produce horizontal as well as
vertical profiles of various physical properties.
A Doppler-sonar system for measuring upper-ocean current velocity
transmits a narrow beam that scatters off drifting plankton and other organisms
in the uppermost strata of the ocean. From the Doppler shift of the backscattered
sound, the component of water velocity parallel to the beam can be determined
to a range of 1,400 metres (about 5,000 feet) from the transmitter with a
precision of 1 cm per 0.1 second (1 cm = 0.394 inch).
Integral to a complete picture of the ocean is a profile of velocity. Various
methods have been devised for measuring currents as dependent on and varying
with depth or horizontal position. Three techniques have been developed to
make such measurements. The first involves acoustically tracking a “sinking
float” as it descends toward the seafloor. The second technique entails the use of
a free-fall device equipped with a current sensor. The third involves a class of
current meter specially designed to move up and down a fixed line attached to a
vessel, mooring, or drifting buoy. One such instrument has a roller block that
couples the front of the instrument to a wire from the vessel. In this way, the
motion of the vessel is decoupled from that of the instrument. Another important
component of this instrument is its hull, a structure that not only furnishes
buoyancy but also serves as a direction vane. In the bottom of the hull is a device
that records velocity, temperature, and depth. The entire system descends at a
rate of approximately 10 cm (about 4 inches) per second, resulting in a vertical
resolution of several metres for the velocity profile produced.

ACOUSTIC AND SATELLITE SENSING


Remote sensing of the ocean can be done by aircraft and Earth-orbiting satellites
or by sending acoustic signals through it. These techniques all offer a more
sweeping view of the ocean than can be provided by slow-moving ships and
hence have become increasingly important in oceanographic research.
Satellite-borne radar altimeters have proved to be especially useful. A radar
system of this type can determine the distance between the satellite and the sea
surface to an accuracy of better than 10 cm by measuring the time it takes for a
transmitted pulse of radio energy to travel to the surface and return. By
combining such a precise distance measurement with information about the
satellite’s orbit, oceanographers are able to produce maps of sea-surface
topography. Moreover, they can deduce the pressure field of the sea surface by
combining the distance measurement with knowledge about the geoid. They can
in turn extrapolate information about the general circulation of the upper stratum
of the ocean from a synoptic view of the surface pressure field.
Another remote-sensing technique involves the use of satellite-borne infrared
and microwave radiometers to measure radiant energy released from the surface
of the ocean. Such measurements are used to determine sea-surface temperature.
High-resolution, infrared images transmitted by polar-orbiting satellites have
provided researchers with an effective means of monitoring wave features in
ocean currents over a wide area, as, for example, long equatorial waves in the
Pacific Ocean and time variations in the flow of the Gulf Stream between
Florida and Cape Hatteras, North Carolina.
Acoustic techniques also have many applications in the study of the ocean,
particularly of those subsurface processes and physical properties inaccessible to
satellite observation. In one such technique, the temperature structure of a water
column from a given point on the seafloor to the surface is studied using an
inverted echo sounder. This instrument, which features both an acoustic
transmitter and a receiver, measures the time taken by a pulse of sound to travel
from the sea bottom to the surface and back again. In most cases, a change in the
average temperature of the water column above the instrument causes a
fluctuation in the time interval between the transmission and the reception of the
acoustic signal.

SONAR

Sonar (from “sound navigation ranging”) is the technique for detecting and
determining the distance and direction of underwater objects by acoustic
means. Sound waves emitted by or reflected from the object are detected by
sonar apparatus and analyzed for the information they contain.
Sonar systems may be divided into three categories. In active sonar
systems an acoustic projector generates a sound wave that spreads outward
and is reflected back by a target object. A receiver picks up and analyzes the
reflected signal and may determine the range, bearing, and relative motion
of the target. Passive systems consist simply of receiving sensors that pick
up the noise produced by the target (such as a ship, submarine, or torpedo).
Waveforms thus detected may be analyzed for identifying characteristics as
well as direction and distance. The third category of sonar devices is
acoustic communication systems, which require a projector and receiver at
both ends of the acoustic path.
Sonar was first proposed as a means of detecting icebergs. Interest in
sonar was heightened by the threat posed by submarine warfare in World
War I. An early passive system, consisting of towed lines of microphones,
was used to detect submarines by 1916, and by 1918 an operational active
system had been built by British and U.S. scientists. Subsequent
developments included the echo sounder, or depth detector, rapid-scanning
sonar, side-scan sonar, and WPESS (within-pulse electronic-sector-
scanning) sonar.
An AQS-13 dipping sonar being deployed in the ocean. C.
Yebba/U.S. Navy

The uses of sonar are now many. In the military field are a large number
of systems that detect, identify, and locate submarines. Sonar is also used in
acoustic homing torpedoes, in acoustic mines, and in mine detection.
Nonmilitary uses of sonar include fish finding, depth sounding, mapping of
the sea bottom, Doppler navigation, and acoustic locating for divers.
A major step in the development of sonar systems was the invention of
the acoustic transducer and the design of efficient acoustic projectors. These
utilize piezoelectric crystals (e.g., quartz or tourmaline), magnetostrictive
materials (e.g., iron or nickel), or electrostrictive crystals (e.g., barium
titanate). These materials change shape when subjected to electric or
magnetic fields, thus converting electrical energy to acoustic energy.
Suitably mounted in an oil-filled housing, they produce beams of acoustic
energy over a wide range of frequencies.
In active systems the projector may be deployed from an air-launched
sonobuoy, hull-mounted on a vessel, or suspended in the sea from a
helicopter. Usually the receiving and transmitting transducers are the same.
Passive systems are usually hull-mounted, deployed from sonobuoys, or
towed behind a ship. Some passive systems are placed on the seabed, often
in large arrays, to provide continuous surveillance.
Other acoustic techniques can be utilized to study ocean variables on a large
scale. A method known as ocean acoustic tomography, for example, monitors the
travel time of sound pulses with an array of echo-sounding systems. In general,
the amount of data collected is directly proportional to the product of the number
of transmitters and receivers, so that much information on averaged oceanic
properties can be gathered within a short period of time at relatively low cost.

COLLECTION OF BIOLOGICAL SAMPLES


Life at the bottom, benthos, is affected by the water column and by the
sediment–water interface; the swimmers, or nekton, are influenced by the water
that they come in contact with; and the floaters, or plankton—phytoplankton
(plant forms) and zooplankton (animal forms)—are influenced by the water and
the transfers that occur at the surface of the sea. Thus, in most cases,
measurements and sampling of marine life is best done in concert with
measurements of the physical and chemical properties of the ocean and the
surface effects of the atmosphere.
As a consequence of the close interaction of sea life and its environment,
marine biologists and biological oceanographers use most of the techniques
mentioned above as well as some specialized techniques for biological sampling.
Investigative techniques include the use of sampling devices, remote sensing of
surface life-forms by satellite and aircraft, and in situ observation of plants and
animals in direct interaction with their environment. The latter is becoming
increasingly important as biologists recognize the fragility of organisms and the
difficulty of obtaining representative samples. The absence of good sampling
techniques means that even today little is known about the distribution, number,
and life cycles of many of the important species of marine life.
Some of the most commonly used samplers are plankton nets and midwater
trawls. Nets have a mesh size smaller than the plankton under investigation;
trawls filter out only the larger forms. The smaller net sizes can be used only
when the ship is either stopped or moving ahead slowly; the larger can be used
while the ship is travelling at normal speeds. Plankton nets can be used to
sample at one or more depths. Qualitative samplers sieve organisms from the
water without measuring the volume of water passing through, whereas
quantitative samplers measure the volume and hence the concentration of
organisms in a unit volume of seawater.
The Clark-Bumpus sampler is a quantitative type designed to take an
uncontaminated sample from any desired depth while simultaneously estimating
the filtered volume of seawater. It is equipped with a flow meter that monitors
the volume of seawater that passes through the net. A shutter opens and closes
on demand from the surface, admitting water and spinning the impeller of the
meter while catching the plankton. When the impeller is stopped by closing the
shutter, the sampler can be raised without contamination from plankton in the
waters above.
The midwater trawl is specially designed for rapid collection at depths well
below the surface and at such a speed that active, fast-swimming fish are unable
to escape from the net once caught. Trawls can be towed at speeds up to 9 km
(5.6 miles) per hour. To counteract the tendency of an ordinary net to surface
behind the towing vessel, a midwater trawl of the Isaacs-Kidd variety uses an
inclined-plane surface rigged in front of the net entrance to act as a depressor.
The trawl is shaped like an asymmetrical cone with a pentagonal mouth opening
and a round closed end. Within the net, additional netting is attached as lining. A
steel ring is fastened at the end of the net to maintain shape. A large perforated
can is fastened by drawstrings on the end of the net to retain the sample in
relatively undamaged condition.
The use of acoustics to record and measure the distribution of biological
organisms is becoming a widely adopted practice. Some organisms can be
tracked directly by their distinctive sounds. By recording and analyzing these
sounds, biologists are able to chart the behaviour and distribution of such life-
forms.
Organisms that passively affect various electronic systems are large
mammals, schools of fish, and plankton that either scatter sound and so appear
as false targets or background reverberation, or that attenuate the acoustic signal.
Some fishes and invertebrates make up layers of acoustic-scattering material,
which may exhibit daily vertical movement related to daily changes in light.
Light in the upper layers of the ocean is crucial to maintaining marine life.
The penetration and absorption of light and the colour and transparency of the
ocean water are indicative of biological activity and of suspended material. In
situ measurements of water transparency and absorption include the submarine
photometer, the hydrophotometer, and the Secchi disk. The submarine
photometer records directly to depths of about 150 metres (about 500 feet) the
infrared, visible, and ultraviolet portions of the spectrum. The hydrophotometer
has a self-contained light source that allows greater latitude in observation
because it can be used at any time of night or day and measures finer gradations
of transparency. The Secchi disk, designed to measure water transparency, is a
circular white disk that is lowered on a cable into the sea. In practice, the depth
at which it is barely visible is noted. The greater the depth reading, the more
transparent is the water.
The primary productivity of the ocean, which occurs in the upper layers, can
be monitored by continuous measurement of absorption by chlorophyll
molecules. This occurs in the red and blue portions of the spectrum, leaving the
green to represent the characteristic colour of biological activity. Satellite
measurements of ocean colour that span a number of wavelengths in the visible
and infrared portions of the spectrum are used to give a large-scale view of the
biological activity and suspended material in the ocean.

EXPLORATION OF THE SEAFLOOR AND THE EARTH’S CRUST


The ocean floor has the same general character as the land areas of the world:
mountains, plains, channels, canyons, exposed rocks, and sediment-covered
areas. The lack of weathering and erosion in most areas, however, allows
geological processes to be seen more clearly on the seafloor than on land.
Undisturbed sediments, for example, contain a historical record of past climates
and the state of the ocean, which has enabled geologists to find a close relation
between past climates and the variation of the distance of Earth from the Sun
(the Milankovich effect).
Because electromagnetic radiation cannot penetrate any significant distance
into the sea, the oceanographer uses acoustic signals, explosives, and
earthquakes, as well as gravity and magnetic fields, to probe the seafloor and the
structure beneath. Such techniques—which now include the capability to
produce a swath, or two-dimensional, description of the seafloor beneath a ship
—are providing increasingly accurate data on the shape of the ocean, its
roughness, and the structure beneath. Satellite techniques are a more recent
development. Because the shape of the sea surface is closely related to that of
the seafloor due to gravity, satellite measurements of surface topography have
been used to provide a global view of the ocean bottom. They also have
provided data for an accurate mapping of such features as seamounts.
Research on marine sedimentation involves the study of deposition,
composition, and classification of organic and inorganic materials found on the
seafloor. Samples of such materials are thoroughly examined aboard research
vessels or in shore-based laboratories, where investigators analyze the size and
shape of constituent particles, determine chemical properties such as pH, and
identify and categorize the minerals and organisms present. From thousands of
reported classifications and collected samples, bottom-sediment charts are
prepared.
Various kinds of equipment are used to obtain samples from the seafloor.
These include grabbing devices, dredges, and coring devices.
Grabbing devices, commonly known as snappers, vary widely in size and
design. One general class of such devices is the clamshell snapper, which is used
to obtain small samples of the superficial layers of bottom sediments. Clamshell
snappers come in two basic varieties. One measures 76 cm (about 30 inches) in
length, weighs roughly 27 kg (60 pounds), and is constructed of stainless steel.
The jaws of this device are closed by heavy arms, which are actuated by a strong
spring and lead weight. It is capable of trapping about a pint of bottom material.
The second type of clamshell snapper is appreciably smaller. Commonly called
the mud snapper, this device is approximately 28 cm (11 inches) long and
weighs 1.4 kg (3.1 pounds). Other grabbing devices include the orange peel
bucket sampler, which is used for collecting bottom materials in shallow waters.
A small hook attached to the end of the lowering wire supports the sampler as it
is lowered and also holds the jaws open. When contact is made with the bottom,
the sampler jaws sink into the sediment and the wire tension is released,
allowing the hook to swing free of the sampler. Upon hoisting, the wire takes a
strain on the closing line, which closes the jaws and traps a sample. The
underway bottom sampler, or scoopfish, is designed to sample rapidly without
stopping the ship. It is lowered to depths less than 200 metres (about 660 feet)
from a ship moving at speeds no more than 28 km (about 17 miles) per hour. The
sampler weighs 5 kg (11 pounds) and can capture samples ranging from mud to
coral.
The second major category of bottom sampler is the dredge, which is
dragged along the seafloor to collect materials. Bottom-dredging operations
require very sturdy gear, particularly when dredging for rock samples. A typical
dredge is constructed of steel plate and is 30 cm (12 inches) deep, 60 cm (24
inches) wide, and 90 cm (35 inches) long. The forward end is open, but the aft
end has a heavy grill of round steel bars that is designed to retain large rock
samples. When finer sized material is sought, a screen of heavy hardware cloth
is placed over the grill.
Coring devices typically have three principal components: interchangeable
core tubes, a main body of streamlined lead weights, and a tailfin assembly that
directs the corer in a vertical line to the ocean bottom. The amount of sediment
collected depends on the length of the corer, the size of the main weight, and the
penetrability of the bottom. One type of coring device, the lightweight Phleger
corer, takes samples only of the upper layer of the ocean bottom to a depth of
about 1 metre (3.3 feet). Deeper cores are taken by the piston corer. In this
device, a closely fitted piston attached to the end of the lowering cable is
installed inside the coring tube. When the coring tube is driven into the ocean
floor, friction exerts a downward pull on the core sample. The hydrostatic
pressure on the ocean bottom, however, exerts an upward pressure on the core
that will work against a vacuum being created between the piston and the top of
the core. The piston, in effect, provides a suction that overcomes the frictional
forces acting between the sediment sample and the inside of the coring tube. The
complete assembly of a typical piston core weighs about 180 kg (about 400
pounds) and can be used to obtain samples as long as 20 metres (about 66 feet).
An improved version of this device, the hydraulic piston corer, is used by deep-
sea drilling ships such as the JOIDES Resolution. Essentially undisturbed cores
of lengths up to 200 metres (about 660 feet) have been obtained with this type of
corer.
Investigators may also make use of wire-line logging tools that are capable
of measuring electrical resistance, acoustic properties, and magnetic and
gravitational effects in the holes drilled. The JOIDES Resolution is equipped
with tools of this sort, including a remote television camera, which are lowered
into a drill hole after the core has been removed. Such wire-line logging
apparatus make data immediately available for scientific analysis and decision
making.
Acoustic techniques have reached a high level of sophistication for
geological and geophysical studies. Such multifrequency techniques as those that
employ Seabeam and Gloria (Geological Long-Range Inclined Asdic) permit
mapping two-dimensional swaths with great accuracy from a single ship. These
methods are widely used to ascertain the major features of the seafloor. The
Gloria system, for example, can produce a picture of the morphology of a region
at a rate of up to 1,000 square km (about 400 square miles) per hour. Techniques
of this kind are employed in conjunction with seismic reflection techniques,
which involve the use of multichannel receiving arrays to detect sound waves
triggered by explosive shots (e.g., dynamite blasts) that are reflected off of
interfaces separating rocks of different physical properties. Such techniques
make it possible to measure the structure of Earth’s crust deep below the
seafloor.

CONCLUSION
Covering the majority of Earth’s surface, the oceans are the planet’s most
prominent features, and in many ways, they are among the most important ones.
The oceans are critical components of Earth’s climate. They are tremendous
absorbers of solar energy and dispersers of heat. Ocean currents move masses of
warm and cold water from their source regions in an attempt to equilibrate
surface heating across the globe. As this water moves, it interacts with processes
in the atmosphere that drive temperature and rainfall patterns across the globe.
Life began in the oceans billions of years ago. Since then, the earliest living
things have evolved into more and more complex forms. Although some of the
descendants of early organisms left the oceans to live on land, many remained.
Today, the oceans teem with crustaceans, corals, mollusks, and other
invertebrates, as well as fishes, mammals, and other vertebrates. From this
bounty, a large number of species are harvested by humans for food.
Like other parts of the living Earth, the oceans are frequently harmed by
human activities. Though populations of many fish species have declined due to
overharvesting, many other fishes and other marine life face the scourge of
pollution. Since Earth’s water courses terminate in the oceans, they are often the
final resting places for trash and other forms of pollution released into rivers and
streams. In addition, the oceans are traversed by large vessels carrying petroleum
and other harmful chemicals. When these vessels are damaged or sink, their
cargoes may be released into the surrounding water. Oil released from damaged
tankers or oil drilling equipment does not sink harmlessly to the ocean floor. It
rises to the surface. Driven by wind and waves toward the shorelines of
continents and other landmasses, it fouls beaches and coastal wetlands, killing
wildlife and threatening the livelihoods of local human residents.
CHAPTER 6
MARINE EXPLORERS AND OCEANOGRAPHERS

The modern scientific understanding of the oceans owes much to the work of
oceanographers and marine explorers. Some of the more influential personages
include Jacques-Yves Cousteau, noted explorer and inventor of the Aqua-Lung,
submersible pioneer Robert Ballard, and celebrated Arctic explorer Fridtjof
Nansen.

ALEXANDER AGASSIZ
(b. Dec. 17, 1835, Neuchâtel, Switz.—d. March 27, 1910, at sea),

Swiss-born American marine zoologist, oceanographer, and mining engineer


Alexander Emmanuel Rodolphe Agassiz made important contributions to
systematic zoology, to the knowledge of ocean beds, and to the development of a
major copper mine.
Son of the Swiss naturalist Louis Agassiz, he joined his father in 1849 in the
U.S., where he studied engineering and zoology at Harvard University. His early
research on echinoderms (e.g., starfish) resulted in his most significant work in
the area of systematic zoology, the Revision of the Echini (1872–74).
From 1866 to 1869 Agassiz was the superintendent of a copper mine near
Lake Superior, at Calumet, Mich.; he eventually became president of the mine,
retaining that position until his death, by which time he had changed an initially
unprofitable operation into the world’s foremost copper mine. Agassiz instituted
modern machinery and safety devices, pension and accident funds for miners,
and sanitary measures for surrounding communities. He donated large sums to
the Harvard Museum, of which he was curator from 1874 to 1885, and to other
institutions engaged in advancing the study of biology.
Agassiz opened a private biological laboratory at Newport, R.I., soon after
closing his father’s experimental school of biology on Penikese Island, off the
coast of Massachusetts, following the death of Louis Agassiz (1873). On a trip
along the west coast of South America (1875), he discovered a coral reef 3,000
feet above sea level, an observation that appeared to contradict Charles Darwin’s
theory of coral-reef formation, in which he postulated a rate of coral formation
identical with the rise of sea level. He continued his marine and coral-reef
studies for more than 25 years, making expeditions in various waters, including
the Caribbean, the South Pacific, and the Great Barrier Reef of Australia.
Alexander Agassiz. National Oceanic and Atmospheric Administration

PRINCE ALBERT I OF MONACO


(b. Nov. 13, 1848, Paris, France—d. June 26, 1922, Paris)

Albert-Honoré-Charles Grimaldi was prince of Monaco from 1889 to 1922. He


was also a seaman, an amateur oceanographer, and a patron of the sciences,
whose contributions to the development of oceanography included innovations
in oceanographic equipment and technique and the founding and endowment of
institutions to further basic research.
Albert’s love of the sea developed at an early age, and as a young man he
served in the Spanish Navy. Later, he conducted his own oceanographic surveys
on a series of increasingly large and well-equipped ships. His active involvement
in oceanography continued even after he became ruler of Monaco—upon the
death of his father, Charles III (1889)—and culminated in his establishment of
the Oceanographic Museum of Monaco (1899) and the Oceanographic Institute
in Paris (1906).
Albert was succeeded as prince of Monaco by his son, Louis II (1870–1949).
ROBERT BALLARD
(b. June 30, 1942, Wichita, Kan., U.S.)

Robert Duane Ballard is an American oceanographer and marine geologist


whose pioneering use of deep-diving submersibles laid the foundations for deep-
sea archaeology. He is best known for discovering the wreck of the Titanic in
1985.
Ballard grew up in San Diego, California, where he developed a fascination
with the ocean. He attended the University of California in Santa Barbara,
earning degrees in chemistry and geology in 1965. As a member of the Reserve
Officers Training Corps, he entered the army following graduation, serving a
two-year tour before requesting a transfer to the navy. In 1967 he was assigned
to the Woods Hole (Massachusetts) Oceanographic Research Institution, where
he became a full-time marine scientist in 1974 after completing his doctoral
degrees in marine geology and geophysics at the University of Rhode Island.
In the early 1970s Ballard helped develop Alvin, a three-person submersible
equipped with a mechanical arm. In 1973–75 he dived 2,750 metres (9,000 feet)
in Alvin and in a French submersible to explore the Mid-Atlantic Ridge, an
underwater mountain chain in the Atlantic Ocean. In 1977 and 1979 he was part
of an expedition that uncovered thermal vents in the Galápagos Rift. The
presence of plant and animal life within these deep-sea warm springs led to the
discovery of chemosynthesis, the chemical synthesis of food energy.
To advance deep-sea exploration, Ballard designed a series of vessels, most
notably the Argo, a 5-metre (16-foot) submersible sled equipped with a remote-
controlled camera that could transmit live images to a monitor. Ballard called
this new technology “telepresence.” To test the Argo, he searched for the Titanic,
which had sunk in 1912 and remained undiscovered despite numerous attempts
to locate it. Working with the Institut Français de Recherche pour l’Exploitation
de la Mer (IFREMER; French Research Institute for the Exploitation of the Sea),
Ballard began the mission in August 1985 aboard the U.S. Navy research ship
Knorr. The Argo was sent some 4,000 metres (13,000 feet) to the floor of the
North Atlantic, sending video to the monitors on the Knorr. On Sept. 1, 1985,
the first images of the ocean liner were recorded as its giant boilers were
discovered. Later video revealed that the Titanic was lying in two pieces, with
the hull upright and largely intact. Ballard returned to the site in 1986, traveling
to the underwater wreckage in the submersible Alvin.
In 1989 Ballard established the JASON project, an educational program that
used video and audio satellite feeds and later the Internet to allow students to
follow various expeditions. In 1997 Ballard, then a commander in the navy, left
Woods Hole to head the Institute for Exploration in Mystic, Conn., a centre for
deep-sea archaeology that he founded. In 2002 he joined the faculty of the
University of Rhode Island’s Graduate School of Oceanography. He continued to
search for shipwrecks, and his notable discoveries include ancient vessels and
World War II ships, including the Bismarck (sunk 1941). A prolific writer,
Ballard described his expeditions in a number of books and articles.

JACQUES-YVES COUSTEAU
(b. June 11, 1910, Saint-André-de-Cubzac, France—d. June 25, 1997, Paris)

Jacques-Yves Cousteau was a French naval officer and ocean explorer, known
for his extensive underseas investigations.
After graduating from France’s naval academy in 1933, he was
commissioned a second lieutenant. However, his plans to become a navy pilot
were undermined by an almost fatal automobile accident in which both his arms
were broken. Cousteau, not formally trained as a scientist, was drawn to
undersea exploration by his love both of the ocean and of diving. In 1943
Cousteau and French engineer Émile Gagnan developed the first fully automatic
compressed-air Aqua-Lung. Cousteau helped to invent many other tools useful
to oceanographers, including the diving saucer (an easily maneuverable small
submarine for seafloor exploration) and a number of underwater cameras.
Cousteau served in World War II as a gunnery officer in France and was a
member of the French Resistance. He later was awarded the Legion of Honour
for his espionage work. Cousteau’s experiments with underwater filmmaking
began during the war. Cousteau helped found the French navy’s Undersea
Research Group in 1945. He also was involved in conducting oceanographic
research at a centre in Marseille, France. When the war ended, he continued
working for the French navy, heading the Undersea Research Group at Toulon.
To expand his work in marine exploration, he founded numerous marketing,
manufacturing, engineering, and research organizations, which were
incorporated in 1973 as the Cousteau Group. In 1950 Cousteau converted a
British minesweeper into the Calypso, an oceanographic research ship, aboard
which he and his crew carried out numerous expeditions. Cousteau eventually
popularized oceanographic research and the sport of scuba diving in the book Le
Monde du silence (1952; The Silent World), written with Frédéric Dumas. Two
years later he adapted the book into a documentary film that won both the Palme
d’Or at the 1956 Cannes international film festival and an Academy Award in
1957, one of three Oscars his films received.
Also in 1957, Cousteau became director of the Oceanographic Museum of
Monaco. He led the Conshelf Saturation Dive Program, conducting experiments
in which men live and work for extended periods of time at considerable depths
along the continental shelves. The undersea laboratories were called Conshelf I,
II, and III.
Cousteau produced and starred in many television programs, including the
American series “The Undersea World of Jacques Cousteau” (1968–76). In 1974
he formed the Cousteau Society, a nonprofit environmental group dedicated to
marine conservation. In addition to The Silent World, Cousteau also wrote Par
18 mètres de fond (1946; Through 18 Metres of Water), The Living Sea (1963),
Three Adventures: Galápagos, Titicaca, the Blue Holes (1973), Dolphins (1975),
and Jacques Cousteau: The Ocean World (1985). His last book, The Human, the
Orchid, and the Octopus: Exploring and Conserving Our Natural World (2007),
was published posthumously.

ROBERT S. DIETZ
(b. Sept. 14, 1914, Westfield, N.J., U.S.—d. May 19, 1995, Tempe, Ariz.)

Robert Sinclair Dietz was an American geophysicist and oceanographer who set
forth a theory of seafloor spreading in 1961.
Dietz was educated at the University of Illinois (B.S., 1937; M.S., 1939;
Ph.D., 1941). After serving as an officer in the U.S. Army Air Corps during
World War II, he became a civilian scientist with the U.S. Navy. In this capacity,
he supervised the oceanographic research on Admiral Richard E. Byrd’s last
Antarctic expedition (1946–47). He subsequently served as oceanographer with
several organizations, including the U.S. Coast and Geodetic Survey (1958–65)
and the Atlantic Oceanography and Meteorology Laboratories (1970–77). He
became professor of geology at Arizona State University, Tempe, in 1977.
Dietz’s discovery in 1952 of the first fracture zone in the Pacific, which he
related to deformation of Earth’s crust, led him to hypothesize that new crustal
material is formed at oceanic ridges and spreads outward at a rate of several
centimetres per year. Subsequent work confirmed this suggestion. He helped to
redevelop the bathyscaphe Trieste of Swiss engineer Jacques Piccard, who
descended about 7 miles (11 km) into the Pacific Ocean in it in 1960. Dietz also
became known for his work in the fields of selenography (study of the Moon’s
physical features) and meteoritics, particularly for his suggestion that certain
shock effects in rocks are indicative of meteorite impact.

VAGN WALFRID EKMAN


(b. May 3, 1874, Stockholm, Swed.—d. March 9, 1954, Gostad, near Stockaryd,
Swed.)

Swedish physical oceanographer Vagn Walfrid Ekman was best known for his
studies of the dynamics of ocean currents. The common oceanographic terms
Ekman layer, denoting certain oceanic or atmospheric layers occurring at various
interfaces; Ekman spiral, used in connection with vertical oceanic velocity; and
Ekman transport, denoting wind-driven currents, derive from his research.
Ekman was the youngest son of Fredrik Laurentz Ekman, a Swedish physical
oceanographer. After finishing secondary school in Stockholm, Ekman studied at
the University of Uppsala, where he majored in physics. But lectures on
hydrodynamics in 1897 by Vilhelm Bjerknes, one of the founders of
meteorology and oceanography, definitely decided the direction of Ekman’s
work.
While still a student at Uppsala, Ekman made important contributions to
oceanography. When it was observed, during the Norwegian North Polar
Expedition, that drift ice did not follow the wind direction but deviated by 20° to
40°, Bjerknes chose Ekman to make a theoretical study of the problem. In his
report, published in 1902, Ekman took into account the balance of the frictions
between the wind and sea surface, within layers of water, and the deflecting
force due to Earth’s rotation (Coriolis force).
After taking his degree at Uppsala in 1902, he joined the staff of the
International Laboratory for Oceanographic Research in Oslo, where he
remained until 1909. During those years he proved to be a skilled inventor and
experimentalist. The Ekman current meter, an instrument with a simple and
reliable mechanism, has been used, with subsequent improvements, to the
present, while the Ekman reversing water bottle is used in freshwater lakes and
sometimes in the ocean to obtain water samples at different depths with a
simultaneous measurement of water temperatures. He displayed his theoretical
and experimental talents in his study of so-called dead water, which causes slow-
moving boats to become stuck because of a thin layer of nearly fresh water
spreading over the sea from melting ice. This phenomenon, frequently occurring
in fjords, seriously impeded the Norwegian explorer Fridtjof Nansen in Arctic
waters. Ekman demonstrated by experiments in a wave tank that the resistance to
the motion of the vessels is increased by the waves that are formed at the
interface between layers of water of different densities.
He also derived an empirical formula for the mean compressibility
(compression ratio divided by pressure) of seawater as a function of pressure and
temperature. This formula is still in use today to determine density of deep
seawater which is compressed by hydrostatic pressure.
From 1910 to 1939, Ekman was professor of mechanics and mathematical
physics at the University of Lund in Sweden, where he pursued his main interest,
the dynamics of ocean currents. He published theories on wind-driven ocean
currents, including the effects of coasts and bottom topography, and on the
dynamics of the Gulf Stream. He also tried, with partial success, to solve the
complex problem of ocean turbulence.
In 1925 Ekman participated in a cruise of a German research ship to the
Canary Islands. On finding that data on currents obtained for several days at
several marine stations between the Bay of Biscay and the Canary Islands were
not sufficient to obtain an average figure, he and a colleague, during the years
1922–29, improved the technique for measuring currents for a prolonged period
by collecting data from an anchored ship. After several preparatory cruises for
that purpose off the Norwegian coast aboard a Norwegian research vessel, they
made a cruise to the trade-wind region off northwestern Africa in the summer of
1930 to determine the average current at various ocean depths at stations
occupied for two weeks or longer. Preliminary reports were published soon after
the cruise, but Ekman wrote the final report in 1953 at the age of 79. The long
delay in publication, partly owing to the loss of important data during the
German occupation of Norway, also indicates the unparalleled care he took with
his work.
Although his name and achievements were well known among
oceanographers, he rarely attended international meetings—but his genuine
kindliness prevented him from becoming a recluse. Most of his teachers and
friends, such as Nansen and Bjerknes, were Norwegian; he spent many vacations
in Bergen. He sang a beautiful bass, spent much time at the piano, and
occasionally composed music. In the fall of 1953, he began a study of turbid
currents, which he continued until a few days before his death.
RICHARD H. FLEMING
(b. Sept. 21, 1909, Victoria, B.C., Can.—d. Oct. 25, 1989, Seattle, Wash., U.S.)

Canadian-born American oceanographer Richard Howell Fleming conducted


wide-ranging studies in the areas of chemical and biochemical oceanography,
ocean currents (particularly those off the Pacific coast of Central America), and
naval uses of oceanography.
Fleming joined the Scripps Institution in La Jolla, Calif., in 1931 and served
as assistant director from 1946 to 1950, when he left to become professor of
oceanography at the University of Washington.
Fleming measured the amounts of chemical elements in sea water and in
marine life. His work on ocean currents included measurement of tidal-current
velocities, description of upwelling of seawater, and observations of seasonal
variations in currents. He also explored sedimentation in moving and stagnant
bodies of water. With H.U. Sverdrup and Martin W. Johnson, Fleming wrote The
Oceans (1942).

BJØRN HELLAND-HANSEN
(b. Oct. 16, 1877, Christiania, Nor.—d. Sept. 7, 1957, Bergen)

Bjørn Helland-Hansen was a Norwegian pioneer of modern oceanography


whose studies of the physical structure and dynamics of the oceans were
instrumental in transforming oceanography from a science that was mainly
descriptive to one based on the principles of physics and chemistry.
Most of Helland-Hansen’s work was done in Bergen, where he was
successively director of the Marine Biological Station, professor at the Bergen
Museum, and first director of the Geophysical Institute, which was established in
1917 largely through his efforts. He was active in international scientific affairs
and in 1945 was elected president of the International Union of Geodesy and
Geophysics.

COLUMBUS O’DONNELL ISELIN


(b. Sept. 25, 1904, New Rochelle, N.Y., U.S.—d. Jan. 5, 1971, Vineyard Haven,
Mass.)
American oceanographer Columbus O’Donnell Iselin served as director of the
Woods Hole Oceanographic Institution (1940–50; 1956–57) in Massachusetts,
expanded its facilities tenfold and made it one of the largest research
establishments of its kind in the world.
The scion of a New York banking family (his greatgrandfather had helped
found the Metropolitan Opera House and the Metropolitan Museum of Art),
Iselin attended Harvard University (A.B., 1926; A.M., 1928). From 1926 on, he
made a series of summer sea excursions toward Labrador and the Arctic with his
own schooner and crew, collecting material and data for Harvard’s Museum of
Comparative Zoology. For Harvard he served as assistant curator of
oceanography (1929–48) and research oceanographer of the Museum of
Comparative Zoology. Concurrently, in 1932 he joined the newly established
Woods Hole institution and, from 1936, taught oceanography at Harvard.
In 1940 he was named director of Woods Hole and, with wartime funds from
the U.S. Navy, vastly increased the institution’s budget and size. Wartime studies
—factors in seaborne invasions, ocean currents, underwater explosions, and
other matters—turned in 1946 to peacetime studies of fisheries, the dynamics of
currents, the profiles of ocean floors, and other oceanographic concerns. Later,
Iselin was professor of oceanography at the Massachusetts Institute of
Technology (1959–70) and Harvard University (1960–70).

MATTHEW FONTAINE MAURY


(b. Jan. 14, 1806, Spotsylvania County, Va., U.S.—d. Feb. 1, 1873, Lexington,
Va.)

Matthew Fontaine Maury was a U.S. naval officer and pioneer hydrographer. He
is considered to be one of the founders of oceanography.
Maury entered the navy in 1825 as a midshipman, circumnavigated the globe
(1826–30), and in 1836 was promoted to the rank of lieutenant. In 1839 he was
lamed in a stagecoach accident, which made him unfit for active service. In 1842
he was placed in charge of the Depot of Charts and Instruments, out of which
grew the U.S. Naval Observatory and Hydrographic Office. To gather
information on maritime winds and currents, Maury distributed to captains
specially prepared logbooks from which he compiled pilot charts, enabling ships
to shorten the time of sea voyages. In 1848 he published maps of the main wind
fields of Earth. Maury’s work inspired the first international marine conference,
held in Brussels in 1853. He was U.S. representative at the meeting that led to
the establishment of the International Hydrographic Bureau. Provided with
worldwide information, Maury was able to produce charts of the Atlantic,
Pacific, and Indian oceans. He also prepared a profile of the Atlantic seabed,
which proved the feasibility of laying a transatlantic telegraph cable. In 1855 he
published the first modern oceanographic text, The Physical Geography of the
Sea. In that year his Sailing Directions included a section recommending that
eastbound and westbound steamers travel in separate lanes in the North Atlantic
to prevent collisions.

Matthew Fontaine Maury. Harris & Ewing Collection/Library of


Congress, Washington, D.C. (Digital File Number: LC-DIG-hec-
07970)

On the outbreak (1861) of the American Civil War, Maury returned to


Virginia to become head of coast, harbour, and river defenses for the
Confederate Navy, for which he attempted to develop an electric torpedo. In
1862 he went to England as a special agent of the Confederacy, and at the war’s
end (1865) he went to Mexico, where the emperor Maximilian made him
imperial commissioner of immigration so that Maury could establish a
Confederate colony there. In 1866, when the emperor abandoned this scheme,
Maury went back to England. He returned to the United States in 1868 and
accepted the professorship of meteorology at Virginia Military Institute, a post
he held until his death. Maury Hall at Annapolis, Md., is named in his honour,
and his birthday is a school holiday in Virginia.

WALTER MUNK
(b. Oct. 19, 1917, Vienna, Austria)

Austrian-born American oceanographer Walter Heinrich Munk pioneered studies


of ocean currents and wave propagation and laid the foundations for
contemporary oceanography.
The child of a wealthy family, Munk was born and raised in Vienna. He
moved to Lake George, N.Y., in 1932 to attend boarding school, as his parents
hoped to prepare him for a career in banking. He worked in banking for several
years but grew dissatisfied and left to take classes at Columbia University. After
earning a bachelor’s degree (1939) in physics from the California Institute of
Technology, Munk convinced Harald Sverdrup, director of the Scripps
Institution of Oceanography at the University of California, Los Angeles, to give
him a summer job. By 1940 he had earned a master’s degree in geophysics from
the California Institute of Technology and by 1947 had completed a doctorate in
oceanography at Scripps. After graduation Scripps hired him as an assistant
professor of geophysics. He became a full professor there in 1954 and was made
a member of the University of California’s Institute of Geophysics.
Distressed by the 1938 occupation of Austria by Germany, Munk had applied
for U.S. citizenship and enlisted in the U.S. Army prior to completing his
doctorate. From 1939 to 1945 he joined several of his colleagues from Scripps at
the U.S. Navy Radio and Sound Laboratory, where they developed methods
related to amphibious warfare. Their method for predicting and dealing with
waves was carried out successfully by the Allied forces on D-Day (June 6, 1944)
during the Normandy invasion. In 1946 Munk helped to analyze the currents,
diffusion, and water exchanges at Bikini Atoll in the South Pacific, where the
United States was testing nuclear weapons. Funded by a Guggenheim
Fellowship, he spent a portion of 1949 at the University of Oslo studying the
dynamics of ocean currents. Throughout the 1950s Munk studied the effect of
geophysical processes on the wobble in Earth’s rotation, publishing the seminal
results in The Rotation of the Earth: A Geophysical Discussion (with G.J.F.
MacDonald, 1960).
In 1959 Munk began campaigning for the creation of what would become the
Cecil H. and Ida M. Green Institute of Geophysics and Planetary Physics (IGPP)
at Scripps. He directed the institute until 1982. Munk worked on the Mid-Ocean
Dynamics Experiment (MODE) from 1965 to 1975; it resulted in significant
improvement in the accuracy of tide prediction. Waves Across the Pacific, a 1967
documentary, depicted his study of how waves generated by storms in the
Southern Hemisphere travel through the rest of the world’s oceans. In 1968 he
became a member of JASON, a panel of scientists who advised the U.S.
government.
Beginning in 1975, Munk had begun experimenting with the use of acoustic
tomography, which uses sound waves to generate images of water. This
culminated in the 1991 Heard Island experiment, in which sound signals were
transmitted from instruments 150 metres (492 feet) below the ocean’s surface to
receivers around the world. The project used the speed at which the signals
transmitted to measure the temperature of the water. He cowrote the definitive
volume on the subject, Ocean Acoustic Tomography (1995). Munk was named
Secretary of the Navy Research Chair in Oceanography in 1984 and continued to
research the implications of global warming for the oceans as part of the
Acoustic Thermometry of Ocean Climate (ATOC) project from 1996 to 2006.
Munk was the recipient of the Royal Astronomical Society’s 1968 Gold
Medal and the American Geophysical Union’s 1989 William Bowie Medal,
among others. He became the first recipient of the annual Walter Munk Award,
given in his honour by The Oceanography Society and several naval offices, in
1993. In 1999 he won the 15th annual Kyoto Prize in Basic Sciences for his
work in physical oceanography and geophysics; he became the first in his field
to be honoured with this award.

SIR JOHN MURRAY


(b. March 3, 1841, Cobourg, Ont., Can.—d. March 16, 1914, near Kirkliston,
West Lothian [now in Edinburgh], Scot.)

Scottish Canadian naturalist Sir John Murray was one of the founders of
oceanography. His particular interests were ocean basins, deep-sea deposits, and
coral-reef formation.
In 1868 Murray began collecting marine organisms and making a variety of
oceanographic observations during an expedition to the Arctic islands of Jan
Mayen and Spitsbergen, off Norway. Murray did much to organize the
Challenger Expedition (1872–76), which made extremely valuable contributions
in charting, surveying, and biological investigation, and he helped outfit it with
equipment for conducting oceanographic studies. As a naturalist with the
expedition, he was placed in charge of the biological specimens collected. Kept
at Edinburgh, they attracted the attention of marine biologists from around the
world for 20 years.
After the death of the expedition’s leader, Sir Wyville Thomson (1882),
Murray completed the publication of the 50-volume Report on the Scientific
Results of the Voyage of H.M.S. Challenger (1880–95). He also directed
biological investigations of Scottish waters (1882–94), surveyed the depths of
Scottish lakes (1906), and took part in a North Atlantic oceanographic
expedition (1910). He was knighted in 1898. His writings include the paper “On
the Structure and Origin of Coral Reefs and Islands” (1880) and, with Johan
Hjort, The Depths of the Ocean (1912).

FRIDTJOF NANSEN
(b. Oct. 10, 1861, Store-Frøen, near Kristiania [now Oslo], Nor.—d. May 13,
1930, Lysaker, near Oslo)

Fridtjof Nansen was a Norwegian explorer, oceanographer, statesman, and


humanitarian who led a number of expeditions to the Arctic (1888, 1893, 1895–
96) and oceanographic expeditions in the North Atlantic (1900, 1910–14). For
his relief work after World War I he was awarded the Nobel Prize for Peace
(1922).
Nansen’s success as an explorer was due largely to his careful evaluation of
the difficulties that might be encountered, his clear reasoning, which was never
influenced by the opinions of others, his willingness to accept a calculated risk,
his thorough planning, and his meticulous attention to detail. Many of these traits
can be recognized in his scientific writings. In 1882 he was appointed curator of
zoology at the Bergen museum. He wrote papers on zoological and histological
subjects, illustrated by excellent drawings. For one of his papers, “The Structure
and Combination of Histological Elements of the Central Nervous System”
(1887), the University of Kristiania conferred upon him the degree of doctor of
philosophy. Though the paper contained so many novel interpretations that the
committee that had to examine it accepted it with doubt, it is now considered a
classic.
Fridtjof Nansen, 1896. © Photos.com/Jupiterimages

On his return from the Fram expedition in 1896, a professorship in zoology


was established for Nansen at the University of Kristiania, but his interests
shifted from zoology to physical oceanography, and in 1908 his status was
changed to professor of oceanography. During 1896–1917 he devoted most of
his time and energy to scientific work. He edited the report of the scientific
results of his expedition and himself wrote some of the most important parts. He
participated in the establishment of the International Council for the Exploration
of the Sea and for some time directed the council’s central laboratory in
Kristiania. In 1900 he joined the Michael Sars on a cruise in the Norwegian Sea.
In 1910 he made a cruise in the Fridtjof through the northeastern North Atlantic;
in 1912 he visited the Spitsbergen waters on board his own yacht Veslemoy; and
in 1914 he joined B. Helland-Hansen on an oceanographic cruise to the Azores
in the Armauer Hansen. In 1913 Nansen traveled through the Barents Sea and
the Kara Sea to the mouth of the Yenisey River and back through Siberia. He
published the results of his cruises in numerous papers, partly in cooperation
with Helland-Hansen. His lasting contributions to oceanography comprise
improvement and design of instruments, explanation of the wind-driven currents
of the seas, discussions of the waters of the Arctic, and explanation of the
manner in which deep-and bottom-water is formed.
Nansen also dealt with other subjects: for instance, his Nord i tåkeheimen, 2
vol. (1911; In Northern Mists) gave a critical review of the exploration of the
northern regions from early times up to the beginning of the 16th century.

JACQUES PICCARD
(b. July 28, 1922, Brussels, Belg.—d. Nov. 1, 2008, La Tour-de-Peilz, Switz.)

Jacques-Ernest-Jean Piccard was a Swiss oceanic engineer, economist, and


physicist, who helped his father, Auguste Piccard, build the bathyscaphe for
deep-sea exploration and who also invented the mesoscaphe, an undersea vessel
for exploring middle depths.
Jacques Piccard was born in Brussels while his Swiss-born father was a
professor at the University of Brussels. After graduating from the École
Nouvelle de Suisse Romande in Lausanne, Switz., in 1943, he studied at the
University of Geneva, taking a year off in 1944–45 in order to serve with the
French First Army. Upon receiving his licentiate in 1946, he taught at the
university for two years before entering private teaching.
Meanwhile, he was helping his father to design bathyscaphes and in 1953
accompanied him in the Trieste on a dive of 10,168 feet (3,099 m) off the island
of Ponza, Italy. In 1956 Jacques Piccard went to the United States seeking
funding; two years later the U.S. Navy bought the Trieste and retained him as a
consultant. On Jan. 23, 1960, he and Lieutenant Don Walsh of the U.S. Navy set
a new submarine depth record by descending 10,916 metres (35,810 feet) into
the Mariana Trench in the Pacific Ocean using the Trieste. He recounted this feat
in Seven Miles Down (1961), written with Robert Dietz. In the early 1960s,
working with his father, he designed and built the first of four mesoscaphes. His
first mesoscaphe, the Auguste Piccard, capable of carrying 40 passengers,
transported some 33,000 tourists through the depths of Lake Geneva during the
1964 Swiss National Exhibition in Lausanne. In 1969 he drifted some 3,000 km
(1,800 miles) along the east coast of North America in the mesoscaphe Ben
Franklin, conducting research on the Gulf Stream for the U.S. Navy.
In his later career Piccard was a consultant scientist for several private
American organizations for deep-sea research, including the Grumman Aircraft
Engineering Corporation, New York (1966–71). In the 1970s he founded the
Foundation for the Study and Protection of Seas and Lakes, based in Cully,
Switz. In 1999 his son Bertrand Piccard, together with Englishman Brian Jones,
completed the first nonstop circumnavigation of the globe in a balloon.

FRANCIS P. SHEPARD
(b. May 10, 1897, Brookline, Mass., U.S.—d. April 25, 1985, La Jolla, Calif.)

Francis Parker Shepard was an American marine geologist whose pioneering


surveys of submarine canyons off the coast of California near La Jolla marked
the beginning of Pacific marine geology.
Shepard studied geology at Harvard under R.A. Daly and at the University of
Chicago (Ph.D., 1922). Most of Shepard’s professional life was spent at the
Scripps Institution of Oceanography, where he was appointed professor of
submarine geology in 1948. Central to his research was the search for an
explanation of the origin of submarine canyons. His series of observations of
depth changes (begun in the early 1950s) at the head of submarine canyons gave
concrete evidence, although not positive proof, of their formation by turbidity
currents (submarine flows of muddy suspensions) and sediment slumping.
Among his principal works are Submarine Geology (1948) and, with R.F. Dill,
Submarine Canyons and Other Sea Valleys (1966).

HENRY MELSON STOMMEL


(b. Sept. 27, 1920, Wilmington, Del., U.S.—d. Jan. 17, 1992, Boston, Mass.)

American oceanographer and meteorologist Henry Melson Stommel became


internationally known during the 1950s for his theories on circulation patterns in
the Atlantic Ocean. Stommel suggested that Earth’s rotation is responsible for
the Gulf Stream along the coast of North America, and he theorized that its
northward flow must be balanced by a stream of cold water moving southward
beneath it. He proposed a global circulation in which surface water sinks in the
far north to feed the deep, south-flowing current, while water rises in the
Antarctic region to supply a northward flow along the eastern coasts of North
and South America. Much of this “conveyor belt” theory has been confirmed. In
addition to his work on ocean currents, Stommel did research on a variety of
problems in oceanography and meteorology.
An anomaly among modern scientists, Stommel became a full professor
without an earned doctorate. He received his B.S. from Yale University (1942)
and served there as instructor in mathematics and astronomy (1942–44). A
research associate at the Woods Hole (Mass.) Oceanographic Institution from
1944 to 1959, he became professor of oceanography at the Massachusetts
Institute of Technology in 1959 and remained there except for the years 1960–
63, when he taught at Harvard. Stommel established several stations to study
ocean currents, including the PANULIRUS station (begun in 1954) in Bermuda.
He was elected to the National Academy of Sciences in 1962 and received the
National Medal of Science in 1989.

GEORG WÜST
(b. June 15, 1890, Posen, Ger. [now Poznán, Pol.]—d. Nov. 8, 1977, Erlangen,
W. Ger.)

Georg Adolf Otto Wüst was a German oceanographer who, by collecting and
analyzing many systematic observations, developed the first essentially complete
understanding of the physical structure and deep circulation of the Atlantic
Ocean.
Wüst received his doctorate from the University of Berlin in 1919. After the
death of his teacher Alfred Merz, Wüst took over as chief oceanographer on the
German Atlantic (1925–27) expedition. He was also in charge of the
International Gulf Stream (1938) expedition. The Atlantic expedition, conducted
from the research vessel Meteor, was the first study of an entire ocean, and it
remains one of the most extensive oceanographic surveys ever undertaken. From
the wealth of data amassed, Wüst constructed cross-sectional profiles that
revealed the Atlantic’s complex temperature and salinity stratification and its
deep-current structure.
After World War II Wüst built up again the Institute for Oceanography, at
Kiel, so that it flourished as a research centre. He was the institute’s director
from 1946 until he retired in 1959.
APPENDIX
GLOSSARY

bathymetry The measurement of ocean depth.


benthos A collection of organisms that live on or at the bottom of a body of
water.
capillary waves Tiny waves at the ocean surface that are direct products of the
wind stress exerted on the sea surface.
desalination A water treatment method that involves the separation of
freshwater from salt water or brackish water.
divergence A meteorological term used to describe the drawing apart of air, as
well as the rate at which each takes place.
downwelling A region of oceanic convergence that forces surface water
downward.
eddy A current in a body of water such as a river that swirls in a different
direction than the main current.
El Niño The anomalous appearance, every few years, of unusually warm ocean
conditions along the tropical west coast of South America associated with
adverse effects on fishing, agriculture, and weather systems.
Gulf Stream A warm ocean current flowing in the North Atlantic northeastward
off the North American coast, which produces a warming effect upon the
climates of adjacent land areas.
gyre A system of ocean currents that spirals around a central zone.
hydrologic cycle The continuous circulation of water through Earth’s
atmosphere, surface, and oceans.
isobaric characterized by constant or equal pressure.
Kuroshio Current A strong surface oceanic current of the northwestern Pacific
Ocean that dramatically affects atmospheric conditions on Earth.
nekton The assemblage of pelagic animals that swim freely, independent of
water motion or wind.
phytoplankton A flora of freely floating, often minute photosynthetic
organisms that drift with water currents.
seiche A rhythmic oscillation of water in a lake or a partially enclosed coastal
inlet, such as a bay, gulf, or harbour.
swell A long often massive and crestless wave or succession of waves often
continuing beyond or after its cause, such as a gale wind.
Southern Oscillation A coherent interannual fluctuation of atmospheric
pressure over the tropical Indo-Pacific region.
thermocline A layer in a thermally stratified body of water that separates an
upper, warmer, lighter, oxygen-rich zone from a lower, colder, heavier,
oxygen-poor zone.
thermohaline circulation The component of general oceanic circulation that is
controlled by horizontal differences in temperature and salinity.
tidal bore A surge of swiftly moving water traveling upstream and headed by a
wave or series of waves.
transpiration A plant’s loss of water, mainly through the stomates of leaves.
tsunamis Catastrophic ocean waves usually caused by submarine earthquakes,
underwater or coastal landslides, or volcanic eruptions.
upwelling The process of upward movement to the ocean surface of deeper,
cold, usually nutrient-rich waters especially along some shores due to the
offshore movement of surface waters.
BIBLIOGRAPHY

GENERAL CONSIDERATIONS
Broad overviews of the oceans are provided by David A. Ross, Introduction to
Oceanography, 4th ed. (1988); Alyn C. Duxbury and Alison B. Duxbury, An
Introduction to the World’s Oceans, 3rd ed. (1991); Ellen J. Prager and Sylvia A.
Earle, The Oceans (2001); Sylvia A. Earle, The World Is Blue: How Our Fate
and the Ocean’s Are One (2009); Deborah Cramer, Smithsonian Ocean: Our
Water, Our World (2008); and M. Grant Gross, Oceanography, a View of the
Earth, 5th ed. (1990). Alastair Couper (ed.), The Times Atlas and Encyclopedia
of the Sea (1989), provides a graphic look at all aspects of the ocean. Ocean
Yearbook (annual), contains essays on resources, transportation, and marine
science, among other topics.
CHEMICAL AND PHYSICAL PROPERTIES OF SEAWATER
As an excellent starting point for the reader interested in an integrated account of
ocean chemistry, physics, and biology, the classic work by H.U. Sverdrup,
Martin W. Johnson, and Richard H. Fleming, The Oceans (1942, reissued 1970),
is highly recommended. An in-depth, but quite readable, account of the general
field of marine chemistry is provided by J.P. Riley and R. Chester, Introduction
to Marine Chemistry (1971). Kenneth W. Bruland, “Trace Elements in Sea-
water,” vol. 8, ch. 45 (1983), pp. 157–220, is another readable account. Modern
treatments include Anthony Gianguzza, Ezio Pelizzetti, and Silvio Sammartano
(eds.), Chemistry of Marine Water and Sediments (2002); R.E. Hester and R.M.
Harrison (eds.), Chemistry in the Marine Environment (2000); and Anthony
Gianguzza, Ezio Pelizzetti, and Silvio Sammartano (eds.), Chemical Processes
in Marine Environments (2000).

OCEAN BASINS AND CONTINENTAL MARGINS


Overviews of the geologic features of the deep-sea floor are given in Eugen
Seibold and Wolfgang H. Berger, The Sea Floor: An Introduction to Marine
Geology (2010); H. Kuenen, Marine Geology (2007); and Alan E.M. Nairn and
Francis G. Stehli (eds.), The Ocean Basins and Margins, 7 vol. in 9 (1973–88).
Specific topics and geographic areas are studied by C.M. Powell, S.R. Roots,
and J.J. Veevers, “Pre-breakup Continental Extension in East Gondwanaland and
the Early Opening of the Eastern Indian Ocean,” Tectonophysics, 155:261–283
(1988); David B. Rowley and Ann L. Lottes, “Reconstructions of the North
Atlantic and Arctic: Late Jurassic to Present,” Tectonophysics, 155:73–120
(1988); T. Simkins et al., (1989); and two essays in E.L. Winterer, Donald M.
Hussong, and Robert W. Decker (eds.), The Eastern Pacific Ocean and Hawaii
(1989), vol. N of the series “The Geology of North America”: Tanya Atwater,
“Plate Tectonic History of the Northeast Pacific and Western North America,”
ch. 4, pp. 21–72; and Ken C. MacDonald, “Tectonic and Magnetic Processes on
the East Pacific Rise,” ch. 6, pp. 93–110.

CIRCULATION OF OCEAN WATERS


Useful books include those by Open University Oceanography Course Team,
Ocean Circulation (1989); George L. Pickard and William J. Emery, Descriptive
Physical Oceanography: An Introduction, 5th enlarged ed. (1990); Rui Xin
Huang, Ocean Circulation: Wind-Driven and Thermohaline Processes (2009);
Stephen Pond and George L. Pickard, Introductory Dynamical Oceanography,
2nd ed. (1983); Henry Stommel, A View of the Sea (1987); and Henry Stommel
and Dennis W. Moore, An Introduction to the Coriolis Force (1989). Joseph
Pedlosky, Ocean Circulation Theory (2010), is an advanced treatment.

WAVES OF THE SEA


Books discussing waves and tides include Joseph Pedlosky, Waves in the Ocean
and Atmosphere: Introduction to Wave Dynamics (2010); Open University
Oceanography Course Team, Waves, Tides, and Shallow-water Processes (1989);
Albert Defant, Ebb and Flow: The Tides of Earth, Air, and Water (1958;
originally published in German, 1953); Blair Kinsman, Wind Waves: Their
Generation and Propagation on the Ocean Surface (1965, reprinted 1984); M.
Grant Gross, Oceanography, 5th ed. (1990), ch. 8, “Waves,” and ch. 9, “Tides,”
pp. 193–241; and David T. Pugh, Tides, Surges, and Mean Sea-level (1987). C.
Garrett and W. Munk, “Internal Waves in the Ocean,” Annual Review of Fluid
Mechanics, vol. 11, pp. 339–369 (1979), is also of interest. Highly technical
treatments of ocean waves are provided in Leo H. Holthuijsen, Waves in Oceanic
and Coastal Waters (2010); and Michel K. Ochi, Ocean Waves: The Stochastic
Approach (2007).

DENSITY CURRENTS IN THE OCEANS


An introductory treatment to the topic of density currents is provided in Marius
Ungarish, An Introduction to Gravity Currents and Intrusions (2009). Other
studies include Gerard V. Middleton and Monty A. Hampton, “Subaqueous
Sediment Transport and Deposition by Sediment Gravity Flows,” ch. 11 in
Daniel Jean Stanley and Donald J.P. Swift (eds.), Marine Sediment Transport
and Environmental Management (1976), pp. 197–218, which describes and
discusses turbidity currents, grain flows, fluidized sediment flows, and debris
flows; Gary Parker, Yusuke Fukushima, and Henry M. Pantin, “Self-accelerating
Turbidity Currents,” Journal of Fluid Mechanics, 171:145–181 (1986); and
Richard J. Seymour, “Nearshore Auto-suspending Turbidity Flows,” Ocean
Engineering, 13(5):435–447 (1986).

IMPACT OF OCEAN-ATMOSPHERE INTERACTIONS ON


WEATHER AND CLIMATE
Overviews of the general relationship between air and ocean may be found in
John Marshall and R. Alan Plumb, Atmosphere, Ocean and Climate Dynamics:
An Introductory Text (2007); Grant R. Bigg, The Oceans and Climate (2004); Y.
Toba (Ed.), Ocean-Atmosphere Interactions (2003); Adrian E. Gill, Atmosphere-
Ocean Dynamics (1982); and Neil Wells, The Atmosphere and Ocean: A
Physical Introduction (1986).
Studies of the El Niño/Southern Oscillation phenomena and their effect on
climatic change are found in P.W. Glynn (ed.), Global Ecological Consequences
of the 1982–83 El Nino-Southern Oscillation (1990); S. George Philander, El
Niño, La Niña, and the Southern Oscillation (1990); Warren S. Wooster and
David L. Fluharty (eds.), El Niño North: Niño Effects in the Eastern Subarctic
Pacific Ocean (1985); Brian Fagan, Floods, Famines, and Emperors: El Niño
and the Fate of Civilizations (2009); Richard T. Barber and Francisco P. Chavez,
“Biological Consequences of El Niño,” Science, 222(4629):1203–1210 (Dec. 16,
1983); Thomas Y. Canby, “El Niño’s Ill Wind,” National Geographic,
165(2):144–183 (February 1984); M.A. Cane, “El Niño,” Annual Review of
Earth and Planetary Sciences, 14:43–70 (1986); David B. Enfield, “El Niño,
Past and Present,” Reviews of Geophysics, 27(1):159–187 (1989); Nicholas E.
Graham and Warren B. White, “The El Niño Cycle: A Natural Oscillator of the
Pacific Ocean-Atmosphere System,” Science, 240:1293–1302 (June 3, 1988); S.
George Philander and E.M. Rasmusson, “The Southern Oscillation and El
Niño,” Advances in Geophysics, vol. 28, part A, pp. 197–215 (1985); and E.M.
Rasmusson, “El Niño and Variations in Climate,” American Scientist, 73(2):168–
177 (March–April 1985). In addition, the entire issue of Oceanus, vol. 27, no. 2
(Summer 1984), is devoted to El Niño studies.

ECONOMIC ASPECTS OF THE OCEANS


An entire issue of Oceanus, vol. 21, no. 4 (Winter 1984/85), is devoted to the
impact of the Exclusive Economic Zone, especially as it pertains to the United
States; coastal fishing, multiple-use management, marine pollution, and
nonliving resources are some of the topics covered. Giulio Pontecorvo, The New
Order of the Oceans: The Advent of a Managed Environment (1986), deals with
the new ocean regime and with the research and technology of marine resources
from a global perspective, particularly emphasizing the international effects of
the 1982 United Nations Law of the Sea Convention. The evolution of
international marine policy and shipping law is compellingly discussed in Edgar
Gold, Maritime Transport (1981), and “Ocean Shipping and the New Law of the
Sea: Toward a More Regulatory Regime,” Ocean Yearbook, vol. 6, pp. 85–96
(1986).
Discussions of the managed production of aquatic organisms include J.F.
Muir, “Aquaculture—Towards the Future,” Endeavour, 9(1):52–55 (1985); and
John Bardach, “Aquaculture: Moving from Craft to Industry,” Environment,
30(2):6–11, 36–41 (March 1988). Desalinization processes are discussed and
illustrated in Alan D.K. Laird, “The Potable Sea: Taking the Salt from
Saltwater,” Oceans, 15(5):25–29 (September–October 1982); and Roberta
Friedman, “Salt-free Water from the Sea,” Sea Frontiers, 36(3):49–54 (May–
June 1990).
Overviews of various ocean energy resources and technologies may be found
in Maxwell Bruce, “Ocean Energy: Some Perspectives on Economic Viability,”
Ocean Yearbook, vol. 5, pp. 58–78 (1985); and Terry R. Penney and Desikan
Bharathan, “Power from the Sea,” Scientific American, 256(1):86–92 (January
1987). Tidal power generation is covered by B. Count (ed.), Power from Sea
Waves (1980), based on conference proceedings; and Michael E. McCormick,
Ocean Wave Energy Conversion (1981). Summaries of ocean thermal energy
conversion techniques and future prospects are provided by R. Cohen, “Energy
from the Ocean,” Philosophical Transactions of the Royal Society of London,
Series A, 307:405–437 (1982); and D.E. Lennard, “Ocean Thermal Energy
Conversion—Past Progress and Future Prospects,” IEE Proceedings, vol. 134,
part A, no. 5, pp. 381–391 (May 1987).
Current and future petroleum and mineral resources in the ocean
environment and the technologies necessary to recover them are addressed in
Gerard J. Mangone (ed.), The Future of Gas and Oil from the Sea (1983);
Elisabeth Mann Borgese, The Mines of Neptune: Minerals and Metals from the
Sea (1985); David Cronan, “A Wealth of Sea-floor Minerals,” New Scientist,
106:34–38 (June 6, 1985); and James M. Broadus, “Seabed Materials,” Science,
235:853–860 (Feb. 20, 1987).
Discussions of the use of the ocean as a site for waste disposal, and the
problems of marine pollution, include Iver W. Duedall et al. (eds.), Wastes in the
Ocean, 6 vol. (1983–85), on industrial, sewage, radioactive, and energy wastes,
dredged-material disposal, and deep-sea and nearshore waste disposal; R.B.
Clark, Marine Pollution, 2nd ed. (1989); Wesley Marx, The Oceans: Our Last
Resource (1981); David K. Bulloch, The Wasted Ocean (1989); and a complete
issue of Oceanus, vol. 33, no. 2 (Summer 1990).
UNDERSEA EXPLORATION
The classical tools of the oceanographer are described by H.U. Sverdrup, Martin
W. Johnson, and Richard H. Fleming, The Oceans (1942, reissued 1970). More
recent texts on oceanography and marine biology include Tom S. Garrison,
Oceanography: An Invitation to Marine Science, 7th ed. (2009); Alan P. Trujillo
and Harold V. Thurman, Essentials of Oceanography, 10th ed. (2010); John A.
Knauss, Introduction to Physical Oceanography, 2nd ed. (2005); M. Grant
Gross, Oceanography, a View of the Earth, 5th ed. (1990); George L. Pickard
and William J. Emery, Descriptive Physical Oceanography, 5th enlarged ed.
(1990); James P. Kennett, Marine Geology (1982); Bruce A. Warren and Carl
Wunsch (eds.), Evolution of Physical Oceanography (1981); James L. Sumich,
An Introduction to the Biology of Marine Life, 5th ed. (1992); M.N. Hill et al.
(eds.), The Sea (1962); and J.P. Riley and G. Skirrow (eds.), Chemical
Oceanography, 2nd ed., 8 vol. (1975–83).
Exploration and discovery are chronicled in Rachel L. Carson, The Sea
Around Us, special ed. (1989); Wolf H. Berger and E.N. Shor, Ocean:
Reflections on a Century of Exploration (2009); and Martin W. Sandler, Atlantic
Ocean: The Illustrated History of the Ocean That Changed the World (2008).
Also informative are the documentaries of famous explorers, such as the books
by William Beebe, Half Mile Down (1934, reissued 1951); Jacques Piccard and
Robert S. Dietz, Seven Miles Down (1961); George Stephen Ritchie, Challenger:
The Life of a Survey Ship (1957); Helen Raitt, Exploring the Deep Pacific, 2nd
ed. (1964); Willard Bascom, A Hole in the Bottom of the Sea: The Story of the
Mohole Project (1961); Francis P. Shepard, The Earth Beneath the Sea, rev. ed.
(1967); and Daniel Behrman, The New World of the Oceans (1969).
Other treatments include Jacques Cousteau and Alexis Siverine, Jacques
Cousteau’s Calypso (1983; originally published in French, 1978); R. Frank
Busby, Manned Submersibles (1976), on their design and operation; and
Kenneth J. Hsü, Challenger at Sea, trans. from German (1992).
Modern undersea explorers are strongly dependent on innovative engineering
and technology, such as that treated in Jane’s Ocean Technology, 1979–80
(1979); John J. Myers (ed.), Handbook of Ocean and Underwater Engineering
(1969); John F. Brahtz (ed.), Ocean Engineering (1968); and Robert L. Wiegel,
Oceanographical Engineering (1964). Other treatments include John Brackett
Hersey (ed.), Deep-Sea Photography (1967); J.F.R. Gower (ed.), Oceanography
from Space (1981); G.A. Maul, Introduction to Satellite Oceanography (1985);
and F. Dobson, L. Hasse, and R. Davis (eds.), Air-Sea Interaction: Instruments
and Methods (1980).
INDEX

A
abyssal hills, 49, 52, 63–64, 102
abyssal plains, 70, 80, 193
Agassiz, Alexander, 226–227
Albert I, Prince of Monaco, 188, 227–228
Antarctica, 2, 3, 31, 58, 59, 60, 80, 91, 96, 99, 113, 120, 123, 191
Antarctic Bottom Water, 31, 123
Antarctic Circumpolar Current, 118–119, 122, 123, 124
Antarctic Intermediate Water, 122, 123, 124
Arctic Ocean, 2, 78, 113
Atlantic Ocean, 60, 64, 65, 79, 95–96, 116, 117, 121, 122, 142, 144, 148, 149,
151, 154, 175, 203, 229
atmosphere, 1, 2, 9, 10, 15, 16, 19, 28, 29, 37, 40, 41, 42, 43, 44, 45, 104, 105,
118, 119, 140, 141, 142, 143, 144, 147, 153, 155, 161, 162, 164, 180, 191,
195, 198, 218, 225
atmospheric circulation, 5, 140, 142, 144, 147–148, 149, 150

B
Ballard, Robert, 226, 228–230
baroclinic field of mass, 107, 114
bathymetry, 188, 190, 193–194, 205
Beagle, 188
Beaufort, Sir Francis, 126
Beaufort scale, 126
Beer’s law, 34, 35, 36
benthos, 218
Blake Plateau, 68, 76, 90

C
Cambrian Period, 9
capillary waves, 125
carbon dioxide, 9, 10, 15, 16, 19, 40, 41, 42, 44, 45, 46, 156, 198, 199
Challenger (HMS), 52, 188, 189, 198, 206, 242
Challenger Expedition, 52, 99, 206, 242
chemical evolution of oceans
early oceans, 28, 39–44
modern oceans, 47–48, 122
the transition stage, 44–47
chemosynthesis, 103, 229
chimneys, 121
chlorofluorocarbons, 119, 156
circulation of the ocean waters, 104–105
climate anomalies, 141–144
climate change, 155, 163
composition of seawater
dissolved inorganic substances, 11–17
dissolved organic substances, 17–19
effects of human activities, 19–20
continental drift, 92, 197, 199
continental margins, 49, 62, 64–70, 72, 74, 75, 78, 79, 90, 117, 123
economic importance of, 69–70
margin types, 65–69
continental rises, 70–71
continental shelves, 71–75
origins of, 72–75
structure of, 71–72
continental slopes, 8, 63, 64, 65, 67, 70, 72, 73, 76–77, 90, 117, 148, 149, 152,
171, 193
convergence, 22, 105, 113, 114, 116, 129, 141, 146
convergence zones, 141
coral reefs, 188, 242
Coriolis, Gustrave-Gaspard, 108
Coriolis force, 106, 108, 109, 110, 111, 112, 115, 116, 150, 151, 233
Cousteau, Jacques-Yves, 226, 230–231
currents, 3, 24, 52, 55, 56, 57, 58, 64, 67, 68, 70, 71, 74, 77, 82, 104, 105–116,
119, 134, 139, 146, 147, 148, 149, 150, 180, 187, 188, 189, 191–193, 202,
203, 206, 211–214, 215, 225, 233, 234, 235, 236, 237, 238, 239, 240, 244,
246, 247

D
Darwin, Charles, 160, 161, 188, 227
dead reckoning, 105, 203, 204
deep-sea drilling, 56, 89, 180, 223
deep-sea sediments
patterns of, 57–58
types of, 56–57
deep-sea trenches, 49, 50, 54, 61, 62, 63, 64, 65, 76, 77–83, 88, 202
origin of, 82–83
structure of, 80–81
types of, 78–80
Deepwater Horizon disaster, 172–173, 179
density currents, 111, 254–255
desalinization, 139, 167, 168, 190, 196–197
Dietz, Robert S., 88, 232–233, 245
divergence, 52, 105, 113, 114, 115, 117, 128, 129
downwelling, 104, 113, 114, 147

E
East Pacific Rise, 83, 90, 92, 94, 99, 100, 101, 102, 103, 176
Ekman, Vagn Walfrid, 112, 188, 233–235
Ekman layer, 112–114, 233
Ekman transport, 113, 114, 115, 116, 117, 233
electromagnetic radiation, 32, 221
El Nino effect, 116, 139, 142, 155–160, 161, 162, 163, 164, 198
energy resources, 139, 168–170
equatorial currents, 115–116
evaporation, 2, 11, 21, 22, 27, 28, 37, 119, 121, 141, 142, 146, 156, 168, 185–
186, 191, 195
Exxon Valdez, 179

F
fertilizers, 19, 157
Fleming, Richard H., 235–236
freezing point, 26, 27, 30
frictional forces, 106, 109–110, 112, 223

G
geostrophic currents, 110–111, 114, 151
Gibraltar, Strait of, 57, 111, 121, 191
glaciers, 1, 2, 51, 66, 174
Great Barrier Reef, 227
greenhouse gases, 156, 199
Greenland, 60, 118, 121, 154
groundwater, 1, 2, 185, 195
Gulf Stream, 24, 116, 117, 139, 148–152, 154, 155, 186, 191, 192, 215, 234,
245, 247, 248
gyres, 105, 114, 116–118

H
Halley, Edmond, 185
heat, 4, 9, 21, 25, 26, 27, 28, 31, 32, 33, 36, 39, 82, 93, 102, 104, 105, 109, 116,
118, 123, 139, 140, 141, 143, 144, 146, 147, 156, 164, 169, 170, 179, 182,
183, 191, 195, 198, 206, 207, 225
poleward transfer of, 119, 153–155
Helland-Hansen, Bjorn, 188, 236, 244
Henry the Navigator, 183–184
horse latitudes, 22
hydrologic cycle, 2, 182–183, 185, 186, 187, 195
hydrothermal vents, 54, 101, 102–103, 229
hypsometry, 4, 5

I
ice ages, 66, 155, 174
Indian Ocean, 2, 5, 59, 60, 61, 76, 79, 96–99, 116, 117, 122, 141, 153, 175, 203,
239
internal waves, 70, 132–134, 152
intertropical convergence zones, 22, 146
Iselin, Columbus O’Donnell, 236–237
island arcs, 60, 61, 79, 80, 81, 86, 100

K
Kelvin waves, 162, 163
Kuroshio Current, 24, 117, 139, 148–153, 154
L
Labrador Sea, 60, 61, 118, 121
lithospheric plates, 50, 61, 75, 78, 82, 123
lunar tide, 153

M
magma chambers, 85, 103
magnetic field, 56, 82, 87, 89, 217, 221
Mariana Trench, 7, 61, 62, 77, 79, 245
marine biology, 196, 199
marine geology, 54, 180, 197, 229, 246
marine magnetic anomalies, 63, 86, 87–90, 93, 99, 101
Maury, Matthew Fontaine, 95, 186, 187, 237–239
Mediterranean Sea, 57, 62, 72, 86, 111, 112, 121, 137, 183, 185, 186, 191, 196,
203
Mid-Atlantic Ridge, 52, 64, 94, 95, 102, 229
Mid-Indian Ridge, 81, 97, 98, 99
minerals, 173–178
momentum, 105, 109, 113, 191
Moon, 125, 134, 135, 183, 190
Munk, Walter H., 114, 193, 239–241

N
Nansen, Fridtjof, 206, 210, 226, 234, 235, 242–244
Nansen bottle, 205, 206, 210
neovolcanic zones, 64, 101, 102, 103
North Atlantic Central Water, 148, 149, 150
North Atlantic Deep Water, 120, 121, 122, 124, 148, 154
Northern Hemisphere, 4, 5, 24, 109, 110, 112, 113, 115, 120, 142, 145, 146, 148
North Pacific Deep Water, 122

O
ocean-atmospheric interactions, 139–140
ocean basins, 25, 41, 49–55, 49, 50, 51, 52–55, 56, 57, 58–62, 63, 66, 76, 81, 89,
92, 94, 95, 111, 180, 188, 189, 202, 205, 242
oceanic crust
investigations of, 85–86
and marine magnetic anomalies, 87–90
and the study of ophiolites, 86–87
oceanic explorations, 52–55
oceanic plateaus, 90–92
oceanic ridges
distribution of major ridges and spreading centres, 95–101
principal characteristics of, 93–95
spreading centres and associated phenomena, 101–103
oceanography
ancient oceanography, 182–183
early oceanography, 183–184
in the 19th century, 186–190
in the 20th century, 190–197
ocean temperature, 141–144
ocean tides, 134–138
ocean trenches, 49, 50, 54, 61, 62, 63, 64, 65, 76, 77–83, 88, 202
offshore drilling, 171, 179
ophiolites, 86–87, 103

P
Pacific basin, 45, 55, 56, 58, 59, 61, 62, 79, 193
Pacific Ocean, 57, 58, 64, 65, 75, 77, 82, 90, 99–101, 122, 130, 139, 141, 144,
152, 154, 156, 175, 177, 215, 232, 245
Pangea, 58
pesticides, 179
petroleum, 69, 139, 171–173, 179, 225
photochemical transformation, 18
photodissociation, 9, 10, 43
photosynthesis, 10, 33, 35, 102
Physical Geography of the Sea, 187, 239
phytoplankton, 19, 35, 147, 218
Piccard, Jacques, 232, 244–246
plate tectonics, 50, 58, 75, 86, 199
plates
Farallon, 59, 62
Izanagi, 59
North American, 61, 68, 99
Pacific, 55, 58, 59, 61, 68, 99
Phoenix, 59
platforms, 200–203, 207, 208, 212, 213
Pleistocene Epoch, 51, 66, 174
pollutants, 180, 209
precipitation, 2, 11, 21, 22, 45, 46, 57, 139, 150, 182, 185, 191, 195
pressure gradients, 104, 107–108

R
radiation
electromagnetic, 32, 221
solar, 1, 4, 23, 33, 35, 36, 169, 170
rainfall, 141, 147–148, 156, 157, 160, 185, 198, 225
rainwater, 185, 186, 195
refraction, 32, 36, 38, 85, 86, 129
reservoirs, 2, 19, 139, 140, 164, 182, 195
Rossby waves, 163
Ross Sea, 118, 123

S
salinity, 11, 13, 15, 20–22, 24, 25, 27, 28, 29, 30, 31, 32, 37, 38, 42, 43, 104,
107, 111, 112, 119, 121, 122, 124, 140, 148, 150, 152, 191, 194, 198, 205–
211, 248
Sargasso Sea, 113, 148, 152
satellites, 54, 55, 180, 192, 200, 204, 205, 214–217
seafloor, 8, 10, 11, 21, 51–52, 53, 56, 57, 58, 59, 60, 61, 62, 63, 72, 74, 80, 81,
83, 84, 85, 88, 89, 90, 92, 93, 94, 99, 101, 102, 103, 105, 115, 119, 121,
122, 165, 171, 173, 174, 177, 178, 180, 187, 190, 191, 193, 194, 197, 199,
202, 214, 215, 221–224, 230, 232
sea level, 5, 7, 30, 49, 51–52, 57, 64, 65, 66, 67, 72, 73, 74, 75, 77, 86, 90, 114,
115, 125, 132, 137, 151, 161, 162, 163, 174, 185, 193, 227
Seasat, 54, 55
seawater
chemistry of, 195–196
composition of, 10–20
physical properties of, 20–39
Secchi disk, 34, 220
seiche/standing waves, 124, 132–133, 134
seismic waves, 85, 103
Shepard, Francis P., 246
solar radiation, 1, 4, 23, 33, 35, 36, 169, 170
solar-tidal forces, 135
sonar, 38, 52, 53, 190, 191, 194, 213, 216–217
Southern Hemisphere, 2, 4, 5, 22, 24, 59, 109, 110, 111, 112, 113, 115, 118, 145,
241
Southern Oscillation, 115, 116, 139, 142, 155, 160–163
standing/seiche waves, 124, 132–133, 134
Stommel, Henry Melson, 116, 120, 246–247
storm tides, 125
subarctic regions, 24, 139
subduction zones, 51, 59, 60, 75, 78, 79, 80, 81, 83
sublimation, 27
subpolar gyres, 114, 117, 118
subterranean water, 182, 184–185
subtropical gyres, 114, 116–117, 118
sulfur-fixing bacteria, 103
Sun, 1, 2, 23, 33, 35, 36, 102, 125, 134, 135, 190, 221
surface gravity waves, 125–129
swell, 125, 128, 129, 131
T
tectonic plates, 50, 58, 65, 75, 77, 86, 94, 199
temperature
ambient, 102
anomalies, 140, 141, 142, 144
and circulation, 104, 111, 119
crustal, 40, 83
and density, 28–32
distribution, 23–26
ocean floor, 189, 194
and salinity, 32, 38, 39
sampling, 205–209, 210
surface water, 1, 4, 9, 116, 140, 142, 143, 144, 147, 154, 155, 163, 215
Tethys seaway, 58, 59, 60
thermal energy, 169–170, 192, 195
thermal properties, 26–28
thermocline, 25, 117, 122, 124, 147
thermohaline circulation, 119–124
thermoremanent magnetization (TRM), 89
tidal bores, 137–138, 183
tidal power, 168–169, 190, 196–197
tides
lunar tides, 153
tidal bores, 137–138, 183
tide-generating forces, 134–137
solar-tidal forces, 135
storm tides, 125
Titanic, 190, 228, 229
tropical cyclones, 139, 145–148, 150
Tropic of Cancer, 23
Tropic of Capricorn, 23
tropospheric circulation, 140, 142
tsunamis, 125, 129, 130, 131–132, 193
U
undersea exploration, 197–224
acoustic sensing, 214–217
basic elements of, 200–205
collection of biological samples, 218–220
exploration of the seafloor, 221–224
measurements of ocean currents, 211–214
methodology and instrumentation, 205–209
navigation, 203–205
platforms, 200–203
satellite sensing, 214–217
temperature and salinity, 205–209
water sampling, 205–211
underwater mountain chains, 49, 61, 65, 71, 73, 75, 76, 78, 81, 92, 93, 95, 99,
102, 184, 185, 194, 221, 229

V
volcanic activity, 61, 78, 80, 81, 101
volcanism, 21, 52, 63, 64, 101, 199

W
waste disposal, 139, 178, 195
water resources, 195–196
water vapour, 9, 20, 27, 28, 41, 42, 141, 170, 199
waves of seismic origin, 125, 129
wave surge, 125, 137, 147
weather, 139, 144, 148, 156, 180, 198, 203
Weddell Sea, 31, 123
wind-driven circulation, 114–119
wind surges, 125, 129
wind waves, 125, 128
Wüst, Georg, 247–248

You might also like