GEOCHEMISTRY: PATHWAYS AND PROCESSES
Second Edition
SECOND EDITION
GEOCHEMISTRY
Pathways and Processes
Harry Y. McSween, Jr.
University of Tennessee, Knoxville
Steven M. Richardson
Winona State University
Maria E. Uhle
University of Tennessee, Knoxville
Columbia University Press
|
NEW YORK
Columbia University Press
Publishers Since 1893
New York
Chichester, West Sussex
First edition © 1989 by Prentice-Hall, Inc.
Second edition © 2003 Harry Y. McSween, Jr.; Steven M. Richardson; and Maria E. Uhle
All rights reserved
This book was previously published by Prentice-Hall, Inc.
Library of Congress Cataloging-in-Publication Data
McSween, Harry Y.
Geochemistry : pathways and processes. — 2nd ed. / Harry Y. McSween, Jr., Steven M.
Richardson, Maria E. Uhle.
p.
cm.
Rev. ed. of: Geochemistry / Steven M. Richardson, c1989.
Includes bibliographical references and index.
ISBN 0-231-12440-6
1. Geochemistry. I. Richardson, Steven McAfee. II. Uhle, Maria E. III. Richardson,
Steven McAfee. Geochemistry. IV. Title.
QE515.R53 2003
551.9—dc21
2003051638
∞ Columbia University Press books are printed on permanent and durable acid-free paper.
Printed in the United States of America
10 9 8 7 6 5 4 3 2 1
For Sue, Cathy, and Mike
CONTENTS
Preface to the Second Edition
ONE
INTRODUCING CONCEPTS IN GEOCHEMICAL SYSTEMS
Overview
What Is Geochemistry?
Historical Overview
Beginning Your Study of Geochemistry
Geochemical Variables
Geochemical Systems
Thermodynamics and Kinetics
An Example: Comparing Thermodynamic and Kinetic Approaches
Notes on Problem Solving
Suggested Readings
Problems
TWO
1
1
1
1
3
4
4
5
5
9
10
10
HOW ELEMENTS BEHAVE
Overview
Elements, Atoms, and the Structure of Matter
12
12
12
Elements and the Periodic Table
12
The Atomic Nucleus and Isotopes
13
The Basis for Chemical Bonds: The Electron Cloud
18
Size, Charge, and Stability
20
Elemental Associations
Bonding
Perspectives on Bonding
22
24
24
Structural Implications of Bonding
26
Retrospective on Bonding
32
Summary
Suggested Readings
Problems
THREE
xv
32
32
34
A FIRST LOOK AT THERMODYNAMIC EQUILIBRIUM
Overview
Temperature and Equations of State
Work
The First Law of Thermodynamics
Entropy and the Second Law of Thermodynamics
Entropy and Disorder
35
35
35
37
37
39
42
vii
viii
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Reprise: The Internal Energy Function Made Useful
Auxiliary Functions of State
Enthalpy
44
The Helmholtz Function
45
Gibbs Free Energy
Cleaning Up the Act: Conventions for E, H, F, G, and S
Composition as a Variable
Components
Changes in E, H, F, and G Due to Composition
Conditions for Heterogeneous Equilibrium
The Gibbs-Duhem Equation
Summary
Suggested Readings
Problems
FOUR
45
47
48
48
51
52
53
53
54
55
HOW TO HANDLE SOLUTIONS
Overview
What Is a Solution?
56
56
56
Crystalline Solid Solutions
57
Amorphous Solid Solutions
60
Melt Solutions
60
Electrolyte Solutions
62
Gas Mixtures
62
Solutions That Behave Ideally
Solutions That Behave Nonideally
Activity in Electrolyte Solutions
The Mean Salt Method
The Debye-Hückel Method
Solubility
FIVE
42
44
62
66
68
69
69
71
The Ionic Strength Effect
74
The Common Ion Effect
74
Complex Species
75
Summary
Suggested Readings
Problems
76
76
77
DIAGENESIS: A STUDY IN KINETICS
Overview
What Is Diagenesis?
Kinetic Factors in Diagenesis
79
79
79
80
Diffusion
80
Advection
85
Kinetics of Mineral Dissolution and Precipitation
The Diagenetic Equation
Summary
87
92
92
Contents
Suggested Readings
Problems
SIX
Preservation by Sorption
94
94
94
96
96
98
98
Degradation in Oxic Environments
99
Diagenetic Alteration
99
Chemical Composition of Biologic Precursors
103
Carbohydrates
103
Proteins
103
Lipids
104
Lignin
Biomarkers
Application of Biomarkers to Paleoenvironmental Reconstructions
105
105
106
Alkenone Temperature Records
106
Amino Acid Racemization
107
Summary
Suggested Readings
Problems
SEVEN
92
93
ORGANIC MATTER AND BIOMARKERS:
A DIFFERENT PERSPECTIVE
Overview
Organic Matter in the Global Carbon Cycle
Organic Matter Production and Cycling in the Oceans
Fate of Primary Production: Degradation and Diagenesis
Factors Controlling Accumulation and Preservation
108
109
110
CHEMICAL WEATHERING:
DISSOLUTION AND REDOX PROCESSES
Overview
Fundamental Solubility Equilibria
111
111
111
Silica Solubility
111
Solubility of Magnesian Silicates
112
Solubility of Gibbsite
114
Solubility of Aluminosilicate Minerals
Rivers as Weathering Indicators
Agents of Weathering
Carbon Dioxide
Organic Acids
Oxidation-Reduction Processes
115
119
121
121
122
124
Thermodynamic Conventions for Redox Systems
124
Eh-pH Diagrams
127
Redox Systems Containing Carbon Dioxide
129
Activity-Activity Relationships: The Broader View
131
Summary
Suggested Readings
Problems
ix
133
134
136
x
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
EIGHT
THE OCEANS AND ATMOSPHERE AS A
GEOCHEMICAL SYSTEM
Overview
Composition of the Oceans
A Classification of Dissolved Constituents
Chemical Variations with Depth
Composition of the Atmosphere
Carbonate and the Great Marine Balancing Act
137
137
137
139
141
143
Some First Principles
143
Calcium Carbonate Solubility
148
Chemical Modeling of Seawater: A Summary
Global Mass Balance and Steady State in the Oceans
Examining the Steady State
151
152
152
How Does the Steady State Evolve?
154
Box Models
154
Continuum Models
158
A Summary of Ocean-Atmosphere Models
158
Gradual Change: The History of Seawater and Air
159
Early Outgassing and the Primitive Atmosphere
159
The Rise of Oxygen
164
Summary
Suggested Readings
Problems
NINE
137
166
166
168
TEMPERATURE AND PRESSURE CHANGES:
THERMODYNAMICS AGAIN
Overview
What Does Equilibrium Really Mean?
Determining When a System Is in Equilibrium
The Phase Rule
Open versus Closed Systems
Changing Temperature and Pressure
169
169
169
169
170
171
172
Temperature Changes and Heat Capacity
172
Pressure Changes and Compressibility
174
Temperature and Pressure Changes Combined
A Graphical Look at Changing Conditions: The Clapeyron Equation
Reactions Involving Fluids
Raoult’s and Henry’s Laws: Mixing of Several Components
Standard States and Activity Coefficients
Solution Models: Activities of Complex Mixtures
Thermobarometry: Applying What We Have Learned
Summary
Suggested Readings
Problems
176
176
177
179
179
181
182
184
185
186
Contents
TEN
PICTURING EQUILIBRIA: PHASE DIAGRAMS
Overview
Ḡ-X2 Diagrams
Derivation of T-X2 and P-X2 Diagrams
T-X2 Diagrams for Real Geochemical Systems
Simple Crystallization in a Binary System: CaMgSi2O6-CaAl2Si2O8
ELEVEN
188
188
190
191
192
193
Solid Solution in a Binary System: NaAlSi3O8- CaAl2Si2O8
194
Unmixing in a Binary System: NaAlSi3O8-KAlSi3O8
195
196
197
199
201
201
204
205
KINETICS AND CRYSTALLIZATION
Overview
Effect of Temperature on Kinetic Processes
Diffusion
Nucleation
206
206
206
208
210
Nucleation in Melts
211
Nucleation in Solids
214
Growth
Interface-Controlled Growth
Diffusion-Controlled Growth
Some Applications of Kinetics
→ Calcite: Growth as the Rate-Limiting Step
Aragonite ←
Iron Meteorites: Diffusion as the Rate-Limiting Step
216
216
218
219
220
220
Bypassing Theory: Controlled Cooling Rate Experiments
222
Bypassing Theory Again: Crystal Size Distributions
223
Summary
Suggested Readings
Problems
TWELVE
188
Formation of a Chemical Compound in a Binary System: KAlSi2O6-SiO2
Thermodynamic Calculation of Phase Diagrams
Binary Phase Diagrams Involving Fluids
P-T Diagrams
Systems with Three Components
Summary
Suggested Readings
Problems
223
224
226
THE SOLID EARTH AS A GEOCHEMICAL SYSTEM
Overview
Reservoirs in the Solid Earth
Composition of the Crust
227
227
227
227
Composition of the Mantle
229
Composition of the Core
231
Fluxes in the Solid Earth
Cycling between Crust and Mantle
xi
233
233
xii
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Heat Exchange between Mantle and Core
Fluids and the Irreversible Formation of Continental Crust
Melting in the Mantle
238
Types of Melting Behavior
238
Differentiation in Melt-Crystal Systems
Fractional Crystallization
240
243
243
Chemical Variation Diagrams
246
Liquid Immiscibility
247
The Behavior of Trace Elements
Trace Element Fractionation during Melting and Crystallization
Compatible and Incompatible Elements
Volatile Elements
248
248
251
255
Crust and Mantle Fluid Compositions
255
Mantle and Crust Reservoirs for Fluids
258
Cycling of Fluids between Crust and Mantle
Summary
Suggested Readings
Problems
259
259
260
262
USING STABLE ISOTOPES
Overview
Historical Perspective
What Makes Stable Isotopes Useful?
Mass Fractionation and Bond Strength
Geologic Interpretations Based on Isotopic Fractionation
Thermometry
263
263
263
264
266
266
266
Isotopic Evolution of the Oceans
270
Fractionation in the Hydrologic Cycle
271
Fractionation in Geothermal and Hydrothermal Systems
275
Fractionation in Sedimentary Basins
278
Fractionation among Biogenic Compounds
278
Isotopic Fractionation around Marine Oil and Gas Seeps
Summary
Suggested Readings
Problems
FOURTEEN
236
237
Thermodynamic Effects of Melting
Causes of Melting
THIRTEEN
236
279
281
282
284
USING RADIOACTIVE ISOTOPES
Overview
Principles of Radioactivity
286
286
286
Nuclide Stability
286
Decay Mechanisms
287
Rate of Radioactive Decay
288
Decay Series and Secular Equilibrium
Geochronology
290
290
Potassium-Argon System
291
Rubidium-Strontium System
292
Contents
Samarium-Neodymium System
294
Extinct Radionuclides
297
Fission Tracks
298
299
Neutron Activation Analysis
299
40Argon-39Argon
300
Geochronology
Cosmic-Ray Exposure
Radionuclides as Tracers of Geochemical Processes
301
302
Heterogeneity of the Earth’s Mantle
302
Magmatic Assimilation
304
Subduction of Sediments
306
Isotopic Composition of the Oceans
307
Degassing of the Earth’s Interior to Form the Atmosphere
Summary
Suggested Readings
Problems
FIFTEEN
294
Uranium-Thorium-Lead System
Geochemical Applications of Induced Radioactivity
308
310
310
312
STRETCHING OUR HORIZONS:
COSMOCHEMISTRY
Overview
Why Study Cosmochemistry?
Origin and Abundance of the Elements
Nucleosynthesis in Stars
Cosmic Abundance Patterns
Chondrites as Sources of Cosmochemical Data
Cosmochemical Behavior of Elements
Controls on Cosmochemical Behavior
Chemical Fractionations Observed in Chondrites
Condensation of the Elements
313
313
313
314
314
316
318
320
320
321
323
How Equilibrium Condensation Works
323
The Condensation Sequence
326
Evidence for Condensation in Chondrites
Infusion of Matter from Outside the Solar System
327
327
Isotopic Diversity in Meteorites
327
A Supernova Trigger?
329
The Discovery of Stardust in Chondrites
The Most Volatile Materials: Organic Compounds and Ices
Extraterrestrial Organic Compounds
Ices—The Only Thing Left
xiii
330
330
330
331
A Time Scale for Creation
Estimating the Bulk Compositions of Planets
332
333
Some Constraints on Cosmochemical Models
333
The Equilibrium Condensation Model
335
The Heterogeneous Accretion Model
336
The Chondrite Mixing Model
336
Planetary Models: Cores and Mantles
339
xiv
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Summary
Suggested Readings
Problems
340
341
342
Appendix A: Mathematical Methods
Appendix B: Finding and Evaluating Geochemical Data
Appendix C: Numerical Values of Geochemical Interest
Glossary
Index
343
348
351
353
359
PREFACE TO THE SECOND EDITION
The various operations of Nature, and the changes which take place in the
several substances around us, are so much better understood by an attention to the laws of chemistry, that in every walk of life the chemist has a
manifest advantage.
Samuel Parkes (1816), The Chemical Catechism
T
he modern geologist with no knowledge of geochemistry will be severely limited. Indeed, geochemistry now pervades the discipline—providing the
basis for measurement of geologic time, allowing critical
insights into the Earth’s inaccessible interior, aiding in
the exploration for economic resources, understanding
how we are altering our environment, and unraveling
the complex workings of geochemical systems on the
Earth and its neighboring planets. Geochemical reasoning reveals both processes and pathways, as the book’s
title implies.
Our book is about chemistry, written expressly for
geologists. To be more specific, it is a text we hope will
be used to introduce advanced undergraduate and graduate students to the basic principles of geochemistry. It
may even inspire some to explore further and to become
geochemists themselves. Whether that happens or not,
our major objective is to help geology students gain that
manifest advantage noted in the quote above by showing
that concepts from chemistry have a place in all corners
of geology. The ideas in this book should be useful to
all practicing geologists, from petrologists to paleontologists, from geophysicists to astrogeologists.
Students exposed to a new discipline should expect
to spend some time assimilating its vocabulary and theoretical underpinnings. It isn’t long, though, before most
students begin to ask, “How can I use it?” Geologists,
commonly being rather practical people, reach this point
perhaps earlier than most scientists. In this book, therefore, we have tried not to just talk about geochemical
theory, but to show how its principles can be used to
solve problems. We have integrated into each chapter a
number of worked problems, solved step by step to
demonstrate the way that ideas can be put to practical
use. In most cases we have devised problems that are
based on published research, so that the student can relate abstract principles to the real concerns of geologists.
Working the problems at the end of each chapter will
further reinforce what has been covered in the text.
This attention to problem solving reflects our belief
that geochemistry, beyond the most rudimentary level, is
quantitative. To answer real questions about the chemical behavior of the Earth, a geologist must be prepared
to manipulate quantitative data and perform calculations.
We address this philosophy directly at the end of chapter 1. It should be understood at the outset, however,
that most questions at the level of this book can be answered by anyone with a standard undergraduate background in the sciences. We assume only that the student
has had a year of college-level chemistry and a year of
calculus. In this second edition, we have added a chapter (chapter 2) to help students put a geological frame
around fundamental concepts in chemistry. As in the earlier edition, we have outlined more advanced methods for
handling a few specialized math concepts, such as partial derivatives, in an appendix. With these aids, most routine approaches to solving geochemical problems should
already be within the student’s reach.
As the book’s title suggests, we have tried from the
beginning to balance the traditional equilibrium perspective with observations from a kinetic viewpoint,
emphasizing both pathways and processes in geochemistry. We have focused initially on processes in which
temperature and pressure are nearly constant. After an
abbreviated introduction to the laws of thermodynamics
and to fundamental equations for flow and diffusive
xv
xvi
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
transport, we spend the first half of the book investigating diagenesis, weathering, and solution chemistry in
natural waters. We have made significant changes in this
second edition of the book by including new chapters on
mineral structure and bonding and on organic geochemistry. These concepts are also integrated into other chapters. In the second half of the book, we return for a closer
look at thermodynamics and kinetics as they apply to
systems undergoing changes in temperature and pressure
during magmatism and metamorphism. Stable and radiogenic isotopes are treated near the end of the book,
where their applications to many of the previous topics
are emphasized. The concept of geochemical systems, introduced in chapter 1, is illustrated in chapters 8, 12, and
15, as we emphasize, respectively, the ocean-atmosphere,
the core-mantle-crust, and the planets as systems. In this
second edition, we have made special efforts to update
examples and worked problems to reflect new discoveries
in the field.
The division of geochemical topics outlined above has
served us well in the classroom. Among other advantages, it spreads out the burden of learning basic theory
and allows the instructor to address practical applications
of geochemistry before students’ eyes begin to glaze over.
For those who might want to leap directly to applicable
equations rather than follow our derivations, we have
placed key equations in boxes. At the other extreme,
we hope that instructors and curious students are intrigued by the occasional advanced or offbeat ideas that
we have included as highlighted boxes in each chapter.
They are intended to provoke discussion and further
investigation.
This book is a distillation of what we ourselves have
learned from our own teachers and from our many colleagues and friends in geochemistry; the book itself is
(eloquent, we hope) testimony to their influences. The
appearance of the second edition results largely from the
unflagging encouragement and contagious enthusiasm
of Holly Hodder, Prentice-Hall editor of the first edition.
We have also greatly benefited from the counsel of our
new Columbia University Press editor, Robin Smith. We
hope that geology students will find this revised edition
stimulating, accessible, and useful.
CHAPTER ONE
INTRODUCING CONCEPTS IN
GEOCHEMICAL SYSTEMS
OVERVIEW
Historical Overview
In this introductory chapter, our goal is to examine the
range of problems that interest geochemists and to
compare two fundamental approaches to geochemistry.
Thermodynamics and kinetics are complementary ways
of viewing chemical changes that may take place in
nature. By learning to use both approaches, you will
come to appreciate the similarities among geochemical
processes and be able to follow the many pathways of
change. To develop some skills in using the tools of geochemistry, we also examine the limitations of thermodynamics and kinetics, and discuss practical considerations
in problem solving.
The roots of geochemistry belong to both geology
and chemistry. Many of the practical observations of
Agricola (Georg Bauer), Nicolas Steno, and other Renaissance geologists helped to expand knowledge of the
behavior of the elements and their occurrence in nature.
As we see in more detail in chapter 2, these observations
formed the basis from which both modern chemistry
and geology grew in the late eighteenth century. The
writings of Antoine Lavoisier and his contemporaries,
during the age when modern chemistry began to take
shape, were filled with conjectures about the oceans and
atmosphere, soils, rocks, and the processes that modify
them. In the course of separating and characterizing the
properties of the elements, Lavoisier, Humphry Davy,
John Dalton, and others also contributed to the growing
debate among geologists about the composition of the
Earth. Similarly, Romé de L’Isle and Réné-Just Haüy
made the first modern observations of crystals, leading
very quickly to advances in mineralogy and structural
chemistry.
The term geochemistry was apparently used for the
first time by the Swiss chemist C. F. Schönbein in 1838.
The emergence of geochemistry as a separate discipline, however, came with the establishment of major
WHAT IS GEOCHEMISTRY?
In studying the Earth, geologists use a number of borrowed tools. Some of these come from physicists and
mathematicians, others were developed by biologists,
and still others by chemists. When we use the tools of
chemistry, we are looking at the world as geochemists.
From this broad perspective, the distinction between
geochemistry and some other disciplines in geology, such
as metamorphic petrology or crystallography, is a fuzzy
and artificial one.
1
2
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
laboratories at the U.S. Geological Survey in 1884; the
Carnegie Institution of Washington, D.C., in 1904; and
in several European countries, notably Norway and the
Soviet Union, between roughly 1910 and 1925. From
these laboratories came some of the first systematic surveys of rock and mineral compositions and some of the
first experimental studies in which the thermodynamic
conditions of mineral stability were investigated. F. W.
Clarke’s massive work, The Data of Geochemistry,
published in 1909 and revised several times over the next
16 years, summarized analytical work performed at the
U.S. Geological Survey and elsewhere, allowing geologists to estimate the average composition of the crust.
The phase rule, which was first suggested through theoretical work by J. Willard Gibbs of Yale University in the
1880s, was applied to field studies of metamorphic rocks
by European petrologists B. Roozeboom and P. Eskola,
who thus established the chemical basis for the metamorphic facies concept. Simultaneously, Arthur L. Day
and others at the Carnegie Institution of Washington
Geophysical Laboratory began a program of experimentation focused on the processes that generate igneous
rocks.
During the twentieth century, the course of geochemistry has been guided by several technological advances.
The first of these was the discovery by Max von Laue
(1912) that the internal arrangement of atoms in a crystalline substance can serve as a diffraction grating to
scatter a beam of X-rays. Shortly after this discovery,
William L. Bragg used this technique to determine the
structure of halite. In the 1920s, Victor M. Goldschmidt
and his associates at the University of Oslo determined
the structures of a large number of common minerals,
and from these structures, formulated principles regarding the distribution of elements in natural compounds.
Their papers, published in the series Geochemische Verteilungsgesetze der Elemente, were perhaps the greatest
contributions to geochemistry in their time.
During the 1930s and into the years encompassing
World War II, new steel alloys were developed that permitted experimental petrologists to investigate processes
at very high pressures for the first time. Percy Bridgeman
of Harvard, Norman L. Bowen of the Carnegie Geophysical Laboratory, and a growing army of colleagues
built apparatus with which they could synthesize mineral assemblages at temperatures and pressures similar to
those found in the lower crust and upper mantle. With
these technological advances, geochemists were finally
able to investigate the chemistry of inaccessible portions
of the Earth.
The use of radioactive isotopes in geochronology
began at the end of the nineteenth century, following the
dramatic discoveries of Ernest Rutherford, Henri Becquerel, and Marie and Pierre Curie. The greatest influence
of nuclear chemistry on geology, however, came with the
development of modern, high-precision mass spectrometers in the late 1930s. Isotope chemistry rapidly gained
visibility through the careful mass spectroscopic measurements of Alfred O. C. Nier of the University of Minnesota; between 1936 and 1939, he determined the isotopic
compositions of 25 elements. Nier’s creation in 1947 of
a simple but highly precise mass spectrometer brought
the technology within the budgetary reach of many geochemistry labs and opened a field that, even today, is still
a rapidly growing area of geochemistry. Beginning with
the work of Harold Urey of the University of Chicago and
his students in the late 1940s, stable isotopes have become
standard tools of economic geology and environmental
geochemistry. In recent years, precise measurements of
radiogenic isotopes have also proved increasingly valuable in tracing chemical pathways in the Earth.
Advances in instrumentation brought explosive
growth to other analytical corners of geochemistry during the last half of the twentieth century. Geochemists
made use of increasingly sophisticated detectors and
electronic signal processing systems to perform chemical
analyses at lower and lower detection levels, opening
the new subfield of trace element studies. As we show in
chapter 12, this has made it possible to refine our understanding of the evolution of the Earth’s mantle and crust
and of many fundamental petrologic processes. The electron microprobe, developed independently by R. Castaing and A. Guinier in France and I. B. Borovsky in
Russia in the late 1940s, became commercially available
to geochemists in the 1960s. Because the new instrument
made it possible to perform analyses on the scale of microns, geochemists were able to study not only microscopic samples but also subtle chemical gradients within
larger samples. The 1970s and 1980s, as a result, saw
increasing interest in geochemical kinetics.
Finally, as exploration for and exploitation of petroleum became a major focus for geology in the early
decades of the twentieth century, the field of organic
geochemistry slowly gained importance. In the 1930s,
Russia’s “father of geochemistry,” V. I. Vernadsky, was
among the first to speculate on the genetic relationships
Introducing Concepts in Geochemical Systems
among sedimentary organic matter, petroleum, and hydrocarbon gases. Alfred Treibs confirmed the biological
origin of oil in the 1930s. Others, including Wallace
Pratt of Humble Oil Company, correctly deduced that
oil is formed by natural cracking of large organic molecules and noted that local differences in source material
and diagenetic conditions create a distinctive chemical
“fingerprint” for each petroleum reservoir. In the latter
half of the century, organic geochemists turned their attention to environmental contaminants, redirecting the
petroleum geochemists’ methods to trace and control
industrial organic matter in soils and groundwater and
reconstruct ancient environments.
The growth of geochemistry has been paralleled by
the development of cosmochemistry, and many scientists
have spent time in both disciplines. From the uniformitarian point of view, the study of terrestrial pathways
and processes in geochemistry is merely an introduction
to the broader search for principles that govern the
general behavior of planetary materials. In the 1920s
Victor M. Goldschmidt, for example, refined many of
his notions about the affinity of elements for metallic,
sulfidic, or siliceous substances by studying meteorites.
(You will read more about this in chapter 2.) In more
recent years, our understanding of crust-mantle differentiation in the early Earth has been enhanced by studies
of lunar samples.
This historical sketch has unavoidably bypassed many
significant events and people in the growth of geochemistry. It should be apparent from even this short treatment, however, that the boundary between geology and
chemistry has been porous. For the generations of investigators who have worked at the boundary during the
past two centuries, it has been an exciting and challenging time. This period has also been marked by recurring
cycles of interest in data gathering and the growth of
basic theory, to which both chemists and geologists have
contributed. Each advance in technology has invited geochemists to scrutinize the Earth more closely, expanded
the range of information available to us, and led to a new
interpretation of fundamental processes.
Beginning Your Study of Geochemistry
In this book, we explore many ideas that geology
shares with chemistry. In doing so, we offer a perspective
on geology that may be new to you. You don’t need to
be a chemist to use the techniques of geochemistry, just
3
as you don’t need to be a physicist to understand the
constraints of seismic data or a biologist to benefit from
the wealth of evolutionary insights from paleontology.
However, the language and methods of geochemistry are
different from much that is familiar in standard geology.
You will have to pick up a new vocabulary and become
familiar with some fundamental concepts of chemistry
to appreciate the role geochemistry plays in interpreting
geological events and environments. We start to introduce some of these concepts in this chapter and the next,
and introduce others as we go along. As we explore environments from magma chambers to pelagic sediment
columns, we hope that your understanding of other disciplines in geology will also be enriched.
As a student of geology, you have already been introduced to a number of unifying concepts. Plate tectonics,
for example, provides a physical framework in which
seemingly independent events such as arc volcanism,
earthquakes, and high-pressure metamorphism can be
interpreted as the products of large-scale movements
in the lithosphere. As geologists fine-tune the theory of
plate tectonics, its potential for letting us in on broader
secrets of how the Earth works is even more tantalizing
than its promise of showing us how the individual parts
behave.
In the same way, as we try to answer the question
“What is geochemistry?” in this book, we look for unifying concepts by discovering how the tools of geochemistry can help us interpret a wide variety of geologic
environments. In chapter 3, for example, we introduce
the concept of chemical potential and demonstrate its
role in characterizing directions of change in chemical
systems. We then encounter chemical potential in various guises in later chapters as we examine topics in
chemical weathering, diagenesis, stable isotope fractionation, magma crystallization, and condensation of gases
in the early solar system, to mention a few. In this way,
you will come to recognize that chemical potential is a
common strand in the fabric of geochemistry. We offer
you the continuing challenge of recognizing other unifying concepts as we proceed through this book.
A glance at the Table of Contents shows you that
geochemistry virtually reaches into all areas of geology.
Because this book is meant to explore broadly applicable
ideas in geochemistry, we do not concentrate on one corner of geology exclusively. Instead, we choose problems
that are of interest to many kinds of geologists and help
you find ways to approach and (we hope) solve them.
4
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
GEOCHEMICAL VARIABLES
Most geochemical processes can be described in rather
broad, qualitative terms if we are presenting them for a
lay audience or we are trying to establish a “big picture”
of how things work. As a rule, however, qualitative arguments in geochemistry are believable only if they are
backed up by more rigorous, quantitative analyses. At a
minimum, it is important to begin any investigation of
a geochemical system by identifying a set of properties
(variables) that may account for potentially significant
changes in the system.
Let’s begin by defining two types of variables, which
we will refer to as extensive and intensive. The first of
these is a measure of the size or extent of the system in
which we are interested. For example, mass is an extensive variable because (all other properties being held
constant) the mass of a system is a function of its size.
If we choose to examine a block of galena (PbS) that is
20 cm on a side, we expect to find that it has a different
mass than a block that measures only 5 cm on a side.
Volume is another familiar extensive property, clearly a
measure of the size of a system. It should be obvious
from these examples that extensive quantities are additive. That is, the mass of galena in the two blocks
described above is equal to the sum of their individual
masses. Volumes are also additive.
Intensive variables, in contrast, have values that are
independent of the size of the system under consideration. Two of the most commonly described intensive
properties are temperature and pressure. To verify that
these are independent of the size of a system, imagine a
block of galena that is uniformly at room temperature
(∼25°C). Experience tells us that if we do nothing to
the block except cut it in half, we will then have two
blocks of galena, but each will still be at 25°C. A similar
thought experiment will confirm that pressure is also
independent of the dimensions of a system, provided
that it too is uniform throughout the system.
It is useful to conceive many geochemical problems in
terms of intensive rather than extensive properties. By
doing so, we free ourselves from considering the size of
a system and think simply about its behavior. For example, if we want to determine whether a mineral assemblage consisting of quartz and olivine is more or less
stable than an assemblage of pyroxene, the answer will
be most useful if it is cast in terms of energy per unit of
mass or volume.
The conversion from extensive to intensive properties
involves normalizing to some convenient measure of the
size of the system. For example, we may choose to describe a block of galena by its mass (an extensive property)
normalized by its volume (another extensive property).
The resulting quantity, density, is the same regardless of
which block we choose to consider, and is therefore an
intensive property of the system. We might choose to
normalize instead by multiplying the mass of the block
by 6.02 × 1023 (Avogadro’s number: the number of formula units per mole of any substance, and in this case an
appropriate scale factor) and then dividing by the number of PbS formula units in it. The result would be the
molecular weight of galena, another intensive property.
In general, then, the ratio of two extensive variables is
an intensive variable. When we start to consider energy
functions, in chapter 3, we will begin using an overscore
on symbols (as in ∆Ḡr) to indicate that they are molar
quantities, and therefore intensive. In that chapter and
beyond, it will be useful to recognize that the derivative
of an extensive variable with respect to either an extensive or an intensive variable is also intensive.
GEOCHEMICAL SYSTEMS
Several times we have used the word system in this
chapter. By this, we mean the portion of the universe
that is of interest for a particular problem, whether it is
as small as an individual clay particle or as large as the
Solar System.
Because the results of geochemical studies will usually
depend sensitively on how a system is defined, it is always
important to begin the study by defining the boundaries
of the system. There are three generic types of systems,
each defined by the condition of its boundaries. Systems
are isolated if they cannot exchange either matter or energy with the universe beyond their limits. A perfectly
insulated thermos bottle is an example of such a system.
Geochemists often make the simplifying assumption
that systems are isolated because it is easier to keep track
of the state of a system both physically and mathematically if they can imagine a well-defined wall around it.
In chapter 3, we also show that the theoretical existence
of isolated systems enables us to put limits on the behavior of fundamental thermodynamic functions. In the
real universe, however, truly isolated systems cannot exist
because there are no perfect insulators to prevent the
exchange of energy across system boundaries.
Introducing Concepts in Geochemical Systems
In nature, systems can either be closed or open. A
closed system is one that can exchange energy, but not
matter, across its boundaries. An open system can exchange both energy and matter. This distinction is actually somewhat artificial, although it is practical in most
cases. For example, provided that the beds confining an
aquifer have an acceptably low permeability, we may consider the aquifer to be a closed system if we are studying
only the relatively rapid chemical reactions within it. If,
however, we are considering a long-term toxic waste
disposal problem, our definition of “acceptably low” permeability would probably be different. The same aquifer
might more realistically be considered an open system.
Often, then, the definition of a system will have to include an assessment of the rates of transfer across its
boundaries. So long as the system is contained within
limits of acceptably low transfer rates, we can usually
justify treating it as either a closed or an isolated system.
Thus, time is an important consideration when we are
defining systems.
5
along which the system may evolve between states of
thermodynamic equilibrium and determine the rates of
change of system properties along those pathways. Most
geochemical systems can change between equilibrium
states along a variety of pathways, some of which are
more efficient than others. In a kinetic study, the task is
often to determine which of the competing pathways
is dominant.
As geochemists, we have an interest in determining
not only what should happen in real geologic systems
(the thermodynamic answer) but also how it is most
likely to happen (the kinetic answer). The two are intimately linked. While the focus of thermodynamics is
on the end states of a system—before and after a change
has taken place—a kinetic treatment of geochemical systems sheds light on what happens between the end states.
Where possible, it is smart to keep both approaches in
mind.
THERMODYNAMICS AND KINETICS
AN EXAMPLE: COMPARING
THERMODYNAMIC AND
KINETIC APPROACHES
There are two different sets of conditions under which
we might study a geochemical system. First, we might
consider a system in a state of equilibrium, a consequence of which is that observable properties of the
system undergo no change with time. (We offer a more
exact definition in terms of thermodynamic conditions
in chapter 3.) If a system is at equilibrium, we are generally not concerned with (and can’t readily determine) how
it reached equilibrium. The tools of thermodynamics
are not useful for asking questions about a system’s
evolutionary history. However, with the appropriate
equations, we are able to estimate what a system at
equilibrium would look like under any environmental
conditions we choose. The methods of thermodynamics
allow us to use environmental properties such as temperature and pressure, plus the system’s bulk composition, to predict which substances will be stable and in
what relative amounts they will be present. In this way,
the thermodynamic approach to a geochemical system
can help us measure its stability and predict the direction
in which it will change if its environmental parameters
change.
The alternative to studying a system in terms of
thermodynamic conditions is to examine the kinetics of
the system. That is, we can study each of the pathways
Figure 1.1a is a schematic drawing of a hypothetical system that we can use to illustrate these two approaches.
It is a closed system consisting of five subsystems or reservoirs (indicated by boxes) that are connected by pathways
(labeled by arrows). To make it less abstract, we have
redrawn this schematic in figure 1.1b so that you can
see the real system it is meant to represent: the floor plan
of a small house without exterior doors or windows (unrealistic, perhaps, but the absence of exterior outlets is
what makes this a closed system). We can use this familiar setting to demonstrate many of the principles that
apply to dynamic geochemical systems.
Assume that you, the owner of the house, have invited several of your friends to a party, and that they are
all hungry. As we take our first look at the party, people
are distributed unevenly in the living room, the dining
room, the kitchen, the bedroom, and the bathroom. Your
guests are free to move from one room to another, provided that they use doorways and do not break holes
through the walls, but there are some rules of population
dynamics (which we will describe shortly) that govern
the rates at which they move.
From a thermodynamic point of view, the system we
have described is initially in a state (i.e., hungry) that
we would have a hard time describing mathematically,
6
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 1.1. (a) Schematic diagram for a house with five rooms.
The number of people in room i is given by A i , and their rates
of movement between rooms by fij , where i is the room being
vacated and j is a destination. The units of fij are people hour −1.
(b) Floor plan for the house described schematically in (a).
but which is clearly different from the final state that you,
as host, would like to approach (i.e., well fed). It is also
clear that if you were to begin serving food to your
friends, they would gradually move from the hungry
state to the well-fed state. If people were as well behaved
as chemical systems, we ought to be able to use an approach analogous to thermodynamics to calculate the
“energy” difference between the two states—just how
hungry are your guests?—and therefore predict the
amount of food you will have to provide to let everyone
go home feeling satisfied. The mathematics necessary
to do this calculation might involve writing a differential
equation for each room, expressing variations in some
function W (a “fullness” function) in terms of variations
in amounts of ham, cole slaw, and beer:
dW = (∂W/∂Ham)dHam + (∂W/∂Slaw)dSlaw
+ (∂W/∂Beer)dBeer.
(1.1)
This equation is in the form of a total differential, described in appendix A. It will become familiar in later
chapters.
Thermodynamics, then, allows us to recognize that
the initial hungry state of your guests is less stable than
the final well-fed state, and provides a measure of how
much change to expect during the party. Even with an
adequate description of a system, however, thermodynamics can give no guarantee that the real world will
reach the equilibrium configuration that we calculate on
paper. Despite our best thermodynamic calculations, for
example, the people at our party may remain indefinitely
in a hungry state. Maybe they don’t like your cooking,
or the large dog in the corner of your kitchen discourages them.
In the geologic realm, minerals or mineral assemblages
may appear to be stable simply because we observe them
over times that are short compared with the rates of
reactions that alter them. Such materials are said to be
metastable. Granites and limestones are metastable at
the Earth’s surface, even though we know them to be less
stable than clays and salt-bearing groundwaters. The
aragonitic tests that are typically excreted by modern
shelled organisms are less stable than calcite, yet aragonite is only slowly converted to the calcite that thermodynamics tells us should be present. Diamonds, stable
only under the high-pressure conditions of the mantle,
do not spontaneously restructure themselves to form
graphite, even over very large spans of time. In each of
these cases, the reaction paths between the metastable
and stable states involve intermediate steps that are difficult to perform and therefore take place very slowly or
not at all. We refer to these intermediate steps as kinetic
barriers to reaction. Given an adequate description of the
compositional and environmental constraints on a system, thermodynamics can tell us what that system should
look like if it is given infinite time in which to overcome
kinetic barriers.
We encounter many geologically important systems
that are in transition between equilibrium states, yet restrained by kinetic barriers. Is it possible to apply thermodynamic reasoning to these systems? It is common to find
that very large systems, such as the ocean, are in a state
of dynamic balance or steady state, in which environmental conditions differ from one part of the system to
another, while the overall system may persist indefinitely
with no net change in its state. Under these conditions,
thermodynamics remains a useful approach, although it
Introducing Concepts in Geochemical Systems
should be clear that no system-wide state of thermodynamic equilibrium can be described.
To return to the example in figure 1.1, let us assume
that randomly moving guests will nibble food whenever
they can find it. Further, suppose that people in any particular room at the party always move around more
rapidly within the room than they do from one room to
another. If that is so, then the proportion of hungry to
moderately well-fed people at different places within that
room will be fairly uniform any time we peek into it. Different rooms, however, may have different proportions
of hungry and well-fed people, depending on the size of
the room and whether it has food in it. Each of the
rooms, in other words, is always in a local equilibrium
state that differs from the states of the other rooms, and
the entire house is a mosaic of equilibrium states. If we
knew how the “thermodynamics” of people works, then
we could apply “fullness” functions such as equation 1.1
within rooms to help describe how hungry the guests in
them were at any time. Furthermore, despite the fact that
the rooms are not in equilibrium with each other, consideration of the mosaic provides valuable information
about the direction of overall change in the house. (Note:
Every example has its limitations. In this case, the statistics of small numbers make it difficult to take this
observation literally. The more people we invite to the
party, however, the more uniform the population is in
any room.)
This is the basis of the concept of local equilibrium
that is often used in metamorphic petrology to discuss
mineral assemblages that are well defined on the scale
of a thin section, but which may differ from one hand
sample to another across an outcrop. Within small parts
of the system, rates of reaction may often be rapid enough
that equilibrium is maintained and thermodynamics can
give useful information about the existence and relative
abundance of minerals in which we are interested. Provided that we can agree on the definition of the mineral
assemblages and therefore the sizes of the regions of
mosaic equilibrium, we can use thermodynamics to analyze the regions individually and to predict directions
of change in the system as a whole.
The point is that thermodynamics remains a useful
part of our tool kit even if we examine a system that is far
from equilibrium. We can’t use thermodynamics, however, to answer questions such as how rapidly the party
will approach its well-fed state, or how people may be
distributed throughout the house at any instant in time.
7
In the world of geochemistry, we may be able to describe
deviations from equilibrium by applying thermodynamics, but we find ourselves incapable of predicting
which of many pathways a system may follow toward
equilibrium or how rapidly it gets there. Those are kinetic problems.
Suppose we address one of the kinetic problems for
the party system: how are people distributed among the
rooms as a function of time? To answer the question,
let’s choose one room at random and look at the operations that will cause its population to change with time.
We do this by taking an inventory of the rates of input
to and output from the room, and comparing them. We
can write this as a differential equation:
dAi /dt = total input − total output
=
Σf − Σf .
ji
(1.2)
ij
Here we are using the symbol Ai to indicate the number of people in the room we have chosen (room i, where
i is a number from one to five). The quantities labeled f
are fluxes; that is, numbers of people moving from one
room to another per unit of time. The order of subscripts
on f is such that the first one identifies where people are
moving from, and the second tells us where they are moving to. This notation is used to label the pathways in figure 1.1a. Room j in each case is some room other than i.
The summations in equation 1.2 are over all values of i.
We have to write an expression of this type for each
room in the example (for dA1/dt, dA2 /dt, and so forth).
If we define the initial values for each Ai and all fij , then
we can solve these flux equations simultaneously for any
time t. The result will be a room-by-room population census as the party proceeds. This step is often intimidating,
but only occasionally impossible, especially with appropriate software for numerically solving sets of differential equations.
Armed with new analytical insight, let’s reexamine
the concept of a steady state for our hypothetical system.
It can now be seen to be a condition in which each of
the flux equations 1.2 is equal to zero. If there is such a
state, the population in each room will not change with
time. Notice that this does not imply that the fluxes are
zero, but that the inputs and outputs from each room
are balanced. That is, steady state may also be defined
as a condition in which the time derivatives dfij /dt are
equal to zero. Is this confusing? Consider worked problem 1.1.
8
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Worked Problem 1.1
In figure 1.1a, we have shown numerical values for each of the
nonzero fluxes in the party example. How can we determine,
using these constant values for fij , whether the five flux equations for this system describe a system in steady state?
The flux equations are
dA1 /dt = f21 + f31 + f41 − f12 − f13 − f14 = 45 + 60 + 100
− 30 − 75 − 100 = 0
dA2 /dt = f12 + f32 − f21 − f23 = 30 + 40 − 45 − 25 = 0
dA3 /dt = f13 + f23 − f31 − f32 = 75 + 25 − 60 − 40 = 0
dA4 /dt = f14 + f54 − f41 − f45 = 100 + 24 − 100 − 24 = 0
dA5 /dt = f45 − f54 = 24 − 24 = 0
Because each of the time derivatives is equal to zero, the
reservoir contents, Ai, must be fixed at the values shown in figure 1.1a. Therefore, although people are free to move between
rooms, the distribution of party guests is in a steady state. If we
solved the five differential equations simultaneously, we would
see that each population census in time looks like the previous
one: no change.
There is no guarantee that all systems have a steady
state, particularly if they are open. A system may continually accumulate or lose mass through time (a dissolving
crystal is one such system). At one time, it was believed
that the salinity of the oceans was a unidirectional function of time, and that the progressive accumulation of salt
in seawater could be used to infer the age of the oceans.
In that view, the ocean is not at or approaching a steady
state. It is also possible that a system may have more than
one steady state. Systems that involve interacting fluxes
for large numbers of species, or that have an oscillatory
behavior (due, perhaps, to the passage of seasons), often
have multiple steady states.
What might cause a system to move away from a
steady state? In our party example, we can define some
laws of population dynamics to describe the rates at
which people move from room to room. These “laws”
are a set of empirical statements that tell us how the
fluxes fij in the set of equations we are developing change
with time. It is important to do this because fluxes are
rarely constant in natural systems. In this problem, for
example, it is fair to assume that the rate at which people
leave any given room to move to another is directly proportional to the number of people in the room. The more
crowded the room becomes, the more likely people are
to leave it. That is,
dAout
i /dt =
Σf = Σk A = A Σk ,
ij
ij
i
i
ij
(1.3)
in which the quantities kij are simple rate constants expressing migrations from room i to room j and the summations are over all values of j. Thus, the rate at which
people leave the living room (dA1out/dt) is equal to the
number of people in the room (A1) times the sum of k12,
k13, k14, and k15, the rate constants that govern movement
out of the living room. The kij have units of inverse time.
This “law” assumes first-order kinetics, which is justifiable in many geochemical problems as well. The rate of
removal of magnesium from seawater at midocean ridges
appears to be directly proportional to the Mg2+ concentration of seawater, for example. Similarly, the rate of formation of halite evaporates is probably also a first-order
function of the concentration of NaCl in the oceans.
By introducing rate laws such as the first-order expression applied here, we recognize that fluxes, in general, are not constant in systems away from their steady
state. If we assume a first-order law, fij = kij Ai in the party
system, this means that we could change the fluxes of
people from one room to another either by changing the
population in one or more of the rooms or by varying
the rate constants kij . For example, without abandoning
the simple first-order law or changing the rate constants,
we could instantly move several people to the dining
room as a way of modeling the effect of a one-time impulse, such as bringing out more food. Equation 1.3 predicts that overcrowding would make the fluxes out of the
dining room higher for a while, and thus that the party
would begin relaxing again to its steady state. If we chose
instead to model the gradual effects as people get full
and mellow, however, we would need to replace the rate
constants with time-variant functions somehow linked
to the “thermodynamic” fullness function we described
earlier. This would make the set of kinetic equations 1.2
much trickier to solve, although probably more realistic.
It is not appropriate to assume that all geologic processes follow simple rate laws. Many do not. Complicated
rate laws may be appropriate for systems that exhibit
seasonal variability or in which the rate of removal of
material from reservoir i is a thermodynamic function
of Ai. Complex processes by which ions detach from
the surfaces of crystals, for example, may cause rates of
solubility to be highly nonlinear functions of undersatu-
Introducing Concepts in Geochemical Systems
ration. In most cases, determination of appropriate rate
laws and constants is a difficult business at best. For this
reason, you will find that many potentially interesting
kinetic problems have been discussed in schematic terms
by geochemists, but solved only in a qualitative or greatly
simplified form.
NOTES ON PROBLEM SOLVING
In this chapter, we have outlined the range of problems
that interest geochemists and introduced in broad terms
the thermodynamic and kinetic approaches to them. In
the remaining chapters, we look in detail at selected
geochemical topics and develop specific techniques for
analyzing them. As you progress, reflect on this introduction. In it, we have tried to emphasize basic principles
that you should always bear in mind as you study a geochemical question.
Remember that predictions made from thermodynamic or kinetic calculations cannot be any better
than our description of the system allows them to be.
In the frivolous example we have been using, we cannot
expect thermodynamics to comment on events outside
of the house, or on the state of hunger being felt by
guests lurking in undescribed closets. We also cannot
comment on the relative efficiency of kinetic pathways
we have failed to include in the model, such as extra
doors between rooms. As we have just shown, this leads
us to make strategic compromises, simplifying models to
make them solvable while keeping them as true to nature
as we can.
This limitation may seem obvious, but it is surprisingly easy to overlook. If we are examining a system consisting of metamorphosed shales, for example, and forget
to include in our description any chemical reactions
that produce biotite, there is no way that we can expect
thermodynamics to predict its existence. This omission
has a serious effect. For a mineral such as biotite, which
may be a volumetrically significant phase in some metamorphosed shales, this error in description will render
meaningless any thermodynamic analysis of the system
as a whole. If we decide to omit a minor phase like apatite, however, the effect on thermodynamic calculations
should be minimal. Thermodynamics will not be able to
comment on its stability, but our overall description of the
system should not be seriously in error. Between these
two extremes, it may not be easy to decide how much
9
simplification is too much. Is it important to include ilmenite and pyrite? The answer may not be clear, but you
should always ask the question.
It is just as easy to forget that this limitation applies
to kinetic studies as well. Because geochemical reactions
commonly proceed along several parallel pathways, a
valid kinetic treatment depends critically on how well we
have identified and characterized those pathways. Most
systems of even moderate complexity, however, include
pathways that are easy to underestimate or to overlook
completely. In chapter 8, for example, we consider the
effect of hydrothermal circulation at midoceanic ridge
crests on the mass balance of calcium and magnesium
in the oceans. Until the mid-1970s, few data suggested
that this was a significant pathway in the geochemical
cycle for magnesium. Kinetic models, therefore, regularly
overestimated the importance of other pathways, such
as glauconite formation in marine sediments.
It is clear, then, that the credibility of any statements
we make regarding the appearance of phases, the stability
of a system, or rates of change depends in large part on
how carefully we have described the system before applying the tools of geochemistry. This may be the single
greatest source of difficulty you will experience in applying geochemical concepts to real-world problems.
Remember also that there are few concepts that are
the exclusive property of igneous or sedimentary petrologists, of oceanographers or soil chemists, of economic
geologists or planetary geologists. The breadth of geochemistry can be overwhelming. However, breadth can
also be advantageous when we try to solve practical problems. Familiarity with the data and methodology used in
one area of geochemistry will often help you see alternative solutions to a problem in another area. Many of
the fundamental quantities of thermodynamics and kinetics should gradually become familiar as they appear
in different contexts. Where possible, we try to point out
areas of obvious overlap. Do not hesitate to do as we
have done in this first chapter, though, and step completely out of the realm of geochemistry to play with an
idea. A simple and apparently unrelated thought problem, such as our party example, can sometimes suggest a
new way to approach a geochemical question.
Finally, we encourage you to develop the simplest
geochemical models that will adequately answer questions posed about the Earth. All too often, problems
appear impossible to solve because we have made them
10
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
unnecessarily complex and treat them as abstract mathematical exercises rather than as investigations of the
real Earth. In most cases, a careful selection of variables
and a set of physically justifiable simplifying assumptions can make a problem tractable without sacrificing
much accuracy. What you already know as a geologist
about the Earth’s properties and behavior should always
serve as a guide in making such choices.
Having said this, we also emphasize that mathematics is an essential tool for the quantitative work of geochemistry. Now is a good time to familiarize yourself
with the contents of the appendices at the end of this
book, before we begin our first excursion into thermodynamics. The first of these appendices is an abbreviated overview of mathematical concepts, emphasizing
methods that may not have been part of your first year
of college-level calculus. Among these concepts are
principles of partial differentiation and an introduction
to numerical methods for fitting functions to data and
finding roots. Refer to appendix A as often as necessary
to make the mathematical portions of this book more
palatable.
Some introductory texts include an appendix with
preferred data needed in problem solving. We have decided not to provide data tables, except when we need
an abbreviated table in a chapter. Instead, appendix B is
a reference list of standard sources for geochemical data
with which you should become familiar. Yes, this is less
convenient than a compact data table would have been.
We hope, though, that you will quickly become accustomed to finding and evaluating data for yourself, and
that our list will be more useful in the long run than a
table might have been.
As a slight concession in the opposite direction, we
have listed the values of selected fundamental constants
and conversion factors in appendix C. The literature
of geochemistry, particularly articles written prior to
the 1990s, uses a rich mixture of measurement units. We
have made no attempt to favor one system over another
in this book. Our practical opinion is that because data
are reported in several conventional forms, you should
be conversant with them all and should know how to
shift readily from one set of units to another.
suggested readings
These first references illustrate the breadth of geochemistry.
They also provide a wealth of data that may be most useful to
professionals.
Fairbridge, R. W., ed. 1972. The Encyclopedia of Geochemistry and Environmental Sciences. New York: Van Nostrand
Reinhold. (A frequently referenced source book for geochemical data.)
Goldschmidt, V. M. 1954. Geochemistry. Oxford: Clarendon
Press. (This book is now dated, but remains a classic.)
Mason, B., and C. Moore. 1982. Principles of Geochemistry.
New York: Wiley. (Probably the most readable survey of the
field of geochemistry now available.)
Wedepohl, K. H., ed. 1969–1978. Handbook of Geochemistry.
Berlin: Springer-Verlag. (A five-volume reference set with
ring-bound pages that are designed for convenient updating.
A definitive source book on the occurrence and natural
chemistry of the elements.)
PROBLEMS
(1.1)
You are already accustomed to describing geologic materials and environments in terms of measurable variables such as temperature or mass. Make a list of 10 such variables and indicate which ones
are intensive and which extensive.
(1.2)
Consider the mass balance and the pathways for change in a residual soil profile.
(a) Sketch a diagram similar to figure 1.1a, and indicate the way in which major species, such as Fe,
Ca, or organic matter, may be redistributed with time.
(b) Write a complete set of schematic flux equations to describe mass balance in your system, following the example of worked problem 1.1.
Introducing Concepts in Geochemical Systems
(1.3)
It has been said that diagenetic reactions generally occur in an open system, whereas metamorphic
reactions more commonly take place in a closed system. Explain what this geochemical difference
means and suggest a reason for it.
(1.4)
Consider the party problem again. What might you have to do to include each of the following extra
factors in the problem?
(a) You run out of food while guests are still hungry.
(b) Guests are free to enter and leave the party.
(c) The bathroom door is locked.
(d) Half of the guests try to avoid being in the same room as the other half, but everyone wants to
spend some time in the kitchen.
The following problems for math review test your skills with concepts we will be using from time
to time.
(1.5)
Write the derivative with respect to t for each of the following expressions:
(a) ln (t)3
(b) exp (2πa/αt) (π, a, and α are constants)
(c) sin (αt 2) (α is a constant)
(d) log10 (α/t) (α is a constant)
(e) t ln t + (1 − t) ln (1 − t)
(f) Σ exp (−Ei /Rt) (R and all Ei are constants)
αt
(g) ∫0 sin (t − z)dz (α and z are constants)
(1.6)
Write down the total differential, df, for each of the following expressions:
(a) f(x, y, z) = x2 + y2 + z2
(b) f(x, y, z) = sin (xyz)
(c) f(x, y, z) = x3y − xy2
(1.7)
Determine which of the following expressions are exact differentials:
(a) df = (y + z)dx + (z + y)dy + (x + y)dz
(b) df = 2xy dx + 2yz2 dy − (1 − x2 − 2y2z)dz
(c) df = z dx + x dy + y dz
(d) df = (x3 + 3x2y2z)dx + (y + 2x3yz)dy + (x3y2)dz
(1.8)
Consider the two differentials, df = y dx − x dy and df = y dx + x dy. One of them is exact; the other
is not. Evaluate the line integral of each over the following paths described by Cartesian coordinates
(x, y):
(a) A straight line from (1, 1) to (2, 2)
(b) A straight line from (1, 1) to (2, 1), then another from (2, 1) to (2, 2)
(c) A closed loop consisting of straight lines from (1, 1) to (2, 1) to (2, 2) and then back to (1, 1)
(1.9)
Change units on the following quantities as directed:
(a) Convert 1.5 × 103 bars to GPa
(b) Convert 3.5 g cm−3 to g bar cal−1
(c) Convert 1.987 cal deg−1 to kJ deg−1
(d) Convert 5.0 × 10−13 cm2 sec−1 to m2(106 yr)−1
11
CHAPTER TWO
HOW ELEMENTS BEHAVE
OVERVIEW
Our goal in this chapter is to refresh concepts that you
may have first encountered in a chemistry course and to
put them in a geologic context, where they will be useful
for the chapters to come. We begin with a review of
atomic structure and the periodic properties of the elements, focusing carefully on the electronic structure of
the atom. This leads us to consider the nature of chemical bonds and the effects of bond type on the properties
of compounds. These should be familiar topics, but they
may look different through the lens of geochemistry.
ELEMENTS, ATOMS, AND THE
STRUCTURE OF MATTER
Elements and the Periodic Table
Long before the seventeenth century, alchemists determined that most pure substances can be broken down,
sometimes with difficulty, into simpler substances. They
reasoned, therefore, that all matter must ultimately be
made of a few kinds of material, commonly called elements. Following the lead of Aristotle, they identified
these as earth, air, fire, and water. Without stretching too
far, we can recognize a parallel to what chemists today
call the three states of matter (solid, liquid, and gas) and
12
energy. These are useful distinctions to make, but unfortunately, medieval alchemists had difficulty explaining
how the Aristotelian elements could combine to form
everyday materials. Two key concepts were missing:
the idea that matter is composed of individual particles
(atoms) and the notion that compounds are made of
atoms held together electrostatically.
Philosophers claim that the atomic theory had its roots
in ancient Greece. The sage Democritus declared that
the motions we observe in the natural world can only be
possible if matter consists of an infinite number of infinitesimal particles separated by void. This was not a scientific hypothesis, however, but a proposition for debate.
Democritus had neither the means nor the inclination to
test his idea. Still, for more than 2000 years, scholars
argued about whether matter was infinitely divisible or
made of tiny, discrete particles. The issue was resolved
only when scientists finally recognized the difference between chemical compounds, on the one hand, and alloys
or mixtures on the other.
Through the seventeenth and eighteenth centuries,
scientists gradually convinced themselves that the Aristotelian earth, in particular, was not a fundamental
substance. Analytical techniques improved and empirical
rules began to emerge. Robert Boyle, for example, declared that a substance cannot be an element if it loses
weight during a chemical change, thus clarifying the
How Elements Behave
distinction between pure elements and compounds. By
the 1790s, systematic purification had yielded roughly a
third of the elements known today. Most of these are
metals: copper, gold, silver, lead, zinc, iron, magnesium,
mercury, and tungsten, to name a few.
Geologists might argue, however, that the greatest
accomplishment of that era was the isolation of nonmetallic elements. In particular, the discovery of oxygen
in the 1780s (attributed independently and with great
controversy to Joseph Priestley, Karl Scheele, and Antoine
Lavoisier) clarified the mysterious relationship between
metallic elements and the clearly nonmetallic minerals
that constitute most of the Earth’s crust. The most common minerals were revealed as complex oxides of metallic elements. This realization, coming in the midst of an
explosion of national investments in mining and metallurgy in Europe, was perhaps the first major step toward
modern geochemistry. It was also a marked change in
direction for science at large. Lavoisier, Joseph Louis
Proust, Jeremias Benjamin Richter, Joseph-Louis GayLussac, and others established by experimentation that
oxygen and other gaseous elements combine in fixed proportions to form compounds. Water, for example, is always 11.2% hydrogen and 88.8% oxygen by weight.
Fixed proportions suggest a systematic arrangement
of discrete particles. In 1805, John Dalton proposed four
postulates:
1. Atoms of elements are the basic particles of matter.
They are indivisible and cannot be destroyed.
2. Atoms of a given element are identical, having the
same weight and the same chemical properties.
3. Atoms of different elements combine with one another
in simple whole-number ratios to form molecules of
compounds.
4. Atoms of different elements may combine in more
than one small whole-number ratio to form more than
one compound.
These postulates have now been revised, particularly in
light of the discovery of radioactivity in the late 1800s.
We now realize, for example, that energy and matter are
interchangeable, and that atoms can be “smashed” into
smaller particles. Still, for most chemical purposes, the
postulates have proved correct.
Once a rational basis for the atomic theory had been
established, the number of newly isolated elements began
to grow. Chemists began recognizing relationships among
the elements. Johann Dobereiner, in 1829, noted several
13
triads of elements with similar chemical properties. In
each triad, the atomic mass of one element was nearly
equal to the average of the atomic masses of the other
two. Chlorine, bromine, and iodine, for example, are all
corrosive, colored, diatomic gases; the atomic mass of
bromine is halfway between the masses of chlorine and
iodine. Over the next 40 years, scientists documented
other patterns when they arranged the elements in order
of increasing atomic mass. Finally, in 1869, Russian
chemist Dmitri Mendeleev made a careful reevaluation
of what was known about the elements, corrected some
erroneous values, and proposed the system known today
as the periodic table (fig. 2.1).
We return several times in this chapter to examine the
periodic properties of elements and to offer explanations
for their behavior in the Earth. For a first look, consider
figure 2.2, in which we compare the melting and boiling
points of elements in the first three main rows of the
periodic table (omitting for now the elements scandium
through zinc: the first transition series). The trend toward increasing values near the center of each row is
unmistakable. Mendeleev was able to predict the properties of germanium, as yet undiscovered in the 1860s,
by using trends like these.
Although the periodic table was developed empirically, its structure suggests that atoms of each element
are built according to regular architectural rules, thereby
offering some hope for figuring out what those rules
are. By the end of the nineteenth century, chemists understood that atoms are made of three fundamental types
of particles. The heaviest of these, protons and neutrons,
are tightly packed into the nucleus of the atom, where
they constitute very little of the atom’s volume but virtually all of its mass. The lightest particles, electrons, orbit
the nucleus in an “electron cloud” that is largely empty
space. Protons and electrons are opposing halves of an
electrical system that controls most of the atom’s chemical properties. By convention, each proton is said to have
a unit of positive charge, and each electron has an equal
unit of negative charge.
The Atomic Nucleus and Isotopes
Three numbers, describing the abundance and type
of nuclear particles in an atom, are in common use. The
first of these, symbolized by the letter Z, counts the
number of protons and is known as the atomic number.
This number identifies which element the atom represents
14
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
IA
1
H
8A
2A
3
Li
4
Be
11
Na
12
Mg
3B
19
K
20
Ca
37
Rb
21
Sc
4B
22
Ti
5B
23
V
6B
24
Cr
7B
25
Mn
38
Sr
39
Y
40
Zr
41
Nb
42
Mo
55
Cs
56
Ba
57
La*
72
Hf
73
Ta
74
W
87
Fr
88
Ra
89
Ac+
104
Db
105
Jl
106
Rf
*Lanthanide series
58
Ce
59
Pr
60
Nd
61
Pm
+Actinide series
90
Th
91
Pa
92
U
93
Np
3A
5
B
4A
6A
6
C
5A
7
N
2
He
8
O
7A
9
F
10
Ne
2B
30
Zn
13
Al
14
Si
15
P
16
S
17
Cl
18
Ar
28
Ni
1B
29
Cu
31
Ga
32
Ge
33
As
34
Se
35
Br
36
Kr
45
Rh
46
Pd
47
Ag
48
Cd
49
In
50
Sn
51
Sb
52
Te
53
I
54
Xe
77
Ir
78
Pt
79
Au
80
Hg
81
Tl
82
Pb
83
Bi
84
Po
85
At
86
Rn
62
Sm
63
Eu
64
Gd
65
Tb
66
Dy
67
Ho
68
Er
69
Tm
70
Yb
71
Lu
94
Pu
95
Am
96
Cm
97
Bk
98
Cf
99
Es
100
Fm
101
Md
102
No
103
Lr
26
Fe
8B
27
Co
43
Tc
44
Ru
75
Re
76
Os
[107] [108] [109]
Bh
Hn
Mt
FIG. 2.1. The elements are described in this periodic table by their atomic number and symbol. The names dubnium (Db), joliotium (Jl),
rutherfordium (Rf), bohrium (Bh), hahnium (Hn) and meitnerium (Mt) for elements 104–109 have been proposed but not yet formally
approved.
FIG. 2.2. Mendeleev suggested that the elements could be
tabulated in a way that emphasizes trends in chemical properties.
For example, with some exceptions (boron’s boiling point and
gallium’s melting point, in column 3), melting and boiling
points increase toward the middle of each row in the periodic
table.
(any atom with Z = 8, for example, is an atom of oxygen, one with Z = 16 is sulfur, and so forth). Because
an electrically neutral atom must have equal numbers of
protons and electrons, Z also tells us the number of electrons in an uncharged atom. The second useful number
is the neutron number, symbolized by the letter N. Within
limits imposed by the attractive and repulsive forces
among nuclear particles, atoms of any element may have
different numbers of neutrons. Atoms for which Z is the
same but N is different are called isotopes.
The numbers Z and N are sufficient to describe atoms,
but chemists find it convenient to use a third quantity, A,
equal to the sum of Z and N. Because protons and neutrons differ in mass by only a factor of one part in 1836,
we assign each the same arbitrary unit mass. The quantity A, therefore, represents the total mass of the nucleus.
Because electrons are so light, A is very nearly the mass
of the entire atom. From a practical perspective, it is
easier to measure the total mass of an atom than to
count its neutrons, so A is a more useful number than N.
In shorthand form, we convey our knowledge about a
How Elements Behave
15
WHAT’S IN A NAME?
A beginning pianist often finds that the greatest obstacle to progress is musical notation. Dotted quarter
notes, the bass clef, and all the strange comments in
Italian are literally a foreign language that must be
mastered by any musician. Beginning chemists often
have a similar problem with the periodic table, their
musical staff. For each element, there are atomic
masses, oxidation states, ionic radii, and a host of
other things to remember. The easiest, perhaps, is the
element’s symbol, but even there a student chemist
may be dazzled by strange conventions.
The classical elements, most of which are metals,
were named so long ago that there is no record of
where their names came from. The words iron, lead,
gold, and tin, for example, are obscure Germanic
names. They were good enough for common folk, but
too plebian for alchemists, who preferred the more
“scientific” sound of the Latin names ferrum, plumbum, aurium, and stannum. From these we get the
symbols Fe, Pb, Au, and Sn. Many modern names have
a Latin or Greek flavor as well, particularly those that
were given as the periodic table went through a season
of explosive growth during the early nineteenth century. Carl Mosander, for example, was so annoyed by
particular atom (or nuclide, as it is called when the focus
is on nuclear properties) by writing its chemical symbol,
a familiar alternative for Z, preceded by a superscript
that specifies A. For a nuclide consisting of 8 protons
and 8 neutrons, for example, we write 16O. For the one
with 26 protons and 30 neutrons, we write 56Fe.
The existence of isotopes was not suspected during
the nineteenth century, but it accounts for some of the
confusion among chemists who were trying to make sense
of periodic properties. Chlorine, for example, has two
common isotopes: 35Cl and 37Cl. Because 35Cl is roughly
three times as abundant as 37Cl, the weighted average
mass of the two in a natural chlorine sample is close to
35.5. A natural sample of any element, in fact, will yield
a nonintegral atomic mass for this reason. (Atomic mass
is measured in atomic mass units [amu], equal to 1/12 of
the mass of a 12C atom, or 1.6605 × 10−24 g.)
the difficulty of isolating the element lanthanum that
he gave it a name derived from the Greek verb lanthanein (“to escape notice”). French chemist Lecoq
de Boisbaudran had a similar experience with dysprosium, for which he crafted a name from the Greek
word dysprositos (“hard to get at”).
A geochemist may find it amusing to track
down elements named for mining districts. In its
first appearance, for example, copper was called aes
Cyprium—literally, “metal from Cyprus.” Magnesium was named for the Magnesia district in Thessaly. The grand prize, however, has to go to a black
mineral collected from a granite pegmatite near Stockholm in 1794 by the Finnish mineralogist Johan
Gadolin. He named it yttria, after the nearby village
of Ytterby. In 1843, Mosander isolated three new
elements from yttria, calling them yttrium, erbium,
and terbium. Almost forty years later, Swiss chemist
Jean Charles de Marignac isolated yet another element, ytterbium, from the same ore. To complete the
package, Marignac isolated one more element in 1880
and named it gadolinium for the original discoverer
of the mineral (which is now known as gadolinite).
Quite a yield from one little mining district!
Worked Problem 2.1
A geochemist analyzes a very large number of minerals containing lead and finds four isotopes in the following relative
abundances:
204Pb
(mass = 203.973 amu) 1.4%
206Pb
(mass = 205.974 amu) 24.1%
207Pb
(mass = 206.976 amu) 22.1%
208Pb
(mass = 207.977 amu) 52.4%
Except for 12C, the masses of nuclides are never integral.
Notice, for example, that the mass of 204Pb is slightly less than
204 amu. This is because a small fraction of the mass of any
atom is actually in the binding energy that holds its nucleus
together. (Recall Albert Einstein’s famous equation, e = mc2.)
Given the nuclide masses in parentheses, then, what is the
atomic mass of lead in nature, as implied by these analyses?
FIG. 2.3. The stable nuclides, shown in black, define a narrow band within a wider band of unstable nuclides in this plot of protons (Z) versus neutrons (N). In general, elements with Z or N
even have more stable nuclides than those with Z or N odd.
How Elements Behave
To calculate, multiply the mass of each nuclide by its proportion in the sample and sum the results:
203.973 × 0.014 + 205.974 × 0.241 + 206.976 × 0.221
+ 207.977 × 0.524 = 207.217 amu.
Roughly 1700 nuclides are known. Of these, only
∼260 are stable. The rest have nuclei that can disintegrate spontaneously to produce subatomic particles and
leave a nuclide in which Z or N has changed. Figure 2.3,
a plot of the number of neutrons versus protons in each
of the known nuclides, illustrates this point. Stable nuclides only occur within a thin diagonal band, in which
N is slightly greater than Z, flanked on both sides by
unstable radioactive nuclides. Careful inspection of a
segment of this band (fig. 2.4) shows that, in most cases,
the stable nuclides have an even number of neutrons
and protons. Stable nuclides with either Z or N as odd
numbers are much less common, and isotopes with both
Z and N odd are generally unstable. This distribution is
shown numerically in table 2.1.
Most of the known radioactive isotopes do not occur
in nature, although all have been produced artificially
in nuclear reactors. Some of the “missing” isotopes may
have occurred naturally in the distant past, but have such
17
TABLE 2.1. Abundance of Stable Nuclides
Z (even)
Z (odd)
N (even)
N (odd)
Total
159
50
53
4
212
54
high decay rates that they have long since become extinct.
The radioactive isotopes of greatest interest to geochemists generally decay very slowly or are replenished
continually by natural nuclear reactions.
Geochemists study both stable and radioactive isotopes in the Earth. In chapter 13, we look carefully at
ways in which stable isotopes are fractionated in nature
and how we can measure their relative abundances to
figure out which processes may have affected geologic
samples. Stable isotopes can also be used as tracers to
infer chemical pathways in the Earth or to follow the
progress of reactions. Radioactive isotopes, the topic of
chapter 14, can also be used as tracers. Their primary
value, however, is in dating geologic samples. Radiometric dating methods, suggested by Ernest Rutherford
in the nineteenth century and put in useful form by
Bertram Boltwood and Arthur Holmes early in the
twentieth, are now the basis for all “absolute” ages in
geology.
FIG. 2.4. Expanded segment of the nuclide chart of figure 2.3, showing stable nuclides and their
percentages of relative natural abundances. The average atomic weight for each element (shown in
the boxes to the left of the figure) is the sum of the weights of the various isotopes times their respective relative abundances.
18
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
The Basis for Chemical Bonds:
The Electron Cloud
An atom is electrically independent of its neighbors
if the number of electrons around it is the same as the
number of protons in its nucleus. If not, then an excess
positive charge on one atom must be balanced by negative charges elsewhere. This is the fundamental basis for
bonding, the force that holds atoms together to form
compounds.
Why should an atom ever have “too many” or “too
few” electrons? Even as chemists at the start of the
twentieth century gained confidence in the predictive
capacity of the periodic table, stoichiometry (the rules
governing how many atoms of each element belong in a
compound) remained a stubborn mystery. It was easy to
see, for example, that the composition of hydrides (and
other simple compounds) follows a simple pattern across
each row of the periodic table. Lithium combines with
one hydrogen atom to form LiH, beryllium forms BeH2,
boron makes BH3, and carbon CH4. From there to the
end of the row, the combining ratio drops: NH3, H2O,
and HF. There is no neon hydride. In the next row, we
see NaH, MgH2, AlH3, SiH4, PH3, H2S, and HCl. There
is no argon hydride. As the transition elements are added
in successive rows, the pattern becomes less obvious, but
is still there. Why? In fact, why are any of the periodic
properties periodic?
The answer lies in the distribution of electrons around
the nucleus. Chemists began to glimpse the structure of
the “cloud” of orbiting electrons in 1913, when Niels
Bohr first successfully attempted to explain why atomic
spectra consist of discrete lines instead of a continuous
blur (fig. 2.5). Building on theoretical advances by Max
Planck and Albert Einstein in the previous decade, he
made the novel assumption that the angular momentum
of an orbiting electron can only have certain fixed values.
The orbital energy associated with any electron, similarly, cannot vary continuously but takes only discrete
quantum values. As a result, only specific orbits are possible. In mathematical terms, Bohr postulated that the
angular momentum of the electron (mevr) must be equal
to nh/2π, where me is the mass of the electron, v is its
linear velocity, r is its distance from the nucleus, h is
Planck’s constant, and n takes positive integral values
from 1 to infinity. Because n can only increase in integral
steps, r defines a set of spherical shells at fixed radial distances from the nucleus.
Clever though this approach was, chemical physicists
soon recognized that an electron cannot be described
correctly as if it were a tiny planet orbiting a nuclear star.
A better model uses the language of wave equations
developed to describe electromagnetic energy. In 1926,
Erwin Schrödinger proposed such a model incorporating not only the single quantum number as Bohr had
suggested, but three others as well. Schrödinger’s equations describe geometrical charge distributions (atomic
orbitals) that are unique for each combination of quantum numbers. The electrical charge that corresponds to
each electron is not ascribed to an orbiting particle, as in
Bohr’s model. Instead, you can think of each electron as
being spread over the entire orbital—a complex threedimensional figure.
The principal quantum number, n, taking integral
values from 1 to infinity, describes the effective volume
of an orbital—commonly calculated as the volume in
which there is a 95% chance of finding the electron at any
instant. This is similar to Bohr’s definition of n. Chemists
FIG. 2.5. Emission spectra for calcium, potassium, and sodium.
How Elements Behave
FIG. 2.6. The three p orbitals in any shell have figure-eight
shapes, oriented along Cartesian x, y, or z axes centered on the
atomic nucleus.
commonly use the word shell to refer to all orbitals with
the same value of n, because each increasing value of n
defines a layer of electron density that is farther from the
nucleus than the last n.
The orbital angular momentum quantum number, l,
determines the shape of the region occupied by an electron. Depending on the value of n, a particular electron
can have values of l that range from zero to n − 1. If l = 0,
the orbital described by Schrödinger’s equations is spherical and is given the shorthand symbol s. If l = 1, the orbital looks like one of those in figure 2.6 and is called a
p orbital. Orbitals with l = 2 (d orbitals) are shown in figure 2.7. The l = 3 orbitals (f orbitals) have shapes that
are too complicated to illustrate easily here.
The third quantum number, ml, describes the orientation of the electron orbital relative to an arbitrary
direction. Because an external magnetic field (such as
might be induced by a neighboring atom) provides a
convenient reference direction, ml is usually called the
magnetic orbital quantum number. It can take any integral value from −l to l.
The fourth quantum number, ms, does not describe
an orbital itself, but imagines the electron as a particle
within the orbital spinning around its own polar axis,
19
much like the Earth does. In doing so, it becomes a tiny
magnet with a north and a south pole. The magnetic
spin quantum number, ms, can be either positive or
negative, depending on whether the electron’s magnetic
north pole points up or down relative to an outside magnetic reference.
Each orbital therefore can contain no more than two
electrons, with opposite spin quantum numbers. This
rule, which affects the order in which electrons may fill
orbitals, is known as the Pauli exclusion principle. The
simplest atom, hydrogen, has one electron in the orbital
closest to the nucleus (n = 1). With each increase in atomic
number, atoms gain an electron in the unfilled orbital
with the lowest energy level. We code electron orbitals
with a label composed of the numerical value of n and a
letter corresponding to the value of l, so the lowest orbital is called the 1s orbital. It can hold up to two electrons, as shown in table 2.2. The atom with atomic
number 3, lithium, has two electrons in the 1s orbital
and a third in 2s, which has the next lowest energy level.
It takes another 7 electrons to fill the n = 2 shell with s
and p electrons, and 18 more to fill the n = 3 shell with
s, p, and d electrons.
You can extend the table to calculate the number of
s, p, d, and f orbitals in the n = 4 shell. The order in
which orbitals beyond 3p fill, however, is not what you
might expect from this exercise. As illustrated in figure 2.8, complex interactions among electrons and the
nucleus reduce the energy associated with each orbital
as atomic number increases. In neutral potassium and
calcium atoms, the first to have >18 electrons, the 4s
orbitals have lower energy levels than the 3d orbitals, so
the nineteenth and twentieth electrons fill them instead.
With increasing atomic number, electrons fill the 3d,
4p, 5s, 4d, and 5p orbitals. This order could only be
anticipated by calculating the relative energy levels of
FIG. 2.7. Three of the d orbitals (dxy, dyz, and dzx) in any shell have four lobes, oriented between the Cartesian axes. A fourth (dx2−y2) also
has four lobes, but along the x and y axes. The fifth (dz2) has lobes parallel to the z axis and a ring of charge density in the x-y plane.
20
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
TABLE 2.2. Configuration of Electrons in Orbitals
n
l
m1
ms
1s orbital
1
1
0
0
0
0
+
−
2s orbital
2
2
0
0
0
0
+
−
2
2
1
1
−1
−1
+
−
2
2
1
1
0
0
+
−
2
2
1
1
1
1
+
−
3
3
0
0
0
0
+
−
3
3
1
1
−1
−1
+
−
3
3
1
1
0
0
+
−
3
3
1
1
1
1
+
−
3
3
2
2
−2
−2
+
−
3
3
2
2
−1
−1
+
−
3
3
2
2
0
0
+
−
3
3
2
2
1
1
+
−
3
3
2
2
2
2
+
−
2p orbitals
3s orbital
3p orbitals
3d orbitals
successive orbitals, as was done in producing figure 2.8.
Once recognized, however, the order provides insight
into the layout of the periodic table. If you compare
table 2.2 with the periodic table, you will see an interesting pattern. Each row in the periodic table begins
with a new s orbital (n increases by 1 and l = 0), and ends
as the last p orbital for that value of n is filled. The s
and p orbitals, in other words, define the basic outline
of the periodic table. Atoms whose last-filled orbitals are
d or f orbitals make up the transition elements that occupy the central block in rows 4 and beyond.
Why should a geochemist care about this level of
detail? One reason is that electronic configuration offers
a way to interpret bonding and, with it, the driving force
for chemical reactions. Site occupancies in minerals and
mineral structures themselves are best understood when
we predict the shapes and orientations of electron orbitals. We explore this general topic later in this chapter.
Another reason is that an atom’s electronic configuration
determines its affinity for atoms of other elements. By
studying the periodic properties of the elements, we can
begin to understand how they are distributed in the Earth.
Size, Charge, and Stability
The outermost electrons in an atom determine its
effective size, or atomic radius, as well as much of its behavior in bonding. These electrons are sometimes called
the valence electrons. Figure 2.9 illustrates two trends
with increasing atomic number. First, as s and p electrons
are added across a single row of the periodic table, atomic
radius decreases. The simple explanation for this is that
the positive charge of the nucleus continues to increase
as atomic number increases. Electrons in the same shell
cannot effectively shield each other from the pull of the
nucleus, so valence electrons are drawn in as nuclear
charge increases, shrinking atoms from left to right across
a row. (Shielding effects among electrons in d and f or-
FIG. 2.8. As atomic number increases, electrons are increasingly
shielded from the charge of the nucleus by the presence of other
electrons. As a result, the energy (E) necessary to stabilize an
electron in each orbital generally decreases with increasing
atomic number. This rule is made more complex by interactions
between the orbitals themselves. Energy levels in the 3d orbitals,
for example, increase across much of row 3 in the periodic table.
Potassium and calcium, which in their ground state have enough
electrons to enter 3d orbitals, instead place them in 4s orbitals,
which have a lower energy.
Radius Increases
0.37
H
How Elements Behave
21
0.80
B
0.77
C
0.74
N
0.74
O
0.72
F
_
He
_
Ne
1.25
Al
1.17
Si
1.10
P
1.04
S
0.99
Cl
Atomic Radii
1.23
Li
0.89
Be
1.57
Na
1.36
Mg
2.03
K
1.74
Ca
1.44
Sc
1.32
Ti
1.22
V
1.17
Cr
2.16
Rb
1.91
Sr
1.62
Y
1.45
Zr
1.34
Nb
2.35
Cs
_
1.98
Ba
_
1.44
Hf
_
Fr
Ra
1.69
La*
_
Ac+
Db
Radius Decreases
1.16
Co
1.15
Ni
1.17
Cu
1.25
Zn
1.25
Ga
1.22
Ge
1.21
As
1.17
Se
1.14
Br
1.29
Mo
1.17 1.17
Mn
Fe
_
1.24
Tc
Ru
1.25
Rh
1.28
Pd
1.34
Ag
1.41
Cd
1.50
In
1.41
Sn
1.41
Sb
1.37
Te
1.33
I
_
Ar
_
Kr
_
Xe
1.34
Ta
_
1.30
W
_
1.28
Re
_
1.26
Os
_
1.26
Ir
_
1.29
Pt
1.34
Au
1.44
Hg
1.55
Tl
1.54
Pb
1.52
Bi
1.53
Po
_
At
_
Rn
Jl
Rf
Bh
Hn
Mt
1.65
Ce
1.64
Nd
+Actinide series
1.65
Th
1.65
Pr
_
_
*Lanthanide series
Pm
_
1.66
Sm
_
1.85
Eu
_
1.61
Gd
_
1.59
Tb
_
1.59
Dy
_
1.58
Ho
_
1.57
Er
_
1.56
Tm
_
1.70
Yb
_
1.56
Lu
_
Np
Pu
Am
Cm
Bk
Cf
Es
Fm
Md
No
Lr
Pa
1.42
U
FIG. 2.9. Atomic radii, indicated here in Ångströms (Å), decrease across each row in the periodic table as nuclear charge increases. Each
added shell of electrons, however, is shielded from the nucleus by inner electrons. Atomic radii therefore increase from one row to the
next. Chemists generally determine the size of atoms by studying the dimensions of their coordination sites in crystal structures (discussed later in this chapter). It is difficult to measure atomic radii for elements that are rare in nature (the actinides, for example) or that
do not bond with other elements to form crystals (the noble elements). For this reason, no radii are reported for those elements in this
figure.
bitals are more complex, so this trend is less consistent
across the transition elements.) Increasing atomic number beyond the end of a row, however, means adding
electrons to a new shell, farther from the nucleus. As a
result, elements in the next row in the periodic table
have larger atomic radii. Within a single column of the
periodic table, therefore, atomic radius increases with
increasing atomic number. This second trend is also
apparent in figure 2.9.
The measure of the energy needed to form a positively charged ion (a cation) by removing electrons from
an atom is called ionization potential (IP), plotted in figure 2.10. When electrons are closest to the nucleus, they
are hardest to remove with energy supplied from the
outside. For this reason, it is easier to produce cations of
atoms toward the bottom or the left side of the periodic
table than from atoms toward the top or the rightmost
columns.
IP is greatest in the right hand column of the periodic
table, among the group of elements known as the noble
FIG. 2.10. The energy required to remove one electron from an
atom increases across each row of the periodic table, but decreases slightly from one row to the next because successive
shells of electrons are held less tightly by the positive charge of
the nucleus. The quantity plotted here is the first ionization potential. A plot of the second or third ionization potential would
be similar to this one, although removal of a second or third
electron requires more energy.
22
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
gases (He, Ne, Ar, Kr, Xe, and Rn). In this group, all appropriate s and p orbitals are filled. This configuration,
known as a complete outer shell, is particularly stable.
Atoms on the left side of the periodic table form cations
readily, losing enough electrons that their outer shell
looks like the noble gas at the end of the previous row.
In the second row, for example, lithium easily loses one
electron to become Li+, beryllium becomes Be2+, and
boron becomes B3+, in each case ending up with an outer
shell that looks like that of neutral helium.
The converse of ionization potential is electron affinity (EA), the amount of energy released if we succeed
in adding a valence electron to a neutral atom to create
a negatively charged anion. Like IP, EA also increases to
the right in the periodic table. The highest affinities are
among the halogens (F, Cl, Br, and I), which only need to
gain a single electron to reach a noble gas configuration.
These and the four elements to their immediate left (O,
S, Se, and Te) are the only ones that commonly form
anions. For all other elements, EA is very small or negative, indicating that they gain little or no stability by
gaining electrons.
Ionization changes the size of an atom. If it loses electrons, the net positive excess charge in its nucleus draws
the remaining electrons in tighter. The overall trend in
cation radii, therefore, is the same as for atomic radii:
they increase toward the bottom and decrease toward the
right side of the periodic table. Gaining electrons has the
opposite effect, so anions are larger than their neutral
counterparts. Figure 2.11 illustrates how ionic radii vary
with atomic number.
Elemental Associations
Trends among the periodic properties suggest logical
ways to group elements. For example, it is customary
to speak of all elements in the leftmost column of the
periodic table as alkali metals and those in the second
column as alkaline earth metals. The rightmost column
we have already identified as the noble gases; the column
to its left contains the halogens. In each of these groups,
elements have the same configuration of valence electrons. The alkali metals (Li, Na, K, Rb, Cs, and Fr), for
example, each have a lone electron in the outermost shell
and therefore readily form cations with a +1 charge.
Chemists also speak of the transition metals, a group of
elements in the middle of rows 4, 5, and 6, across which
the d electron orbitals are gradually filled. Similarly, the
FIG. 2.11. As electrons are removed from an atom, the remaining
electrons are drawn more tightly by the charge of the nucleus, so
that ionic radius decreases with increasing positive charge. Compare, for example, the radii of Fe2+ and Fe3+ or Mn2+ and Mn4+.
Also, compare each of the ionic radii in this figure with the
atomic radii in figure 2.9.
lanthanides and actinides in rows 6 and 7 are groups
across which the f orbitals are filled. Each of these groups,
labeled in figure 2.12, contains elements that, by virtue
of their common electron configuration, behave in similar ways during chemical reactions and therefore form
similar compounds.
Appreciation for the periodic properties of the elements has cast light on many geochemical mysteries. One
of the most fundamental observations that geologists
make about the Earth, for example, is that it is a differentiated body with a core, mantle, and crust that are
chemically distinct. Why? Is there a rational way to
explain why such elements as potassium, calcium, and
strontium are concentrated in the crust and not in the
core? Or why copper is almost invariably found in sulfide ores rather than in oxides?
How Elements Behave
23
FIG. 2.12. Elements with similar properties fill similar roles in nature and are therefore commonly recognized as members of groups, as
indicated in this figure. The lanthanides are often referred to as rare earth elements (REE).
Early in the twentieth century, Victor Goldschmidt
was prompted by a study of differentiated meteorites
to propose a practical scheme for grouping elements
according to their mode of occurrence in nature. His
analyses of three kinds of materials—silicates, sulfides,
and metals—suggested that most elements have a greater
affinity for one of these three materials than for the other
two. Calcium, for example, can be isolated as a pure
metal with great difficulty, and its rare sulfide, oldhamite
(CaS), occurs in some meteorites. Calcium is most common, however, in silicate minerals. Gold, however, is
almost invariably found as a native metal.
Goldschmidt’s classification, summarized below, ultimately included a fourth type of material not found in
abundance in meteorites:
1. Lithophile elements (from the Greek lithos, meaning
“rock”) are those that readily form compounds with
oxygen. The most common oxygen-based minerals
in the Earth are silicates, but lithophile elements also
dominate in oxides and other “stony” minerals.
2. Chalcophile elements are those that most easily form
sulfides. The Greek root for the name of this group is
chalcos, meaning “copper,” a prominent element in
this category.
3. Siderophile elements prefer to form metallic alloys
rather than stoichiometric compounds. The most
abundant siderophile element is iron (in Greek,
sideros), for which this group is named.
4. Atmophile elements are commonly found as gases,
whether in the atmosphere or in the Earth.
24
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 2.13. Goldschmidt’s classification of the elements emphasizes their affinities with three fundamental types of Earth materials (sulfides, silicates, or native metals) or with the atmosphere. Several elements—the most abundant being iron—have affinities with more
than one group.
As figure 2.13 suggests, these four categories overlap
quite a bit. Tin, for example, is commonly found as an
oxide (the mineral cassiterite, SnO2) and less often as a
sulfide (stannite, Cu2FeSnS4). Carbon, as graphite or
diamond, occurs as a pure element and it alloys readily
with other metals, as in the manufacture of steel, but it
is also common as a lithophile element in carbonates.
Iron forms abundant silicate and sulfide minerals, but
it is also the major constituent of the Earth’s metallic
core. Nevertheless, most elements belong primarily in
one group, and Goldschmidt’s classification is a guide to
geochemical behavior.
Why does this geochemical classification work? Again,
it is based loosely on the periodic properties of elements
and, in particular, on those properties that control bonding between elements. We pursue this topic in the final
sections of this chapter.
BONDING
Perspectives on Bonding
A bond exists between atoms or groups of atoms
when the forces between them are sufficient to create an
atomic association that is stable enough for a chemist
to consider it an independent entity. This definition gives
us a lot of room for subjective judgment about what
“stable enough” and “independent” mean, and it covers
a range of situations in which the types and strength of
forces differ. In general, chemists find this flexibility to
be an advantage. Students of geochemistry, however,
can be led into uncertainty about the role of individual
atoms in natural materials. In crystalline solids, for
example, the “independent entity” is not a freestanding
molecule but an extended periodic array of atoms in
How Elements Behave
which bonds may range subtly both in strength and in
character. One source of confusion is the apparently
sharp distinction between ionic and covalent bonds that
many students learn in basic chemistry courses. Briefly
and in the simplest terms, you may have learned that an
ionic bond is one in which one or more electrons is transferred from a cation to an anion; a covalent bond is one
in which electrons are shared between adjacent atoms.
In both cases, the goal is for each atom to end up with a
noble gas configuration in its outer shell of electrons. In
fact, the difference between ionic and covalent bonding
is more a matter of degree than a clear distinction.
A common way to determine whether a bond is ionic
or covalent is based on electronegativity (EN). EN is
proportional to the sum of IP and EA. Because affinity
is difficult to measure, however, EN is usually calculated
directly from bond energies. Figure 2.14 summarizes EN
values for many elements.
The greater the EN difference between atoms, the
more likely it is that electrons will be transferred from
one atom to the other, forming an ionic bond. Typically,
a bond is considered ionic if the EN difference in it is
>2.1. For example, when a chlorine atom (EN = 3.16) is
adjacent to a sodium atom (EN = 0.93), it easily draws
the lone electron from sodium to itself. This is apparent in the electronegativity difference (∆EN) of 2.23 between the atoms. The Mg-O and Ca-O bonds that are
common in rock-forming silicates are also ionic (∆EN =
2.13 and 2.44, respectively). Si-O and Al-O bonds in the
same minerals have ∆EN = 1.54 and 1.83, however, and
are therefore covalent, by this definition.
25
Nobel Prize-winning chemist Linus Pauling suggested
an alternative way to describe a bond by calculating its
percentage of ionic character with the empirical formula:
p = 16|ENa − ENb | + 3.5|ENa − ENb | 2,
in which a and b are atomic species to be bonded. For
Na-Cl, we find that p = 54.7% and for Ca-O, p = 59.9%,
whereas for Si-O, p = 32.9% and for Al-O, p = 41.0%.
If we look at bonding in this way, we see “ionic” and
“covalent” as relative terms on a continuum of bond
types. This offers a more realistic perspective on variations in bonding, but can present conceptual difficulties
when we consider mineral structures. It is customary,
for example, for geochemists to talk about the “ionic
radius” of Mg2+, as we have earlier in this chapter. If an
Mg-O bond is described in Pauling’s terms, however, p =
49.9%. Roughly half of the electron charge transferred
from the Mg2+ ion, therefore, is actually shared with the
O2− ion, so it is difficult to conceive of either ion as an
independent entity with a clearly defined radius. The
standard practice of referring to “ions” in mineral structures as if they were the same as free ions in aqueous
solution is clearly misleading.
The percentage that Pauling’s formula estimates can
be thought of in another way: as the polarity of the bond.
Geochemists who deal with gases or aqueous solutions
find that a knowledge of polarity helps to explain solubility, reactivity, and other key properties. Those who deal
with solid materials can use the same sort of information
to interpret optical, thermal, and electric properties. Polarity can greatly affect the boiling point and conductivity
FIG. 2.14. Electronegativity generally increases with atomic number, within a single row of the
periodic table. It may be thought of as a measure of the ease with which an atom can add electrons to complete its outer shell.
26
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
of a substance, for example, and the way it combines
with other substances, as in H2O (see the box on this
topic, later in the chapter).
Structural Implications of Bonding
If bonds in a substance have a highly ionic character,
it is convenient to view the atoms like marbles arranged
according to size and electrical charge. Each ion must be
surrounded by oppositely charged ions, so that electrical
charge is balanced locally, but the number of surrounding ions depends on their relative sizes. This is the case
because ionic bonding leads to the most stable structure
when the ions are packed together as closely as possible.
In the halite (NaCl) structure, for example, each sodium
atom is at the center of a cluster of six chloride ions and
each chloride ion is surrounded by six sodium ions. The
positive charge on each sodium ion is distributed evenly
over six Na-Cl bonds, and so is the negative charge on
each chloride ion. But why six ions?
Figure 2.15a shows a perspective drawing of the halite
structure, with cations and anions drawn as spheres. The
distance between ions, which touch each other in the actual structure, has been exaggerated for ease of viewing.
Figure 2.15b isolates a portion of the structure: a cation
at the geometric center of an octahedral array of anions.
In figure 2.15c, we have removed the top and bottom
anions for clarity, and in figure 2.15d, we have rotated
that view to be viewed directly from above. In this final
perspective, we have restored interatomic distances to
show the cation-anion contact.
If we were to decrease the ionic radius ratio r+/r−
somehow, it should be apparent that the distance d between the anions would gradually shrink to zero, at
which point they would be touching. At that point, the
length of each side of the square connecting their centers
would be 2r+. By the Pythagorean Theorem, it would
also be equal to √2
(r+ + r−). Algebraic manipulation of
the two expressions reveals that the value of r+/r− would
have become √2
− 1 = 0.414. We could not decrease r+/r−
further without either allowing the anions to overlap or
creating free space between the anions and the cation, so
that they no longer touched. The first of these conditions
is self-limiting, because the valence electrons of the anions
repel each other, preventing further overlap. The second
condition is far more likely. In fact, it is not uncommon
for a small cation to rattle around in an oversized site sur-
FIG. 2.15. (a) The structure of sodium chloride—the mineral
halite—is based on a cubic face-centered lattice. Na+ and Cl− ions
alternate along all three coordinate axes. As a result, each ion is
surrounded by six ions of the other element. (b) A single Na+ ion
and the six Cl− ions surrounding it, “exploded” to illustrate the
octahedral arrangement of anions. (c) The top and bottom Cl− ions
have been removed for clarity, revealing a square planar array of
Cl− ions around the Na+. (d) A view straight down on the planar
arrangement.
rounded by anions that touch each other. In such cases,
however, the cation-anion bonds are longer and thus
weaker than they are when the anions are farther apart.
This can be seen, for example, in the abnormally low
melting points of lithium halides (table 2.3).
For ionic compounds with r +/r − < 0.414, fourfold
(tetrahedral) coordination is more likely than the sixfold (octahedral) coordination of the NaCl structure,
because the cation no longer rattles around in an oversized site. We have to be cautious about applying geometric arguments for such compounds, however, because
cations tend to form bonds with a more covalent character when they are in fourfold coordination. Unlike
purely ionic bonds, which are electrostatic and therefore
nondirectional, covalent bonds are directional. As we
show shortly, this adds further complexity to the study
of structure.
What if we increase r+/r−, though, instead of decreasing it? If the central cation becomes large enough, it is
eventually possible to create a more stable site by surrounding it with eight, rather than six anions. In prob-
How Elements Behave
27
TABLE 2.3. Melting Points of Some Alkali Halides
CsBr
RbBr
KBr
NaBr
LiBr
Melting
Point (°C)
r+/r−
627
681
748
740
535
0.89
0.80
0.70
0.52
0.39
Melting
Point (°C)
r+/r−
621
638
693
653
450
0.79
0.71
0.63
0.46
0.35
CsI
RbI
KI
NaI
LiI
CsCl
RbCl
KCl
NaCl
LiCl
Melting
Point (°C)
r+/r−
638
717
768
803
606
0.96
0.86
0.76
0.56
0.42
Ionic radii are from Shannon (1976).
Cesium ions are in eightfold, rather than sixfold, coordination in halide compounds.
lem 2.9, at the end of this chapter, you are challenged
to show that this arrangement is preferred when r+/r− >
√3
− 1 = 0.732. As indicated in table 2.3, this is the case
for each of the cesium halides.
We can anticipate the coordination number and geometric arrangement of anions around any ion in an ionic
solid, then, by referring to the limiting ratios r+/r− shown
in table 2.4. These coordination guidelines are only approximate, because ions are not like the rigid marbles we
have assumed in this discussion. Instead, ions expand or
shrink slightly to fit their environment. This makes it
difficult to apply the guidelines when r+/r− is close to one
of the limiting values, as is the case for RbBr and RbCl,
which have the halite structure rather than the predicted
cesium halide structure. Still, it is surprising how well
simple radius ratios can help us anticipate a coordination number even for groups of ions with relatively little
ionic character.
Another approach to understanding the geometric
implications of bonding builds on valence-bond theory,
an extension of the quantum model described earlier
in this chapter. According to valence-bond theory, a
chemical bond occurs when electron orbitals in adjacent
atoms overlap, creating new valence electron orbitals
that encompass both atoms. Only certain orbitals are allowed to overlap, however. The geometric arrangement
TABLE 2.4. Coordination Relationships in Ionic Compounds
Radius Ratio
r+/r−
0.15–0.22
0.22–0.41
0.41–0.73
0.73–1.0
1.0
Arrangement of Anions
Coordination
Number
Corners of a triangle
Corners of a tetrahedron
Corners of an octahedron
Corners of a cube
Corners and edges of a cube
3
4
6
8
12
of those orbitals determines how atoms cluster around
each other.
To explore briefly how this theory works, let’s look
back at the quantum model for a single atom. Recall that
for each combination of n, l, and ml, there are two possible values of ms, the magnetic spin quantum number.
Two electrons with the same n, l, and ml, but opposite
values of ms are said to be paired. Each shell, for example, can hold up to six electrons in the orbitals labeled
px, py, and pz in figure 2.6. Because the second electron
in an orbital occupies a slightly higher energy level than
the first, electrons find an advantage to being distributed
evenly among the three orbitals. If an atom doesn’t have
enough electrons to fill all three orbitals completely, then
it accepts only one electron each in px, py, and pz before
adding a second electron to any of them. Why is all of
this important? Because the electrons in the unpaired orbitals are the ones that can overlap between atoms to
form bonds with a covalent character.
A neutral oxygen atom, for example, has eight electrons. Two of them fill the s orbital in the first shell, another two enter the s orbital in the second shell, and the
other four are in second shell p orbitals. Two are paired
in 2px, one is in 2py, and one in 2pz. Oxygen is therefore
divalent; covalent bonds can form at right angles, along
the partially occupied py and pz orbitals. Nitrogen, which
has one fewer electron than oxygen, is trivalent; its 2p
electrons are unpaired—one each in px, py, and pz. In a
molecule such as NH3, mutually perpendicular covalent
bonds form along each orbital lobe.
Valence-bond theory predicts molecular structures,
then, by examining the number and arrangement of
unpaired electrons in each atom. It also takes into account the energy level of each orbital and their orientations as they overlap. The union of new valence orbitals
28
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
constitutes a new entity, a bonding orbital, that has a
lower energy and is therefore more favored than the valence orbitals considered independently. In a molecule
of H2, for example, each atom contributes an unpaired
1s electron, and their orbitals overlap to form s-s bonding orbitals. The H2 molecule is stable because the pair
of bonding orbitals has a lower potential energy than the
combination of two partially occupied valence orbitals.
This will be a useful perspective to recall in chapter 3,
where we introduce the concept of Gibbs free energy and
explore the driving force for chemical reactions.
There are many ways to form bonding orbitals, and
a discussion of them is beyond the scope of this chapter. The following example emphasizes the usefulness of
valence-bond theory in geochemistry—perhaps enough
to encourage you to make a more rigorous study of the
topic.
two electrons in partially filled orbitals—silicon’s configuration
is 1s2 2s2 2px2 2py2 2pz2 3s2 3px1 3p1y —so, by analogy with what
we have just illustrated about oxygen and nitrogen, silicon
should be divalent. In fact, however, it has a formal valence of
+4, as if the atom had two more unpaired electrons. Why?
The solution to this paradox lies in the formation of hybrid
valence orbitals (fig. 2.16). In compounds of silicon, an electron
is “promoted” from an s orbital to a vacant p orbital, giving
the outer shell the potential structure 3s1 3px1 3py1 3p1z . It takes
energy to redistribute electrons in this way, however, and this
energy is obtained in part by making each of the four unpaired
electrons equivalent to the others. The new entities are sp3 hybrid
orbitals, oriented toward the four corners of a regular tetrahedron. As these form bonding orbitals to oxygen atoms, they
become the basis for the SiO4 tetrahedral unit that is common
to all silicate minerals.
We might have tried to predict this tetrahedral arrangement
of bonds by comparing ionic radii. The ionic radius of Si4+ is
0.26 Å; the radius of O2− is 1.35 Å. A quick check of table 2.3
indicates that their ratio (= 0.19) is only slightly smaller than
expected for fourfold coordination. Although this result is reassuring, it is somewhat misleading. Using r+/r− ignores the
fact that the Si-O bond is highly covalent and that the Si4+ ion,
therefore, cannot be considered as spherically symmetrical.
Valence-bond theory is a more appropriate tool in this case.
Worked Problem 2.2
Silicon, a key element in rock-forming minerals, poses an interesting conceptual challenge in valence-bond theory. It has only
py
s
Four tetrahedral
sp3 orbitals
px
An sp3 orbital
pz
FIG. 2.16. Bonding electrons in a silicon atom are in 3px and 3py orbitals. There are no electrons in 3pz. The paired electrons in 3s are
not available for bonding. If the s and p orbitals are combined, then each of the four electrons becomes a bonding electron in a hybrid
sp3 orbital. Each hybrid orbital has two lobes, one of which is larger than the other. The four large lobes point toward the corners of a
tetrahedron.
How Elements Behave
So far, we have emphasized interactions between pairs
of atoms. In extended structures, such as crystals and
most organic molecules, however, each atom is affected
not only by its nearest neighbors but also by those farther away. Because electrons are generally bound closely
to their parent atoms, these distant forces are minimal.
In some materials, though, extended molecular orbitals
may include several atoms. For example, the benzene
(C6H6) molecule (fig. 2.19a) is described classically as a
ring of carbon atoms held together by covalent single
and double bonds. As the structural diagrams in figure
2.19b indicate, there are five different ways to arrange
the six bonds in the ring. Valence-bond theory suggests
that each of them is equally probable and that the true
WATER IS A STRANGE SUBSTANCE
It is virtually impossible to find a geochemical environment in which water does not play a role, whether
at low or high temperature, on the Earth’s surface, or
deep in the crust. Although water is a familiar substance, it has unusual properties that can only be
understood by studying hybrid bonding orbitals.
A neutral oxygen atom contains unpaired electrons
in 2py and 2pz orbitals. When these bond to hydrogen atoms, the result should be a pair of H-O bonds
at right angles. Instead, interelectronic repulsion between the filled 2s2 and 2px2 orbitals favors the formation of sp3 hybrid orbitals such as those discussed
in worked problem 2.2. The oxygen atom in a water
molecule therefore finds itself at the center of a tetrahedral arrangement of hybrid orbitals. Two of these
contain lone pairs of electrons; the other two bond to
hydrogen atoms.
29
A water molecule, therefore, is asymmetrical. The
center of positive charge is offset quite a bit from
the center of negative charge. Because H-O bonds lie
on one side of the oxygen atom, the water molecule
has a strong dipole moment and can exert an orientation effect on nearby charged particles. As a result,
water is an excellent solvent for ionic substances.
In a sodium chloride solution, for example, each
Na+ ion is surrounded by a hydration tetrahedron
of water molecules, each one with its negative end
drawn toward the Na+ ion by a weak ion-dipole
interaction (fig. 2.17). Each Cl− ion attracts the positive end of a water dipole and is also enclosed in a
hydration sphere. Sodium and chloride ions are thus
shielded from each other and prevented from precipitating. We investigate solubility more fully in chapters 4 and 7.
FIG. 2.17. Visualize a hydrated Na+ or Cl− ion at the center of a cube with water molecules at four of
the corners. The negative end of each water dipole is attracted to the Na+ ion, so the hydrogen ions
at the positive end point outward. A Cl− ion attracts the positive end of the water dipole, so that the
hydrogen ions point inward.
30
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Water molecules also attract each other through
dipole-dipole interactions that we refer to as hydrogen bonds. The attraction is greatest in ice, in which
each hydrogen atom is drawn electrostatically to a
lone electron pair on a neighboring water molecule
(fig. 2.18). The result is an open network of molecules that has a dramatically lower density than liquid water. This attraction also accounts for water’s
unusually high surface tension, boiling point, and
melting point, as well as many other distinctive properties we will explore elsewhere in this book.
FIG. 2.18. When water freezes, sp3 orbitals on adjacent H2O
units align so that there is one hydrogen atom along each
oxygen-oxygen axis. The hydrogen bonds (dashed lines)
between H2O units thus make an electrostatic connection
between hydrogen ions on one H2O molecule and paired (nonbonding) electrons on the other molecule. These weak bonds
are nevertheless strong enough to support the open crystal
structure of ice.
structure of benzene, therefore, is a resonant mixture of
all five. Some of the bonding electrons, in other words,
do not belong to specific pairs of carbon atoms but to
the entire ring.
Taken to an extreme, this multiatom example suggests a parallel to the popular description of bonding in
metals. We observed earlier that the distinction between
ionic and covalent bonding is not sharp, but a matter of
degree. Metallic bonding is yet another mode along a
continuum of atomic interactions, a mode in which electrons are even more free to roam from their parent atoms
FIG. 2.19. (a) Bonding electrons in a benzene molecule bind six
carbon atoms in a ring and connect a hydrogen ion to each one.
The remaining bonding electron from each carbon atom occupies
a p orbital perpendicular to the plane of the ring. (b) At any
instant, electrons in the projecting p orbitals (technically, π
orbitals) can form stable pairs (hence a “second” or “double”
bond) in any of five ways. The benzene structure is a resonant
mixture of these equally probable arrangements.
How Elements Behave
than in covalently bonded materials. In a pure metal, all
atoms are the same size, so the coordination number for
any one of them is 12 (see table 2.4). Statistically, therefore, each atom should have 12 equivalent bonds with
its neighbors. None of the metals, however, has 12 unpaired electrons or can generate that many hybrid orbitals. This apparent problem can be resolved if we
consider the bonding orbitals to be an extensive, nondirected resonant network, often described as an electron
“gas” that permeates the structure. As in the benzene
molecule, valence electrons belong to the entire structure,
not to specific atoms. As a result, metals have a higher
electrical and thermal conductivity than nonmetals.
This simplistic “free electron” model for metallic
bonding is limited in ways that would become apparent
if we studied the electrical potential across a metallic
structure carefully. The quantum nature of the structure
makes some energy levels accessible to the free electrons
and prohibits others. Even at this basic level, however,
the concept of nonlocalized bonding offers a useful starting point for understanding the behavior of metals.
31
Yet another variation on this theme of nonlocalized
bonding is Van der Waals bonding, which may also be
described as an extended, nondirectional field of forces
between atoms or molecular units. Geologists may know
of Van der Waals bonding primarily through their familiarity with the structure of graphite, which consists of
stacked sheets of covalently bonded carbon (fig. 2.20).
Overlapping sp2 hybrid orbitals shape the hexagonal
array of atoms within a sheet. Electrons in these orbitals
are tightly bound to pairs of atoms. Electrons in the remaining unhybridized p orbital on each carbon atom
stabilize the sheet by being shared more extensively
within the array. Because all of these interatomic forces
are confined within sheets, each sheet can be thought of
as a separate “molecule” of graphite. The unhybridized
p orbitals, however, are perpendicular to the sheets
and thus contribute to local charge polarization, analogous to the polarity on a water molecule. This polarization provides the weak cohesion—the Van der Waals
attraction—that binds sheets together in crystals of
graphite.
FIG. 2.20. Carbon atoms within a sheet of the graphite structure are bonded by electrons in overlapping sp2 hybrid orbitals and by electrons shared among p orbitals perpendicular to the sheet, as
in benzene (see fig. 2.19). The weak Van der Waals attraction between sheets is due to local charge
polarization on those p orbitals.
32
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
TABLE 2.5. Properties Associated with Interatomic Bonds
Type of Bond
Property
Interatomic
force is . . .
Bonds are
mechanically . . .
In response to
temperature
change, phases
have . . .
In an electrical
field, phases
are . . .
In response to
light, phases . . .
Ionic
Covalent
Metallic
Van der Waals
nondirected
spatially directed
nondirected
nondirected
strong, making
hard crystals
fairly high melting
point, low coefficients
of expansion, ions in
melt
moderate insulators,
conducting by ion
transport in melt
have properties similar
to individual ions,
and therefore also
similar in solution
strong, making
hard crystals
high melting point,
low coefficients
of expansion,
molecules in melt
insulators in solid
and in melt
variable; gliding
is common
variable melting
points
weak, making soft crystals
conductors, by
electron transport
insulators in solid and in
melt
have high refractive
indices; absorption
very different in
solution or gas
are opaque; properties
similar in melt
have properties similar to
individual molecules,
and therefore also
similar in solution
low melting points, large
coefficients of expansion
After Evans (1964).
Retrospective on Bonding
Unlike other treatments of chemical bonding that
you are likely to have seen, we have tried to call attention to similarities among bond types rather than differences. As so often is true in science, the simple categories
you may have learned to recognize in early chemistry or
mineralogy courses have blurred edges. Ionic, covalent,
metallic, and Van der Waals bonds constitute a continuum of forces that hold atoms together. It is particularly
important to emphasize this, because the structures of
the most abundant geologic materials include bonds from
more than one of the standard categories, as well as
some that cannot be placed easily in any of them. As a
result, many geologic materials do not behave the way
you may have learned to expect ideal ionic, covalent, or
metallic substances to behave.
We do not mean to imply that the distinctions among
the bond types are meaningless. As table 2.5 indicates,
there are some fundamental differences between compounds based primarily on each of these types of bonds.
These differences are apparent across broad categories
of Earth materials that we consider in later chapters.
SUMMARY
In this chapter, we have established a vocabulary for the
chemical language we use throughout this book, with-
out straying too far from our peculiar interest in the
chemistry of the Earth. With luck, much of what you
have just read is a review of familiar material. We hope,
though, that we have succeeded in presenting it from a
novel perspective. The structure of atoms, particularly
the configuration of their electrons, controls the ways
that they combine to form natural materials. The type
and strength of chemical bonds in a substance determine
the amount of energy needed to involve it in reactions.
The masses, sizes, and electrical charges on atoms influence the rates at which they react with each other. All of
these are topics we explore in the chapters ahead.
suggested readings
Ahrens, L. H. 1965. Distribution of the Elements in our Planet.
New York: McGraw-Hill. (This concise paperback, now
dated, is still one of the best overviews of the principles affecting element distribution in the Earth.)
Barrett, J. 1970. Introduction to Atomic and Molecular
Structure. New York: Wiley. (A rigorous introduction to
the topic, without the heavy reliance on mathematics
that can make quantum chemistry inaccessible for nonspecialists.)
Faure, G. 1991. Principles and Applications of Inorganic Geochemistry. New York: Macmillan. (A well-written, broad
introduction to geochemistry, with particular attention to
nuclear chemistry.)
McMurray, J. and R. C.Fay. 1995. Chemistry. Englewood Cliffs:
Prentice-Hall. (Every geochemistry student should have a
How Elements Behave
basic comprehensive chemistry text; we consider this one to
be the best.)
Pauling, L. 1960. The Nature of the Chemical Bond, 3rd ed.
Ithaca: Cornell University Press. (This is the classical treatise
on bonding, surprisingly easy to follow.)
Van Melsen, A.G. 1960. From Atomos to Atom. New York:
Harper and Brothers. (For students who are curious about
the history of science, this book is a highly readable account
of the evolution of atomic theory.)
33
The following sources were referenced in this chapter. The
reader may wish to examine them for further details.
Evans, R. C. 1964. An Introduction to Crystal Chemistry.
Cambridge: Cambridge University Press.
Shannon, R. D. 1976. Revised effective ionic radii and systematic studies of interatomic distances in halides and chalcogenides. Acta Crystallographica, Section A 32:751–767.
34
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
PROBLEMS
(2.1)
Magnesium has three naturally occurring isotopes: 24Mg (78.99% abundance, mass = 23.985 amu),
(10% abundance, mass = 24.986 amu), and 26Mg (10% abundance, mass = 25.983amu). What
is the average atomic weight of magnesium?
25Mg
(2.2)
The average atomic weight of silicon is 28.086 amu. It has three stable isotopes, two of which are
(92.23% abundance, mass = 27.977 amu) and 30Si (3.10% abundance, mass = 29.974 amu).
What is the mass of the third stable isotope?
28Si
(2.3)
14C,
important in radiometric dating, has the same atomic mass number as 14N. How is this possible?
(2.4)
Arrange the following bonds in order of their increasing ionic character, using Pauling’s formula:
Be-O, C-O, N-O, O-O, Si-O, Se-O. Repeat, using DEN values. How do the two sets of results
compare?
(2.5)
What electrically neutral atoms have each of following the ground state configurations?
(a) 1s2 2s2 2p6 3s2 3p6 4s2
(b) 1s2 2s2 2p6 3s2 3p6 4s2 3d6
(c) 1s2 2s2 2p6 3s2 3p6 4s2 3d10 4p5
(2.6)
What is the ground state configuration for each of the following ions?
(a) Mg2+
(b) F−
(c) Cu+
(d) Mn3+
(e) S2−
(2.7)
Predict the coordination number for each element in the following compounds or complexes:
(a) CuF2
(b) SO3
(c) CO2
(d) MnO2+
(2.8)
Considering their EN values, explain why potassium, strontium, and aluminum are classified as
lithophile rather than chalcophile elements in Goldschmidt’s classification.
(2.9)
Show algebraically that eightfold coordination for a cation is likely only if the ionic radius ratio
− 1 = 0.732.
r+/r− > √3
(2.10) According to one empirical set of definitions, atoms joined by covalent bonds form molecules; those
joined by ionic or metallic bonds do not. What might be the justification for these definitions? Are
they valid? What are their limitations?
(2.11) Graphite will conduct an electrical current applied parallel to its sheet structure, but will not conduct
a current perpendicular to the sheets. Why?
CHAPTER THREE
A FIRST LOOK AT
THERMODYNAMIC EQUILIBRIUM
OVERVIEW
In this chapter, we introduce the foundations of thermodynamics. A major goal is to establish what is meant by
the concept of equilibrium, described informally in the
context of the party example in chapter 1. The idea of
equilibrium is an outgrowth of our understanding of
three laws of nature, which describe the relationship
between heat and work and identify a sense of direction
for change in natural systems. These laws are the heart
of the subject of thermodynamics, which we apply to a
variety of geochemical problems in later chapters.
To reach our goal, we need to examine some familiar
concepts like temperature, heat, and work more closely.
It is also necessary to introduce new quantities such as
entropy, enthalpy, and chemical potential. The search
for a definition of equilibrium also reveals a set of four
fundamental equations that describe potential changes in
the energy of a system in terms of temperature, pressure,
volume, entropy, and composition.
TEMPERATURE AND
EQUATIONS OF STATE
It is part of our everyday experience to arrange items
on the basis of how hot they are. In a colloquial sense,
our notion of temperature is associated with this ordered
arrangement, so that we speak of items having higher or
lower temperatures than other items, depending on how
“hot” or “cool” they feel to our senses. In the more rigorous terms we must use as geochemists, though, this
casual definition of temperature is inadequate for two
reasons. First of all, a practical temperature scale should
have some mathematical basis, so that changes in temperature can be related to other continuous changes in
system properties. It is hard to use the concept of temperature in a predictive way if we have to rely on disjoint,
subjective observations. Second, and more important, it
is hard to separate this common perception of temperature from the more elusive notion of “heat,” the quantity
that is transferred from one body to another to cause
changes in temperature. It would be helpful to develop a
definition of temperature that does not depend on our
understanding of heat, but instead relies on more familiar
thermodynamic properties.
Imagine two closed systems, each of which is homogeneous; that is, each system consists of a single substance
with continuous physical properties—from now on, we
will call this kind of substance a phase—that is not undergoing any chemical reactions. If this condition is met,
the thermodynamic state of either system can be completely described by defining the values of any two of its
35
36
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
intensive properties. These might be pressure and viscosity, for example, or acoustic velocity and molar volume.
To study the concept of temperature, let’s examine the relationship between pressure and molar volume, using the
symbols p and v̄ in one system, and P and V̄ in the other.
(Refer back to chapter 1 to confirm that molar volume,
the ratio of volume to the number of moles in the system, is indeed an intensive property. Notice, also, that
we are now beginning to use the convention of an overscore to identify molar quantities, as described in chapter 1.)
If we place the two containers in contact, changes of
pressure and molar volume will take place spontaneously
in each system until, if we wait long enough, no further
changes occur. When the two systems reach that point,
they are in thermal equilibrium. Let’s now measure their
pressures and molar volumes and label them P1, V̄1 and
p1, v̄1. There is no reason to expect that P1 will necessarily be the same as p1 or that V̄1 will be equal to v̄1. In
general, the systems will not have identical properties.
If we separate the systems for a moment, we will find
it possible to change one of them so that it is described
by new values P2 and V̄2 that still lead to thermal equilibrium with the system at p1 and v̄1. There are, in fact,
an infinite number of possible combinations of P and V̄
that satisfy this condition. These combinations define a
curve on a graph of P versus V̄ (fig. 3.1a). Each combination of P and V̄ describes a state of the system in thermal equilibrium with p1 and v̄1. All points on the curve
are, therefore, also in equilibrium with each other. This
observation has been called the zeroth law of thermodynamics: “If A is in equilibrium with B and B is in
equilibrium with C, then A is in equilibrium with C.”
Notice—this is important—that figure 3.1a could describe either system. We could just as well have varied
pressure and molar volume in the other system and
found an infinite number of values of p and v̄ that lead
to thermal equilibrium with the system at P1,V̄1. These
values would define a curve on a graph of p versus v̄,
shown in figure 3.1b.
Curves like these are called isothermals. What we
have just demonstrated is that states of a system have the
same temperature if they lie on an isothermal. Furthermore, two systems have the same temperature if their
states on their respective isothermals are in thermal equilibrium. These two statements, together, define what we
mean by “temperature.” You could express these two
statements algebraically by saying that the each of isotherms in our example can be described by a function:
f(p, v̄) = t or F(P, V̄) = T,
and that the systems are in thermal equilibrium if t = T.
Functions like these, which define the interrelationships
among intensive properties of a system, are called equations or functions of state.
Because many real systems behave in roughly similar
ways, several standard forms of the equation of state
have been developed. We find, for example, that many
gases at low pressure can be described adequately by the
ideal gas law:
PV̄ = RT,
in which R is the gas constant, equal to 1.987 cal mol−1
K−1. Other gases, particularly those containing many
atoms per molecule or those at high pressure, are characterized more appropriately by expressions like the Van
der Waals equation:
(P + a/V̄ 2)(V̄ − b) = RT,
FIG. 3.1. (a) All states of a system for which f(P, V) = T are said
to be in thermal equilibrium. A line connecting all such states is
called an isothermal. (b) If another system contains a set of
states lying on the isothermal f(p, v) = t, and if t = T, then the
two systems are in thermal equilibrium.
in which a and b are empirical constants. We apply these
and other equations of state in chapter 4.
Our notion of temperature, therefore, depends on
the conventions we choose to follow in writing equations
of state. A practical temperature scale can be devised
simply by choosing a well-studied reference system (a
thermometer), writing an arbitrary function that remains
constant (that is, generates isothermals) for various states
of the system in thermal equilibrium, and agreeing on a
convenient way to number selected isothermals.
A First Look at Thermodynamic Equilibrium
same at all places in the gas and the gas is not expanding
in mechanical equilibrium with its surroundings.
WO R K
In general terms, work is performed whenever an object
is moved by the application of a force. An infinitesimal
amount of work dw, is therefore described by writing:
dw = F dx,
in which F is a generalized force and dx is an infinitesimal displacement. By convention, we define dw so that
it is positive if work is performed by a system on its
surroundings, and negative if the environment performs
work on the system. We know from experience that an
object may be influenced simultaneously by several
forces, however, so it is more useful to write this equation as:
dw =
37
ΣF dx .
i
i
THE FIRST LAW OF THERMODYNAMICS
During the 1840s, a series of fundamental experiments
were performed in England by the chemist James Joule.
In each of them, a volume of water was placed in an insulated container, and work was performed on it from
the outside. Some of these experiments are illustrated in
figure 3.2. The paddle wheel, iron blocks, and other
mechanical devices are considered to be parts of the
insulated container. The temperature of the water was
monitored during the experiments, and Joule reported
the surprising result that a specific amount of work performed on an insulated system, by any process, always
results in the same change of temperature in the system.
The forces may be hydrostatic pressure, P, or may be directed pressure, surface tension, or electrical or magnetic
potential gradients. All forms of work, though, are equivalent, so the total work performed on or by a system can
be calculated by including all possible force terms in the
equation for dw. The thermodynamic relationships that
build on the equation do not depend on the identity of
the forces involved.
This description of work is broader than it often needs
to be in practice. Work is performed in most geochemical systems when a volume change dV is generated by
application of a hydrostatic pressure:
dw = P dV.
This is the equation we most often encounter, and geochemists commonly speak as if the only work that
matters is pressure-volume work. Usually, the errors
introduced by this simplification are small. It is always
wise, however, to examine each new problem to see
whether work due to other forces is significant. Later
chapters of this book discuss some conditions in which
it is necessary to consider other forces.
Note that we have defined work with a differential
equation. The total amount of work performed in a process is the integral of that equation between the initial
and final states of the system. Because forces in geologic
environments rarely remain constant as a system evolves,
the integral becomes extremely difficult to evaluate if we
do not specify that the process is a slow one. The work
performed by a gas expanding violently, as in a volcanic
eruption, is hard to estimate, because pressure is not the
FIG. 3.2. English chemist James Joule performed experiments to
relate mechanical work to heat. (a) A paddlewheel is rotated and
a measurable amount of work is performed on an enclosed water
bath. (b) Two blocks of iron are rubbed together. (c) Electrical
work is performed through an immersion heater.
38
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
What changed inside the container, and how do we
explain Joule’s results? In physics, work is often introduced in the context of a potential energy function. For
example, a block of lead may be lifted from the floor to
a tabletop by applying an appropriate force to it. When
this happens, we recognize not only that an amount of
work has been expended on the block, but also that its
potential energy has increased. Usually, problems of this
type assume that no change takes place in the internal
state of the block (that is, its temperature, pressure, and
composition remain the same as the block is lifted), so
the change in potential energy is just a measure of the
change in the block’s position. When Joule performed
work on the insulated containers of water, however, it
was not the position of the water that changed, but its
internal state, as measured by a change in its temperature. The energy function used to describe this situation
is called the internal energy (symbolized by E) of the
system, and Joule’s results can be expressed by writing
dE = −dw.
(3.1)
Thus, there are two ways to change the internal energy of a system. First, as we implied when considering
the meaning of temperature, energy can pass directly
through system walls if those walls are noninsulating.
Energy transferred by this first mode is called heat. If
Joule’s containers had been noninsulators, he could have
produced temperature changes without performing any
work, simply by lighting a fire under them. Second, as
Joule demonstrated, the system and its surroundings can
perform work on each other.
To investigate these two modes of energy transfer,
we need to distinguish between two types of walls that
can surround a closed system. What we have spoken of
loosely as an “insulated” wall is more properly called
an adiabatic wall. (The Greek roots of the word adiabatic mean, appropriately, “not able to go through.”)
Adiabatic processes, like those in Joule’s experiments,
do not involve any transfer of heat between a system and
its surroundings. Perfect adiabatic processes are seldom
seen in the real world, but geochemists often simplify
natural environments by assuming that they are adiabatic. A nonadiabatic wall, however, allows the passage
of heat. It is possible to perform work on a system that
is bounded by either type of wall.
Equation 3.1 is an extremely useful statement, referred to as the First Law of Thermodynamics. In words,
it tells us that the work done on a system by an adiabatic
process is equal to the increase in its internal energy, a
function of the state of the system. We may conclude further that if a system is isolated, rather than simply closed
behind an adiabatic wall, no work can be performed on
it from the outside, and its internal energy must remain
constant.
Equation 3.1 can be expanded to say that for any
system, a change in internal energy is equal to the sum
of the heat gained (dq) and the work performed on the
system (−dw):
dE = dq − dw.
(3.2)
Notice that dE is a small addition to the amount of internal energy already in the system. Because this is so,
the total change in internal energy during a geochemical
process is equal to the sum of all increments dE:
dE = Efinal − Einitial = ∆E.
The value of ∆E, in other words, does not depend on
how the system evolves between its end states. In fact, if
the system were to evolve along a path that eventually
returns it to its initial state, there would be no net change
whatever in its internal energy. This is an important
property of functions of state, described in more mathematical detail in appendix A.
In contrast, the values of dq and dw describe the
amount of heat or work expended across system boundaries, rather than increments in the amount of heat or
work “already in the system.” When we talk about heat
or work, in other words, the emphasis is on the process
of energy transfer, not the state of the system. For this
reason, the integral of dq or dw depends on which path
the system follows from its initial to its final state. Heat
and work, therefore, are not functions of state.
Worked Problem 3.1
Consider the system illustrated in figure 3.3, whose initial state
is described by a pressure P1 and a temperature T1. This might
be, for example, a portion of the atmosphere. Recall that the
equation of state that we use to define isothermals for this system tells us what the molar volume, V̄1, under these conditions
is. For ease of calculation, assume that the equation of state for
this system is the ideal gas law, V̄ = RT/P. Compare two ways
in which the system might slowly evolve to a new state in which
P = P2 and T = T2. In the first process, let pressure increase
slowly from P1 to P2 while the temperature remains constant, then
A First Look at Thermodynamic Equilibrium
39
ENTROPY AND THE SECOND LAW
OF THERMODYNAMICS
FIG. 3.3. As a system’s pressure and temperature are adjusted
from P1T1 to P2T2, the work performed depends on the path that the
system follows.
let temperature increase from T1 to T2 at constant pressure P2.
In the second process, let temperature increase first at constant
pressure P1, and then let pressure increase from P1 to P2. (If this
were, in fact, an atmospheric problem, the isobaric segments of
these two paths might correspond to rapid surface warming on
a sunny day, and the isothermal segments might reflect the passage of a frontal system.) Is the amount of work done on these
two paths the same?
This problem is similar to problem 1.8 at the end of chapter 1.
If the only work performed on the system is due to pressurevolume changes, then the work along each of the paths is defined by the integral of PdV̄. For path 1,
w1 =
P2T1
养
P1T1
PdV̄ +
养
P2T2
PdV̄ +
养
P2T2
P2T1
PdV̄.
Along path 2,
w2 =
P1T2
养
P1T1
P1T2
PdV̄.
To evaluate these four integrals, find an expression for dV̄ by
expressing the equation of state as a total differential of V̄(T, P):
dV̄ = (∂V̄/∂T)P dT + (∂V̄/∂P)T dP = (R/P)dT − (RT/P2)dP,
and substitute the result into w1 and w2 above. The result, after
integration, is that the work performed on path 1 is:
w1 = R[(T2 − T1) + T1 ln(P1 /P2)]
but the work performed on path 2 is:
w2 = R[(T2 − T1) + T2 ln(P1 /P2)],
which is clearly different. Similarly, if we had a function for dq,
the heat gained, we could integrate it to show that q1 and q2, the
amounts of heat expended along these paths, must also be different. We are about to do just that.
Taken by itself, equation 3.2 tells us that heat and work
are equivalent means for changing the internal energy of
a system. In developing that expression, however, we
have engaged in a little sleight-of-hand that may have led
you, quite incorrectly, to another conclusion as well. By
stating that heat and work are equivalent modes of energy transfer, the First Law may have left you with the
impression that heat and work can be freely exchanged
for one another. This is not the case, as a few commonplace examples will show.
A glass of ice water left on the kitchen counter gradually gains heat from its surroundings, so that the water
warms up and the room temperature drops ever so
slightly. In the process, there has been an exchange of
heat from the warm room to the relatively cool water.
As Joule showed, the same transfer could have been accomplished by having the room perform work on the ice
water. It is clearly impossible, however, for a glass of
water at room temperature to cool spontaneously and
begin to freeze, although we can certainly remove heat
by transferring it first from the water to a refrigeration
system. In a similar way, a lava flow gradually solidifies
by transferring heat to the atmosphere, but this natural
process cannot reverse itself either. It is impossible to melt
rock by transferring heat to it directly from cold air.
If we were shown films of either of these events, we
would have no trouble recognizing whether they were
being run forwards or backwards. Equation 3.2, however, does not provide us with a means of making this
determination of direction theoretically. On the basis
of experience with natural processes, therefore, we are
driven to formulate a Second Law of Thermodynamics.
In its simplest form, first stated by Rudolf Clausius in the
middle of the nineteenth century, the Second Law says
that heat cannot spontaneously pass from a cool body to
a hotter one. Another way of stating this, which may also
be useful, is that any natural process involving a transfer
of energy is inefficient, with the result that a certain
amount is irreversibly converted into heat that cannot be
involved in further exchanges. The Second Law, therefore, is a recognition that natural processes have a sense
of direction.
Two difficult concepts are embedded in what we have
just said. One is spontaneity and the other is reversibility. A spontaneous change is one that, under the right
40
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
if they were reversible. The conceptual challenge arises
because true reversibility is beyond our experience with
the natural world.
Another perspective on the Second Law, then, is that
it tells us that nature favors spontaneous change, and
that maximum work is never performed in real-world
processes. Just as the internal energy function was introduced to present the First Law in a quantitative fashion,
we need to define a thermodynamic function that quantifies the sense of direction and systematic inefficiency that
the Second Law identifies in natural processes. This new
function, known as entropy and given the symbol S, is
defined in differential form by:
dS = (dq/T)rev,
FIG. 3.4. Two weights are connected by a rope that passes over
a frictionless pulley. If one weight is infinitesimally heavier than
the other, then it will perform the maximum possible amount of
work (the work of lifting the other weight) as it falls. The smaller
the weight difference between them is, the greater the amount of
work and the more nearly reversible it is.
conditions, can be made to perform work. The heavy
weight in figure 3.4 can perform the work of raising the
lighter weight as it falls, so that this change is spontaneous. If the lighter weight raised the heavier one, we
would recognize the change as nonspontaneous and we
would assume that some change outside the system had
probably caused it. When we face a conceptual challenge
in telling whether a change is spontaneous, it is usually
because we haven’t defined the bounds of the system
clearly. Notice also that a spontaneous change doesn’t
have to perform work; it just has to be capable of it. The
heavy weight will fall spontaneously whether it lifts the
lighter weight or not.
This leads us to consider reversibility. A change is
reversible if it does the maximum amount of possible
work. The falling weight in figure 3.4, for example, will
be undergoing a reversible change if it lifts an identical
weight and loses no energy to friction in the pulley and
rope that connects it to the other weight. Experience tells
us that reversibility is an unattainable ideal, of course,
but that doesn’t stop engineers from designing counterweighted elevators to use as little extra energy as possible.
It also doesn’t stop geochemists from considering gradual
changes in systems that are never far from equilibrium as
(3.3)
in which dq is an infinitesimal amount of heat gained by
a body at temperature T in a reversible process. It is a
measure of the degree to which a system has lost heat
and therefore some of its capacity to do work. Many
people find entropy an elusive concept to master, so it is
best to become familiar with the idea by discussing examples of its use and properties.
Figure 3.5 illustrates a potentially reversible path for
a system enclosed in a nonadiabatic wall. Suppose that
the system expands isothermally from state A to state
B. In doing so, it performs an amount of work on its
surroundings that can be calculated by the integral of
PdV. Graphically, this integral is represented by the area
AA′B′B. Because this is an isothermal process, state A
and state B must be in equilibrium with each other,
FIG. 3.5. The work performed by isothermal compression from
state A to state B is equal to the area AA′B′B. Because dE = 0, an
equivalent negative amount of heat is gained by the system. If
this change could be reversed precisely, the net change in entropy
dS would be zero.
A First Look at Thermodynamic Equilibrium
41
Remember, though, that the Second Law tells us that heat
can only pass spontaneously from a hot body to a cooler
one, so this result is only valid if T2 > T1. Therefore, the
entropy of an isolated system can only increase as it
approaches internal equilibrium:
dSsys > 0.
FIG. 3.6. Two bodies in an isolated system are separated by a
nonadiabatic wall and may, therefore, exchange quantities of heat
dq1 and dq2 as they approach thermal equilibrium.
which is another way of saying that the internal energy
of the system must remain constant along the path between them. Therefore, the work performed by the system must be balanced by an equivalent amount of heat
gained, according to equation 3.2. Suppose, now, that it
were possible to compress the system isothermally, thus
reversing along the path B → A without any interference
due to friction or other real-world forces. We would find
that the work performed on the system (the heat lost by
the system) would be numerically equal to the work on
the forward path, but with a negative sign. That is, the
net change in entropy around the closed path is:
dS = dq/T − dq/T = 0.
Finally, look at the closed system illustrated in figure 3.7. It is similar to the previous system in all respects,
except that is bounded by a nonadiabatic wall, so that
the two bodies can exchange amounts of heat dq1′ and
dq2′ with the world outside. As they approach thermal
equilibrium, the heat exchanged by each is given by:
(dq1)total = dq1 + dq1′,
and
(dq2)total = dq2 + dq2′.
Any heat exchanged internally must still show up either
in one body or the other, so as before:
dq1 = −dq2.
The net change in system entropy is, therefore,
dSsys = (dq1)total + (dq2)total
= (dq1′/T1) + (dq2′/T2) + dq1([1/T1] − [1/T2]).
It has already been shown that the Second Law requires
the last term of this expression to be positive. Therefore,
dSsys > (dq1′/T1) + (dq2′/T2).
The entropy function for a system following a reversible
pathway, therefore, is a function of state, just like the
internal energy function.
Next, consider an isolated system in which there are
two bodies separated by a nonadiabatic wall (fig. 3.6).
Let the two bodies initially be at different temperatures
T1 and T2. If we allow them to approach thermal equilibrium, there must be a transfer of heat dq2 from the
body at temperature T2. Because the total system is isolated, an equal amount of heat dq1 must be gained by the
body at temperature T1. The net change in entropy for
the system is:
dSsys = dS1 + dS2,
or
dSsys = dq1 ([1/T1] − [1/T2]).
FIG. 3.7. The system in figure 3.6 is allowed to exchange heat
with its surroundings. The net change in internal entropy depends
on the values of dq1′ and dq2′.
42
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
This result is more open ended than the previous one.
Although the entropy change due to internal heat exchanges in any system will always be positive, it says
that we have no way of predicting whether the overall
dSsys of a nonisolated system will be positive unless we
also have information about dq1′, dq2′, and the temperature in the world outside the system.
To appreciate this ambiguity, think back to the difference between placing a glass of ice water on the kitchen
counter and placing it in the freezer. In both cases, internal heat exchange results in an increase in system entropy. When the glass is refrigerated, however, (dq1′/T1)
and (dq2′/T2) become potentially large negative quantities,
with the result that the change due to internal processes
is overwhelmed and the entropy within the glass decreases. The Second Law, therefore, does not rule out
the possibility that thermodynamic processes can reverse
direction. It does tell us, though, that a spontaneous
change can only reverse direction if heat is lost to some
external system. A local decrease in entropy must be accompanied by an even larger increase in entropy in the
world at large. When we have difficulty with this notion,
as we commented earlier, it is usually because we have
not been clear about defining the size and boundaries of
the system. Therefore, except for those cases in which a
system evolves sufficiently slowly that we are justified in
approximating its path as a reversible one, its change in
entropy must be defined not by equation 3.3, but by:
dS > (dq/T )irrev.
(3.4)
Entropy and Disorder
Entropy is commonly described in elementary texts as
a measure of the “disorder” in a system. For those who
learned about entropy through statistical mechanics, following the approach pioneered by Nobel physicist Max
Planck in the early 1900s, this may make sense. This is
sometimes less than satisfying for people who address
thermodynamics as we have, because the mathematical
definition in equations 3.3 and 3.4 doesn’t seem to address
entropy in those terms. To see why it is valid to think of entropy as an expression of disorder, consider this problem.
Worked Problem 3.2
Two adjacent containers are separated by a removable wall.
Into one of them, we introduce 1 mole of nitrogen gas. We fill
the other with 1 mole of argon. If the wall is removed, we expect that the two gases will mix spontaneously rather than remaining in their separate ends of the system. It would require
considerable work to separate the nitrogen and argon again.
(See problem 3.5 at the end of this chapter.) By randomizing the
positions of gas molecules and contributing to a loss of order
in the system, we have caused an increase in entropy.
Suppose that the mixing takes place isothermally, and that
the only work involved is mechanical. Assume also, for the sake
of simplicity, that both argon and nitrogen are ideal gases. How
much does the entropy increase in the mixing process? First,
write equation 3.2 as:
dq = dE + PdV.
Because we agreed to carry out the experiment at constant
temperature, dE must be equal to zero. Using the equation of
state for an ideal gas, we can see that dq now becomes:
dq = (nRT/V)dV,
from which:
dS = dq/T = (nR/V)dV.
(Notice that this problem is cast in terms of volume, rather than
molar volume, because we need to keep track of the number of
moles, n.) We can think of the current problem as one in which
each of the two gases has been allowed to expand from its initial volume (V1 or V2 ) into the larger volume V1 + V2 . For each,
then, the entropy change due to isothermal expansion is:
∆Si = ∆ni R ln([V1 + V2 ]/Vi ),
and the combined entropy change is:
∆S1 + ∆S2 = n1R ln([V1 + V2 ]/V1) + n2 R ln([V1 + V2 ]/V2 ).
This can be simplified by defining the mole fraction, Xi, of either
gas as Xi = ni /(n1 + n2) = Vi /(V1 + V2 ); thus,
∆S̄ = −R(X1 ln X1 + X2 ln X2),
where the bar over ∆S̄ is our standard indication that it has
been normalized by n1 + n2 and is now a molar quantity. The
entropy of mixing will always be a positive quantity, because X1
and X2 are always <1.0. The answer to our question, therefore,
is that entropy increases by:
2 × ∆S̄ = 2 × −(1.987)(0.5 ln 0.5 + 0.5 ln 0.5)
= 2.76 cal K−1 or 11.55 J K−1.
REPRISE: THE INTERNAL ENERGY
FUNCTION MADE USEFUL
With the First and Second laws in hand, we can now
return to the subject of equilibrium, which was defined
informally at the beginning of this chapter as a state in
A First Look at Thermodynamic Equilibrium
43
ENTROPY AND DISORDER: WORDS OF CAUTION
The description of entropy as a measure of disorder
is probably the most often used and abused concept
in thermodynamics. In many applications, the idea
of “disorder” is coupled with the inference from the
Second Law that entropy should increase through
time. Like all brief descriptions of complex ideas,
though, the statement that “entropy measures system
disorder” must be treated with some care. It is common practice among chemical physicists to estimate
the entropy of a system by calculating its statistical
degrees of freedom—computing, in effect, the number of different ways that atoms can be configured in
the system, given its bulk conditions of temperature,
pressure, and other intensive parameters. A less rigorous description of disorder, however, can lead to
very misleading conclusions about the entropy of a
system and thus the course of its evolution.
Creationists, for example, have grasped at the idea
of entropy as disorder as a vindication of their belief
that biologic evolution is impossible. In their view,
“higher” organisms are more ordered than the more
primitive organisms from which biologists presume
that they evolved. Because this interpretation presents
evolution as a historical progression from a less ordered to a more ordered state, it describes an apparent
violation of the Second Law. There are so many flaws
in this argument that it is difficult to know where to
begin to discuss it, and we will mention only a few.
The first is the question of the meaning of “order”
as applied to organisms or families of organisms. This
is far from a semantic problem, because it is by no
means clear that higher organisms are more ordered
than primitive ones, from the point of view of thermo-
which no change is taking place. The First Law restates
this condition as one in which dE is equal to zero. From
the Second Law, we find that entropy is always maximized during the approach to equilibrium, and that dS
becomes equal to zero once equilibrium is reached.
The results expressed in equations 3.3 and 3.4 can be
used to recast equation 3.2 as:
dE ≤ TdS − dw.
dynamics. The biologic concept of order, although
potentially connected to thermodynamics in the realm
of biochemistry, is based largely on functional complexity, not equations for energy utilization. The situation becomes even murkier when applied to groups
of organisms, rather than to individuals.
Furthermore, from a thermodynamic perspective,
living organisms cannot be viewed as isolated systems,
separated from their inorganic surroundings. As we
have tried to emphasize, the definition of system
boundaries is crucial if we are to interpret system
changes by the Second Law. A refrigerator, for example, might be mistakenly seen to violate the Second Law if we failed to recognize that entropy in the
kitchen around it increases even as the contents of
the refrigerator become more ordered.
Finally, we note that there may be instances in
which entropy and disorder in the macroscopic sense
can be decoupled, even in a properly defined isolated
system. For example, if two liquids are mixed in an
adiabatic container, we might expect that they will
mix to a random state at equilibrium. We know,
however, that other system properties may commonly
preclude a random mixture. Oil and vinegar in a salad
dressing unmix readily, as a manifestation of differences in their bonding properties. If the attractive force
between similar ions or molecules is greater than that
between dissimilar ones, then complete random mixing would imply an increase, rather than a decrease, in
both internal energy and entropy. A macroscopically
ordered system, in other words, can easily be more
stable than a disordered one without shaking our
faith in the Second Law of Thermodynamics.
Under most geologic conditions, mechanical (pressurevolume) work is the only significant contribution to dw,
and it is reasonable to substitute PdV for dw:
dE ≤ TdS − PdV.
(3.5)
This is a practical form of equation 3.2, although only
correct under the conditions just discussed. In most geologic environments, we assume that change takes place
44
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
very slowly, so that systems can be regarded as following
nearly reversible paths. Generally, therefore, the inequality in equation 3.5 can be disregarded, even though it is
strictly necessary.
Because much of the remaining discussion of thermodynamics in this book derives from equation 3.5, we
should recognize three other relationships that follow
directly from it and from the fact that the internal energy
function is a function of state. First, equation 3.5 is a differential form of E = E(S, V). Assuming that changes take
place reversibly, it can be written as a total differential:
dE(S, V) = (∂E/∂S)V dS + (∂E/∂V)S dV.
Comparison with equation 3.5 yields the two statements:
T = (∂E/∂S)V ,
and
P = −(∂E/∂V)S .
In other words, the familiar variables temperature
and pressure can be seen as expressions of the manner
in which the internal energy of a system responds to
changes in entropy (under constant volume conditions)
or volume (under adiabatic conditions). Second, because
E(S, V) is a function of state, dE(S, V) is a perfect differential (perfect differentials are reviewed in appendix A).
That is,
(∂2E/∂S∂V) = (∂2E/∂V∂S),
or
Enthalpy
The first of these equations can be derived by writing
equation 3.5 as:
dE + PdV ≤ TdS.
If we restrict ourselves by looking only at processes
that take place at constant pressure, this is the same as
writing:
d(E + PV ) ≤ TdS,
because d(PV) = PdV + VdP, which equals PdV if dP = 0.
The quantity E + PV is a new function, called enthalpy and commonly given the symbol H. It is useful as
a measure of heat exchanged under isobaric conditions,
where dH = dq. Reactions that evolve heat, and therefore
have a negative change in enthalpy, are exothermic. Those
that result in an increase in enthalpy are endothermic.
The left side of this last equation can be expanded to
reveal a differential form for dH:
dH = d(E + PV ) = dE + PdV + VdP,
or
dH ≤ TdS + VdP.
(3.7)
As we did with equation 3.5, we can compare the total
differential dH(S, P) and equation 3.7 under reversible
conditions to recognize that:
T = (∂H/∂S)P,
(∂T/∂V)S = −(∂P/∂S)V.
(3.6)
This last equation is known as one of the Maxwell relationships. These and other similar expressions to be
developed shortly can be used to investigate the interactions of S, P, V, and T. Some of these will be discussed
more fully in chapter 9, when we look in more detail at
the effects of changing temperature or pressure in the
geologic environment.
AUXILIARY FUNCTIONS OF STATE
Equation 3.5 is adequate for solving most thermodynamic problems. In many situations, however, it is possible to use equations of greater practical interest, which
can be derived by imposing environmental constraints
on the problem.
and
V = (∂H/∂P)S.
Enthalpy, like internal energy and entropy, can be
shown to be a function of state, because it experiences
no net change during a reversible cycle of reaction paths.
This makes it possible, among other things, to determine
the amount of heat that would be exchanged in geologically important reactions, even when those reactions may
be too sluggish to be studied directly at the low temperatures at which they take place in nature. The careful determination of such values is the business of calorimetry.
Because H is a function of state, it is also possible to
extract one more Maxwell relation like equation 3.6 from
its cross-partial derivatives (also reviewed in appendix A):
(∂T/∂P)S = (∂V/∂S)P.
(3.8)
A First Look at Thermodynamic Equilibrium
The Helmholtz Function
45
The total differential dF(T, V), when compared with
equation 3.9, yields the useful expressions:
By analogy with the way we introduced enthalpy,
we can discover another useful function by writing equation 3.5 as:
S = −(∂F/∂T )V,
and
P = −(∂F/∂V)T .
dE − TdS ≤ −PdV.
Under isothermal conditions, this expression is equivalent to:
Because F is a function of state and dF(T, V ) is therefore a perfect differential, we also gain another Maxwell
relation:
d(E − TS) ≤ −PdV.
The function F = E − TS is best referred to as the Helmholtz function, although you may see it referred to elsewhere as Helmholtz free energy or the work function.
Unfortunately, a variety of symbols have been used
for the internal energy, enthalpy, and Helmholtz functions, as well as Gibbs free energy, which is discussed
shortly. This has led to some confusion in the literature.
In the United States, F is commonly used to designate
Gibbs free energy (for example, in publications of the
National Bureau of Standards). The International Union
of Pure and Applied Chemistry, however, has recommended that G be the standard symbol for Gibbs free
energy. This usage, if not yet standard, is at least widespread. In the United States, those who use F for Gibbs
free energy generally use the symbol A for the Helmholtz
function. To complete the confusion, internal energy,
which we have identified with the symbol E, is frequently referred to as U, to avoid confusing it with total
(internal plus potential plus kinetic) energy. Always be
sure you know which symbols you are using. Nicolas
Vanserg has written an excellent article on the subject,
which is listed among the references at the end of this
chapter (Vanserg 1958).
Because it might be mistaken for Gibbs free energy,
which is ultimately a more useful function in geochemistry, it is best to avoid the term Helmholtz free energy.
The name work function, however, is fairly informative.
The integral of dF at constant temperature is equal to
the work performed on a system. This expression has its
greatest application in mechanical engineering.
It is easy to derive the differential form dF:
(∂S/∂V)T = (∂P/∂T)V .
(3.10)
Although the Helmholtz function is rarely used in
geochemistry, the Maxwell relationship equation 3.10 is
quite widely used. When we return for a second look at
thermodynamics in chapter 9, equation 3.10 will be
discussed as the basis of the Clapeyron equation, a means
for describing pressure-temperature relationships in
geochemistry.
Gibbs Free Energy
The most frequently used thermodynamic quantity
in geochemistry can be derived by writing equation 3.5 in
the form:
dE − TdS + PdV < 0.
Under conditions in which both temperature and pressure
are held constant, this expression becomes:
d(E − TS + PV) < 0.
In the now familiar fashion in which we have already
defined H and F, we designate the quantity E − TS + PV
with the symbol G and call it the Gibbs free energy, in
honor of the Josiah Willard Gibbs, a chemistry professor
at Yale University, who wrote a classic series of papers in
the 1870s in which virtually all of the fundamental equations of modern thermodynamics appeared for the first
time. Colloquially, among geochemists, it is common to
speak simply of “free energy.”
The differential dG can be written:
dG = d(E − TS + PV) = dE − TdS − SdT
+ PdV + VdP,
dF = d(E − TS) = dE − TdS −SdT,
or
or
dF ≤ −SdT − PdV.
(3.9)
dG ≤ −SdT + VdP.
(3.11)
46
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
As with the previous fundamental equations, we can
apply our knowledge of the total differential dG(T, P) of
this state function to find that:
−S = (∂G/∂T )P , V = (∂G/∂P)T ,
and
−(∂S/∂P)T = (∂V/∂T )P .
(3.12)
It can be seen from equation 3.11 that the Gibbs free
energy is the first of our fundamental equations written
solely in terms of the differentials of intensive parameters. This and the ease with which both temperature and
pressure are usually measured contributes to the great
practical utility of this function.
However, what is “free” about the Gibbs free energy? Consider the intermediate step in equation 3.11.
At constant temperature and pressure, this reduces to:
dG = dE − TdS + PdV.
Substituting dE = dq − dw, we have:
dG = dq − dw − TdS + PdV.
If the quantity of heat dq is transferred to the system
isothermally and any changes are reversible, then dq = TdS
and dw = dwrev , so we can rewrite this last expression as:
−dG = dwrev − PdV.
The decrease in free energy of a system undergoing a
reversible change at constant temperature and pressure,
therefore, is equal to the nonmechanical (i.e., not pressurevolume) work that can be done by the system. If the
condition of reversibility is relaxed, then:
−dG > dwirrev − PdV.
In either case, the change in G during a process is a measure of the portion of the system’s internal energy that is
“free” to perform nonmechanical work.
The free energy function provides a valuable practical
criterion for equilibrium. At constant temperature, the
net change in free energy associated with a change from
state 1 to state 2 of a system can be calculated by integrating dG:
∫ dG = ∫ d(E − TS + PV ) = ∫ d(H − TS),
2
2
2
1
1
1
from which:
∆G = ∆H − T∆S,
(3.13)
where ∆G = G2 − G1, ∆H = H2 − H1, and ∆S = S2 − S1.
Energy is available to produce a spontaneous change
in a system as long as ∆G is negative. According to equation 3.13, this can be accomplished under any circumstances in which ∆H − T∆S is negative. Most exothermic
changes, therefore, are spontaneous. Endothermic processes (those in which ∆H is positive) can also be spontaneous, but only if they are associated with a large
positive change in entropy. This possibility was not appreciated at first by chemists. In fact, in 1879 the French
thermodynamicist Marcelin Berthelot used the term
affinity, defined by A = −∆H, as a measure of the direction of positive change. According to his reasoning, a
spontaneous chemical reaction could only occur if A > 0.
If an endothermic reaction turned out to be spontaneous,
he assumed that some unobserved mechanical work
must have been done on the system. As chemists became familiar with Gibbs’s papers on thermodynamics,
however, it became clear that Berthelot and others had
misunderstood the concept of entropy. Affinity was a
theoretical blind alley.
In summary, a thermodynamic change proceeds as
long as there can be a further decrease in free energy.
Once free energy has been minimized (that is, when dG
= 0), the system has attained equilibrium.
Worked Problem 3.3
To illustrate how enthalpy, entropy, and temperature are related to G, consider what happens when ice or snow sublimates
at constant temperature. Those who live in northern climates
will recognize this as a common midwinter phenomenon. Figure 3.8 is a schematic representation of the way in which various energy functions change as functions of the proportions of
water vapor and ice in a closed system. (Strictly speaking, snow
does not sublimate into a closed system, but into the open atmosphere. On a still night, close to the ground, however, this
is not a bad approximation.) Notice that the enthalpy of the
system increases linearly as vaporization takes place. This can
be rationalized by noting we have to add heat to the system to
break molecular bonds in the ice. If enthalpy were the only factor involved in the process, therefore, the system would be most
stable when it is 100% solid, because that is where H is minimized. Entropy, however, is maximized if the system is 100%
water vapor, because the degree of system randomness is greatest there. It can be shown from statistical arguments, however,
that entropy is a logarithmic function of the proportion of vapor,
rising most rapidly when the vapor fraction in the system is low.
Therefore the free energy of the system, G = H − TS, is less than
the free energy of the pure solid until a substantial amount of
A First Look at Thermodynamic Equilibrium
FIG. 3.8. The quantity G = H − TS varies with the amount of
vapor as sublimation takes place in an enclosed container. The
free energy when vapor and solid are in equilibrium is Gmin.
(Modified from Denbigh 1968.)
vapor has been produced. Sublimation occurs spontaneously.
At some intermediate vapor fraction, H − TS reaches a minimum, and solid and vapor are in equilibrium at Gmin. If the
vapor fraction is increased further, the difference H − TS exceeds Gmin again, and condensation occurs spontaneously. This
is the source of the beautiful hoar frost that develops on tree
branches and other exposed surfaces on still winter mornings.
CLEANING UP THE ACT: CONVENTIONS
FOR E, H, F, G, AND S
Except in the abstract sense that we have just used G in
worked problem 3.3, there is no way we can talk about
absolute amounts of energy in a system. To say that a
beaker of reagents contains 100 kJ of enthalpy or 55 kcal
of free energy is meaningless. Instead, we compare the
value of each of the energy functions to its value in a
mixture of pure elements under specified temperature
and pressure conditions (a standard state). For example,
we would measure the enthalpy of a quantity of NaCl at
298 K relative to the enthalpies of pure sodium and pure
chlorine at the same temperature and at one atmosphere
of pressure, using the symbol ∆Hf0. The superscript 0 indicates that this is a standard state value and the subscript f indicates that the reference standard is a mixture
of pure elements. By convention, the ∆Hf0 or ∆Gf0 for a
pure element at any temperature is equal to zero. The
same notation is used for internal energy and the Helmholtz function as well, although they are less frequently
encountered in geochemistry.
47
In our discussions so far, we have perhaps given the
impression that whatever thermodynamic data we might
need to solve a particular problem will be readily available. The references in appendix B do, in fact, include
data for a large number of geologic materials. We have
largely ignored the thorny problem of where they come
from, however, and where we turn for data on materials
that have not yet appeared in the tables. As an example
of how thermodynamic data are acquired, we now show
how solution calorimetry is used to measure enthalpy of
formation. In chapter 10, we show how to extract thermodynamic data from phase diagrams.
It is not necessary (or even possible) to make a direct
determination of ∆Hf0 for most individual phases, because we can rarely generate the phases from their constituent elements. Instead, we measure the heat evolved
or absorbed as the substance we are interested in is produced by a specific reaction between other phases for
which we already have enthalpy data. Because most reactions of geologic interest are abysmally slow at low
temperatures, even this enthalpy change may be almost
impossible to measure directly. However, because enthalpy is an extensive variable, we can employ some
sleight-of-hand: we can measure the heat lost or gained
when the products and reactants are each dissolved in separate experiments in some solvent at low temperature,
and then add the heats of solution for the various products and reactants to obtain an equivalent value for the
heat of reaction.
Worked Problem 3.4
Consider the following laboratory exercise. We wish to determine the molar enthalpy for the reaction:
→
2MgO + SiO2 ←
periclase quartz
Mg2SiO4,
forsterite
to which we assign the value ∆H̄1. This can be determined by
measuring the heats of solution for periclase, quartz, and forsterite in HF at some modest temperature:
→ (2MgO, SiO )
2MgO + SiO2 + HF ←
2 solution ,
which gives ∆H̄2, and:
→ (2MgO, SiO )
Mg 2SiO4 + HF ←
2 solution ,
which gives ∆H̄3. If the solutions are identical, we can see that
the first of these equations is mathematically equivalent to the
second minus the third, so that:
∆H̄1 = ∆H̄2 − ∆H̄3.
48
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
In this way, solution calorimetry can be used to determine the
enthalpy of formation for forsterite from its constituent oxides
at low temperature. As you might expect, however, solvents
suitable for silicate minerals (such as hydrofluoric acid or molten
lead borate) are generally corrosive and hazardous to handle.
Consequently, these experiments require considerable skill and
specialized equipment. Most geochemists refer to published
collections of calorimetric data rather than making the measurements themselves.
What we have just described is not a measurement of ∆H̄f0,
because the reference comparison was to constituent oxides,
not pure elements. If we wanted to determine ∆H̄f0, we would
search for tabulated data for reactions forming the oxides from
pure elements (typically measured by some method other than
solution calorimetry). Again, recalling that enthalpies are additive, we could recognize that ∆H̄f0 for MgO is defined by the
reaction:
→ 2MgO,
2Mg + O2 ←
and ∆H̄f0 for SiO2 is defined by:
→ SiO .
Si + O2 ←
2
Label these two values ∆H̄4 and ∆H̄5. Therefore, for the enthalpy ∆H̄6 of the net reaction:
→ Mg SiO ,
2Mg + 2 O2 + Si ←
2
4
we obtain:
∆H̄6 = ∆H̄2 − ∆H̄3 + ∆H̄4 + ∆H̄5.
The value ∆H̄6 in this case is the enthalpy of formation for forsterite derived from those for the elements, ∆H̄f0.
Although most thermodynamic functions are best
defined as relative quantities like ∆Hf0, entropy is a major
exception. The convention observed most commonly is
an outgrowth of the Third Law of Thermodynamics,
which can be stated, in paraphrase from Lewis and Randall (1961): If the entropy of each element in a perfect
crystalline state is defined as zero at the absolute zero of
temperature, then every substance has a nonnegative
entropy; at absolute zero, the entropy of all perfect crystalline substances becomes zero. This statement, which
has been tested in a very large number of experiments,
provides the rationale for choosing the absolute temperature scale (i.e., temperature in Kelvins) as our standard
for scientific use. For now, the significance of the Third
Law is that it defines a state in which the absolute or
third law entropy is zero. Because this state is the same
for all materials, it makes sense to speak of S0, rather
than ∆Sf0.
With this new perspective, return for a moment to
equation 3.13. It should now be clear that the symbol ∆
has a different meaning in this context. In fact, although
it is rarely done, it would be less confusing to write equation 3.13 as:
∆(∆G) = ∆(∆H) − T∆S,
in which the deltas outside the parentheses and on S refer
to a change in state (that is, a change in these values during some reaction), and the deltas inside the parentheses
refer to values of G or H relative to some reference state.
We examine this concept more fully in later chapters.
Although we have not felt it necessary to prove, it
should be apparent that each of the functions E, H, F, G,
and S is an extensive property of a system. The amount of
energy or entropy in a system, therefore, depends on the
size of the system. In most cases, this is an unfortunate
restriction, because we either don’t know or don’t care
how large a natural system may be. For this reason, it is
customary to normalize each of the functions by dividing them by the total number of moles of material in the
system, thus making each of them an intensive property.
COMPOSITION AS A VARIABLE
Up to now, the functions we have considered all assume
that a system is chemically homogeneous. Most systems
of geochemical interest, however, consist of more than
one phase. The problem most commonly faced by geochemists is that the bulk composition of a system can be
packaged in a very large number of ways, so that it is
generally impossible to tell by inspection whether the assemblage of phases actually found in a system is the most
likely one. To answer questions dealing with the stability of multiphase systems, we need to write a separate set
of equations for E, H, F, and G for each phase and apply
criteria for solving them simultaneously. We will do this
job in two steps.
Components
To describe the possible variations in the compositions and proportions of phases, it is necessary to define
a set of thermodynamic components that satisfy the following rules:
1. The set of components must be sufficient to describe
all of the compositional variations allowable in the
system.
A First Look at Thermodynamic Equilibrium
49
2. Each of the components must vary independently in
the system.
As long as these criteria are met, the specific set of
components chosen to describe a system is arbitrary,
although there may often be practical reasons for choosing one set rather than another.
These rules set very stringent restrictions on the way
components can be chosen, so it is a good idea to spend
time examining them carefully. First, notice that the solid
phases we encounter most often in geochemical situations
have compositions that are either fixed or are variable
only within bounds allowed by stoichiometry and crystal
structure. This means that if we are asked to find components for a single mineral, they must be defined in such
a way that they can be added to or subtracted from the
mineral without destroying its identity. A system consisting only of rhombohedral Ca-Mg carbonates, for
example, cannot be described by entities such as Ca2+,
MgO, or CO2, because none of them can be independently added to or subtracted from the system without violating its crystal chemistry. The most obvious, although
not unique, choice of components in this case would be
CaCO3 and MgCO3.
When a system consists of more than one phase, it is
common to find that some components selected for individual phases are redundant in the system as a whole.
This occurs because it is possible to write stoichiometric
relationships that express a component in one phase as
some combination of those in other phases. For each
stoichiometric equation, therefore, it is possible to remove one phase component from the list of system components and thus to arrive at the independent variables
required by rule 2. It is also possible, and often desirable,
to choose system components that cannot serve as components for any of the individual phases in isolation. Such
a choice is compatible with the selection rules if the
amount of the component in the system as a whole can
be varied by changing proportions of individual phases.
Petrologists usually refer to nonaluminous pyroxenes, for
example, in terms of the components MgSiO3, FeSiO3,
and CaSiO3, even though CaSiO3 cannot serve as a component for any pyroxene considered by itself.
Worked Problem 3.5
Olivine and orthopyroxene are both common minerals in basic
igneous rocks. For the purposes of this problem, assume that
FIG. 3.9. Olivine and pyroxene solid solutions can each be represented by two end-member components.
olivine’s composition varies only between Mg2SiO4 (Fo) and
Fe2SiO4 (Fa), and that orthopyroxene is a solid solution between MgSiO3 (En) and FeSiO3 (Fs). What components might
be used to describe olivine and orthopyroxene individually,
and how might we select components for an ultramafic rock
consisting of both minerals?
The simplest mineral components are the end-member
compositions themselves. The end-member compositions are
completely independent of one another in each mineral and, as
can be seen at a glance in figure 3.9, any mineral composition
in either solid solution can be formed by some linear combination of the end members. Compositions corresponding to Fo,
Fa, En, and Fs, therefore, satisfy our selection rules.
A choice of FeO or MgO would not be valid, because
neither one can be added to or subtracted from olivine or orthopyroxene unless we also add or subtract a stoichiometric
amount of SiO2. Changing FeO or MgO alone would produce
FIG. 3.10. If we choose to describe a system containing olivines
and pyroxenes, we need three components. Phase compositions
lying outside the triangle of system components require negative
amounts of one or more components.
50
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
phase with the composition FeSiO3. Monatomic components such as F, S, or O are also legitimate, even though
fluorine, sulfur, and oxygen invariably occur as molecules
containing two or more atoms. For some petrological
applications, it makes sense to use such components as
CaMg−1, which clearly do not exist as real substances. In
fact, the carbonates discussed above can be characterized
quite well by MgCO3 and CaMg −1, as can be seen from
the stoichiometric relationship:
MgCO3 + CaMg −1 = CaCO3.
FIG. 3.11. An alternative selection of components for the system
in figure 3.10. All olivine and pyroxene compositions now lie
within the triangle.
system compositions that lie off of the solid solution lines in
figure 3.9.
If olivine and orthopyroxene are not isolated phases but are
constituents of a rock, however, we need to choose a different
set of components. The mineral components are still appropriate, but one of them is now redundant. We can eliminate it by
writing a stoichiometric relationship involving the other three.
For example:
MgSiO3 = 0.5Mg2 SiO4 + FeSiO3 − 0.5Fe2 SiO4.
Notice that this is only a mathematical relationship among
abstract quantities, not necessarily a chemical reaction among
phases. We have tried to emphasize this by using an equal sign
rather than an arrow. The solid triangle in figure 3.10 illustrates
the system as defined by the components on the right side.
As required by the negative amount of Fe2SiO4 in the
equation above, magnesium-bearing orthopyroxenes have compositions outside of the triangle defined by the system components. There is nothing wrong with this representation, but it is
awkward for most petrologic applications. A more conventional selection of components is shown in figure 3.11. The
mineral compositions at the ends of the olivine solid solution
are still retained as system components, but the third component, SiO2, does not correspond to either mineral in the system.
It is very important to recognize that components are
an abstract means of characterizing a system. They do not
need to correspond to substances that can be found in nature or manufactured in a laboratory. Orthopyroxenes,
for example, are frequently described by the components
MgSiO3 and FeSiO3, even though there is no natural
Components of this type, known as exchange operators,
have been used to great advantage in describing many
metamorphic rocks (Thompson et al. 1982; Ferry 1982).
Because components are abstract constructions, we are
not required to use them in positive amounts. Fe2O3, for
example, can be described by the components Fe3O4 and
Fe, even though we need to add a negative amount of
Fe to Fe3O4 to do the job:
3Fe3O4 − Fe = 4Fe2O3.
Worked Problem 3.6
The ratio K/Na in coexisting pairs of alkali feldspars and alkali
dioctahedral micas can be used to infer pressure and temperature conditions for rocks in which they were formed. We see
how this is done in chapter 9. For now, let’s ask what selection
of components might be most helpful if we were interested in
→ Na exchange reactions between these two minerals.
K←
The exchange operator KNa−1 is a good choice for a system
component in this case. Feldspar compositions can be generated
from:
NaAlSi3O8 + x KNa−1 = Kx Na1−x AlSi3O8 ,
for any value of x between 0 and 1. Similarly, mica compositions
can be derived from:
NaAl3Si3O10(OH)2 + yKNa −1
= Ky Na1−yAl3Si3O10 (OH)2.
If we also select the two sodium end-member compositions as
components, then all possible compositional variations in the
system can be described. Figure 3.12a shows one way of illustrating this selection. A more subtle diagram, using the same set
of components, is presented in figure 3.12b.
This is not a contrived example. We have selected it to emphasize that chemical components are abstract mathematical
entities, but it is also meant to illustrate how a very common
class of geochemical problems can be reconceived and simplified by choosing components creatively.
A First Look at Thermodynamic Equilibrium
51
Further infinitesimal changes in the amounts of forsterite
or fayalite in the system would result in a small change
in E:
dE = ĒFo dn Fo + ĒFa dn Fa.
This differential equation can be written in the form of
a total differential, from which it can be seen that:
ĒFo = (∂E/∂n Fo )n Fa,
and
ĒFa = (∂E/∂n Fa )n Fo.
This is an idealized process, of course, chosen to demonstrate that the internal energy of a phase can be changed—
in addition to the ways we have already discussed—by
varying its composition. It is more realistic to recognize
that the mixing process as described involves increases
in both entropy and volume. Notice also that the total
internal energy of the phase, E, is not a molar quantity,
because it has not been divided by the total number of
moles in the system. The relationships dEolivine = ĒFodn or
dEolivine = Ē Fa dn can only be valid if the olivine is either
pure forsterite or pure fayalite. In between, as we see in
later chapters, dE takes a nonlinear and generally complicated form. The quantities labeled Ē Fo and Ē Fa above
are more useful if written with these restrictions in mind:
FIG. 3.12. (a) Alkali micas and feldspars can be described by a
component set that includes KNa−1. (b) An orthogonal version of
the diagram in (a). Its vertical edges both point to KNa−1.
CHANGES IN E, H, F, AND G
DUE TO COMPOSITION
Consider an open system containing only magnesian
olivine, a pure phase. The internal energy of the system
is equal to E = ĒFonFo, where ĒFo is the molar internal energy for pure forsterite and n Fo is the number of moles of
the forsterite component in the system. Suppose that it
were possible to add a certain number of moles of the
fayalite component, nFa, to the one-phase system without
causing any increase in energy as a result of the mixing
itself. Because E is an extensive property, the internal energy of the phase would then be equal to:
E = ĒFo n Fo + ĒFa nFa.
Ei = (∂E/∂ni)S,V,n j ≠ i.
(3.14)
We have now identified a new thermodynamic quantity, the partial molar internal energy, which describes
the way in which total internal energy for a phase responds to a change in the amount of component i in the
phase, all other quantities being equal. You may think of
it, if you like, as a chemical “pressure” or a force for energy change in response to composition, in the same way
that pressure is a force for energy change in response to
volume. To emphasize the importance of this new function, it has been given the symbol µ i and is called the
chemical potential of component i in the phase.
The internal energy of a phase, then, is:
E = E(S, V, n1, n2, . . , nj).
We can rewrite equation 3.5 to include the newfound
chemical potential terms:
dE ≤ TdS − PdV + µ1dn1 + µ2dn2 + . . .
+ µjdnj ≤ TdS − PdV + µidni.
Σ
(3.15a)
52
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
In the same way, it can be shown that the auxiliary functions H, F, and G are also functions of composition:
dH ≤ TdS + VdP + µ1dn1 + µ2dn2 + . . .
+ µj dnj ≤ TdS + VdP + µi dni ,
(3.15b)
dF ≤ −SdT − PdV + µ1dn1 + µ2dn2 + . . .
+ µj dnj ≤ −SdT − PdV + µi dni ,
(3.15c)
Σ
Σ
dG ≤ −SdT + VdP + µ1dn1 + µ2dn2 + . . .
+ µj dnj ≤ −SdT + VdP + µidni .
(3.15d)
Σ
Chemical potential, therefore, can be defined in several
equivalent ways:
It is easiest to examine the equilibrium condition
among phases if we consider only the simple (and geochemically unlikely) situation in which system components correspond one-for-one with components of the
individual phases. We also consider only the equilibrium
conditions for an isolated system. We are doing this only
for the sake of simplicity, however. We would arrive at
the same conclusions if we took on the more difficult
challenge of a closed or open system, or if we considered
a set of system components that differ from the set of all
phase components.
Because the system is isolated, we know that its extensive parameters must be fixed. That is,
ΣdE
= ΣdS
= ΣdV
= Σdn
= Σdn
µi = (∂E/∂ni)S,V,n j ≠ i
(3.16a)
dEsys =
Φ
=0
= (∂H/∂ni)S,P,n j ≠ i
(3.16b)
dSsys
Φ
=0
= (∂F/∂ni)T,V,n j ≠ i
(3.16c)
dVsys
Φ
=0
= (∂G/∂ni)T,P,n j ≠ i
(3.16d)
dn1,sys
and can be correctly described as the partial molar enthalpy, the partial molar Helmholtz function, or the partial molar free energy, provided that the proper variables
are held constant, as indicated in equations 3.16a–d.
CONDITIONS FOR
HETEROGENEOUS EQUILIBRIUM
We are now within reach of a fundamental goal for this
chapter. Having examined the various ways in which internal energy is affected by changes in temperature, pressure, or composition, we may now ask: what conditions
must be met if a system containing several phases is in
internal equilibrium? This is a circumstance usually referred to as heterogeneous equilibrium.
A system consisting of several phases can be characterized by writing an equation in the form of equation 3.15a for each individual phase:
dEΦ ≤ TΦdSΦ − PΦdVΦ +
Σ (µ
iΦ dniΦ ),
where we are using the subscript Φ to identify properties
with the individual phase. For example, if the system contained phases A, B, and C, we would write an equation
for dEA, another for dE B, and a third for dEC . The final
term of each equation is the sum of the products µi dni
for each system component (1 through i) for the individual phase.
dn2,sys
.
.
.
dni,sys =
1,Φ =
0
2,Φ =
0
Σdn
i,Φ
= 0.
There is one dni,sys equation for each component in the
system. Despite these equations, there is no constraint
that prevents the relative values of E, S, V, or the various
n’s from readjusting themselves, as long as their totals
remain zero. That is, the various individual values of EΦ,
VΦ, and the ni,Φ are not independent. Whatever leaves
one phase in the system must show up in at least one of
the others. On the other hand, there are constraints on
the intensive parameters in the system. To see what they
are, let’s write an equation for the total internal energy
change for the system, dEsys:
dEsys =
Σ dE = 0 = Σ(T dS
+ Σ(Σ (µ dn )).
Φ
Φ
i,Φ
Φ)
−
Σ(P dV )
Φ
Φ
iΦ
The only way to guarantee that Σ dEsys = 0 is for each
of the terms on the right side of this equation to equal
zero. We have already agreed that dS, dV, and each of
the dni might differ from phase to phase. It would be a remarkable coincidence if TΦ, PΦ, and each of the µiΦ could
also vary among phases in such a way that Σ(TΦdSΦ),
Σ(PΦdVΦ), and Σ(Σ(µiΦ dniΦ)) were always equal to
zero. Fortunately, we do not need to rely on coincidence.
Unlike extensive properties, intensive properties are not
free to vary among phases in equilibrium. We showed
A First Look at Thermodynamic Equilibrium
earlier in this chapter that this is true for temperature
and pressure; at equilibrium
TA = TB = . . . = TΦ
PA = PB = . . . = PΦ.
53
respect to the mole fraction of FeSiO3 is identical in pyroxene
and in quartz. If it were possible to add the same infinitesimal
amount of FeSiO3 to each, the free energies of the two phases
would each change by the same amount:
dG = µFeSiO dnFeSiO .
3
3
It should be evident now that the same is true for chemical potentials; at equilibrium
The Gibbs-Duhem Equation
µ1A = µ1B = . . . = µ1Φ
µ2A = µ2B = . . . = µ2Φ
.
.
.
µiA = µiB = . . . = µiΦ.
Because this conclusion is so crucial in geochemistry, we
emphasize it again: At equilibrium, the chemical potential of any component must be the same in all phases in a
system. To be sure that these general results are clear, let’s
examine heterogeneous equilibrium in a specific system.
Worked Problem 3.7
An experimental igneous petrologist, working in the laboratory,
has produced a run product which consists of quartz (qz) and
a pyroxene (cpx) intermediate in composition between FeSiO3
and MgSiO3. Assuming that the two minerals were formed in
equilibrium, what conditions must have been satisfied?
To answer this question, first choose a set of components
for the system. Several selections are possible, some of which
we considered in an earlier problem. This time, let’s choose the
end-member mineral compositions FeSiO3, MgSiO3, and SiO2.
Heterogeneous equilibrium then requires that the intensive parameters be:
Tqz = Tcpx = T
Pqz = Pcpx = P
µ SiO ,qz = µ SiO ,cpx = µ SiO
2
2
2
µ FeSiO ,qz = µ FeSiO ,cpx = µ FeSiO
3
3
3
µMgSiO ,qz = µMgSiO ,cpx = µMgSiO .
3
3
3
Notice that the chemical potentials of FeSiO3 and MgSiO3 are
defined in quartz, despite the fact that they are not components
of the mineral quartz itself, and µ SiO2 is defined in pyroxene,
although SiO2 is not a component of pyroxene. All three are
system components. The chemical potential of any component
is a measure of the way in which the energy of a phase changes
if we change the amount of that component in the phase. For
example, if chemical potentials are defined by equation 3.16d,
the constraint on FeSiO3 should be read to mean that at constant temperature and pressure, the derivative of free energy with
One final, very useful relationship can be derived from
this discussion of chemical potentials. Consider equation 3.15a for a phase that is in equilibrium with other
phases around it. It is possible to write an integrated
form of equation 3.15a:
E = TS − PV +
Σµ n .
i i
Therefore, the differential energy change, dE, that takes
place if the system is allowed to leave its equilibrium
state by making small changes in any intensive or extensive properties is:
dE = TdS + SdT − PdV − VdP +
+ ni dµi .
Σ
Σ µ dn
i
i
To see how the intensive properties alone are interrelated,
subtract equation 3.15a, which was written in terms of
variations in extensive parameters alone, from this equation to get:
0 = SdT − VdP +
Σ n dµ .
i
i
(3.17)
This expression is known as the Gibbs-Duhem
equation. It tells us that there can be no net gradients in
intensive parameters at equilibrium. Of more specific interest in geochemical problems, at constant temperature
and pressure, there can be no net gradient in internal
energy as a result of composition in a phase in internal
equilibrium. This can be seen, in a way, as another way
of stating the conclusions from our discussion of heterogeneous equilibrium.
SUMMARY
What is equilibrium? Our discussion of the three laws
of thermodynamics has led us to discover several characteristics that answer this question. First, systems in
equilibrium must be at the same temperature: this is the
condition of thermal equilibrium. Second, provided that
54
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
work is exclusively defined as the integral of PdV, there
can be no pressure gradient between systems in equilibrium: this is the condition of mechanical equilibrium.
Finally, the chemical potential of each component must
be the same in all phases at equilibrium. This is the least
obvious of the conditions we have discussed, and is the
focus of many discussions in subsequent chapters.
We have also repackaged the internal energy function
in several ways and examined the role of entropy, which
will always be maximized during the approach to equilibrium. To demonstrate that each of the equilibrium
criteria is a direct consequence of the three laws, we used
the internal energy function, E(S, V, n), as the basis of
our discussion. As we progress, though, we will quickly
abandon E in favor of the Gibbs free energy. Any of the
energy functions, however, can be used to clarify the conditions of equilibrium.
With the ideas we have introduced in this chapter, we
are now ready to begin exploring geologic environments.
This has only been a first look at thermodynamics, however. We return to the topic many times.
suggested readings
There are many good introductory texts in thermodynamics,
although most are written for chemists rather than for geochemists. Among the best, or at least most widely used, texts for
geochemists are listed here, along with excellent but challenging articles and historical accounts.
Denbigh, K. 1968. The Principles of Chemical Equilibrium.
London: Cambridge University Press. (Heavy going, but almost anything you want to know is in here somewhere.)
Ferry, J. M., ed. 1982. Characterization of Metamorphism
through Mineral Equilibria. Reviews in Mineralogy 10.
Washington, D.C.: Mineralogical Society of America. (An
enlightening selection of articles designed for classroom use.
Chapters 1–4 are particularly relevant to our first discussion
of thermodynamics. Be prepared to stretch.)
Fraser, D. G., ed. 1977 Thermodynamics in Geology. Dordrecht,
Holland: D. Reidel. (Topical chapters showing both fundamentals and applications of thermodynamics.)
Goldstein, M., and I. F. Goldstein. 1978. How We Know: An
Exploration of the Scientific Process. New York: Plenum.
(The historical development of our notions of heat and temperature is discussed in chapter 4 of this thought-provoking
book.)
Greenwood, H. J., ed. 1977. Short Course in Application of
Thermodynamics to Petrology and Ore Deposits. Mineralogical Association of Canada. (A very readable book, with
many sections of interest to economic geologists.)
Kern, R., and A. Weisbrod. 1967. Thermodynamics for Geologists. San Francisco: Freeman, Cooper. (Perhaps the simplest
treatment of thermodynamics available for the geologist,
although now a bit dated.)
Lewis, G. N., and M. Randall. 1961. Thermodynamics, revised
by K. S. Pitzer and L. Brewer. New York: McGraw-Hill. (A
classic text still in widespread use.)
Nordstrom, D. K., and J. Munoz. 1985. Geochemical Thermodynamics. Menlo Park: Benjamin Cummings. (An excellent
text for the serious student.)
Smith, E. B. 1982. Basic Chemical Thermodynamics, 3rd ed.
New York: Oxford University Press. (A remarkably clear text
for the serious beginner.)
Thompson, J. B., J. Laird, and A. B. Thompson. 1982. Reactions
in amphibolite, greenschist, and blueschist. Journal of Petrology 23:1–27. (A good study for the motivated student. Reactions in amphibole-bearing assemblages are discussed in
terms of exchange operators.)
Vanserg, N. 1958. Mathmanship. The American Scientist 46:
94A–98A. (This article does not deal with thermodynamics
directly, but is a commentary on the use of symbols. Its author was better known by his real name, Hugh McKinstry,
professor of geology at Harvard for many years.)
A First Look at Thermodynamic Equilibrium
PROBLEMS
(3.1)
Refer to figures 3.9 and 3.10. Notice that by adding or subtracting SiO2, we change the proportions
of olivine and pyroxene in a rock, but not their compositions. Changing the amount of any other
system component in figure 3.9 or 3.10, however, alters both the compositions and the proportions
of minerals. Can you find another set of components that includes one which, when varied, changes
the compositions but not the proportions of olivine and pyroxene?
(3.2)
Show that the work done by 1 mole of a gas obeying the Van der Waals equation during isothermal
expansion from V1 to V2 is equal to w = RT ln([V̄2 − b]/[V̄1 − b] + a([1/V̄2] − [1/V̄1]).
(3.3)
How much work can be obtained from the isothermal, reversible expansion of one mole of chlorine
gas from 1 to 50 liters at 0°C, assuming ideal gas behavior? What if chlorine behaves as a Van der
Waals gas? Use the result of problem 3.2; a = 6.493 l 2 atm mol−2, b = 0.05622 l mol−1.
(3.4)
What is the maximum work that can be obtained by expanding 10 grams of helium, an ideal gas,
from 10 to 50 liters at 25°C? Express your answer in (a) calories, (b) joules.
(3.5)
Suppose that a mixture of two inert gases is divided into two containers separated by a wall. Imagine
a microscopic valve placed in the wall that allows molecules of one gas to move into container A and
molecules of the other gas to move into container B, but does not allow either gas to move in the
opposite direction. Such an imaginary device is called Maxwell’s demon. How would such a device
violate the Second Law of Thermodynamics?
(3.6)
What conditions must be satisfied if hematite, magnetite, and pyrite are in equilibrium in an ore
assemblage?
(3.7)
Construct a triangular diagram for the alkali feldspar-mica system similar to figure 3.12a, but in
which you swap the positions of NaAl3Si3O10(OH)2 and KAlSi3O8. What component must now take
the place of KNa−1 at the top corner of the diagram? Where would this component have been plotted
on figure 3.12a? Where does KNa−1 plot on your diagram?
(3.8)
Using the Maxwell relations, verify that −(∂V/∂T )P /(∂P/∂T )V = (∂V/∂P)T .
55
CHAPTER FOUR
HOW TO HANDLE SOLUTIONS
OVERVIEW
This is a chapter about solutions, important to us because most geological materials have variable compositions. They are, in fact, mixtures at the submicroscopic
scale between idealized end-member substances such as
albite, water, grossular, dolomite, or carbon dioxide,
which lose their molecular identities in the mixture. To
see how thermodynamics can be used to predict the
equilibrium state in a system dominated by phases with
variable compositions, we examine first the structure of
solutions of solids, liquids, and gases, and discuss ways
in which this architecture is reflected in mole fractions
of end-member components.
At the end of chapter 3, we developed thermodynamic equations that describe the state of equilibrium
for a heterogeneous system. Among these equations are
constraints implying that the compositions of phases in
equilibrium are controlled by the system’s overall drive
toward a minimum ∆G. As we study the structure of
solutions in this chapter, we look explicitly at the compositions of phases in equilibrium and apply the thermodynamic principles from chapter 3. Specifically, we begin
to develop equations that relate mole fractions and other
mixing parameters (which we introduce) to ∆G.
The results of this investigation can be applied to many
problems in the realm of geochemistry. We begin applying
56
them in this chapter by studying aqueous fluids and the
phenomenon of solubility. This study not only presents
an opportunity to use the equations we develop, but also
prepares the way for our later discussions of the oceans,
diagenesis, and weathering reactions.
WHAT IS A SOLUTION?
Most phases of geochemical interest do not have fixed
compositions. With the conspicuous exception of quartz,
each of the major rock-forming minerals is a solid solution between two or more end-member molecules. The
range of allowable compositions in any solution is dictated by rules of structural chemistry that consider both
crystal architecture and the size and electronic configuration of ions. For many minerals, these rules permit
liberal replacement of iron by magnesium or calcium,
sodium by potassium, or silicon by aluminum. At first
glance, the degree of flexibility that we observe in mineral
compositions looks random. It seems to spoil any hope
that we might be able to apply thermodynamic principles
to geologic materials. Liquids and gases follow even fewer
structural guidelines and have even more variable compositions. Can the principles of equilibrium thermodynamics we have just explored in chapter 3 make sense
in the more complicated world of these real phases? To
answer, we first need to know more about solutions.
How to Handle Solutions
Crystalline Solid Solutions
Most geologically important materials are crystalline
solids; that is, solids in which atoms are arranged in a
periodic three-dimensional array that maintains its gross
structural identity over fairly large distances. A “fairly
large” distance is not easily defined, however. If we see a
mineral grain in thin section that has continuous optical
properties, it is clearly large enough. Even submicroscopic
grains are judged to be crystalline if their structures are
continuous over distances sufficient to yield an X-ray
diffraction pattern. We can get reliable results from thermodynamic calculations for most geochemical purposes
if the materials we study are crystalline at this scale. We
can still apply principles of thermodynamics at smaller
scales, but the rules become more complex. In very finegrained crystalline aggregates, for example, surface properties make a major contribution to free energy. In highly
stressed materials, too, dislocations or defects constitute
a significant volume fraction of the material and call for
special methods. Finally, intimately intergrown materials,
such as those in figure 4.1, in which two or more phases
are structurally compatible and randomly intermixed, are
difficult to model. We take a look at some of these special
cases in chapter 10, but for now we stick with crystalline
materials that have uniform properties over distances of
thousands of unit cell repeats.
57
The rules that dictate where atoms sit in a crystalline
substance are quite flexible. Two major factors govern
which atoms may occupy a given site: their ionic charges
and their sizes. Charge is important because of the need
to obtain electrical neutrality over the structure. Size is
important because it influences the degree of “overlap”
between the orbitals of valence electrons of the ion and
those that surround it. Recall our discussion of these
principles in chapter 2 and scrutinize figure 2.11 again.
You will find that several ions available to a growing
crystal will satisfy these constraints. In addition, the
charge balance requirement can be satisfied by the formation of randomly distributed vacancies or by inserting ions into defects or into positions that are normally
unoccupied. Local charge imbalances caused when an
inappropriately charged ion occupies a site, therefore,
can be averaged out over the structure as a whole. As a
result, minerals are most properly viewed as crystalline
solutions in which many competing ions substitute freely
for one another. We must consider both the bulk composition of a solution and the way in which it is mixed
as we describe it with thermodynamics.
To illustrate, let us compare two possible mixing
schemes, using the clinopyroxene solid solution CaMgSi2O6 (diopside)-NaAlSi2O6 (jadeite) as an example. First,
take a look at figure 4.2 as we do a quick overview of
pyroxene structural chemistry. The atoms in silicates are
FIG. 4.1. High-resolution transmission electron microscope (HRTEM) image of an intergrowth of
ortho- and clinopyroxenes. This material is crystalline, but the identity of the structural unit
changes after a random, rather small number of repeats (arrows). Thermodynamic characterization
of such material is very difficult.
58
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 4.2. The crystal structure of C2/c pyroxene, showing the geometry of M1 and M2 cation sites,
each of which is surrounded by six oxygen atoms, and their relationship to tetrahedral (T) sites,
which are largely occupied by silicon atoms. This diagram, a projection onto the (100) crystallographic plane, is drawn to emphasize the geometry of cation sites. (Modified from Cameron and
Papike 1981.)
arranged so that each silicon is surrounded by four
oxygens to form a tetrahedral unit with a net electrical
charge of −4. These are connected to each other and to
other atoms by bonds that are largely ionic in nature.
The manner in which the units are linked together, and
the identity of other atoms that occupy sites between
them, determine the mineral’s structure and establish its
bulk chemical identity. (Review our discussion of coordination polyhedra in chapter 2.) In a pyroxene, SiO4
tetrahedra are linked by corners to form long single
chains. Diopside and jadeite are C2/c pyroxenes, in which
cations occupy two other types of sites, designated M1
and M2, between these chains. M1 is surrounded by six
oxygens and is nearly a regular octahedron. It is occupied by the relatively small cations Mg2+ and Al3+. The
M2 site is eight-coordinated, 20–25% larger, and irregular in shape. Ca2+ and Na+, relatively large cations,
prefer this site. For a detailed but very readable description of pyroxene crystal chemistry, see the articles by
Cameron and Papike (1980, 1981).
So much for the general context. Now, if we consider
only the pure end-member pyroxenes, we should have
no trouble figuring out how sites are occupied, because
there is only one type of cation available for each site.
Once we investigate intermediate compositions, however,
two possible cation arrangements suggest themselves.
The first, called molecular mixing, assumes that charge
balance is always maintained over very short distances.
That is, calcium and magnesium ions substitute for
sodium and aluminum ions as a coupled pair: each time
How to Handle Solutions
we find a Ca2+ in an M2 site, there is a Mg2+ in an adjacent M1 site. The end-member pyroxenes mix as discrete
molecules and complete short-range order results. In this
case, the mole fraction of CaMgSi2O6 in the solid solution (XCaMgSi2O6,cpx) is equal to the proportion of the
diopside molecule present. Until recently, molecular
mixing was widely assumed by petrologists.
The other possibility, known as mixing on sites, assumes that Ca2+ and Na+ are randomly distributed on M2
sites, unaffected by whether adjacent M1 sites are filled
with Mg2+ or Al3+ ions. This style of mixing may produce
local charge imbalances, but the average stoichiometry
is satisfied over long distances even in the absence of
short-range order. Because each of the sites, according to
this model, behaves independently of the others, we
must now calculate the mole fraction XCaMgSi2O6,cpx as the
product of XCa,M2,cpx and XMg,M1,cpx. This follows from
a basic principle of statistics: if the probability of finding
a Mg2+ ion in M1 is some value x and the probability of
finding a Ca2+ in M2 is y, then the combined probability
of finding both ions in the structure is xy. Thus, if 80%
of the ions in the M1 site are Mg2+ and 80% of the M2
atoms are Ca2+, XCaMgSi2O6,cpx is equal to 0.64. If the
tetrahedral site were partially occupied by something
other than silicon, we would have to multiply by XSi,T,cpx
as well.
How can we tell which of these two mixing models
is the correct one? It is logical to infer that the degree of
randomness among atoms distributed on M1 and M2
affects the molar entropy of a phase, and therefore each
of the energy functions E, H, F, and G. In theory, then,
it should be possible to use these two extreme models to
calculate the relative stabilities of clinopyroxene solutions and see which one matches what we see in the natural world. To do this requires calculating bond energies
from electrostatic equations and detailed crystallographic
data. Large-scale computer models exist for this purpose, but are successful only with very simple structures.
Ronald Cohen (1986), however, has done the next best
thing. By evaluating a large body of calorimetric data
and phase equilibrium studies for aluminous clinopyroxenes, he concluded that the macroscopic behavior of
pyroxene solid solutions provides no observational justification for molecular mixing. His conclusions suggest
that other common silicate solutions (feldspars, micas,
amphiboles) behave similarly. Unless crystallographic evidence suggests otherwise, therefore, we may assume that
mixing takes place on sites.
59
This is a significant finding. It is not the end of the
discussion about how to model clinopyroxenes, however. We have assumed that each of the cations occupies
only one type of site, for example. In fact, this is not the
case. Although Mg2+ prefers the smaller M1 site, it may
also be present in M2. Fe2+ and Mn2+ may also be found
in either M1 or M2. Other possible complications include the stabilizing effect felt by certain transition metal
cations as a result of site distortions (more about this in
chapter 12) and the degree of covalency of bonds in the
structure (refer again to chapter 2). Because we generally
have no way of testing each of these effects directly,
thermodynamic interpretation of crystalline solutions is
an empirical process, based largely on the macroscopic
behavior of materials. Models of site occupancy rely
strongly on observations made by spectroscopic and
X-ray diffraction methods, and on our understanding
of the rules of crystal chemistry that govern element
partitioning. These still leave considerable uncertainty,
so most calculations can only be performed by making
simplifying assumptions.
Worked Problem 4.1
To illustrate how we can calculate mole fractions in a disordered mineral structure, consider the following analysis of a
C2/c omphacite (pyroxene), which was first reported by Clark
et al. (1969). The abundances are recorded as numbers of cations
per six oxygens in the clinopyroxene unit cell. (The analysis in
Clark et al. [1969] also includes 0.002 Ti3+, which we have
omitted for simplicity. Including Ti3+ would change the contents of the M1 site slightly.)
Atom
Abundance
Si
Al
Fe3+
Fe2+
Mg
Ca
Na
1.995
0.238
0.123
0.116
0.582
0.583
0.325
To estimate the site occupancy, we first assume that all tetrahedral sites are filled with either Si or Al (the only two atoms in
the analysis that regularly occupy tetrahedral sites). Because
the ratio of tetrahedral cations to oxygens in a stoichiometric
pyroxene should be 2/6, we see that 0.005 Al must be added to
the 1.995 Si to fill the site. An alternative, which might also be
justifiable, is to assume that Si is the only tetrahedral cation,
and simply normalize the Si in the analysis to 2.00.
Of the remaining cations, we may assume that Ca and Na
(both of which are large) are restricted to the M2 site, whereas
60
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Al and Fe3+ (both relatively small) prefer the more compact M1
site. X-ray structure refinement by Clark et al. indicates, further, that the ratio Fe/Mg in M1 is 0.40 and in M2 is 0.54.
Using these pieces of information, we proceed as follows:
1. The amount of Fe2+ in M1 is some unknown quantity that
we call x. The amount of Mg in M1, also unknown, we call
y. We know, however, that (Fe3++ x)/y = 0.4.
2. The amount of Fe2+ in M2 is unknown, but must be equal
to 0.116 − x, just as the amount of Mg in M2 must be
0.582 − y. We know, therefore, that (0.116 − x)/(0.582 − y)
= 0.54.
3. If we solve the two equations in x and y simultaneously, we
conclude that x = 0.092 and y = 0.538.
4. The total cation abundance in M1 (ΣM1) is, therefore,
ΣM1 = (Al) + (Fe
3+) + x + y = 0.233 + 0.123 + 0.092
+ 0.538 = 0.986.
The total cation abundance in M2 (ΣM2), similarly, is:
ΣM2 = (Ca) + (Na) + (0.116 − x) + (0.582 − y) = 0.583
+ 0.325 + 0.024 + 0.044 = 0.976.
5. To calculate the mole fractions on each site, we normalize
their respective contents by 1/ΣM1 or 1/ΣM2 to get:
M1
Cation
Al
Fe3+
Fe2+
Mg
M2
X,M1
Cation
X,M2
0.236
0.125
0.093
0.546
Ca
Na
Fe2+
Mg
0.598
0.333
0.025
0.045
One way to model this pyroxene is to choose likely molecular end members such as NaAlSi2O6 (jadeite), CaMgSi2O6
(diopside), NaFe3+Si2O6 (aegirine), among others. We might
then calculate mole fractions of each on the assumption that
an M1 site occupied by Al is always adjacent to an M2 site
containing Na, an M2 occupied by Na is either adjacent to an
Al or an Fe3+ in M1, and so forth. We have just argued that molecular mixing is generally a poor way to proceed, however. Instead, if we wish to consider a reaction in which the NaAlSi2O6
component of this pyroxene is involved, we calculate its mole
fraction as the cumulative probability of finding both Al and
Na at random on their respective sites:
XAl,M1,cpx XNa,M2,cpx = (.236)(.333) = 0.079.
Likewise,
XAeg,cpx = XFe3+,M1,cpx XNa,M2,cpx = (.125)(.333)
= 0.042,
and
XDi,cpx = XMg,M1,cpx XCa,M2,cpx = (.546)(.598) = 0.327.
Amorphous Solid Solutions
Glasses and other amorphous geological materials
have structures that differ from crystalline materials only
in their degree of coherency. An amorphous solid has no
long-range structural continuity but may still consist of
numerous very small-scale ordered domains. Many of the
concerns we had about mixing in crystalline solids are
therefore also valid for amorphous materials. Their greatest importance for geochemists, however, is that they
provide a conceptual bridge toward our understanding
of liquids.
Melt Solutions
The only geologically significant melt systems that
we understand in any detail are silicate magmas. Our
knowledge of silicate melt structures has been acquired
primarily since the late 1970s, with the help of improved X-ray diffraction and spectroscopic methods.
Geochemists now understand that silicon and oxygen
atoms continue to associate in SiO4 tetrahedral units beyond the melting point, and that the units tend to form
polymers. Depending on the composition of the melt,
these units may be chain polymers similar to the ones
found in pyroxenes, or they may be branched structures
of various sizes—the remnants (or precursors) of sheet
and framework crystalline structures.
The degree of polymerization depends on temperature, but is most strongly influenced by the proportion
of network-formers such as silicon and aluminum and
network-modifiers such as Fe2+, Mg2+, Ca2+, K+, and
Na+. In silica-rich melts, the highly covalent nature of
the Si-O bond tends to stabilize structures in which a
large number of oxygen atoms are shared between SiO4
units. As a result, these melts are highly viscous and have
low electrical conductivities. Adding even a small proportion of Na2O or another network modifier, however,
reduces viscosity drastically and increases melt conductivity. As in crystalline silicates, the bonds between oxygen and the larger cations are ionic. These break more
easily than the Si-O bonds, thus reducing the melt’s
overall tendency to form large, tenacious polymeric structures. Water also acts as a network modifier in silicarich melts, apparently by reacting with the bridging
oxygens in a polymer. This reaction may be described
generically by:
How to Handle Solutions
FIG. 4.3. Relationship between melt viscosity and composition at
1150°C. The variable on the horizontal axis is the atomic ratio of
oxygen atoms to network-formers (Si + Al + P). The filled circles
represent compositions of natural magmas.
Component
.75 < XSiO2 < .81
Di
.55 < XSiO2 < .65
Di
SiO2
TiO2
FeO
MnO
MgO
CaO
MgAl2O4
CaAl2O4
NaAlO2
KAlO2
10.50
−3.61
6.17
−6.41
−2.23
−3.61
−4.82
−1.74
15.1
15.1
9.25
−4.26
−6.64
−5.41
−4.27
−5.54
−1.71
−0.22
7.48
7.48
To illustrate this method, we consider two different melt
compositions, a granite and a diabase. For each, we have tabulated a bulk analysis in weight percentage of oxides and a recalculated analysis in mole percentage. In the recalculation, all
iron has been converted to FeO, phosphorus has been added to
SiO2, water has been omitted (a serious omission, but at least
we can compare anhydrous melt viscosities this way), and the
total has been adjusted to 100 mole percent.
|
|
|
→ 2(Si-OH).
H2O + Si-O-Si ←
|
|
|
Figure 4.3 illustrates the degree to which melt structure
influences thermodynamic properties by showing the qualitative relationship between viscosity and composition.
Worked Problem 4.2
Before much of the modern experimental work to examine to
relationship between melt structure and viscosity was begun,
Jan Bottinga and Daniel Weill (1972) developed a simple empirical method for predicting melt viscosity. They noted that
the natural logarithm of viscosity (η) may be approximated by a
function of composition that is linear within restricted intervals
of SiO2 content. Specifically, they showed that the empirical
coefficients Di in the equation:
ln η =
Component
Granite
Wt. %
Diabase
Wt. %
SiO2
TiO2
Al2O3
Fe2O3
FeO
MnO
MgO
CaO
Na2O
K2O
H2O
P2O5
70.18
0.39
14.47
1.57
1.78
0.12
0.88
1.99
3.48
4.11
0.84
0.19
50.48
1.45
15.34
3.84
7.78
0.20
5.79
8.94
3.07
0.97
1.89
0.25
Mole %
Mole %
80.12
0.38
1.83
0.11
1.02
0.00
0.47
2.41
7.68
5.97
58.45
1.26
10.83
0.12
9.95
4.76
0.00
6.28
6.85
1.43
SiO2
TiO2
FeO
MnO
MgO
CaO
MgAl2O4
CaAl2O4
NaAlO2
KAlO2
ΣXiDi
are constants at a given temperature and a given value of XSiO .
2
In the table below are values of Di at 1400oC for melts with
two different silica contents, calculated statistically by Bottinga
and Weill from the measured viscosities of a large number of
silicate melts. Notice from the arithmetic signs on Di that
some melt components tend to increase viscosity (that is, they
act as network-formers), whereas others tend to decrease it (by
acting as network-modifiers). Notice also that some components,
such as FeO, behave differently in silica-rich melts than in silicapoor ones.
61
For the granite,
ln η =
Σ Xi Di = (.8013)(10.5) + (.0038)(−3.61)
+ (.0183)(6.17) + (.0011)(−6.41)
+ (.0102)(−2.23) + (.0047)(−4.82)
+ (.0241)(−1.74) + (.0768)(15.1)
+ (.0597)(15.1) = 10.48
η
= 3.56 × 104 poise = 3.56 × 103 kg m−1 sec−1.
62
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
For the diabase,
ln η =
η
Σ
Xi Di = (.5845)(9.25) + (.0126)(−4.26)
+ (.1083)(−6.64) + (.0019)(−5.41)
+ (.0995)(−4.27) + (.0476)(−5.54)
+ (.0628)(−0.22) + (.0685)(7.48)
+ (.0143)(7.48) = 4.54
= 9.37 × 101 poise = 9.37 kg m−1 sec−1.
To an even greater extent than is true for crystalline
solids, our limited understanding of melt structures makes
it impossible to calculate thermodynamic quantities from
models that consider the potential energy contributions
of individual atoms. Enthalpies and free energies of
formation for melts are determined on macroscopic systems, and therefore represent an average of the contributions from an extremely large number of local structures.
Thermodynamic models that make simple assumptions
about how mixing takes place between melts of different
end-member compositions are surprisingly successful, but
tend to be more descriptive than predictive. We discuss
these more fully in chapter 9.
Electrolyte Solutions
An electrolyte solution is one in which the dissolved
species (the solute) are present in the form of ions in a
host fluid that consists primarily of molecular species
(the solvent). In the geological world, these are usually
aqueous fluids, although CO2-rich solvents may be increasingly important with increasing depth in the crust
and upper mantle. Electrolyte solutions do not exhibit
the large-scale periodic structures found in either melts
or solids. The ions and solvent molecules, however,
cannot generally be treated as a mechanical mixture of
independent species. Electrostatic interactions among
solute and solvent species place limits on the concentration of free solute ions and can produce complexes that
influence the thermodynamic behavior of a solution.
These effects are most obvious as the concentration of
dissolved species increases and require a rather complicated mathematical treatment. Many of the later sections
of this chapter are devoted to understanding the behavior
of aqueous electrolyte solutions.
Gas Mixtures
Gases at low pressure mix nearly as independent
molecules. In this way, they are perhaps the simplest of
geologic materials and therefore the best to study initially.
Except for transient species in high-energy environments,
most gas species are electrically neutral. Their neutrality
and the relatively large distances between molecules at
low pressure allow us to treat atmospheric and most
volcanic gases as mechanical mixtures. As such gas mixtures interact with other geological materials (say, in
weathering reactions), the thermodynamic influence of
each of the individual gas species is directly proportional
to its abundance in the mixture. Only at elevated pressures do gas molecules interfere with each other and begin
to alter the thermodynamic behavior of the mixture.
SOLUTIONS THAT BEHAVE IDEALLY
For each type of solution we have considered in this brief
discussion, there is some range of composition, pressure,
and temperature over which the end-members mix as
nearly independent species. For many, this range is distressingly narrow, but it provides us with a simplified
system from which we can extrapolate to examine the
behavior of “real” solutions. When the end-member constituents of solutions act nearly as if they were independent, we refer to their behavior as ideal.
In this section and the following one on nonideal solutions, we begin with a discussion of gases. This is done
partly because gases are found in every geological environment, but largely because their behavior is the easiest
to model.
We have already seen (equation 3.11) that the Gibbs
free energy for a one-component system can be written as
G(T, P). In differential form:
dG = −SdT + VdP.
Furthermore, dG can be recognized as dG = ndḠ = ndµ.
Suppose we are interested in the difference in chemical
potential between some reference state for the system
(for which we will continue to use the superscript 0) and
another state in which temperature or pressure is different. To extract this information, we must integrate
dG between the two states:
∫
Ḡ
Ḡ 0
ndḠ = n
∫
µ
µ0
∫
dµ = −
T
T0
SdT +
∫
P
VdP.
P0
The integration, unfortunately, is harder than it looks.
We can move the variable n outside of the integrals for
Ḡ and µ because the amount of material in a system is
not a function of its molar free energy. We could not do
How to Handle Solutions
the same with S or V, however, because they are, respectively, functions of T and P. To make the problem solvable, we need to supply S(T ) and V(P). The first of these
is difficult enough to define that we will avoid it for now
by declaring that we are only interested in isothermal
processes. In this chapter, therefore, the integral of SdT
is equal to zero. Later, in chapter 9, we consider temperature variations.
The pressure-volume integral is easier to handle. The
function V(P) is obtained from an equation of state, the
most familiar of which is the ideal gas equation: V =
nRT/P. The integral equation for an isothermal ideal gas,
therefore, reduces to:
n
∫
µ
µ0
dµ = nRT
∫
P
P0
dP.
If we do the integration, we see that:
µ − µ0 = RT ln(P/P 0 ).
It is standard practice to choose the reference pressure as
unity, so that this equation takes the form
µ = µ0 + RT ln P.
(4.1)
Notice that because P in equation 4.1 really stands for
P/P 0, it is a dimensionless number. We have now derived
an expression for the chemical potential of a single component in a one-component ideal gas at a specified temperature and pressure.
Next, consider a phase that is a multicomponent
mixture of ideal gases. There must be a stoichiometric
equation that describes all of the potential interactions
between gas species in the mixture:
νj+1 A j+1 + . . . + νi−1 A i−1 + νi A i
= ν1A1 + ν2A 2 + . . . + νj A j .
This is a standard equation that you have seen many
times, but the notation may look a little strange, so let’s
pause to explain it. We keep track of each gas species
by associating it with a unique subscript. Subscripts 1
through j refer to product species and j + 1 through i
refer to reactants. The quantities ν1 through νi are stoichiometric coefficients and A1, A2, A3, . . . , Ai are the
individual gases. If the gas were a two-component system of carbon and oxygen species, for example, and if we
were to ignore all possible species except for CO2, CO,
and O2, then the stoichiometric equation would be
CO + 1–2 O2 = CO2.
63
The species CO, O2, and CO2 here are A1, A2, and A3,
and the stoichiometric coefficients are ν1 = ν2 = 1 and
ν3 = 0.5. We use this notation in stoichiometric equations
throughout the rest of this book.
Returning now to the gas mixture, we find that the
expanded form of equation 3.11 to describe variations in
free energy looks like:
dG = −SdT + VdP +
j
i
Σ µ dn − Σ µ dn ,
k=1
k
k
k=j+1
k
k
(4.2)
in which the chemical potentials are properly defined as
partial molar free energies (equation 3.16d):
µi = (∂G/δni)T,P,nj≠i = Ḡi.
The summation terms in equation 4.2 do not contradict equation 3.15d, where positive and negative compositional effects on free energy were combined implicitly in
a single term. Here, we separate them explicitly merely to
emphasize that quantities associated with product species
(1 through j) are meant to be added to the free energy of
the system; those associated with reactant species (j + 1
through i) are subtracted.
If you have been following this discussion closely, it
might bother you that we said we would use the subscript i in stoichiometric equations to identify real chemical species and have now reverted to the convention of
using i to identify components. As we have emphasized
before, the relationship between species and components
is one of the hardest concepts in this book to understand,
so we encourage you to re-examine the final sections of
chapter 3 if you are confused. The practice of using i in
equation 4.2 to count species implies that there are stoichiometric equations that relate species compositions to
system components. For example, if we were describing
the two-component system of carbon and oxygen species
it would be fair to incorporate µCO, µO2, and µCO2 in
equation 4.2, even though only two of the quantities
CO, O2, and CO2 can be system components. The parallel to the stoichiometic equation CO + 1–2 O2 = CO2 is a
Gibbs-Duhem equation (3.17), which assures us that at
equilibrium µCO + 0.5µO2 = µCO2. The individual quantities dni in equation 4.2, in other words, are not independent, but are tied to the stoichiometric coefficients νi
and related to each other by:
(dn1 /ν1) = (dn2 /ν2) = . . . = (dnj /νj) = −(dnj+1 /νj+1)
= . . . = −(dni−1 /νi−1) = −(dni /νi ) = dζ.
The important point here is that changes in the amount
of each species in the system are proportional by a
64
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
common factor to their stoichiometric coefficients in the
reaction. That factor, dζ, is a convenient measure of how
far a reaction among species has proceeded.
With the insight that νi = dni/dζ, we can rewrite equation 4.2 at constant T and P as:
dG = (ν1µ1 + ν2 µ2 + . . . + νj µj − νj+1 µj+1 − . . .
− νi−1 µi−1 − νi µi )dζ
=
(Σ
j
k=1
νk µk −
i
k k
= ∆Ḡr dζ.
We have just derived a very useful quantity,
∆Ḡr = dG/dζ ,
(4.3)
known as the free energy of reaction or the chemical reaction potential. At equilibrium, it has the value zero.
If ∆Ḡr < 0, then the reaction proceeds to favor the product species; if ∆Ḡr > 0, then the reaction is reversed to
favor reactant species. This is a quantitative explanation
of LeChatlier’s Principle, which states that an excess of
species involved on one side of a reaction causes the reaction to proceed toward the opposite side.
Because this is still a discussion of ideal gas mixtures,
the equation of state for the gas phase remains V = nRT/P,
but the total amount of material in the system, n, is now
equal to the sum of the amounts of species, ni. The pressure exerted by any individual species in the mixture is,
therefore, a partial pressure, defined by:
Pi = ni RT/V = Pni /n = PX i .
For each species in a gas mixture, then, we can write an
expression like equation 4.1:
µi = µ0i + RT ln Pi .
(4.4)
For the mixture as a whole, equations 4.3 and 4.4 tell us
that the free energy of reaction can be divided into two
parts, one of which refers to the free energy in some
standard state (often one in which solutions with variable
compositions consist of a limiting end member) and the
other of which tells us how free energy changes as a result of compositional deviations from the standard state.
In summary, the free energy of reaction is written as:
Σν µ
= Σν µ
∆Ḡr =
i
i
= ∆Ḡ0
or
The new term, Keq, is called the equilibrium constant, and
equation 4.6 is the first answer to our question: “What
controls the compositions of phases at equilibrium?”
It is time to apply what we have discussed to a specific example.
Worked Problem 4.3
The Soviet Venera 7 planetary probe measured the surface
temperature on Venus to be 748 K and determined that the
total pressure of the atmosphere is 90 bar. The Pioneer Venus
mission, which parachuted four probes to the surface in 1978,
determined that the lower atmosphere consists of 96% CO2
and contains 20 ppm of CO. Other gases in the atmosphere
(primarily N2) have a marginal effect on chemical equilibria in
the C-O system. Given these data and assuming that CO2, CO,
and O2 behave as ideal gases, what should be the partial pressure of O2 at the surface of Venus?
The first step is to write a balanced chemical reaction among
the species:
→ CO .
CO + 1–2 O2 ←
2
Equation 4.6 for this example, then, should be written:
0/RT
Keq = e− ∆Ḡ
= PCO2 /(PCO PO1/2
).
2
We can calculate the value of ∆Ḡ0 from the standard molar free
energies of formation of each of the reactant and product
species from the elements at 475°C (748 K):
1
∆Ḡ0 = ∆Ḡf,CO − ∆Ḡf,CO − –2 (∆Ḡf,O )
2
2
= −94.516 − (−42.494) − 1–2 (0.0)
= −52.022 kcal mol−1.
The value of Keq, therefore, is:
Keq = exp[52.022 kcal mol−1/(0.001987 kcal deg−1 mol−1)
(748 K)] = 1.59 × 1015.
From this and the expression for Keq, we calculate that:
i
Σ ln P
+ RT Σ ln P
0
i
νj+1
νi−1 νi
−∆Ḡ0/RT = ln(P1ν1 P2ν2 . . . Pj νj /Pj+1
. . . Pi−1
Pi ),
e−∆Ḡ0/RT = Keq
νj+1 . . . Pνi−1P νi). (4.6)
= (P1ν1P2ν2 . . . Pj νj /Pj+1
i−1 i
Σ ν µ ) dζ
k=j+1
If the gas phase is in internal (homogeneous) equilibrium,
then ∆Ḡr = 0 and:
+ RT
i
i
νi
(4.5)
νj+1 . . . Pνi−1P νi).
= ∆Ḡ0+ RT ln(P1ν1P2ν2 . . . Pj νj /Pj+1
i−1 i
PO2 = (PCO 2 /Keq PCO)2 =
(86.4 bar/(1.59 × 1015)(.00002)(90 bar))2 =
9.11 × 10−22 bar.
This quick “back of the envelope” answer is consistent with
what we know from observations: the amount of free oxygen in
How to Handle Solutions
Venus’s atmosphere is below the detection limits of our planetary
probes. Except as an example of how to use Keq, however, you
should not take this numerical result very seriously. Several
other factors would need to be considered if we were to attempt
a rigorous calculation, including equilibria involving solid silicate and carbonate minerals in the Venusian crust.
With only small modifications, this derivation for
gaseous systems can yield a similar expression that can
be used for ideal liquid or solid solutions. If we stipulate
that both pressure and temperature are constant for these
solutions, then:
dPi = d(PXi) = P dXi.
Worked Problem 4.4
During greenschist facies metamorphism (∼400°C), rocks
containing crysotile, calcite, and quartz commonly react to
form tremolite, releasing both CO2 and water. Assume, for the
moment, that calcite is pure CaCO3, that the other two solids
are pure magnesian end members of crystalline solutions, and
that water and CO2 mix as an ideal gas phase. This reaction is
strongly affected by total pressure, and we have not yet discussed
adjustments for pressure. Suppose, however, that you tried to
simulate this reaction by heating crysotile, calcite, and quartz to
400°C in an open crucible. What would ∆Ḡr be under these
circumstances?
The balanced reaction for this example is:
5 Mg3Si2O5(OH)4 + 6 CaCO3 + 14 SiO2
→ 3 Ca Mg Si O (OH) + 6 CO + 7 H O.
←
2
5 8 22
2
2
2
The equation analogous to 4.4 is therefore:
µi = µi* + RT ln Xi ,
(4.7)
Data relevant to this problem are available in Helgeson (1969):
Tremolite
and the free energy of reaction is:
∆Ḡr = ∆Ḡ* + RT ln [X1ν1X 2ν2 . . . Xjνj /
νj+1 . . . Xνi −1X ν i )].
(Xj+1
i−1 i
CO2
H2O
Crysotile
∆H̄0 −2894140 −87451 −54610 1012150
(4.8)
The superscript * on both µ* and ∆Ḡ* is a reminder
that each is a function of temperature and pressure,
unlike µ0 and ∆Ḡ0, which depend only on temperature.
When we discuss the effects of pressure on mineral equilibria in chapter 9, we will see that it is necessary to add
an extra term to the free energy equations for solid and
liquid phases to allow for compressibility. For gases, that
adjustment is made in the RT ln Keq term.
Because equations 4.5 and 4.8 are similar in form, they
can be combined to formulate a general equation relating
the compositions of several ideal solutions or gas mixtures in a geochemical system to their molar free energies:
∆Ḡr = (∆Ḡ0 + ∆Ḡ*) + RT ln [X1ν1X2ν2 . . .
νj+1 . . .
Xjνj P1ν1P2ν2 . . . Pj νj /(Xj+1
65
(4.9)
ν i−1 X νi Pνj+1 . . . Pνi−1P νi)].
Xi−1
i−1 i
i j+1
Each of the free energies, mole fractions, and partial pressures should be understood to carry a subscript to identify the phase to which they refer. For example, the mole
fraction of CaAl2Si2O8 in plagioclase should be shown
as XCaAl2Si2O8,Pl. Except where they are necessary to avoid
ambiguity, however, we usually follow standard practice
and omit phase subscripts. Let’s try a problem involving
both a gas phase and condensed phases to see how equation 4.9 works in practice.
S̄ 0
258.8
65.91
51.97
118.55
Calcite
Quartz
−279267 −212480
cal
mol−1
41.65
20.87 cal
deg−1
mol−1
The partial pressure of CO2 in the atmosphere is ∼3.2 × 10−4
atm, and the partial pressure of H2O, although variable, is ∼1
× 10−2 atm. If the crucible is open to the atmosphere and mixing is rapid, then these partial pressures should remain roughly
constant.
The molar free energy of reaction can be calculated from:
∆Ḡr = ∆H̄ 0 − T∆S̄0 + RT ln Keq,
in which:
6
X3Ca2Mg5 Si8 O22 (OH),tremPCO
P7
2 H2O
Keq = ———————————————
.
6
5
(XMg3Si2O5 (OH) 4,cry XCaCO3,cc X14
SiO2,qz )
From the data above, we calculate:
∆H̄ 0 = 3(−2894140) + 6(−87451) + 7(−54610)
− 5(−1012150) − 6(−279267) − 14(−212480)
= 121256 cal mol−1
∆S̄ 0 = 3(258.8) + 6(65.91) + 7(51.97) − 5(118.55)
− 6(41.65) − 14(20.87) = 399.32 cal deg−1 mol−1.
Because we assumed that the solid phases have end-member
compositions, each of the mole fractions is equal to one. To
reach the final answer, then, we calculate:
∆Ḡr = 121256 − 673(399.32) + 1.987(673)[6 ln(3.2
× 10−4) + 7 ln(1 × 10−2)] = −255161 cal mol−1.
The reaction, therefore, favors the products under the conditions
specified. We should expect CO2 and water vapor to be released
from our open crucible.
66
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Suppose that we performed the same experiment, but used
an impure calcite with a composition (Ca0.8Mn0.2CO3). How
would you expect the free energy of reaction to change, assuming that Ca-Mn-calcite behaves as an ideal solid solution?
The expressions for ∆Ḡr and Keq remain the same, as do the
values for ∆H̄0 and ∆S̄0. The mole fraction of CaCO3 in calcite
is no longer 1.0, however, so the value of (−6 ln XCaCO ) is no
3
longer zero. We find, now, that:
∆Ḡr = 121256 − 673(399.32) + 1.987(673)
(6 ln(3.2 × 10−4) + 7 ln(1 × 10−2) − 6 ln(0.8)
= −253371 cal mol−1.
The reaction still favors the products, but ∆Ḡr has increased by
1790 cal mol−1.
SOLUTIONS THAT BEHAVE NONIDEALLY
In many cases, it is appropriate to assume that real gases
follow the ideal gas equation of state. More generally,
however, interactions between molecules force us to
modify the ideal gas law by adding correction terms to
adjust for nonideality. Before we worry about the precise
nature of these adjustments, let us illustrate how nonideality complicates matters by writing the corrected
equation of state as a generic power function:
2
3
4
PV = RT + BP + CP + DP + EP + . . . ,
(∂µ/∂P)T = RT (∂ ln a/∂P)T .
(Remember that µ0 is a function of T only, so that
(∂µ0/∂P)T is equal to zero.) This can be recast another way
by recalling from equation 3.11 that (∂µ/∂P)T = (∂Ḡ/∂P)T
= V̄, so that:
RT d ln a = V̄dP.
Subtract the quantity RT d ln P from both sides to get:
RT d ln(a/P) = V̄dP − RT d ln P
in which the coefficients B, C, D, and so forth are empirical functions of T alone. Assuming isothermal conditions and following the procedure by which we derived
equation 4.1, we now get:
µ = µ0 + RT ln P + B(P − 1) + C(P 2 − 1)/2 + . . . .
This is a clumsy equation to handle, particularly if we
are headed for a generalized form of equation 4.9, and it
is only valid for the empirical equation of state we chose.
The way we avoid this problem is to repackage all the
terms containing P and introduce a new variable, f, so
that:
µ = µ0 + RT ln(f/f 0 ).
pure substance at 1 atmosphere total pressure and at a
specified temperature. (A variety of standard states are
possible. For any given problem, the one we choose is
partly governed by convenience and partly by convention. In chapter 9, we discuss a number of alternate standard states.) For a pure (one-component) gas, f 0 = P 0 =
1 atm. The ratio f/f 0 is given the special name activity,
for which we use the symbol a. The activity of a gas is,
therefore, a dimensionless value, numerically equal to its
fugacity.
We have certainly made the integrated power function
look simpler, but this is an illusion because we have only
redefined variables. We have not improved our understanding of nonideal gases unless we can evaluate fugacity. To do this, we differentiate equation 4.10 with respect
to P at constant temperature:
(4.10)
This new quantity is known as fugacity, and it will
serve in each of the equations as a “corrected” pressure;
that is, as a pressure which has been adjusted for the effects of nonideality. Notice that equation 4.10 is identical
in form to equation 4.1. The reference fugacity, f 0, used
for comparison is usually taken to be the fugacity of the
= (V̄ − [RT/P])dP,
and then integrate the result between the limits zero and
P. This procedure yields the expression:
ln a = ln P +
∫ (V̄/RT − 1/P)dP,
P
0
which is just what we were looking for. The adjustment
for nonideality is expressed as an integral that can be
evaluated by replacing V̄ with an appropriate equation
of state, V̄(P). Notice that if V̄(P) = RT/P (the ideal gas
equation), then the integral has the value zero at any
pressure. The quantity:
γ = exp
[∫ (V̄/RT − 1/P)dP],
P
0
(4.11)
defines the activity coefficient, γ, which is equal to the
ratio a/P.
Activities of species in nonideal gas mixtures are
defined the same way we followed in deriving equation
4.5, except that all partial pressures Pi must now be
How to Handle Solutions
ACTIVITY COEFFICIENTS AND EQUATIONS OF STATE
One good way to illustrate the nonideal behavior of
real gases with increasing pressure is to plot values
of pressure against the experimentally-determined
quantity PV̄/RT, which serves as a measure of compressibility. For an ideal gas, PV̄/RT will, of course,
be equal to 1 at all pressures; values >1 indicate that
a gas is more compressible than an ideal gas; values
>1 indicate that it is less compressible. In figure 4.4,
we show the behavior of molecular hydrogen and
oxygen on this type of plot.
At low pressures, most gases occupy less volume
than we would expect from the ideal gas equation of
state, suggesting that attractive forces between gas
molecules reduce the effective mean distance between
them. In 1879, the Dutch physicist Van der Waals
recognized this phenomenon and adjusted the equation of state by adding a term a/V̄ 2 to the observed
pressure, where a is an empirically determined constant for the gas. This modification correctly describes
the deviation shown in figure 4.4 at very low pressures, although research has shown that a is not a
constant, in fact, but depends on both pressure and
temperature.
This correction for intermolecular attractions,
however, predicts that real gases will become even
more compressible as pressure is increased. This is
clearly not the case. Consequently, the full form of
Van der Waals’ equation includes a second empirical
adjustment on the molar volume:
(P + [a/V̄ 2])(V̄ − b) = RT.
As pressure increases, an increasingly significant volume in the gas is occupied by the molecules themselves. The factor b, therefore, can be thought of as a
measure of the excluded volume that already contains
molecules. The remaining volume, in which molecules
are free to move, is much less than the total volume,
and pressure is higher than it would be for an ideal
gas. The full equation, therefore, reflects a balance between attractive and repulsive intermolecular forces
that affect the molar volume of a real gas.
To calculate an activity coefficient using the Van
der Waals equation of state, we need to rearrange it
in terms of V̄:
V̄ 3 − V̄ 2(b + RT/P) + V̄(a/P) − ab/P = 0.
FIG. 4.4. Total pressure versus PV¯/RT for H2 and O2. A gas for
which PV¯/RT < 1 is more compressible than an ideal gas; if
PV¯/RT > 1, the gas is less compressible.
This cubic equation can be shown to have three real
roots at temperatures and pressures below the critical
point, of which the largest root is the true molar volume for the gas. (We do not define the critical point
until chapter 10. For now, think of it as a condition
in temperature and pressure beyond which there is no
physical distinction between liquid and vapor states.
Virtually all near-surface environments are below the
critical point.) It is this solution that should replace V̄
in equation 4.11. In practice, equation 4.11 is solved
numerically by computing molar volumes iteratively
over the range of pressures from zero to P as part of
the algorithm that solves the integral. This is a reasonably straightforward and reliable procedure at moderate temperatures and pressures.
At the more elevated pressures at which many
igneous or metamorphic reactions may involve gas
phases, the Van der Waals equation of state is less
67
68
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
successful. Redlich and Kwong (1949) found that they
could only predict the nonideal behavior of gases at
very high pressures by modifying the attractive force
term. Their equation,
P = RT/(V̄ − b) − a/([V̄ + b]V√T
),
has been widely used in petrology, and has been
modified further in many studies. Good reviews of
these modifications, and of the use of RedlichKwong type equations of state in calculating activity
coefficients, can be found in Holloway (1977) and
Kerrick and Jacobs (1981).
multiplied by appropriate activity coefficients γi. The resulting equilibrium constant is given by:
Keq = e−∆Ḡ*/RT = [a1ν1 a 2ν2 . . . a jνj /
νj+1 . . . a νi−1 νi
ν1 ν2
νj
(aj+1
i−1 ai )][P1 P2 . . . Pj /
ν
ν
ν
ν
ν
(Pj+1j+1 . . . Pi−1i−1Pi i )][γ1 1 γ 2 2 . . . γ jνj /
νj+1
νi−1 νi
(γ j+1
. . . γ i−1
γ i )].
(4.12)
In a similar way, nonideal liquid or crystalline solutions can be treated by modifying equation 4.9. As with
gases, activity is defined as the ratio f/f 0. Most species in
liquid or solid solution, however, have negligible vapor
pressures at 1 atmosphere total pressure, so that f 0 is
rarely equal to one. Furthermore, because fugacities in
solution are so small, it is impractical to use the fugacity
ratio to measure quantities like the activity of Fe2SiO4 in
a magma. Instead, it is common to write:
Keq = e−∆Ḡ*/RT = [a1ν1 a 2ν2 . . . ajνj /
νj+1 . . . a νi−1 a νi )][X ν1X ν2 . . . Xνj /
(aj+1
2
j
1
i−1 i
νj+1 . . . X νi−1 X νi )][γ ν1 γ ν2 . . . γ νj /
(X j+1
j
i
1 2
i−1
νj+1 . . . γ νi−1 γ νi )].
(γ j+1
i−1 i
(4.13)
in which we declare that ai = γi Xi. In practice, it is common to specify, further, that ai is equal to 1 for a pure
substance at any chosen temperature. More specifically,
ai → Xi as Xi → 1.
Following this convention and applying equation 4.13,
the activity of MgSiO3, for example, would be equal to
1 in pure enstatite (XMgSiO = 1). Other conventions are
3
also common, however. Aqueous geochemists are more
comfortable with compositions expressed in molal units
(moles per kilogram of solution), rather than in mole
fractions. Therefore, in very dilute aqueous solutions,
we usually define the activity of solute species by the relation ai = γi mi , in which mi is the molality of species i.
This produces an equation like equation 4.13 in which
FIG. 4.5. Schematic representation of the various contributions
to the free energy of a nonideal solution.
the Xi quantities are replaced by mi . We generally apply
this expression by specifying a standard state in which
mi = 1 and then extrapolating to infinite dilution, where:
ai → mi as mi → 0.
We discuss this standard state more fully in chapter 9.
According to either of these conventions and others
that we have not mentioned, solutions approach ideal behavior (γi → 1) as their compositions approach a singlecomponent end member. The greatest difficulty arises
when we have to consider nonideal solutions that are
far from a pure end-member composition. We consider
that problem for nonelectrolyte solutions and examine
standard states for activity more fully in chapter 9. Figure 4.5 summarizes graphically the various contributions
to ∆Gr we have discussed so far.
ACTIVITY IN ELECTROLYTE SOLUTIONS
Species in aqueous solutions such as seawater or hydrothermal fluids present an unusual challenge. Although
natural fluids commonly contain some associated (that
is, uncharged) species, it is much more common to find
How to Handle Solutions
free ions. In a typical problem, for example, we may be
asked to evaluate the solubility of sodium sulfate in seawater. Ignoring the fact that crystalline sodium sulfate
usually takes the form of a hydrate, Na2SO4⋅H2O, the
relevant chemical reaction is:
→ 2Na+ + SO 2−,
Na2SO4 ←
4
which we always write with the solid phase as a reactant
and fully dissociated ions as products. For this type of
reaction, we refer to the equilibrium constant as an equilibrium solubility product constant, Ksp:
Ksp = (aNa+)2aSO42 /aNa2SO4.
(The activity of Na2SO4 is unity, because it is a pure solid
phase.)
In general, ionic species in solution do not behave
ideally unless they are very dilute, because ions tend to
interact electrostatically. They also generally associate
with water molecules to produce “hydration spheres” in
which the chemical potential of H2O is different from
that in pure water. (Look again at figure 2.17 and accompanying discussion.) The result is that the free energies
of both the solvent (H2O) and the solute (dissolved ions)
differ from their standard states, and their activities differ from their concentrations.
Unfortunately, there is no way to measure the activity
of a dissolved ion independently. Because charge balance
must always be maintained, we cannot vary the concentration of a cation such as Na+ without also adjusting
the anions in solution. It is impossible, for example, to
determine how much of the potential free energy change
during evaporation of seawater is due to increasing aNa+
and how much is the result of parallel increases in aCl−,
aSO42−, and other anion activities.
The Mean Salt Method
Several approaches have been taken to solve this problem.
One is to define the mean ionic activity, given by:
a± = (a+ν+ a−ν−)1/ν,
in which a+ and a− are the individual activities of cations
and anions, respectively, and the stoichiometric coefficients ν = ν+ + ν − count the number of ions formed per
dissociated molecule of solute. For example, the dissociation reaction for MgCl2 is:
→ Mg2+ + 2Cl−,
MgCl2 ←
69
so ν+ = 1, ν− = 2, and ν = 3. The mean ionic activity of
MgCl2 in aqueous solution is therefore equal to:
a± (MgCl2) = (aMg 2+ aCl−2)1/3.
Mean activities, unlike the activities of charged ions, are
readily measurable. From these, it is possible to calculate
mean ion activity coefficients by using the relationship:
γ± = a± /m±.
The quantity m± is the mean ionic molality, which is related to the molal concentration of total (nondissociated
solute), m, by:
m± = m(ν+ν+ ν−ν− )1/ν.
(4.14)
Again, for MgCl2 in aqueous solution,
m±(MgCl2) = mMgCl2([1]1[2]2)1/3.
It is customary to calculate individual ion activity coefficients, γ+ and γ −, by referring to a standard univalent
electrolyte in a solution of the same effective concentration as the one we are studying. Potassium chloride is a
common standard electrolyte for this purpose because
it has been determined experimentally that γK+ = γCl−, so
that:
γ± (KCl) = (γK+ γCl −)1/2 = γ K+ = γCl − .
If we wanted to calculate an individual ion activity coefficient for Mg 2+ by this mean salt method, then, we
would assume that the value of γCl − in an MgCl2 solution
is the same as the value of γ ± (KCl) that we measured in
a KCl solution. In other words,
γ ± (MgCl2) = (γMg 2+[γCl −]2)1/3 = (γMg 2+ γ± (KCl)2)1/3,
and, therefore:
γMg 2+ = γ± (MgCl2)3/γ± (KCl)2.
Geochemists use the mean salt method to estimate activity coefficients for anions as well as cations. Because
the basis for the method is an empirical comparison to
a well studied electrolyte such as KCl, and because the
comparison is always done between solutions with the
same effective concentration, the method is reliable over
a very wide range of conditions.
The Debye-Hückel Method
Because the mean salt method relies on the use of
γ± from simple electrolyte solutions, you need access to
70
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
measurements on a large number of electrolytes over a
wide range of concentrations. Fortunately, there are
extensive tables and graphs of experimentally derived
activity coefficients in the chemical literature (see the
highlighted discussion of nonideality in natural waters
for one example). For those who prefer to generate their
own values, an alternative nonempirical method for
calculating ion activity coefficients in dilute solutions is
provided by Debye-Hückel theory, which attempts to calculate the effect of electrostatic interactions among ions
on their free energies of formation. The disadvantage of
the theory is that it is only reliable in dilute solutions.
Nevertheless, because very many solutions of geologic
interest fall within its range of reliability, geochemists
find the Debye-Hückel method extremely useful.
To evaluate the cumulative effect of attractive and repulsive forces, we define first a charge-weighted function
of species concentration known as ionic strength (I):
I = 1–2
Σm z .
2
i i
In this way, we recognize that ions in solution are influenced by the total electrostatic field around them and
that polyvalent ions (|z| > 1) exert more electrostatic
force on their neighbors that monovalent ions. A 1 m
solution of MgSO4, for example, has an ionic strength of
4, whereas a 1 m solution of NaCl has an ionic strength
of only 1.
In its most commonly used form, the Debye-Hückel
equation takes into account not only the ionic strength
of the electrolyte solution, but also the effective size of
the hydrated ion of interest, thus estimating the number
of neighboring ions with which it can come into contact.
Values of the size parameter, å, for representative ionic
species, are given in table 4.1, along with values of empirical parameters A and B, which are functions only of
pressure and temperature. We calculate the activity coefficient for ion i from:
/(1 + B å √I
).
log10 γi = −Azi2 √I
Pytkowicz (1983) has discussed the theoretical justification for this equation and a critique of its use in aqueous
geochemistry.
Worked Problem 4.5
What is the activity coefficient for Ca2+ in a 0.05 m aqueous
solution of CaCl2 at 25°C? How closely do answers obtained
by the mean salt method and the Debye-Hückel equation agree?
To calculate γCa2+ by the Debye-Hückel method, we first determine the ionic strength of the solution:
I = 1–2
Σm z
2
i i
= 1–2 ([0.05][2]2 + [0.1][−1]2) = 0.15.
Then, using the data in table 4.1, we find that:
log10 γCa 2+ = −Az i2 √I
/(1 + B å √I
)
= −(0.5085)(2)2
√
(0.15)/[1 + (0.3281 × 108)
(6 × 10−8)
√
(0.15)]
= −0.447.
So,
γCa2+ = 0.357.
How does this compare with γCa 2+ calculated by the mean
salt method? The mean ionic activity coefficient γ±CaCl in 0.05 m
2
TABLE 4.1. Values of Constants for Use in the Debye-Hückel Equation
Temperature
°C
A
B
å × 108
0
5
10
15
0.4883
0.4921
0.4960
0.5000
0.3241
0.3249
0.3258
0.3262
2.5
3.0
3.5
4.0–4.5
20
25
30
0.5042
0.5085
0.5130
0.3273
0.3281
0.3290
4.5
5.0
6
35
40
45
50
55
60
0.5175
0.5221
0.5271
0.5319
0.5371
0.5425
0.3297
0.3305
0.3314
0.3321
0.3329
0.3338
8
9
11
Data from Garrels and Christ (1982).
Ion
Rb+,
Cs+,
NH4+,
− −
Tl+,
Ag+
K+, CL−, Br , I , NO3−
OH−, F−, HS−, BrO3−, IO4−, MnO4−
Na+, HCO3−, H2PO4−, HSO3−, SO42−,
SeO42−, CrO42−, HPO42−, PO43−
Pb2+, CO32−, SO32−, MoO42−
Sr2+, Ba2+, Ra2+, Cd2+, Hg 2+, S2−, WO42−
Li+, Ca2+, Cu2+, Zn2+, Sn2+, Mn2+, Fe2+,
Ni2+, Co2+
Mg2+, Be2+
H+, Al3+, Cr3+, trivalent rare earths
Th4+, Zr4+, Ce4+, Sn4+
How to Handle Solutions
71
Therefore,
γCa2+= (γ±CaCl 2)3/[γ±KCl]2) = (0.577)3/(0.649)2 = 0.347.
This is very close to the 0.357 value calculated by DebyeHückel theory. In figure 4.7, we have repeated this pair of calculations for a range of ionic strengths from 0.05 to 3.0. It is
apparent that the methods agree at ionic strengths less than
∼0.3, but diverge strongly in more concentrated solutions, as
Debye-Hückel theory fails to model changes in the structure
of the solute adequately. Other calculation schemes have been
used to model highly concentrated solutions, particularly at
elevated temperatures (see, for example, Harvie et al. [1984], or
the Helgeson references in appendix B), but most researchers
continue to use values obtained by some variant of the mean
salt method in those situations.
FIG. 4.6. Activity coefficients for common aqueous species as a
function of ionic strength.
SOLUBILITY
aqueous solution has been reported by Goldberg and Nuttall
(1978) to be 0.5773. To use the mean salt method, we need to
know the mean ion activity coefficient for KCl at the same effective concentration as the CaCl2 solution; that is, at the same
ionic strength. A KCl solution of ionic strength 0.15 is 0.15
molal. Robinson and Stokes (1949) report that γ±KCl = 0.744 in
0.15 m aqueous KCl. You can verify this, although with less
precision, from the graph in figure 4.6. From this value, we
calculate that:
γ±CaCl 2 = (γCa2+ [γCl−]2)1/3 = (γCa2+ [γ± KCl]2)1/3.
Many of the interesting questions facing aqueous geochemists have to do with mineral solubility. How much
sulfate is in groundwater that is in equilibrium with
gypsum beds? How much fluoride can be carried by an
ore-forming hydrothermal fluid? What sequence of
minerals should be expected to form during evaporation
of seawater? What processes can cause deposition of ore
minerals from migrating fluids? How are the compositions of residual (zonal) soils affected by the chemistry
of soil water? How might a chemical leak from a waste
disposal site affect the stability of surrounding rocks and
soils? To address these questions and others like them in
coming chapters, we must first look at broad conditions
that affect solubility.
In worked problems 4.6–4.10, we will make successive refinements in our approach to a typical problem,
illustrating by stages the major thermodynamic factors
influencing solubility.
Worked Problem 4.6
What is the solubility of barite in water at 25°C? The simplest
answer can be calculated from data for the molar free energies
of formation of species in the reaction:
→ Ba2+ + SO 2−.
BaSO4 ←
4
We find the following values in Parker et al. (1971):
FIG. 4.7. Ion activity coefficient for Ca2+ as a function of ionic
strength, calculated from the Debye-Hückel equation and by the
mean salt method. Notice the deviations at higher ionic strengths.
∆Ḡf (kcal mol−1)
BaSO4
Ba2+
SO42−
−325.6
−134.02
−177.97
72
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
NONIDEALITY IN NATURAL WATERS
The water in streams and the aquifers that we typically draw on for irrigation and public water supplies
is in fact a dilute electrolyte solution. People and plants
are sensitive to small variations in composition. Some
electrolytes affect cosmetic qualities of water such as
taste and smell, and others can have profound effects
on health. Inorganic solutes can also interact with
pipes and machinery, producing corrosion or scale.
For this reason, the composition of natural water in
most parts of the world is monitored quite closely.
As we discuss in greater detail in chapters 5 and 7,
the composition of natural waters is controlled largely
by weathering reactions. The dominant cations in
surface and ground waters (Ca2+, Mg2+, Na+, and K+)
and the dominant anions (HCO3−, Cl−, and SO42−) are
derived from silicate and carbonate minerals and the
oxidation of metal sulfides. Hydrogeologists commonly trace the chemical history of water from an
aquifer or reservoir by studying the relative abundances of these ions, expressed as concentration ratios in either molalities (moles/kg) or equivalencies
(moles × charge/liter). Using diagrams such as those
in figure 4.8, for example, we can distinguish between
waters that have equilibrated with limestone (relatively Ca2+, Mg2+, and HCO3−rich) and those that are
more likely to have equilibrated with shale (relatively
Na+, and K+ rich). These are useful qualitative indicators of where a water sample has been.
The more quantitative thermodynamic and kinetic
analyses that we introduce in chapters 5 and 7 will
FIG. 4.8. Classification diagram for anion and cation facies in terms of major ion percentages. Groundwater types are designated according to the domain in which they occur on diagram segments. (After
Freeze and Cherry 1979.)
How to Handle Solutions
require ion activities, rather than concentrations. You
can use either the mean salt method or the DebyeHückel method to calculate ion activity coefficients.
To decide which method is most likely to give a reliable result for a given water sample, calculate its ionic
strength. Except when some other ion is present in
significant amounts, this is equal to:
I = 21– [mNa+ + mK+ + 4mCa2+ + 4mMg2+ + mHCO3−
+ mCl− + 4mSO2−
].
4
73
As table 4.2 suggests, the ionic strengths of natural
waters range from about 1.0 × 10−3 in rivers and
lakes to as much as 1 × 10−1 in groundwater. Oilfield
brines and saline lakes have significantly higher ionic
strengths. The ionic strength of seawater is 7 × 10−1.
Because the Debye-Hückel method is unreliable
above ionic strengths of 1.0 × 10−1 (see figure 4.7),
therefore, you should use the mean salt method for
any solution more saline than groundwater.
TABLE 4.2. Analyses of Groundwater and River Waters
Source
Ion
+
K
Na+
Ca2+
Mg2+
HCO3−
Cl−
SO42−
I
1
2
3
4
5
6
7
1.0
7.9
56.0
12.0
160.0
12.0
53.0
0.0065
9.0
37.0
60.0
60.0
417.0
27.0
96.0
0.0146
0.6
2.0
4.8
1.5
24.0
0.6
1.1
0.0006
0.2
0.3
2.1
0.1
4.9
—1
2.1
0.0002
0.2
0.9
1.7
0.4
1.6
0.5
6.2
0.0003
1.5
6.5
20.1
4.9
71.4
7.0
14.9
0.0026
1.6
6.6
16.6
4.3
66.2
7.6
9.7
0.0022
Values in mg/l. Sources are: 1, limestone aquifer, Florida (Goff 1971); 2, dolomite aquifer,
Manitoba (Langmuir 1971); 3, granitic glacial sand aquifer, Northwest Ontario (Bottomley
1974); 4, streams draining greenstone, South Cascade Mountains, Oregon (Reynolds and Johnson 1972); 5, streams draining granite, metamorphics, and glacial till, Hubbard Brook watershed, New Hampshire (Likens et al. 1977), analysis also contains trace amounts of NO3−; 6, average river waters, corrected for pollution, North America (Meybeck 1979); 7, average river
waters, corrected for pollution, Asia (Meybeck 1979).
1
No measurement.
The standard molar free energy of reaction is calculated by:
∆Ḡr = −134.02 + (−177.97) − (−325.6)
= 13.61 kcal mol−1.
If the fluid is saturated with respect to BaSO4, then the reaction
above is perfectly balanced and ∆Ḡr = 0. Therefore,
log Ksp = − ∆Ḡr0/RT = −13.61/([298][0.001987])
= −9.98,
Ksp = 1.04 × 10−10.
If we assume that the solution behaves ideally, then the solubility of barite can be calculated by recognizing that charge
balance in the fluid requires that m Ba2+ = mSO42−, and that:
m Ba2+mSO42− = Ksp.
From this, we get:
m Ba2+ = mSO42− = 1.02 × 10−5 mol kg−1.
The result in worked problem 4.6 is quite close to the
value of 1.06 × 10−5 mol kg−1 measured in experiments
by Blount (1977), even though we assumed the fluid was
ideal. This agreement may be fortuitous, however, because we should anticipate some nonideal behavior. Let’s
do the problem again.
Worked Problem 4.7
What are the activity coefficients for Ba2+ and SO42− in a solution saturated with respect to barite, and how much does the
calculated solubility change if we allow for nonideal behavior?
The easiest way to solve this problem is to design a numerical algorithm and iterate toward a solution by computer. In
outline form, the procedure looks like this:
1. Using the ideal solubility as a first guess, calculate the ionic
strength of a saturated solution.
74
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
2. Calculate activity coefficients for Ba2+ and SO42− from the
Debye-Hückel equation, using constants in table 4.1.
3. Using the activity coefficients and the value of K derived from
∆Ḡr0, calculate improved values of mBa2+ and mSO42−.
4. Repeat steps 1 through 3 until mBa2+ and mSO42− do not change
in successive iterations by more than some acceptably small
amount.
When we did this calculation, it took four cycles to satisfy
the condition |mBa2+ (old) − mBa2+ (new)| < 10−9, which is a calculation error of less than one part in 104 (much better than is
justified by the uncertainties in ∆Ḡr0). Notice that charge balance
constraints require that mBa2+ = mSO42−.
Iteration
1
2
3
4
I
10−5
2.41 ×
4.18 × 10−5
4.21 × 10−5
4.21 × 10−5
γBa2+
γSO 2−
γ±
mBa2+
0.9775
0.9705
0.9704
0.9704
0.9774
0.9704
0.9703
0.9703
0.9774
0.9705
0.9704
0.9704
1.0443 × 10−5
1.0519 × 10−5
1.0520 × 10−5
1.0520 × 10−5
4
As we suspected in worked problem 4.6, a solution saturated
with BaSO4 is nearly ideal. By taking the slight nonideality into
account, however, the computed solubility increased to within
1% error of the value measured by Blount (1977).
The Ionic Strength Effect
The reason that there is progressive improvement
with each iteration in worked problem 4.7 is that each
cycle calculates the ionic strength with an improved set
of activity coefficients. Suppose, however, that there were
other ions in solution as well as Ba2+ and SO42−. Natural
fluids, in fact, rarely contain only a single electrolyte in
solution. Most hydrothermal fluids or formation waters
are dominated by NaCl. To calculate the ionic strength
correctly, we must consider not only the amount of Ba2+
and SO42− but also include the concentrations of all other
ions in solution. This, in turn, leads to an adjustment in
the solubility of BaSO4.
Worked Problem 4.8
What effect do other solutes have on the solubility of barite? In
particular, what would be the solubility of barite in a solution
at 25°C that is already 0.2 m in NaCl? Assuming that there is
little tendency for Ba2+ to associate with Cl− (Sucha et al. 1975)
or for Na+ to associate with SO42− (Elgquist and Wedborg
1978), we should expect that barite solubility changes only as
the result of increased ionic strength. To verify this, we repeat
the calculation of worked problem 4.7, this time calculating
ionic strength in each iteration by I = 1–2 (mBa2+ × 4 + mSO42− × 4 +
mNa+ + mCl−).
I
0.2000
γBa2+
γSO42−
γ±
mBa2+
0.2986
0.2671
0.2824
3.615 × 10−5
These values, once again, compare favorably with Blount (1977),
who measured barite solubility in 0.2-m NaCl-H2O at 25°C
and found mBa2+ = 3.7 × 10−5 mol kg−1 and γ± = 0.2754.
The Common Ion Effect
In many natural fluids, the solubility of minerals like
barite increases with ionic strength, as we saw in worked
problem 4.8. This is not the case, however, if ions produced by dissociation of the solute are also present from
some other source in solution. For example, if barite is
in equilibrium with a solution that contains both NaCl
and Na2SO4, the charge balance constraint on the fluid
becomes:
2mBa2+ + mNa+ = 2mSO42− + mCl−.
Sulfate is a common ion in BaSO4 and Na2SO4. Extra
SO42− ions in solution, whether they come from Na2SO4
or some other source, force the reaction:
→ Ba2+ + SO 2−
BaSO4 ←
4
toward the left (remember equation 4.3?), decreasing the
solubility of barite. To demonstrate this common ion effect, we offer the following problem.
Worked Problem 4.9
How does the addition of successive amounts of Na2SO4 to an
aqueous fluid affect the solubility of barite?
The molal concentration of Ba2+ can be calculated from this
expression and the equilibrium constant, K:
K = mBa2+ mSO42− γBa2+ γSO 42−
= mBa2+ γBa2+ γSO 42−(2mBa2+ + mNa+ − mCl−)/2
mBa2+ = 2K/(γBa2+ γSO 42−[2mBa2+ + mNa+ − mCl−]).
This equation can be solved by the same numerical procedure
we applied in worked problems 4.7 and 4.8. We can simplify
the calculation in this case, because the solubility of barite has
been shown to be quite low. We can assume that all SO42− in solution is due to dissociation of Na2SO4, as long as we examine
only solutions of moderately high mSO 42−. The charge balance
constraint, in other words, is approximately:
mNa+ = 2mSO 42− + mCl−.
The ionic strength calculated with this simplification will differ
only very slightly from the “true” ionic strength.
How to Handle Solutions
To illustrate the effect of a common ion (SO42− in this case),
we have calculated mBa2+ in a variety of NaCl-Na2SO4-H2O
solutions with ionic strength equal to 0.2.
mBa2+
3.91 × 10−6
1.31 × 10−6
6.53 × 10−7
1.96 × 10−7
9.80 × 10−8
6.53 × 10−8
mSO 42−
mNa+
mCl −
3.33 × 10−4
1.00 × 10−3
2.00 × 10−3
6.67 × 10−3
1.33 × 10−2
2.00 × 10−2
2.00 × 10−1
1.99 × 10−1
1.98 × 10−1
1.93 × 10−1
1.86 × 10−1
1.80 × 10−1
1.99 × 10−1
1.97 × 10−1
1.94 × 10−1
1.80 × 10−1
1.60 × 10−1
1.40 × 10−1
As expected, the addition of sulfate, even in very small amounts,
dramatically lowers the solubility. Barite will precipitate if the
product mBa2+ mSO 42− equals or exceeds Keq (1.31 × 10−9 at ionic
strength 0.2). With increasing amounts of sulfate in solution,
this condition is met with vanishingly small amounts of dissolved Ba2+. The same effect would be observed if we were to
provide barium ions from some source other than barite (for
example, from barium chloride).
Complex Species
In stream waters (ionic strength < 0.01 on average)
and in most surficial environments on the continents, dissolved species are highly dissociated. As ionic strength
increases, however, electrical interactions between ions
also increase. The result is that many ions in concentrated solutions exist in groups or clusters known as
complexes. For ease of discussion, these may be divided
into three broad classes: ion pairs, coordination complexes, and chelates. By convention (not universally
observed, unfortunately), the stability of any complex is
expressed by a stability constant, Kstab, which is written
for the reaction forming a complex species from simpler
dissolved species. Thus, for example, Mg2+ and SO42− ions
can associate to form the neutral ion pair MgSO40 by the
reaction
→ MgSO 0,
Mg2+ + SO42− ←
4
for which Kstab = aMgSO 0 /(aMg2+ aSO42−) = 5.62 × 10−3 at
4
25°C. The larger a stability constant, the more stable the
complex it describes.
The distinctions among different types of complexes
are not easy to define and, for thermodynamic purposes,
not very important. Ion pairs are characterized by weak
bonding; hence, small stability constants. The dominant
complex species in seawater and most brines fall into
this category. In chapter 8, we examine the effect of ion
pair formation on the chemistry of the oceans.
75
Coordination complexes involve a more rigid structure of ligands (usually anions or neutral species) around
a central atom. This well-defined structure contributes
to greater stability. Transition metals, such as copper,
vanadium, and uranium have been shown to exist in
aqueous solution primarily as coordination complexes
like Cu(H2O)62+, in which the negatively charged ends of
several polar water molecules are attracted to the metal
cation to form a “hydration sphere” like the one we illustrated in figure 2.17. Other polar molecules also commonly form coordination complexes with metal cations.
One that may be familiar to you from analytical chemistry is ammonia, which combines readily with Cu2+ in
solution to form the bright blue complex Cu(NH3)42+.
The stability constant for this species is 2.13 × 1014.
Molecules like water, ammonia, Cl−, and OH−, which
form simple coordination complexes because they can
only attach to one metal cation at a time, are called
unidentate ligands. Other species, usually organic molecules or ions, are called multidentate or polydentate
ligands. Chelates, which form around these, have larger,
more complicated structures that include several metal
cations. Chelates differ from coordination complexes
only in their degree of structure. Some, like the hexadentate anion of ethylenediaminetetraacetic acid (mercifully
known as EDTA) are widely used by analytical chemists
to scavenge metals from very dilute aqueous solutions.
In nature, the vast number of humic and fulvic acids
and other dissolved organic species also serve as chelating
agents. In general, the larger and more complicated these
are, the more stable the complexes they form. Because
they have very high stability constants and can coordinate
several cations at once, chelating agents can contribute
significantly to the chemistry of transition metals and of
cations such as Al3+ in natural waters, even if both they
and the metal cations are in very low concentration.
The formation of complex species increases the solubility of electrolytes by reducing the effective concentration of free ions in solution. In a sense, we can think of
ions bound up in complexes as having become electrostatically shielded from oppositely charged ions and
therefore unable to combine with them. Because the
activity product of free ions is lowered, equilibria that
involve those ions favor the further dissolution of solid
material. Thus, although the activity product of free ions
at saturation will still be equal to Ksp, the total concentration of species in solution will be much higher than in
a solution without complexing.
76
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Worked Problem 4.10
Calcium and sulfate ions can associate in aqueous solution to
form a neutral ion pair, CaSO40. This is one of several ion pairs
that affect the concentration of species in seawater, as we shall
see in chapter 8. How does it affect the solubility of gypsum?
The solubility product constant, Ksp, for gypsum is calculated from the free energy of the reaction:
→ Ca2+ + SO 2− + 2 H O.
CaSO4 ⋅2H 2O ←
4
2
In reasonably dilute solutions, the activity of water can be assumed to be unity. Gypsum, a pure solid, is in its standard state
and also has unit activity. Ksp, then, is equal to aCa2+ aSO 42−. At
25°C, it has the value 2.45 × 10−5.
The stability constant, Kstab, for CaSO40 is derived by experiment for the reaction:
Ca2+
+
SO42−
→
←
CaSO40.
Kstab = aCaSO 40 /(aCa2+ aSO 42−) = 2.09 × 102
at 25°C. By combining the two expressions for Ksp and Kstab
and rearranging terms, we can calculate:
aCaSO 40 = Kstab Ksp = 5.09 × 10−3.
Because CaSO40 is a neutral species, its concentration in
solution will be nearly the same as its activity. The concentration of the ion pair, therefore, is ∼5 mmol kg−1. It can be
shown, using the iterative procedure in worked problem 4.7,
that the concentration of Ca2+ in the absence of complexing
is ∼10 mmol kg−1. The total calcium concentration in solution
is thus ∼15 mmol kg−1, or ∼50% higher than the concentration
of Ca2+ alone.
SUMMARY
We see now that the structure and stoichiometry of solutions play a key role in determining how they act in
chemical reactions. The free energy of any reaction can
be expressed in terms of a standard state free energy and
a contribution due to mixing of end-member species in
solutions. Relatively few solutions of geochemical interest
are ideal; in most, interactions among dissolved species
or among atoms on different structural sites produce
significant departures from ideality. In many cases, however, we gain valuable first impressions of a system by
assuming ideal mixing and then adding successive corrections to describe its actual behavior.
We discuss solution models more fully in chapters 9
and 10, where, in particular, we address the problem
of calculating activity coefficients for nonideal systems of
petrologic interest. In the next few chapters, however,
we make use of the Debye-Hückel equation and the mean
salt method, which yield the activity coefficients we need
to explore dilute aqueous solutions. Solubility equilibria
will dominate our discussion of the oceans, of chemical
weathering, and of diagenesis.
suggested readings
Several excellent textbooks are available for the student interested in the thermodynamic treatment of solutions. The ones
listed here are a selection of some that the authors have found
particularly useful.
Drever, J. I. 1997. The Geochemistry of Natural Waters. New
York: Simon and Schuster. (This is a very good book for
the student who is new to aqueous geochemistry. There are
quite a few carefully written case studies to show the use of
geochemistry in the real world.)
Garrels, R. M., and C. L. Christ. 1982. Solutions, Minerals,
and Equilibria. San Francisco: Freeman, Cooper. (Perhaps the
most often used text in its field.)
Powell, R. 1978. Equilibrium Thermodynamics in Petrology.
London: Harper and Row. (A good text for classroom use.
Chapters 4 and 5 consider the calculation of mole fractions
in crystalline and liquid solutions.)
Pytkowicz, R. M. 1983. Equilibria, Nonequilibria, and Natural
Waters, vol. I. New York: Wiley. (An exhaustive treatment
of the theoretical basis for the thermodynamics of aqueous
solutions. Chapter 5 has more than 100 pages devoted to
the determination of activity coefficients.)
Robinson, R. A., and R. H. Stokes. 1968. Electrolyte Solutions.
London: Butterworths. (A reference for chemists rather
than geologists, but widely quoted by aqueous geochemists.
Highly theoretical, but not difficult to follow.)
Stumm, W., and J. J. Morgan. 1995. Aquatic Chemistry. New
York: Wiley. (A detailed, yet highly readable text with an
emphasis on the chemical behavior of natural waters. Chapter 6, dealing with complexes, is particularly well written.)
The following articles were referenced in this chapter. An interested student may wish to explore them more thoroughly.
Blount, C. W. 1977. Barite solubilities and thermodynamic quantities up to 300°C and 1400 bars. American Mineralogist
62:942–957.
Bottinga, J., and D. F. Weill. 1972. The viscosity of magmatic
silicate liquids: a model for calculation. American Journal of
Science 272:438–475.
Bottomley, D. 1974. Influence of Hydrology and Weathering on
the Water Chemistry of a Small Precambrian Shield Watershed. Unpublished MSc. Thesis, University of Waterloo.
How to Handle Solutions
Cameron, M., and J. J. Papike. 1980. Crystal chemistry of silicate pyroxenes. Mineralogical Society of America Reviews
in Mineralogy 7:5–92.
Cameron, M., and J. J. Papike. 1981. Structural and chemical
variations in pyroxenes. American Mineralogist 66:1–50.
Clark, J., D. E. Appleman, and J. J. Papike. 1969. Crystalchemical characterization of clinopyroxenes based on eight
new structure refinements. Mineralogical Society of America
Special Paper 2:31–50.
Cohen, R. E. 1986. Thermodynamic solution properties of
aluminous clinopyroxenes: Non-linear least-squares refinements. Geochimica et Cosmochimica Acta 50:563–576.
Elgquist, B., and M. Wedborg. 1978. Stability constants of
NaSO4−, MgSO4 , MgF+, MgCl+ ion pairs at the ionic strength
of seawater by potentiometry. Marine Chemistry 6:243–252.
Freeze, R. A., and J. A. Cherry. 1979. Groundwater. Englewood
Cliffs: Prentice-Hall.
Goff, K.J. 1971. Hydrology and Chemistry of the Shoal Lakes
Basin, Interlake Area, Manitoba. Unpublished Master’s
thesis, University of Manitoba.
Goldberg, R. N., and R. L. Nuttall. 1978. Evaluated activity
and osmotic coefficients for aqueous solutions: The alkaline
earth halides. Journal of Physical and Chemical Reference
Data 7:263.
Harvie, C. E., N. Moller, and J. H. Weare. 1984. The prediction
of mineral solubilities in natural waters: The Na-K-Mg-CaH-Cl-SO4-OH-HCO3-CO3-CO2-H2O system to high ionic
strengths at 25°C. Geochimica et Cosmochimica Acta 48:
723–752.
Helgeson, H. C. 1969. Thermodynamics of hydrothermal systems at elevated temperatures and pressures. American Journal of Science 267:729–804.
Holloway, J. R. 1977. Fugacity and activity of molecular species
in supercritical fluids. In D. G. Fraser, ed. Thermodynamics
in Geology, pp. 161–180. Boston: D. Reidel.
Kerrick, D. M., and G. K. Jacobs. 1981. A modified RedlichKwong equation for H2O, CO2, and H2O-CO2 mixtures at
elevated pressures and temperatures. American Journal of
Science 281:735–767.
Langmuir, D. 1971 The chemistry of some carbonate groundwaters in Central Pennsylvania. Geochimica et Cosmochimica Acta 35:1023–1045.
Likens, G. E., F. H. Bormann, R. S. Pierce, J. S. Eaton, and
N. M. Johnson. 1977. Biogeochemistry of a Forested Ecosystem. New York: Springer-Verlag.
Livingstone, D. A. 1963. Chemical Composition of Rivers and
Lakes. U.S. Geological Survey Professional Paper 440G.
Meybeck, M. 1979. Concentrations des eaux fluviales en éléments majeurs et apports en solution aux océans. Revues
de Géologie Dynamique et de Géographie Physique 23:
215–246.
Parker, V. B., D. D. Wagman, and W. H. Evans. 1971. Selected
Values of Chemical Thermodynamic Properties: Tables for
the Alkaline Earth Elements (Elements 92 through 97 in the
Standard Order of Arrangement). Technical Note 270-6.
Washington, D.C.: U.S. National Bureau of Standards.
Redlich, O., and J.N.S. Kwong. 1949. An equation of state:
Fugacities of gaseous solutions. Chemical Review 44:
233–244.
Reynolds, R. C., and N. M. Johnson. 1972. Chemical weathering in the temperate glacial environment of the Northern
Cascade Mountains. Geochimica et Cosmochimica Acta 36:
537–544.
Robinson, R. A., and R. H. Stokes. 1949. Tables of osmotic
and activity coefficients of electrolytes in aqueous solutions
at 25°C. Transactions of the Faraday Society 45:612.
Sucha, L., J. Cadek, K. Hrabek, and J. Vesely. 1975. The stability of the chloro complexes of magnesium and of the
alkaline earth metals at elevated temperatures. Collected
Czechoslovakian Chemical Communications 40:2020–2024.
PROBLEMS
(4.1)
Use the Debye-Hückel equation to calculate values for the ion activity coefficients for Cl− and Al3+
at 25°C in aqueous solutions with ionic strengths ranging from 0.01 to 0.2. Plot these on a graph of γ
versus log I. What physical differences between these two ions account for the differences in their activity coefficients?
(4.2)
Using the procedure introduced in worked problem 4.5, calculate the solubility of gypsum in water
at 25°C.
(4.3)
To produce a solution that is saturated with respect to fluorite at 25°C, you need to dissolve 6.8 × 10−
g of CaF2 per 0.25 liter of pure water. Assuming that CaF2 is 100% dissociated, and that the
solution behaves ideally, calculate the Ksp for fluorite. What is the free energy for the dissociation
reaction?
3
77
78
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
(4.4)
The Ksp for celestite is 7.6 × 10−7 at 25°C. How many grams of SrSO4 can you dissolve in a liter of
0.1 m Sr(NO3)2? Assume that the activity coefficients for all dissolved species are unity, and that no
complex species are formed in solution.
(4.5)
Repeat the calculation in problem 4.4, again assuming that complex species are absent, but correcting
for the ionic strength effect by calculating activity coefficients with the Debye-Hückel equation.
(4.6)
The following analysis of water from Lake Nipissing in Ontario was reported by Livingstone (1963):
HCO3−
Cl−
Ca2+
Na+
26.2 ppm
1.0
9.0
3.8
SO42−
NO 3−
Mg2+
8.5 ppm
1.33
3.6
What is the ionic strength of this water? (Recall that concentrations in ppm are equal to concentrations in mmol kg−1 multiplied by formula weight.)
(4.7)
Calculate the mole fractions of forsterite and fayalite in an olivine with the composition
Mg0.7 Fe1.3 SiO4.
(4.8)
A garnet has been analyzed by electron microprobe and found to have the following composition:
SiO2
Al2O3
FeO
MnO
MgO
39.08 wt %
22.66
29.98
1.61
6.67
Calculate the mole fractions of almandine, pyrope, and spessartine in this garnet.
(4.9)
Calculate the mole fractions of enstatite, ferrosilite, diopside, CaCrAlSiO6 (Cr-CATS), and
Mg1.5AlSi1.5O6 (pyrope) in the orthopyroxene analysis below. Assume complete mixing on sites in
such a way that all tetrahedral sites are filled by either Si or Al; M1 sites by Al, Cr, Fe3+, Fe2+, or Mg;
and M2 sites by Ca, Fe2+, or Mg. Assume also that the ratio of XFe to XMg is that same in M1 and
M2 sites. (This problem, modified from Powell 1978, requires a fair amount of effort.)
SiO2
Al2O3
Fe2O3
CaO
57.73 wt %
0.95
0.42
0.23
Cr2O3
FeO
MgO
0.46 wt %
3.87
36.73
CHAPTER FIVE
DIAGENESIS
A Study in Kinetics
OVERVIEW
WHAT IS DIAGENESIS?
Diagenesis embraces all of the changes that may take
place in a sediment following deposition, except for
those due to metamorphism or to weathering at the
Earth’s surface. Intellectual battles have been waged
over these two environmental limits to diagenesis.
Rather than join these battles, we focus on some of the
geochemical processes that affect sediments after burial
and consider some of the pathways along which they
may change.
Diagenetic changes take place slowly at low temperature. The assemblages we see in sediment, therefore,
usually represent some transition between stable states.
In previous chapters, we began to study thermodynamic
principles that allow us to interpret those stable states.
This chapter, instead, focuses exclusively on building
our understanding of kinetic methods—the best tools we
have for examining the transition between stable states.
We develop a set of fundamental equations that describe
the transport of chemical species by diffusion and advection, and consider ways to describe dissolution and precipitation reactions from a kinetic perspective. Our goal
is a general diagenetic equation that summarizes the various processes controlling the redistribution of chemical
species in sediments after burial.
Many diagenetic processes are associated with lithification. Among these are compaction, cementation, recrystallization, and growth of new (authigenic) minerals.
Through these mechanisms, unconsolidated sediment
undergoes a general reduction in porosity and develops
a secondary framework that converts it to solid rock.
Other changes also occur, often independent of the progress of lithification. Chief among these are the many
processes that modify buried organic matter, which we
consider in chapter 6.
Each of these diagenetic processes commonly takes
place between ∼20°C and 300°C and occurs close
enough to the surface that the total pressure is less than
∼1 kbar. Almost invariably, diagenetic processes also involve the participation of fluids in the interstitial spaces
between sediment particles. In freshly buried sediments,
this fluid has the same bulk composition as the waters
from which the sediments were deposited. As time proceeds, these formation waters accumulate and transmit
the products of reactions within the sediment column.
The thermodynamic principles we introduced in earlier
chapters are sometimes useful in examining these reactions. In most cases, however, sediments and fluids undergoing diagenesis are highly nonuniform; that is, their
79
80
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
compositions vary from place to place within a distance
of a few centimeters to a few meters. Because of this
variation, it is possible to extract kinetic information by
studying concentration-distance profiles in a diagenetic
environment.
KINETIC FACTORS IN DIAGENESIS
Transport of material can take place in either of two
ways: by diffusion or advection. When we focus on diffusive transport, we are considering the dispersion of
ions or molecules through a medium that is not itself
moving. The path followed by any single particle is random, but a net movement of material may result if there
is a gradient in some intensive property such as temperature or chemical potential across the system. In contrast,
advective transport takes place when the ions or molecules are carried in a medium, like an intergranular fluid,
that is moving. The driving force in this case is simply a
gradient in hydrostatic pressure.
In this section, we develop a general expression to illustrate how the concentration of any mobile species in
a diagenetic environment may change as the result of
diffusive or advective transport, or as the result of chemical reactions between the fluid and solid phases in a sedimentary column.
Diffusion
In chapter 3, we introduced thermodynamic functions
that describe the behavior of systems in bulk, rather than
developing a statistical method for describing systems
at the atomic level. This phenomenological approach
addresses the thermodynamic properties of systems without acknowledging the interactions among atoms, or even
their existence. In the same way, nineteenth-century scientists such as Joseph Fourier and Georg Ohm described
the transport of heat and electrical charge by assuming
that these entities travel through a continuous medium
rather than one consisting of discrete atoms. This assumption leads to a description of transport properties
by differential equations, just as it did when we adopted
this perspective for thermodynamics. Diffusive transport
was first described in this way in the 1850s by Adolf Fick,
a German chemist.
According to Fick’s First Law, the flux (J) of material
along a composition gradient at any specific time, t, is
directly proportional to the magnitude of the gradient.
That is, if the ions or molecules in which we are interested move along a direction, x, parallel to a gradient in
concentration, c, then:
J = −D(∂c/∂x)t .
(5.1)
The quantity D is called the diffusion coefficient, and is
commonly tabulated in units of cm2 sec−1. Equation 5.1
satisfies the empirical observation that the flux drops to
zero in the absence of a concentration gradient. It is
applicable in any system experiencing diffusive transport,
but is most appropriate when the system of interest is in
a steady state; that is, when (∂c/∂t)x is a constant. (Recall
our brief introduction to the idea of a steady state in the
party example of chapter 1.) Equation 5.1 is similar to
Ohm’s law or to fundamental heat flow equations in that
D can be shown experimentally to be independent of
the magnitude of (∂c/∂x)t . Because all of the quantities in
equation 5.1 are positive numbers, we must place a negative sign on the right hand side to indicate that material
diffuses in the direction of decreasing concentration.
If the concentration at a given point in the system is
changing with time, equation 5.1 still holds, but it is not
the most convenient way to describe diffusive transport.
To derive a more appropriate differential equation, consider the schematic sediment column in figure 5.1, in
which we have measured the flux of some species at two
depths x1 and x2, separated by a distance ∆ x, and have
found J1 to be different from J2. So long as ∆x is small, we
can describe the relationship between J1 and J2 by writing:
J1 = J2 − ∆x(∂J/∂x).
Because the fluxes at the two depths differ, the concentration of the species of interest in the sediment volume
between them must change during any time interval we
choose. If we assume a unit cross-sectional area for the
column, then the affected volume is 1 × ∆x and the change
can be written as:
(J1 − J2) = ∆x(∂c/∂t) = −∆x(∂J/∂x).
We can now substitute this result into equation 5.1 to
obtain
( )
∂c
∂
∂c
—– = —– D—– .
∂t
∂x
∂x
(5.2)
This equation is Fick’s Second Law, sometimes referred
to as a continuity equation, because it arises directly
from the need to conserve matter during transport.
Diagenesis: A Study in Kinetics
81
illustrate common situations in diagenetic environments
of interest to geochemists.
Worked Problem 5.1
FIG. 5.1. Schematic sediment column with a unit cross-sectional
area. The flux J1 of some chemical species at depth x1 is greater
than the flux J2 at depth x2. This means that concentration of the
species between x1 and x2 must vary as a function of time.
As they are presented here, Fick’s laws describe diffusion in a straight line and assume that the medium for
diffusion is isotropic. In most diagenetic environments,
where diffusion takes place in pore waters, this formulation is adequate. Lateral variations in the composition of
interstitial waters in a sediment, in most cases, are much
less pronounced than changes with depth, so the driving
force for diffusion is primarily one dimensional. In general, however, diffusion takes place in three dimensions,
(x, y, z), and equation 5.2 is more properly written as:
( )
∂c
∂
∂c
—– = —– D—–
∂t
∂x
∂x
y,z
( )
∂
∂c
+ —– D—–
∂y
∂y
x,z
( )
∂
∂c
+ —– D—–
∂z
∂z
x,y
.
In this chapter, we deal exclusively with the onedimensional form. We also generally assume that the
diffusion coefficient, D, does not vary from place to place.
These simplifications make mathematical modeling much
easier, and in most cases we lose very little accuracy by
adopting them.
If D is not a function of distance, then Fick’s Second
Law can be written more simply as:
( )
∂c
∂2c
—– = D —— .
∂t
∂x2
(5.3)
Solutions for this equation in a nonsteady state system
(the most general situation) involve determining concentration as a function of distance and time, c(x, t), for a
given set of initial and boundary conditions. Even in this
simplified form of Fick’s Second Law, the task is often
quite difficult. Many of the most useful solutions are
described in Crank (1975). The two following problems
Suppose a potentially mobile chemical species is deposited in
a very thin layer of sediment and subsequently buried. Can we
describe how it becomes redistributed in the surrounding sediments during diagenesis?
Geochemists commonly encounter problems like this when
they are asked to evaluate a site where heavy metals, arsenic, or
similar toxic materials from a contaminant spill have entered
soils. Imagine, for example, that copper-rich mud from a mining operation has accidentally been flushed into a lake, where it
settles in a thin layer and is then quickly covered with other sediment. As a result, there is soluble copper in an infinitesimally
thin sediment layer at some distance x0 below the sedimentwater interface. At the start of the problem, this marker horizon contains a quantity, a, of Cu2+ that is not found elsewhere
in the column.
The solution to Fick’s Second Law under these conditions is:
( )
α
x2
c(x, t) = ——–— exp ——– ,
2√
πDt
4Dt
where x is a distance either above or below the reference horizon x0. Unless you want to practice your skills at solving differential equations, you can look up solutions like this in Crank
(1975) or any other reference in which common solutions are
tabulated. This particular one, called the thin-film solution, is
most valuable as the basis for solving more complex problems.
We can verify that it is the correct solution for this problem,
because it can be differentiated to yield equation 5.2 again and
because it satisfies the initial conditions we specified:
for |x| > 0, c(x, t) → 0 as t → 0,
and
for x = 0, c(x0, t) → ∞ as t → 0,
in such a way that the total amount of the mobile species is
fixed:
∫
∞
−∞
c(x, t)dx = α.
Suppose, then, that the marker horizon contains 50 mg of
Cu2+ per cm2, and assume that the diffusion coefficient for Cu2+
in water is 5 × 10−6 cm2 sec −1. If diffusion proceeds outward
from the marker horizon for 2 months, what will be the concentration of copper in pore waters 1 cm away? Applying the
thin-film solution, we find that:
c(1 cm, 5.1 × 106 sec) = 2.74 mg cm−3
= 2.74 × 103 mg kg−1
= 2740 ppm.
82
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
in worked problem 5.1, xP = 10 cm. This result is often
generalized as a rule called the Bose-Einstein relation, describing the penetration distance over which a diffusing
species has traveled:
(5.4)
xP = k√
Dt.
The size of the proportionality constant, k, may vary from
one problem to another, depending on the geometry of
the system and how strictly we want to define the practical limits of the c(x, t) curve.
Now let’s try another common problem, more complicated than the first.
Worked Problem 5.2
FIG. 5.2. Diffusion away from a thin-film source creates a symmetrical concentration profile that flattens and stretches out with
time t. The areas under the curves are identical because the total
amount of diffusing material is constant.
The total mass of diffusing material in a thin-film
problem is typically very small and is concentrated initially in a layer with “zero” thickness. In a real setting,
of course, the source layer has a finite thickness. For that
case, we would need to apply a slightly different solution
to equation 5.2.
We can examine the general consequences of thinfilm geometry by studying figure 5.2, in which we have
plotted c(x, t) against x at successive times after diffusion
outward from the marker horizon has begun. The area
under the c(x, t) curve remains constant (because the
total amount of material is fixed), but the diffusing
species spreads out symmetrically above and below the
source layer as time advances. Referring to the solution
in worked problem 5.1, we can see that the concentration at x = 0, c(x0, t), decreases at a rate proportional
to 1/√t. We also see that the curve has inflection points
(where ∂2c/∂x 2 = 0) above and below x0 , and that these
separate a zone of solute depletion near the source bed
from zones of solute enrichment at a distance. One of
the other questions we might ask, then, is How far will
material have diffused by some time t? This is a difficult
question to answer, because c(x, t) never reaches zero at
any distance once diffusion has begun. One practical
measure of the limit to e−x2 is the point at which it decays
to 1/e (∼0.36) times its value at x0. In this example, that
distance is equal to xP = 2√
Dt. For the numerical example
Suppose that a diffusing species is initially concentrated in one
rather thick layer or a body of water, and then begins to travel
across a boundary into a second layer. How does the concentration of diffusing ions vary in the second layer as a function
of position and time?
Consider the situation illustrated in figure 5.3a. Here, lake
water containing Cu2+ at a concentration c = c′ overlies a sediment column that initially contains no copper. This is the configuration we might expect if, for example, acid mine drainage
were diverted into a previously clean lake, so that the lake
water was suddenly copper-rich while the pore water in the
underlying sediment was still clean. The boundary between the
two water volumes is at the lake bed, where x = x0 = 0. Verify
for yourself that the initial conditions are:
for x > 0 at t = 0, c = 0
(in the sediment),
FIG. 5.3. (a) At the beginning of worked problem 5.2, the concentration of the mobile species is equal to zero at all depths in
the sediment, and is equal to c′ throughout the overlying water.
(b) Diffusion across the water-sediment interface produces profiles
described by equation 5.5. Notice that the concentration at the
interface remains equal to c′/2 at all times after t = 0.
Diagenesis: A Study in Kinetics
[
and
for x < 0 at t = 0, c = c′ (in the lake).
To address this problem, we need something more complicated than the thin-film solution, but we can use the thin-film
solution as a guide. We imagine a column of lake water with
unit cross-sectional area, divided into a set of n horizontal slices,
each of thickness ∆a. Each slice acts as a thin-film source for
solute molecules that eventually reach a depth xP in the sediment. We can therefore calculate the total concentration at xP
after a time t has passed by taking the sum of all the contributing thin-film solutions.
From worked problem 5.1, we know these solutions to have
the form:
( )
c′∆ a
ai2
c(x, t) = ——–— exp –——
,
2√
πDt
4Dt
where ai is the distance from depth x to the center of the ith slice.
The term α, which we used in worked problem 5.1 to equal the
total amount of Cu2+ in the system, has been replaced by c′∆a,
where ∆a is the depth of the lake. If each slice is infinitesimally
thin and we treat the infinite sum of all slices as an integral,
c′
c(x, t) = ——–—
2√
πDt
∫
∞
x
(
(
)]
x
c(x, t) = 1 − erf ———
2√
Dt
)
z2
exp − –—— dz.
4Dt
It is customary to rewrite this analytical result in a form that is
easy to evaluate numerically with a computer or even a pocket
calculator. To do this, we substitute the variable η = z/2√
Dt, so
that:
83
.
Suppose, then, that we add enough Cu2+ to the lake water
to raise its concentration to 500 ppm. If the diffusion coefficient
for Cu2+ in water, as before, is 5 × 10−6 cm2 sec−1, what should
be the copper concentration in pore waters 10 cm below the
water-sediment interface after a year? Applying these values in
our final equation for c(x, t), we get:
c(10 cm, 3.1 × 107 sec) = 500 ppm [1 −
−6 cm
2 sec−1
erf(10 cm /2√
[5 × 10
][3.1 ×
107 sec
])]
= 285 ppm.
In the discussion so far, we have considered only the
diffusion of dissolved material through water. This is
appropriate, because diffusion through any solid medium
is almost negligibly slow compared with diffusion in
water. The bulk diffusive flux within a sediment column
therefore depends on how much volume is not occupied
by solid particles (that is, it depends on the porosity (ϕ).
Diffusion may also be slowed significantly because of the
geometrical arrangement of particles, because the effective path length for diffusion must increase as dissolved
material detours around sediment particles. This second
factor is known as the tortuosity (θ), defined by:
θ = dl/dx,
c′
c(x, t) = —–
√π
∫
∞
2
x/(2√
Dt
)
[
[
c′
= —– 1 −
√π
∫
exp(−η )dη
]
x/(2√
Dt
)
exp(−η2)dη
0
(
(5.5)
)]
c′
x
= —– 1 − erf ———
2
2√
Dt
.
This may not look like much of a simplification, but it is. The
error function, erf(x), is a standard mathematical function that
can be approximated algebraically (see appendix A).
Figure 5.3b illustrates the concentration profile generated
by this equation at several times after the onset of diffusion.
Notice that the profile is symmetrical around the sedimentwater interface, where concentration at all times is equal to
c′/2. Unless the water column is unusually stagnant, as in some
swamps, we would not expect to see a measurable gradient like
this above the interface. Instead, the lake water should be uniformly stirred by winds or thermal convection, with the result
that the concentration for all values of x ≤ 0 is c′ at all times.
We won’t use the space to prove it here, but it can be shown that
c(x, t) in the sediment column with this additional boundary
condition differs only by a factor of two from our first answer:
where dl is the actual path length a diffusing species must
traverse in a given distance dx. Tortuosity cannot be
measured directly, but is usually determined by comparing values of D determined in open water with those
measured in pore water (Li and Gregory 1975), or by
comparing the electrical conductivities of sediments
and pore waters (Manheim and Waterman 1974). Taking
both porosity and tortuosity into account, the diffusion
coefficient in the sediment column, Ds , is related to the
value of D in pure water by:
Ds = ϕD/θ 2.
(5.6)
Worked Problem 5.3
Measurements of Mg2+ concentration are made in the pore
waters of a marine red clay, and it is found that Mg2+ decreases
linearly from 1300 ppm at the sea floor to 1057 ppm at a depth
of 1 m. Assuming that the profile is entirely due to diffusion,
how rapidly is Mg2+ diffusing into the sediment?
84
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Geochemists Yuan-Hui Li and Sandra Gregory (1975) have
reported that the diffusion coefficient for Mg2+ in pure water
close to 0°C is 3.56 × 10−6 cm2 sec−1 and the average porosity
and tortuosity in marine red clay are 0.8 and 1.34, respectively.
We calculate from equation 5.6 that the effective diffusion coefficient, Ds, is therefore:
Ds = (0.8)(3.56 × 10−6 cm2 sec−1)/(1.34)2
= 1.6 × 10−6 cm2 sec−1.
Because ∂c/∂x does not vary with depth, we conclude that
the profile represents a steady-state flux of Mg2+ into the sediment. Thus, we can easily calculate J at the interface from Fick’s
First Law. The concentration gradient is 243 ppm per meter, or
1.0 × 10−2 mol 1−1 m−1, or 0.1 µmol cm−4. Therefore,
J = −Ds∂c/∂x
= (1.6 × 10−6 cm2 sec−1)(0.1 µmol cm−4)
= 1.6 × 10−6 µmol cm−2 sec−1
= 4.96 µmol cm−2 yr−1
= 0.12 mg cm−2 yr−1.
In addition to porosity and tortuosity, other chemical
factors affect diffusion in a diagenetic environment. These
arise because most material transported in pore fluids is
in the form of charged particles. A flux of cations into a
sediment column must be accompanied by an equal flux
of anions or must be balanced by a flux of cations of
some other species in the opposite direction. We may expect that electrostatic attractions or repulsions between
ions will affect their mobilities. The greater the charge
(z) on an ion, the more likely it is that it will be slowed
by interaction with other ions. Figure 5.4 illustrates the
magnitude of this effect for most simple aqueous ions by
plotting the value of D against the ionic potential (|z|/r)
for a variety of common ionic species. Geochemist Antonio Lasaga (1979) has shown that ionic cross-coupling
can change the value of D for some species by as much
as a factor of ten.
Charged species not only interact with each other but
also interact with sediment. Clay particles, in particular,
develop a negatively charged surface when placed in
water, as interlayer cations are lost into solution. Several
important properties of clays arise from this behavior.
Some cations from solution are adsorbed onto the clay
surface to form a fixed layer, which may be held purely
by electrostatic forces or may form complexes. Other
cations form a diffuse layer farther from the clay-water
interface, in which ions constitute a weakly bonded
FIG. 5.4. Tracer diffusion coefficients D for common ionic species
as a function of ionic potential. (Modified from Li and Gregory
1975.)
cloud. This double layer of positively charged ions repels
similar structures around adjacent clay particles. In open
water, these electrostatic effects support small particles
in stable suspensions known as colloids and prevent them
from settling to the bottom. For example, we observe
that sediment particles are carried as colloids in river
water but are commonly deposited as they enter the
ocean, where their diffuse layer is overwhelmed by the
abundance of free ions in salt water. In general, this
tendency to maintain a colloid decreases with increasing
ionic strength. In a sediment column, the double layer
promotes a very open structure, like a house of cards,
and maintains a high porosity.
The way in which surface properties of clays affect
ionic diffusion can be shown by considering the adsorption of a mobile ion A onto a clay particle as a chemical
reaction:
→A
.
A free ←
bound
Diagenesis: A Study in Kinetics
85
Advection
Because there is also no movement of fluid relative to the
sediment if it, too, is simply buried, we call its apparent
flow pseudoadvection. To avoid being misled by pseudoadvection, we can choose an alternate reference frame
centered on the bed itself instead of one that is centered
on the sediment-water interface.
True fluid advection in marine or lacustrine sediments
is most commonly driven by the gradual reduction in
porosity that occurs as sediments are compacted with
depth. If we choose a bed-centered reference frame, we
usually anticipate a gentle upward flow, as water is
squeezed out of sediments in deeper layers. This effect
should be most prominent in clay-rich sediments, which
initially have high porosities because of electrostatic
repulsion between particles. Within a few tens of meters
of the surface, these weak repulsive forces are overwhelmed by the weight of accumulating sediment, and
porosity drops from 70–80% to <40%, gradually approaching an asymptotic limit within a few tens of meters of the surface. Porosity decreases very rapidly within
a few tens of meters of the sediment surface, and then only
very slowly for the next few thousand meters, as shown
in the schematic profile in figure 5.5. For our purposes,
it is reasonable to treat the porosity below a depth of
∼30–50 m as a constant.
Even in high-porosity clays, however, the gentle upward flow of fluid is too slow to measure, because clays
typically accumulate and compact very slowly. In sandy
sediments, it is even slower. Particularly in those sands
with poor sorting, porosity is initially low and does not
Water can flow through pores in a sediment as the result of compaction or a regional hydraulic gradient. The
first of these is most likely to dominate in a marine or
lacustrine environment and is therefore most commonly
associated with early diagenesis. Regional flow, however,
can be important when dissolved matter is transported
and interacts with sediments in an aquifer or tectonic
pressure induces fluid flow in deep sediments.
As sediment accumulates, we can examine the progress of diagenesis in a given bed in either of two reference frames. In one frame, we locate the bed by referring
to its depth below the sediment-water interface. With
continued deposition, the depth to the bed increases.
Viewed from this perspective, both the sediment and any
interstitial fluid seem to move with time. Unless the sediment is compacted as it is buried, however, the sediment
particles are not actually moving relative to each other.
FIG. 5.5. Porosity in clastic sediments decreases to a nearly
asymptotic limit within a few tens of meters of the surface. This
approximation makes it possible to estimate advective flow rates
caused by compaction.
Ignoring any effects of nonideality, the equilibrium constant for this reaction is given by:
K = C/c,
where c is the concentration of A in solution and C is its
concentration in bound form, in units of moles per liter
of interstitial solution. Because the total amount of A is
constant, the rate of change in c must be equal and opposite to the rate of change in C. For a rapid exchange
between solution and solids, then, we can write:
K(∂c/∂t) = −(∂C/∂t).
Equation 5.2, with this modification, becomes:
∂c/∂t = D∂2c/∂x2 + ∂C/∂t
= D∂2c/∂x2 − K∂c/∂t,
or
∂c/∂t = [D/(1 + K)]∂2c/∂x2.
The rate of diffusion, therefore, becomes much slower as
the role of ion exchange or adsorption by clay surfaces
increases. If we take chemical interactions with solids
into account, the effective diffusion coefficient defined in
equation 5.6 now becomes:
ϕD
Ds = ———–—.
θ 2(1 + K)
(5.7)
86
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
change much with depth. The angularity of coarse clastic particles makes reorientation and compaction difficult.
Regardless of particle size and burial rate, then, we cannot directly measure fluid advection due to compaction
during marine or lacustrine diagenesis. We can only estimate its rate by constructing theoretical models of the
compaction process. The simplest model we can devise
(and the only one we consider in this book) assumes that
the sedimentation rate is constant and ∂ϕ/∂t = 0 at any
depth in the sediment column. This is very nearly true
within a few meters of the sediment-water interface.
Let us adopt the symbols ω and u to indicate the rates
at which sediment and pore waters are buried, respectively. If there were no compaction, ω would be equal to
u, and both would simply be equal to the rate of sediment
accumulation. Under steady-state compaction, however,
we find that as porosity decreases toward its asymptotic
limit (ϕd), ω and u each decrease as well, gradually approaching a limit where ωd = ud . As long as the change
in porosity with depth is fairly small (as it is below about
10 meters in the sediment), we conclude that:
uϕ = ϕd ud = ϕd ωd .
Assuming steady-state compaction, therefore, it is possible to calculate the rate of fluid burial at any depth if
we have a porosity-depth profile and can estimate ω.
Clearly, however, at any depth less than the one where
ϕ, u, and ω all approach their asymptotic limits, fluids
are buried less rapidly than sediment. Expelled waters,
therefore, flow upward at a rate (U) equal to:
Yale University geochemist Robert Berner (1980) showed
that if we assume steady-state compaction, ∂[(1 − ϕ)ω]/∂x = 0.
That is, [1 − ϕ(x)]ω(x) has a fixed value. To find what that value
is, note that the sediment burial rate at the sediment interface,
ω0, is equal to the sediment deposition rate. We can evaluate
ϕ(x) at x = 0 to see that ϕ(0) = 0.5e (0.0) + 0.2 = 0.3. Therefore,
[1 − ϕ(x)]ω(x) = 0.3ω0.
The burial rate of the sediment in this column, therefore, is
given by:
ω(x) = 0.3ω0 /[1 − (0.5e (−0.05x) + 0.2)]
= 3ω0 /[8 − 5e (−0.05x)].
In Puzzle Pond, the sedimentation rate is 0.2 cm yr−1.
Applying our expressions for ϕ(x) and ω(x), we calculate that
ϕ and ω approach nearly constant values (ϕd = 0.22 and ωd =
0.077 cm yr−1) ∼60 meters down. These, then, can be inserted
in equation 5.7 to yield:
U(x) = [0.22(0.077 cm yr−1)/ϕ(x)] − ω(x).
The fluid advection rate at selected depths can be calculated by
substituting appropriate values for ϕ(x) and ω(x):
x (m)
ϕ(x)
ω(x)
U(cm yr−1)
5
10
15
20
30
40
50
60
0.59
0.50
0.44
0.38
0.31
0.27
0.24
0.22
0.15
0.12
0.11
0.097
0.087
0.082
0.079
0.077
−0.117
−0.087
−0.068
−0.053
−0.033
−0.019
−0.009
−0.002
Notice that the advection rates are all negative, indicating upward flow.
U = u − ω.
If we multiply both sides of this expression by ϕ and combine it with the previous equation, we see finally that it
is possible to calculate the rate of upward flow from:
U = (ϕdωd/ϕ) − ω.
(5.8)
Worked Problem 5.4
A team of researchers has measured porosity as a function of
depth in a sediment column at the bottom of mythical Puzzle
Pond. They find that they can describe the data adequately with
the mathematical function:
ϕ(x) = 0.5e (−0.05x) + 0.2,
where x is depth in meters. How can we use this information
to estimate the rate of compaction-driven fluid flow in the
sediments?
Because these rates of fluid advection depend so
strongly on grain size, sorting, and sedimentation rate, it
is hard to devise reliable models that are more detailed
than those we have described above. Given a semiempirical model like this, however, it is easy to compare rates of diffusion and advection in a sediment and
see which is the major mechanism for transport during
diagenesis.
Worked Problem 5.5
Which is the dominant process for interstitial transport of
Mg2+ in a typical marine mud, diffusion or advection? In
worked problem 5.3, we calculated that the effective diffusion
coefficient (Ds ) for Mg2+ under these circumstances is ∼1.6 ×
10−6 cm2 sec−1. To answer the question, then, we compare that
value of Ds against the advection rates we calculate with equa-
Diagenesis: A Study in Kinetics
tion 5.8, using the same parameter values we used in worked
problem 5.4. To express Ds and U in compatible units, we will
first need to adjust U for porosity, multiplying it by ϕ to produce
an effective advection rate, Us. To convert its dimensions of
cm yr−1 into cm sec−1, we then divide Us by 3.1 × 107 sec yr −1.
The ratio Ds /Us has the dimensions of distance (L), in centimeters. To evaluate the balance between diffusion and advection, then, we can divide Ds /Us by L to create the Peclet
number, Ds /LUs , which is dimensionless. If the Peclet number
for an environment is <<1, then advection dominates; if Ds /LUs
>> 1, then diffusion is the more effective transport process.
The critical distance Lc at which the Peclet number equals 1.0,
therefore, defines the physical boundary between diffusiondominated and flow-dominated regimes in a sediment. Because
Us is a function of depth, Lc also varies with depth. In this
problem, we find:
x (m)
5
10
15
20
30
40
50
60
Lc (m)
7.2
11.3
16.8
24.3
48.6
99.4
234
1058
From these results, we conclude that because the actual distance
Mg2+ has to travel to escape from the sediment column into the
overlying water is always significantly less than Lc, diffusion is
the major transport process during diagenesis of marine muds.
This result would not be significantly different if we had chosen
to follow a different ion (see fig. 5.4).
To illustrate how much this inference depends on the value
of U, consider the transport of Mg2+ in a sandy aquifer, where
the rate of flow is imposed by some regional pressure gradient.
To calculate such a rate, we might use a form of Darcy’s law, the
flow equation familiar to hydrologists:
Us = k (1/h),
where k is a measure of permeability and 1/h is the local hydraulic gradient. It would not be unusual to find flow rates in
excess of 100 cm yr−1. Substituting in the expression above, we
find that the Peclet number is equal to 1.0 when Lc = 0.5 cm.
Not surprisingly, then, we verify that flow, not diffusion, is the
major transport mechanism in an aquifer.
KINETICS OF MINERAL DISSOLUTION
AND PRECIPITATION
In our discussion so far, we have considered changes in
the sediment column from a mechanical point of view, in
which diffusing or advecting ions are carried as inert particles. Variations in fluid composition during diagenesis,
however, are not due to diffusion or advection alone. In
87
addition, dissolved material can be added to solution (or
lost from it) in a variety of chemical reactions. Reactions
between primary minerals in the sediment may produce
or consume dissolved species, as can the growth of new
(authigenic) minerals.
Biological processes associated with decay or respiration can act in a number of ways to change the abundance of organic compounds in the sediment column and
modify the composition and redox state of pore waters.
We look carefully at some of these processes in chapter 6. We consider dissolution-precipitation reactions
from a thermodynamic perspective in chapters 7 and 8.
The thermodynamics of redox reactions, which may
affect the concentration of transition metal ions like
Mn2+ or Fe2+ in solution, is also the subject of chapter 7.
Although we devote a lot of attention to thermodynamics in these chapters, it is important to recognize
that chemical reactions at low temperature are commonly
arrested in midflight. That is, the phase assemblages
and compositions we observe in a sediment column frequently represent some intermediate step along the way
toward equilibrium rather than equilibrium itself. Because this is so, thermodynamics is often inadequate to
interpret the chemical variability observed in sediments.
Instead, geochemists find that they can apply principles
of kinetics to minerals and fluids in the sediment and
thus deduce the reaction pathways that dominate during
diagenesis, as well as the physical or transport pathways.
We demonstrate two very different approaches in this
final section of the chapter, first considering kinetics at
the molecular level and later at a macroscopic level.
Antonio Lasaga (1998) makes a useful distinction
between elementary chemical reactions, which occur as
written at the molecular level, and overall reactions,
which describe a net change that involves several intermediate steps and take place along several competing,
parallel pathways. We have also encountered this distinction briefly in chapter 1, where thermodynamic and kinetic approaches to geochemistry were contrasted. As we
show here, this distinction is also reminiscent of the contrast between atomistic and empirical perspectives on
chemical change, with which we opened this chapter. Let
us consider a few short examples to see how this distinction affects the way we treat reactions during diagenesis.
The chemical reactions that occur during diagenesis
include some processes like carbonate precipitation, for
which we can write an elementary reaction:
Ca2+ + CO32− → CaCO3.
88
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
The rate at which molecules of CaCO3 are produced is
directly proportional to the probability that a Ca2+ ion
will collide with a CO32− ion in solution. Therefore, the
appropriate rate law in this case is:
dnCaCO3/dt = knCa2+nCO32−,
in which nCa2+ , nCO32− , and nCaCO3 are the concentrations
of each species and k is a linear rate constant. In fact, for
any elementary reaction, the rate is simply proportional
to the abundance of reactant species.
To put this into a more theoretical framework, remember that in chapter 4, we introduced notation for
writing a generic chemical reaction in the form:
→νA +ν A
νj+1A j+1 + . . . + νi−1A i−1 + νi A i ←
1 1
2 2
+ . . . + νj A j .
We introduced the notion of flux in chapter 1 in equation 1.3:
dAout
/dt =
i
i
Applying this equation, it can be shown that the rate of
a generic forward reaction (vf) is represented by:
ν
ν
ν
j+1
j+2
vf = dAforward /dt = kf (aj+1
aj+2
. . . a i i),
Reactions involving carbon dioxide and water are important in
many geologic environments, including those undergoing diagenesis. One important reaction in this system involves the
equilibrium between dissolved CO2 gas and the bicarbonate ion:
→ HCO − + H+.
CO2(aq) + H2O ←
3
vr = dAreverse /dt = kr (a1ν1 a 2ν2 . . . ajνj),
in which the quantities kf and kr are first-order rate constants. Thus, as in the party example of chapter 1, vf and
vr are each proportional to the abundance of reactant (or
product) species.
At equilibrium, the forward and reverse reaction rates
must be equal if there is to be no net change in the system.
Therefore, there is a direct relationship between the rate
constants and the equilibrium constant for the reaction:
Keq = kf /kr .
Furthermore, the degree of disequilibrium can be estimated from the ratio vf /vr and the activities of species in
the system, because:
ν
ν
ν
ν
ν
j+1 a j+2 . . . a i )/(a 1 a 2 . . . a j ).
kf /kr = (vf /vr )(aj+1
j
1 2
j+2
i
This expression can be written, for greater clarity, in the
form:
kf /kr = Keq = (vf /vr )Q.
→ H CO ,
CO2(aq) + H2O ←
2
3
an elementary reaction, and then much more quickly by dissociation:
→ HCO − + H+,
H2CO3 ←
3
and the rate of the reverse reaction (vr) by:
ν
Worked Problem 5.6
This is actually an overall reaction that proceeds first by the
slow step of hydration:
Σk A .
ij
more than that. If Keq /Q > 1, then the rate of the forward reaction exceeds the rate of the reverse reaction. If
Keq /Q < 1, the opposite is true. Furthermore, we can see
that a reaction proceeds most rapidly (in either a forward or a backward direction) when it is farthest from
equilibrium; that is, when Keq /Q is very much greater or
less than 1. For this reason, the direction and rate of reaction are often good indicators of how far a system may
be from equilibrium.
(5.9)
This expression degenerates to Keq = Q (equation 4.12)
for the equilibrium case, where vf /vr = 1, but it tells us
another elementary reaction. Hydration is such a slow step,
comparatively, that CO2(aq)/H2CO3 is ∼600 at 25°C. The overall reaction, therefore, behaves as if it were a much simpler elementary reaction and follows approximately a first-order rate
law. For the sake of this worked problem, then, let’s pretend
that it really is an elementary reaction. Can we use its kinetic
behavior to describe the approach to equilibrium and estimate
the value of Keq?
The combined rate constants for the two-step reaction
(Kern 1960) are kf = 3 × 10−2 sec−1 and kr = 7.0 × 104 sec−1.
Figure 5.6 shows the progress of reaction in a system presumed
to consist initially of 1 × 10−5 mol liter−1 CO2(aq) at pH = 7(aH+
= 1 × 10−7). Hypothetically, the concentrations of CO2(aq) and
HCO3− could be measured as functions of time, but this is a
nearly impossible task in the CO2-H2O system. To construct
figure 5.6, therefore, the concentrations were calculated algebraically from the experimentally determined rate constants.
For details of the method, see Stumm and Morgan (1995).
Our interest here is in the rates of reaction as the overall reaction approaches equilibrium. These can be calculated at any
given time from:
vf = kf aCO2(aq),
and
vr = kraHCO3− aH+.
Diagenesis: A Study in Kinetics
89
The results are shown in figure 5.7. Notice that the rate profiles for the forward and reverse reactions are not mirror images
of each other. During the approach to equilibrium, vf /vr > 1, so
Keq /Q > 1. Both ratios decrease with time, however, and within
roughly 20 seconds, vf and vr approach the same value (2.4 ×
10−4 mole sec−1) to <1% error. The dashed line in figure 5.7
indicates the progress of the net reaction rate, vf + vr, which
approaches zero. The value of Keq, calculated from the rate constants by equation 5.8, is 4.3 × 10−7. This is virtually the same
as the value (4.4 × 10−7) calculated from the free energies of formation of dissolved species.
FIG. 5.6. Computed concentrations of CO2(aq) and HCO3− as a
→ HCO − + H+
function of time for the reaction CO2(aq) + H2O ←
3
at 25°C in a closed aqueous system. The total concentration
of carbon species remains constant at 1 × 10−5 moles. (Modified
from Stumm and Morgan 1995.)
Unlike the elementary reactions in this simple example, unfortunately, the rates at which overall reactions
occur are commonly nonlinear functions of concentration. They may be influenced by processes that occur at
the surface of a growing or dissolving particle, by the production of transient, metastable species, or by transport
of dissolved material toward or away from the interface.
As a result, rate laws may be quite complex, reflecting a
variety of inhibiting or enhancing processes. Robin Keir
(1980), for example, has determined by experiment that
the rate of calcite dissolution, R, in seawater follows the
nonlinear relationship:
R = k[1 − (aCa2+ aCO32− /Kcal)]4.5,
in which Kcal is the Keq for the reaction:
→ Ca2+ + CO 2−,
CaCO3 (calcite) ←
3
FIG. 5.7. Computed reaction rates for the forward and reverse
directions of the approximate first-order reaction illustrated in
figure 5.6 as functions of time. At equilibrium, both reaction
rates approach 2.4 × 10−4 mole sec−1. (Modified from Stumm and
Morgan 1995.)
and k is an empirical constant that varies widely among
samples and is most strongly a function of grain size
(fig. 5.8). The large exponential factor has been hypothesized to arise from the absorption of phosphate ions
on crystal surfaces, where they serve as dissolution inhibitors. Similarly, Robert Cody and Amy Hull (1980)
have demonstrated in the laboratory that some organic
acids are preferentially adsorbed on nuclei of gypsum,
inhibiting further growth and leading instead to lowtemperature formation of anhydrite from saturated calcium sulfate solutions. In both of these cases, we cannot
predict the rate law without detailed knowledge of the
reaction mechanism or, as is more often the case, an exhaustive set of empirical observations.
For many overall reactions that affect diagenesis,
we have little theoretical knowledge and few relevant
observations, and we cannot therefore adopt an atomistic approach to studying them. The formation of sulfide
in sediments, for example, is associated with bacterial
90
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 5.8. The rate of dissolution of calcite (expressed as the rate
constant) increases rapidly with decreasing grain size because of
the increased surface area in fine-grained materials. (Modified
from Keir 1980.)
reduction of sulfate in pore waters. We can write a reasonable reaction to express this process,
→
7CH2O + 4SO42− ←
2S2− + 7CO2 + 4OH− + 5H2O,
but it is valuable only as a guide to system mass balance.
The species labeled “CH2O” is a generalized compound
with the average stoichiometry of organic matter in sediments. The rate of sulfide formation during diagenesis
depends on a large number of factors that are not indicated in the schematic form of the overall reaction. These
may include the identity of specific organic molecules in
the fresh sediment and any that form after burial, surface reactions on growing sulfides, and absorption by clay
surfaces. We examine these factors in greater detail in
chapter 6. Our understanding of reaction kinetics in such
cases is primitive.
Fortunately, it is often possible to model diagenetic
environments that are dominated by mineral growth
without dealing with overall reactions at an atomistic
or molecular level. Before we finish this section, let us
consider another specific diagenetic problem and demonstrate an empirical approach to handling the kinetics of
overall reactions.
Bottom waters throughout most of the ocean are oxygenated. Within the sediment column below the ocean
floor, however, oxygen is rapidly consumed by bacteria
that feed on organic debris. As mineral dissolution reactions take place in the sediment, then, any iron and
manganese become mobilized as reduced Fe2+ and Mn2+
rather than remaining oxidized and relatively immobile.
These ions migrate, primarily by diffusion, toward the
sediment surface, where they become reoxidized to Fe3+
and Mn3+ or Mn4+ and precipitate to form a stable layer
of Fe-Mn-oxyhydroxides. Manganese nodules are formed
on the ocean floor by a closely related process. Examined in detail, this redistribution of metal ions to form
new minerals involves a complicated network of organic
and inorganic reactions, many of which are nonlinear.
Burdige and Gieskes (1983), however, developed an analytical solution that does not make explicit reference to
the overall reactions that are occurring in the sediment
column. Figure 5.9a describes Mn2+ variations in pore
water with depth in a sediment core from the east equatorial Atlantic Ocean. The solid curve is Burdige and
Gieskes’s solution to a differential equation that includes
measured values of porosity, sedimentation rate, and
other key parameters in the sample area. Rather than lay
out the details of their rather complicated model, let’s
follow the example of Robert Berner (1980) by using a
simpler model to illustrate the same principles.
Worked Problem 5.7
How does the distribution of manganese in pelagic sediments
change with time? Berner’s model, illustrated in figure 5.9b, assumes that the zone of mineral precipitation grows upward
from the sediment-water interface (x = 0) and consists only of
oxyhydroxide minerals and fluid-filled pore spaces. This is possible only if both the sedimentation rate and bioturbation are
insignificant. Berner further assumes that there are no porosity
gradients, and that Mn2+ is not adsorbed on clay surfaces, so
that advective flow is negligible and the rates of dissolution reactions that initially mobilize Mn2+ are rapid enough to not
limit the eventual precipitation process. He has therefore reduced the complex set of elementary reactions in this overall
system to a single precipitation reaction. He describes this reaction implicitly by monitoring how rapidly the zone of mineral precipitation grows instead of including the reaction rate
itself in the model. Finally, because the concentration gradient
in the transition zone between the precipitation layer and the
top of the source region (x = L) is nearly linear, he makes the
approximation:
∂C/∂x = (CL − C0 )/L,
where CL and C0 are the concentrations of Mn2+ at x = L and
x = 0, respectively.
Diagenesis: A Study in Kinetics
91
FIG. 5.9. (a) Plot of the manganese content of pore water as a function of depth in a sediment
core from the equatorial Atlantic Ocean. Solid curve is the model profile calculated by Burdige and
Gieskes (1983). (b) Concentration profile in the simplified redox model for marine Mn redistribution
discussed by Berner (1980).
The concentration profile in this model consists, then, of an
uninterestingly constant segment below depth L, and a linear
segment above it, the slope of which changes with time. It bears
a crude resemblance to the measured curve in figure 5.9a, offset vertically so that the zone of mineral deposition lies entirely
above the sediment. Rather than using this model simply to
calculate the concentration profile, however, Berner shows how
it is possible to estimate how long it takes to grow a monomineralic surface layer of any given thickness.
The flux of dissolved Mn at x = 0 at any given time can be
calculated from Fick’s First Law, adjusted for sediment porosity
and tortuosity:
If we then insert this result in the expression for dl/dt and integrate for the boundary conditions l = 0 when t = 0 and l = l when
t = t, we have derived an expression for t:
t = l2Fp2Vd /(2Vp2 Fd Ds [CL − C0 ]).
We can now determine how long it takes to form an oxidized surface layer of an arbitrary thickness (1 cm) by placing
reasonable values for each of the variable parameters in the last
two equations:
Ds = 2.4 × 10−6 cm2 sec−1,
Vp = Vd = 20 cm3 mol−1,
J0 = −Dx ∂C/∂x = −Dx(CL − C0 )L.
The thickness of the transition zone (that is, the value of L)
varies as a function of time that is directly proportional to J0:
Fp = 0.2,
and
C0 = 0.0 M.
dL/dt = Vd Ds(CL − C0 )/FdL.
Here, Vd = the volume of dissolving mineral matter that contains
one mole of Mn2+ and Fd = the volume fraction of dissolving
matter in the bulk sediment.
In a similar way, the thickness of the oxyhydroxide layer (l)
grows according to:
dl/dt = Vp Ds(CL − C0 )/Fp L,
in which Vp and Fp are similar to Vd and Fd , but refer to the mineral matter in the precipitation zone.
The thickness of the transition zone can be calculated by integrating dL/dt for the boundary conditions L = 0 when t = 0
and L = L when t = t:
L = [(2Vd Ds /Fd )(CL − C0 )t]1/2.
For several pairs of Fd and CL values, we calculate t and L:
Fd
0.001
0.002
0.005
0.010
0.010
CL (mol cm−3)
−7
1 × 10
1 × 10−7
1 × 10−7
1 × 10−7
1 × 10−6
L (cm)
t (yr)
200
100
40
20
20
133,500
66,770
26,700
13,350
1335
The geometry of this model is similar in many ways to the
more complex example presented by Burdige and Gieskes, and
the numerical results may be generally comparable as well. Notice, however, that the predicted time scale for growth of the
authigenic layer is long enough to stress some of the initial assumptions of the model. In nature, it is unlikely that a shallow
92
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
sediment column would remain undisturbed by bioturbation
and chemical perturbations for 103–105 years.
The primary appeal of this approach is that it avoids
the necessity of describing directly each of the many
chemical reactions in the overall system. By dealing instead with the observed transport of chemical species and
focusing on the few elementary reactions whose rates
limit the overall system, this approach makes an otherwise impossibly difficult problem easy. Depending on the
complexity of the dissolution/precipitation problem, then,
geochemists can adopt the atomistic or molecular-level
perspective that we have identified with Lasaga and with
the CO2–H2O example in worked problem 5.6, or they
can choose a more empirical perspective, such as the one
demonstrated in worked problem 5.7.
THE DIAGENETIC EQUATION
With this brief survey, we have introduced each of the
important processes that affect the mass balance in a
sediment during diagenesis. Our final task is to combine
them, in what Robert Berner (1980) has called a general
diagenetic equation. The rate at which concentration of
some interesting chemical species in a particular volume
of sediment changes can be expressed by:
∂C/∂t = −∂J/∂x +
ΣR.
Here, the first term on the right side describes the total
flux of the species through the volume per unit time, and
the second term includes the rates, R, of all relevant chemical reactions proceeding in the volume. The distance
variable, x, is taken relative to the sediment-water interface. From discussions in previous sections, we know that
J must include both the diffusive and the advective flux:
J = −D∂C/∂x + UC.
The chemical reaction rates, R, may be explicit rates
of elementary reactions or may be generalized in a more
empirical fashion, as we have just shown. The general
form of the diagenetic equation, then, becomes:
( )
∂C
∂ D—–
∂C
∂x
∂(UC)
—– = ———— − ——— +
∂t
∂x
∂x
ΣR.
(5.10)
This general equation implicitly includes each of the factors we have discussed in previous sections, as well as
many we have not, such as the effect of bioturbation.
SUMMARY
We have spent this chapter discussing the kinetics of
transport and reaction, a departure from our emphasis
on thermodynamics in earlier chapters. This is appropriate, because so many of the changes we associate with
diagenesis are incomplete—limited by the rates at which
reactants move from one place to another or by how
slowly they combine. We have introduced and contrasted
diffusion and advection as transport mechanisms, and
have experimented with mathematical methods for
evaluating both. We use these methods in later chapters
as well, particularly in the very different context of melt
crystallization, in chapter 11. We have tried to show
how these tools can be applied in the form of a general
diagenetic equation to problems of practical interest in
sedimentary geochemistry. In this survey of diagenesis,
we have deliberately postponed most references to the
fate of organic matter, which is a large enough topic to
deserve much of our attention in chapter 6.
suggested readings
Few books have been written on the geochemical side of diagenesis, although there is a great deal of interest among geologists. Most of the informative writing is disseminated in journal
articles. For general background, you may wish to consult the
following sources.
Berner, R. A. 1980. Early Diagenesis: A Theoretical Approach.
Princeton: Princeton University Press. (This is the best single
reference in the field. Much of what is written today about
the chemistry of diagenesis, including what we have said in
this chapter, was first considered by Berner. The first five
chapters are particularly useful; they establish the fundamental principles we have introduced here.)
Lasaga, A. C. 1998. Kinetic Theory in the Earth Sciences. Princeton: Princeton University Press. (This is a comprehensive
volume, presenting what geochemists know about kinetics
over a wide range of environments. It is heavy on theory, but
very clearly written.)
Scholle, P. A., and P. R. Schluger, eds. 1979. Aspects of Diagenesis. Special Publication 26. Tulsa: Society of Economic
Paleontologists and Mineralogists. (This volume contains
widely quoted papers from two symposia on clastic diagenesis.)
Stumm, W., and J. J. Morgan. 1995. Aquatic Chemistry: Chemical Equilibria and Rates in Natural Waters, 3rd ed. New
York: Wiley. (A detailed yet highly readable text with an
emphasis on the chemical behavior of natural waters. Chapter 6, dealing with precipitation and dissolution, is particularly well written.)
Diagenesis: A Study in Kinetics
The following sources were referenced in this chapter. The
reader may wish to examine them for further details.
Burdige, D. J., and J. M. Gieskes. 1983. A pore water/solid
phase diagenetic model for manganese in marine sediments.
American Journal of Science 283:29–47.
Cody, R. D., and A. Hull. 1980. Experimental growth of primary anhydrite at low temperatures and water salinities.
Geology 8:505–509.
Crank, J. 1975. The Mathematics of Diffusion. London: Oxford University Press.
Keir, R. S. 1980. The dissolution of biogenic calcium carbonate
in seawater. Geochimica et Cosmochimica Acta 44:241–254.
93
Kern, D. B. 1960. The hydration of carbon dioxide. Journal of
Chemical Education 37:14–23.
Lasaga, A. C. 1979. The treatment of multi-component diffusion and ion pairs in diagenetic fluxes. American Journal of
Science 279:324–346.
Li, Y.-H., and S. Gregory. 1975. Diffusion of ions in sea water
and in deep-sea sediments. Geochimica et Cosmochimica
Acta 38:703–714.
Manheim, F. T., and L. S. Waterman. 1974. Diffusimitry (diffusion coefficient estimation) on sediment cores by resistivity
probe. Initial Reports of the Deep Sea Drilling Project 22:
663–670.
PROBLEMS
(5.1)
During a catastrophic flood, a barrel of liquid nuclear waste containing radium is carried away by a
major stream and dropped on the floodplain, where it bursts and is spread in a thin layer. Assume
that overbank deposits cover the material rapidly, and that no process affects the distribution of
radium other than diffusion. Using the diffusion coefficient for Ra2+ from figure 5.4, calculate the
effective vertical distance over which radium would be distributed in six months.
(5.2)
A zinc smelter is required to dispose of slurry from its pickling operation, containing 50 ppm Cd2+,
into a “safe” environment. Management wants to pump it into a clay pit abandoned by a neighboring brick factory. Use equation 5.5 to calculate the one-dimensional distribution of cadmium in the
clay after the pit has been used for 3 years. Use a value of D from figure 5.4 and assume that porosity
is 0.1 and tortuosity is 1.25. Assume that no process other than diffusion takes place.
(5.3)
It is determined that porosity in sediments at the bottom of a lake follows the relationship:
ϕ = 0.8e (− 0.07x) + 0.08.
Assume that sediment accumulates at the rate of 0.15 cm yr−1 and compaction is steady state. Calculate and graph the porosity, rate of sediment burial, ω, and vertical fluid advection rate as functions
of depth.
CHAPTER SIX
ORGANIC MATTER AND BIOMARKERS
A Different Perspective
OVERVIEW
Organic geochemistry is built on the application of organic chemical principles and analytical techniques to
sedimentary geology. Carbon, nitrogen, oxygen, sulfur,
hydrogen, calcium, and iron are the primary elements
that living organisms utilize in their structural tissues,
for energy harvesting, and for replication. Accumulation
of organic matter in recent and ancient sediments is the
most important link between the biosphere and geology.
Not only are the organic materials in sedimentary rocks
an economically important resource (coal and petroleum)
but, as we shall see, they also provide a molecular record
of life. The sedimentary burial of organic matter is also
important to the global cycles of carbon, sulfur, and oxygen over geologic time.
In this chapter, we focus on understanding the production, degradation, and preservation of organic matter
in sedimentary environments. We introduce the role of
the carbon cycle in organic matter production, diagenesis, and preservation. Our focus will be on biochemically
important compounds and the relationship of specific
biomarkers to their precursor biota. We discuss mechanisms and factors that influence the accumulation of
organic-rich sediments and evaluate biomarker distribution in recent sediments and ultimately in the sedimentary
record. We close this chapter with a few examples illus94
trating how biomarkers are used to reconstruct paleoclimatic and paleoenvironmental conditions.
ORGANIC MATTER IN THE GLOBAL
CARBON CYCLE
Organic geochemistry emerged in 1936, when Alfred
Treibs discovered and described porphyrin pigments isolated from shale, oil, and coal. Treibs demonstrated that
porphyrins originated from the degradation of chlorophyll, thereby linking the biochemicals in living organic
matter to the ancient sedimentary record.
To understand the fate of organic matter in the biosphere and in sediments, it is useful to have an appreciation of how carbon is distributed among different
geochemical environments and how it is transferred from
one of them to another. We examine this topic at greater
length in chapter 8, where we consider various ways to
characterize environments and transfer processes for
the purpose of creating analytical models of ocean chemistry. As you will see, there is no “best” way to model the
global carbon cycle. Rather than begin that discussion
now, let’s just consider the highly simplified model in figure 6.1, which is reminiscent of the party illustration we
used in chapter 1. The inputs and outputs from the various reservoirs are roughly in balance, resulting in what
can be considered a steady state system.
Organic Matter and Biomarkers: A Different Perspective
95
FIG. 6.1. Simplified version of the global carbon cycle indicating the principal reservoirs (boxes;
see table 6.1 for amounts of carbon in most boxes) and pathways (arrows).
The Earth contains ∼1023 g of carbon, disseminated
between sedimentary materials and active surficial reservoirs. Most of this carbon is sequestered in carbonate
rocks (6.5 × 1022 g C) and organic materials, such as
kerogen or coal (1.56 × 1022 g C) (Schlesinger 1997).
The inventories show that in shales and other sedimentary rocks, essentially one out of every five carbon atoms
is organic (table 6.1) and ∼90% of this preserved organic
matter is amorphous kerogen, with the remaining 10%
bitumen. Kerogen is insoluble, high molecular weight
organic matter derived from algae and woody plant
material that can yield petroleum products when heated.
Bitumen is soluble in organic solvents and is formed from
kerogen during petroleum generation.
Although a large fraction of the Earth’s organic and
inorganic carbon is sequestered, only ∼0.1% (40 × 1018 g)
of the carbon in the Earth’s upper crust is cycled throughout active surface reservoirs (table 6.1). The largest active reservoir is dissolved inorganic carbon (DIC) contained in the global ocean. Other inorganic reservoirs,
such as soil carbonates and atmospheric carbon, and
the organic reservoirs (soil humus, land plant tissue, dissolved organic carbon [DOC] in seawater, and carbon
preserved in marine sediments) are one to two orders of
magnitude smaller than the DIC reservoir.
Atmospheric carbon exists mainly as carbon dioxide,
which is used by plants and other photosynthetic organisms, thereby linking the atmosphere with the biosphere
TABLE 6.1. Major Reservoirs of Inorganic and Organic Carbon
Reservoir Type
Sedimentary rocks
Inorganic
Carbonates
Organic
Kerogen, coal, etc.
Active (surficial) reservoirs
Inorganic
Marine DIC
Soil carbonate
Atmospheric CO2
Organic
Soil humus
Land plant tissue
Seawater DOC
Marine surface sediments
After Hedges and Keil (1995).
Amount
(1018 g C)
Reference
60,000
Berner (1989)
15,000
Berner (1989)
38
1.1
0.66
1.6
0.95
0.60
0.15
Olson et al. (1985)
Olson et al. (1985)
Olson et al. (1985)
Olson et al. (1985)
Olson et al. (1985)
Williams and Druffel (1987)
Emerson and Hedges (1988)
96
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
and oceans (fig. 6.1). The ocean DIC reservoir moderates changes in atmospheric CO2 concentration and, at
equilibrium, the oceans contain ∼56 times the carbon in
the atmosphere.
The organic reservoirs constitute only ∼8% of the
total active surficial reservoirs and are typically composed
of a complex mixture of heavily degraded substances
(kerogen and bitumen), in which there are generally few
readily identifiable biochemicals. The soil reservoir contains the largest concentration of organic and inorganic
carbon in the terrestrial environment. Organic matter is
almost totally recycled in terrestrial soils; therefore,
terrestrial organic-rich deposits are typically confined
to peat formations in bogs and low-lying swamps. Although the modern peat reservoir is small compared with
organic-rich marine sediments, the presence of large coal
deposits in the sedimentary record suggests that peat accumulation may have been an important process in the
geologic past.
The amount of carbon in land plant tissue is similar
to that contained in ocean waters and sediments. Together, these reservoirs contain approximately half as
much carbon as is sequestered in soils. We have more to
say about organic matter in soils in chapter 7, and we
return to the topic later in this chapter, when we focus
on diagenesis. For now, let’s turn our attention to the
areas of the globe where most of the organic biomass is
produced—the oceans.
ORGANIC MATTER PRODUCTION AND
CYCLING IN THE OCEANS
Organic matter is derived from the tissues of living
organisms. Photosynthetic organisms (land and marine
plants) capture sunlight energy, store it in organic compounds, and release the O2 that sustains other organisms.
The organic matter produced by these organisms is almost completely utilized in the biosphere, but a small
portion is preserved. Phytoplankton are the primary
producers in the oceans, much as vegetation is in the
terrestrial realm. Compared with the massive forests,
marine phytoplankton are easily overlooked, owing to
their small size and ephemeral nature. However, these
organisms occupy almost 362 × 1012 m2 (Schlesinger
1997) of the Earth’s surface and consequently their presence in the open ocean accounts for almost half of the
Earth’s photosynthesis (table 6.2). A simple version of the
carbon cycle within the oceans is shown in figure 6.2.
Phytoplankton production takes place in the upper
100 m of the ocean, where enough sunlight and nutrients
delivered by rivers are available for phytoplankton to
reduce carbon dioxide to form carbohydrate while splitting a mole of water and releasing a mole of oxygen:
sunlight
6CO2 + 6H2O → H2O + O2.
chlorophyll
The most productive regions of the oceans are upwelling areas (table 6.2), where cold nutrient-rich bottom
waters are brought to the surface and warm surface
waters are pushed offshore by onshore winds. Coastal
zones are also highly productive (table 6.2). Although
the open ocean has the lowest mean production, its vast
area is one to two orders of magnitude greater than the
coastal and upwelling areas. Consequently, the open
ocean constitutes 42% of the total global production,
whereas coastal and upwelling zones make up only 9%
and <1% of the global total, respectively. Oceanographers refer to the total mass of photosynthetic organisms
at the ocean’s surface as primary production.
FATE OF PRIMARY PRODUCTION:
DEGRADATION AND DIAGENESIS
Regardless of where the organic matter is produced—in
large terrestrial forests, the soil reservoir, or lacustrine
or marine environments—it is subject to degradation
during deposition and to chemical changes after deposition (diagenesis), as we discussed in chapter 5. In the
ocean water column, most marine primary production
is consumed by animal plankton (zooplankton) and free-
TABLE 6.2. Estimates of Marine Primary Production
Province
Open ocean
Coastal zone
Upwelling area
Percentage
of Ocean
Area
(1012 m2)
Mean Production
(g C m−2 yr−1)
Total Global Production
(1015 g C m−2 yr−1)
90
9.9
<0.1
326
36
0.36
130
250
420
42
9.0
0.15
After Hedges and Keil (1995).
(6.1)
Organic Matter and Biomarkers: A Different Perspective
97
FIG. 6.2. Schematic depicting the fate of organic matter in the oceans. Boxes represent organic
reservoirs and arrows indicate pathways and processes.
floating bacteria. The bacteria also decompose a large
fraction of dissolved organic carbon and organic colloids produced by phytoplankton. Cole and coworkers
(1988) found that net bacterial production is about
twice that of zooplankton and accounts for the disappearance of 30% of primary production from the
photic zone and, in some areas, as much as 70% of primary production may be degraded. Bacterial communities are important to the biogeochemical cycling of
nutrients and organic matter in the oceans. The organic
matter produced by phytoplankton may be consumed
by zooplankton and eventually make its way up the food
chain to higher trophic level organisms, such as fish.
Alternatively, phytoplankton may be consumed by bacteria, which, in turn, may be consumed by bacteriovores.
This process ultimately mineralizes nutrients and releases
CO2 back to the surface waters. If the bacterial abundance
is high, then a large fraction of the carbon fixed during
photosynthesis is not passed to higher trophic levels. In
chapter 8, we confirm that between 80 and 90% of the
primary production is degraded to inorganic compounds
(CO2, NO3, PO2, and the like) in surface waters and the
remainder sinks below the euphotic zone (the depth to
which light penetrates, where most primary production
occurs) and into the deep ocean as particulate organic
matter.
Most of the organic matter sequestered in marine
sediments is ultimately derived from organic matter
synthesized by marine organisms inhabiting the surface
waters of the oceans and transported to the seafloor
(fig. 6.2). However, only a small fraction of the sinking
organic matter survives transport to the seafloor to be
preserved in the sediments. Extensive alteration of the
organic matter in the water column can yield sedimentary
98
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
organic matter having a chemical composition markedly
different from that of the original material. Despite this
alteration, a small fraction of particulate and dissolved
organic material eventually reaches the sediment-water
interface. There, the organic matter either undergoes
further degradation by microbial communities or is preserved as kerogen and possibly petroleum if conditions
are favorable.
FACTORS CONTROLLING
ACCUMULATION AND PRESERVATION
In general, the preservation of organic materials depends
on a complex interaction between the oxygen content of
a system and the type of organic matter deposited within
it (fig. 6.3). In the presence of oxygen, organic matter
can be remineralized or converted to CO2. However, if
this material is protected in some manner or is deposited
in a suboxic or anoxic environment, it is likely to be preserved (fig. 6.3). Preservation is also enhanced in sediments underlying highly productive surface waters. In
these environments, production of organic matter is much
greater than its oxic degradation, so that much more
organic matter reaches the sediments and is preserved.
Although we have identified the key variables that influence organic matter preservation, you should realize that
the mechanisms governing this process remain unclear.
In this section, we address some of the mechanisms proposed to explain how organic matter is preserved in the
sedimentary record.
Preservation by Sorption
Most organic matter in marine environments is concentrated in deltaic and continental shelf/slope sediments,
whereas lower organic carbon (OC) concentrations are
associated with high productivity areas and anoxic basins
(table 6.3). Despite differences in OC concentration and
key variables that influence organic matter preservation,
a common mechanism appears to control organic matter preservation in these environments. More than 90%
of the sedimentary OC from marine environments
cannot be physically removed from its mineral matrix
(Hedges and Keil 1995). Consequently, organic matter
must be sorbed to mineral surfaces. The concentration
of sorbed organic matter covaries with mineral surface
area, such that more OC is associated with sediments that
have highly irregular surfaces with small pores (Mayer
1994). Sorption of organic matter forms protective
coatings that typically approach a single molecule-thick
covering (often expressed as a monolayer equivalent,
∼0.5–1.0 mg OC m−2; Hedges and Keil 1995). These
coatings contain both refractory (not easily degraded)
and labile (easily degraded) organic matter that is pre-
FIG. 6.3. The fate of organic matter in different redox environments. Arrows show pathways along
which organic matter is produced or altered to various byproducts by different processes.
Organic Matter and Biomarkers: A Different Perspective
TABLE 6.3. Organic Carbon Preservation
in Various Marine Environments
Sediment Type
Deltaic sediments
Shelves and upper slopes
High-productivity slope and pelagic zone
Shallow-water shelf carbonates
Low-productivity pelagic sediments
Anoxic basins (e.g., Black Sea)
Organic Content
(1012 g C yr−1)
70
68
10
6
5
1
From Hedges and Keil (1995).
served in the underlying anoxic sediments, because it is
protected from mineralization during transport through
the water column and oxygenated surface sediments.
Formation of monolayer equivalents of organic matter
on marine sediments implies that, at one time, the organic matter was in the dissolved phase, because it is
unlikely that particulate organic matter would spread
uniformly over mineral surfaces.
Surface area is the primary control on organic preservation within continental shelf and slope environments,
despite differences between these settings in primary production, bottom water oxygen concentration, sediment
accumulation rate, and water depth (Keil et al. 1994;
Mayer 1994). This control plays an important role in
global organic matter preservation, because continental
shelf and slope environments account for ∼45% of the
Earth’s carbon burial (table 6.3).
Organic matter sorption is also the key mechanism
in deltaic sediments, which account for another 45% of
the total preserved OC, but the coatings on deltaic sediments are typically less than monolayer equivalents.
Some of the sorbed OC to river sediments is removed
when the sediments are deposited in a marine environment. The exact mechanism that causes this loss of
coatings is unclear, but it may be related to desorption
by seawater or direct oxidation (discussed later) on the
mineral surface during sediment transport to the marine
environment.
Sediments underlying highly productive O2-poor
waters are typically enriched in organic matter, but these
marine environments are rare and constitute only ∼5%
of the global organic preservation. Monolayer equivalent sorption cannot explain the high organic matter
concentrations (>5%) observed in these sediments. Less
than 15% of the OC in these sediments can be separated
from the mineral matrix, so these mineral surfaces must
have coatings equivalent to several monolayers.
99
Many factors may lead to the formation of these
thicker coatings. Among these are elevated DOC in sediment porewaters, condensation reactions, the presence
of sulfides, or very brief O2 exposure times. Elevated
DOC concentrations may lead to increased organic
sorption to the mineral surfaces. DOC may also form
organic-bearing sulfides that are resistant to degradation. Degraded biomolecules may combine to form
complex high-molecular-weight substances (condensation reactions). The coating that these substances produce is most likely resistant to microbial degradation
and strongly bound to the mineral surface. In addition,
the coating may promote further condensation reactions,
thereby protecting the organic matter sorbed to the
mineral surface. Rapid sedimentation rates in highly
productive areas may enhance preservation, as the O2
exposure time may be limited. This limited exposure may
permit the preservation of labile organic matter and
enhance the formation of resistant substances through
condensation.
Degradation in Oxic Environments
Organic matter preservation in deep-sea sediments
is typically poor (table 6.3), which suggests that some
degradation mechanism may overcome surface area protection. Oceanic sediments have coatings with less mass
than monolayer-equivalents, potentially formed from
direct oxidation during transport. Slow accumulation
rates and oxygenated waters in deep-sea environments
give rise to long O2 exposure times and may inhibit organic preservation.
The most direct evidence of this oxygen degradation
is seen in “oxidation fronts” in deep-sea turbidites, where
slumping exposes sedimentary organic matter to O2-rich
bottom waters for long periods of time. Molecular O2
reacts with OC and reduced minerals along sharp redox
fronts, which can reduce the organic matter content by
up to 75%. Oxidation continues until the deposit is
covered by either another turbidite flow or by gradual
pelagic sedimentation (Hedges and Keil 1995).
Diagenetic Alteration
We have seen how organic matter is produced and
where it is likely to accumulate and be preserved. Now
we examine the transformations that organic matter
undergo during diagenesis, as described in chapter 5.
100
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Microbial communities and other biologic agents primarily control diagenetic transformations, although some
chemical transformations catalyzed by mineral surfaces
do occur.
During diagenesis, sediments undergo compaction and
consolidation, with a simultaneous decrease in water
content and increase in temperature. Biologic alteration of OC eventually ceases as temperatures increase
(>50°C) during the later stages of burial, a process
called catagenesis. The boundary between diagenesis and
catagenesis is not well defined but it essentially coincides
with the onset of oil formation.
As we have seen, degradation of OC begins in the
water column and continues after sedimentation (fig. 6.2).
Different compounds in the organic matter degrade at
different rates and only some compounds survive in a
recognizable form. We examine this aspect of degradation more closely in the next section. During diagenesis,
residual organic matter from microbial degradation undergoes condensation to yield macromolecules of insoluble brown organic material. The result of diagenesis is
a condensed organic residue, or geopolymer, which contains varying amounts of largely unaltered refractory
organic material. In soil environments, diagenesis yields
humin, brown coal (or lignite) in coal-forming swamps,
and kerogen in marine and deep lacustrine sediments.
Humic substances are found in soils, terrestrial and
marine sediments, coal deposits, and all aquatic environments. Humics can account for almost all the organic
carbon in freshwater environments, giving waters a
brown color or possibly a green tint. Humin in soils is
derived from terrestrial plants that decompose, forming
OC that condenses during diagenesis to form the insoluble humic residue.
Coals may also form during diagenesis of organic
matter, and the type of coal formed typically reflects its
precursor plant material. For example, humic coals are
formed mainly from the woody tissues of vascular plant
remains. These coals are typically stratified, have a lustrous, black/dark brown appearance, and go through a
peat-forming stage. Sapropelic coals, on the other hand,
are not stratified and are dull in appearance. They form
in quiet, anoxic shallow waters from fine-grained, organic-rich sediments that contain varying amounts of
algal remains and degraded peat swamp plants. Unlike
humic coals, formation of sapropelic coals does not
typically involve a peat-forming stage. Instead, these
coals follow a pathway similar to kerogens, which we
discuss next.
Coals typically form as a result of two main processes:
peatification, dominated largely by biologic activity, followed by coalification, in which heat and pressure are
the most important agents of change. Peatification and
early coalification are equivalent to diagenesis. Late-stage
coalification, which is essentially catagenesis, parallels
the onset of metamorphism. The boundary between the
early and late stages of coalification is not sharp, because
the actions of the biologic and physicochemical agents
may overlap. The principal stages of coal formation begin
with peat formation followed by the formation brown
coal, bituminous coal, and eventually anthracite.
Kerogen is the highly complex organic material from
which hydrocarbons are produced as the material is
subject to increasing burial and heating. It is by far the
most abundant form of organic carbon in the Earth’s
crust and it occurs primarily in sedimentary rocks as
finely disseminated organic macerals (chunks of organic
matter, analogous to minerals in rocks). Until recently,
kerogen was thought to form during the late stages of diagenesis from the alteration of huminlike material. This
material was supposedly derived from the condensation
of insoluble geopolymers, which formed from microbial
degradation of organic matter. Kerogen is now thought to
form during the early stages of diagenesis from mixtures
of partially altered, refractory biomolecules (Tregelaar
et al. 1989).
Kerogen is modified by temperature ultimately to yield
petroleum hydrocarbons. The potential yield of petroleum products typically depends on the type of kerogen.
There are four types of kerogens distinguished by their
maceral groups: lignite, exinite, vitrinite, and inertinite.
Type I kerogens (lignite), which are derived from algal
or bacterial remains, are relatively rare but have a high
oil potential. These materials formed in fine-grained,
organic-rich muds deposited under anoxic conditions in
quiet, shallow water environments, such as lagoons and
lakes. Type II kerogens (exinite) are the most common
and are usually formed in marine environments from
mixtures of phytoplankton, zooplankton, and microbial
organic matter under reducing conditions, but they can
also be formed from higher plant debris. Type II kerogen
has a lower yield of hydrocarbons than Type I but it has
still produced oil shales of commercial value and sourced
a large number of oil and gas fields (Killops and Killops
Organic Matter and Biomarkers: A Different Perspective
101
A CRASH COURSE IN ORGANIC NOMENCLATURE
You need some knowledge of organic chemistry to
understand the principal aims, methods, and results
of organic geochemical investigations, so let’s briefly
examine some geochemically important structures
and outline their nomenclature.
Organic compounds consist principally of carbon
atoms linked to nitrogen, oxygen, sulfur, and other
carbon atoms. The basic carbon skeleton can be
arranged in simple straight chains, branched chains,
one or more rings, or combinations of these structures. The simplest organic compounds are hydrocarbons, which consist solely of carbon and hydrogen
atoms. Carbon atoms have four valence electrons
that must be satisfied by either single covalent bonds
(see chapter 2) to four separate atoms or some combination of single and double bonds. A double bond
is one that involves two valence electrons. Hydrocarbons that have only single bonds are saturated,
whereas those that contain double bonds are referred
to as unsaturated.
The carbon atoms can be linked together to form
either straight-chain (aliphatic) or simple cyclical (alicyclic) structures. Saturated aliphatic hydrocarbons
are alkanes, where the suffix -ane refers to the lack
of any double bonds in the structure. An unsaturated
aliphatic hydrocarbon is called an alkene, where the
suffix -ene denotes a double bond. Alkanes (and alkenes) are named according to the number of carbon
atoms in the chain: methane (1), ethane (2), propane
(3), butane (4), pentane (5), hexane (6), heptane (7),
octane (8), and so on. These compounds are typically
represented by two-dimensional (“sawtooth”) drawings, in which the kink at each line segment represents
a carbon atom. Leftover valence electrons (that is,
ones not involved in bonding carbon atoms to each
other) attach hydrogen atoms to the structure. For example, figure 6.4a shows dodecane, a straight chain,
aliphatic hydrocarbon with 12 carbon atoms and no
double bonds. It is understood that two hydrogen
atoms attach to the carbon atom at each kink in the
chain, and three attach at each end of the chain. The
compound shown in figure 6.4b, dodecene, is essentially the same compound shown in figure 6.4a with
one double bond. The carbon atoms at each end of
the double bond would have only one attached hydrogen atom.
Stable configurations of carbon atoms also occur
where double (C=C) bonds alternate with single
(CC) bonds to form a pattern like ∼CC=CC=
CC=C∼. This configuration, referred to as conjugated, is common in polyunsaturated compounds.
If conjugation occurs in a ring structure with three
double and three single bonds (fig. 6.4c), the configuration is aromatic. Aromatic compounds are unsaturated and are typically very stable.
The groups of elements bonded to the carbon backbone (whether chain or ring) are called functional
groups, because they are usually reactive and thus
influence the chemical behavior of the compound.
Functional groups involve combinations of hydrogen
atoms with one or more atoms of another element,
most commonly oxygen, nitrogen, or sulfur. Some
of these, like hydroxyl (−OH), are familiar from inorganic chemistry. Alcohols and phenols are based
FIG. 6.4. Structural representations of some organic molecules. (a) Dodecane; (b) dodecene; (c) benzene; (d) phenol;
(e) L- and D-enantiomers of isoleucine.
102
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
TABLE 6.4. Some Functional Groups in Organic Molecules
Symbol1
Function Group
ROH
Hydroxyl
Alcohol (R = aliphatic group)
Phenol (R = aromatic group)
C O
R
Carbonyl
Aldehyde (R = H)
Ketone (R = aliphatic group)
Alkanone (R = aromatic group)
C O
OH
Carboxyl
Acid
NH2
Amino
Amine
Pyrryl
Pyrrole
⁄⁄
HCC ⁄
⁄⁄⁄
⁄⁄
CH
HC ⁄⁄
N
H
Compounds
1
R denotes groups of atoms that may be joined to the function group.
on combinations of −OH with aliphatic or aromatic
groups, respectively (see the phenol in fig. 6.4d). Carbohydrates, such as sugars and starches, are a major
class of aliphatic compounds with a large number
of hydroxyl groups in place of hydrogen atoms. Organic compounds are often grouped under the name
of the functional group that they contain. For example, carboxylic acids are compounds that bear a
carboxyl group. Some geochemically important functional groups are listed in table 6.4.
Although we typically represent organic compounds with two-dimensional drawings, these compounds actually have three-dimensional structures.
Different spatial configurations of atoms attached to
a central carbon atom can exist. This phenomenon is
called stereoisomerism. Several compound classes
exhibit this property, in which two structurally identical compounds exist as mirror images that cannot
be superimposed. Your hands provide a straightforward illustration of this property: each hand has the
same structure of four fingers and a thumb, but you
1993). Type III kerogen (vitrinite) is derived from vascular plants and may contain identifiable woody plant
debris. Vitrinite is usually not extensively altered by
microbial degradation, which may be a result of the
material’s rapid sedimentation and burial. Vitrinite gen-
cannot superimpose them (palms facing up). Similarly, certain amino acids and carbohydrates have
identical structures but differ by the placement of
their functional groups on one side of the central carbon, called the chiral carbon. These mirror images,
called enantiomers, have similar physical properties
but rotate light in different directions. Compounds
that rotate light clockwise are called dextrorotatory
and are denoted by the prefix d-, whereas counterclockwise rotation is referred to as levorotatory and
denoted by the prefix l-. For example, the amino acid
isoleucine can exist in either l-or d-forms (fig. 6.4e).
The only difference between these two compounds is
the placement of an amino group, which is on opposite sides of the central carbon atoms. (As you can
see, isoleucine has two chiral carbon atoms, which is
a special case. When the l-form of this compound
changes to its d-form, this process is called epimerization and the d-form is referred to as alloisoleucine.
You will encounter an example of this in worked
problem 6.3.)
erally occurs as coals or coaly shales; therefore, it is
similar to coal in terms of its composition and behavior
with increasing burial. Type IV kerogen (inertinite) is
composed of primarily black opaque debris potentially
formed from the deposition of highly oxidized, higher
Organic Matter and Biomarkers: A Different Perspective
plant material. It has no hydrocarbon generating potential and is sometimes not considered a true kerogen.
CHEMICAL COMPOSITION OF
BIOLOGIC PRECURSORS
Despite all the potential changes that may alter organic
matter along various pathways from production to preservation, some organic compounds still reflect their original source materials, whether those are higher plants,
marine or lacustrine algae, or bacteria. It is important to
understand the chemical composition of these sources so
that we can relate these compounds to those preserved
in the sedimentary record.
All organisms are composed of compounds derived
from carbohydrates, proteins, and lipids. In addition, significant amounts of lignin (structural tissues) are present
in higher plants. Let’s now review the composition of
the main chemical classes of compounds and their biochemical functions in living organisms. The structural
representations and important functional groups for these
compounds are discussed in an accompanying box.
Carbohydrates
The carbohydrates are aliphatic compounds containing only carbon, hydrogen, and oxygen, with a ratio
of hydrogen to oxygen similar to water. Their general
formula is represented by (Cn(H2O)n). Organisms use
carbohydrates as food reserves and structural materials.
Polysaccharides (many saccharide units) are major components of cell membranes in plants, bacteria, and fungi.
Cellulose, the main structural unit in higher plants, is
the most common carbohydrate. Cellulose is composed
of roughly 10,000 glucose (C6 monosaccharide) units that
are, in essence, energy reserves. This energy is stored as
starch, which is a component of the cellulose. Energy is
derived from this reserve by breaking down the polysaccharides into individual glucose units.
Hemicellulose compounds are the next most abundant group of carbohydrates. These compounds are
complex mixtures of polysaccharides containing between
50 and 2000 monosaccharide units that form a matrix
around the cellulose fibers in plant cell walls. These
compounds are present in nonwoody tissues and fruits
but are only a minor component in woody tissues.
Whereas cellulose is the dominant structural material
in plants, the polysaccharide chitin is the principal struc-
103
tural compounds in most fungi, some algae, mollusks,
and arthropods (insects and crustaceans). This compound
is composed of chains of glucose units with an attached
amino (NH2) functional group. Eubacterial cell walls
contain murein, another polysaccharide. It is possible
to distinguish some eubacteria based on the compounds
that are contained on the exterior of their cell walls.
Similarly, other carbohydrates (specifically enantiomers; see accompanying box) may be characteristic of,
but not necessarily exclusive to, different organisms.
Fungal polysaccharides are predominantly d-glucose,
d-galactose, and d-mannose, whereas marine algae
such as Chlorophyta contain a large proportion of lrhamanose, and Phaeophyta are rich in d-ribose. Freshwater algae and higher aquatic plants, however, contain
significant amounts of l-arabinose and d-xylose (Killops
and Killops 1993).
Proteins
Proteins, which consist of linked amino acids, are the
principal storage sites for nitrogen in organisms. The
amino acids contain an amino group (NH2) and a carboxyl group (COOH) that can bond through the elimination of water to form peptide linkages. Structurally,
all amino acids are chiral molecules with the exception
of glycine, which has two hydrogen atoms bonded to
the chiral carbon atom. There are only 20 amino acids
involved in protein synthesis and the l-configuration of
these compounds is utilized exclusively, owing to the
stereospecificity of enzymes in all organisms except
bacteria. The predominant use of d-amino acids by bacteria may help identify the source of these compounds
in complex mixtures, provided that diagenesis has not
occurred.
Plants are able to synthesize all the amino acids necessary for biomass production, and most amino acids are
derived from glutamic acid. Animals, on the other hand,
are not able to synthesize all the amino acids required
for protein synthesis; therefore, their essential amino
acids are supplied by eating plants.
Proteins typically constitute a substantial portion of
the bulk OC in an organism. Owing to their ability to
form fibers, proteins provide supportive tissues such as
skin and bone (collagen) or hooves and claws (keratin)
in animals. The function of these protein-derived structural materials is analogous to that of cellulose and lignin
materials in plants.
104
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Lipids
Lipids are those compounds that are insoluble in water
but are extractable in solvents known to dissolve fats
(such as chloroform, hexane, toluene, and acetone). Most
organic geochemists use this broad definition of lipids,
because it encompasses all the compound classes of geologic importance. The most abundant and geochemically important lipids (such as those that are principal
components of oil source rocks) include the following
groups.
Glycerides are esters of the alcohol glycerol that can
bond with one, two, or three carboxylic acids (-COOH)
to form mono-, di-, and triglycerides, respectively. Among
the triglycerides, fats are important compounds.
Fats are composed of fatty acids, which are straightchain carboxylic acids (RCOOH) that typically contain
between 12 and 36 carbon atoms. Saturated fatty acids
are found in animals, whereas unsaturated and polyunsaturated fatty acids are common in plants. Unsaturated acids have a lower melting point for a given chain
length, which is why some unsaturated plant-derived
fats are oils whereas saturated animal fats are solids. The
C16 and C18 saturated fatty acids are most abundant in
animals, whereas the C18 mono-, di-, and triunsaturated
fatty acids are the dominant forms in plant tissues.
Polyunsaturated fatty acids are more common in algae
than in higher plants. Fatty acids are predominantly evennumbered carbon compounds, because their biosynthetic
pathway incorporates two carbon atoms (acetyl units)
that are derived from glucose. Similar to carbohydrates,
fats are used as an energy reserve by animals and plants,
but the principal difference is that fats provide twice the
energy as carbohydrates during oxidation. This is particularly useful during processes that require a substantial
amount of energy, such as seed and fruit generation.
Waxes serve mostly to protect organism membranes,
as in, for example, the waxes that coat plant leaves. These
biochemicals are mixtures of fatty acid esters and fatty
alcohols, whose carbon chains range from C24 to C28.
Waxes are mostly even-carbon–numbered owing to their
synthesis from fatty acids (also even-carbon numbered).
Hydrocarbons, principally long-chain n-alkanes, are also
a major constituent of waxes, but these compounds
contain predominantly odd numbers of carbon atoms in
the range of C23 to C33 with the majority of these being
C27, C29 and C31. The odd-number preference arises
because n-alkanes are formed from an aliphatic acid
(CH3(CH2)nCH2COOH) by the elimination of the carboxyl group (COOH), which contains one carbon atom.
Higher plants and fungi contain a similar distribution of
n-alkanes, whereas these compounds are almost absent
in most bacteria.
Terpenoids are a class of lipids whose structures and
functions vary markedly from the basic building block
of chlorophyll and gums of higher plants to volatile sex
pheromones. All the different terpenoids, however, have
a similar structural component made of five carbon atoms
called an isoprene unit. Terpenoids are classified by the
number of constituent isoprene units.
Monoterpenoids consist of two isoprene units (hence,
they are C10 compounds) and some are highly volatile.
They are usually found in algae and constitute the essential oils of higher plants. These oils, like citronella or
menthol in peppermint, give the plant a specific odor.
Due to their volatility, some of these compounds, such
as insect pheromones, act as attractants. Others, such as
chrysanthemic acid in pyrethrum flower heads, are natural insecticides.
Sesquiterpenoids (C15 compounds), formed from three
isoprene units, function not only as the essential oils of
higher plants but also as fungal antibiotics. Farnesol is a
common acyclic compound found in many plants and in
the chlorophyll of some bacteria.
Diterpenoids (C20 compounds) contain four isoprene
units and form resins, which seal damaged tissues with
protective coatings and inhibit insect and animal attack.
Other compounds in this class cause some bitter taste in
plants, which can also act as a protective mechanism.
Triterpenoids (C30) contain six isoprene units and are
believed to develop from squalene, which is a ubiquitous
30-carbon compound in many organisms. Most of the
triterpenoids are either tetracyclic (four rings) or pentacyclic (five rings). The pentacyclic triterpenoids typically
constitute the resins of higher plants, but another group
of these compounds (hopanoids) is found in bacteria.
In addition, other triterpenoids have been identified as
the precursors to certain petroleum hydrocarbons. Although you may be not familiar with the majority of
these compounds, you probably recognize the tetracyclic
triterpenoids known as steroids (or sterols). These compounds contain four rings in their structure, which result
from the oxidation of squalene followed by cyclization
(ring formation). This reaction produces two different
compounds that are the precursors to all plant (cycloartenol) and animal (lanosterol) sterols. Oxidation of
Organic Matter and Biomarkers: A Different Perspective
lanosterol produces cholesterol (C27 compound), which
is the precursor to all other animal steroids. The majority of the sterols that are found in the geologic record
range between C27 and C30. Sterols are rare in bacteria,
but hopanoids are abundant, which may provide a means
to distinguish the source of organic matter in sedimentary environments.
Tetraterpenoids (C40) consist of eight isoprene units.
The most important group of these compounds is the
carotenoid pigments. Carotenoids are essential for the
production of vitamin A and are found most organisms.
Carotenoid pigments are responsible for the red coloration in “red tide” blooms of dinoflagellates. Carotenes
and xanthophylls (from the Greek for contain oxygen)
are highly unsaturated and can absorb relatively short
wavelengths. These compounds are major photosynthetic
pigments that aid in the capture of sunlight energy and
transfer this energy to chlorophyll. Certain carotenoids
are characteristic of different photosynthetic organisms.
Tetrapyrrole pigments consist of four pyrrole units (a
five-member ring structure, in which four of the atoms
are carbon and the fifth member is a NH group linked by
a double-bonded CH group). These structures can either
be a ring or large open chain. The pigments are involved
in photosynthesis as either primary or accessory pigments. Chlorophylls are members of this group that have
a large ring structure called a porphyrin. Chlorophyll is
found in all algae, higher plants, and cyanobacteria.
Lignin
Lignin, characterized by phenolic structures (aromatic
six-member carbon rings bonded to OH groups, illustrated in fig. 6.4d), is a major, unique component of the
structural tissues of higher plants. These compounds are
derived from monosaccharides. Lignin forms around the
fibrous woody core (xylem) of terrestrial plants and acts
as a support structure as the plant grows. Cellulose
constitutes up to 40–60% of wood and lignin essentially
makes up the remainder. This high-molecular-weight
material has a highly complex structure consisting of
many different types of alcohols that undergo condensation reactions to ultimately form lignin.
BIOMARKERS
Now that we are familiar with some of the biochemically
important compounds, we can use these as biomarkers to
105
gain insight into the types of organisms that contribute
to the total organic matter in sedimentary environments.
Biomarkers survive diagenesis almost intact or their diagenetic pathways can be traced, whereas other organic
compounds are altered and may not necessarily reflect
their original source. Here we investigate the role of
biomarkers as source indicators in sedimentary environments, using lignin and certain lipids as examples.
There are many differences in the biochemical compositions of the principal groups of organisms. For example, the presence of lignin in sediments may indicate
a higher plant contribution to the total organic matter of
a sedimentary environment, because, as we have seen,
lignin is present only in higher plants. We have also seen
that higher plants contain protective waxes, in which
even-carbon-number fatty acids predominate, on their
leaves. The predominance of even- over odd-carbonnumber fatty acids in organic residue, then, is a further
biomarker for higher plants. Hydrocarbons (n-alkanes)
are also a major constituent of the waxes, but these compounds contain predominantly odd numbers of carbon
atoms, so an odd-over-even predominance of n-alkanes
also suggests input from higher plants.
Compounds derived from the steroid class can also be
indicative of certain groups of organisms. Phytoplankton primarily consist of abundant C28 sterols, whereas
zooplankton typically contain C27 sterols, particularly
cholesterol. In contrast, major plants are dominated by
the C29 sterols, and C27–C29 sterols are typically associated with fungi.
Lignin and its associated phenolic units can be used
to differentiate among the types of vascular plants that
contribute organic matter in complex sedimentary environments. To do this in the laboratory, a geochemist
liberates the phenolic units from the lignin structure
through an oxidation process. Oxidation produces
three groups of structurally related compound groups,
vanillyl (V), syringyl (S), and cinnamyl (C), which are
diagnostic of lignin. These compound groups are absent
from nonvascular plants and vary according to vascular
plant type (table 6.5), which makes source determination possible. For example, nonwoody angiosperms can
be distinguished from the other three groups of vascular
plants based on their S/V ratios, and woody angiosperms have a distinct C/V ratio. Another approach is
to plot the S/V ratios versus C/V ratios, which can also
provide a means to differentiate these sources. Let’s see
how this works with a real example.
106
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
TABLE 6.5. Distribution of Lignin Oxidation Products in Various Plant Tissues
Syringyl
(mg/g)
Plant Type
Nonvascular plants
Nonwoody angiosperms
Woody angiosperms
Nonwoody gymnosperms
Woody gymnosperms
1
Cinnamyl
(mg/g)
0
1–3
7–18
0
0
0
0.4–3.1
0
0.8–1.2
0
The syringyl, cinnamyl, and vanillyl concentrations have been
measured at different depths within a lacustrine sediment core.
The calculated S/V and C/V ratios are shown below. How can
we determine the source of organic matter that has accumulated in this lake basin over time?
0
2
4
6
8
10
12
0
0.6–3.0
2.7–8.0
1.9–2.1
4–13
S/V
1
—
0.5–2.5
0
0
0
C/V
—
0.4–1.1
1.1–5.2
0.4–0.6
0
—, not applicable.
Worked Problem 6.1
Sample Depth (m)
Vanillyl
(mg/g)
Age (kyr)
S/V
C/V
1.0
0.44
0.65
1.00
0.97
0.96
0.95
0.05
0.22
0.20
0.08
0.10
1.00
0.90
0.38
7.2
10.0
13.8
We can compare the S/V and C/V values with table 6.5 to
determine planet sources for each core sample and investigate
any vegetation changes. The basal portion of the core has low
S/V and C/V ratios, which suggests input from nonwoody gymnosperms, such as ferns. Between 8 and 10 m depth, both the
S/V and C/V ratios are high, reflecting a change to nonwoody
angiosperms (marsh or grassland plants) and possibly a tundra
landscape. High S/V and low C/V ratios between 4 and 6 m
indicate input from woody angiosperms, such as deciduous
trees, making another vegetation change to a boreal biome. A
decrease in S/V ratios from 4 m to the top of the core with relatively constant C/V ratios suggests a progressive enrichment in
woody gymnosperms (conifers), mostly likely at the expense of
the woody angiosperms (deciduous trees). Changes in the lignin
parameters (S/V and C/V) thus indicate that climate changed
from a fern-dominated landscape 13 kyr ago to tundra by 10 kyr.
Climate change forced the tundra to give way to boreal forest
by 7 kyr, which eventually gave way to a pine forest.
APPLICATION OF BIOMARKERS
TO PALEOENVIRONMENTAL
RECONSTRUCTIONS
Biomarkers can also be used to quantify specific environmental conditions. For example, the distributions of some
biomarkers are sensitive to temperature (because of the
metabolic pathways involved in their synthesis) and time
(because of the kinetics of their reactions).
Alkenone Temperature Records
One family of temperature-sensitive lipids is the (C37–
C39) unsaturated ketones or alkenones. These compounds
have a structure similar to long chain alkenes (remember, the -ene suffix denotes double bonds) and can have
between two and four double bonds. The number of
double bonds is a function of temperature. A higher
proportion of unsaturated compounds is associated with
colder temperatures. For C37:3 and C37:4, 37 indicates
the number of carbon atoms and 3 or 4 is the number of
double bonds, respectively. These unsaturated alkenones
are produced by the coccolithophore algae Emiliani
huxleyii, which first appeared in the geologic record
during the late Pleistocene approximately 250,000 years
ago (Killops and Killops 1993). By culturing these organisms at various temperatures, Simon Brassell and coworkers (1986) established a linear relationship between
the degree of unsaturation and growth temperature. This
K
index:
relationship is known as the U37
[C37:2] − [C37:4]
K = ———————————,
U37
[C37:2] + [C37:3] + [C37:4]
where the square brackets are the concentrations of the
di-, tri-, and tetra- unsaturated ketones. The di- and triunsaturated compounds are dominant in most sediments,
so this index can be simplified to:
[C37:2]
K′ = ————–——,
U37
[C37:2 + C37:3]
(6.2)
K can be related to the growth temperature
Values of U37
and hence sea surface temperature (SST) in the case of
marine sediments through the following relationship
(Prahl et al. 1988):
Organic Matter and Biomarkers: A Different Perspective
K′ = 0.034(SST) + 0.039.
U37
(6.3)
Equation 6.3 holds well for most oceans, but there are
some, such as the Southern Ocean, for which a slightly
different temperature calibration has been developed.
K index appears to
Regardless of the calibration, the U37
provide reasonable temperatures in the modern oceans
and consequently has been used to determine SST on
glacial/interglacial time scales.
Worked Problem 6.2
How can we determine changes in SST in the Pacific Ocean
over glacial/interglacial time scales at midlatitudes and in the
tropics? Oceanographers have recovered ocean sediment cores
that span the past 50,000 years, and we have determined the
K values from the concentrations of the di- and tri-unsaturated
U37
K
C37 alkenones for several radiocarbon dated samples. The U37
values for the tropical core for the last glacial maximum and the
Holocene are 0.925 and 0.99, respectively. In the midlatitudes,
K values are signifithe last glacial maximum and Holocene U37
cantly smaller at 0.53 and 0.65, respectively.
Using equation 6.3, the SST during the last glacial maximum
in the tropical Pacific Ocean was:
K′
(U37
− 0.039)/0.034 = (SST) = (0.925 − 0.039)/0.034
= 26.2°C.
and the Holocene temperature was:
(SST) = (0.99 − 0.039)/0.034 = 28.0°C.
This corresponds to a 1.8°C temperature increase from the
last glacial maximum, which is consistent with other ways of
estimating SST. If we examine the temperature changes in the
midlatitudes, we see that during the last glacial period,
(SST) = (0.53 − 0.039)/0.034 = 14.4°C,
and Holocene temperatures were on the order of:
(SST) = (0.65 − 0.039)/0.034 = 18.0°C,
which is ∼4°C warmer. This is also consistent with other proxy
records that suggest that midlatitude glacial oceans were much
colder than tropical oceans and that the glacial/Holocene temperature rise was more pronounced in the midlatitudes.
Amino Acid Racemization
We noted earlier that all living organisms (except bacteria) manufacture l-amino acids, rather than d-amino
acids. The chiral (l- versus d-) configuration of amino
acids is sensitive to both time and temperature. lamino acids convert or racemize (for one chiral carbon
107
atom and epimerize for two chiral carbon atoms) to their
d-amino acid equivalents over time after the proteins are
formed and effectively protected from the biologic processes of the organism. The change from l- to d-amino
acids continues until an equal mixture of the two forms
exists.
Bacteria preferentially produce d-amino acids. These
compounds are nearly ubiquitous in the biosphere and
can come into contact with a potential sample of preserved nonbacterial material, thereby introducing amino
acids not indigenous to the sample. To reduce the risk
of contamination, most organic geochemists prefer to
sample materials that are least likely to have been accessible to bacteria. They sample shells of marine organisms
(mollusks and foraminifera), land snails, and ostrich eggs,
because the amino acids in these materials are protected
from the surrounding environment owing to a layer of
calcium carbonate deposited over protein layers in the
shell.
Reliable ages and temperatures can be calculated
from the D/L values of a sample provided that a reasonable kinetic model is established to quantify the relationships among time, D/L ratios, and temperature. This
is typically accomplished through artificial heating experiments of age-dated fossil shells. The relationship
between the age of a sample and measured D/L value is
expressed as (McCoy 1987):
t = ln([1 + D/L]/[1−K′D/L]) − ln([1 + D0 /L0 ]/
[1−K′D0 /L0 ])/(1+K′)Ae−Ea/RT,
(6.4)
where t is the age of the sample, D/L is the ratio of d to
l enantiomers, K′ is 1/Keq (equilibrium constant ∼0.77
for alloisoleucine/isoleucine [A/I]), D0 /L0 is the D/L ratio
of the sample at t = 0 (∼0.01, owing to a small amount
of racemization during sample preparation), Ea (the activation energy in J mol−1), and A (the entropy factor)
are the Arrhenius parameters of the racemization reaction, which are determined experimentally. R is the gas
constant (1.9872 cal K−1 mol−1), and T is the effective
diagenetic temperature in K.
Application of this equation to generate numerical
age estimates is limited by the temperature dependence
of the racemization reaction and the variable temperature history of many sample sites. This variability leads
to imprecise and potentially inaccurate age dates. However, if samples are collected from different environments
with similar postdepositional temperature histories, then
differences in D/L ratios can be used to identify relative
108
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
age differences between samples. This approach is known
as aminostratigraphy.
Age estimates may also be made from an empirical
relationship between the measured D/L ratios and the
sample age. The D/L ratios of independently dated samples can be used to calibrate ages of other samples on the
basis of their D/L ratios, provided that all the samples
have had similar postdepositional temperature histories.
By solving equation 6.4 for temperature (T ), we can
estimate the effective diagenetic temperature of a sample
terms of its age, measured D/L values, and the Arrhenius
parameters (McCoy 1987):
T = −Ea/R ln(ln{([1 + D/L]/[1 − K′(D/L])
− ([1 + D/L]/[1 − K′D/L])}/At[1 + k′]).
(6.5)
This equation can be simplified for an interval of time
bracketed by two independently dated samples to:
T(t2−t1) = Ea/R ln(A/k),
(6.6)
where
k = (k2t2 − k1t1)/(t2 − t1)
= ln(c/b)/[(1 − K′)(t2 − t1)],
(6.7)
in which b = (1 + D1/L1)/(1 − K′D1/L1) and c = (1 + D2/L2)/
(1 − K′D2/L2). D1/L1 and D2/L2 refer to the D/L ratios
for the younger and older samples, respectively. The numerical ages of the younger and older samples are represented by t1 and t2, respectively.
Worked Problem 6.3
How can we determine the effective diagenetic temperature of
fossil shells deposited at different times following the last glacial
maximum? Eric Oches and coworkers (1996) collected fossil
gastropod shells from silt in the Mississippi Valley and measured
their D/L ratios. They also determined the Arrhenius parameters
(A = 1.658 × 1017, K′ for alle/ile, and Ea = 29,235 cal mol−1),
measured the equilibrium constant for isoleucine epimerization
(K′ = 0.77), and obtained radiocarbon dates for each sample.
Locality
Temp.
(oC)
D1/L1
(alle/ile)
Age
(yr)
D2/L2
(alle/ile)
Age
(yr)
Finley, Tenn.
Vicksburg, Miss.
Vicksburg, Miss.
Vicksburg, Miss.
Pond, Miss.
15.7
18.4
18.4
18.4
18.8
0.11
0.11
0.12
0.11
0.14
26,040
20,820
20,820
20,820
18,320
0.13
0.14
0.13
0.14
0.16
32,230
25,110
25,110
27,930
25,610
The values of b and c in equation 6.7, for the Finley, Tenn.,
samples, are therefore:
b = (1 + 0.11)/(1 − [0.77 × 0.11]) = 1.21,
c = (1 + 0.13)/(1 − [0.77 × 0.13]) = 1.25.
Substituting these values into equation 6.7, we can determine
the value of k:
k = ln(1.25/1.21)/[(1 + 0.77)(32,320 − 26,040)]
= 3.18 × 10−6.
Using this value in equation 6.6, we can find T:
T = (29,235 cal mol−1)/(1.9872 cal K−1 mol−1)ln(1.658
× 1017/3.18 × 10−6) = 279.7 K = 6.5oC.
Determining the temperatures for all sample sites using
this procedure (see below), we find that temperatures between
20,000 and 30,000 years ago were much cooler than the mean
annual temperature (MAT) for all locations. These results agree
with the trend in midlatitude sea-surface temperatures that we
calculated in worked problem 6.2. These temperatures were
much cooler during the last glacial maximum than at present.
Locality
Finley, Tenn.
Vicksburg, Miss.
Vicksburg, Miss.
Vicksburg, Miss.
Pond, Miss.
MAT (oC)
Effective Diagenetic Temp. (oC)
15.7
18.4
18.4
18.4
18.8
6.5
12.3
6.4
9.5
7.2
SUMMARY
Although the field of organic geochemistry has its origins
in the study of petroleum, here we have concentrated on
the application of organic geochemistry to paleoenvironmental and paleoclimatic reconstructions. We have examined the origin of organic matter and how it is cycled
within the geosphere. The organic carbon reservoir is
only a small fraction of the total global carbon content
and represents only 8% of the active surface reservoirs.
Despite its relatively small size, this reservoir is intricately linked to the biosphere and geosphere.
We showed that the oceans are the primary source of
organic carbon in the sedimentary record and discussed
how organic matter is transformed and altered during
deposition and diagenesis. Most of the organic matter
produced in surface waters is completely recycled within
the water column and only a very small fraction gets preserved in sediments. Further alteration of this material
occurs during diagenesis, owing to microbial degradation
of labile organic matter. Degradation is typically followed
by condensation to form kerogen, which is ultimately
preserved in the sedimentary record.
Organic Matter and Biomarkers: A Different Perspective
We examined several mechanisms that may influence
the accumulation and preservation of organic matter in
marine environments. Preservation of organic matter is
controlled by sorption to mineral surfaces. Organic matter preservation may actually be a competition between
sorptive preservation and oxic degradation. We highlighted the geochemically relevant biologic compounds
to give you an idea of which compounds can be used to
identify inputs of organic matter from specific organisms.
The lipid and lignin compounds are potentially the most
useful biomarkers. Using alkenones, we were able to
reconstruct sea surface temperatures from midlatitude
and tropical regions of the Pacific over the past 35,000
years. We also examined how amino acid racemization
can be used as a paleoenvironmental tool.
There are many more questions that organic geochemists try to answer, not only in the petroleum and
paleoclimate fields but also from the modern environmental/contamination point of view. To gain insight
into these fields, you may consider investigating some
of the sources listed in the suggested readings for this
chapter.
suggested readings
Only a few books dedicated to organic geochemistry have been
written, and these are already dated. Most findings, especially
on the biomarker front, are disseminated in articles in the following technical journals: Organic Geochemistry, Geochimica
et Cosmochimica Acta, Marine Geochemistry, and Paleoceanography. For general background, you may wish to consult the
following texts.
Eglinton, G., and M. T. J. Murphy. 1969. Organic Geochemistry: Methods and Results. New York: Springer-Verlag. (This
book was the first definitive text in organic geochemistry
and still provides background information on the occurrence and evolution of organic matter in sediments.)
Engel, M. H., and S. A. Macko, eds. 1993. Organic Geochemistry: Principles and Applications. New York: Plenum. (This
text nicely shows applications of organic geochemistry to the
formation of petroleum, reconstructing paleoenvironments,
and understanding the fate of organic compounds.)
Killops, S. D., and V. J. Killops. 1993. An Introduction to Organic Geochemistry. New York: Wiley. (Top-of-the-line text
that introduces basic concepts of organic geochemistry in an
easily digestible manner.)
109
The following articles were referenced in this chapter.
Berner, R. A. 1989. Biogeochemical cycles of carbon and sulfur
and their effect on atmospheric oxygen over Phanerozoic
time. Palaeogeography, Palaeoclimatology, Palaecology 73:
97–122.
Brassell, S. C., G. Eglinton, I. T. Marlowe, U. Pflaumann, and
M. Sarnthein. 1986. Molecular stratigraphy: A new tool for
climatic assessment. Nature 320:129–133.
Cole, J. J., S. Findlay, and M. L. Pace. 1988. Bacterial production in fresh and saltwater ecosystems: A cross-system overview. Marine Ecology-Progress Series 43:1–10.
Emerson, S., and J. Hedges. 1988. Processes controlling the
organic carbon content of open ocean sediments. Paleoceanography 3:621–634.
Hedges, J. I., and R. G. Keil. 1995. Sedimentary organic-matter
preservation—an assessment and speculative synthesis. Marine Chemistry 492–493:81–115.
Keil, R. G., E. Tsamakis, C. B. Fuh, C. Giddings, and J. I. Hedges.
1994. Mineralogical and textural controls on the organic
composition of coastal marine sediments—hydrodynamic
separation using SPLITT-fractionation. Geochimica et Cosmochimica Acta 582:879–893.
Mayer, L. M. 1994. Surface area control of organic carbon accumulation in continental shelf sediments. Geochimica et
Cosmochimica Acta 114:347–363.
McCoy, W. D. 1987. The precision of amino acid geochronology and paleothermometry. Quantitative Science Review 6:
43–54.
Oches, E. A., W. D. McCoy, and P. U. Clark. 1996. Amino acid
estimates of latitudinal temperature gradients and geochronology of loess deposition during the last glaciation, Mississippi Valley, United States. Geological Society of America
Bulletin 1087:892–903.
Olson, J. S., R. M. Garrels, et al. 1985. The Natural Carbon
Cycle. Atmospheric Carbon Dioxide and the Global Carbon
Cycle. J. R. Trabalka, ed. Washington, D.C.: U.S. Department of Energy, pp. 175–213.
Prahl, F. G., L. A. Muehlhausen, and D. L. Zahnle. 1988. Further evaluation of long-chain alkenones as indicators of
paleoceanographic conditions. Geochemica et Cosmochemica Acta 52:2303–2310.
Schlesinger, W. H. 1997. Biochemistry: An Analysis of Global
Change. San Diego: Academic.
Tregelaar, E. W., J. de Leeuw, S. Derenne, and C. Largeau. 1989.
A reappraisal of kerogen formation. Geochimica et Cosmochimica Acta 53:3103–3106.
Williams, P. M., and E. R. M. Druffel. 1987. Radiocarbon in
dissolved organic matter in the central North Pacific Ocean.
Nature 330(6145):246–248.
110
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
PROBLEMS
(6.1)
Briefly discuss how the chemical compositions of higher plants differ from those of bacteria and
determine which biochemical compounds you would expect to survive diagenesis and why.
(6.2)
You recover a sediment core from a Miocene age lake deposit. Knowing that Miocene woody forests
were replaced by grasslands, discuss: (a) which organic compounds you would analyze to determine
the organic matter sources to sinkhole sediments; and (b) what you would expect to find if indeed
forests were replaced by grasslands.
(6.3)
Identify three areas of the oceans where monolayer equivalents of various masses would potentially
form, and discuss the key factors that control their formation in each environment.
(6.4)
Draw the structures of the following organic compounds: octane, pentene, ethanol (ethyl alcohol), a
phenol, and an alkenone with 37 carbon atoms and 3 double bonds (an unsaturated straight-chain
ketone).
(6.5)
K = 0.040(SST) − 0.110, recalculate the values in worked problem 6.2.
Using the relationship U37
What difference does this calibration make in the resulting values, and is this significant?
(6.6)
Using the information in worked problem 6.3, calculate the age of fossil snail shells that have alle/ile
ratios of: 0.09, 0.13, 0.19, 0.23, 0.35. Is this a linear relationship?
CHAPTER SEVEN
CHEMICAL WEATHERING
Dissolution and Redox Processes
OVERVIEW
The geochemical processes that most profoundly affect
the surface of continents—the environment with which
we have the closest personal familiarity—can be classed
together as chemical weathering processes. These include a variety of reactions involving water and either
acids or oxygen in the decomposition of rocks. In chapter 5, we referred to these implicitly as the source of
sediments and ionic species for diagenetic processes. In
this chapter, we examine them explicitly. Our focus is
on the weathering of silicates, which constitute the bulk
of continental rocks.
We begin the chapter by considering simple reactions
that dominate the low-temperature SiO2-MgO-Al2O3(Na2O, K2O)-water system, to show how these control
the stability of key minerals such as quartz, feldspars,
and clays on the Earth’s surface. We then look more
closely at the roles of organic compounds and CO2 in the
weathering process, and finish by examining oxidationreduction reactions. Throughout the chapter, we emphasize the graphical representation of equilibria, using
activity-activity diagrams.
The results of this theoretical treatment can be applied
to two closely related geochemical processes: the formation of residual minerals and the chemical evolution of
surficial and groundwaters. We introduce both of these
by example in this chapter and end the chapter by showing how our knowledge can also be applied to problems
of economic importance.
FUNDAMENTAL SOLUBILITY
EQUILIBRIA
Silica Solubility
Equilibria between water and silica are important in
two contexts. First, many sedimentary rocks and most
igneous and metamorphic rocks contain quartz. Chemical weathering of these involves the removal of SiO2
from a discrete phase. In the second, broader context,
because virtually all common crustal minerals contain
SiO2, our appreciation of the simple silica-water system is
important to understanding more general reactions with
silicates.
For any silica mineral, the primary reaction with
water is:
→ H SiO .
SiO2 + 2H2O ←
4
4
(7.1)
If the silica phase is quartz, the equilibrium constant Kq
at 25°C is 1 × 10−4. If it is opaline or amorphous quartz,
the equilibrium constant Kqa at 25°C is 2 × 10−3. Assuming that both water and the silica phase have unit activity, Kq (or Kqa ) is also numerically equal to the activity
111
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
of H4SiO4 that enters the solution as a result of reaction 7.1.
H4SiO4 is a weak acid, by which we mean that it
does not break down readily to release hydrogen ions in
aqueous solution. It dissociates first to form H3SiO4− by
the reaction:
→ H + + H SiO −,
H4SiO4 ←
3
4
(7.2)
for which:
K1 = aH+ aH3 SiO4− /aH4 SiO4 = 1.26 × 10−10 at 25°C,
and then to form H2SiO42− by the reaction:
→ H+ + H SiO 2−,
H3SiO4− ←
2
4
(7.3)
for which:
K2 = aH+ aH2SiO42− /aH3SiO4− = 5.01 × 10 −11 at 25°C.
No further dissociation takes place. With the exception of some polymerization reactions that occur only in
very alkaline solutions and have a negligible effect on
total solubility, the concentration of all silica species is
given by:
ΣSi = m
H4SiO4
+ mH3SiO4− + mH2SiO42−.
(7.4)
We can now combine equations 7.2 through 7.4 to write:
ΣSi = m
H4SiO4 (1
+ K1 /aH+ + K1K 2 /a2H+ ),
Log Activity of Dissolved Silica Species
112
FIG. 7.1. Activities of aqueous silica species at 25°C in equilibrium with quartz and with amorphous silica as a function of pH.
The lines are ΣSi. (Modified from Drever 1997.)
ever, so we will consider it in greater depth in chapter 8,
where we study the influence of CO2-H2O on seawater
chemistry.
(7.5)
in which all activity coefficients (γH SiO4− , γH2SiO42− , etc.)
3
have been omitted for simplicity. This equation is
shown graphically in figure 7.1, in which we follow
standard chemical convention by using the quantity pH
(= −log10aH+) as our measure of acidity. At pH values
below ∼9, only H4SiO4 contributes significantly to ΣSi.
At pH 9.9, however, the first and second terms in equation 7.5 are numerically equal, so mH4SiO4 = mH3SiO4− . In
increasingly alkaline solutions, mH4SiO4 becomes even
less important. At pH 11.7, according to equation 7.5,
mH3SiO4− = mH2SiO42− . The silica-water system, then, has
a pH buffering capacity in natural waters. By this, we
mean that pH is resistant to change when we add an acid
or base to a solution near either 9.9 or 11.7. At pH 9.9,
any added hydrogen ions are likely to be involved in
forcing reaction 7.2 to the left, producing H4SiO4 rather
than remaining as free H+ in solution. Similarly, at pH
11.7, “extra” H+ forces reaction 7.3 to produce H3SiO4−.
Because pH values as high as 9 are uncommon in nature, SiO2-H 2O buffering has only a small influence in
natural waters. Buffering is an important concept, how-
Solubility of Magnesian Silicates
Just as for silica solubility, it is convenient to write
reactions between magnesian silicates and water in such
a way that they eventually lead to solubility expressions
in terms of dissolved silica and some function of aH+. For
now, we consider only congruent reactions (that is, those
that produce only dissolved species) rather than those that
produce an intermediate hydrated phase. We disregard,
for example, the reaction:
→ 2Mg Si O (OH) ,
3Mg2SiO4 + 2H2O + H4SiO4 ←
3 2 5
4
forsterite
serpentine
and consider only:
→ 2Mg2+ + H SiO ,
Mg2SiO4 + 4H+ ←
4
4
for which:
2
4
2 2
Kfo = aH4SiO4aMg
2+ /a + = a
H
H4SiO4 (aMg2+ /a H+) ,
and
2
log aH4SiO4 = log Kfo − 2log(aMg2+ /a H
+).
Chemical Weathering: Dissolution and Redox Processes
One justification for ignoring incongruent reactions
is that reactions between magnesian silicates and water
in nature strongly favor total dissolution, so that little
magnesium is retained in solid phases after weathering.
Petit et al. (1987) have verified this experimentally by
means of a highly sensitive resonant nuclear imaging
technique. They found that dissolution of diopside
(CaMgSi2O6) takes place as water diffuses slowly into
the silicate to produce a hydrated zone ∼1000 Å thick.
This zone is very porous near the grain surface, but
shows no evidence of secondary precipitates. This observation strongly suggests that diopside dissolves congruently, and lends support to the notion that other
magnesian silicates may do so as well.
We can write many such congruent reactions. For
example, we have:
→ 3Mg2+ + 2H SiO + H O,
Mg3Si2O5(OH)4 + 6H + ←
4
4
2
serpentine
for which (assuming that aH
2O
= 1):
2
6
2
2 3
Kserp = a H
a3 2+ /a H
+ = a
H4SiO4(aMg2+ /a H+) ,
4SiO4 Mg
113
Finally, we consider the solubility of brucite, although it
is not a silicate, and find that:
→ Mg2+ + 2H O
Mg(OH)2 + 2H+ ←
2
brucite
yields:
Kbr = aMg2+ /a 2H+,
so that log aH SiO4 is independent of log(aMg2+ /a 2H+).
4
Notice how cleverly we have described each of these
equilibria as linear, first-order equations in log aH4SiO4
2
and log (aMg2+ /a H
+), so that they can be shown easily in
graphical form, as in figure 7.2. The solubility of quartz,
which is independent of log aH4SiO4 is also shown. We
have already shown that H4SiO4 dissociates appreciably
above pH 9, so this diagram is not reliable for highly
alkaline environments. Each of the magnesium silicate
equilibria plots as a straight line whose y intercept (at
infinitesimal H4SiO4 activity) is equal to a reaction K,
calculated from standard free energies of formation. Their
slopes in each case depend on the stoichiometry of the
and
2
log aH4SiO4 = 1–2 log Kserp − –23 log(aMg2+ /a H
+).
The similar reaction for talc,
→
Mg3Si4O10(OH)2 + 6H+ + 4H 2O ←
talc
3Mg2+ + 4H4SiO4 ,
yields:
4
4
K tc = a H
a 3 2+ /a6H+ = a H
(a 2+ /a 2H+)3
4SiO4 Mg
4SiO4 Mg
and
log a H4SiO4 = 1–4 log K tc − log(aMg2+ /a 2H+).
For sepiolite, the reaction:
→
Mg4Si6O15(OH)2 ⋅ 6H2O + 8H+ + H2O ←
sepiolite
4Mg2+ + 6H4SiO4
leads to the equations:
2 )4,
6
6
4
Ksep = a Mg
(a 2+ /a H
/a 8 = a H
+
2+ a
H4SiO4 H+
4 SiO4 Mg
and
2
log aH4SiO4 = 1–6 log Ksep − –23 log(aMg2+ /a H
+).
FIG. 7.2. Solubility relationships among magnesian silicates in
equilibrium with water at 25°C. Lines represent reactions discussed in the text. The dissolution of forsterite lies too far to the
right to appear on this diagram (see problem 7.7 at the end of
the chapter). (Modified from Drever 1997.)
114
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
solubility reaction (that is, on the Mg/Si ratio in the silicate mineral). The serpentine reaction, for example, has
a slope of − –32 , and talc that of − –34.
Notice that the dissolved species are stable on the
lower left side of each equilibrium line, so that each line
divides the diagram into a region in which fluid is supersaturated with respect to a silicate mineral (to the upper
right) and one in which it is undersaturated (to the lower
left). Therefore, the bold boundary of the solution field
shown in figure 7.2, combining segments of the brucite,
serpentine, talc, and quartz equilibria, represents the only
2
set of log aH4SiO4 and log(aMg2+ /a H
+) values for which fluid
and a mineral are stable. Sepiolite and forsterite, whose
reaction lines are entirely to the right of the boundary are
unstable (or metastable) in contact with water at 25°C.
Solubility of Gibbsite
Leaving the relatively simple realm of magnesium
silicate equilibria, let’s now try our luck with aluminosilicates. Dissolved aluminum in acid solutions exists as
the free ion Al3+. With increasing alkalinity, however,
Al3+ reacts with water to form ion pairs, which dominate
at pH >5. When we considered the solubility of magnesian silicates, there was no mention of complex species,
because they have little effect in that system. The ion pair
Mg(OH)+, for example, is abundant only for pH >11,
outside the range of acidity in most natural waters. By
contrast, the solubility relations of Al(OH)3 and aluminosilicate minerals are greatly influenced by the balance
among complex species, and therefore are more complicated than those of Mg(OH)2 and magnesium silicates.
Beginning in highly acid solutions, dissolution of
gibbsite is described by:
→ Al3+ + 3H O,
Al(OH)3 + 3H+ ←
2
for which:
K1 =
3 .
aAl3+ /a H
+
As pH increases beyond 5, Al(OH)+2 dominates instead.
The dissolution equilibrium is described by:
→ Al(OH) + + H O,
Al(OH)3 + H+ ←
2
2
and the constant:
K2 = aAl(OH)2+ /aH+ .
Finally, above pH 7, the major species in solution is
Al(OH)4−, which forms by:
→ Al(OH) − + H+,
Al(OH)3 + H2O ←
4
and for which:
K3 = aAl(OH)4− aH+.
Thus, to calculate the total concentration of dissolved
aluminum (ΣAl) in waters in equilibrium with gibbsite,
we write:
ΣAl = a
Al3+
+ aAl(OH)2+ + aAl(OH)4−,
and substitute appropriately from the equations for K1,
K2, and K3 to get:
3
1 H+
ΣAl = K a
+ K2 aH+ + K3 /aH+ .
(7.6)
Equation 7.6 differs from equation 7.5 (for ΣSi) in that it
predicts high solubility in both very alkaline and very acid
solutions, but low solubility throughout most of the range
of pH observed in streams and most groundwaters. To
see this quantitatively, let’s do a worked problem.
Worked Problem 7.1
Decomposition of organic matter in the uppermost centimeters
of a soil may decrease the pH of percolating water to 4 or lower.
As weathering reactions consume H+, we expect the dominant
form of dissolved aluminum and the total solubility, ΣAl, to
change. Ignoring any possible reactions involving the organic
compounds themselves, by how much should ΣAl vary between,
say, pH 4 and pH 6?
The three equilibrium constants in equation 7.6 can be
calculated from tabulated values of ∆Ḡf°. We use molar free
energies for water from Helgeson and coworkers (1978) and
for aluminum species at 25°C from Couturier and associates
(1984), except for Al(OH)2+, which is from Reesman and coworkers (1969):
Species
Al(OH)3
Al3+
Al(OH)2+
Al(OH)4−
H2O
∆Ḡf° (kcal mol−1)
−276.02
−117.0
−216.1
−313.4
−56.69
We calculate K1 from:
∆Ḡ1 = ∆Ḡf°(Al3+) + 3∆Ḡf°(H2O) − ∆Ḡf°(Al(OH)3)
= −117.0 + 3(−56.69) − (−276.02) = −11.05,
K1 = e−∆Ḡ1 /RT = 1.27 × 108,
Chemical Weathering: Dissolution and Redox Processes
Solubility of Aluminosilicate Minerals
K2 from:
∆Ḡ2 =
115
∆Ḡf°(Al(OH)2+ )
+ ∆Ḡf°(H2O) − ∆Ḡf°(Al(OH)3)
= −216.1 + (−56.69) − (−276.02) = 3.23
K2 = e−∆Ḡ2 /RT = 4.28 × 10−3,
and K3 from:
∆Ḡ3 = ∆Ḡf°(Al(OH)4− ) − ∆Ḡf°(Al(OH)3) − ∆Ḡf°(H2O)
= 311.4 − (−276.02) − (−56.69) = 21.31,
Σ
K3 = e∆Ḡ3 /RT = 2.34 × 10−16.
At pH 4, aH+ = 1 × 10 , so we calculate from equation 7.6
that:
−4
ΣAl = 1.27 ×
10−4
+ 4.28 ×
= 1.27 × 10−4,
In the same way that we constructed figure 7.2 to
illustrate the solubility of magnesium silicates, it would
be convenient to diagram the solubility behavior of aluminum silicates. Now that we have studied Al(OH)3 solubility, however, we see that this cannot be done easily.
For example, consider the dissolution of kaolinite. Over
the range of natural pH values, each of the following
three reactions contributes to Al:
10−7
+ 2.34 ×
10−12
and conclude that aAl3+/aAl(OH) + = 297. At pH 6 (aH+ = 1 × 10−6),
2
+ 4.28 × 10−9 + 2.34 × 10−10
ΣAl = 1.27 × 10−10
−9
1
– Al Si O (OH)
2
2 2 5
4
→ Al3+ + H SiO + 1– H O,
+ 3H+ ←
4
4
2 2
1
– Al Si O (OH)
2 2 5
4
2
→
+ H+ + –32 H2O ←
Al(OH)2+ + H4SiO4,
1
– Al Si O (OH)
2 2 5
4
2
→
+ –72 H2O ←
Al(OH)4− + H+ + H4SiO4.
kaolinite
= 4.64 × 10 ,
Equation 7.6 now takes the form:
so that:
aAl3+ /aAl(OH) + = 2.97 × 10−2.
2
With a shift of two pH units, total aluminum concentration
has dropped by four orders of magnitude and, as expected,
Al(OH)2+ has eclipsed free Al3+ as the major dissolved species.
Figure 7.3 shows the solubility of gibbsite, calculated from the
data in this worked problem over the range pH 3–12.
3
1 H+
ΣAl = (K a
+ K2aH+ + K3 /aH+)/a H 4SiO4.
This result is shown graphically in figure 7.4. The
solubility surface for kaolinite has cross sections perpendicular to the aH SiO4 axis that are qualitatively similar
4
to figure 7.3. With increasing silica activity, however,
kaolinite solubility at all pH values decreases. A threedimensional diagram such as figure 7.4 or contours drawn
at constant aH4SiO4 on diagrams as in figure 7.3 can help
us visualize the stability fields for such minerals as kaolinite or pyrophyllite. More complicated minerals like feldspars and micas, however, cannot be considered in this
way. Furthermore, because the aluminum concentration
in most natural waters is too low to measure reliably
with ease, diagrams such as figures 7.3 and 7.4 are less
practical than figure 7.2 was for magnesian silicates. For
this reason, geochemists commonly write dissolution reactions for aluminum silicates that produce secondary
solid aluminous phases instead of dissolved aluminum.
Such reactions are said to be incongruent.
One example of such a reaction is:
1
– Al Si O (OH)
2
2 2 5
4
FIG. 7.3. Activities of aqueous aluminum species in equilibrium
with gibbsite at 25°C. The bold curve, marking the boundary of
the fluid field, is ΣAl. (Modified from Drever 1997.)
→ Al(OH) + H SiO ,
+ –52 H 2O ←
3
4
4
which can be viewed loosely as an intermediate to any
of the kaolinite reactions just considered. Notice that
the relative stabilities of kaolinite and gibbsite, written
this way, depend only on the activity of H4SiO4 (that is,
116
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
for which:
Kksp-pyp = (aK+ /aH+)aH 4 SiO4;
→
KAlSi3O8 + –23 H+ + 4H 2O ←
K-feldspar
1
– KAl Si O (OH) + –2 K+ + 2H SiO ,
3
3 3 l0
2
3
4
4
muscovite
for which:
Kksp-mus = (aK+ /aH+)a3H 4SiO4;
→
KAl3Si3O10(OH)2 + H + + –32 H2O ←
muscovite
–3 Al Si O (OH) + K+,
2
2 2 5
4
kaolinite
for which:
Kmus-kao = (aK+ /aH+);
FIG. 7.4. Schematic diagram illustrating the kaolinite solubility
surface as a function of log(aK+/a H2+), pH, and log aH4SiO4. Notice
the similarity in form between crosssections at constant aH4SiO4
and the gibbsite solubility curve in figure 7.3. (Modified from
Garrels and Christ 1965.)
Kkao-gib= aH SiO4). Any information about dissolved alu4
minum species is “hidden,” because this reaction focuses
entirely on the solid phase Al(OH)3. We have no information about the stability of either mineral with respect
to a fluid unless we also know Al. We emphasize this
point before proceeding to develop equations and diagrams commonly used in aluminosilicate systems because, unfortunately, it is easy to overlook and can lead
to misinterpretation of those equations and diagrams.
With that warning, let’s now consider the broader
set of common minerals that includes not only silica
and alumina, but other cations as well. In the system
K2O-Al2O3-SiO2-H2O, for example, we can write the following set of reactions between minerals:
Σ
→
KAlSi3O8 + H+ + –92 H 2O ←
K-feldspar
1
– Al Si O (OH) + K+ + 2H SiO ,
2
2 2 5
4
4
4
kaolinite
for which:
Kksp-kao = (aK+ /aH+)a2H4SiO4;
→
KAlSi3O8 + H+ + 2H 2O ←
K-feldspar
1
– Al Si O (OH) + K+ + H SiO ,
2
2 4 10
2
4
4
pyrophyllite
and
1
– Al Si O (OH)
2
2 4 10
2
pyrophyllite
→
+ –52 H 2O ←
1
– Al Si O (OH)
2
2 2 5
4
kaolinite
+ H4SiO4,
for which:
Kpyp-kao = aH4SiO4.
Each of these equations plots as a straight line on a
graph of log aH SiO versus log(aK+ /aH+), as shown in fig4
4
ure 7.5. This diagram looks superficially like the one we
constructed for magnesium silicates. Because figure 7.5
has been constructed on the basis of incongruent reactions, there is one major difference between the two
diagrams: instead of marking the conditions under which
a mineral is in equilibrium with a fluid, each line segment in figure 7.5 indicates that two minerals coexist
stably. Consequently, four reactions (kao-gib, mus-kao,
ksp-kao, and pyp-kao) bound a region inside which kaolinite is stable relative to any of the other minerals we have
considered. Each of the other areas bounded by reaction
lines similarly represents a region in which one aluminosilicate mineral is stable.
Several times in earlier chapters we warned that thermodynamic calculations are valid only for the system as
it has been defined, and we raise the same caution flag
again. Just as figure 7.2 showed both forsterite and sepiolite to be unstable in the presence of water, you would
find that reactions between muscovite and pyrophyllite,
K-feldspar and gibbsite, or pyrophyllite and gibbsite are
Chemical Weathering: Dissolution and Redox Processes
FIG. 7.5. Stability relationships among common aluminosilicate
minerals at 25°C. Precise locations of field boundaries may vary
because of uncertainties in thermodynamic data from which the
equilibrium constants (K ) are calculated.
all metastable with respect to the reactions we have
plotted in figure 7.5. You can verify this for yourself (or
simply assume) that we have already written expressions
for all potential equilibrium constants in the system and
inserted free energy data to determine which reactions
are favored. It’s not an easy task. Particularly as other
components like CaO and Na2O are added and graphical
representation once again becomes difficult, it is no simple matter to identify each of the reactions that bounds
a mineral stability field. Figure 7.6 shows how unwieldy
graphical methods become with the addition of even one
more compositional variable (aNa+ /aH+). Today, computer
programs do the job by brute force, but accurately, so
that no significant reaction is inadvertently overlooked.
The most likely sources of error, then, are those that arise
because the thermodynamic data we input to the program have experimental uncertainties or we accidentally
omit phases from consideration.
Both of these sources of error are more likely than
you might think. The crystal chemistry of phyllosilicates
is complex: cationic and anionic substitutions are common, as are mixed-layer architectures in which two or
more mineral structures alternate randomly. Among
117
(Ca, Na)-aluminosilicates, this is particularly true. Natural smectites, for example, include both dioctahedral
(pyrophyllite-like) clays like montmorillonite and trioctahedral (talc-like) clays such as saponite. All show high
ion-exchange capacity and readily accept Ca2+, Na+,
K+, Fe2+, Mg2+, and a host of minor cations in structural
sites or interlayer positions. Their variable degrees of
crystallinity and generally large surface areas also contribute to large uncertainties in their thermodynamic
properties. As a result, thermodynamic data for many
aluminosilicates are gathered on phases that bear only a
loose resemblance to minerals we encounter in the world
outside the laboratory. The exact placement of field
boundaries in diagrams such as figures 7.5 and 7.6 is
subject to considerable debate. Complicating the picture are a number of common zeolites, like analcime and
phillipsite, which are stable at room temperature and
should therefore appear on diagrams of this sort, but are
remarkably easy to overlook.
These comments are not intended to be discouraging,
but should point out the necessity of choosing data carefully when you try to model a natural aluminosilicate
system rather than a hypothetical one. One moderately
successful approach to verifying diagrams like figure 7.5
involves the assumption that subsurface and stream
FIG. 7.6. Stability relationships among common sodium and
potassium aluminosilicate minerals at 25°C. (Modified from Drever
1997.)
118
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Worked Problem 7.2
FIG. 7.7. Activity diagram illustrating the compositions of water
samples from the Rio Tanama, Puerto Rico. Solid circles indicate
that sediment samples associated with the water contain both
kaolinite and calcium montmorillonite. Open circles identify
samples containing only kaolinite. The dashed boundary line,
based on these observations, compares well with the solid bold
boundary line, which was calculated from thermodynamic data.
Rio Tanama water lies to the right of the gibbsite stability field.
(Modified from Norton 1974.)
waters are in chemical equilibrium with the weathered
rocks over which they flow. If this is so, it should be possible to place field boundaries by analyzing water from
wells and streams for pH, dissolved silica, and major
cations and plotting the results as representative of the
coexisting minerals identified in the rocks at the sampling site. Norton (1974) followed this method in his
analysis of springs and rivers in the Rio Tanama basin,
Puerto Rico. As indicated in figure 7.7, the theoretical
boundary between stability fields for kaolinite and Camontmorillonite corresponds closely to the measured
composition of natural waters in equilibrium with both
minerals. Drever (1997) has questioned the assumption
of equilibrium in studies of this type, but the agreement
often shown between theory and nature in such studies
justifies hesitant optimism.
So much for theory. Let’s see how to use what we
have learned.
Assuming that reliable stability diagrams can be drawn to illustrate solubility relationships among aluminosilicate minerals,
how can we use this information to predict the course of chemical weathering? As an example, consider the reactions that
take place as microcline reacts with acidified groundwater in a
closed system.
In principle, we can predict a pathway by treating the problem as a titration of acid by microcline, just as you can predict
how solution pH varies as you add Na2CO3 to a dilute acid in
chemistry lab. We consider exactly that problem in chapter 8.
The difference here is that incongruent mineral reactions occur
along the reaction path, so that species other than aqueous ions
have to be taken into account. This is a cumbersome exercise,
commonly performed today by computer. The results shown in
figure 7.8 and discussed here were produced by the program
EQ6 (Wolery 1979).
As microcline dissolves in a fixed volume of water, aH SiO
4
4
and the ratio aK+ /aH+ are both initially very low. The fluid composition, then, begins beyond the lower left corner of the diagram and gradually creeps upward and to the right. If pH is
below ∼4, the dominant reaction at this early stage is:
→ K+ + Al3+ + 3H SiO .
KAlSi3O8 + 4H 2O + 4H + ←
4
4
But remember our earlier warning. No secondary phase
can appear until Al is high enough to saturate the fluid,
even though the water composition lies in the field marked
Σ
FIG. 7.8. Evolution of fluid composition during dissolution of
microcline in a closed system at 25°C, calculated by the program
EQ6 (Wolery 1979). Reactions along each segment of the path
indicated by arrows are discussed in worked problem 7.2.
Chemical Weathering: Dissolution and Redox Processes
119
“Gibbsite.” Once enough aluminum is in solution, gibbsite
forms while microcline continues to dissolve. The net reaction
releases potassium ions, OH −, and silica to the solution, causing
the fluid composition to move toward the upper right as both
aH SiO and aK+ /aH+ increase:
muscovite fields. The fluid path would then move vertically to
the K-feldspar field without aH SiO ever reaching 2 × 10−3, where
4
4
amorphous silica could precipitate.
→ Al(OH) + K+ + OH − + 3H SiO .
KAlSi3O8 + 8H2O ←
3
4
4
Although we devised this problem to illustrate weathering in a closed system, chemical weathering more commonly takes place in an open system. If water flows
through a parent material instead of stagnating in it, then
fluid composition not only evolves with time but also
varies spatially. As a result, residual minerals accumulate
in a layered weathering sequence that can provide direct
information about the fluid reaction path. A feldspathic
boulder, for example, might be surrounded by concentric
zones of muscovite-quartz, kaolinite-quartz, and gibbsite,
suggesting that open-system weathering had followed the
fluid path indicated in worked problem 7.2.
4
4
Eventually, the fluid composition reaches the gibbsitekaolinite boundary. At this point, microcline dissolution continues to release H4SiO4. Instead of continuing to accumulate,
however, that H4SiO4 is now consumed in the production of
kaolinite at the expense of gibbsite:
→
2KAlSi3O8 + 4Al(OH)3 + H 2O ←
3Al2Si2O5(OH)4 + 2K+ + 2OH −.
The fluid composition cannot move farther to the right as long
as some gibbsite remains. Potassium ions continue to accumulate in solution, however, so the fluid composition follows the
gibbsite-kaolinite boundary toward higher aK+ /aH+.
When all of the gibbsite is finally converted to kaolinite, the
fluid composition is no longer fixed along the field boundary, and
both aH SiO and aK+/aH+ can increase as kaolinite precipitates:
4
RIVERS AS WEATHERING INDICATORS
4
→
2KAlSi3O8 + 11H 2O ←
Al2Si2O5(OH)4 + 2K+ + 2OH − + 4H4SiO4.
Once the activity of H4SiO4 reaches 1 × 10−4 (the value of
Kq) it cannot increase further, because the fluid is saturated with
respect to quartz, which begins to precipitate as microcline dissolution continues. This reaction does not change the identity
of the stable aluminosilicate. Kaolinite remains in equilibrium
with the fluid as aK+ /aH+ rises.
At the muscovite-kaolinite boundary, the situation is similar
to that at the gibbsite-kaolinite boundary reached earlier, except that neither aH SiO nor aK+ /aH+ can change until all kaolinite
4
4
is converted to muscovite by the reaction:
→
KAlSi3O8 + A12Si2O5(OH)4 + H 2O ←
KAl3Si3O10 (OH)2 + SiO2.
Still prevented from becoming more silica-rich by the saturation
limit for quartz, the fluid continues to follow a vertical pathway
across the muscovite field:
→
3KAlSi3O8 + 2H2O ←
KA13Si3O10 (OH)2 + 2K+ + 2OH − + 6SiO2,
until it finally reaches the edge of the microcline stability field.
No further dissolution takes place, because the fluid is now saturated with respect to microcline.
It is important, once again, to emphasize that the preferred
pathway in nature may be distinctly different from what we
have calculated in this worked problem. For example, quartz
generally nucleates less readily than amorphous silica, so that
H4SiO4 might continue to accumulate when aH SiO = 10−4. In
4
4
that case, the reaction path would continue diagonally across
the kaolinite field, eventually reaching the kaolinite-pyrophyllite
boundary instead of moving vertically through the kaolinite and
As we have shown in several contexts so far, the type
and intensity of weathering reactions is reflected in an
assemblage of residual minerals and the composition of
associated waters. Sometimes it is appropriate to examine the pathways of chemical weathering by comparing
soil profiles with mineral sequences that we derive from
model calculations, as in worked problem 7.2. In other
cases, it is more reasonable to study the waters draining
a region undergoing weathering. This might be true if
soils were poorly developed or mechanically disturbed,
or if we were interested in alteration in a subsurface environment from which residual weathering products are
not easily sampled. It might also be true if we were concerned about weathering on a regional scale and wanted
to integrate the effects of several types of parent rock, or
if we were interested in seasonal variations in weathering.
In preceding sections, we focused on the SiO2-Al2O3
framework of silicate minerals and its stability in a weathering environment. We have seen, however, that mineral
stabilities during chemical weathering are functions of
2 , so this has
activity ratios such as aK+ /aH+ or aMg2+ /a H
+
also been a study of alkali and alkaline earth element
behavior. What happens, however, when several cations
are released along a reaction pathway? How are they
partitioned, relative to each other, among the clay products and the fluid?
Nesbitt and coworkers (1980) analyzed a progressively weathered sequence of samples from the Toorongo
120
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Granodiorite in Australia, which provides an interesting
perspective. The parent rock consists largely of plagioclase, quartz, K-feldspar, and biotite, with some minor
hornblende and traces of ilmenite. During initial stages of
weathering, sodium, calcium, and strontium in the rock
decrease rapidly in comparison with potassium. The primary reason for this decline is the early dissolution of
plagioclase, which is much less stable than K-feldspar or
biotite, the major potassic minerals. The clay mineral
that appears in the least altered samples is kaolinite.
As weathering progresses, vermiculite and then illite
join the clay mineral assemblage. These further influence
the loss or retention of cations. As the initially low pH
of migrating fluids begins to increase, charge balance on
surfaces and in interlayer sites in clays is satisfied by
large cations (K+, Rb+, Cs+, and Ba2+) rather than by H +.
The smaller cations (Ca2+, Na+, and Sr2+) are less readily
accepted in interlayer sites and so are quantitatively
removed in groundwater and surface runoff. The sole
exception is Mg2+, which is largely retained in the soil,
as biotite is replaced by vermiculite, which has a small
structural site to accommodate it. As a result, progressive weathering of the Toorongo Granodiorite produces
a residual that is greatly depleted in calcium and sodium
and only slightly depleted in potassium and magnesium.
FIG. 7.9. Chemical changes during weathering of the Toorongo
Granodiorite. Measured compositions of rock samples are indicated
by solid circles. Average unweathered rock compositions are indicated by small open boxes, and average stream water composition
by the larger boxes. With progressive weathering, rock compositions move away from the Ca + Na2 corner on both diagrams,
becoming more aluminous and relatively enriched in potassium
and magnesium. Water compositions show complementary trends.
(Modified from Nesbitt et al. 1980.)
FIG. 7.10. Clay minerals in residual soils in California formed on
(a) felsic and (b) mafic igneous rocks. (After Barshad 1966.)
Figure 7.9 illustrates this trend, and shows that the composition of local stream waters is roughly complementary
(that is, they are relatively rich in Na+ and Ca2+ and poor
in K+ and Mg2+). Here, then, is an illustration of how
analyses of natural fluids can confirm what geochemists
learn about weathering by analyzing residual minerals.
These observations are generally consistent with the
thermodynamic predictions we made in the last section
and with observed clay mineral abundances on a global
scale (see, for example, fig. 7.10). Gibbsite only appears
in weathered sections that are intensively leached (that
is, ones in which the associated waters are extremely dilute). Kaolinite dominates where waters are less dilute,
and smectites where waters have higher concentrations
of dissolved cations. Vermiculite and illite are common
as intermediate alteration products where soil waters
have concentrations between these extremes, as is true in
most regions of temperate climate and moderate annual
rainfall.
Chemical Weathering: Dissolution and Redox Processes
AGENTS OF WEATHERING
Carbon Dioxide
All surface environments are exposed to atmospheric
CO2 and H2O. The chemical combination of the two is
a powerful and ubiquitous weathering agent. Data compiled by Heinrich Holland (1978) suggest that water in
wells and springs draining limestone terrains contains
between 50 and 90 mg Ca2+ kg−1, due almost entirely to
contact with CO2-charged waters. When we study carbonate equilibria in greater detail in chapter 8, however,
we show that this is much more dissolved calcium than
we ought to expect from weathering by atmospheric CO2
alone. The measured calcium concentrations reported by
Holland suggest that the partial pressure of CO2 in equilibrium with limestones in nature is 10–100 times higher
than atmospheric PCO2. These surprisingly high values are
actually common in air held in the pore spaces of soil.
Measurements made in arable midlatitude soils indicate
that PCO2 is between 1.5 × 10−1 and 6.5 × 10−3 atm—
roughly 5–20 times that in the atmosphere. Analyses of
a tropical soil reported in Russell (1961) indicate PCO2 as
high as 1 × 10−1 atm—330 times the atmospheric value!
These findings indicate that if we want to study the
effectiveness of CO2 as a chemical weathering agent, we
must look primarily to those processes that control its
121
abundance in soils. A small amount of CO2 is generated
when ancient organic compounds in sediments are oxidized during weathering. This amount, however, is almost
negligible. Most of the CO2 is produced by respiration
from plant roots and by microbial degradation of buried
plant matter (see, for example, Wood and Petratis 1984;
Witkamp and Frank 1969). Cawley and coworkers
(1969) provided dramatic confirmation of this CO2
source by sampling streams in Iceland. They found that
streams draining vegetated regions had bicarbonate concentrations two to three times higher than streams in areas
with the same volcanic rock but without plant cover.
CO2 generation follows an annual cycle in response to
variations in both temperature and soil moisture. Values
rise rapidly at the onset of a growing season and fall again
as lower air and soil temperatures mark the end of the
growing season. The rate of chemical weathering, therefore, should be greatest in most regions during the summer. Even in midwinter, however, soil PCO2 may sometimes
be several times the atmospheric value. Solomon and
Cerling (1987) found that snow cover can trap CO2 in
some intermontane soils, so that PCO2 in winter can actually exceed the value in summer at shallow depths (see
fig. 7.11). Because carbon dioxide is more soluble in cold
than in warm water, this seasonal pattern may lead to
weathering rates that are significantly enhanced during
the winter.
FIG. 7.11. Soil CO2 as a function of depth and time from April 1984 to July 1985 at Brighton,
Utah. In general, CO2 levels are highest during the warm months, when biological activity in the
soil is highest. Also, PCO2 decreases toward the surface due to diffusive loss to the atmosphere.
Notice, however, that PCO2 in shallow soils is higher during the winter months than during the
summer, even though biological CO2 production is lower. The snow may act as a cap to prevent
CO2 from escaping to the atmosphere. (After Solomon and Cerling 1987.)
122
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Organic Acids
Despite considerable study, the role of organic matter in chemical weathering is poorly understood. We suggested in the previous section that some of the elevated
CO2 concentrations in soil are the result of biodegradation, so that organic compounds affect weathering rates
indirectly by serving as a source for CO2. Laboratory
and field studies, however, indicate that they have a much
stronger direct influence.
Upon decomposition, the organic remains of plants
and animals yield a complex mixture of acids. A number
of simple compounds, such as acetic and oxalic acid, are
commonly found around active root systems and fungi.
These and other acids (such as malic and formic acids)
serve the same role as HCl or H 2SO4 and act as H +
donors. Depending on the acidity of groundwaters, any
of these acids can increase mineral solubility.
Low-molecular-weight organic acids may play a role
in mineral dissolution by forming complexes with polyvalent cations on mineral surfaces. Organic ligands such
as oxalate, malonate, succinate, salicylate, phthalate, and
benzoate have been shown to accelerate dissolution rates
of simple oxides (Al2O3 and iron [III] oxides) compared
with rates determined for inorganic acids alone. This
mechanism appears to have a limited effect on quartz
solubility, although Bennett (1991) observed direct evidence of increased quartz dissolution in a hydrocarboncontaminated aquifer and inferred surface complexation.
Additional laboratory studies show that quartz dissolution at 25°C increases in the presence of citric and
oxalic acids (0.02 M) with the greatest effect between
pH 5.5 and 7. In contrast, organic complexes on the surfaces of feldspars appear to greatly enhance solubility.
The Si/Al ratio of the mineral influences the extent to
which organic ligands enhance feldspar dissolution, with
the greatest solubility corresponding to the lowest ratio.
This inverse relation suggests that the organic ligand enhances dissolution by forming a complex with available
Al3+ ions. Amrhein and Suarez (1988) found that at pH 4
and above, 2.5 mM oxalate enhanced anorthite (Si/Al = 1)
dissolution by a factor of two, and Stillings and others
(1996) showed that in 1 mM oxalate solutions, feldspar
dissolution rates were enhanced by a factor of 2–5 at
pH 3 and by factors of 2–15 at pH 5–6.
The role of high-molecular-weight acids is less well
understood. Humic and fulvic acids are composed of a
bewildering array of aliphatic and phenolic structural
units, with a high proportion of −COOH and phenolic
OH radicals, as suggested schematically in figure 7.12.
At least some of these acids are soluble or dispersible in
water, where limited dissociation of these radicals can
occur, particularly if soil pH is low. The effect is twofold.
First, dissociation releases hydrogen ions into solution,
where they influence solubility equilibria in the same way
that inorganic acids do. Second, the organic acid molecule, following dissociation, develops a surface charge
that helps to repel other similarly charged particles. It
therefore adopts a dispersed or colloidal character, which
increases its mobility in soil waters.
This second role is particularly significant, because
several studies indicate that organic colloids may be
important in controlling both solubility and transport
of mineral matter (Schnitzer 1986). Provided that free
cations are not so abundant that they completely neutralize the surface charge, humic and fulvic acids can
attach significant amounts of Fe3+ and Al3+, and lesser
amounts of alkali and alkaline earth cations, without
losing their mobility as colloids. In this way, the stability
of colloidal organic matter may, for example, be a controlling factor in the downward transport of iron and
aluminum in acidic forest soils (spodosols and alfisols)
(DeConick 1980).
Finally, organic acids in soils commonly act as chelating agents. Chelates of fulvic acids and of oxalic, tartaric,
and citric acid are thought to be responsible for most of
the dissolved cations in forest soils (Schnitzer 1986). As
we first saw in chapter 4, dissolved cations can contribute
significantly to the mobility of metals. Fulvic and humic
acids also both form mixed ligand complexes with phosphate and other chemically active groups.
Although each of these factors can contribute to mineral dissolution, their net effect is slight at best. In fact,
high-molecular-weight organic molecules are more refractory than light molecules (that is, they are not as easily
metabolized by extant microbial communities), and they
appear to decrease the feldspar dissolution rates rather
than increase them, except at low pH. Ochs and coworkers (1993) investigated the pH dependence of
α-Al2O3 dissolution in the presence of humic acids and
found that at pH 3, there was a slight enhancement,
whereas at pH 4 and 4.5, the acids inhibited dissolution.
They concluded that under strongly acidic conditions,
most of the functional groups on the humic molecule are
protonated and can therefore detach Al3+ ions by forming mononuclear complexes. At higher pH, the surface
Chemical Weathering: Dissolution and Redox Processes
CH
O
CH
OH
H2
CH
O
C
C
H2
CH2O
C
O
H 2C
C
C
C
C
O
CH3
O
O
N
O
HN
H
H
H
H
NH3
C
C
C
C
C
NH2 H
H
H
O
O
HN
O
H2N
CH2 C
H
C
O
O
CH3
CH3
H2COH
O
C
CH3
O
O
C
CH3O
CH2
C
O
C
H
O
HCOH
O
C
CH2
C
O
O
CH2 CH
CH2
2
CH2
CH2 CH2
O
C
O
O
O
H2C
C
H
CH2
CH3
N
S
S
O
O
OH
OCH3
CH3
CH3
O
C
C
CH2
CH3
O
O
O
C20H39OCCH2
CH2
N
C
C
O
O
CH3
H3NCH
O
O
O
OH
O
H3NCH
C
H3N
O
C
H
O
O
C
O
C
CH2
OH
O
O
O
OH
NH
CH2
CH2
CH2 CH2 CH2 CH2
CH2
CH2
C
CH2
CH2
CH2 CH2 CH2 H
O
O
CH2
C
O
OH
HO
O
CH2
OH
CH2 CH
C
2
O
C
O
C
C
CH3
O
NH3
CH3
O
O
O
CH3
H 3O
O
123
CH3
H
O
C
O
HO
C
N
O
CH2 OH
HO
OH
FIG. 7.12. Chemical-structural model of a humic acid with a molecular weight of ∼5000. Boxes outline various biological precursor units
incorporated into the humic acid. Light shaded boxes are lignin degradation products. Dark shaded boxes are nucleic acids, boxes with
horizontal lines are vitamins, boxes with diagonal lines are sugars, and white boxes are amino acids or other compounds. (Modified from
Leventhal 1986.)
of the mineral is coated with poorly protonated humic
molecules that block the release of cations and decrease
the dissolution rate.
Solubility in some lab settings can be striking. For
example, Baker (1986) performed a series of experi-
ments in which various minerals were held in a 500-ppm
solution of humic acids extracted from bracken fern. Because of the possibility that CO2 derived from organic
acids (rather than the acids themselves) affects mineral
solubility in nature, Baker also performed dissolution
124
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
TABLE 7.1. Dissolution of Selected Minerals
after 24 Hours at 25°C
Mineral
Ion
Atomspheric
H2O/CO2
Humic Acid
Calcite
Dolomite
Dolomite
Hematite
Pyrite
Galena
Gold
Ca
Ca
Mg
Fe
Fe
Pb
Au
0.65
0.25
0.14
0.06
0.04
0.06
10 ng/mL
17.40
7.50
5.65
0.58
0.24
2.62
190 ng/mL
Humic Acid
(H2O/CO2)
2.75
0.45
0.33
0.24
All solution concentrations are in mg L–1 except as noted.
Based on data selected from Baker (1986).
In a great many cases, however, oxygen is not a reactant
in redox systems. In fact, it is misleading to emphasize
the role of oxygen, because it is the change in the transition metal ion that characterizes these systems. It is more
practical in many cases, therefore, to write redox reactions to emphasize the exchange of electrons between
reactant and product assemblages. The example above
can be rewritten as:
2Fe3O4 + H 2O → 3Fe2O3 + 2H + + 2e − ,
in which the fundamental change involves the oxidation
of ferrous iron:
Fe2+ → Fe3+ + e − .
experiments in water charged with enough CO2 to simulate the effect of converting all of the organic matter in
the humic acid solution to carbon dioxide. Some of his
results are shown in table 7.1. The data indicate that organic acids can greatly enhance the dissolution process,
especially when reactions take place in an open, flowthrough system. Unfortunately, stability constants for
most relevant organic complexes are not known, so measurements of this sort cannot generally be used to constrain thermodynamic models of chemical weathering.
Also, in general, silicate dissolution rates derived from
experiments are much faster than those observed in natural settings. We conclude that the quantitative effect
of organic acids is highly variable, but generally nowhere
near as great as the effect of other weathering agents.
OXIDATION-REDUCTION PROCESSES
Thermodynamic Conventions for Redox Systems
In previous sections and chapters, we confined our
discussion to reactions that involve atoms with only one
common oxidation state. Solubility equilibria in these
systems are characterized by either hydrolysis or acidbase reactions. Most natural systems, however, also
contain transition metal elements, which may form ions
in two or more different oxidation states.
Reactions in which the oxidation states of component
atoms change are called redox reactions. These are not
restricted to reactions involving transition metals, although it is most often those systems that are of interest.
In many cases, redox reactions involve the actual exchange of oxygen between phases. For example:
→ 3Fe O .
2Fe3O4 + 1–2 O2 ←
2 3
Geochemists may write reactions according to either
convention. At low temperatures or when a fluid phase
is present, it is usually convenient to write redox reactions
in terms of electron transfer. As we shall see, reaction potentials in aqueous systems are easily measured as voltages. At high temperatures, where these measurements
are much more difficult to perform, it is generally easier
to determine oxygen fugacity and write redox equilibria
in terms of oxygen exchange. In this chapter, we devote
most of our attention to aqueous systems.
A reaction written to emphasize the production of
free electrons, although conceptually useful, cannot take
place in isolation in nature, because an electron cannot
exist as a free species. (To emphasize this important
distinction, we are using a single arrow [→] to indicate
→]
reactions of this type, instead of the double arrows [←
used for all other reactions in this book.) Such a reaction
should be viewed in tandem with another similar reaction that consumes electrons. Together, the pair of halfreactions represents the exchange of electrons between
atoms in one phase, which become oxidized by loss of
electrons, and those in another, which gain electrons and
are therefore reduced. A companion to the half-reaction
above, for example, might be:
1
–H
2 2
→ H + + e −.
A natural reaction involving half-reactions but resulting
in no free electrons is:
2Fe3O4 + H 2O → 3Fe2O3 + H 2.
By writing these illustrative reactions, we introduce
two standard conventions. First, geochemists usually
write half-reactions with free electrons on the right side;
Chemical Weathering: Dissolution and Redox Processes
FIG. 7.13. Schematic diagram of the standard hydrogen electrode
1
(SHE). The half-cell reaction is –2 H2 → H+ + e −. Platinum metal
serves as a catalyst on the electrode. The salt bridge is a tube
filled with a KCl solution and closed with a porous plug.
that is, they are written as oxidation reactions. It should
be clear that the progress of a pair of half-reactions depends on their relative free energies of reaction, so that
there is no special significance in this convention. Fe3O4
may actually be oxidized to form Fe2O3 if it is coupled
with a half-reaction with a higher free energy of reaction, but may be reversed if its companion half-reaction
has a lower ∆Ḡr.
The second standard convention involves the oxidation of H 2 to form H +, which we used above. It is arbitrarily chosen to serve as an analytical reference for all
other half-reactions. To explain this standard and its
value in solution chemistry, we have drawn a schematic
illustration (fig. 7.13) of a lab apparatus known as the
standard hydrogen electrode or SHE. It consists of a
piece of platinum metal immersed in a 25°C solution
with pH = 1, through which H2 gas is bubbled at 1 atm
pressure. The platinum serves only as a reaction catalyst
and a means of making electrical contact with the world
outside the apparatus. If the SHE had a means of exchanging electrons with some external system, it would
be the physical embodiment of the half-reaction:
1
–H
2 2
→ H+ + e−.
By agreement, chemists declare that the ∆Ḡf of H +
and of e− is zero. Hence, the SHE has its free energy of
reaction defined arbitrarily as zero. Because aH+ = 1 in
125
the SHE, ae− = 1 as well. The “activity of electrons” is a
slippery concept, because it is a measure of a quantity
that has no physical existence. In this sense, ae− is not
equivalent to ion activities we have considered before.
Instead, it is best to treat it as a measure of the solution’s
tendency to release or attract electrons from a companion half-reaction cell. By agreeing that there is no electrical potential between the platinum electrode and the
solution, we have created a laboratory device against
which the reaction potential of any other half-reaction
cell can be measured.
Let’s see how this works. Figure 7.14 illustrates an
experiment in which the SHE is paired with another system containing both Fe2+ and Fe3+ in solution. The platinum electrode of the SHE is connected by a wire with a
similar electrode in the Fe2+ | Fe3+ cell, and electrical contact between the solutions is maintained through a KCl
solution in a U-tube between them. When the circuit is
completed, a current will flow, as electrons are transferred
from the solution with the higher activity of electrons
to the one with the lower activity. If a voltage meter is
placed in the line, it will measure the difference in electrical potential (E) between the SHE and the Fe2+ | Fe3+
cell.
In this example and most others of interest to aqueous geochemists, the voltage measurement (E) is given
the special symbol Eh, where the h informs us that the
reported value is relative to the standard hydrogen
electrode. The implication of the last few paragraphs is
that chemical potential and electrical potential are interrelated, so that it is fair to speak either of the Eh of a
half-cell or of its free energy. To be explicit, recall from
chapter 3 that the decrease in free energy of a system
undergoing a reversible change at constant temperature
and pressure is equal to the nonmechanical (i.e., nonpressure–volume) work that can be done by the system.
That is, dG = −dwrev. The reversible work done by an
electrochemical half-cell is equal to the product of its
electrochemical potential (Eh) and the current (Q) flowing from it. Q, in turn, is simply equal to the product of
Faraday’s constant (F = 23.062 kcal volt −1 eq −1) and the
number of moles of electrons released (n):
Q = −nF.
The negative sign appears because the charge on an electron is negative. This gives us, then:
∆G = −∆w = −EhQ = nFEh.
(7.7)
126
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 7.14. A pair of half-cells described by the SHE and the reaction Fe2+ → Fe3+ + e −. When the
switch is closed, electrons can be transferred through the salt bridge and wires between the two
cells. When it is opened, the voltage meter measures the potential difference between them. The
direction of current flow depends on the Fe2+/Fe3+ ratio in the left hand cell.
For the two half-cells in this example, we can write
equilibrium constants:
Keq = aFe3+ ae− /aFe2+ ,
and
KSHE = aH+ ae − /PH2.
We calculate the value of Keq in now familiar fashion
from tabulated ∆Ḡf values for Fe2+ and Fe3+, and we see
that ae− is proportional to the ratio aFe2+ /aFe3+. For the net
reaction,
→ Fe3+ + 1– H ,
Fe2+ + H+ ←
2 2
the free energy of reaction is given by:
∆Ḡr = ∆Ḡr0 + RT ln(Keq /KSHE),
which is simply:
∆Ḡr = ∆Ḡr0 + RT ln(aFe3+ /aFe2+).
From equation 7.7, we see that this result is equivalent to:
( )
RT
Eh = E0 + —— ln(aFe3+ /aFe2+),
nF
or
(
)
2.303RT
Eh = E0 + ———–— log10(aFe3+ /aFe2+).
nF
The quantity E0, called the standard electrode potential,
is analogous to the standard free energy of reaction. It
is the Eh that the system would have if all of the chemical species in both half-cells were in their standard
states (that is, if each had unit activity). The logarithmic
term is simply the ratio of the activity product of oxidized species (here, aFe3+) to the activity product of reduced
species (here, aFe2+). This equation, in its general form, is
called the Nernst equation. (We have used the notation
log10 for base 10 logarithms sparingly elsewhere in this
book. We use it for the rest of this chapter, however, to
call attention to the numerical conversion factor between
base 10 and natural logarithms in most equations.)
From this general discussion, it should now be clear
that the thermodynamic treatment of redox reactions in
terms of Eh is fully compatible with the procedure we
have observed beginning with chapter 4. In the same way
that the direction of a net reaction depends on whether
the products or the reactants have the lower free energy,
Chemical Weathering: Dissolution and Redox Processes
a half-cell reaction may be viewed as “oxidizing” or “reducing” relative to the SHE, depending on the arithmetic
sign on Eh. The advantage of using Eh rather than ∆Ḡr is
that electrochemical measurements are relatively easier
to perform than is calorimetry.
Worked Problem 7.3
Iron and manganese are commonly carried together in river
waters, but are separated quite efficiently upon entering the
ocean. Ferric oxides and hydroxides precipitate in near-shore
environments or estuaries, but manganese remains as soluble
Mn2+ even in the open ocean, except where it is finally bound
in nodules on the abyssal plain. Some of this behavior is due to
colloid formation. Can we also describe it in terms of electrochemistry?
Two half-reactions are of interest in this problem. Iron may
be converted from fairly soluble Fe2+ to relatively insoluble Fe3+
by the half-reaction:
Fe2+ → Fe3+ + e−,
for which E 0 = +0.77 volt. The manganese half-reaction is:
Mn2+ + 2H 2O → MnO2 + 4H + + 2e−,
for which E 0 = +1.23 volt. We construct the net reaction by
doubling the stoichiometric coefficients in the Fe2+ | Fe3+ reaction
and subtracting it from the manganese reaction to eliminate
free electrons:
→ Mn2+ + 2H O + 2Fe3+.
MnO2 + 4H+ + 2Fe2+ ←
2
We do not calculate the standard electrode potential for the
net reaction in the way you might expect. Unlike enthalpies and
free energies, voltages are not multiplied by the coefficients we
used to generate the net reaction. The reason is apparent from
equation 7.7, from which we see indirectly that:
∆Gr0
E 0 = —–—
.
nF
As the stoichiometric coefficients of the Fe2+ | Fe3+ reaction are
doubled, the value of ∆Ḡr0 and the number of electrons in the
reaction, n, also double. The half-cell E 0, therefore, is unaffected.
The standard potential for the net reaction is +0.77 − (+1.23) =
−0.46 volt. This corresponds to a net ∆Ḡr0 of −21.2 kcal mol−1.
Eh for this problem is calculated from:
2
2.303R(298)
aMn2+ a Fe
3+
Eh = −0.46 + —————— log10 ———
——
4+ 2
2F
a H a Fe2+
2
0.059
aMn2+ a Fe
3+
= −0.46 + ———– log10———
—— .
4+ 2
2
a H a Fe2+
Sundby and coworkers (1981) found that waters entering
the St. Lawrence estuary have a dissolved manganese content of
127
10 µg/L−1, or approximately 1.8 × 10−7 mol L−1. On average,
the ratio of total iron to manganese in the world’s rivers is
about 50:1, approximately the ratio in average crustal rocks. We
estimate, therefore, that aFe2+ + aFe3+ = 9 × 10−6. Unfortunately,
river concentrations of Fe3+ are rarely reported, so we must
choose a value for the activity ratio of Fe3+ to Fe2+ arbitrarily.
Divalent iron is almost certainly the more abundant species but,
for reasons that soon become apparent, aFe3+ cannot be much
below 1 × 10−6. For the purposes of this problem, we assume
that aFe3+ /aFe2+ is 1:8, so that:
aFe3+ = 1.0 × 10−6,
and
aFe2+ = 8.0 × 10−6.
We also assume that the pH of water entering the St. Lawrence
system is 6.0, which is typical of streams draining forest-covered
terrain in temperate climates.
By making these assumptions and substituting appropriate
values for R and F in our earlier expression for Eh, we calculate
that:
0.059
1.8 × 10−7(l.0 × 10−6)2
Eh = −0.46 + ——— log10 —————————— .
2
a 4H+ (8.0 × 10−6)2
At pH = 6, the value of Eh predicted by this expression is
very close to zero. This indicates, therefore, that the net reaction
we have written is very nearly in equilibrium. However, ferric
ions cannot accumulate by much before the solubility product
for a number of ferric oxides and hydroxides is exceeded. As
ferric iron is lost to sediments, the bulk reaction favors gradual
oxidation of Fe2+ and thus a reduction in the total amount of
dissolved iron in the estuary. The same reaction favors the production of Mn2+ ions, which remain in solution and are washed
out to sea.
Eh-pH Diagrams
As we have shown in the previous section, many redox
reactions are functions of both Eh and pH. The net reaction in worked problem 7.3, for example, consumes four
hydrogen ions and involves the transfer of two electrons.
It can be easily shown that reactions in many systems
of interest to aqueous geochemists can be illustrated in
diagrams of the type we have already seen in figures 7.5
and 7.8, but using Eh and pH as variables instead of the
activities of H4SiO4 and aqueous cations. We show this
by discussing figure 7.15, which represents a number of
equilibria among iron oxides and water.
The uppermost and lowest lines in figure 7.15 establish the stability limits of water in an atmosphere with a
128
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
The activity of water, as a pure phase, is unity. The
partial pressure of O2 in the atmosphere at sea level is
0.2 atm. We make only a small difference in the final result if we choose instead to let PO = 1 atm to simplify
2
the expression for Eh:
Eh = E 0 + (0.059/4)log10 a 4H+
= E 0 − 0.059 pH.
This is the equation for a straight line with slope −0.059
and y intercept (E 0 ) equal to +1.23 volts, which is a
graphical representation of the half-reaction producing
O2 from H 2O. As shown in figure 7.15, it divides the
Eh-pH plane into a region in which water is stable and
one in which O2 is.
The lower stability limit for water is derived from the
SHE half-reaction itself, because it is that half of the net
reaction that defines the oxidation of H 2. The appropriate expression is:
Eh = E 0 + (0.059/2) log10 a 2H+.
Because E 0 for the SHE is defined to be zero, and PH
2
cannot be higher than the total atmospheric pressure
of 1 atm, this expression simplifies to:
FIG. 7.15. Stability fields for hematite, magnetite, and aqueous
ions in water at 25°C and 1 atm total pressure. Boundaries involving aqueous species are calculated by assuming that the total activity of dissolved species is 10−6. Lighter-weight lines indicate
the field boundaries when the sum of dissolved species activities
is 10−4. The boxed area near pH 4–6 indicates the range of rainwater compositions in many parts of the world.
total pressure of 1 atm (that is, at the surface of the Earth).
These limits are derived from the net reaction:
→ 2H + O ,
2H 2O ←
2
2
which is the sum of the half-reaction:
2H 2O → O2 + 4H + + 4e−,
and the reaction that defines the SHE. Eh for this reaction
at 25°C, therefore, is given by:
2.303RT
Kq
Eh = E0 + ————– log10 ———
4F
KSHE
0.059
PO42aH+
= E0 + ——— log10 ———–
.
4
aH2O
Eh = −0.059 pH.
Again, this is a line in the Eh-pH plane (fig. 7. 15) representing the SHE graphically; in this case, it defines a region in which H 2 is stable and another in which H + (and
therefore H 2O) is stable. The net reaction, therefore, is
represented in the Eh-pH plane as a region between the
half-reactions, in which H 2O is stable. As a point of reference within this large region, it may be convenient to
focus on the shaded box around Eh = +0.5 and pH = 5.5,
which is typical of rainwater in many parts of the world
and thus a reasonable starting point for many weathering reactions.
The stability fields of iron oxides are also bounded by
half-reactions. The reaction forming hematite by oxidation of magnetite, for example, can be written:
→ 3Fe O ,
2Fe3O4 + 1–2 O2 ←
2 3
or can be written in terms of electron exchange by combining this equation with the half-reaction:
2H 2O → O2 + 4H + + 4e−.
We saw the result earlier as:
2Fe3O4 + H 2O → 3Fe2O3 + 2H + + 2e−.
Chemical Weathering: Dissolution and Redox Processes
Following the procedure we just used for the stability of
water, we find that:
3
2
a 2 + /[a Fe
a ])
Eh = E 0 + (0.059/2) − log10(a Fe
2O3 H
3O4 H 2O
2
= E 0 + (0.059/2) log10 a H
+
= E 0 − 0.059 pH.
The value of E 0, calculated from standard free energies,
is +0.221 volt.
The remaining half-reactions shown in figure 7.15
do not have a slope of −0.059, because they all involve
water as a reactant or product. Consider, for example,
the half-reaction forming hematite from dissolved Fe2+:
2Fe2+ + 3H 2O → Fe2O3 + 6H + + 2e−.
Here the equilibrium constant involves not only hydrogen
ions and electrons, but ferrous ions as well, so that:
6
2
Eh = E 0 + (0.059/2)log10(a H
+ /a 2+)
Fe
= E 0 − 0.059log10 aFe2+ − 0.177 pH,
where E 0 = +0.728 volt. To draw a line to represent
this reaction in the Eh-pH plane, therefore, we need to
specify an activity of ferrous ions. In drawing the Fe2+hematite boundary in figure 7.15, we have assumed that
aFe 2+ = 10−6, and have indicated by use of a lighter line
the way that the boundary would shift if aFe 2+ were 10−4.
The boundary between Fe2+ and magnetite can be calculated in the same way.
The line between the aqueous Fe3+ and Fe2+ fields indicates the conditions under which activities of the two
species are identical. The redox reaction is:
Fe2+ → Fe3+ + e−,
which involves no hydrogen ions and is therefore independent of pH. We see easily that:
Eh = E 0 + 0.059log10 (aFe3+ /aFe2+).
Because aFe2+ = aFe3+ , however, this simply reduces to
Eh = E0, which in this case is +0.771 volt.
Finally, the boundary between aqueous Fe3+ and hematite is defined by the net reaction
→ 2Fe3+ + 3H O.
Fe2O3 + 6H+ ←
2
We could write this as the sum of two half-reactions, but
we can save ourselves a lot of trouble by noticing that
this net reaction does not involve a change of oxidation
state for iron. The field boundary, therefore, must be a
129
horizontal line, independent of Eh. The standard molar
free energy of reaction, ∆Ḡr0, is equal to 2 kcal mol−1, so
that:
2 /a 6 )
log10 Keq = −∆Ḡr0/(2.303RT) = log10(a Fe
3+
H+
= −1.45.
Rearrange this result to solve for pH, and we conclude
that:
pH = −0.24 − 1–3 log10 aFe3+ ,
which plots as a vertical line with a position that depends,
again, on the activity of dissolved iron.
Redox Systems Containing Carbon Dioxide
Redox reactions in weathering environments always
occur in an atmosphere whose PCO2 is at least as high as
the atmospheric value and, as we saw earlier in this chapter, may be several hundred times greater. Under some
conditions, therefore, we might expect to find siderite
(FeCO3 ) stable in the weathered assemblage. Including
carbonate equilibria in the redox calculations we have
just considered is not difficult. It merely requires us to
write half-reactions to include the new phases. For the
equilibrium between siderite and magnetite, we write:
3FeCO3 + H 2O → Fe3O4 + 3CO2 + 2H + + 2e−,
and
Eh = 0.319 + 0.089 log10 PCO2 − 0.059 pH.
For the equilibrium between siderite and hematite:
2FeCO3 + H 2O → Fe2O3 + 2CO2 + 2H + + 2e−,
and
Eh = 0.286 + 0.059 log10 PCO2 − 0.059 pH.
Worked Problem 7.4
What values of Eh, pH, and PCO define the stability limits for
2
siderite formed in weathering environments? The answer is
shown in graphical form by the shaded area in figure 7.17. To
see how it is derived, take a look at the two equations for Eh
above. Notice that each of these would plot as a diagonal line
with slope of −0.059 on an Eh-pH diagram if we were to choose
a fixed value of PCO . For successively higher values of PCO , the
2
2
boundary lines would be drawn through higher values of Eh at
any selected pH.
130
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
HOW DO WE MEASURE pH?
Electrodes are commercially available in a variety of
physical configurations, and can be designed for sensitivity to a particular ion or group of ions, to the exclusion of others. Regardless of the details of design,
however, all rely on ion exchange to do their job. The
transfer of ions between the electrode and a test solution results in a current, which can then be registered
in a metering circuit.
One standard design is known as the glass-membrane electrode; the basic operating principles of this
design are illustrated in figure 7.16. The device consists
of a thin, hollow bulb of glass, filled with a solution
with a known activity of a cation to be measured. The
glass of the bulb contains the same cation, introduced
during manufacture. A wire immersed in the filling
solution completes the electrode.
In operation, the bulb is placed in a solution to be
analyzed. Because the solution inside the bulb and
the solution outside generally have different activities
of the cation to be measured, a chemical potential
gradient is immediately established between them.
Ions on the more concentrated side minimize the difference by attaching themselves to the glass, whereas
those on the other side leave the glass to join the less
concentrated solution. The glass thus serves as a
semipermeable membrane. Because charged particles
are passing from one side to the other, an electrical
imbalance is created. By connecting the wire inside
the bulb to an inert or reference electrode immersed
in the test solution, we allow electrons to flow and
neutralize the charge difference. A voltage measured
in this system is the electrode potential.
A standard pH electrode, found in almost any
chemistry lab, is a glass electrode of this type, filled
with an HCl solution of known aH + . The outer bulb
is made of a Na-Ca-silicate glass, from whose surfaces
sodium ions are readily leached and replaced by H +.
When it is placed in a test solution, the glass gains
extra hydrogen ions on one side and yields them on the
other. Instead of a simple wire, its internal electrical
connection is actually a second electrode, which consists of a silver wire coated with silver chloride. This
serves as an internal reference electrode. The charge
imbalance that develops within the filling solution in
the glass bulb is resolved by the silver wire in the inner
electrode, which reacts with the filling solution to release or gain an electron in the half-reaction:
Ag + Cl− → AgCl + e−.
FIG. 7.16. A standard glass-membrane pH electrode (on the
right) consists of a Na-glass bulb filled with 0.1 M HCl, at
the tip of a Ag-AgCl reference electrode. Inset shows the ion
exchange path between the electrode and a solution with
unknown pH. This electrode must be paired with a second one,
which serves as a reference half-cell to complete the electrical
circuit. The one shown on the left in this figure is a calomel
electrode, and is based on the redox pair Hg0 | Hg+.
The charge imbalance in the test solution is ultimately
resolved by transferring electrons between the silver
wire and a reference electrode (such as the SHE or a
more portable equivalent) that is also immersed in the
test solution. In this rather roundabout way, the aH+
difference between the test solution and the filling
solution is expressed as a measurable voltage.
Chemical Weathering: Dissolution and Redox Processes
istic limit, perhaps, is 10−1 atm, measured by Russell (1961) in
a tropical soil. At this PCO value, the siderite-hematite bound2
ary is given by:
+1.0
-4
Fe3+aq
+0.8
Eh = 0.227 − 0.059 pH,
O2
+0.6
which is the upper edge of the shaded siderite field in figure 7.17.
In very alkaline solutions, siderite is stable over the range of
Eh and PCO values we have just found. In more acidic solutions,
2
however, siderite breaks down according to the reaction:
H2O
+0.4
Fe2+aq
+0.2
Eh
Hematite + Water
-4
→ Fe2+ + CO + H O.
FeCO3 + 2H + ←
2
2
0.0
H2O
-0.2
Ma
gne
ti
Sid te + W
eri
te + ater
Wa
ter
H2
-0.4
-0.6
-0.8
-1.0
0
2
4
131
6
8
10
12
14
pH
FIG. 7.17. Stability of hematite, magnetite, siderite, and aqueous
species at 25°C and 1 atm total pressure. The limits of the siderite
field (shaded) are drawn to show its probable maximum stability
in surficial weathering environments. Boundaries involving aqueous species have been drawn by assuming that the total activity
of dissolved species is 10−6.
To form siderite in a weathering environment by either of
these reactions, water must be present. Consequently, siderite’s
lower stability limit is imposed by the lower stability limit of
water, which we found to be:
Eh = 0.0 − 0.059 pH.
The siderite-hematite line coincides with this limiting condition when both lines have the same y intercept; that is, when
0.0 = 0.286 + 0.059 log10 PCO , which occurs when PCO = 1.4 ×
2
2
10−1 atm. The siderite-magnetite line, however, coincides with
the breakdown of water when 0.0 = 0.319 + 0.089 log10 PCO ,
2
which is true when PCO is almost 20 times higher, at 2.6 × 10−4
2
atm. We conclude, therefore, that if the partial pressure of CO2
in a weathering environment < 2.6 × 10−4 atm, no siderite is
formed. The stable phase, instead, would be magnetite.
The upper stability limit of siderite produced during weathering is set by the natural availability of CO2. An absolute
upper limit on PCO is 1 atm, which we also adopted earlier as
2
the maximum possible value for either PO or PH . A more real2
2
No free electrons are involved, so this boundary is a vertical line.
For the maximum PCO we chose above, assuming that aFe2+ =
2
10−6, this line lies at pH = 6.79. This condition puts the sideriteFe2+ boundary at almost the same place as the magnetite-Fe2+
boundary shown in figure 7.16.
We conclude from these various constraining equations that
siderite does not form easily on exposed rock and soil, where
PCO is at the atmospheric value of 3.2 × 10−4 atm. In more CO22
rich environments such as tropical soils, however, siderite may
form under roughly the same low Eh conditions in which magnetite is normally stable. These might exist, for example, where
decay of organic matter creates a local reducing environment.
If we drew figure 7.17 at PCO values between 2.4 × 10−4 atm and
2
1 × 10−1 atm and assembled the drawings in an animated sequence, we would see the siderite field grow at the expense of
the magnetite field, squeezing it totally out of existence when
PCO reached ∼8 × 10−2 atm.
2
Activity-Activity Relationships:
The Broader View
Instead of thinking of figure 7.17 as a single frame
in an animated sequence, as suggested in worked problem 17.4, you might think of it as a slice out of a threedimensional diagram that recognizes PCO2 as a variable
in addition to Eh and pH. Similar diagrams for systems
containing sulfur species or phosphate or nitrate are also
common in the geochemical literature. Two-dimensional
sections through any of these can be drawn and interpreted by analogy with the Eh-pH diagram for the system
iron-water. Depending on how you slice the multivariable figure, your finished product might be a PCO2-pH
diagram, a PCO2-PS 2 diagram, an Eh-PS 2 diagram, or any
number of other possibilities. It is worth noting, in fact,
that all of the redox equations we have discussed are similar in form to the solubility expressions we considered
in the opening sections of this chapter, and therefore lead
to similar graphical representations. You may have been
132
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
struck with the similar appearance of Eh-pH diagrams
and earlier figures such as figure 7.5. In each case, the
stabilities of minerals or fluids have been expressed in
terms of relative activities of key species. Generically,
these are called activity-activity diagrams. We will introduce a number of such diagrams in later chapters.
We want to emphasize that activity-activity diagrams
are not the exclusive property of geochemists who study
chemical weathering, even though we have presented
them in that context. In fact, figure 7.5 first gained broad
use among geochemists through studies of wall-rock alteration by hydrothermal fluids (Hemley and Jones 1964;
Meyer and Hemley 1967), and is in wide use by economic geologists. Eh-pH diagrams also are applied to a
variety of problems in mining and mineral extraction.
Consider the following problem as an example.
Worked Problem 7.5
Much of the world’s mineable uranium occurs as UO2 (uraninite) in sandstone-hosted deposits, in association with both
vanadium and copper minerals. What redox conditions are
necessary to form such a deposit, and how can that information be used as a guide to mining uranium?
Before we address the question directly, let’s review a few
features of solution chemistry relevant to the geologic occurrence of uranium. Uranium can exist in several oxidation states,
of which U 4+ and U6+ are most important in nature. In highly
acidic solutions, U 4+ can occur as a free ion under reducing con-
ditions or can take part in fluoride complexes. More commonly,
however, tetravalent uranium combines with water to form
soluble hydroxide complexes. Under reducing conditions, uranium may also be present in solution in its 5+ state as UO2+. The
tendency for uranium to combine with oxygen is even more
pronounced when it is in the hexavalent state. Throughout
most of the natural range of pH, U6+ forms strong complexes
with oxygen-bearing ions such as CO32−, HCO3−, SO42−, PO43−,
and AsO43−, which are present in most oxidized stream and
subsurface waters. At 250°C and with a typical groundwater
PCO of ∼10−2 atm, the most abundant of these are the uranyl
2
carbonate species, which are stable down to a pH of ∼5. Langmuir (1978), however, has shown that phosphate and fluoride
complexes can be at least as abundant as the uranyl carbonate
species in some localities (see fig. 7.18). Below pH = 5, U6+ is
generally in the form of UO22+. In addition to these inorganic
species, a large number of poorly understood organic chelates
also contribute to uranium solubility.
In most near-surface environments, therefore, uranium is
easily transported in natural waters. Very little uranium can be
deposited in the 6+ state. In some localities where PCO is prob2
ably close to the atmospheric value, uranyl complexes combine with vanadate to precipitate K2(UO2 )2 (VO4 )2 (carnotite), or
Ca(UO2)2 (VO4)2 (tyuyamunite), or with phosphate or silica to
form autunite or uranophane. These constitute only a small
fraction of mineable uranium, however. Most uranium is deposited as U4+ in the major ore minerals uraninite, UO2 and
coffinite, USiO4.
Field observations that form the basis for this brief overview
are consistent with figure 7.19, in which we have calculated the
stability fields of abundant species in terms of Eh and pH from
experimentally determined stability constants. This diagram,
FIG. 7.18. Distribution of uranyl complexes versus pH for some typical ligand concentrations in
groundwaters of the Wind River Formation, Wyoming, at 25°C. PCO2 = 10−2.5 atm, ΣF = 0.3 ppm, ΣCl
= 10 ppm, ΣSO42− = 100 ppm, ΣPO43− = 0.1 ppm, ΣSi = 30 ppm. (After Langmuir 1978.)
Chemical Weathering: Dissolution and Redox Processes
133
→
4[UO2(CO3)34−] + HS − + 15H + ←
4UO2 + SO42− + 12CO2 + 8H2O,
or Fe2+ may be oxidized to form an insoluble hydroxide mineral by:
→
UO2(CO3)34− + 2Fe2+ + 3H2O ←
UO2 + 2Fe(OH)3 + 3CO2.
In either case, a uranyl complex is reduced to yield insoluble
UO2. Another commonly hypothesized reduction mechanism
involves the adsorption of uranyl complexes on hydrocarbons,
leading to gradual oxidation of the organic matter and precipitation of uraninite.
One mining technique used at several prospects in the southwestern United States involves reversing these precipitation
pathways by artificially increasing Eh within an ore body. Typically, a system of wells is drilled into a uranium-bearing sandstone unit. Through some of these, miners inject fluorine-rich
acidic solutions with a moderately high Eh. Fluids are then withdrawn by pumping from the remaining wells. The extracted
solution contains hexavalent uranium as a uranyl fluoride complex or as UO22+. In this way, uranium can be mined efficiently
without a significant amount of excavation.
FIG. 7.19. Stability of uraninite and aqueous uranium complexes
at 25°C for PCO2 = 10−2 atm. Boundaries involving aqueous species
are drawn by assuming the total activity of dissolved species is
10−6. (Modified from Langmuir 1978.)
although quantitative, has been constructed to show only those
species in the U-O2-CO2-H2O system that are stable at 25°C
with PCO = 10−2 atm, ignoring the role of other species, such as
2
those shown in figure 7.18. Even with this simplification, however, we can deduce qualitative fluid paths that may characterize
the formation of sandstone-hosted uranium deposits.
The primary source of sedimentary uranium is probably
granitic rocks, which may have as much as 2–15 ppm uranium. Surface waters, as we have found, have pH values that
typically range from ∼5 to 6.5. Eh values of these waters are
usually >0.4 volt. Therefore, during weathering of the igneous
source rocks, uranium is readily carried in solution as neutral
UO2CO30 or as UO2(CO3)22−. In earlier discussions (for example,
in worked problem 7.2), we showed that progressive weathering of aluminosilicate minerals causes an increase in pH. When
this happens, UO2(CO3)34− may also play an important role in
solution. Any of these mobile uranyl complexes, which occupy
a wide range of moderate to high Eh and pH values in figure 7.19,
may be reduced to precipitate uraninite. These often involve the
simultaneous oxidation of iron, carbon, or sulfur to create a low
Eh environment. For example, Langmuir (1978) suggests that
HS− generated by bacterial processes may be oxidized at pH 8
to SO42− by the reaction:
SUMMARY
In this chapter, we explored ways to apply our knowledge
of solubility equilibria to problems in chemical weathering. We have shown that, although silica minerals and
magnesian silicates dissolve congruently, most major
rock-forming minerals decompose to produce secondary
minerals. Prominent among these secondary products
are clay minerals and various oxides and hydroxides.
We can examine the chemical pathways along which
these changes occur by studying either the residual
phases or associated waters. To appreciate the driving
forces for weathering, however, we need to understand
processes that control the abundance of major weathering agents: CO2, organic acids, and oxygen. We have
seen that it is often desirable to express the last of these
in redox reactions, written to show the exchange of electrons between oxidized and reduced species.
Finally, we have shown that the concepts and methods
used in this chapter can be used for problems other
than those in chemical weathering. Economic geologists,
in particular, are fond of using Eh-pH diagrams and
other activity-activity diagrams to interpret the environments of ore transport and deposition. We explore other
types of activity-activity diagrams and their uses in later
chapters.
134
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
suggested readings
The literature on chemical weathering is voluminous. Many
good references are available in the applied fields of soil science,
water chemistry, and economic geology. The following list is
intended to be representative, not exhaustive.
Barnes, H. L., ed. 1997. Geochemistry of Hydrothermal Ore Deposits, 3rd ed. New York: Wiley Interscience. (This massive
volume contains basic papers on a variety of geochemical
topics of interest to the economic geologist.)
Bowers, T. S., K. J. Jackson, and H. C. Helgeson. 1984. Equilibrium Activity Diagrams. New York: Springer-Verlag. (This
book contains only eight pages of text, but is the most extensive compilation of activity-activity diagrams available.)
Brookins, D. G. 1988. Eh-pH Diagrams for Geochemistry.
New York: Springer-Verlag. (This is an excellent, slim
book, crammed with E h-pH diagrams for geologically relevant systems.)
Colman, S. P., and D. P. Dethier, eds. 1986. Rates of Chemical
Weathering of Rocks and Minerals. New York: Academic.
(This is an excellent compendium of papers at the introductory and advanced level on mechanisms and kinetics of
chemical weathering.)
Drever, J. I. 1997. The Geochemistry of Natural Waters, 3rd ed.
New York: Simon and Schuster. (This text provides a good
introduction to the geochemistry of natural waters. Chapters 7, 8, and 12 are particularly relevant to topics in this
chapter.)
Garrels, R. M., and C. L. Christ. 1965. Solutions, Minerals, and
Equilibria. New York: Harper and Row. (This book, now out
of print, was the standard text in aqueous geochemistry for
a generation of geologists. It is still a model of clarity.)
Langmuir, D. 1996. Aqueous Environmental Chemistry. Upper
Saddle Brook: Prentice-Hall. (A good introductory level text
with examples from biochemistry and environmental chemistry as well as geology.)
The following papers were referenced in this chapter. You may
wish to consult them for further information.
Amrhein, C., and D. L. Suarez. 1988. The use of a surface complexation model to describe the kinetics of ligand-promoted
dissolution of anorthite. Geochimica et Cosmochimica Acta
51:2785–2793.
Baker, W. E. 1986. Humic substances and their role in the
solubilization and transport of metals. In D. Carlisle, W. L.
Berry, I. R. Kaplan, and J. R. Watterson, eds. Mineral Exploration: Biological Systems and Organic Matter. Englewood
Cliffs: Prentice-Hall, pp. 377–407.
Barshad, I. 1966. The effect of variation in precipitation on the
nature of clay mineral formation in soils from acid and basic
igneous rocks. Proceedings of the International Clay Conference (Jerusalem) 1:167–173.
Bennett, P. C. 1991. Quartz dissolution in organic-rich aqueous
systems. Geochimica et Cosmochimica Acta 55:1781–1797.
Cawley, J. L., R. C. Burrus, and H. D. Holland. 1969. Chemical weathering in central Iceland: An analog of pre-Silurian
weathering. Science 165:391–392.
Couturier, Y., G. Michard, and G. Sarazin. 1984. Constantes
de formation des complexes hydroxydés de l’aluminum en
solution aqueuse de 20 à 70°C. Geochimica et Cosmochimica Acta 48:649–660.
DeConick, F. 1980. Major mechanisms in formation of spodic
horizons. Geoderma 24:101–128.
Helgeson, H. C., J. M. Delany, H. W. Nesbitt, and D. K. Bird.
1978. Summary and critique of the thermodynamic properties of rock-forming minerals. American Journal of Science
278A:1–229.
Hemley, J. J., and W. R. Jones. 1964. Chemical aspects of hydrothermal alteration with emphasis on hydrogen metasomatism. Economic Geology 59:538–569.
Holland, H. D. 1978. The Chemistry of the Atmosphere and
Oceans. New York: Wiley Interscience.
Langmuir, D. 1978. Uranium solution-mineral equilibria at low
temperatures with applications to sedimentary ore deposits.
Geochimica et Cosmochimica Acta 42:547–569.
Leventhal, J. S. 1986. Roles of organic matter in ore deposits.
In W. E. Dean, ed. Organics and Ore Deposits. Wheat Ridge,
Colorado: Denver Region Exploration Geologists Society,
pp. 7–20.
Li, Y.-H. 2000. A Compendium of Geochemistry. Princeton:
Princeton University Press.
Meyer, C., and J. J. Hemley. 1967. Wall rock alteration. In H. L.
Barnes, ed. Geochemistry of Hydrothermal Ore Deposits,
1st ed. New York: Holt, Rinehart, and Winston, pp. 166–
235.
Ochs, M., I. Brunner, W. Stumm, and B. Cosovic. 1993. Effects
of root exudates and humic substances on weathering kinetics. Water, Air and Soil Pollution 681–682:213–229.
Nesbitt, H. W., G. Markovics, and R. C. Price. 1980. Chemical
processes affecting alkalis and alkaline earths during continental weathering. Geochimica et Cosmochimica Acta 44:
1659–1666.
Norton, D. 1974. Chemical mass transfer in the Rio Tanama system, west-central Puerto Rico. Geochimica et Cosmochimica
Acta 38:267–277.
Ohmoto, H., and R. O. Rye. 1979. Isotopes of sulfur and carbon. In H. L. Barnes, ed. Geochemistry of Hydrothermal
Ore Deposits, 2nd ed. New York: Holt, Rinehart, and Winston, pp. 509–567.
Petit, J.-C., G. DellaMea, J.-C. Dran, J. Schott, and R. A. Berner.
1987. Mechanism of diopside dissolution from hydrogen
depth profiling. Nature 325:705–707.
Reesman, A. L., E. E. Pickett, and W. D. Keller. 1969. Aluminum ions in aqueous solutions. American Journal of
Science 267:99–113.
Russell, E. W. 1961. Soil Conditions and Plant Growth. New
York: Wiley.
Chemical Weathering: Dissolution and Redox Processes
Schnitzer, M. 1986. Reactions of humic substances with metals
and minerals. In D. Carlisle, W. L. Berry, I. R. Kaplan, and
J. R. Watterson, eds. Mineral Exploration: Biological Systems and Organic Matter. Englewood Cliffs: Prentice-Hall,
pp. 408–427.
Solomon, D. K., and T. E. Cerling. 1987. The annual carbon
dioxide cycle in a montane soil: Observations, modeling,
and implications for weathering. Water Resources Research
23:2257–2265.
Stillings, L. L., J. I. Drever, S. L. Brantley, Y. T. Sun, and
R. Oxburgh. 1996. Rates of feldspar dissolution at pH 3–7
with 0–8 mM oxalic acid. Chemical Geology 1321(4):
79–89.
135
Sundby, B., N. Silverberg, and R. Chesselet. 1981. Pathways of
manganese in an open estuarine system. Geochimica et
Cosmochimica Acta 45:293–307.
Witkamp, M., and M. L. Frank. 1969. Evolution of CO2 from
litter, humus, and subsoil of a pine stand. Pedobiologia 9:
358–365.
Wolery, T. J. 1979. Calculation of Chemical Equilibrium between Aqueous Solution and Minerals: The EQ3/6 Software Package. NTIS Document UCRL-52658. Livermore:
Lawrence Livermore Laboratory.
Wood, W. W., and M. J. Petratis. 1984. Origin and distribution of
carbon dioxide in the unsaturated zone of the southern high
plains of Texas. Water Resources Research 20:1193–1208.
136
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
PROBLEMS
(7.1)
Using appropriate data from this chapter or literature sources, calculate the value of aH4SiO4 for which
gibbsite and kaolinite are in equilibrium with an aqueous fluid at 25°C.
(7.2)
Following the approach used in worked problem 7.2, give a qualitative description of the fluid reaction path for pure water during dissolution of potassium feldspar, assuming that amorphous silica
rather than quartz is precipitated. Write all relevant reactions.
(7.3)
How much total silica (ΣSi in ppm) is in equilibrium with quartz at pH 9? At pH 10? At pH 11?
(7.4)
A typical bottled beer has a H + activity of ∼2 × 10−5. Assuming that it remains in the bottle long
enough to reach equilibrium with the glass and that the effect of organic complexes is negligible,
what concentrations of H4SiO4, H3SiO4−, and H2SiO42− should we expect to find in the beer?
(7.5)
Assume that two species, A and B, can each exist in either an oxidized or a reduced form. For the
→ A + B , in which only one electron is transferred, what must be the
redox reaction Aox + Bred ←
red
ox
0
value of E at 25°C if the equilibrium ratios Ared /Aox and Box /Bred are found to be 1000:1?
(7.6)
What is the oxidation potential (Eh) for water in equilibrium with atmospheric oxygen (PO2 =
0.21 atm) if its pH is 6?
(7.7)
Using data from Helgeson (1969) or Li (2000), calculate the value of log10Kfo and plot a line representing the congruent dissolution of forsterite on figure 7.2. Explain why forsterite is metastable with
respect to the other magnesian silicates shown.
(7.8)
Using data from worked problem 7.1, estimate the total dissolved aluminum content (ΣAl in ppm) of
water near the mouth of the Amazon River. Assume that runoff is in equilibrium with kaolinite, ΣSi
is 11 ppm, and the pH is 6.3.
(7.9)
By writing reactions in the system Fe-O2-H2O in terms of transfer of oxygen rather than transfer of
electrons, redraw figure 7.15 as an fO2-pH diagram.
(7.10) E0 for the reaction Ce3+ → Ce4+ + e− is −1.61 volt. The ratio Ce3+/Ce4+ in surface waters in the ocean
has been estimated to be 1017. If the redox half-reaction for cerium is in equilibrium with the halfreaction Fe2+ → Fe3+ + e− at 25°C, what should be the Fe2+/Fe3+ ratio in these waters? How does
→ 4Fe2+
this value compare with the ratio you would calculate from the reaction 4Fe3+ + 2H2O ←
+ 4H+ + O2, assuming that the pH of seawater is 8.2?
(7.11) Write the half-reaction for the reference electrode shown on the left side of figure 7.17. What is its
standard potential at 25°C? What is the net reaction for this pH measurement system?
CHAPTER EIGHT
THE OCEANS AND ATMOSPHERE AS A
GEOCHEMICAL SYSTEM
OVERVIEW
In this chapter, we introduce three broad problems that
have occupied much of the attention of geochemists who
deal with the ocean-atmosphere system. The first of these
concerns the composition of the oceans and the development of what has been called a “chemical model” for
seawater. This is a comprehensive description of dissolved species, constrained by mass and charge balance
and the principles of electrolyte solution behavior we
discussed in chapter 4.
Closely linked to this problem are questions about
the dynamics of chemical cycling between the oceanatmosphere and the solid Earth. Aside from the geologically rapid processes that control the electrolyte species
in seawater, there are large-scale processes that maintain
its bulk chemistry by balancing inputs from weathering
sources against a variety of sinks. We discussed many
of these inputs and outputs in chapters 5 and 7. We
consider their global effect here, and add hydrothermal
reactions at midocean ridges and the formation of evaporites as examples of this class of problem.
Finally, to assure you that oceanic geochemistry has
room for visionaries, we discuss several ideas relating to
the history of seawater and air. This third area of study
draws on our understanding of electrolyte chemistry and
our understanding of cycling processes to help interpret
ancient marine environments and events.
COMPOSITION OF THE OCEANS
A Classification of Dissolved Constituents
The oceans provide geochemists with a very large
workshop in electrolyte solution chemistry. In one sense,
this workshop is boringly uniform: its composition is
dominated by six major elements (Na, Mg, Ca, K, Cl,
and S). As early as 1819 Alexander Marcet, professor of
Chemistry in Geneva, determined that the relative abundances of these six elements are very nearly the same all
over the world. This is true even though the total dissolved salt content (the salinity) of the open ocean varies
from 33‰ to 38‰. (It is common practice to report salinity measurements in parts per thousand by weight, or per
mil [‰]. Thus, seawater that has a total dissolved salt
content of 3.45 wt % has a salinity of 34.5‰.) Of these
six elements, only calcium has been shown to vary in its
ratios to the other five by a small but measurable amount.
Constituents whose relative concentrations remain constant in seawater are identified as conservative. Variations
in their absolute abundance can be attributed solely to the
addition or subtraction of pure water to the oceans.
137
138
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Most often, geochemists focus their attention on the
remaining elements, the minor constituents of seawater.
Some of them (such as B, Li, Sb, U, and Br) are also
conservative, but many vary dramatically relative to local
salinity. The latter dissolved species are, therefore, nonconservative. Most of these variations are associated with
the oceans’ major role as a home for living organisms.
Plants live only in near-surface waters, where they can
receive enough sunlight to permit photosynthesis. They
and the animals that feed on them deplete those waters
in nutrient species such as phosphate, nitrate, dissolved
silica (ΣSi), and dissolved inorganic carbon (ΣCO2 = CO2
+ HCO3− + CO32−), which are converted to organic and
skeletal matter. These materials settle toward the ocean
floor in dead organisms or in fecal matter. As they settle,
hard body parts made of carbonate redissolve and tissues
are rapidly consumed by bacteria, causing a relative enrichment of nutrients in bottom water. Of the elements
associated with ocean life, the only significant exception
to this pattern is oxygen, which is released by photosynthesis at the surface and consumed by respiration at
depth. Since the mid-1970s, geochemists have gradually
discovered that a number of trace elements, such as Sr,
Cd, Ba, and Cr, also follow nutrient-correlated abundance patterns, although it is not clear in many cases
how these elements are involved in the oceans’ biochemical mechanisms. From a geochemical viewpoint, these
measurable gradients are essential to our ability to trace
the pathways and processes of biologic, sedimentary, and
circulatory behavior in the oceans.
We have indicated the abundance and behavior of a
number of elements in the ocean in table 8.1, dividing
them into conservative and nonconservative groups.
Dissolved gases (except for those like O2, CO2, H2S,
and to some degree N2 and N2O, which are involved in
metabolic processes) follow abundance patterns that are
controlled largely by their solubility. To a good first approximation, concentration in surface waters is fixed by
the temperature of the water and the atmospheric partial
pressure of the gas. Because atmospheric gases are also
trapped in bubbles during stirring by waves, however,
their measured abundances are not controlled solely by
solubility, but vary by an unpredictable (but usually
small) amount. Human activity contributes significantly
to the abundance or distribution of a few elements. Lead,
injected into the ocean-atmosphere system by automobile emissions, is a good example.
It is apparent from the last column of table 8.1 that
organisms play a central role in determining the chemistry of the oceans, in addition to the six major conservative elements already mentioned. Of the elements
listed as nutrients or nutrient-related, five are particularly noteworthy. Phosphate, nitrate, silica, Zn, and Cd
are very nearly depleted in surface waters, and are often
referred to as biolimiting constituents of seawater, because they are scavenged so efficiently by microorganisms and, once depleted, place a limit on the biomass of
surface waters. Only one of these five is universally biolimiting, however. Phosphorus (as phosphate) alone is
required in the metabolic cycles of all organisms, and
therefore is the ultimate biolimiting element. The supply
of silica limits only the population of organisms (diatoms
and radiolarians) that form hard body parts from SiO2,
but not the vast population of those that build skeletons
with calcium carbonate. Nitrate is a limiting nutrient for
a great many species, but not for the large family of bluegreen algae, which can fix nitrogen directly from dissolved N2. Cadmium and zinc are apparently depleted
drastically in surface waters only because plants mistakenly incorporate these elements in place of phosphorus.
One element that is conspicuously absent from the
list of biolimiting constituents is carbon, about which we
have much more to say as this chapter unfolds. Notice
that, although surface waters have a lower concentration
of total dissolved carbon species than deep waters, organisms deplete surface waters by only ∼10%—not a startlingly large depletion. They are therefore leaving ∼90%
of the dissolved carbon around them untouched.
One way to appreciate the nonlimiting nature of
carbon more fully is to compare its abundance with that
of elements that are biolimiting. The atomic ratio of
phosphorus to nitrogen to carbon in organic tissue formed
in surface waters is very nearly constant at 1:15:105, a
proportion known as the Redfield ratio, after Alfred C.
Redfield, the oceanic geochemist who first determined it.
As organic matter settles into the deep oceans, it is almost entirely consumed and returned to solution as dissolved species. It is not surprising, then, that the atomic
ratio P:N in deep waters is 1:15. The deep water ratio of
P:C, however, is roughly 1:1000. This implies that ∼90%
of the dissolved carbon in deep water cannot have been
carried down in organic matter. Some was carried down
as carbonate, but this, too, is a rather small fraction. For
every four atoms of carbon fixed in organic matter, ap-
The Oceans and Atmosphere as a Geochemical System
139
TABLE 8.1. Abundance and Behavior of Selected Elements in Seawater
Element
Surface Concentration
Li
B
C
N (N2)
(NO3− )
O (O2)
F
Na
Mg
Al
Si
P
S
Cl
K
Ca
Cr
Mn
Fe
Ni
Cu
Zn
As
Br
Rb
Sr
Cd
Sn
I
Cs
Ba
Hg
Pb
U
178 µg kg
4.4 mg kg−1
2039 µmol kg−1
Concentration at Depth
−1
<3 × 10−7 mol kg−1
205 µmol kg−1
1.3 mg kg−1
10.781 g kg−1
1.284 g kg−1
1.0 µg kg−1
2.0 µg kg−1
0.08 µg kg−1
2.712 g kg−1
19.353 g kg−1
399 mg kg−1
417.6 mg kg−1
268 ng kg−1
34 ng kg−1
8 ng kg−1
146 ng kg−1
34 ng kg−1
6–7 ng kg−1
1.1 µg kg−1
67 mg kg−1
124 µg kg−1
7.404 mg kg−1
0.3 ng kg−1
1.0 ng kg−1
48 µg kg−1
0.3 µg kg−1
4.8 µg kg−1
3 ng kg−1
13.6 ng kg−1
3.2 µg kg−1
2264 µmol kg−1 (714 m)
560 µmol kg−1 (995 m)
41 µmol kg−1 (891 m)
47 µmol kg−1 (891 m)
0.6 µg kg−1 (1000 m)
100.8 µmol kg−1 (891 m)
2.84 µmol kg−1 (891 m)
413.6 mg kg−1 (1000 m)
296 ng kg−1 (1000 m)
38 ng kg−1 (985 m)
45 ng kg−1
566 ng kg−1 (985 m)
130 ng kg−1 (985 m)
438 ng kg−1 (985 m)
1.8 µg kg−1 (1000 m)
7.720 mg kg−1 (700 m)
117 ng kg−1 (985 m)
60 µg kg−1 (1036 m)
13.3 µg kg−1 (997 m)
4 ng kg−1 (1000 m)
4.5 ng kg−1 (1000 m)
Behavior
Conservative1
Conservative1
Nutrient
Nonnutrient gas
Nutrient
Nutrient-related
Conservative
Conservative
Conservative
?
Nutrient
Nutrient
Conservative
Conservative
Conservative
Conservative2
Nutrient-related
Nutrient-related
Nutrient-related
Nutrient-related
Nutrient-related
Nutrient-related
Nutrient-related
Conservative
Conservative
Nutrient-related
Nutrient-related
Anthropogenic
Nutrient-related
Conservative
Nutrient-related
Nutrient-related
Anthropogenic
Conservative
Modified from Quimby-Hunt and Turekian (1983).
1These values are for illustration only. They are not to be taken as average values, which in
most cases are poorly known.
2Although these data indicate that typical surface and deep waters have similar calcium
contents, calcium is commonly removed from surface waters by biological precipitation of
carbonates. The net loss of Ca, however, is a small fraction of the total amount present and is
not easily discerned from global values. Therefore, Ca is nearly conservative.
proximately one atom is combined with a calcium ion
and precipitated biochemically as carbonate. (This process accounts, qualitatively, for the slightly nonconservative behavior of calcium.) We must conclude that, per
phosphorus atom in the ocean, there are roughly 870
more carbon atoms than are necessary to satisfy the
growth demands of the biomass. If you were skeptical
before, it should now be clear that carbon is by no means
biolimiting, despite its central role in the architecture of
living matter.
Chemical Variations with Depth
It should be clear that the major compositional gradients in the oceans are vertical, driven by the biologic
cycling process we have just outlined. As a way of
building a greater appreciation of these gradients, we
presented in figure 8.1 several sets of graphical data
from the northern Pacific Ocean gathered by the Geochemical Ocean Sections program (GEOSECS) during
the 1970s.
140
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 8.1. Variations in chemical and physical properties of the ocean with depth at GEOSECS
station 214 in the North Pacific Ocean. (a) Potential temperature, (b) salinity, (c) total organic
carbon, (d) silica, (e) oxygen, and (f) nitrate. (Modified from Broecker and Peng 1982.)
The first two graphs in the figure, showing vertical
gradients in temperature and salinity, are important for
understanding oceanic environments, but are controlled
by physical rather than geochemical processes. The thermal gradient shown in figure 8.1a is the result of solar
heating at the surface. Water is so opaque to visible and
infrared radiation that sunlight penetrates only a few
tens of meters before it is absorbed almost completely.
This opacity and thermal exchange with the atmosphere
result in significant warming of the top of the ocean.
Mixing by winds and currents causes limited downward
heat transport, so that the water temperature throughout most of the deep ocean is nearly constant at 2°C.
(Seawater experiences a minor temperature increase
with depth [∼0.1°C/km] due to compression, but this is
an adiabatic effect. No net internal energy change would
result if the water were slowly brought to the surface
again. For that reason, water temperatures at depth are
generally reported as potential temperatures, from which
adiabatic heating has been subtracted.) It is convenient,
therefore, to describe the ocean in thermal terms as if it
consisted of two zones: a warm surface zone, mixed by
The Oceans and Atmosphere as a Geochemical System
winds to an average depth of 70 m, and a deep water
zone of constant cold temperatures that begins at ∼1 km
below the surface. The transitional region between these
zones is called the thermocline. Depending on distance
from the equator and the time of year, the thermocline
may be very thin and close to the surface or (as in
fig. 8.1a) much thicker and deeper. Because warm water
is, in general, less dense than cold water, the shape of
the temperature profile suggests that the oceans should
be stable against thermal convection.
Salinity profiles are more complex than temperature
profiles. At most latitudes, solar heating causes enough
evaporation from the surface to cause an increase in
salinity in waters above the thermocline. Despite their
higher salt content, these waters are warm and thus less
dense than the less saline waters below. Near the poles,
however, evaporation is minimized and the production
of sea ice tends to form cold, dense saline water, which
descends to the ocean floor. Global circulation draws
large amounts of this cold bottom water (largely from
the North Atlantic Ocean and the Weddell Sea off of
Antarctica) into more temperate latitudes, where it gradually displaces local bottom water and causes a gentle
upward circulation. Over large areas of the world, this
pattern is further complicated by the lateral transport of
cold, low-salinity melt waters (as in the North Pacific,
for which the profile in fig. 8.1b was constructed) or
warm, high-salinity waters from shallow seas (as in the
mid-Atlantic, where Mediterranean waters enter through
the Straits of Gibraltar). As a result, the oceans have
multiple layers of high and low salinity, created as global
circulation injects waters of varying density into the
water column.
Although figure 8.1a,b is important to our appreciation of the physical behavior of the oceans, the remaining
profiles (fig. 8.1c–f) are more interesting geochemically,
because they help to illustrate the effects of the biologic
“pumping” mechanism. Total inorganic carbon and silica show depth profiles that are roughly mirror images
of the thermal profile. Together, these delineate the zone
of photosynthesis and surficial biologic activity. The only
difference between them is that silica, being biolimiting,
begins at nearly zero concentration at the surface, whereas
carbon begins at ∼90% of its value in bottom water, as
we discussed earlier. Barium is shown in table 8.1 as an
example of a nutrient-related, but not biolimiting, constituent. It is not yet clear how most elements like barium
are incorporated into organic matter or skeleta, but their
141
association with such constituents as silica (in hard body
parts) and carbon (largely in tissues) is quite apparent
from their similar depth profiles.
Oxygen and nitrate deserve particular attention, because their profiles demonstrate the influence of both
biologic mechanisms and global circulation. Near the
surface, the oxygen content of seawater is constrained
by its solubility in thermodynamic equilibrium with the
atmosphere. In fact, surface waters generally have excess
O2 as the result of local photosynthesis. Just below the
thermocline, however, the oxygen level drops precipitously as O2 is consumed by the oxidation of organic
matter raining down from above. Nitrate is released during this process, so its profile is almost a mirror image of
that for oxygen. Dissolved oxygen content has not been
found to reach zero at any place in the modern oceans,
but there are inland seas such as the Black Sea that are
anoxic below the thermocline. The sedimentary record
indicates that anoxia was more common in pre-Cenozoic
oceans.
In polar waters, the story is somewhat different. Because of the low temperatures, organisms do not consume
all of the nutrient nitrate and phosphate at the surface.
The rain of organic matter toward to bottom is therefore
less intense. As a result, the oxygen minimum below the
thermocline is less pronounced and the nutrient content
of polar bottom water is slightly lower than is observed
at the base of the thermocline in warmer latitudes. When
this water is transported by the deep global currents discussed earlier, the result is that average bottom water is
much more oxygen-rich and somewhat less nutrient-rich
than average seawater. This is illustrated in figure 8.1e,f.
A profile of phosphate would be qualitatively similar to
that for nitrate.
COMPOSITION OF THE ATMOSPHERE
The three major reactive components of the atmosphere—
oxygen, water, and carbon dioxide—are exchanged
readily between the air and the oceans. Equilibrium or
kinetic factors that influence their distribution, therefore, have an important effect on marine chemistry. In
chapter 7, we considered how the atmosphere engenders
weathering on land and thus provides further controls
on the bulk composition of streams and the ocean. We
look at some of these controls again shortly.
Although some photochemical reactions that take
place in the upper atmosphere may affect geochemical
142
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
processes on the Earth, we can largely disregard the composition above ∼12 km. The lower portion of the atmosphere, known as the troposphere, contains the bulk
of its mass and is vigorously stirred by convection due to
solar heating at the planetary surface. As a result, its
composition, presented in table 8.2, is nearly uniform.
Water vapor, of course, is a major exception. Because
atmospheric PH2O is, on average, very close to the saturation vapor pressure, relatively small local changes in
temperature can make the difference between evaporation and precipitation. The conversion of liquid water to
vapor and back again is a major means for energy transport at the Earth’s surface.
In addition to the species listed in table 8.2, there are
many gases with concentrations <1 ppm. Variations in
some of these, such as ozone (O3), can have a significant
effect on organisms, but rarely have a measurable role in
controlling geochemical environments.
The noble (also called rare) gases are chemically inert,
so they have no tendency to form compounds during
chemical weathering or rock-forming processes. To a first
approximation, then, they are gradually accumulating
in the atmosphere with time. The picture is much more
complicated than this, however. Helium, for example, is
light enough to escape from the top of the atmosphere in
a geologically short time, argon is produced continually
in the crust by radioactive decay, and all of the rare gases
have been shown to bond weakly to clay surfaces and
undergo limited reburial during sedimentation. With allowances for these complexities, geochemists have used
rare gas abundances to interpret the origin and evolution
of the atmosphere and of the Earth itself, precisely because they are nonreactive. This is a topic of discussion
TABLE 8.2. The Composition of the Atmosphere
at Ground Level
Constituent
Nitrogen (N2)
Oxygen (O2)
Argon
Water
Carbon dioxide
Neon
Helium
Methane (CH4)
Krypton
Modified from Holland (1978).
Concentration
(ppm by volume)
78.084 (±0.004) × 104
20.946 (±0.002) × 104
9340 (±10)
40 to 4 × 104
320
18.18 (±0.04)
5.24 (±0.04)
1.4
1.14 (±0.01)
in chapters 14 and 15, and we preview it briefly at the
end of this chapter.
Nitrogen is also nearly inert, although a limited
amount of N2 is fixed directly by microorganisms and
incorporated by growing plants. Harvard geochemist
Heinrich Holland (1978) estimated that of the 4700 ×
1012 g of nitrogen used in marine photosynthesis each
year, only 85 × 1012 g (< 2%) is converted from atmospheric N2. Most nitrogen in the nutrient cycles of the
ocean or continents, therefore, is recycled within the
biosphere. The rates of nitrate burial in sediments and
nitrogen release by weathering and volcanism are very
small and nearly balanced, although estimates are difficult to verify. Additional values from Holland indicate
that less than one part of the atmospheric mass of N2 in
2 × 108 is lost to sedimentation each year. As a result, it
is a very good approximation to call the atmospheric
mass of nitrogen inert.
The two remaining constituents of interest in this
quick survey are oxygen and carbon dioxide, both of
which owe their abundance in the atmosphere to the
processes of the biosphere. Each year, ∼83 × 1015 g of
carbon (12% of the atmospheric reservoir) is used in
photosynthesis. The process liberates 2.2 × 1017 g of
O2. Estimating the total mass of atmospheric O2 to be
1.2 × 1021 g, it is easy to calculate that the biosphere
could produce that amount of oxygen once every 5500
years. The shorter time in which all CO2 could theoretically be consumed by photosynthesis (∼9 yr) is largely a
function of the relative amounts of both gases in the atmosphere. In fact, the biogeochemical cycles for O2 and
CO2 are very nearly balanced. The yearly mass of CO2
fixed in organic matter is almost quantitatively oxidized
by the yearly mass of “new” O2 and returned to the
atmosphere. Neither gas, in any case, can be regarded
nonreactive, in comparison with nitrogen and the noble
gases.
Despite the dominance of biologic processes in controlling the fates of oxygen and CO2, both gases also
play a major role in the inorganic cycle of weathering
and sedimentation. In a year, ∼3.6 × 1014 g of carbon (as
CO2) are consumed in weathering continental rocks. This
amounts to <0.5% of the mass of carbon fixed annually
in organic matter. The O2 consumed in all weathering
reactions is ∼4 × 1014 g yr−1—again, a very small mass
compared with the annual amount linked to photosynthesis. These small amounts, however, dominate the
weathering process.
The Oceans and Atmosphere as a Geochemical System
It is almost true that the biologic cycle for oxygen
and carbon dioxide is independent of the weatheringsedimentation cycle. Less than 0.1% of the carbon fixed
yearly in organic matter is trapped in sediments rather
than immediately reoxidized, and an equivalent amount
is weathered out of continental rocks. The constituents
of both cycles meet in the oceans, however, and are a
major source of its chemical stability. In the next section, we concentrate on the chemistry of marine carbonate species.
CARBONATE AND THE GREAT MARINE
BALANCING ACT
143
(corresponding to equation 8.2), and
KCO2 = a HCO3− /PCO2 = 10 −1.47
(corresponding to equation 8.3).
We have now written three equations to relate five
unknowns (PCO and the activities of H 2CO3, HCO3−,
2
CO32−, and H+). If we are given any two other pieces of
nonredundant information about the system, therefore,
it should be possible to determine the activities of all five
species. The other two pieces of information vary from
one problem to another, because they are conditions of
the particular environment we wish to study.
Worked Problem 8.1
Some First Principles
A good way to begin is by considering a simplified
“ocean” that consists only of pure water in equilibrium
with carbon dioxide. We can then write four major
reactions relating species in the system to one another:
→ CO (aq),
CO2 (gas) ←
2
→ H CO ,
CO2 (aq) + H2O ←
2
3
→ HCO − + H+,
H2CO3 ←
3
HCO3−
→
←
CO32−
+
H+.
(8.1)
(8.2)
As we demonstrated in chapter 7, the first two of these
reactions can easily be combined into one. Because the
hydration of CO2 is more than two orders of magnitude
faster than the formation of H2CO3 from CO2(aq), the
combined reaction follows nearly a first-order kinetic expression. It is customary, therefore, to use the expression:
→ H CO
CO2 (gas) + H 2O ←
2
3
Suppose that we do not know the partial pressure of CO2 in the
atmosphere. During this century, after all, atmospheric PCO has
2
increased significantly, so a published value is likely to be out of
date (see fig. 8.2). How can we determine it?
One way is to collect a beaker of pure rainwater and measure its pH, on the reasonable assumption that rainwater is in
equilibrium with atmospheric CO2. (For simplicity, we ignore
other gases commonly dissolved in “pure” rainwater, even
though some of them do affect pH.) The activity of H+, therefore, is one piece of information we did not have before. One of
our two remaining equations, then, is simply:
aH+ = 10−pH,
where pH is the measured value.
(8.3)
to describe the hydration of CO2 and to speak of all
CO2(aq) as H 2CO3. No thermodynamic validity is lost
by this simplification, because the equilibrium constant
for the overall reaction is simply the product of the
constants for the two elementary reactions. The set of
thermodynamic equations to describe this system, therefore, consists of a solubility expression for CO2 and two
dissociation reactions. At 25°C, the equilibrium constants for these reactions have the values:
K1 = a HCO3− a H+ /aH2CO3 = 10 −6.35
(corresponding to equation 8.1),
K2 = aCO32− aH+ /a HCO3− = 10 −10.33
FIG. 8.2. Atmospheric CO2 concentrations recorded at the Mauna
Loa Observatory, Hawaii. Annual oscillations around the increasingly CO2-rich average (solid line) follow the growing cycle in the
Northern Hemisphere. (After Sundquist 1985.)
144
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Because it is necessary to maintain electrical neutrality, we
can often write a charge balance equation that provides another
constraint. In this case, that equation (our fifth and final one) is:
(aH+)2 = K1 aH CO .
2
3
By combining this with the expression for KCO , we get:
mH+ = mHCO − + 2mCO 2−.
3
The expression for K1, therefore, may be written as:
2
3
To calculate PCO , we solve the five equations simultane2
ously. Rainwater is a very dilute solution, so we simplify the
calculation by assuming that the activities of all aqueous species
are equal to their concentrations and the activity of water is
equal to 1:
1. From reaction 8.2, we know that:
(aH+)2 = K1KCO PCO ,
2
2
or:
pH = − 0.5 log10 (K1 KCO PCO ).
2
2
Using a little chemical sophistication, therefore, we have derived
an analytical solution that is easy to use with paper, pencil, and
a pocket calculator.
mCO 2− = K2 mHCO − /m H+.
3
3
By combining this with the charge balance expression and
performing a little algebra, we find that:
(m H+)2 = 2K2 m HCO − /(1 − m HCO −).
3
3
2. Reaction 8.1 gives us a way to express m HCO − as a function
3
of m H CO and m H+ . If we combine it with the expression for
2
3
KCO , the result is:
2
m HCO − = PCO KCO K1 /m H+.
3
2
2
3. Finally, we can substitute this expression into the result of
step one. A little more algebraic manipulation yields the
answer:
PCO = (10−pH)3/(KCO K1[10−pH + 2K2]).
2
2
To calculate the atmospheric partial pressure of CO2, then, we
insert the measured pH. If we had measured a pH of 5.65, for
example, we would conclude that PCO = 331 ppm.
2
While on this topic, let’s consider the inverse of this problem. Suppose that we had measured atmospheric PCO and we
2
wanted to know what rainwater pH to expect. The final equation produced above cannot be rewritten to get a H+ by itself on
one side, so we are forced to extract pH by solving a cubic
equation. There are several ways to do this. One way is to place
all of the terms on one side, so that the equation is in the form:
f(aH+) = 0,
write an expression for its first derivative:
df/daH+ = 3(aH+)2 − (KCO K1PCO ),
2
2
and use the Newton-Raphson method described in appendix A.
We recommend this as a simple programming exercise.
A much more common method, however, involves using our
knowledge of the H2O-CO2 system to simplify the problem.
When carbon dioxide is dissolved in water at room temperature, the resulting solution always has a pH <7. As we show
shortly, this implies that aCO 2− is negligibly small. (In fact, it is
3
∼2 × 10−5 times aHCO − under the conditions we are using.) If
3
this is the case, the charge balance equation is very nearly:
aH+ = aHCO − .
3
How do we know when it is safe to make simplifying
approximations like the one we used in worked problem
8.1? Figure 8.3a is a plot illustrating the relative importance of H2CO3, HCO3−, and CO32− in solutions of
varying pH but a fixed total activity of dissolved carbon
species. This sort of diagram was introduced in 1914 by
the Swedish chemist Niels Bjerrum and gained wide use
in chemical oceanography through the many studies of
physical chemist Lars Sillén in the 1960s. From it, we see
that there are two pH values at which the activities of a
pair of carbon species are equal. We can derive these
easily. The expression for K1 can be written to give a
ratio of:
a H2CO3 /a HCO3− = a H+ /K1.
This ratio will be equal to 1.0 if a H+ is equal to K1. Similarly, the expression for K2 can be written in the form:
a HCO3− /aCO32− = a H+ /K2,
from which we see that the activities of bicarbonate and
carbonate will be equal if a H+ equals K2. At 25°C, the
magic pH values are 6.35 and 10.33. Because these crossover points on the Bjerrum plot are several pH units
apart, we are justified in neglecting aCO32− at pH values
much below 8 or aH2CO3 at pH values much above 9.
Because the activities of two dissolved carbon species
are roughly the same in the vicinity of the crossover
points, a CO2-H2O solution is said to have a high buffering capacity in these regions. By this, we mean that the
solution pH is extremely resistant to change when we
add an acid or a base to it. We can start to explain this
feature of solution chemistry qualitatively by recognizing that hydrogen ions added around pH 6.35 or pH
10.33 are likely to be involved in reactions 8.1 or 8.2,
The Oceans and Atmosphere as a Geochemical System
145
Worked Problem 8.2
Imagine a closed container filled with one liter of a 10−3 m
Na2CO3 solution. How does the pH in the container change as
we add successive small amounts of 1.0 m HCl? To make the
calculation easier, assume that there is never any gas phase in
the container, and pretend that the volume of solution always
remains 1.0 L. Furthermore, let’s assume that all species in this
system behave ideally. If we are ever concerned about this last
assumption, we can use the procedure of worked problem 4.7
to calculate the necessary activity coefficients.
Before we tackle the general problem, let’s address a
warm-up question: what is the pH before we add any acid?
There are four unknown quantities in this particular problem:
mCO 2− , m HCO − , m H CO , and pH. We therefore need to find four
3
3
2
3
equations to define the system. Two are the defining equations
for K1 and K2 (we cannot use KCO , because there is no gas
2
phase present). Charge balance must be maintained, so the
third equation is:
m Na+ + m H+ = m HCO − + 2mCO 2− + m OH− .
3
3
We can simplify this equation by drawing on observations in
the laboratory, where we know from experience that a Na2CO3
solution is quite alkaline. Before we begin adding acid, therefore, mH + should be negligibly small, so the charge balance
expression is very nearly:
mNa+ = m HCO − + 2mCO 2− + mOH− .
3
FIG. 8.3. (a) Bjerrum plot for a solution with ΣCO2 = 1 × 10−3 m.
(b) Titration curve, showing the effect on pH of adding measured
volumes of 1.0 m HCl to the solution in (a). (c) d(pH)/dV versus
volume of added acid for the titration in (b). Notice the changes
in each curve at the titration end points.
3
(Don’t worry about the sudden appearance of OH− in the problem. We have, of course, added another unknown species, but
we have also implicitly gained another equation to describe the
dissociation of water:
Kw = a H+ aOH− /aH O,
2
or, in this problem:
Kw = m H+ mOH − = 10−14.
rather than remaining as free ions in solution. Charge
balance is maintained in the solution, even as hydrogen
ions (or OH−) are added, simply by adjusting the ratio of
H2CO3 to HCO3− or the ratio of HCO3− to CO32−. This
qualitative explanation should not satisfy you, however,
because reactions 8.1 and 8.2 must consume or release
hydrogen ions in any pH region. To see why buffering is
strongest around pH 6.35 and 10.33, we consider the
following titration problem, which was alluded to in
chapter 7 in the discussion of the weathering of carbonate rocks.
We therefore include the activity of OH − and Kw explicitly as a
fifth unknown and a fifth constraint without really making the
problem more difficult.)
The final equation could be a statement fixing the value of
m H CO , m HCO − , or mCO 2− at some measured value, as we did
2
3
3
3
in worked problem 8.1. In this case, however, we would find
it difficult to measure any of these three concentrations independently. Instead, we write a statement of mass balance. Because the system is closed, the only source of dissolved carbon
species is the Na2CO3. We know, therefore, that:
ΣCO2 = mH CO
2
3
+ m HCO − + mCO 2− .
3
3
This, too, can be simplified, because we know that m H CO must
2
3
be negligible in alkaline solutions. Therefore:
ΣCO2 = mHCO
−
3
+ mCO 2− .
3
146
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
We know both ΣCO2 and mNa+ from the stoichiometry of
Na2CO3 and the initial condition of the problem we are
solving:
ΣCO
2
= 1–2 m Na+ = 10−3 m.
The equations for K1, K2, Kw, charge balance, and mass
balance, solved simultaneously, yield a quadratic expression
in m H +:
m Na+ (m H +)2 − 2Kw(m H + + K2 ) = 0.
By inserting numerical values for mNa+, K2, and Kw, we can solve
for m H +. The initial pH turns out to be 10.56. This is clearly
high enough, in retrospect, to justify the simplifications we made
in writing the charge balance and mass balance equations.
Now, back to the original question: what happens when we
start to add 1.0 m HCl? To begin, we need to rewrite the charge
balance equation to accommodate the addition of Cl−. In doing
this, we should no longer ignore m H +, which becomes increasingly significant as more acid is added. Therefore,
m Na+ + m H + = m HCO − + 2mCO 2− + mOH − + mCl − .
3
3
We use the symbol V to represent the volume of acid added (in
milliliters) by the end of any stage in our experiment. If V is
small compared to the 1.0 L volume we started with, then mCl−
in the container at any time is defined by:
mCl −(mol L−1) = 1.0(mol L−1) × V(ml)/[103 + V](ml)
≈ 10 −3V.
The charge balance equation, therefore, can be recast in
terms of V instead of mCl − . This is a convenient substitution,
because the volume of added acid is the control variable we
want to remain at the end of this calculation. The charge balance expression now becomes:
V = (m Na + + m H + − m HCO − − 2mCO 2− − mOH −) × 103.
3
3
We now combine this equation with the expressions for K1, K2,
and Kw to reduce the number of compositional variables, producing the intermediate result:
V = (m Na+ + m H + − [K1m H CO /m H+] − [2K2 mHCO − /m H +]
2
3
3
− [Kw /m H+]) × 103.
We’re almost finished, but we need to rewrite the mass balance among dissolved carbon species to include mH CO , be2
3
cause it is increasingly important as we add acid to the system.
Therefore, we manipulate the equations for K1 and K2 once
more to derive expressions for m HCO − and mCO 2− in terms of
3
3
m H CO , m H + , and the equilibrium constants. We substitute the
2
3
results in the original mass balance equation and get:
ΣCO
2
= m H CO + (K1m H CO /m H+)
2
3
2
3
+ (K1K2 m H CO /[m H+]2).
2
3
The last step is to eliminate m H CO from this equation by com2
3
bining it with the intermediate result to get:
(
Σ
)
CO2K1(m H+ + 2K2)
V = mNa+ − —————————–—
− Kw /m H + + m H +
(m H +)2 + K1m H + + K1K2
× 103.
This equation describes a titration curve for the Na2CO3 solution. We can examine it graphically by inserting the values for
m Na+ and ΣCO2 and solving for V at a number of selected values of pH. The result is shown in figure 8.3b. Below it, in figure 8.3c, is a plot of d(pH)/dV.
While the experience of deriving and plotting the
titration curve is still fresh, we can ask again how buffering in the carbonate system works. We see now that in
the immediate vicinity of pH 6.35, m H + is very nearly the
same as K1. This simplifies the titration equation greatly:
(
)
ΣCO2(m H + + 2K2) − K /m + + m +
V ≈ m Na+ − ———————–—
w
H
H
2m H + + K2
× 103,
or, since K2 << m H + and K w << m H + under these circumstances,
V ≈ (mNa+ − Σ CO2 /2 + mH +) × 103.
Around pH 6.35, therefore, m H + is almost directly
proportional to V. As we move away from pH 6.35 in
either direction, however, this approximation breaks
down and the quadratic behavior of mH+ in the denominator of the second term of the titration equation gradually speeds up the rate of change of pH with V. From
a mathematical perspective, then, the buffering region
around pH 6.35 arises because the titration curve becomes steeper and more nearly linear when:
m H 2CO3 /m HCO3− = m H + /K1 = 1.
A similar situation arises when pH is ∼10.3, for which
value the solution is buffered by nearly equal amounts of
carbonate and bicarbonate, so that m H + ≈ K2. Beyond
pH 11, the largest changes in the titration equation are
due to the Kw /mH + term; below about pH 3, the last term
is most significant. In both regions, pH is again relatively
less affected by the second term and, therefore, changes
slowly with the addition of acid.
Notice, however, that there are two regions in which
the pH of the solution changes very rapidly with only a
small addition of acid—that is, where d(pH)/dV is maximized. These regions are centered on the end points of
The Oceans and Atmosphere as a Geochemical System
the titration at pH 8.35 (where HCO3− is the dominant
species) and 4.32 (where H2CO3 is dominant). At these
points, the (m H +)2 in the denominator of the second term
in the titration equation makes the titration curve very
sensitive to small added volumes of acid. It is commonly
said that in these regions, all of the dissolved carbon
species have been titrated to form either H2CO3 or
HCO3−. In fact, this is not true, because we can always
find all three species at any pH. However, it is true that
one dissolved carbon species is significantly more abundant than the other two.
Earlier, we used the terms conservative and nonconservative to separate elements in seawater on the
basis of whether their relative abundances remain constant from place to place in the ocean. Now we see that
the same concept can be applied to ions as well, and that
it has a special significance when we consider the ocean’s
response to processes that generate or consume hydrogen ions. Most ions in seawater, such as Na+, Ca2+, K+,
Mg2+, SO42−, NO3−, and Cl −, are unaffected by changes
in acidity, because they are not associated with species
(like bicarbonate) that contain one or more hydrogen
atoms. These are conservative ions. A handful of others,
including not only HCO3− and CO32− but also B(OH)4−,
H3 SiO4−, and a variety of organic ions, are nonconservative. Their relative abundances can therefore vary
(sometimes considerably) from one part of the ocean to
another.
To understand further how the carbonate system helps
maintain charge balance in the oceans, let’s sum the
charge equivalents in seawater due to dominant conservative cations and compare them with the summed charge
equivalents due to conservative anions:
Species
Concentration (mol/kg)
Charge (eq/kg)
Cations
Na+
0.470
Mg2+
0.053
K+
0.010
Ca2+
0.010
Sum of cation charges
0.470
0.106
0.010
0.020
0.606
Anions
Cl−
0.547
SO42−
0.028
Br−
0.001
Sum of anion charges
0.547
0.056
0.001
0.604
We commit only a small error by ignoring the remaining conservative ions, which are much less abundant.
147
The difference between our totals, therefore, is 0.002
eq kg−1. This charge deficit, referred to as the alkalinity
of the seawater, must be balanced by some combination
of nonconservative anions. To a good first approximation, this means that:
alkalinity = m HCO3− + 2mCO32− ,
because these two ions make up the bulk of the remaining dissolved species in seawater. More correctly, this
quantity is known as the carbonate alkalinity. Unless we
are making very careful measurements, the only difference between it and the true alkalinity is the omission of
m B(OH)4− , which is a small but nonnegligible quantity in
the oceans.
With this perspective, it should now be apparent
why the titration we described at length in worked problem 8.2 is often called an alkalinity titration. The amount
of acid necessary to bring a solution’s pH to its first end
point (8.35) is a measure of mCO32− , because that is the
pH at which “all” CO32− has been converted to bicarbonate. Titration to the second end point (pH 4.32)
gives us an indirect measure of ΣCO2, because at this
point, “all” of the original CO32− and HCO3− have been
converted to H2CO3. Because ΣCO2 = mCO32− + mHCO3− ,
we can use the last equation above to calculate the alkalinity of the solution.
Why should we expect mHCO3− and mCO32− in seawater
to be different from one part of the ocean to another?
Early in this chapter, we discussed how carbon is carried
from surface to deep waters in the organic and inorganic
remains of planktonic organisms. Two sorts of chemical changes take place in this process. On the one hand,
photosynthesis allows plants to consume carbon from
seawater and fix it in organic matter. As a result, ΣCO2
is lower in the surface ocean than it is in the deep ocean.
A very slight alkalinity increase results from the uptake
of dissolved NO3− by tissues. Except for that minor
contribution, however, the alkalinity of surface water is
unaffected by the formation of organic matter, because
the balance of other conservative ions is left intact. In
contrast, as carbonate shells form, both ΣCO2 and the
alkalinity of surface seawater change. Each mole of
carbon that leaves the surface ocean as a constitutent of
calcite takes a mole of calcium ions with it, carrying two
moles of positive charge. Formation of calcite, therefore,
causes the alkalinity (expressed in milliequivalents per
148
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
liter) to change by exactly twice the amount that ΣCO2
changes (in mmoles per liter).
To appreciate the combined effects of tissue and shell
formation, recall that four out of every five carbon atoms
transferred from surface to deep ocean waters are carried
in organic matter. Therefore, if we momentarily ignore
the alkalinity change due to nitrate, we find that every
five millimoles of carbon leaving the surface ocean are
accompanied by one mole of Ca2+ ions, thus reducing
the alkalinity by two milliequivalents. Even correcting
for the slight charge increase due to nitrate reduction, we
conclude that alkalinity decreases only 28% as fast as
ΣCO2 during growth of marine organisms. In deep ocean
waters, the reverse takes place. As respiration, fermentation, and dissolution return the constituents of dead
organisms to seawater, alkalinity increases at 28% of the
rate of total carbon increase.
These relative changes can also be seen as changes in
the ratio of carbonate to bicarbonate with depth. With a
little algebra, we can express mCO32− and m HCO3− in terms
of ΣCO2 and alkalinity:
mCO32− = Alkalinity −
ΣCO ,
2
FIG. 8.4. Plot of alkalinity versus ΣCO2 for major surface and
deep waters in the world’s oceans. The slope of the solid line
confirms our deduction that particulate carbon is transferred from
the surface ocean to the bottom primarily in the form of organic
debris. The ratio of CaCO3 to organic carbon transported vertically
is 1:4. (Modified from Broecker and Peng 1982.)
(8.4)
and
Σ
m HCO3− = 2 CO2 − Alkalinity.
(8.5)
This little algebraic shuffling shows us, therefore, that
the ratio of mCO 2− to mHCO3− increases as surface organ3
isms consume carbon, nitrogen, and calcium ions. In deep
waters, mCO32− /mHCO3− decreases by the same factor. Relative variations in dissolved carbon species with depth,
then, are a direct result of the life-death cycle in the
oceans. These variations occur laterally as well, as a result of differences in biologic activity from place to place
on the globe, as illustrated in figure 8.4.
Calcium Carbonate Solubility
By introducing carbon species from a source other
than atmospheric CO2, worked problem 8.2 brought
us closer to understanding carbonate equilibria in the
oceans, where the abundance of dissolved carbonate
species depends not only on equilibrium with the atmosphere, but also on reactions involving continental silicates
and marine carbonate sediments. Chemical weathering
provides a continuous source of dissolved ions, including
bicarbonate, to river water by processes already discussed
in chapter 7. These accumulate in seawater until they
reach steady-state limits imposed by solubility or adsorption equilibria. In addition to the reactions that give us
K1, K2, and KCO2, then, two solubility equilibria are important. These are:
→ Ca2+ + CO 2−,
CaCO3 ←
3
calcite
and
→ Ca2+ + CO 2−.
CaCO3 ←
3
aragonite
Carbonates of Na, Mg, and K, the other major cations
in seawater, are significantly more soluble than either calcite or aragonite, so they have little effect on alkalinity
or ΣCO2. For kinetic reasons, dolomite does not precipitate in the open ocean, so we can disregard it, too.
Calcite and aragonite, therefore, dominate carbonate
equilibria in most ocean chemistry problems. Because
most CaCO3 is precipitated from surface waters by plankton, the relative abundances of calcite and aragonite follow variations in the population of microbiota. Their
solubility equations, then, are each important in different
The Oceans and Atmosphere as a Geochemical System
marine settings. At 25°C, the solubility product constants for these reactions are:
Kcal =
10−8.34,
With the addition of CaCO3, problems in oceanic carbon chemistry become only slightly more complicated.
There are now six potentially variable quantities: PCO
2
and the activities of H2CO3, HCO3−, CO32−, H +, and Ca2+.
Their values can be determined, therefore, by solving
any six independent equations that relate them. Four of
these are the equations for K1, K2, KCO2, and either Kcal
or Karag (depending on which CaCO3 polymorph is present). As in previous examples, the remaining two may
include explicit values for one or two of the activities, or
for some combination of two or more activities (ΣCO2,
for example), or a charge balance expression.
Worked Problem 8.3
A sample of limestone consisting of pure calcite is placed in a
beaker of water and allowed to reach equilibrium with the
water and a CO2 atmosphere above it. Suppose that we can
control the partial pressure of CO2 experimentally, and that
mHCO − can be measured. How does the solubility of limestone
3
vary as a function of mHCO − and PCO ?
3
2
We know that:
K1 = a H+ a HCO − /a H
3
2
K2 a HCO
−
3
aCO2− = ———–—
.
3
K1 a H CO
If we make the simplifying assumption that the acidity of surface ocean waters is controlled by equilibrium between calcite,
seawater, and atmospheric CO2 and we use a value of 340 ppm
(3.4 × 10−4 atm) for PCO , what is the equilibrium value of pH?
2
Once again, our challenge is to solve several equations simultaneously to reduce the problem to a single equation that
will let us calculate a dependent variable (pH) as a function of
an independently measurable quantity (PCO ). By combining the
2
expressions for K1 and KCO , we find that:
2
aHCO − = KCO PCO K1 /a H +.
2
2
3
we can eliminate aHCO − by substitution to get:
3
2+
aCO 2− = K1K2 KCO PCO /a H
.
3
2
2
We can now eliminate aCO 2− as a variable by combining this
3
expression with the solubility product constant for calcite:
2
Kcal = aCa2+ aCO 2− ,
3
if the limestone is pure calcite, we can once again combine expressions to eliminate aCO 2− :
3
Kcal K1 a H CO
2
3
aCa2+ = ——————–.
2
K2 a HCO
−
3
In addition, PCO is related to a H
2
The remaining uncontrolled variable is aCa2+. The best method
for isolating it involves the charge balance equation:
Because we also know that:
2
Worked Problem 8.4
2+
aH
= aCa2+ K1 K2 KCO PCO /Kcal .
3
/PCO ,
To obtain the most reliable numerical result from the
final expression in worked problem 8.3, it is best to
perform the calculation iteratively, as in worked problem 4.7, to refine γ Ca2+ and γ HCO3−. To see how this is done,
we consider another worked problem.
3
If we divide one of these by the other to eliminate a H + and then
rearrange the result to get an expression for the activity of
CO32−, we find that:
2
3
K2 = a H + aCO 2− /aHCO − ,
3
2
3
Because we also know that:
K2 = a H + aCO 2− /a HCO − .
3
Kcal K1KCO PCO
1
2
2
mCa2+ = ——————–—
————–—.
2
2
K2 a HCO −
γCa2+ γ HCO
−
3
,
2CO3
and
2CO3
KcalK1KCO PCO
2
2
aCa2+ = ———————–
.
2
K2 a HCO
−
3
Karag = 10−8.16.
2
so we can write:
Finally, we want an answer written in terms of the measurable quantities mCa2+ and m HCO − instead of activities, so we
3
must include the appropriate activity coefficients:
and
KCO = a H
149
2CO3
by:
2mCa2+ + m H+ = 2mCO 2− + m HCO − + m OH −.
3
3
We defined a simplified “seawater” for the purposes of this
problem. You should be a little worried now, because you know
that the bulk composition of natural seawater is dominated by
many species that are missing from this charge balance expression. Stay worried; we address this shortcoming in a moment.
Meanwhile, we anticipate that pH will be somewhere between
7 and 9, so it is fair to simplify the charge balance expression
150
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
by disregarding m H + , m OH − , and mCO 2− , each of which should
3
be much smaller than either mCa2+ or m HCO −. Therefore,
3
2mCa2+ ≈ m HCO − ,
3
and
a HCO − γCa2+
3
aCa2+ ≈ —————
.
2γHCO −
3
As before, we eliminate a HCO − from this equation by replacing
3
it with KCO PCO K1 /a H+ . This yields an expression for aCa2+ that,
2
2
when inserted into the intermediate result above, finally yields:
2 K 2 K2 K γ
PCO
CO2 2 Ca2+
2 1
a 3H+ = —————————,
2Kcal γHCO −
3
from which we can extract pH. We’re done, almost.
Let’s get back to that worry. It is tempting to suggest that
we should iterate toward “best” values for γCa2+ and γHCO −, as
3
we did in chapter 4. Because we cannot ignore all of the other
species in seawater, however, the exercise should not be taken
lightly. Each cycle of iteration requires us to calculate improved
values not only for mCa2+ and the concentrations of carbonate
species, but also for those of the other major ions and complexes in seawater, plus each of their activity coefficients. The
product of this computationally intense exercise has been called
a chemical model for seawater. Several such models have been
produced and refined by geochemists, beginning with Robert
Garrels and Mary Thompson (1962), whose model we describe
briefly in a later section. Rather than expend the effort to construct and justify such a model here, we complete this worked
problem by using activity coefficients derived from experiments
with synthetic seawater (for example, those in Pytkowicz
[1983], table 5.89). Using γCa2+ = 0.261 and γHCO − = 0.532, we
3
calculate that surface ocean water should have an equilibrium
pH of 8.39. The presence of calcite in the ocean, therefore,
should maintain a pH very near to the HCO3− /CO32− titration
end-point (pH 8.32) we discussed in worked problem 8.2. In
fact, pH measurements in surface seawater (between 8.1 and
8.4) cluster very close to this value.
The solubility of calcite varies as a function of depth
in the oceans in ways that are directly related to the
steady downward rain of dead organisms. To see how
this works, consider our earlier discussion of alkalinity
and ΣCO2 variations with depth, in which we concluded
that the ratio mCO32− /m HCO3− decreases from the surface
downward. In fact, although the total concentration of
dissolved carbon is 10% higher in bottom waters, mCO32−
is lower by almost two-thirds. Figure 8.5 illustrates a
vertical profile of mCO32− at a station in the South Atlantic
Ocean, in which this decrease is apparent. Notice that
the most dramatic change occurs within the wave-mixed
FIG. 8.5. Carbonate ion variations with depth at a station in the
South Atlantic Ocean. The depth at which measured concentrations are the same as the theoretical saturation limit for calcite
is the carbonate compensation depth (CCD). (After Broecker and
Peng 1982.)
surface zone, and that rather gradual changes continue
below ∼500 m.
The two solid lines on figure 8.5 indicate the theoretical values of mCO32− in equilibrium with calcite and
aragonite. The slight increase in each with depth is due
primarily to pressure, but is also a function of decreasing temperature. To the right of either line, seawater is
supersaturated with respect to a carbonate mineral; to
the left, it is undersaturated. Notice that the measured
profile, in this example, lies above both solid lines at
depths less than ∼3300 m. Therefore, a carbonate particle falling through the upper portion of the ocean should
act as a nucleus for further growth. Below 3300 m,
though, aragonite should dissolve in seawater. Below
about 4300 m, calcite also is unstable. The lower of these
two saturation levels has been called the carbonate
compensation depth (CCD), representing the level below
which no carbonate mineral is stable.
In practice, the CCD is generally defined in kinetic
terms as the level at which the rates of carbonate growth
The Oceans and Atmosphere as a Geochemical System
151
SEARCHING FOR THE CCD
At first glance, it might seem that the easiest way to
locate the saturation depth in the oceans would be to
look for dissolution features in pelagic sediment. The
abyssal plains, however, have remarkably low relief.
Except on the flanks of seamounts, much of the ocean
floor is well below the CCD or the lysocline, and is
therefore virtually carbonate-free.
Several people have devised clever ways to measure the CCD in situ experimentally. An early, elegant method was introduced by Melvin Peterson of
Scripps Oceanographic Institute (1966), who hung
preweighed spheres of calcite at various depths in the
Pacific Ocean, suspending them on cables from a
current-monitoring buoy. After 250 days, he retrieved
the spheres and reweighed them to assess their net
gain or loss.
Several scientific teams repeated and improved
upon Peterson’s original experiment. One of the most
elegant was performed by Susumo Honjo and J. Erez
(1978), who reasoned that there could be a significant uncertainty in Peterson’s measurements if the
carbonate samples had seen different current flow
rates. Their solution was to use a small pump at
each sample depth to pull water at a constant rate
and dissolution are balanced. Rates of reaction at 2°C
are slow enough that some carbonate persists in pelagic
sediments even below the CCD. Therefore, many studies
refer instead to the lysocline, the depth at which dissolution features are first observed in carbonate sediments.
The position of the lysocline varies from place to place
in the oceans, depending on the rate of carbonate supply
and burial, its composition, and the local degree of
undersaturation.
Chemical Modeling of Seawater: A Summary
Seawater is the most widely available natural electrolyte solution and serves both to control and to record
the effects of many global geochemical processes. As we
have seen, biologic processes cause significant variations
in seawater chemistry from place to place, particularly
in vertical profiles. It is not surprising, then, that geo-
through a container with a preweighed mass of calcite or aragonite.
Direct measurements have also been made by an
apparatus that oceanographer Peter Weyl (1961) has
called a saturometer. In this apparatus, seawater is
pumped at a high flow rate through a tube filled with
calcite crystals and then stopped abruptly. The pH of
water trapped in the tube is monitored by an embedded glass electrode. If the water is initially undersaturated, then the pH rises as its acidity is titrated by
dissolving calcite. If it is initially supersaturated, then
calcite precipitation causes a drop in pH. With time,
the pH value in the tube approaches a saturation
value asymptotically, and the initial CO32− concentration can be calculated from:
mCO32−(initial) = mCO 32− (final)
× m H + (final)/m H + (initial).
This apparatus has been modified since the mid1970s, so that it is now possible to make in situ measurements by lowering a saturometer from a ship into
the deep waters (see, for example, Ben-Yaakov et al.
1974). Figure 8.5 was constructed from measurements
of this type.
chemists have invested a considerable amount of work
in understanding how ionic species are distributed in the
oceans.
To this point in the chapter, we have emphasized the
role of carbonate equilibria in controlling the chemistry
of many natural waters, including seawater, through
the mechanism of charge balance. As we have seen in
worked problem 8.4, however, the other major ions in
seawater (Mg2+, Na+, K+, Cl −, and SO42−) may modify the
simple equilibria we have described. In addition, the
high ionic strength of the oceans (∼0.7) favors the formation of ion pairs, which may influence solution chemistry significantly.
Garrels and Thompson (1962) recognized ten complexes whose association constants are large enough that
they contribute significantly to the ionic strength of seawater: MgHCO3+, MgCO30, MgSO40, CaHCO3+, CaCO30,
CaSO40, NaHCO30, NaCO3−, NaSO4−, and KSO4−. They
152
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
estimated the activity coefficients for these and for the
six major ions by using the mean salt method (see chapter 4) or, in the case of uncharged species, they assumed
that the coefficients were equal to unity. Then, by extension of the methods we have just discussed, they wrote
a series of mass and charge balance equations. These,
together with appropriate expressions for dissociation
and association constants, were solved simultaneously
to calculate the equilibrium abundances of all major
dissolved species.
Many geochemists have applied Garrels and Thompson’s method to other brine systems, or have developed
more sophisticated means of estimating activity coefficients in seawater. Rudolph Pytkowicz and coworkers
have made some of the most impressive advances in
the laboratory and in theoretical modeling of activityconcentration relationships. It is now apparent that other
complex species (particularly those involving chloride)
are also present in significant concentrations. Instead of
the mean salt method, investigators now generally employ a variety of comprehensive theoretical equations
(in the spirit of the Debye-Hückel equation, although different in form) to calculate activity coefficients. A good
review of seawater models can be found in chapter 5 of
Pytkowicz (1983).
Despite the sophistication of these studies, a variety
of problems are not easily interpreted by any thermodynamic model of seawater. The relative magnitudes of
Kcal and Karag for example, suggest that aragonite should
always be metastable relative to calcite. We know, however, that both carbonates are common in marine sediments and aragonite is particularly abundant in tropical
waters. To a certain degree, this observation can be explained by recognizing that Mg2+ enters most readily into
the structure of calcite, where it increases its solubility
slightly. Aragonite accepts magnesium ions only sparingly, and is therefore only marginally less stable than
“impure” calcite. That explanation is only part of the
answer, however. Virtually all carbonate precipitated from
the sea is nucleated biochemically, and aragonite depends
on the peculiarities of ocean life for its tenuous stability.
Some organisms, particularly among the scleractinian
corals, pelecypods, and gastropods, build shells and skeletons exclusively with aragonite. Others precipitate only
calcite, and some build with both polymorphs interchangeably. Once formed, aragonite persists metastably
until it is gradually replaced in ancient carbonate rocks
by calcite. This, then, is a kinetic problem outside the
scope of thermodynamic models of seawater.
Geochemists also observe that the concentration of
CO32− in the surface ocean is roughly five times greater
than should be expected for seawater in equilibrium with
calcite (see fig. 4.5). Carbonate precipitation by organisms, as we found earlier in this chapter, is limited by the
availability of phosphate. Why, then, doesn’t excess
carbonate precipitate inorganically? Once again, the apparent answer lies with kinetics and biology, rather than
with thermodynamics. Plankton are such voracious and
nondiscriminating eaters that they ingest not only nutrients, but also a large fraction of the suspended submicroscopic calcite particles that might otherwise act as
seeds for inorganic crystal growth. When these are later
excreted by the plankton, they carry a thin coating of
organic matter that is sufficient to retard further growth.
Carbonate ions, therefore, accumulate in surface seawater
well past the concentration at which Kcal is exceeded.
In summary, the distribution of major chemical species
in the ocean can be understood with an impressive degree of confidence if we recognize the central role of
carbonate equilibria and manipulate the proper thermodynamic equations. Few other geochemical systems have
been as well described. As the two brief examples given
above show, however, there are still problems in seawater chemistry that defy solution by standard thermodynamic modeling.
GLOBAL MASS BALANCE AND STEADY
STATE IN THE OCEANS
Examining the Steady State
To this point, this chapter has emphasized the equilibrium behavior of nonconservative species in the oceans.
Another class of geochemical problems is equally challenging. Put simply, these problems compel us to ask what
kinetic factors control the abundance of conservative or
nonconservative species. Because dissolved matter is constantly being supplied to oceans in river water, the influx
of each conservative species must be exactly balanced by
removal processes. Without such a balance, a species
abundance cannot be in steady state: it must either accumulate or vanish with the passage of time. In a similar
way, nonconservative species such as HCO3− and CO32−
may vary in concentration from one place to another in
The Oceans and Atmosphere as a Geochemical System
the ocean, but there are kinetic factors that operate to
keep the HCO3− /CO32− ratio nearly constant at each location. A steady state for nonconservative species is therefore maintained among various smaller reservoirs within
the ocean. Thus, it is a matter of great importance to
identify each of the sources and sinks for dissolved material in seawater and estimate the magnitudes of these
sources and sinks, so that we can evaluate the chemical
pathways for change in the oceans.
Worked Problem 8.5
What does the oceanic budget for sodium look like? At first
glance, this appears to be an easy problem. Sodium is the most
abundant cation in seawater (see table 8.1). Its concentration,
on average, is 10.78 g kg−1. The total mass of sodium in the
oceans, then, is:
A Na+ = 10.78 g kg−1 × 1.4 × 1021 kg = 15.1 × 1021 g.
The average concentration of Na+ in river water is 6.9 mg
kg−1. Rivers supply ∼4.6 × 1016 kg of water to the oceans each
year, so the rate of sodium supply is:
dA Na+ /dt = 6.9 mg kg−1 × 4.6 × 1016 kg yr−1
× 0.001 g mg−1 = 3.2 × 1014 g yr−1.
At that rate, it would take the world’s rivers only 4.7 × 107
years to deliver the ocean’s entire present mass of sodium. Using
earlier and somewhat less accurate numbers, the nineteenthcentury Irish geologist James Joly calculated a value of 9.0 ×
107 years, which he then concluded was the age of the Earth.
This was a clever calculation, and the results were remarkably
good for the period. His interpretation was flawed, however,
because he failed to recognize the existence of sinks as well as
sources for sodium.
We now know the age of the Earth to be much greater than
47 or even 90 million years. Furthermore, by studying the
compositions of formation brines and the stratigraphic record
of evaporate formation, Heinrich Holland (1984) concluded
that the sodium content of the oceans probably has not varied
by >30% during the Phanerozoic. This implies that the river
input of sodium is very nearly balanced by processes that remove sodium from the oceans; that is, the oceanic concentration of Na+ has been at or very close to a steady state value
during the past 600 million years. Holland has presented further observations that suggest that the concentration of NaCl in
seawater has varied by less than a factor of ±2 in the past 3.5 ×
109 years. Instead of yielding the age of the oceans, as Joly supposed, the balanced source and loss rates give us a measure of
the mean residence time for sodium in the ocean. The average
sodium ion, in other words, spends 4.7 × 107 years in the ocean
before it is removed by some loss process.
153
The best available data (Holland 1978) suggest that a little
more than 40% of the sodium in river water is derived from the
solution of halite in evaporates, ∼35% is derived from weathering of all other rocks, and the remaining 25% comes from sea
spray carried inland by winds. The major removal processes
from seawater include precipitation in evaporates, burial in
pore waters in marine sediments, reaction with silicates in submarine geothermal systems, and atmospheric transport of sea
spray. On the basis of experimental results reported from several labs, Holland estimated that the oceans annually lose between 1.0 and 1.4 × 1014 g Na+ as seawater circulates through
midocean ridge systems. This corresponds roughly to 30 or
40% of the annual river flux. Burial of sodium in interstitial
waters in the sediment column may account for another 15%
of the annual river flux. Estimates of sea salt transport by winds
are highly variable, but it is fairly clear that sodium derived
from sea spray is carried onshore at roughly the same annual
rate that rivers return it to the oceans. We reported this value
above to be ∼25% of the river flux of sodium per year. The
remainder, ∼40% of the annual input of sodium, must be lost
by formation of evaporates.
We present these estimates here and in figure 8.6 as an abbreviated example of the sort of mass balance exercise that is
needed to account for the provenance of each dissolved constituent in seawater and to evaluate the geochemical pathways
they follow. You should not be misled, however, by the apparent confidence with which we have reported magnitudes of the
loss processes. The rates of these processes for most dissolved
species are poorly known. By some estimates, for example,
formation of evaporates may account for as much as 60% of
the annual river flux of sodium. Consequently, this implies that
the importance of other loss processes has been exaggerated.
FIG. 8.6. A simple box model illustrating major reservoirs, pathways, and fluxes for sodium.
154
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Even the oceanic budget for sodium, then, is not a simple matter to calculate.
reservoir fluxes that have large uncertainties, and we must
be prepared to revise them frequently.
How Does the Steady State Evolve?
On occasion, ignorance about some fundamental
pathways has led geochemists to overestimate greatly the
importance of other pathways. The mass balance cycle
for magnesium is a good example. Rivers annually carry
1.4 × 1014 g of Mg2+ per year to the oceans (Meybeck
1979), the result of weathering both silicates and dolomitic carbonates on the continents. If the magnesium
budget of the oceans is in a steady state, then some combination of processes must remove an equal amount of
magnesium each year. Until the early 1970s, geochemists
believed that the most important loss process was the
formation of authigenic clays (glauconite) in marine sediments. The only other known loss process was dolomitization, widely recognized as negligible in today’s oceans.
As more analyses of marine clays became available,
however, it quickly became apparent that authigenic clay
formation is insufficient to balance the river flux. The
“magnesium problem” became most acute when James
Drever’s (1974) research could not authenticate any
clearly authigenic magnesian marine clays. Laboratory experiments and the observations gathered during the Deep
Sea Drilling Program of the 1970s and 1980s have shown
that reactions between seawater and hot basalt, previously thought to be inconsequential, remove virtually all
of the magnesium from seawater as it cycles through midocean ridges. Estimates of the cycling rate itself depend on
the inferred distribution of heat and its rate of loss around
ridges. Because these estimates involve significant uncertainties, calculations of the total amount of magnesium
removed from the oceans by this process per year are still
imprecise. It is now clear, however, that a major portion
of the annual river flux may be balanced this way. Some
observations suggest that a substantial amount of magnesium is also lost during alteration of basalts at bottom
water temperatures (∼0°C) far from ridge systems. Minor
amounts are also deposited in magnesian calcite or are
involved in exchange reactions with marine clays, and
some, in fact, may be lost during authigenic mineralization. Although uncertainties persist, in any event, the
“magnesium problem” seems to have vanished.
Each of these two examples—the sodium and magnesium cycles—illustrates the difficulty of describing a
steady state in a system of global scale. In most cases,
we must accept estimates of reservoir contents and inter-
Geochemists have an interest in kinetic modeling that
extends beyond understanding the steady state. Clearly,
geologic conditions change through time. Paleogeographic studies indicate that the rates of seafloor
spreading and volcanism have not remained constant,
and that climatic fluctuations have induced changes in
rates of weathering, erosion, and evaporation. If the ocean
has managed to remain nearly in steady state through all
of these fluctuations, it is logical to ask what keeps the
system under control. What sorts of feedback mechanisms
allow seawater to recover from external perturbations,
and how rapidly do these mechanisms operate?
As an example of a familiar elemental cycle whose
evolving steady state has been debated quite intensely,
let’s consider the fate of carbon. In a sense, this is an
arbitrary choice, because our purpose here is to demonstrate how global-scale cycles stay in balance, not necessarily to study carbon. However, as we have already seen
in earlier parts of this chapter, the carbon cycle is of
pivotal importance to the chemistry of the oceans and
the ocean biosphere. Thus, by looking closely at the
carbon cycle, we accomplish two goals at once.
Since the late 1960s, ocean chemists have been interested in models that predict how the global carbon balance is likely to adjust to the increase in atmospheric
PCO2 illustrated in figure 8.2. These calculations often
also attempt to estimate the climatic changes that may
occur as more of Earth’s outgoing infrared radiation is
absorbed by CO2 and retained by the atmosphere (the
so-called greenhouse effect). Such predictions have great
value for agricultural, economic, and political planners.
A large number of these models have been produced,
each differing from the others in either the number of
reservoirs and pathways it considers or the numerical
approach it follows in evaluating them. We briefly consider a few of these to demonstrate that no model is
“best,” but that each is appropriate for addressing a different problem.
Box Models
The models most commonly used by geochemists describe the oceans and adjoining environments as a finite
The Oceans and Atmosphere as a Geochemical System
set of reservoirs or “boxes,” each of which is internally
homogeneous, trading mass along discrete kinetic pathways. This is the approach we used for the party example in chapter 1. As in that example, oceanic box
models assume that the mass inputs and outputs from
each reservoir normally balance each other, so that the
system as a whole is in steady state. A geochemist can
then examine variations in ocean chemistry with time by
introducing a perturbation in one or more of the reservoirs or kinetic pathways. In studying twentieth-century
anthropogenic CO2, for example, geochemist Eric Sundquist (1985) developed a series of models in which the
world ocean is divided first into four geographically distinct regions and then conceptually into a dozen or more
reservoirs. Figure 8.7 is a schematic illustration of one of
155
these models. Reservoirs within each geographic region
(the Atlantic, Pacific, Arctic, and Antarctic Oceans) represent vertical differences in water chemistry and the
intensity and style of biologic activity. These oceanic
reservoirs are supplemented by an atmospheric box and
several boxes on the continents, each characterized by a
distinct biologic community with a different metabolic
role in the carbon cycle. Sundquist then wrote a set of
differential equations to describe the transfer of carbon
among these reservoirs. The equations describe geochemical and biochemical processes (such as weathering, photosynthesis, decay) as well as physical transfer by
currents, particle settling, and advection.
Box models like Sundquist’s, describing short-term
chemical variations, can be complex. In addition to the
FIG. 8.7. A schematic box model for the short-term carbon cycle described by Sundquist (1985). The world ocean is divided into four
separate reservoirs, each of which is further divided vertically. The indicated pathways between oceanic reservoirs describe the physical
movement of water and dissolved carbon by currents and advection. The five continental reservoirs are all in the biosphere, and the
pathways between them are biochemical. In this model, continental and oceanic reservoirs communicate only through the atmosphere.
(Modified from Sundquist 1985.)
156
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
reservoirs shown in figure 8.7, for example, some shortterm models recognize geologic and biologic differences
among the continents. Others emphasize processes along
the continental margins or in marine sediments. Changes
in seawater chemistry that take place in less than a few
thousand years (the time scale over which global stirring
occurs) do not affect the entire ocean at once. The same
also is true on the continents. Therefore, the shorter our
attention span is, the more we must acknowledge that
the ocean behaves as a series of semi-independent reservoirs, and the more boxes we build into the model.
Oceanic box models that focus on changes over a few
years to a few hundred years also have to be sensitive to
biologic processes with rapid response times. The surface ocean boxes in Sundquist’s models adjust to changes
in atmospheric CO2 within two to seven years, in part
because populations of living organisms respond quickly
to the environmental pressures of food supply and climate. In the twentieth century, we saw rapid environmental changes caused by human activity. By comparison,
most geologic processes, particularly those that cause
slow changes in the positions of continents or the average
height of mountains, have operated nearly at a constant
rate during our lifetimes. Although rates of tectonic processes change, they change so slowly that biologic communities do not notice. This means that box models
designed to study the oceans over short time spans can
ignore many geologically distinct boxes and pathways
that are crucial in long-term models.
As the time scale of interest becomes longer, biologic
factors have less effect than geologic factors on system
kinetics. To develop a midrange box model, Sundquist
simplified the model in figure 8.7 by lumping the atmosphere together with all surface ocean reservoirs, land
plants, and detritus to make a single super-reservoir box.
The mean residence times of these reservoirs are all so
short that, from the perspective of several hundred years
or more, they all appear to be in equilibrium. With increasing time scale, however, Sundquist found it necessary
to add reservoirs for marine and terrestrial sediments,
which interact with the ocean-atmosphere system along
weathering and diagenetic pathways. Response times for
the modified geochemical cycles are appropriately longer
as a result.
For a more detailed account of what happens when
we expand our time frame beyond a few hundred years,
let’s take a closer look at another midrange box model.
In 1982, Wallace Broecker at Columbia University hy-
pothesized that the late Pleistocene ice sheets may have
fluctuated in response to changes in the distribution of
carbon and nutrients in the oceans. He reasoned that
because organic carbon burial takes place almost exclusively on continental shelves, it should be least effective
during a glacial epoch, when the sea level is lowest.
During deglaciation, however, the sea level rises, the
proportion of shelf area increases, and more organic
carbon should be buried. As submarine shelf area fluctuated, therefore, the balance of carbon species in shallow and deep waters should also have shifted, as should
the amount of CO2 in the atmosphere. Because global
temperatures follow PCO2, this constitutes a climatic feedback system. Broecker observed that the carbon contents
of pelagic and shelf sediments and of atmospheric gases
trapped in ancient polar ice follow qualitative trends
consistent with this scheme.
Robin Keir and W. H. Berger (1983, 1985) tested
Broecker’s hypothesis with the box model shown in figure 8.8, which consists of seven reservoirs rather than
the eighteen shown in figure 8.7. Four of these are peripheral to the ocean/atmosphere system, and are large
enough that changes in their carbon contents over 105
years are almost certainly negligible: (1) continental rocks
provide a continuous flux (Hriv ) of dissolved HCO3− to
the oceans, (2) the shelves act as a sink for CaCO3 (Ecarb )
and (3) organic matter (Eorg ), and (4) deep-sea sediments
exchange both organically and inorganically derived carbon with seawater in the deep oceans. These last interactions involve inorganic (Icarb ) and organic (Iorg ) carbon
deposition in the ocean floor, and carbonate dissolution
(ξ) from pelagic sediments. Keir and Berger characterized the transfer of carbon among reservoirs with a set
of mathematical functions and then wrote four differential equations to examine the evolution of ΣCO2 and
alkalinity (A) in the remaining two boxes (surface and
deep ocean waters). Rather than describe their approach
in detail, we invite you to recall the definitions of alkalinity and ΣCO2 and compare their equations with figure 8.8. The atmosphere and surface waters follow the
relations:
(ΣCO ) /dt = U(ΣCO
Vs d
2(tot)
2(d)
−
ΣCO ) − I
2(s)
org
− Icarb − Eorg − Ecarb + Hriv ,
and
Vs d(As)/dt = U(Ad − As ) − 2Icarb − 2Ecarb
+ νo Iorg + 2Hriv.
The Oceans and Atmosphere as a Geochemical System
FIG. 8.8. A box model for the intermediate-term carbon cycle
described by Keir and Berger (1983) to test Broecker’s (1982)
glacial hypothesis. Inorganic solids follow the paths identified as
“Carb”; organic solids follow the “Org” path. Other pathways are
dominated by dissolved or gaseous carbon species.
In the deep ocean,
(ΣCO )/dt = U(ΣCO
Vd d
2(d)
2(d)
−
ΣCO )
2(s)
+ Iorg + ξ,
and
Vd d(Ad)/dt = −U(Ad − As ) − νo Iorg + 2ξ.
By inspection, you can see that each equation consists of
source terms (with positive signs) and loss terms (with
negative signs), which may combine to produce a net increase or a decrease during any time interval of interest.
The total surface ΣCO2 (ΣCO2(tot)) is partitioned into an
atmospheric fraction and a dissolved fraction (ΣCO2(s))
on the basis of solubility equilibria, and the rate of dissolution (ξ) is controlled by the solubility of calcite in
deep waters and the rate of pelagic sedimentation. Other
parameters appearing in these equations are the volumes
of water in surface and deep ocean (Vs and Vd), the rate
of physical overturning of deep and surface waters (U),
and the ratio of alkalinity loss to CO2 production due to
oxidation of organic matter (νo ). This last factor takes
into account any changes in alkalinity due to conversions between NO3− and organic N.
157
To test the model, Keir and Berger introduced perturbations in Eorg and Icarb, as inferred from the geologic
record of Pleistocene sea levels. Given these functions
and reasonable initial choices for each of the model parameters, they then solved the four differential equations
repeatedly at closely spaced time steps to watch the evolution of atmospheric PCO and the preservation rate for
2
pelagic carbonate. Their numerical results are consistent
with the carbonate record in deep-sea drill cores and with
the CO2 trapped in air bubbles in polar ice caps, supporting Broecker’s hypothesized link between Pleistocene glacial fluctuations and the global carbon cycle.
Notice that Keir and Berger’s model ocean has only a
vertical dimension. Because global currents tend to homogenize waters within the surface and deep layers, this
simplification is characteristic of midrange models. Midrange box models also characteristically recognize that
the ocean’s slow rate of overturn (the parameter U) means
that chemical perturbations within the surface reservoir
are not readily communicated to the deep ocean, and
vice versa. Consequently, short-term biologic perturbations within surface and deep reservoirs, such as those
addressed by Sundquist, constitute a background “noise”
that is insignificant over the time scale that Keir and
Berger consider. Their model not only has far fewer
boxes, therefore, but it is also largely geologic.
Again, the longer the time span to be modeled, the
less important biologic factors are. Processes that change
only with geologic slowness act as natural feedback mechanisms to keep short-term fluctuations from getting out
of hand. For example, a planktonic community that overeats may grow quickly, but it grows only until it reaches
the limits of its food supply. The river supply of phosphate or nitrate is a function of the global weathering
rate, which remains nearly constant over periods of
hundreds or thousands of years. Boxes dominated by
biologic activity can be quite noisy, therefore, but their
effects on the ocean-atmosphere system as a whole are
severely limited by the slower kinetics of geologic
processes.
If we consider even longer time scales, biology and
structure in the ocean-atmosphere system can be ignored
altogether. For example, figure 8.6, which we used in
discussing the long-term geochemical cycle for sodium,
includes only two boxes: one for the ocean and one for
the continents. In a similar vein, Yale geochemist Robert
Berner and his colleagues (1983, 1985) have shown that
variations in the seafloor spreading rate can be related to
158
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
long-term fluctuations in atmospheric PCO2 during the
past 100 million years. Their model, in its simplest form,
includes an undifferentiated ocean reservoir with no hint
of biologic control. Mass transfer between the ocean
and various solid-earth reservoirs takes place only by the
slow geologic processes of weathering, metamorphism,
volcanism, and seawater cycling at ridge crests. The response of seawater to a perturbation, in this model, is
ultimately governed by the mean residence time for bicarbonate (105 yr), calcium (107 yr), and magnesium
(107 yr) in the oceans. Even the interglacial fluctuations
considered by Keir and Berger generate only negligible
“noise” at this time scale, and they are therefore ignored.
As in Keir and Berger’s model, the atmosphere follows changes in ocean chemistry very closely. Berner
and associates, however, have incorporated an empirical
equation in their long-term model that describes the rate
of chemical weathering as a nonlinear function of PCO2.
As the amount of carbon dioxide in the atmosphere increases, both temperature and rainfall increase as well,
and the rates of chemical weathering for continental rocks
rise rapidly. This natural mechanism provides a negative
feedback that prevents PCO2 from varying by more than
a factor of two or three. Calculated values of PCO2 since
the late Mesozoic era, from Berner et al. (1983), are
shown in figure 8.9. Kasting and Richardson (1986) have
used this model’s prediction of a rise in PCO2 ∼45 million
years ago to account for the 2–3°C global warming
that is generally inferred to have taken place during the
Eocene. Considered over the appropriate time scale,
therefore, the kinetic behavior of the oceans can be exploited to show a causal link between plate tectonics and
paleoclimate.
Continuum Models
If the general rule is that oceanic box models are most
complex when they are designed to examine processes
within short time frames, it should follow that the ultimate short-term model must identify an infinite number
of boxes. Before the advent of high-speed computers, such
an idea would have been impractical. A growing class
of continuum models, however, does essentially that by
describing chemical variations as continuous functions
that can be sampled at any random point in the ocean.
Rather than describing box-to-box mass transfer, differential equations in a continuum model describe the
physical, biologic, and chemical processes that occur in
any part of the ocean. The steady state is described by a
global-scale database of measured temperatures, salinities, ΣCO2 values, and so forth that anchor the differential equations at specific sites. Assuming that the various
parameters vary smoothly from one site to another, a
computer program solves numerical forms of the differential equations to determine the variables at intermediate locations. A chemical oceanographer can then test the
response of the model system to perturbations by altering the parameters at one or more sites and monitoring
the propagation of the calculated changes to other locations. Models of this type are similar to the global atmospheric circulation models that have steadily improved
the accuracy of weather forecasts over the past 25 years.
Although they are of growing interest to those who study
the short-term sensitivity of the ocean to natural and
human events, continuum models are highly complex
mathematically and are beyond the scope of this book.
Curious readers may want to consult issues of the journals Chemical Oceanography or Ocean Modelling.
A Summary of Ocean-Atmosphere Models
FIG. 8.9. Variations in PCO2 during the Cenozoic, as calculated by
the model of Berner et al. (1983, 1985).
We have had two goals in this discussion of models.
One is simply to introduce you to concepts that geochemists consider in modeling the ocean-atmosphere
system. No single model is appropriate for studying the
full range of kinetic behavior in the system. Depending
on the processes that interest us, time scales can range
from less than a year to many tens of millions, and reservoirs can be grouped and regrouped in many ways. The
ultimate tests of success are internal consistency and the
ability to elucidate some facet of the system’s behavior
in a way that is compatible with the geologic record. Our
The Oceans and Atmosphere as a Geochemical System
other goal is to demonstrate some dynamic features of
the chemistry of the ocean-atmosphere system and, in
particular, of the carbon cycle. The chemistry of the
oceans is resilient. If we examine the dynamic redistribution of an element such as carbon, we find that its
abundance in seawater is rarely at a true steady-state
value, but that it is also rarely far from such a state.
Short-term perturbations have a limited effect on the
ocean-atmosphere system, because the response times
of reservoirs dominated by biology are very short and
long-term geologic processes place stringent limits on the
growth of biologic communities. The geologic processes
that tend to disturb the ocean’s chemistry are themselves
gently cyclical and have periods that are long compared
with the mean residence times for many dissolved species
in the oceans. Therefore, their abundances and rates of
movement into and out of neighboring reservoirs follow
gradual changes in global tectonics from one geologic
period to another without deviating markedly from the
evolving steady state.
GRADUAL CHANGE: THE HISTORY OF
SEAWATER AND AIR
Beyond the time scales we have just considered lie broader
geochemical questions about the evolution of our planet.
What did the earliest ocean and atmosphere look like?
What pathways did they follow in arriving at their
present compositions? How has evolution of the oceanatmosphere system been linked with evolution of the crust
and the history of life on the Earth? These are highly
speculative realms, but by careful interpretation of the
rock record and the application of thermodynamic constraints, geochemists have been able to peer through the
fog of time and glimpse the outlines of the early state of
the ocean-atmosphere system.
Early Outgassing and the Primitive Atmosphere
Among the most useful bits of information we have
to work with is the Earth’s excess volatile inventory. This
tabulation, first suggested in a landmark paper by William W. Rubey (1951), is an estimate of the total amount
of carbon, nitrogen, sulfur, and water that have been
released from the solid Earth since the time of its formation. A certain fraction of this inventory is now in the
ocean-atmosphere system, but much of it has been recombined with crustal rocks during chemical weathering
159
and now resides primarily in the sedimentary column as
a constituent of carbonates, clay minerals, and fossil organic matter. In particular, roughly a third of the outgassed water and nitrogen now reside in the crust, along
with almost all of the carbon and sulfur. These are regularly recycled through the surface reservoirs of the Earth
as a result of volcanism, metamorphism, and erosion.
Geochemists’ understanding of the history of volatiles
has been dominated by two schools of thought. According to one, the volatile inventory has accumulated gradually through geologic time as the result of volcanic
emanations. The other holds that degassing of the planet
was an early, catastrophic event associated with accretion and differentiation. Clearly, some outgassing has
continued throughout the Earth’s history, so the answer
lies somewhere between these two extremes. The best evidence we have comes not from the major volatiles listed
in table 8.2, but from the noble gases that are presumed
to have followed similar outgassing histories. The advantage of working with noble gases is that they are chemically nonreactive and therefore accumulate in the atmosphere through time rather than being partially recycled
into the crust. Abundance ratios of their stable and radiogenic isotopes can, in many cases, yield valuable clues to
the history of outgassing.
When we discuss radiogenic isotopes in chapter 14,
we show how geochemists can read these clues. For now,
we present a summary diagram (fig. 8.10) derived from
Holland (1984) to illustrate outgassing curves that result
from several different assumptions regarding the present
volatile content and homogeneity of the mantle. A key
consideration is the degree to which primordial gases
(those that were part of the Earth at its time of formation) have escaped from the mantle. We show in chapter 14 that, although the mantle has lost some of its
primordial gas content, it has by no means lost all of it.
Figure 8.10 indicates that unless <30% of those gases
remain in the mantle today, most of the planet’s volatile
inventory must have been outgassed catastrophically during the earliest Archean era. In other words, although
outgassing has continued throughout the Earth’s history,
most of the volatiles now in the atmosphere and oceans
and trapped in crustal rocks were released quite early.
This result is compatible with our modern conception of
planetary formation, which is discussed in chapter 15.
We should keep in mind, though, that it depends on many
glaring uncertainties and may change dramatically as we
continue to gather data.
160
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
No samples of volcanic rock from the first half billion
years of the Earth’s existence have survived, so we have
no direct information about their composition. It is almost certainly true, however, that the oxidation state of
volcanic gases during the early Archean was no higher
than it is today. Modern volcanic gases behave as if their
oxidation state is controlled by the reaction:
→ 3Fe SiO + O ,
3SiO2 + 2Fe3O4 ←
2
4
2
quartz magnetite
fayalite
FIG. 8.10. Accumulation of noble gases in the atmosphere. The
upper curve was calculated by assuming a heterogeneous mantle.
The others were calculated by assuming that the mantle is homogeneous. Notice that the two models of mantle structure have
indistinguishable effects unless the noble gas inventory in a
homogeneous mantle is at least 70% outgassed. (Based on calculations by Holland 1984.)
In addition to determining the composition of the
Earth’s excess volatile inventory and the rate at which it
was released, a significant effort has been made to deduce the species abundances in the ocean-atmosphere
system through geologic time. Although some attempts
have been made, geochemical cycling models such as
those for the carbon cycle that we discussed earlier have
not proven useful in pre-Phanerozoic studies. Our knowledge of initial reservoir contents and rate constants is
too scanty. Also, we understand so little about indirect
controls, such as the configuration of continents and the
average depth of seas, that it is very difficult to construct
a reliable kinetic model. Instead, the chemical history
of the Precambrian oceans and atmosphere has been reconstructed by studying a succession of presumed equilibrium states through which they passed.
During planetary accretion, it seems likely that primordial gases were released from the solid Earth in equilibrium with silicate magmas. The species abundances in
the gas mixture, therefore, can be estimated theoretically
by assuming that their bulk composition was that of the
present excess volatile inventory and then using the tools
of thermodynamics to find the mixture whose free energy, in association with the magma, is lowest. The major
controlling parameter is oxygen fugacity, which is determined in most magmatic systems by the oxidation state
of iron.
commonly referred to as the QFM oxygen fugacity buffer.
This places an upper limit on the oxidation state of the
primordial atmosphere. The lower limit is probably best
defined by the reaction:
→ Fe SiO ,
SiO2 + 2Fe + O2 ←
2
4
quartz iron
fayalite
which may have been dominant before metallic iron was
segregated to form the Earth’s core. The oxygen fugacity
controlled by this assemblage, known as QFI, is about
four orders of magnitude lower than QFM at any temperature of interest in a magmatic system (see fig. 8.11).
As we shall see in chapter 15, prevailing ideas about the
differentiation of the primitive Earth favor conditions
near this lower limit.
FIG. 8.11. Calculated curves indicating the oxidation state of
magmatic systems buffered by quartz + fayalite + magnetite
(QFM) or quartz + fayalite + iron (QFI) as a function of temperature. The two curves are approximately four orders of magnitude
apart over the temperature range for most magmas.
The Oceans and Atmosphere as a Geochemical System
mospheres. The five equations can therefore be solved simultaneously to yield:
Worked Problem 8.6
Consider a closed system containing a gas phase and the solid
QFM oxygen buffer at 1100°C. The total pressure on the gas is
1.0 atm, and it consists of a mixture of C-H-O species. What
are the partial pressures of each of those species if the system is
at equilibrium and the bulk atomic ratio C:H is 0.05?
As phrased, this is a very difficult problem to do on paper,
because it involves a great number of potentially stable species.
To make it easier, let’s assume that there are only four gases in
the mixture, other than O2: H2, H2O, CO, and CO2. The partial pressures of these five, then, constitute the unknown values
we must calculate.
To solve for the five unknowns, we must write five equations. Two of these follow directly from the bulk chemistry of
the system. Because total pressure is 1.0 atm, we know that:
PH + PH O + PCO + PCO + PO = 1.0.
2
2
2
2
Because the atomic ratio C:H is 0.05, we also know that:
(PCO + PCO )/(PH + PH O ) = 0.1.
2
2
2
The partial pressure of O2 is defined by the QFM buffer at
the temperature specified. An empirical equation commonly used
for this purpose is:
log fO = 9.00 − (25738/T),
2
where T is the temperature in kelvins (Eugster and Wones 1962).
Our third equation, then, is
fO = 1.8 × 10−10.
2
The final two equations are based on the stoichiometry of
the gas species. From the potential reaction:
→ 2CO + O ,
2CO2 ←
2
we calculate that:
fCO /fCO = (fO /Kc )1/2,
2
2
in which Kc is the equilibrium constant for this reaction between carbon species. It has a value, under these conditions,
of 3.79 × 10−13.
The only other independent chemical reaction that can be
written is:
→ 2H + O ,
2H2O ←
2
2
from which:
fH O /fH = (fO /Kh )1/2.
2
2
161
2
The equilibrium constant, Kh, has the value of 8.82 × 10−14 at
1373 K.
Assuming ideality, each of the fugacities in these last three
equations is numerically equivalent to a partial pressure in at-
PH = 1.97 × 10−2 atm,
2
PH O = 0.89 atm,
2
PCO = 4.00 × 10−3 atm,
PCO = 8.70 × 10−2 atm.
2
This, of course, is a rough answer, because we limited the selection of gas species and then made the simplifying assumption of
ideality. To do the problem more carefully, we would use the
tabulated ∆Ḡf0 values and fugacity coefficients for a large number of possible C-O-H gas species to compute equilibrium mole
fractions of each for the specified bulk composition and fO .
2
The method used in worked problem 8.6 becomes
cumbersome if the number of potential gas species is
large, so a computer is a more appropriate tool than
pencil and paper. To investigate model atmospheres for
the very early Archean, we have calculated the mole
fractions of 84 gas species in the system C-O-H-S, with
its bulk composition derived from the volatile inventory
in table 8.3. The mixture, therefore, has atomic ratios
C:H = 0.032, C:N = 16.5, and C:S = 42.0. We have
assumed a total atmospheric pressure of 1.0 bars and a
magma temperature of 1200°C, but have found that
the results of the model calculation are very insensitive to
changes in either parameter. The results are applicable,
therefore, throughout the probable range of outgassing
pressures and temperatures. Values were computed at
both the QFM and QFI oxidation state limits. Here are
the concentrations of the major species and some important abundance ratios:
Species
H2O
H2
CO2
CO
SO2
H2S
N2
O2
CH4
CO2/CO
SO2/H2S
H2O/H2
QFM
QFI
0.91
0.02
0.057
0.0031
0.0013
1.3 × 10−1
0.0018
3.4 × 10−9
4.7 × 10−14
18.4
100.0
45.5
0.30
0.64
0.0097
0.051
2.0 × 10−7
0.0014
0.0018
4.0 × 10−13
2.7 × 10−7
0.19
1.4 × 10−4
0.46
Today’s volcanic gases are similar to the calculated
mixture in the QFM column. They are dominated by
162
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
TABLE 8.3. Distribution of Carbon, Nitrogen, Sulfur, and Water
in Near-Surface Geochemical Reservoirs
Carbon
(×1018 g)
Nitrogen
(×1018 g)
Sulfur
(×1018 g)
Water
(×1021 g)
Atmosphere
Biosphere
Hydrosphere
Crust
0.69
1.1
40
90,000
3950
0.02
22
2000
<0.001
<0.1
1300
15,000
≤0.2
1.6
1400
600
Inventory
90,041.79
5972.02
16,300.1
2001.8
After Holland (1984), table 3.5.
H2O. Among the carbon and sulfur species, CO2 and
SO2 are most abundant, and N2 is the major nitrogen
species. The values in the QFI column, in contrast, indicate that if metallic iron were present in very early
Archean magmas, H2, a major reduced gas species,
should have been roughly twice as abundant as water
vapor in erupting gases. Carbon monoxide should have
been five times as abundant as CO2, and H2S should
have been the dominant sulfur species. We are only partway to our goal, however. Partial pressures of atmospheric gas species should differ from those in volcanic
gases because of reactions between gases during cooling
or as the result of various loss processes. Therefore, to
the extent that thermodynamic rather than kinetic factors determined the state of the primitive atmosphere,
we can estimate the relative abundances of gas species
during the earliest Archean (before ∼4.0 × 109 years ago)
by allowing the mixture in the QFI column to reequilibrate at low temperature.
Carbon monoxide, for example, is not stable at low
temperatures in a hydrogen-rich atmosphere, and should
react to form CH4 plus water or, as PH2 drops below 10−4
atm, to form graphite plus water. Therefore, the dominant carbon species in the earliest Archean atmosphere
should have been methane. Similarly, as long as PH 2 exceeded ∼10−3 atm, the dominant nitrogen species at 25°C
should have been NH3 rather than N2. Figure 8.12 illustrates the way in which each of these equilibria depends
on the partial pressure of H2. The actual equilibrium
partial pressures of CH4 and NH3, of course, would also
have depended on the amount of outgassing, as suggested by the pair of curves in figure 8.12b.
Hydrogen is lost from the upper atmosphere to space
at a rate that is roughly proportional to its partial pressure and is rapid enough that the mean residence time for
atmospheric H2 today is four to seven years. To maintain
the high initial partial pressure suggested by our calcula-
tions, the rate of volcanic supply of H2 had to meet or
exceed its rate of loss by chemical reactions and escape
to space. The change from QFI to QFM buffering was
probably complete within the first 5 × 108 years of
the Earth’s history, when segregation of free iron from the
Earth’s mantle to the core had ended. For the next 109
years, PH2 was probably sustained at its early high value by
photodissociation of water in the upper atmosphere and
other photochemical reactions. Kasting (1986) suggested
that the dominant reactions were the photochemical reaction producing hydrogen peroxide from water vapor,
→H O +H ,
2H 2O ←
2 2
2
FIG. 8.12. Equilibrium partial pressures of major carbon- and
nitrogen-bearing gases as functions of PH2 at 25°C. (a) PCO2 and
PCH calculated for a total atmospheric carbon concentration of 1.5
4
moles cm−3. (b) PNH calculated for two values of total nitrogen
3
abundance. (After Holland 1984.)
The Oceans and Atmosphere as a Geochemical System
and the photooxidation of ferrous iron in seawater,
→ Fe3+ + 1– H .
Fe2+ + H + ←
2 2
Several lines of evidence suggest that the delicate balance
between H2 production and loss was maintained until
3.7 × 109 years ago, when PH began to decrease.
2
Studies of ancient basalts indicate that the oxidation
state of the mantle has remained essentially unchanged
since at least 3.7 × 109 years ago and possibly 5 × 108
years earlier. The average composition of volcanic gases
derived from the mantle, then, has probably also stayed
roughly the same throughout much of geologic history.
NASA scientists David Catling, Kevin Zahnle, and Christopher McKay (2001), therefore, have ascribed the atmospheric PH2 decrease during the mid- to late Archean
not to a change in the composition of outgassed volatiles,
but to a gradual increase in the rate at which H2 was lost
to space. Noting that significant amounts of organic carbon began to appear in sediments ∼3.7 × 109 years ago,
they hypothesized that anaerobic decomposition would
have released increasing amounts of methane into the
atmosphere, even though new volcanic gases buffered by
QFM should have contained a vanishingly low volume
fraction of CH4. For >109 years, until ∼2.5 × 109 years
ago, PCH4 may have been as high as 102 –103 ppm by volume, compared with 1.7 ppmv today. Such a CH4-rich
atmosphere, they estimate, would have lost H 2 to space
orders of magnitude faster than today’s atmosphere does.
This, coupled with a decrease in the total amount of mantle outgassing as the crust became progressively thicker
during the late Archean, would have led to gradually
lower atmospheric PH2.
Hydrogen sulfide, easily destroyed by solar ultraviolet
radiation or consumed in redox reactions with the crust,
should also have become less abundant during this period. Also, as PH2 continued to decrease, PCO2 should have
increased as a result of the reactions:
→ CO + 4H ,
CH4 + 2H 2O ←
2
2
followed by:
→ CO + 2H ,
C(graphite) + 2H 2O ←
2
2
as shown in figure 8.12a. As we discuss below, however,
the continual burial of CO2 as organic carbon probably
delayed the rise of atmospheric PCO until almost 2.7 ×
2
109 years ago. Only as the Archean period drew to a
close, then, did the atmosphere gradually become a mixture of CO2 and N2.
163
It is important to reemphasize that this interpretation is based on thermodynamic arguments and may not
bear any resemblance to the actual history of Archean
atmosphere-ocean chemistry. Today’s atmosphere, in fact,
does not contain the mixture of NOx gases that would be
predicted by equilibrium thermodynamics. It is likely that
the early atmosphere’s composition was also controlled
by reaction kinetics rather than thermodynamics. The true
course of events may never be known. Regardless of the
pathways, however, it is fairly certain that the atmosphere had become predominantly a CH4-CO2-N2 mixture by the time our stratigraphic record began 3.7 ×
109 years ago. The earliest surviving sedimentary rocks
already contain mature chemical weathering products,
which are formed most readily by the interaction of CO2charged rainwater with crustal rocks.
The greenhouse effect may have been important in
stabilizing surface temperatures. Of the likely atmospheric gases in the C-O-H-S system, only water, CO2,
and CH4 are effective in this role. If any of these species
is present, outgoing radiation is partially trapped, and the
surface temperature of the planet is raised. The degree
of warming is a nonlinear function of the abundance of
greenhouse gases and the amount of incoming solar radiation. Using the estimated Pleistocene global surface
temperatures as a guide, James Kasting (1986) calculated
that a CO2 partial pressure of ∼0.05 bars, or roughly
100 times the present atmospheric level, would have been
necessary to generate the same surface temperature by
the greenhouse effect during the Huronian glacial event.
Archean paleosols, however, show no evidence of the
intense acidic weathering that should have occurred in
a high-PCO2 atmosphere. Catling and coworkers (2001)
suggest that the greenhouse was largely supported instead by methane. It seems likely, then, that the evolution
of PCO2 during the Precambrian followed a path closer to
the bottom than the center of the envelope hypothesized
by Kasting in 1986 (see fig. 8.13).
As we discussed in chapter 7, H2CO3 produced in
CO2-charged rainwater is one of the most common
participants in reactions like:
→
2KAlSi3O8 + 2H 2CO3 + 9H 2O ←
K-feldspar
2K+ + Al 2Si 2O5(OH)4 + 4H 4 SiO4.
kaolinite
Clay-forming reactions, therefore, would have served
as a sink for CO2 and water from the early Archean and
164
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 8.13. Limits on the evolution of atmospheric CO2, inferred
primarily from the greenhouse-supported temperatures necessary
to avoid runaway glaciation. (Modified from Kasting 1986.)
would have increased in efficiency as PCO increased. In
2
the same way, oxidation of volcanic gases would have
served as a sink for any free molecular oxygen, creating
a primitive “acid rain” that, in turn, helped to sequester
both oxygen and sulfur in sediments during weathering.
Therefore, as chemical weathering progressed, a sizable
portion of the excess volatile inventory was transferred to
residual minerals and buried in sedimentary rocks. The
mean residence times for most dissolved species in the
oceans are short enough that these weathering reactions
would probably also have led to seawater of nearly modern composition within a few hundred million years. As
we remarked earlier, the earliest sedimentary record contains carbonates, evaporates, and shales that appear to
have been deposited from an ocean similar to today’s.
The Rise of Oxygen
The final story we have to tell in this chapter concerns the evolution of oxygen, which has been parallel to
the evolution of life. Prebiotic O2 levels were very low,
partly because there were no volumetrically significant
oxygen sources prior to photosynthesis and partly because crustal and atmospheric sinks were plentiful. The
dissociation of H 2O by solar ultraviolet radiation, followed by the escape of hydrogen into space, would have
provided a small but continuous supply of O2 in the
upper atmosphere. Molecular oxygen is easily consumed
by back reaction with H2, however, so much of the
oxygen would have been consumed before it reached
the Earth’s surface. As we noted earlier, atmospheric PH2
remained high even after hydrogen had become a less
abundant component in volcanic emanations. Kasting
(1986) has shown that photochemical reactions alone
would have guaranteed a PH2 high enough to preclude an
O2 abundance greater than about one part in 1012 at the
Earth’s surface during the prebiotic Archean.
Stromatolites resembling modern algal mats, which
are the product of cyanobacteria (blue-green algae), have
been identified in rocks ∼3 × 109 years old. The earliest
of these, however, may have been formed by photosynthetic bacteria that use reduced compounds like H 2S and
thiosulfate rather than H 2O as electron donors. There
is a growing consensus that cyanobacteria, and thus
photosynthetically generated O2, first appeared ∼2.7 × 109
years ago.
The oxidation state of various weathering products in
Proterozoic and pre-Proterozoic detrital sediments and
soils suggests that PO2 rose very slowly, if at all, for the
next 400 million years. Several rock units in Precambrian
terrain, most notably conglomerates in the Witwatersrand Basin in South Africa and at Blind River and Elliot
Lake in Canada, contain grains of uraninite and pyrite
that are almost certainly detrital. Over most of the natural range of pH and oxidation conditions today, pyrite
is readily oxidized to form insoluble iron hydroxides.
Uraninite, at pH less than ∼5, combines with O2 and
bicarbonate to form the soluble uranyl carbonate complex, UO2(CO3) 22−, and is thus easily destroyed. Between pH 4 and 6, reactions with O2 and H + also yield
UO2(OH)+. As a result, except in very arid regions or
under conditions of unusually rapid erosion and burial,
neither pyrite nor uraninite commonly accumulates in
modern sediments. The presence of both in these ancient
sediments strongly suggests that the end of the “O2-free”
era did not come until 2.3 × 109 years ago.
Until recently, geologists assumed that PO2 rose only
slowly because cyanobacteria struggled to gain a foothold for much of that period. The carbon isotopic record
in sediments, however, indicates that the rate at which
organic matter is buried has been more or less constant
since at least 3.2 × 109 years ago. This suggests that the
slow rise in atmospheric PO2 was not due to a biologic
limit on O2 production but instead to highly efficient
sinks for O2.
The primary sink seems to have been thick marine deposits of banded iron formation (BIF), which formed extensively during this period, but not since. Briefly, oxygen
The Oceans and Atmosphere as a Geochemical System
165
ARCHEAN ATMOSPHERIC GASES AND THE ORIGIN OF LIFE
The composition of the early atmosphere is of interest not only to geochemists but also to biologists, who
would like to know what environmental conditions
were most likely to have prevailed when the first biologic molecules were formed on the Earth. The Russian biochemist A. I. Oparin opened discussion on the
topic in the 1920s by theorizing that lightning or intense solar ultraviolet radiation could have produced
amino acids and other organic building blocks from
an early atmosphere rich in CH4, H2, and NH3. These
amino acids then accumulated in the oceans, forming
a primordial “soup” in which more complex hydrocarbons, such as proteins and nucleic acids, finally
developed. This notion was further reinforced as it
became apparent that the atmospheres of Jupiter and
the other outer planets were mixtures of these same
reduced C-H-N gases, plus helium. Earth’s present
atmosphere was assumed by many scientists of the
period to have evolved from a gravitationally accreted
primordial one like Jupiter’s.
The first modern experiments to shed light on
this problem were performed by Stanley Miller (1953),
who was then a graduate student in chemistry. Miller
FIG. 8.14. Sketch of the apparatus used by Stanley Miller
(1953) to generate prebiotic molecules from mixtures of CH4,
NH3, H2, and water vapor.
built a closed apparatus, illustrated in figure 8.14, into
which he introduced various mixtures of CH4, H2, and
NH3. Water was boiled in the small flask to add water
vapor to the gas mixture and encourage circulation
through the larger flask, where the gases were subjected to a spark discharge. The resulting mixture was
then condensed in a cold trap and fluxed back through
the smaller flask. After a week of operation, during
which nonvolatile products gradually accumulated in
the small flask, samples were withdrawn and analyzed
by gas chromatography.
Miller’s experiments produced a number of fairly
simple organic compounds, many of which are found
in living organisms. In all, 15% of the carbon in the
system had reacted to form amino acids, aldehydes,
or polymeric hydrocarbons. This was a dramatic and
somewhat surprising confirmation of Oparin’s hypothesis, which seemed to confirm the importance of
an early, highly reduced atmosphere.
It is a big leap from producing organic molecules
to creating life, but these experiments at least suggested that some organic building blocks for life could
have existed in the Archean ocean-atmosphere system.
Unfortunately, that’s all they tell us. Miller’s now
classic experiment has been repeated countless times
since 1953. Variants using a wide range of starting gas
compositions and different energy sources have shown
that prebiotic molecules can be formed from almost
any nonoxygenated atmosphere. These results have
blunted the scientific community’s excitement about
Miller’s initial experiment, because they do not yield
unique answers to our questions about the composition of the Earth’s early atmosphere or the earliest
organic compounds. We have learned that the proportion and absolute abundance of amino acids decreases
with an increasing oxidation state of the atmosphere,
to the extent that none have been found in experiments
on CO2, N2, H2O mixtures. Even in these experiments,
however, some biologically significant aldehydes and
simple molecules such as HCN have been observed.
So far, then, this line of investigation has shed little
light on Oparin’s original question. They suggest that
there was no lack of organic material early in Earth’s
history, but they bring us no closer to understanding
the environmental conditions for early life.
166
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
released by cyanobacteria in surface ocean waters during
the Late Precambrian—rather than entering the atmosphere as it does today—was consumed in oxidizing Fe2+
to Fe3+. The oxygen, then, was deposited in iron oxides,
carbonates, and silicates on the sea floor.
For many years, there has been a debate concerning
the primary source of iron in the Precambrian oceans
and the ultimate reason why BIF ceased to be an effective oxygen sink. According to one view (for example,
Holland 1984), marine Fe2+ was derived largely from the
global river flux, fed ultimately by continental weathering
reactions. As green-plant photosynthesis gained dominance, the river supply decreased, because iron was retained in the soils as Fe3+, and Fe2+ remaining in the deep
ocean was gradually oxidized. From this perspective, continental soils gradually displaced BIF as an oxygen sink
and then were themselves overwhelmed because atmospheric O2 could only easily affect the upper few meters
of soil. This possibility seems to be less in favor today
than one articulated in 1983 by Dutch geochemist Jan
Veizer. He proposed that iron was introduced into the
oceans by seafloor volcanism, presumed to have been
more vigorous during the first half of Earth history than
it is today. James Kasting (2001) embellished this notion
by focusing on the role of seafloor hydrothermal vent
activity and the serpentinization of oceanic basalts. Whatever the specific mechanism was, Fe2+ was released to the
oceans rapidly enough to overwhelm the photosynthetic
production of O2. The rise of PO2 occurred as the Earth’s
plate tectonic engine gradually slowed down and less iron
was being injected into the world ocean. At that point,
the Fe2+ oxygen sink, BIF, could no longer dominate
O2 production. The rather rapid PO2 increase ∼2.3 × 109
years ago, then, was triggered by a geologic change—
gradual cooling of the mantle—rather than a biologic
one.
Whether ferrous iron in the oceans was a cause or a
side effect of low atmospheric PO2, it is clear that O2 was
a minor species in the atmosphere as long as the deep
oceans were anoxic. Only after the supply of oxygen to
the seafloor exhausted the Fe2+ sink did it become possible for PO2 to rise to modern levels. The depositional
history of banded iron formation—virtually complete by
2.3 × 109 years ago—is therefore a record of the transition to an oxygen-rich atmosphere.
Antonio Lasaga and Hiroshi Ohmoto (2002) have
developed an intricate box model of the coupled C-O
geochemical cycle, with which they conclude that PO2
has stayed within a narrow range from 0.6 to 2 times its
present level since 2.3 × 109 years ago.
SUMMARY
The field of ocean-atmosphere geochemistry is extensive, and we cannot claim to have given it more than a
quick glimpse in this chapter. By giving examples, however, we have shown three of the major directions for
investigation pursued by geochemists and have suggested
other trends in research.
The composition of seawater and its variations with
depth and global location are of great interest as indicators of environmental control. In particular, chemical
species such as HCO3− and CO32− play a central role in
the ocean’s pH and alkalinity regulation and thus affect
not only geologic but also biologic processes. The pathways along which change occurs, and the network of
kinetic factors that define steady-state conditions, are
our gateway to understanding the recent history of the
ocean-atmosphere system. Geochemical cycle models can
also help to elucidate the connections between geochemistry, biochemistry, and geophysics that force long-term
changes in the system. Finally, by examining the chemistry of seawater and air from their earliest days, we gain
a perspective on the Earth’s history that bears on the
evolution of life.
suggested readings
Many books have been written on ocean-atmosphere chemistry.
The four listed here were written by some of the undisputed
leaders of research in the area and are impressive for their clarity of presentation.
Broecker, W. S., and T.-H.Peng. 1982. Tracers in the Sea. New
York: Eldigio Press, Columbia University. (A comprehensive
book on chemistry of the oceans, written as a textbook, with
many worked examples.)
Gray, J. C., and J. C. G. Walker. 1990. Numerical Adventures
with Geochemical Cycles. London: Oxford University Press.
(A good “how to” book for those interested in learning more
about modeling geochemical cycles.)
Holland, H. D. 1978. The Chemistry of the Atmosphere and
Oceans. New York: Wiley Interscience. (An exceptionally
readable discussion of processes that relate to the compositions of the ocean and atmosphere.)
Holland, H. D. 1984. The Chemical Evolution of the Atmosphere and Oceans. Princeton: Princeton University Press.
(This is the best single volume of data and ideas regarding
the history of the ocean-atmosphere system.)
The Oceans and Atmosphere as a Geochemical System
The following articles were referenced in this chapter. An interested student may wish to explore them further.
Ben-Yaakov, S., E. Ruth, and I. R. Kaplan. 1974. Carbonate
compensation depth: Relation to carbonate solubility in
ocean waters. Science 184:982–984.
Berner, R. A., A. C. Lasaga, and R. M. Garrels. 1983. The
carbonate-silicate geochemical cycle and its effect on carbon
dioxide over the past 100 million years. American Journal
of Science 283:641–83.
Berner, R. A., A. C. Lasaga, and R. M. Garrels. 1985. An improved geochemical model of atmospheric CO2 fluctuations
over the past 100 years. In E. T. Sundquist and W. S. Broecker,
eds. The Carbon Cycle and Atmospheric CO2 : Natural
Variations Archean to Present. American Geophysical Union
Monograph 32. Washington, D.C.: American Geophysical
Union, pp. 397–441.
Broecker, W. S. 1982. Glacial to interglacial changes in ocean
chemistry. Progress in Oceanography 11:151–197.
Catling, D. C., K. J. Zahnle, and C. P. McKay. 2001. Biogenic
methane, hydrogen escape, and the irreversible oxidation of
early Earth. Science 293:839–842.
Drever, J. I. 1974. The magnesium problem. In E. D. Goldberg,
ed. The Sea, vol. 5. New York: Wiley Interscience, pp. 337–
357.
Eugster, H. P., and D. R. Wones. 1962. Stability relations of the
ferruginous biotite, annite. Journal of Petrology 3:82–125.
Garrels, R. M., and M. E. Thompson. 1962. A chemical model
for sea water at 25°C and one atmosphere total pressure.
American Journal of Science 260:57–66.
Honjo, S., and J. Erez. 1978. Dissolution rates of calcium carbonate in the deep ocean: An in situ experiment in the North
Atlantic. Earth and Planetary Sciences Letters 40:226–234.
Kasting, J. F. 2001. The rise of atmospheric oxygen. Science
293:819–820.
Kasting, J. F. 1986. Theoretical constraints on oxygen and carbon dioxide concentrations in the Precambrian atmosphere.
Precambrian Research 34:205–229.
Kasting, J. F., and S. M. Richardson. 1986. Seafloor hydrothermal activity and spreading rates: The Eocene carbon
dioxide greenhouse revisited. Geochimica et Cosmochimica
Acta 49:2541–2544.
167
Keir, R. S., and W. H. Berger. 1983. Atmospheric CO2 content
in the last 120,000 years: The phosphate extraction model.
Journal of Geophysical Research 88:6027–6038.
Keir, R. S., and W. H. Berger. 1985. Late Holocene carbonate
growth in the equatorial Pacific: Reef growth or neoglaciation? In E. T. Sundquist and W. S. Broecker, eds. The
Carbon Cycle and Atmospheric CO2: Natural Variations
Archean to Present. American Geophysical Union Monograph 32. Washington, D.C.: American Geophysical Union,
pp. 208–220.
Lasaga, A. C., and H. Ohmoto. 2002. The oxygen geochemical
cycle: Dynamics and stability. Geochimica et Cosmochimica
Acta 66:361–381.
Meybeck, M. 1979. Concentration des eaux fluviales en éléments majeurs et apports en solution aux oceans. Revues
de Géologie Dynamique et de Géographie Physique 21:
215–246.
Miller, S. L. 1953. A production of amino acids under possible
primitive Earth conditions. Science 117:528–529.
Peterson, M.N.A. 1966. Calcite: Rates of dissolution in a vertical profile in the Central Pacific. Science 154:1542–1544.
Pytkowicz, R. M. 1983. Equilibria, Nonequilibria, and Natural
Waters, vol. 1. New York: Wiley Interscience.
Quimby-Hunt, M. S., and K. K. Turekian. 1983. Distribution
of elements in sea water. EOS, Transactions of the American
Geophysical Union 64:130–131.
Rubey, W. W. 1951. Geologic history of seawater, an attempt
to state the problem. Bulletin of the Geological Society of
America 62:1111–1147.
Sundquist, E. T. 1985. Geological perspectives on carbon dioxide and the carbon cycle. In E. T. Sundquist and W. S.
Broecker, eds. The Carbon Cycle and Atmospheric CO2 :
Natural Variations Archean to Present. American Geophysical Union Monograph 32. Washington, D.C.: American
Geophysical Union, pp. 5–60.
Veizer, J. 1983. Geologic evolution of the Archean-early Proterozoic Earth. In J. W. Schopf, ed. The Earth’s Earliest
Biosphere: Its Origin and Evolution. Princeton: Princeton
University Press, pp. 240–259.
Weyl, P. K. 1961. The carbonate saturometer. Journal of Geology 69:32–43.
168
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
PROBLEMS
(8.1)
A naive geochemist might expect the nitrate concentration in surface waters of the ocean to be maintained at an equilibrium concentration for the reaction:
→ 2H + + 2NO −.
N2 + –52 O2 + H 2O ←
3
Use tabulated thermodynamic data and your knowledge of PO2, PN 2, and seawater pH to estimate the
expected oceanic NO3− concentration. How does this value compare with the concentration recorded
in table 8.1? What factors might account for this difference?
(8.2)
What is the pH of a 3 × 10−3 M Na2CO3 solution?
(8.3)
Using the procedure described in worked problem 8.2, construct titration curves to show the change
in solution pH as a 3 × 10−3 m Na 2CO3 solution is titrated by (a) 0.01 m HCl, (b) 1.0 m HCl, and
(c) 10.0 m HCl. Do the end points depend on the molality of the HCl solution?
(8.4)
Suppose you were to add Fe2+ to the Na2CO3 solution in problem 8.2. How high can you raise the
Fe2+ concentration before FeCO3 begins to precipitate? Assume that the Ksp for FeCO3 is 2.0 × 10−11
and that there are no kinetic barriers to precipitation.
(8.5)
How is the alkalinity of a beaker of seawater affected by the addition of small amounts of the following substances: (a) NaHCO3, (b) NaCl, (c) HCl, and (d) MgCO3?
(8.6)
In worked problem 8.1, we made the simplifying assumption that aCO32− is negligible in rainwater,
and derived an expression to calculate pH, given PCO2. Use that expression to calculate rainwater pH
if PCO2 = 380 ppm. Can you estimate the percentage error in your approximate solution?
(8.7)
In worked problem 8.6, we calculated the composition of a gas mixture under the moderately oxidizing conditions of the QFM buffer. Redo this calculation, using instead the oxygen fugacity defined by
QFI. A useful empirical equation for this buffer is:
log fO 2 = 7.51 − 29382/T.
CHAPTER NINE
TEMPERATURE AND
PRESSURE CHANGES
Thermodynamics Again
OVERVIEW
you should be comfortable with the idea that the observable properties of a system at equilibrium will undergo no
change with time, as long as physical conditions imposed
on the system remain constant. Although this statement is
commonly used to define equilibrium, it is actually a consequence of equilibrium rather than a strict definition of
it. A more rigorous definition of the equilibrium state that
we have not seen before is: a condition of minimum energy for that portion of the universe under consideration.
We have also learned that Gibbs free energy is a convenient way of describing the energy of a system. The
relationship of this thermodynamic quantity to temperature, pressure, and the amounts of the constituent components of a system is given by the following:
This chapter expands on the thermodynamic principles
introduced in chapter 3 and provides the preparation for
the next chapter on phase diagrams. We first give a thermodynamic definition for equilibrium. We also derive the
phase rule and explore its use as a test for equilibrium.
The effects of changing temperature and pressure on the
free energy of a system at equilibrium are analyzed, and
a general equation for dḠ at P and T is formulated. P-T
and Ḡ-T diagrams for one-component systems are introduced. The utility of the Clapeyron equation in
constructing or interpreting phase diagrams is explored.
We then introduce concepts that are necessary for understanding systems with more than one component.
Raoult’s law and Henry’s law are defined, and expressions
for the chemical potentials of components that obey
these laws are derived. The concept of standard states
is introduced, and we examine how activity coefficients
are defined in mixing models. The usefulness of these
concepts is illustrated by discussing formulation of a
geothermometer and a geobarometer.
where ni represents the number of moles of component
i. The quantities (∂G/∂ni)P,T,n ≠i are chemical potentials,
j
symbolized by µi and defined in chapter 3. The goal of
this chapter is to learn how to evaluate all of the quantities in this equation.
WHAT DOES EQUILIBRIUM
REALLY MEAN?
DETERMINING WHEN A SYSTEM
IS IN EQUILIBRIUM
In previous chapters, we have introduced and made use of
the concept of thermodynamic equilibrium. By this time,
How can we determine whether a geologic system is
in a state of equilibrium? Although we can never be
dG = (∂G/∂T )P,ni dT + (∂G/∂P)T,ni dP
+ (∂G/∂ni)P,T,nj≠i dni,
(9.1)
169
170
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
completely sure, we can at least argue for equilibrium
on the basis of our experience with reactions in the laboratory and the absence of any observational evidence for
disequilibrium. Consider, for example, a metamorphic
rock consisting of biotite, garnet, plagioclase, sillimanite,
and quartz. We know from experiments that these minerals can constitute a stable assemblage under conditions
of high temperature and pressure. Let’s assume that textural observations of this rock provide no indication of
arrested reactions between any of the constituent minerals. Furthermore, grains of all five minerals are observed
to be touching each other, and no zoning or other kind
of chemical heterogeneity within grains is present. In this
case, we might infer that the rock system is in a state of
equilibrium, but this indirect evidence may be less than
convincing. Is there another way to test for equilibrium?
The Phase Rule
Let’s describe the equilibrium configuration of a system in more quantitative terms. The general relationships
among phases, components, and physical conditions are
given by the phase rule, first formulated by J. Willard
Gibbs more than a century ago. The phase rule describes
the maximum number of phases that can exist in an equilibrium system of any complexity. As we will see, this
provides an additional test for equilibrium.
The variance (f ), also called the degrees of freedom,
of a system at equilibrium is the number of independent
parameters that must be fixed or determined to specify
the state of the system. The number of independent parameters equals the number of unknown variables minus
the number of known (dependent) relationships between
them. Assume that there are p phases and every phase
contains one or more of the c chemical components that
constitute the system. For each phase p in the system, we
can write a Gibbs-Duhem equation:
0 = S̄dT − V̄dP +
Σ n dµ ,
c
c
where µc is the chemical potential of each component in
the phase. Recall also some of the major conclusions of
chapter 3; namely, that at equilibrium, the temperatures
and pressures in all phases must be the same, and µc must
be the same in all phases that contain component c. The
total number of potentially unknown variables, therefore, is c (the number of components) plus two (for
temperature and pressure, whatever their values). The
number of known relationships among them is p (the
number of Gibbs-Duhem equations). The variance, therefore, is the difference:
f = (c + 2) − p.
(9.2)
Equation 9.2 is a mathematical statement of the phase
rule. To illustrate its use, consider an ordinary glass of
water at room temperature and atmospheric pressure.
The number of components c necessary to describe the
compositions of all the phases in this system is one (H2O),
and the number of phases p is also one (liquid water).
Thus, from equation 9.2, the variance of the system is
two. This means that two variables (in this case, temperature and pressure) can be changed independently with
no resulting effect on the system. Our everyday experience tells us that this is true—it would still be a glass of
water if transported from sea level to a mountain top,
where the temperature is cooler and atmospheric pressure is less—but there is some finite limit to the permissible variation. For example, the phase in this system
(liquid water) would certainly change if we were to lower
the temperature drastically by putting the glass of water
in a freezer. However, in the case of a frozen glass of
water, the variance would still be two, because there is
still only one component (H2O) and one phase (ice in this
case). Are there situations in which the variance of this
system is other than two?
The phase diagram for water, illustrated in figure 9.1,
summarizes how this simple system reacts to changing
conditions of temperature and pressure. The diagram
consists of fields in which only one phase is stable, separated by boundary curves along which combinations of
phases coexist in equilibrium. The three boundary lines
intersect at a point, called the triple point, at which three
phases occur. The boundary between the liquid water and
water vapor fields terminates in a critical point, above
which there is no distinction between liquid and vapor.
From inspection of this figure, it is obvious that a single
phase is stable under most combinations of temperature
and pressure. Within these one-phase fields, the variance
is two, and the fields are said to be divariant.
However, there are certain special combinations of
temperature and pressure at which several phases coexist at equilibrium. For example, because two phases
coexist along the boundary curves in this diagram, the
variance at any point along the boundary curves is one.
(Prove this yourself using the phase rule.) These boundary
curves are thus univariant. Coexistence of three phases
Temperature and Pressure Changes: Thermodynamics Again
171
sion between the Ca3 Al2Si3O12 component of garnet, Al2SiO5,
SiO2, and the CaAl2Si2O8 component of plagioclase (it does not
matter that such a reaction might never take place in nature):
Ca3Al2Si3O12 + 2Al2SiO5 + SiO2 = 3CaAl2Si2O8.
FIG. 9.1. Schematic phase diagram for the system H2O. Most of
the area of the diagram consists of regions in which only one
phase is stable at equilibrium. Boundary curves are combinations
of pressure and temperature along which two phases are stable,
and three phases occur at the triple point. The boundary between
liquid water and water vapor terminates in a critical point, above
which the distinction between these phases is meaningless.
at the triple point produces a variance of zero, forming
an invariant point.
For a more complex system, such as a rock, of course,
more components are required to describe the system,
and there will probably be more phases. In an assemblage
of minerals, it is unlikely that every component will occur
in every phase. However, the phase rule is a general result and is valid for systems in which some components
do not occur in all phases.
Worked Problem 9.1
What is the variance for the metamorphic rock described at the
beginning of this section? There are five phases, and the eight
components that describe the compositional variation of each
phase are listed beside each mineral below. From our discussion
of components in chapter 3, you should recognize that the components listed below are not unique, and you might try devising
your own set of components to replace the ones given here. Note
that it would be necessary to choose additional components if
some of these minerals were more complex solid solutions:
biotite: KMg 3AlSi3O10(OH)2, KFe3 AlSi3O10(OH)2
garnet: Fe3 Al2Si3O12, Ca3Al2 Si3O12
sillimanite: Al2SiO5
Thus we can express anorthite in terms of the other three
components. In fact, any one of these four components can be
expressed in terms of the other three by rearranging the equation, and so we can delete one of these (take your choice) as a
system component. We then are left with five phases and seven
components, which give a variance of four. This means that
temperature and pressure can be varied independently (within
some reasonable limits), as well as the compositions of two
phases (for example, biotite and garnet, or garnet and plagioclase), without changing the number of phases in the system
at equilibrium.
Open versus Closed Systems
The discussion so far assumes that the system is closed.
However, there are examples of geologic systems in which
some components are free to migrate into and out of the
system. A common example of such a mobile component is the fluid that is present during the metamorphism
of some rocks. We may describe this situation using the
same Gibbs-Duhem equations that we formulated before,
but in this case, we have an additional restriction. In a
closed system, the chemical potentials of all components
are the same in every phase, but in an open system, the
chemical potentials of mobile components are controlled
by phases or other parameters outside of the system. Consequently, we have an additional constraining equation
for each mobile component. These additional chemical
restrictions r are reflected in the phase rule as additional
known relationships. The number of known relationships
therefore becomes p + r. Substitution of this value into
equation 9.2 produces the relation:
f = (c + 2) − (p + r).
(9.3)
This open-system modification provides an alternative
form of the phase rule for systems in which mobile components can be handled.
plagioclase: NaAlSi3O8, CaAl2Si2O8
quartz: SiO2
As we learned in chapter 3, not all of these components may
be necessary to define the rock system, if they are not all independent. For example, we can write a mathematical expres-
Worked Problem 9.2
Consider a rock containing calcite, quartz, and wollastonite,
which is undergoing metamorphism in equilibrium with a fluid
172
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
of H2O and CO2. Does the variance of this assemblage change
if the system is closed or open to fluid migration?
The system has four phases (H2O and CO2 in the fluid are
miscible in all proportions under these conditions). The components of each of the phases can be expressed as:
calcite: CaCO3
quartz: SiO2
wollastonite: CaSiO3
fluid: H2O, CO2
But we can reduce the number of system components
from the five listed above to four, because of the following
relationship:
CaCO3 = CaSiO3 + CO2 − SiO2 .
Under closed-system conditions, the variance of this system is
then given by equation 9.2:
f = (4 + 2) − 4 = 2.
Now let’s suppose that this rock is metamorphosed in an
open system with a mobile fluid, whose composition is controlled from outside the rock itself. (If this is hard to visualize,
consider a situation in which there is a vast reservoir of fluid
and only a thin layer of carbonate rock, so that the reaction in
the rock exerts virtually no influence on the composition of the
fluid that passes through.) As before, we again have four phases
and four system components. However, we also have the additional restriction r that the proportions of CO2 and H2O in the
fluid are fixed. Solving equation 9.3, we find that the variance is:
f = (4 + 2) − (4 + 1) = 1.
Therefore, metamorphism of this rock under closed-system
conditions is divariant, but the system becomes univariant if
the mobile fluid composition is controlled from outside the
system.
Now let’s return to our discussion of the utility of
the phase rule as a test for equilibrium. We still cannot
demonstrate equilibrium with certainty, but by examining
the variance the system, we can now cite an additional
observation that is consistent with equilibrium. In the
phase diagram for water (fig. 9.1), it is apparent that
most of the area of the diagram is occupied by divariant
fields, and those parts of the diagram with lower variance comprise only a tiny fraction of the total area. The
same holds true for complex systems such as rocks. Although univariant and invariant mineral assemblages
surely are formed under just the right combinations of
temperature and pressure, the probability of finding such
assemblages in the field is slight compared with the likelihood of sampling divariant assemblages that are stable
over a range of temperatures and pressures. For example,
locating univariant mineral reactions precisely in a metamorphic terrain may be difficult, because most outcrops
will contain either reactants or products, but not both.
Some isograds may mark the locations of univariant
reactions, but most are probably the first outcrops to be
found on the other side of the reaction. We can therefore
conclude that most rocks should have variances of two
or higher if they have equilibrium mineral assemblages.
Those assemblages that appear to be univariant or invariant probably represent disequilibrium in most cases,
unless it can be shown by field mapping that these rocks
lie between appropriate divariant assemblages.
CHANGING TEMPERATURE
AND PRESSURE
You are probably aware that water boils at different
temperatures at the seashore and in the mountains. The
boundary curve between liquid water and water vapor in
figure 9.1 defines the boiling point of water at any given
pressure, and its slope explains this phenomenon. Changing conditions of temperature and pressure obviously
influence more complex geological systems as well.
It is most convenient to describe modifications in the
state of a geochemical system in terms of changes in free
energy (∆G), because differentials of ∆G with respect to
T and P are easily evaluated. Recall that the concept of
free energy was developed in chapter 3 for fixed T and P
conditions. In the following sections, we see how to evaluate ∆G under different sets of temperature and pressure
conditions.
Temperature Changes and Heat Capacity
First we examine the effect of changing only temperature on a system at equilibrium. Differentiating the
expression d∆Ḡ = ∆V̄dP − ∆ S̄dT with respect to temperature at constant pressure gives:
(∂∆Ḡ/∂T )P = −∆S̄.
(9.4)
Because it is often not easy to evaluate the entropy
change ∆ S̄ directly, we must resort to using another
more readily measurable quantity, enthalpy (∆H̄). Since
Temperature and Pressure Changes: Thermodynamics Again
∆Ḡ = ∆H̄ − T∆S̄ (equation 3.13), equation 9.4 is equivalent to:
(
173
)
∆H̄0
− ———— = −9215,
2.303R
and
(∂∆Ḡ/∂T )P = (∆Ḡ − ∆H̄)/T.
Because we are holding pressure is constant in this example, we can rewrite this as:
∆H̄0 = 2.303(1.987 cal K−1mol−1)(9215K)
= 42.2 kcal mol−1.
d∆Ḡ = (∆Ḡ − ∆H̄)dT/T.
Dividing both sides of this expression by T and rearranging yields:
d∆Ḡ/T − (∆Ḡ)dT/T 2 = −(∆H̄ )dT/T 2.
The left side of this equation is equivalent to d(∆Ḡ/T ),
so that:
d(∆Ḡ/T) = −(∆H̄)dT/T 2.
If we consider the expression above at standard conditions, then ∆Ḡ0 and ∆H̄ become ∆Ḡ0 and ∆H̄0, respectively. For a system at equilibrium, ∆Ḡ0 = −RT ln Keq, so
that:
d ln Keq /dT = ∆H̄ 0/RT 2,
(9.5a)
−d ln Keq = [∆H̄ 0/R]d(1/T ).
(9.5b)
or
Equations 9.5a,b, known in either form as the van’t
Hoff equation, provide a useful way to determine enthalpy changes.
Worked Problem 9.3
Determine ∆H̄ for the breakdown of annite (the Fe end-member
component of biotite) by the reaction:
KFe3AlSi3O10(OH)2 → KAlSi3O8 + Fe3O4 + H2.
annite
sanidine magnetite
From equation 9.5b, we can see that ∆H̄ 0 can be obtained from
a plot of ln Keq versus 1/T. The slope of this line is equivalent to
−∆H̄0/R.
Hans Eugster and David Wones (1962) found experimentally that the equilibrium constant for this reaction could be
expressed by:
log Keq = − (9215/T ) + 10.99.
In other words, a plot of log Keq versus 1/T gave a straight line
of slope −9215K. Therefore,
The van’t Hoff equation works well for standard conditions, but how can we determine enthalpy differences
under nonstandard conditions? First let us define a new
variable, the heat capacity at constant pressure, C̄P :
C̄P = (dq/dT )P = (∂H̄/∂T)P.
If a material is heated at constant pressure, then its
molar enthalpy increases with temperature, according to
the relationship
(9.6)
d∆H̄ =∆C̄P dT,
where ∆H̄ is the increase in enthalpy through a temperature interval dT. Tabulations of thermodynamic data
for minerals normally give enthalpy at 298 K, and it is
necessary to integrate equation 9.6 to obtain enthalpy at
any other temperature. Thus, we have:
0
∆H̄T = ∆H̄298
∫
T
298
∆C̄P dT.
(9.7)
However, heat capacities are not generally independent of temperature, so the integrated form of the expression for ∆C̄P is somewhat complex. The experimental
data from which integrated C̄P values are determined are
commonly given in the form of a power series:
C̄P = a + bT + c/T 2,
(9.8)
where a, b, and c are experimentally determined constants.
Coincident with a change in enthalpy with temperature (equation 9.6), the system will also experience a
change in entropy, because
d∆H̄ = T d∆S̄
at constant pressure. Consequently, entropy can also be
expressed in terms of the heat capacity:
d∆S̄ = (∆C̄P dT )/T.
We can obtain the entropy at any temperature by integrating this equation:
0 =
∆S̄T = ∆S̄298
∫
T
298
(∆C̄P /T )dT.
(9.9)
174
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
We now have all the necessary tools to determine the
effect of a change in temperature on the free energy of a
system. ∆Ḡ at any temperature can be evaluated by
substituting enthalpy (equation 9.7) and entropy (equation 9.9) values at the appropriate temperature into the
familiar equation ∆Ḡ = ∆H̄ − T∆S̄:
0 +
∆ḠT = ∆H̄298
−T
(
∫
T
298
0
∆S̄ 298
+
∆C̄P dT
∫
T
298
)
(∆C̄P /T )dT .
(9.10)
From the relationship:
∆Ḡ = ∆Ḡ0 + RT ln Keq,
we obtain a value of ∆Ḡ = −28,833.65 cal mol−1.
Pressure Changes and Compressibility
The effect of changing pressure on a system at equilibrium can be determined by differentiating d∆Ḡ = ∆V̄dP −
∆S̄dT with respect to pressure at constant temperature:
(∂∆Ḡ/∂P)T = ∆V̄.
Worked Problem 9.4
As an example of the use of equation 9.10, determine the free
energy change for the following reaction at 800°C and atmospheric pressure:
MgSiO3 + CaCO3 + SiO2 → CaMgSi2O6 + CO2.
enstatite calcite quartz
diopside
(En)
(Cc)
(Qz)
(Di)
The thermodynamic data for these phases are given in table 9.1.
After confirming that the reaction is balanced, we calculate
∆ values for products (Di + CO2) minus reactants (En + Cc + Qz):
Consequently, the change in free energy of a reaction with
respect to pressure alone is equal to the change in the
molar volume of products and reactants. Molar volume
changes can be determined from simple physical measurements of product and reactant molar volumes. For reactions involving only solid phases under most geologic
conditions, the effects of temperature and pressure on
∆V̄ are small and can be ignored. In this case, integration
of equation 9.10 gives:
0 + ∆V̄
∆Ḡ = ∆ḠT,1
0 = 14,720 cal mol−1
∆H̄298
0 = 37.01 cal mol−1 deg−1
∆S̄ 298
(9.11)
∫ dP,
P
(9.12)
1
0 is the free energy for the reaction at 1 bar
where ∆ḠT,1
and temperature T.
10−3T
∆C̄P = 2.69 − 8.24 ×
− 2.62
× 105/T 2 cal deg−1 mol−1
Substitution of these values, along with T = 1073K, into equation 9.10 gives:
∆Ḡ1073 = 14,720 − 6.38 − 1073(37.01 − .0059)
= −23,304 cal mol−1.
An interesting variation on this problem is to specify the
value of PCO , say, at 0.1 bar. To solve this problem now, we
2
must first write an expression for the equilibrium constant Keq,
which in this case is:
Worked Problem 9.5
Consider the polymorphic transformation of aragonite to calcite. Use equation 9.12 to calculate the pressure at which these
two minerals coexist stably at 298 K.
At equilibrium, ∆Ḡ = 0, so:
∫
P
1
0 /∆V̄,
dP = −∆ḠT,1
or
0 /∆V̄.
P − 1 = −∆ḠT,1
Keq = PCO = 0.1 bar.
2
TABLE 9.1. Thermodynamic Data for Worked Problems 9.4 and 9.7
Mineral
En
Cc
Qz
Di
CO2
0 (cal mol−1)
H̄298
−370,100
−288,420
−217,650
−767,400
−94,050
From Helgeson et al. (1978).
0
S̄298,1
(cal mol−1 K−1)
16.22
22.15
9.88
34.20
51.06
V̄298,1 (cal mol−1 bar−1)
C̄p
0.752
0.883
0.542
1.579
584.73
24.55 + 4.74 × 10 T − 6.28 × 105 T −2
24.98 + 5.24 × 10 −3 T − 6.20 × 105 T −2
11.22 + 8.20 × 10 −3 T − 2.70 × 105 T −2
52.87 + 7.84 × 10 −3 T − 15.74 × 105 T −2
10.57 + 2.10 × 10 −3 T − 2.06 × 105 T −2
−3
Temperature and Pressure Changes: Thermodynamics Again
0
(kJ mol−1) =
Appropriate thermodynamic data are: ∆Ḡ298,1
−1127.793 for aragonite and −1128.842 for calcite; V̄ (J bar−1)
= 3.415 for aragonite and 3.693 for calcite. For the reaction aragonite → calcite, ∆Ḡ0 = −1049 J mol−1 and ∆V̄ = 0.278 J bar−1.
By substituting these values into the equation above, we calculate that P = 3773 bars.
Aragonite has the lower molar volume and is thus the phase
that is stable on the high-pressure side of this reaction. Therefore, at 298 K, calcite is the stable polymorph to ∼3.7 kbar pressure. From this computation, we can readily see why aragonite
precipitated by shelled organisms at low pressure is metastable.
Equation 9.12 is an extremely useful relationship for
determining the effect of pressure on many equilibria of
geologic interest. For any reactions involving fluids (such
as many metamorphic reactions), however, or for solid
phase reactions at very high pressures and temperatures
(such as those under mantle conditions), we cannot assume that ∆V̄ is independent of temperature and pressure.
In these cases, we must take into account the expansion
or contraction of a fluid or crystal lattice as temperature
changes and the compression or relaxation due to changing pressure. For gases, it is permissible to use an equation
of state, such as the ideal gas law V̄ = RT/P, to describe
the effects of changing temperature or pressure:
(∂V̄/∂P)T = −RT/P,
and
(∂V̄/∂T)P = R/P.
More complex equations of state for fluids have already
been introduced in chapter 3.
For solid phases, we can correct ∆V̄for changing temperature using the thermal expansion, αP, defined as:
αP = 1/V̄(∆V̄/∆T )P,
where V̄ is the molar volume of the phase at the temperature at which αP is measured. The temperature dependence of αP is large near absolute zero, but becomes small
at the temperatures at which most geologic processes
operate. If we ignore this small variation, this expression
can be rearranged to give:
0
∆V̄ = V̄298,1
(1 + αP ∆T),
where V̄ 0 is the molar volume at 298 K and 1 bar, and
∆T = T − 298.
The effect of changing pressure on ∆V̄ is corrected by
using the compressibility βT , defined by:
175
β T = −1/V̄(∆V̄/∆P)T ,
where V̄ is the molar volume of the phase at the pressure
at which βT is measured and ∆P = P − 1. The effect of
pressure on βT can as a first approximation be neglected
at pressures < 2 or 3 GPa, although it is very important
in interpreting mantle seismic data. If we regard β T as
constant in the pressure range where most observable
geologic processes operate, this expression becomes
0
∆V̄ = V̄298,1
(1 − β T ∆P),
where V̄ 0 is the molar volume at l bar and 298°C. By
combining the definitions of α P and β T in a total differential expression for V̄, we obtain
dV̄ = (∂V̄/∂T )P dT + (∂V̄/∂P)T dP
= α PV̄dT − β TV̄dP
or
∫
V̄T,P
dV̄/V̄ =
V̄298,1
∫
T
298
αP dT −
∫β
P
1
T
dP.
This produces the result:
0
V̄T,P = V̄298,1
exp(αP ∆T − β T ∆P).
Because the term in parentheses is very small, we can use
the approximation exp x ≈ 1 + x to yield:
0
V̄T,P = V̄298,1
(1 + αP ∆T − β T ∆P).
(9.13)
This last expression allows us to gauge the effect of
changing thermal expansion and compressibility on the
volume change for a reaction occurring under hightemperature and high-pressure conditions.
Worked Problem 9.6
What is the change in molar volume of forsterite in going from
25 to 800°C and from 1 bar to 5 kbar? To answer this question,
we need the following data for forsterite:
0
= 4.379 cm3
V̄298,1
α P = 41 × 10−6 K−1
β T = 0.8 × 10−6 bar−1
Inserting these values into equation 9.13 gives the following
result:
V̄800,5000 = 4.379[1 + (41 × 10−6)(1073 − 298)
− (0.8 × 10−6)(5000 − 1)] = 4.501 cm3.
176
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
A GRAPHICAL LOOK AT
CHANGING CONDITIONS:
THE CLAPEYRON EQUATION
The change in molar volume is thus:
∆V̄ = 4.501 − 4.379 = 0.122
cm3
mol−1.
Temperature and Pressure
Changes Combined
A general equation for the effect of both temperature
and pressure on free energy can be obtained by simply
combining the equations we have already derived for the
effects of temperature alone (equation 9.10) or pressure
alone (equation 9.12), as follows:
0 +
∆ḠP,T = ∆H̄ 298
+
∫
T
298
∫
T
298
(
0
∆CP dT − T ∆S̄ 298
)
(∆CP /T )dT + ∆V̄
∫
P
1
dP.
(9.14)
In using this equation, remember that care must be
exercised in obtaining the integrated form of ∆CP , normally done by employing the relationship in equation 9.8.
As written, equation 9.14 assumes that thermal expansion and compressibility do not need to be taken into
account. However, evaluating ∆V̄ in the last term for reactions at very high temperatures and pressures or when
a fluid phase is present may require equation 9.13.
Worked Problem 9.7
Consider again the reaction enstatite + calcite + quartz → diopside + CO2. In worked problem 9.4, we calculated the free
energy of this reaction at 800°C and 1 bar. Now calculate the
value of ∆Ḡ at 800°C and 2000 bars pressure, assuming that
PCO = 0.1 bar.
2
To do this, we need to make some assumption about
(∂V̄/∂P)T . It is clear that ∆V̄ is not constant, because CO2 is
quite compressible. Let’s assume instead that all of the compressibility is due to CO2, and pretend that it is an ideal gas, so
that ∆V̄ = RT/∆P. Which is stable under these conditions, reactants or products?
0
From the data in table 9.1, we can calculate ∆H̄298,1
, ∆CP ,
0 . (In fact, we have already done this in worked proband ∆S̄298
lem 9.4). The change in molar volume ∆V̄ of CO2 between 1 bar
and 2000 bar pressure is (1.987 cal K−1 mol−1)(1073K)/2000
bar = 1.066 cal bar−1. Substituting these values into equation
9.14 gives:
A phase diagram of temperature versus pressure summarizes a large quantity of data on the stabilities of various phases and combinations of phases in a system. It
is important to recognize, however, that such a diagram
gives information about the thermodynamic properties
of these phases as well. Let’s see how.
The phase diagram for the Al2SiO5 system is illustrated in figure 9.2. Like the phase diagram for water,
this is a one-component system containing three possible
phases: the polymorphs kyanite, andalusite, and sillimanite. Ignore, for the moment, the dashed horizontal lines.
At the beginning of this chapter, we noted that equilibrium is defined as a condition of minimum energy for a
system. Therefore, the divariant fields in figure 9.2 represent regions of P-T space where the free energy of the
system is minimized by the occurrence of only one phase,
and the univariant curves and the invariant point define
special combinations of pressure and temperature at
which two or more coexisting phases provide the lowest
free energy.
To illustrate this relationship between free energy
and phase diagrams graphically, we construct qualitative
Ḡ-T diagrams at constant pressure for the aluminosilicate system. The traces of such isobaric sections are represented by the horizontal dashed lines in figure 9.2. At
pressure P1 and any temperature below T1, kyanite is the
stable phase, so it must have a lower free energy than the
∆Ḡ800,200 = −30,027.05 cal mol−1.
Because ∆Ḡ is negative, the reaction proceeds as written and
diopside + CO2 is stable under these conditions.
FIG. 9.2. Schematic phase diagram for the system Al2SiO5.
Dashed lines represent the traces of isobaric sections used to
construct Ḡ -T diagrams.
Temperature and Pressure Changes: Thermodynamics Again
177
which is equivalent to:
dP/dT = ∆S̄/∆V̄ = ∆H̄/ T∆V̄.
(9.15)
This expression, known as the Clapeyron equation, is
extremely useful in constructing and interpreting phase
diagrams. Notice that the left side of the Clapeyron equation is the slope of any line in a P-T phase diagram. Thus,
this relationship enables us to calculate phase boundaries
from thermodynamic data, provided that one point on
the boundary is known. Alternatively, ratios of thermodynamic quantities can be obtained from experimentally
determined reactions.
Worked Problem 9.8
FIG. 9.3. Ḡ -T diagrams for the system Al 2SiO5 at constant
pressure: (a) corresponds to pressure P1 in figure 9.2, and
(b) corresponds to P2. At any temperature, the stable phase or
combination of phases is that with the lowest free energy.
other Al 2SiO5 polymorphs. This is shown in figure 9.3a.
Because (∂Ḡ/∂T )P = −∆ S̄ and the entropy of any phase
must be greater than zero, free energy decreases in all
phases as temperature increases. At temperature T1, the
Ḡ-T lines for kyanite and sillimanite intersect, and the coexistence of both phases provides the lowest free energy
configuration. At temperatures above T1, sillimanite has
the lowest free energy and is the stable phase. Andalusite
never appears at this pressure because its free energy is
higher than those of the other polymorphs. The Ḡ-T diagram at pressure P2, shown in figure 9.3b, is similar to
the P1 case except that the Ḡ-T line for andalusite is depressed so that it intersects those of kyanite and sillimanite at temperature T2, the aluminosilicate triple point. At
temperatures other than T2, however, the other polymorphs still offer lower free energy configurations.
There are other, even more useful relationships between phase diagrams and thermodynamic quantities. In
chapter 3, we derived a Maxwell relationship from dH̄
(equation 3.8):
(∂P/∂T)V̄ = (∆S̄/∆V̄)T ,
To illustrate how the Clapeyron equation is used, calculate the
approximate slope of the kyanite-andalusite boundary curve in
figure 9.2. The data we need are: V̄ (J bar−1) = 4.409 for kyanite,
5.153 for andalusite, and 4.990 for sillimanite; S̄800 (J mol −1 K−1)
= 242.42 for kyanite, 251.31 for andalusite, and 252.94 for
sillimanite. Inspection of these data indicates the relative positions of the phase fields for the three polymorphs. Kyanite,
which has the smallest molar volume, should be stable at the
highest pressures, and sillimanite should be stable at the highest temperatures because of its high molar entropy. Substituting
values of S̄ and V̄ for kyanite and andalusite into equation 9.15,
we obtain:
dP/dT = (251.37 − 242.42 J mol−1 K−1)/
(5.153 − 4.409 J bar−1) = 12.03 bar K−1
for the slope of the kyanite-andalusite boundary curve.
To complete the construction of this line, we must fix its
position by determining the location of some point along it. The
triple point has been located by experiment, although not without controversy (see the accompanying box), and its position
can be used to locate this boundary in P-T space. There is a
considerable amount of potential error in most calculated P-T
slopes because of uncertainties in thermodynamic data. For this
reason, it is preferable to determine phase diagrams experimentally. However, calculations such as these can be valuable in estimating how systems behave under conditions other than those
under which experiments have been carried out. Calculations
also enable the geochemist to limit experimental runs to the
probable location of a line or point, thus saving time and money.
REACTIONS INVOLVING FLUIDS
Many geochemical reactions involve production of a fluid
phase; that is, they are dehydration or decarbonation
178
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
PHASE EQUILIBRIA IN THE ALUMINOSILICATE SYSTEM
The tortuous history of attempts to determine phase
equilibria in the aluminosilicate system illustrates the
difficulties that can be encountered in this kind of
work. E-An Zen summarized the state of affairs in
1969, after decades of experiments, in the diagram
shown as figure 9.4. Because of the constant shifting
of the position of the aluminosilicate phase boundaries with each new experimental determination, this
situation came to be known popularly as the “flying
triple point.” Zen discussed a number of causes for
this experimental difficulty. One possibility was experimental error, such as incorrect pressure calibration. This might be due to failure to take into account
the strength of the material, friction, and other factors
in converting the gauge pressure to actual pressure in
FIG. 9.4. Experimental determinations of phase equilibria in
the system Al2SiO5. These data, summarized by Zen (1969),
illustrate the difficulties in these measurements. The phase
relationships determined by Holdaway and Mukhopadhyay
(1993), shown in heavy lines, provide the most commonly
accepted value for the aluminosilicate triple point.
reactions. At low pressures and high temperatures, H2O
and CO2 are not very dense, so reactions that liberate
these fluids have large, positive ∆V̄. Consequently, the
slopes of the reaction lines on P-T diagrams, dP/dT =
∆S̄/∆V̄, are small. If the same reactions occur at higher
pressures, fluids are more compressed, so that ∆V̄ is
smaller and the slopes of the reaction lines are steeper.
For this reason, dehydration or decarbonation reactions
on P-T diagrams typically are curved upward. Because
fluids are more compressible than solids, high pressures
the sample. He also observed that most of the experiments to that time were synthesis experiments; that
is, they were based only on the first appearance of a
phase and either failed to demonstrate reversibility
or represented some synthesis reaction that did not
occur in metamorphic rocks. For example, in one set
of experiments, sillimanite was synthesized from gels
or kaolinite. Another potential source of error was
that some of the polymorphs were synthesized in the
presence of water. If they contained small quantities
of H2O, the polymorphs would not be true onecomponent phases.
The problem has been evaluated more recently by
Daniel Kerrick (1990), who summarized data obtained
since Zen’s review and critically evaluated potential
problems with each experimental determination. Kerrick also considered constraints on the position of
the triple point from consistency with thermodynamic
data and field calibrations from metamorphic mineral
assemblages.
Many metamorphic petrologists now accept the
experimental determination of Michael Holdaway
and Biswajit Mukhopadhyay (1993) as the best available phase equilibria for this system. The phase diagram of Holdaway and Mukhopadhyay is shown in
heavy lines in figure 9.4. Its triple point is located at
504 ± 20°C and 3.75 ± 0.25 kbar. Discrepancies
between previous data and Holdaway’s experiments
may be explained by the difficulty in measuring small
amounts of reaction, the occurrence of fibrolite (a
rapidly grown, fine-grained sillimanite often intergrown with quartz), and the presence of Fe2O3 in
sillimanite.
may ultimately so compress the fluid that ∆V̄ for these
reactions becomes negative. However, ∆S̄ is still positive,
so the reaction curves bend back on themselves with negative slopes. This phenomenon is nicely illustrated by the
reaction:
NaAlSi2O6 ⋅H 2O + SiO2 → NaAlSi3O8 + H 2O.
analcite
quartz
albite
For most fluid-release reactions, the slope reversal occurs at very high pressures, but the analcite dehydration
Temperature and Pressure Changes: Thermodynamics Again
FIG. 9.5. The dehydration reaction for analcite changes slope
(dP/dT ) and finally bends back on itself with increasing pressure,
reflecting the large change in volume of the fluid phase.
179
Interactions between molecules or ions, however, may
cause the chemical potentials of individual species to
increase or decrease on a certain site, resulting in nonideal mixing. In this case, ∆H̄mixing is not equal to zero,
and the activities will deviate from ideal behavior, as
illustrated in figure 9.6. However, as the nonideal component becomes increasingly diluted (as Xi becomes much
less than one), the component becomes so dispersed that
eventually it is surrounded by a uniform environment of
other ions or molecules. Therefore, at high dilution, its
activity becomes directly proportional (but not equal) to
its concentration, as shown by the straight dashed line in
figure 9.6. In other words,
ai = hi Xi ,
reaction, common in low-pressure metamorphic rocks,
exhibits a bend-back at very modest pressures (fig. 9.5).
The reaction has a small slope at very low pressure, which
rapidly becomes negative as pressure increases.
RAOULT’S AND HENRY’S LAWS:
MIXING OF SEVERAL COMPONENTS
So far in this chapter, we have focused mostly on the
effects of changing temperature and pressure on systems
with one component. In systems with several components, it is necessary to understand how the various components interact. For example, if we define our system as
a grain of olivine, we must determine the behavior of the
components Mg2SiO4 and Fe2SiO4 when mixed together.
The mixing characteristics of these two components may
be very different from the mixing behavior of some trace
olivine component such as Ni2SiO4. The interaction of
various components in combinations of coexisting minerals, magmatic liquid, or hydrous fluid is in some cases
nonideal, resulting in their selective concentration in one
phase or another. We now examine the thermodynamic
basis for various kinds of mixing behavior.
In an ideal solution, mixing of components with similar volumes and molecular forces occurs without any
change in the energy states or total volume of the system.
Under these conditions, mixing is neither endothermic
nor exothermic, that is, ∆H̄mixing = 0. The activities of
components (ai ) mixing on ideal sites are thus equal to
their concentrations:
ai = Xi ,
which is a statement of Raoult’s Law.
where hi is a proportionality constant. This is a statement of Henry’s Law.
These two laws can be used to explain the mixing behavior of the components of many minerals. Raoult’s Law
behavior is a common assumption for many major components of solid solution series, as in the case of the
Mg2SiO4 (forsterite) and Fe2SiO4 (fayalite) components
of olivine. Henry’s Law is commonly used to describe
the behavior of many trace components, such as nickel
in olivine.
STANDARD STATES AND
ACTIVITY COEFFICIENTS
Chemical potentials are functions of temperature, pressure, and composition. It is common practice to define
the chemical potential relative to a chemical potential at
FIG. 9.6. Relations between activity and mole fraction for solutions obeying Raoult’s Law and Henry’s Law. Raoult’s Law may be
a good approximation for nonideal solutions in which X i is large,
and Henry’s Law is applicable to highly diluted components.
180
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
some reference value of temperature and pressure (called
the standard state), adding a term that corrects this value
for deviations in temperature, pressure, and composition:
µi = µ0i + RT ln ai ,
(9.16)
where µi is the chemical potential of component i at T
and P in the phase of interest, µ0i is its value at some
standard set of conditions, and ai is the activity of the
component. Note that when the component of interest is
in its standard state, its activity ai must equal one, so that
its logarithm is zero and the second term drops out.
The standard state can be chosen arbitrarily, because
it has no effect on the final result as long as we are consistent in its use. It might seem most logical to choose
the standard state as a pure end-member component at
298 K and 1 bar. However, because of the availability of
tabulated thermodynamic data, it is often more convenient to select the standard state as a pure component at
the temperature and/or pressure of interest.
The relationships between chemical potential and the
activity term in equation 9.16 can be most readily seen
in graphical form. Figure 9.7a shows µi versus ln Xi . The
line illustrated is not straight (because it is for a real substance that does not mix ideally at all concentrations):
it has two linear segments (shown extended with dashed
lines). In the straight line segment extending from pure
component i (Xi = 1, so that ln Xi = 0) to slight dilution
with another component, component i obeys Raoult’s
Law (ai = Xi). The slope of this segment is RT and its intercept on the µi axis is µ0i . In the region where Raoult’s
Law holds, chemical potential can be expressed as:
µi = µ0i + RT ln Xi.
(9.17)
The Henry’s Law region extends from infinite dilution
of component i to some modest value and, because ai =
hi Xi , it is also represented by a straight line in figure 9.7a.
Its slope is also RT, but its intercept on the µi axis is not
µ0i . Instead, this intercept is now µ0i plus the added term
RT ln hi , which is a function of temperature, pressure, and
the composition of the component with which i is diluted,
but is not a function of Xi . The expression for chemical
potential in the Henry’s Law region, therefore, is:
µi = µ0i + RT ln Xi + RT ln hi ,
where hi is the Henry’s Law constant for component i in
this phase. Combining the two logarithmic terms in this
equation gives:
FIG. 9.7. (a) Plot of µ i versus ln X i showing the dependence of
chemical potential on composition. In the Raoult’s Law region,
µ i = µ0i ; in the Henry’s Law region, µ i = µ0i + RT ln hi . The slope of
the Raoult’s and Henry’s Law regions is RT. The normal standard
state for component i in a pure substance at P and T is µ0i .
(b) Plot of µ i versus ln X i showing how the intermediate region
changes with activity coefficient. With ideal mixing (γ i = 1),
the Raoult’s and Henry’s Law regions lie along the same trend.
As γ i becomes larger, the inflection appears and the intermediate
region becomes larger.
µi = µ0i + RT ln hi Xi .
(9.18)
The intermediate region between Henry’s and Raoult’s
Law behavior is characterized by a curved segment that
connects the two straight line segments (fig. 9.7a). This
region is thus expressed by an equation that is intermediate between equations 9.17 and 9.18. To formulate
such an equation, we can describe the degree of nonideality by using γi, the activity coefficient, introduced in
chapter 4. The activity coefficient is a function of composition, such that in the Raoult’s Law region,
γ i → 1, so that ai → Xi ,
and in the Henry’s Law region,
γ i → hi , so that ai → hi Xi .
Temperature and Pressure Changes: Thermodynamics Again
Thus we can formulate a general equation for chemical potential of a pure component i in any phase at any
temperature and pressure:
µi = µ0i + RT ln γ i Xi .
(9.19)
We can see that as component i becomes more similar
to the other component with which it is mixing, hi → 1.
Moreover, greater similarities in these components will
increase the size of the Raoult’s and Henry’s Law regions
in figure 9.7a, to the point where the intermediate region
disappears and the Raoult’s and Henry’s Law regions connect. In such a case, we would then have one straight
mixing line in this diagram (equivalent to the line that
extrapolates to µ0i ), and the mixing would be ideal. The
effects of changing activity coefficients on the various
mixing regions are illustrated in figure 9.7b.
Besides the standard state we have just considered,
there are others that could be used, and in fact, it is advantageous to do so under certain conditions. Obviously,
we must use another standard state if it is not possible to
make the phase of interest out of pure component i. For
181
example, we cannot synthesize a garnet of lanthanum
silicate, so µ0La in garnet based on the pure phase cannot
be determined directly. In this case, the standard state
could be a hypothetical garnet with properties obtained
by extrapolating the Henry’s Law region to the pressure
and temperature of interest. Such a situation is illustrated in figure 9.8a. This is the standard state we assumed without any explanation in chapter 4. In fluids, it
is not possible to measure the activity of pure Na+ ions,
for example, so we must extrapolate from the dilute salt
solution.
We might also envision a case where component i
cannot be studied directly in the phase of interest, but its
chemical potential can be analyzed in another phase.
This could occur, for example, if we were interested in
determining the chemical potential of Al2O3 in a magma
but found it easier to study glass quenched from the
magma. We could use the activity of Al2O3 in a glass of
pure Al2O3 composition to define the standard state.
Such a standard state is illustrated in figure 9.8b. The
expressions for chemical potentials using other standard
states can be derived from equation 9.19.
SOLUTION MODELS: ACTIVITIES OF
COMPLEX MIXTURES
FIG. 9.8. Diagrams of µ i versus ln Xi illustrating (a) the standard
state for component i in a dilute solution extrapolated from the
Henry’s Law region, and (b) the standard state for component i
taken as the activity of i in another phase at P and T.
Activity coefficients are functions of temperature, pressure, and composition. They can be greater or less than
one. The values for activity coefficients cannot be determined from thermodynamic properties but must be obtained experimentally from measurements of ai and Xi
or estimated empirically from mineral data. Many different expressions for activity coefficients as a function of
composition—often called solution models—have been
formulated. The only restrictions on them are that they
must obey the constraints given above for the Raoult’s
and Henry’s Law regions; that is, they must explain mixing at the extreme ends of the compositional spectrum,
where Xi → 1 and Xi → 0.
One solution model that is often used in solving geochemical problems assumes ideal mixing. This is a trivial
model, because the expression for γi = 1. From figure 9.7,
we can see that the assumption will be correct in the
Raoult’s Law region, but will become increasingly less
accurate in moving toward the intermediate and Henry’s
Law regions. Thus, this solution model works best in cases
for which Xi is large.
182
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Another solution model that is commonly used is that
for a symmetric regular solution. For this model,
RT ln γ i = WXj2 = W(1 − Xi )2,
(9.20)
where W is an interaction parameter that is a function of
temperature and pressure but not composition, and Xj
is the mole fraction of the other component with which
component i is mixing. Substitution of this expression in
equation 9.20 gives the chemical potential for component i using a symmetric regular solution model:
µi = µ0i + RT ln Xi + WXj2.
(9.21)
Note that this equation obeys the necessary restrictions
in the Raoult’s and Henry’s Law regions. In the case of
Raoult’s Law, as Xi → 1, W Xj2 → 0 and γ i → 1. For
Henry’s Law, as Xi → 0, W Xj2 → W and γ i → hi .
In the regular solution model, the inflection point in
the intermediate region on the µi-ln Xi plot is flanked by
symmetric limbs, as illustrated in figure 9.9. However, in
real systems, mixing behavior need not be this regular.
The asymmetric regular solution model is a somewhat
more complex formulation that is also very useful in some
situations. This model involves two interaction parameters, Wi and Wj , rather than one as in the regular solution
model. For the asymmetric regular solution model,
RT ln γ i = Xj2(Wi + 2Xi (Wj − Wi )).
(9.22)
This model can be used in situations in which mixing is
not symmetric, as illustrated in figure 9.9. Note that when
Wi = Wj , the asymmetric regular solution model reduces
to the symmetric regular solution model. This mixing
model also obeys the necessary restriction in the Raoult’s
and Henry’s Law regions. However, in the Henry’s Law
region, RT ln γ i → Wi and RT ln γ j → Wj .
Worked Problem 9.9
Calculate the relationship between activity and mole fraction of
Fe3Al2Si3O12 (almandine) in a garnet solid solution. Although
some geochemists have argued that almandine mixes nearly
ideally with Mg3Al2Si3O12 (pyrope), others have advocated a
symmetric regular solution model with a temperature-dependent
interaction parameter of the form:
WFe-Mg = 3480 − 1.2T,
where T is the temperature in °C and W has units of cal g−1
atom−1. Strictly speaking, regular solution models do not allow
for variation of W with temperature, but let’s ignore this minor
problem and use this solution model to illustrate how interaction parameters work.
The garnet we consider has the following composition:
XFe = 0.710 and XMg = 0.088, along with minor amounts of
Ca and Mn. Let’s calculate the activity of Fe in garnet at 827 K
or 554°C (the reason this temperature was chosen will become
clear shortly). The value of the interaction parameter is therefore:
WFe-Mg = 3480 − 1.2(554) = 2815 cal g−1 atom−1.
Applying this value to equation 9.20, we have:
RT ln γFe = 2815(1 − X Fe )2,
or
ln γFe = (2815)(1 − 0.710)2/(1.987)(827) = 0.144
and
γFe = 1.155.
Fe mixes on three crystallographic sites in garnet, so aFe =
3 . Therefore, the activity of Fe in this garnet is:
γFeX Fe
aFe = 1.155(0.710)3 = 0.413.
For comparison, for an ideal solution (γFe = 1), the activity of Fe
in this garnet would be:
aFe = 1(0.710)3 = 0.446.
THERMOBAROMETRY: APPLYING WHAT
WE HAVE LEARNED
FIG. 9.9. The difference in mixing behavior for the symmetric and
asymmetric regular solution models is illustrated by the shapes of
the intermediate regions. The symmetric solution model produces
symmetrical limbs flanking the inflection point, and the asymmetric model does not.
The formulation and application of geothermometers
and geobarometers involves many of the concepts just
introduced. We use one rather simple example of each of
these techniques to illustrate how useful the concepts can
Temperature and Pressure Changes: Thermodynamics Again
be in solving geologic problems. For a particular reaction
to be used as a geothermometer, it must be a strong function of temperature but be nearly independent of pressure. Conversely, a geobarometer should be sensitive to
pressure but not to temperature. We have already seen
that at equilibrium:
∆G 0 = − RT ln K eq = ∆H̄ − T∆S̄ + P∆V̄.
This equation assumes that ∆V̄ is independent of temperature and pressure; otherwise we would have to include expressions for thermal expansion and isothermal
compressibility. The form of this equation indicates that
if Keq is constant, there is only one equilibrium pressure
at any given temperature, and fixing Keq determines the
position of the equilibrium reaction curve in P-T space.
The factors that control the slopes of equilibrium curves
in P-T space can be evaluated by differentiating the above
equation with respect to temperature at constant pressure
and with respect to pressure at constant temperature:
(∂ ln Keq /∂T )P = ∆H̄/RT 2
(9.23)
and
(∂ ln Keq /∂P)T = − ∆V̄/RT.
(9.24)
If the right side of equation 9.23 is larger than that of
9.24, then the equilibrium depends more on temperature
than pressure and would be a suitable geothermometer.
Thus, we can see that geothermometer reactions should
have large values of ∆H̄ and, conversely, that geobarometer reactions should be characterized by large ∆V̄values.
It will also be necessary to have accurate standard-state
thermodynamic data for the reaction available, as well
as formulations for activity coefficients.
Worked Problem 9.10
183
partitioning between these phases is for the most part only a
function of T (and, to a lesser extent, P) so long as we assume
ideal mixing in both garnet and biotite. No formulation for
their activity coefficients is necessary in this case. (Recall that
some workers have suggested that a symmetric regular solution
model provides a better expression for garnet mixing, as we
explored in worked problem 9.9.)
There are quite a few expressions for the garnet-biotite geothermometer; we will consider the experimental calibration of
John Ferry and Frank Spear (1978). They measured the progress of this exchange reaction at a series of temperatures, and
obtained the following expression for T (in kelvins):
ln KD = −2109/T + 0.782,
(9.25)
where KD equals (XMg /XFe)garnet /(XMg/XFe)biotite. (For equilibria in which the same phase or phases appear on both sides of
the equation, Keq is commonly called KD, the distribution coefficient.) The coefficients were determined by a linear leastsquares fit of the experimental values of ln KD versus 1/T,
determined at a pressure of ∼2 kbar. (To see how this is done,
refer to appendix A.) Using the Clapeyron equation, Ferry and
Spear were also able to calculate dP/dT and, therefore, to estimate the dependence of the reaction on pressure. Their expression for this Fe-Mg exchange reaction at any pressure and
temperature is:
ln KD = −(2089 + 0.0096P)/T + 0.782
(9.26)
where P is in bars and T is in kelvins. From this equation, we
can see that the effect of pressure on this exchange equilibrium
is small, and it can be ignored if no pressure estimate is available. To verify that the error generated by ignoring pressure is
small, let’s solve for temperature in two ways.
First, we use this geothermometer to calculate the temperature of equilibration for a metamorphic rock containing garnet
and biotite with these compositions:
garnet: XFe = 0.710, XMg = 0.088
biotite: XFe = 0.457, XMg = 0.323
The value of KD is then:
KD = (0.088/0.710)/(0.323/0.457) = 0.1743
The partitioning of Fe and Mg between coexisting biotite and
garnet serves as a very useful geothermometer. The exchange
reaction is:
→
Fe3Al2Si3O12 + KMg3AlSi3O10(OH)2 ←
almandine
phlogopite
Mg3Al2Si3O12 + KFe3AlSi3O10(OH)2.
pyrope
annite
The enthalpy change for this reaction is large but ∆V̄ is quite
small, as appropriate for a geothermometer. Available data suggest that Fe and Mg mix almost ideally in biotite and garnet, so
we assume that they obey Raoult’s Law and ai = Xi. The Fe-Mg
Since we do not know the pressure, we can use equation 9.25
to obtain T:
ln 0.1743 = −2089/T + 0.782;
T = 827 K = 554°C.
Let’s now say that we have an independent measure of
pressure, determined to be 5 kbar. What effect does including
the pressure term have on the calculated temperature? Applying
equation 9.26:
−1.743 = −(2089 + 0.0096[5000])/T + 0.782.
184
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Solving this expression for T (our second temperature estimate),
we obtain:
T = 846 K = 573°C,
a difference of only 11°C, or about 2°C per kbar of pressure.
Worked Problem 9.11
A useful geobarometer is based on the reaction:
→ 3FeTiO + Al SiO + 2 SiO .
Fe3Al2Si3O12 + 3TiO2 ←
3
2
5
2
almandine
rutile
ilmenite
Al
quartz
silicate
This reaction is almost independent of temperature, but has a
large ∆V̄ that varies slightly depending on which aluminosilicate
polymorph (kyanite, andalusite, or sillimanite) is produced by
the reaction. The equilibrium constant for this reaction is:
Keq = [(ailm)3 (aAl SiO )(aqz)2 ]/[(aFe gar)(aru)3].
2
5
(9.27)
The great advantage of this particular geobarometer is that
so many of the activity terms in equation 9.27 are equal to unity
(rutile, quartz, and the aluminosilicate polymorphs form essentially pure phases). Ilmenite forms a solid solution with hematite, but it rarely contains more than 15 mol % Fe2O3 and can
be treated as an ideal solution or, in some cases, a pure phase.
Fe in garnet can be treated as an ideal solution or as a symmetrical regular solution, as we saw in worked problem 9.8. Consequently, the activity terms for garnet and ilmenite depend on
their compositions.
The end-member reaction (involving pure almandine garnet)
was calibrated experimentally by Steven Bohlen and coworkers
(1983). They then calculated the relationship between P, T, and
Keq, which of course takes into account other compositions of
garnet and ilmenite. To do this, Bohlen and his collaborators
used available data on molar volumes, thermal expansion, and
compressibility, and the relationships we have already derived
in this chapter. They presented their results in graphical form,
contouring log Keq for the reaction on a P-T diagram, as shown
in figure 9.10. You can see that the slopes of the log Keq contours in this diagram are very shallow, demonstrating that the
reaction is very sensitive to pressure but almost independent
of temperature. Also note that the slopes change slightly when
contours move from the kyanite field into the sillimanite field or
when they cross the α-quartz–β-quartz transition, reflecting
changes in ∆V̄ of the reaction.
Let’s use this diagram to determine the equilibration pressure of a metamorphic rock containing rutile, kyanite, ilmenite,
and quartz, as well as garnet and biotite with the compositions
as in worked problem 9.10. Note that we have already used this
same garnet-biotite pair to calculate the temperature (554°C).
To keep the problem simple, when calculating Keq, we stipulate
that ilmenite is a pure phase. Bohlen and coworkers recom-
FIG. 9.10. P-T diagram showing the experimentally determined
→ ilmenite + Al SiO + quartz, and
reaction almandine + rutile ←
2
5
contours of log K eq for the reaction with impure garnet and ilmenite compositions, after Bohlen et al. (1983). The slopes of the
contours show that the reaction is strongly dependent on pressure
but not on temperature (as appropriate for a geobarometer).
Boundary curves for the polymorphic transformation of andalusite,
kyanite, and sillimanite, and of α- and β-quartz are also shown.
mended a symmetric regular mixing model for Fe in garnet, and
we have conveniently already used this model (and a temperature
of 554°C) to determine aFe in worked problem 9.9. Substituting
this value into equation 9.27 gives:
Keq = 1/0.413 = 2.421.
Therefore,
log Keq = 0.384.
Applying this value and a temperature of 554°C to figure 9.10,
we obtain a pressure of ∼6 kbar.
Before leaving this problem, let’s see how much difference it
would make if instead we had used an ideal solution model for
Fe in garnet. From equation 9.27, the value of Keq becomes:
Keq = 1/(0.710)3 = 2.793,
and
log Keq = 0.446.
The log Keq value corresponds to a pressure of about 5 kbar, a
difference of 1 kbar.
SUMMARY
We have reaffirmed that equilibrium is a condition of minimum free energy for a system. The phase rule provides
Temperature and Pressure Changes: Thermodynamics Again
a useful test for equilibrium. Most geologic systems at
equilibrium have variances of two or higher (reflecting the
fact that most rocks are stable over a range of temperatures and pressures). The effects of changing temperature
and pressure on the free energy of a system can be determined, and the direction of a particular reaction under
different conditions can be predicted from appropriate
thermodynamic data. The Clapeyron equation illustrates
the usefulness of thermodynamic data in assessing the
slopes of reactions on P-T diagrams.
Systems containing more than one component require analysis of the way such components interact in
the phases that constitute the system. Components that
mix ideally obey Raoult’s Law; that is, their activities
equal their concentrations. Components at high dilution
obey Henry’s Law; that is, their activities are directly
proportional to (but not equal to) their concentrations.
The chemical potentials of components are generally
expressed in the form µi = µ0i + RT ln γi Xi (the chemical
potential at some standard state plus an activity term
that corrects this value for deviations in conditions and
composition). Activity coefficients are equal to one for
Raoult’s Law behavior and hi for Henry’s Law behavior.
Various solution models have been formulated for activity coefficients for components exhibiting intermediate
(nonideal) behavior. Solution models find practical application in the formulation of geothermometers and
geobarometers.
With this thermodynamic background, we are now
ready to consider phase diagrams for multicomponent
systems. These are introduced in chapter 10.
suggested readings
There are a number of excellent textbooks that provide an introduction to thermodynamics, as well as various geological
applications. These generally contain significantly more detail
than has been presented here, and are highly recommended for
the serious student.
Denbigh, K. 1971. The Principles of Chemical Equilibrium.
London: Cambridge University Press. (A classic thermodynamic text containing rigorous derivations but no examples of geologic interest; a superb reference, but not for the
fainthearted.)
Ernst, W. G. l976. Petrologic Phase Equilibria. San Francisco:
Freeman. (A very readable introductory text; chapter 3 illustrates the computational approach to phase diagrams, and
chapter 4 contains descriptions of Ḡ-T diagrams.)
185
Essene, E. J. 1982. Geologic thermometry and barometry. In J. M.
Ferry, ed. Characterization of Metamorphism through
Mineral Equilibria. Reviews of Mineralogy 10. Washington, D.C.: Mineralogical Society of America, pp. 153–206.
(An excellent review of the calibration, assumptions, and
precautions in determining metamorphic temperatures and
pressures.)
Kretz, R. 1994. Metamorphic Crystallization. New York:
Wiley. (Chapter 2 provides a good description of relations
between thermodynamic properties and metamorphic
equilibria.)
Philpotts, A. R. 1990. Principles of Igneous and Metamorphic
Rocks. Englewood Cliffs: Prentice-Hall. (Chapter 9 of this
comprehensive textbook gives an especially good section on
solution models.)
Powell, R. 1978. Equilibrium Thermodynamics in Petrology.
London: Harper and Row. (An excellent introductory text
that dispenses with proofs; chapter 2 describes Ḡ-T diagrams,
and chapter 3 gives useful formulations for various standard
states not presented in the present book.)
Saxena, S. K. 1973. Thermodynamics of Rock-Forming Crystalline Solutions. New York: Springer-Verlag. (A more
advanced book that provides a wealth of information on
mixing models and solution behavior for real mineral
systems.)
Wood, B. J., and D. G. Fraser. 1977. Elementary Thermodynamics for Geologists. Oxford: Oxford University Press.
(A clear and concise introduction to thermodynamics, with
numerous worked examples of geologic interest; chapter 3
treats regular solution models, and chapter 4 provides a much
more detailed look at geothermometers and geobarometers
than is given in the present book.)
These papers were also referenced in this chapter.
Bohlen, S. R., V. J. Wall, and A. L. Boettcher. 1983. Experimental investigations and geological applications of equilibria
in the system FeO-TiO2-Al2O3-SiO2-H2O. American Mineralogist 68:1049–1058.
Eugster, H. P., and D. R. Wones. 1962. Stability relations of
the ferruginous biotite, annite. Journal of Petrology 3:82–
125.
Ferry, J. M., and F. S. Spear. l978. Experimental calibration
of the partitioning of Fe and Mg between biotite and garnet. Contributions to Mineralogy and Petrology 66:113–
117.
Helgeson, H. C., J. M. Delany, H. W. Nesbitt, and D. K. Bird.
1978. Summary and critique of the thermodynamic properties of the rock forming minerals. American Journal of
Science 278A:1–229.
Holdaway, M. J., and B. Mukhopadhyay. 1993. A reevaluation
of the stability relations of andalusite: Thermochemical
data and phase diagram for the aluminum silicates. American Mineralogist 78:298–315.
186
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Kerrick, D. M., ed. 1990. The Al2SiO5 polymorphs. Reviews of
Mineralogy 22. Washington, D.C.: Mineralogical Society
of America.
Zen, E. 1969. The stability relations of the polymorphs of aluminum silicates: A survey and some comments. American
Journal of Science 267:297–309.
PROBLEMS
(9.1)
Using figure 9.2, construct qualitative Ḡ-P diagrams for the Al2SiO5 system at temperatures T1 and T2.
(9.2)
From the understanding that (∂∆Ḡ/∂T )P = −∆S̄ and (∂∆Ḡ/∂P)T = ∆V̄, construct a schematic S̄-T diagram at pressure P1 and a V̄-P diagram at temperature T2 for the Al2SiO5 system (fig. 9.2).
(9.3)
Consider an isolated system containing the following phases at equilibrium: (K,Na)Cl,
(Na,K)AlSi3O8, and aqueous fluid. (a) Choose a set of components for this system. (b) State the conditions necessary for the system to be in complete equilibrium.
(9.4)
Clinoenstatite (MgSiO3) melts incongruently to forsterite (Mg2SiO4) plus liquid at 1557°C at atmospheric pressure. (a) Using thermodynamic data from some reliable source, calculate the chemical potential of SiO2 in the liquid, when these three phases coexist at equilibrium. (b) What will happen to
the system if µSiO2 is increased while T and P are held constant?
(9.5)
Using the data provided,
0
0
0
(a) Calculate ∆H̄800
, ∆S̄800
, and ∆Ḡ800
for the following reaction:
NaAlSi3O8 → NaAlSi2O6 + SiO2.
albite
jadeite
quartz
albite
jadeite
quartz
0
H̄298
(cal mol−1)
0
S̄ 298
(cal deg−1 mol−1)
C̄P
(cal deg−1 mol−1)
−937.146
−719.871
−217.650
50.20
31.90
9.88
61.7 + 13.9T/103 − 15.01 × 105/T2
48.2 + 11.4T/103 − 11.87 × 105/T2
11.2 + 8.2T/103 − 2.7 × 105/T2
0 for the reaction assuming ∆C̄ = 0. Compare the two values and comment.
(b) Calculate ∆Ḡ800
P
(9.6)
In worked problem 9.8, we calculated the slope of the andalusite-kyanite boundary in P-T space. Its
location was fixed using the experimentally determined triple point of Holdaway and Mukhopadhyay discussed in a highlighted box. Using the Clapeyron equation and data in worked problem 9.8,
construct the remainder of the aluminosilicate phase diagram. Assume that ∆C̄P = 0 and that V̄ is independent of P and T.
(9.7)
For the reaction:
Mg(OH)2 → MgO + H2O,
brucite
periclase
(a) Calculate the equilibrium temperature for the reaction at 1000 bars pressure using thermodynamic data from some appropriate source. Assume that ∆C̄P = 0, ∆V̄solids = constant, and that
H2O behaves as an ideal gas.
Temperature and Pressure Changes: Thermodynamics Again
187
(b) Calculate the equilibrium temperature under the same conditions as before, but this time with
PH2O = 0.5.
(9.8)
Using the values for molar volume, thermal expansion, and isothermal compressibility given below,
calculate V̄ for each of these phases at 800°C and 3 kbar.
(a) Almandine:
V̄298,1 = 118.2 cm3 mol−1;
αP = 25 × 10−6 deg−1;
β T = 1.27 × 10−6 bar−1.
(b) Quartz:
V̄298,1 = 22.69 cm3 mol−1;
αP = 69 × 10−6 deg−1;
β T = 26.71 × 10−6 bar−1.
(9.9)
Calculate ai as Xi increases, using Raoult’s Law and a symmetric regular solution model with W = 10 kJ,
and determine how small Xi can become before Raoult’s Law is no longer a good approximation.
(9.10) Given the reactions below, discuss what standard states you would choose for the chemical potentials:
(a) leucite + SiO2 (melt) → orthoclase, at 1 bar and 1170°C;
(b) EuAl 2Si2O8 (melt) → EuAl 2Si2O8 (plagioclase), at 1300°C and 1 bar;
(c) gypsum → anhydrite + water, at 100°C and 1 kbar.
CHAPTER TEN
PICTURING EQUILIBRIA
Phase Diagrams
OVERVIEW
Even after years of experience, many geologists find
thermodynamics difficult to understand unless they can
draw pictures that relate their equations to tangible
systems. Phase diagrams provide a convenient and powerful way to picture equilibria. In this chapter, we introduce the principles of Ḡ-X2 diagrams for binary systems
and illustrate graphically the minimization of free energy. We then construct T-X2 diagrams by stacking Ḡ-X2
sections atop one another, using examples of real geochemical systems. These diagrams illustrate simple crystallization, the formation of chemical compounds, solid
solution, and exsolution. We review equilibrium crystallization paths and show how binary phase diagrams can
be constructed from thermodynamic data. In addition, we
examine phase diagrams that employ fluid composition
or partial pressure as variables. With this background, we
then describe P-T diagrams for binary systems. Finally,
we briefly introduce graphical relationships for systems
with three components.
G¯-X2 DIAGRAMS
In chapter 9, we learned that a phase diagram is simply
a graphical representation of the stability fields of one
or more phases, drawn in terms of convenient thermo188
dynamic variables. In a one-component (unary) system,
the equilibrium state is determined when we specify two
thermodynamic variables, such as some combination
of P, T, S̄, or V̄. In our earlier discussion of the unary
system H2O, we considered a phase diagram with pressure and temperature as Cartesian axes, which showed
single-phase fields for water, ice, and vapor, as well as
lines along which two phases coexisted stably. To determine the state of a two-component (binary) system
completely, an additional compositional variable must
be specified. Thus, a complete phase diagram for a binary system is drawn in three dimensions. Once again,
we are free to choose any three thermodynamic functions as Cartesian axes, but for most geochemical applications, it is most convenient to select P, T, and
composition, normally expressed as X2 (the mole fraction of component 2). Of course, we could just as easily
choose X1 as our compositional variable, because X1 =
1 − X2.
Molar Gibbs free energy is a function of all three of
these variables (that is, Ḡ = Ḡ(P, T, X2)), so let us consider a diagram (fig. 10.1) in which a single phase with a
continuous compositional range between X2 = 0 and
X2 = 1 is represented. The curved line on this diagram is
a graph of the function Ḡ = Ḡ(P, T, X2), considered at a
fixed P and T. Now let’s examine some significant features of this diagram.
Picturing Equilibria: Phase Diagrams
FIG. 10.1. Ḡ -X2 diagram at fixed T and P for a phase with a continuous compositional range from X2 = 0 to 1. At any particular
composition, the ends of a tangent to the Ḡ curve define the
values for µ1 and µ2 in that phase, as illustrated by the dashed
line.
We can, of course, define the value of Ḡ for this phase
at any specific composition X2 by reading it directly
from the ordinate on the graph. It is more instructive,
however, to consider the individual contributions of the
chemical potentials µ 1 and µ 2 to Ḡ. We do this by constructing a tangent to the Ḡ-X2 curve at X2, as illustrated
in figure 10.1 by the dashed line. Because the first derivative of Ḡ with respect to X2 equals µ 2, the point at
which this tangent intersects the ordinate at X2 = 1 is
the value for µ 2 in the phase; similarly, the point at which
it intersects the ordinate at X2 = 0 (X1 = 1) is the value
of µ1 in the phase. Note that, although as drawn in figure 10.1, Ḡ reaches a minimum value at some point
between X2 = 0 and X2 = 1, the value of µ2 increases
monotonically (and the value of µ 1 decreases monotonically) as X2 increases. Also note that the value for either
chemical potential approaches a maximum as the composition approaches that of the pure chemical component. Therefore, µ02 represents the chemical potential of
component 2 in the phase at its pure X2 = 1 end member.
In this example, we have chosen a phase exhibiting
complete solid solution behavior. There are, of course,
many pure substances with no compositional variation
whatsoever. For such substances, the Ḡ-X2 curve reduces
to a point. (This point could be represented equally well
by a sharp vertical spike in the Ḡ-X2 diagram.)
If several phases are possible in the system, each will
have its own Ḡ-X2 curve or point. The equilibrium state
for the system is that phase or set of phases that yields a
minimum value of Ḡ for any specified value of X2 (still,
remember, at constant P and T). We can best visualize
the equilibrium states for various values of X2 by imag-
189
ining a tangent line or set of tangent lines that touch
the set of Ḡ-X2 curves or points, forming a sort of floor
to the diagram as illustrated by the dashed lines in figure 10.2a. In this example, we have one pure phase B and
two phases (A and C) that exhibit solid solution behavior.
Note that if we start at the left side of the diagram and
move gradually toward the right, the tangent to the left
limb of the Ḡ-X2 curve for phase A changes slope monotonically, which is to say that both µ1A and µ2A vary
monotonically. This situation continues until we reach
the value of X2 at which the tangent to the curve for
phase A bumps against the Ḡ-X2 point for phase B. Now
µ1A = µ2A and µ2A = µ1B. This is, of course, a condition
defined originally in chapter 3 as chemical equilibrium
between phases A and B. Beyond this value of X2, the
free energy of the system can be found along the common tangent and is lower than the free energy for either
FIG. 10.2. (a) Ḡ -X2 diagram at fixed T and P for a system containing three phases. Phase B is pure, but phases A and C show
limited ranges of composition. The equilibrium states for various
values of X2 are those phases which yield the minimum Ḡ for the
system. These can be readily determined from tangents to the Ḡ
curves, as shown by the dashed lines. (b) The chemical potential
for component 2 in the stable phase or phase assemblage in (a)
increases with X2, as illustrated in this diagram. µ2 is fixed by the
coexistence of two phases, but changes monotonically if only one
phase is present.
190
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
phase A or B alone. The two-phase assemblage A + B
continues to be stable until we reach the value of X2 corresponding to the composition of B. At this point, only
phase B is stable, and the tangent pivots instantaneously
at B to become the tangent to both B and C. Now, B + C
becomes the stable phase assemblage. A similar pattern
occurs as we continue to the right and encounter a region in which phase C alone is the stable phase.
We see, therefore, that the chemical potential of either
component 1 or 2 changes monotonically across singlephase fields as we vary X2, but both chemical potentials
remain fixed in two-phase fields. This can be seen graphically in figure 10.2b, a schematic construction of the
variation of µ2 in the stable phase or phase assemblage
as X2 ranges from 0 to 1 in the system in figure 10.2a.
When two phases are stable, a binary system has only two
degrees of freedom (T and P), as dictated by the phase
rule.
DERIVATION OF T-X 2 AND
P-X 2 DIAGRAMS
A Ḡ-X2 diagram such as those we have just considered
is valid only for the fixed values of T and P that were
specified when drawing the diagram. To construct a
phase diagram for the system that takes into account
changing temperature and pressure conditions, we must
recall from chapter 9 how Ḡ varies with T and with P at
constant X2:
(∂Ḡ/∂T )P, X2 = −S̄,
(10.1)
and
(∂Ḡ/∂P)T, X2 = V̄.
(10.2)
Because S̄ and V̄ are inherently positive quantities, all
Ḡ-X2 curves will shift upward or downward simultaneously with increasing P or T, respectively. However,
because the specific S̄ and V̄ values for each phase are
different and because S̄ and V̄ are themselves functions
of composition, relative movements of the Ḡ-X2 curves
and changes in their shapes will generally occur. Consider, for example, a sequence of Ḡ-X2 diagrams drawn
at T1, T2, and T3 , a set of arbitrarily decreasing temperatures, illustrated in figure 10.3. With increasing temperature, the Ḡ-X2 curves for all three phases in the system
move upward, but in this example, the curve for phase B
shifts more than do the other two curves. As a result, the
curves for all three phases have a common tangent at the
unique temperature T2, and we have the condition µ1A =
µ1B = µ1C; that is, a three-phase field A + B + C exists at
FIG. 10.3. A series of three Ḡ -X2 diagrams at fixed T and P for a succession of temperatures decreasing from T1 to T3 . The Ḡ curves for phases A, B, and C shift downward at different rates with
decreasing temperature, and thus different combinations of phases are stable.
Picturing Equilibria: Phase Diagrams
FIG. 10.4. A “stack” of G-X2 diagrams at various temperatures
produces a T-X2 diagram. This particular diagram results from
stacking the three Ḡ -X2 diagrams shown in figure 10.3. The compositions of coexisting phases are connected by horizontal lines,
because at equilibrium, the temperatures of all coexisting phases
must be the same.
T2, replacing the fields A + B, B only, and B + C that were
stable at temperatures below T2. A further increase in
temperature lifts the curve for phase B above the common tangent to A and C, and thus phase B is no longer
stable either alone or in combination with A or C. The
diagram for T1 illustrates this condition. Phase B is said
to be metastable with respect to A + C.
Now let us visualize Ḡ-X2 diagrams drawn over a
continuous range from T1 to T3, and extending to higher
and lower temperatures beyond these limits. The one-,
two-, and three-phase fields in these diagrams will be
seen to change shapes continuously. If we now envision
a “stack” of all such diagrams and rotate it so that it can
be viewed from the side with temperature as the vertical
axis and X2 the horizontal one, we have created a T-X2
diagram for the system. Such a diagram is illustrated in
figure 10.4.
Notice that as in the Ḡ-X2 diagrams, phase fields for
single phases and pairs of phases alternate from left to
right across the composition axis. The three-phase field
A + B + C exists only at temperature T2. Phase B does not
exist for any value of X2 at temperatures above the line
A + B + C. The appearance and disappearance of phases
are probably the most complicated and confusing parts
of phase diagrams. Other kinds of phase changes are illustrated and explained in the next section.
Note also the set of tie-lines parallel to the X2 axis in
figure 10.4. These connect the compositions of phases
A + B, B + C, or A + C that are in equilibrium in the twophase fields at each temperature. Because it is a necessary
191
condition of equilibrium that the temperatures of all
phases be the same, tie-lines must always be isothermal.
It may not be immediately apparent that a point on
any tie-line has a significance in terms of the relative
proportions of phases in the two-phase field. We can, in
fact, define a lever rule using such a point. The proportion
of each phase in the field at that point varies inversely as
the length of the tie-line from the point to the composition of that phase. For example, consider the two-phase
field A + C in figure 10.4. A point on the tie-line in this
field at a composition one quarter of the way from the
left edge (the composition of phase A) represents a mixture of 75% phase A and 25% phase C, and the tie-line
segments to the left and right of the point are 25% and
75% of the total length, respectively. This configuration
is illustrated in figure 10.5. Proportions of the phases in
a two-phase field can thus be determined by inspection.
The process by which we just developed a T-X2 diagram from Ḡ-X2 sections could also be duplicated to
generate a P-X2 diagram. In general, however, the appearance of a P-X2 diagram will be inverted from that
of a T-X2 diagram, because increases in temperature and
pressure have inverse effects on most materials; that is,
high temperature favors higher V̄ and higher S̄, but high
pressure favors lower V̄ and lower S̄. P-X2 diagrams are
not commonly used in solving geochemical problems.
Instead, the effects of pressure are generally gauged from
P-T diagrams, considered in a later section. For the time
being, we focus on the characteristics of T-X2 diagrams.
T-X 2 DIAGRAMS FOR REAL
GEOCHEMICAL SYSTEMS
Now we look at some Ḡ-X2 and corresponding T-X2
diagrams for real binary systems, to illustrate the geochemically important features commonly encountered in
FIG. 10.5. This balance schematically illustrates the lever rule.
For the two-phase field A + C in figure 10.4, a point on the tieline one quarter of the way from the composition of A represents
a mixture of 75% phase A and 25% phase C, and the tie-line segments on either side of this point are 25% and 75%, respectively.
192
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
phase diagrams for geologic materials. In most of the
following examples, a liquid phase is stable at high temperatures and solidifies to form other phases as the
temperature is lowered. Each of these systems will be
considered only at atmospheric pressure, at least for the
moment. We present a series of Ḡ-X2 diagrams for various temperatures at constant pressure, and then “stack”
these to construct the appropriate T-X2 diagram for that
system. We can then explore the characteristics of each
T-X2 diagram in detail.
Simple Crystallization in a Binary System:
CaMgSi2O6-CaAl2Si2O8
The simplest case we consider is an example of a twocomponent system that is liquid at high temperature and
crystallizes to form two pure solid phases, diopside
CaMgSi2O6 and anorthite CaAl2Si2O8, as shown in figure 10.6. First let us examine some Ḡ-XAn diagrams for
this system at T1 through T4 , representing a progression
of decreasing temperatures. At T1, a tangent to the Ḡ-XAn
curve for liquid (L) defines the lowest free energy for the
system at any XAn, so that liquid is the only stable phase.
By T2 , the Ḡ point for anorthite has dropped farther
than the curves and points for the other phases, so that
a tangent to anorthite and liquid offers the lowest free
energy configuration for the system at high XAn values.
By T3 , the Ḡ for diopside has lowered to the point at
which Di + L or An + L are the stable phases over nearly
the entire compositional range, but at one special composition, all three phases are stable. At T4, liquid is metastable, because a tangent to the two solid phases lies
below the liquid Ḡ curve.
By combining the information on phase stability in
these Ḡ-XAn diagrams, we can construct a T-XAn diagram
for the system diopside-anorthite, as illustrated at the bottom of figure 10.6. Examination of this diagram reveals
some interesting features, probably already familiar to
petrology students. The curve defining the upper boundary of the area in which liquid coexists with a solid phase,
called the liquidus, represents the onset of crystallization.
The lower boundary of this area denotes a region in
which liquid is not a stable phase. This boundary, called
the solidus, defines the temperature below which the
system is fully crystallized. The point at which the liquidus touches the solidus is called a eutectic. It represents the composition at which the liquid can persist to
the lowest temperature, as well as the point at which
FIG. 10.6. Ḡ -XAn diagrams for the system diopside-anorthite at
P = 1 atm and at various temperatures can be stacked to produce
the T-XAn diagram at the bottom. Phases and their abbreviations
are: diopside (Di), anorthite (An), and liquid (L). This diagram
contains a eutectic, representing the three-phase equilibrium
between Di, An, and L.
both diopside and anorthite crystallize together with falling temperature.
Worked Problem 10.1
What is the equilibrium crystallization sequence for a liquid in
the diopside-anorthite system with composition XAn = 0.2? The
system, initially all liquid, begins to crystallize diopside as it
cools and intercepts the diopside liquidus. From figure 10.6,
we can see that at XAn = 0.2, this occurs at a temperature of
approximately 1350°C. During subsequent cooling, the composition of the remaining liquid follows the liquidus down slope,
changing as diopside is removed. Finally, the liquid reaches the
eutectic, at which anorthite and diopside crystallize together.
The system can cool no further until the liquid is fully crystallized. Following solidification of the last drop of liquid, the
mixture of diopside and anorthite crystals cools to a lower
temperature without further change. The final proportions of
phases can be determined from the position of this composition
along the horizontal axis of the diagram, using the lever rule:
80% diopside and 20% anorthite.
Picturing Equilibria: Phase Diagrams
Formation of a Chemical Compound in a
Binary System: KAlSi2O6-SiO2
Now we examine a system in which the Ḡ-X 2 curves
for two phases having the same composition intersect.
Figure 10.7 illustrates Ḡ-Xsilica diagrams for the system
leucite KAlSi2O6-tridymite SiO2 at various temperatures.
At T1 , liquid is the stable phase at any composition. First
the Ḡ point for cristobalite, and then that for leucite,
drop with lowering temperature to produce the situation
at T2. So far, this system does not differ in its behavior
from the diopside-anorthite system we have just exam-
193
ined; however, this changes at T3 . The composition of
orthoclase KalSi3O8 can be represented as a mixture of
KAlSi2O6 and SiO2, as shown by its position on the horizontal axis of figure 10.7. At T3 , the orthoclase Ḡ point
coincides with the Ḡ curve for liquid (that is, orthoclase
and liquid have the same free energy), and by T4 , the
orthoclase point has pierced the liquid curve to replace
liquid as the stable phase over part of the compositional
range. Liquid has become metastable by T5.
The T-Xsilica diagram produced from this information
is shown at the bottom of figure 10.7. The liquidus is
indicated by the curves touching the liquid (L) field, and
the solidus is defined by the lines that separate liquidbearing fields from fields with only solid phases. The right
side of the diagram contains a eutectic and is similar to
the diopside-anorthite system discussed earlier. However, cristobalite transforms to tridymite at 1470°C,
because the tridymite Ḡpoint descends more rapidly than
that for cristobalite and eventually overtakes it with
lowering temperature. The left side of the T-Xsilica diagram contains a new feature, called a peritectic (a point
of reaction between solid and liquid), produced when
the orthoclase Ḡ point intersects the Ḡ curve for liquid.
Note that in the Ḡ-Xsilica section at T3 , the tangent between leucite and liquid has been replaced by tangents
between each of these phases and orthoclase. This means
that leucite and liquid are no longer stable together at
this temperature; thus, these phases react to produce a
new phase with intermediate composition, orthoclase.
Worked Problem 10.2
FIG. 10.7. Ḡ -XSiO2 diagrams for the system leucite-silica at P = 1
atm and at a progression of temperatures are stacked to make the
T-XSiO2 diagram at the bottom. Phases and their abbreviations are:
leucite (Lc), orthoclase (Or), cristobalite (Cr), tridymite (Tr), and
liquid (L). This diagram contains a eutectic representing equilibrium between Or, Tr, and L, and a peritectic representing the
reaction between Lc and L to form Or.
What is the equilibrium crystallization sequence for a composition Xsilica = 0.3 in the system leucite-silica? The first phase
to crystallize from a liquid of this composition is leucite, as
the temperature drops to ∼1425°C. With continued cooling, the
liquid follows the liquidus down slope until intersecting the
peritectic. At this point, already crystallized leucite reacts with
liquid to produce orthoclase (at T3, a tangent joining the compositions of orthoclase and liquid replaces that between leucite
and liquid), and the temperature of the system cannot decrease
until all of the leucite is consumed. When the reaction is
complete, the liquid resumes its descent along the liquidus, all
the while crystallizing more orthoclase. When it reaches the
orthoclase-tridymite eutectic, these two phases crystallize together, and after final solidification, the mixture of orthoclase
and tridymite crystals cools to a lower temperature. The final
proportion of phases can be ascertained from the position of
the initial liquid composition along the horizontal axis of figure 10.7. Application of the lever rule, this time using tridymite
194
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
and orthoclase (rather than leucite) as the ends of the tie-line,
gives ∼87% orthoclase and the remainder tridymite.
Let’s now try the crystallization sequence for a composition
Xsilica = 0.l in this system. Minerals crystallize in the same sequence as before, but in this case, the liquid will never be able
to persist past the peritectic. The final product for a liquid of
this composition must be a mixture of leucite + orthoclase,
because that is the mixture of solids in the two-phase field at
Xsilica = 0.1 below temperature T5. Therefore, the liquid phase
must be consumed before all of the leucite reacts to form
orthoclase.
Solid Solution in a Binary System:
NaAlSi3O8-CaAl2Si2O8
Albite (NaAlSi3O8) and anorthite (CaAl2Si2O8) exhibit
complete solid solution in plagioclase. Some Ḡ-XAn sections for this system are shown in figure 10.8. In this
case, the plagioclase composition that is stable at any
temperature is determined by the interaction of the Ḡ-XAn
curves for plagioclase and liquid. At temperatures above
FIG. 10.8. Ḡ -XAn diagrams for the system albite-anorthite at
P = 1 atm and at various temperatures are combined to make the
T-XAn diagram at the bottom. Plagioclase shows a complete solid
solution between these end members, forming a loop.
T1, only liquid is stable. At T1, the plagioclase Ḡ curve
intersects the liquid Ḡ curve at XAn = 1. As the temperature decreases, the plagioclase Ḡ curve sweeps downward
across the liquid Ḡcurve, intersecting it obliquely at lower
and lower XAn values. Thus, at T2 and T3 , compositions
with high XAn have only plagioclase stable, intermediate
compositions have both plagioclase and liquid, and those
with low XAn contain only liquid. At some value between
T3 and T4 , the plagioclase and liquid Ḡcurves intersect at
XAn = 0. At T4 , liquid is metastable at any composition.
The corresponding T-XAn diagram for the plagioclase
system is shown at the bottom of figure 10.8. This diagram contains a loop that defines the boundaries for the
liquid-plagioclase field. The upper curve is thus the liquidus, and the lower curve is the solidus. The liquidus
gives the composition of the liquid coexisting with a
solid whose composition is specified by the solidus at the
same temperature.
Worked Problem 10.3
Describe the equilibrium crystallization path for a liquid composition XAn = 0.65. The final product will obviously be crystals with composition An65, as this is a solid solution series. The
path by which this product is formed is illustrated in figure 10.9.
Crystallization begins when the liquid intersects the liquidus at
point a, and the first crystals to form have composition b. As the
liquid cools further to point c, the intersections of a horizontal
line through this point with the liquidus (point d) and solidus
(point e) give the compositions of the liquid and solid, respectively, that are in equilibrium at this temperature. Recall that
two phases in equilibrium must have the same temperature, so
their compositions are connected by a horizontal tie-line in this
diagram. As cooling continues, the composition of the liquid
follows the liquidus down slope, and that of crystallizing
plagioclase follows the solidus. The intersection of liquid and
solid Ḡ curves seen in the previous diagram indicates that a reaction takes place between liquid and already crystallized solid,
so that anorthite-rich plagioclase continually re-equilibrates to
form more albite-rich plagioclase that is stable at lower temperatures. Finally, the liquid reaches point f, its farthest point of
descent along the liquidus. A horizontal line through this point
passes through point g, the composition of the solid at this temperature. Because this composition is XAn = 0.65, the required
final composition for plagioclase, reaction with liquid must end
here and the liquid disappears. Crystals of An65 composition
then cool to low temperature without further change. (Actually,
there are some subsolidus reactions, similar to those that take
place in alkali feldspars discussed in the next example, which
may take place in slowly cooling plagioclases, but we will ignore those here.)
Picturing Equilibria: Phase Diagrams
FIG. 10.9. An example of an equilbrium crystallization path in
the system albite-anorthite. The liquid cools to point a, where
crystallization begins with crystals of composition b. With falling
temperature, liquid changes composition progressively from a to f,
and solid from b to g.
Unmixing in a Binary System:
NaAlSi3O8-KAlSi3O8
Finally, we consider a system with an inflection in a
Ḡ-X2 curve. Such inflections cause a single homogeneous phase to separate into two new phases as it
cools. A natural example is the system albite NaAlSi3O8-
195
orthoclase KAlSi3O8, a solid solution at high temperature that unmixes into discrete alkali feldspars at lower
temperature. For our purposes here, we are only interested in what happens below the solidus in this system,
so we ignore temperatures above the solidus.
Several Ḡ-XOr sections are illustrated in figure 10.10.
At T1, the alkali feldspars exhibit complete solution at
any composition. At T2, a small inflection in the alkali
feldspar Ḡ curve appears. Notice that a tangent providing the lowest system free energy touches the curve at two
places, indicating that two stable feldspar phases should
exist, at least for a small compositional range near the
middle of the diagram. With decreasing temperature, the
inflection widens and thereby increases the compositional range affected, as shown at T3.
The T-XOr diagram generated from this phenomenon
is illustrated at the bottom of figure 10.10. The solidus
is shown at the top of this diagram to emphasize the
point that we are looking at what happens in the subsolidus region. The gradually widening tangent points in
the Ḡ-XOr diagrams define a concave-downward curve,
called a solvus, which identifies a region of unmixing or
exsolution. Above the solvus, one homogeneous phase is
stable, and below it, this phase separates into two phases
with distinct compositions.
Worked Problem 10.4
What is the subsolidus crystallization path for the composition
XOr = 0.6 in the alkali feldspar system? With cooling, a homogeneous feldspar of composition XOr = 0.6 reaches the solvus at
approximately 630°C (fig. 10.10). At this point, the homogeneous feldspar begins to unmix in the solid state, producing
two coexisting feldspars. At any temperature below the solvus,
intersections of a horizontal (isothermal) line with the two limbs
of the solvus give the compositions of the two stable phases. For
example, at 550°C, the coexisting phases will have compositions of XOr = 0.14 and 0.72. Notice that the compositions of
these phases become further separated as temperature decreases
and the solvus widens. At some low temperature, the diffusion
of ions becomes so sluggish that unmixing ceases, and the compositions of the exsolution products are frozen in at that point.
A photograph of an exsolved alkali feldspar formed in this way
is illustrated in figure 10.11.
FIG. 10.10. Ḡ -XOr diagrams for the system albite-orthoclase at
P = 1 atm and at various temperatures are stacked to produce the
T-XOr diagram at the bottom. An inflection in the Ḡ curve with
lowering temperature creates a region of unmixing in the solid
state, called a solvus. Abbreviations for phases are: feldspar (Fsp)
and liquid (L).
We have now examined the common features of binary T-X2 diagrams. Simple crystallization of pure phases
produces a eutectic. Reactions between liquid and solid
either form a compound at a peritectic, or result in solid
196
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
arate equations to express the chemical potentials of albite and
anorthite in the plagioclase solid solution and in the liquid:
µAb, S = µ0Ab,S + RT ln XAb,S,
FIG. 10.11. Photomicrograph of perthite, an intergrowth of sodic
and potassic alkali feldspars, formed by exsolution. The white
lamellae are albite, and the twinned host phase is K-feldspar. This
sample is from the Winnsboro granite, South Carolina. Width of
the photomicrograph is ∼2 mm.
µAb, L =
+ RT ln XAb, L,
(10.3b)
µAn, S =
µ0An, S
+ RT ln XAn, S,
(10.3c)
µAn, L = µ0An, L + RT ln XAn, L.
(10.3d)
The two phases, liquid (L) and solid (S), are indicated by the
appropriate subscripts above. In each case, the chemical potentials with the 0 superscripts are standard state values (using the
conventional standard state defined by pure plagioclase or pure
liquid of end-member compositions). For this exercise, we
assume that both the liquid and the solid are ideal solutions, so
that ai = Xi.
Within the loop (the two-phase field) we know that µAb, S
must equal µAb, L (in other words, equations 10.3a and 10.3b
must be equal), and likewise that equations 10.3c and 10.3d must
also be equal. The consequence of two-phase equilibrium, therefore, is that:
µ0Ab, S − µ0Ab, L = RT ln(XAb, L /XAb, S )
solution described by a loop. Unmixing produces a solvus.
Although binary systems are relatively simple, they nevertheless provide important insights into the crystallization
of multicomponent systems of geologic interest.
THERMODYNAMIC CALCULATION
OF PHASE DIAGRAMS
Up to this point, we have presented without proof a series
of qualitative Ḡ-X2 diagrams and their corresponding
T-X2 diagrams. We confess that to do this, we first worked
the problems backwards, by using experimentally determined T-X2 diagrams to construct qualitatively correct
Ḡ-X2 sections. However, if we had enough good thermodynamic data, we could do each of the examples in the
preceding pages just as we presented them; that is, we
could calculate rigorous Ḡ-X2 diagrams, and from them,
construct T-X2 diagrams. In practice, it is often easier to
omit the Ḡ-X2 diagrams and to calculate the T-X2 sections directly. This problem is similar to the calculation
of phase boundaries in the P-T diagram for the unary
system Al2SiO5 that was introduced in chapter 9, with
the added complication that the mixing properties of the
two binary components must be modeled.
Worked Problem 10.5
Let’s see how accurately we can construct the albite-anorthite
phase diagram from thermodynamic data. We can write four sep-
(10.3a)
µ0Ab, L
(10.4a)
and
µ0An, S − µ0An, L = RT ln(XAn, L /XAn, S ).
(10.4b)
The chemical potentials with the 0 superscripts are simply
the free energy values for pure albite, pure anorthite, and the
pure liquids of end-member composition. The left sides of equations 10.4a and 10.4b therefore correspond to the free energies
of melting pure albite and pure anorthite. Because ∆Ḡ = ∆H̄ −
T∆S̄ (∆H̄ and ∆S̄ in this context refer to the enthalpy or entropy
changes associated with the melting process), we now have:
µ0Ab, S − µ0Ab, L = ∆H̄Ab − T∆S̄Ab
(10.5a)
and
µ0An, S − µ0An, L = ∆H̄An − T∆S̄An.
(10.5b)
These can be substituted into equations 10.4a and 10.4b, respectively, to give
RT ln(XAb, L /XAb, S ) = ∆H̄Ab − T∆S̄Ab
(10.6a)
and
RT ln(XAn, L /XAn, S ) = ∆H̄An − T∆S̄An.
(10.6b)
Let us now take the exponential of both sides of equations
10.6a and 10.6b to solve for the mole fractions. From 10.6a
we get:
(XAb, L /XAb, S) = exp((∆H̄Ab − T∆S̄Ab )/RT ),
or
XAb, L = XAb, S exp((∆H̄Ab − T∆S̄Ab )/RT ).
(10.7a)
Equation 10.6b similarly yields:
XAn, L = XAn, S exp((∆H̄An − T∆S̄An )/RT).
(10.7b)
Picturing Equilibria: Phase Diagrams
Now recall that the mole fractions XAb and XAn are not independent, but are related by:
XAn, L = 1 − XAb, L
(10.8a)
and
XAn, S = 1 − XAb, S.
(10.8b)
From equations 10.7a, 10.7b, and 10.8a, we can derive:
XAn, S exp((∆H̄An − T∆S̄An )/RT ) =
1 − XAb, S exp((∆H̄Ab − T∆S̄Ab )/RT ).
We can eliminate XAn, S from this equation by substituting for it
from equation 10.8b:
(1 − XAb, S )exp((∆H̄An − T∆S̄An )/RT ) =
1 − XAb, S exp((∆H̄Ab − T∆S̄Ab )/RT ).
If we rearrange this result to solve for XAb, S̄,
XAb, S = [1 − exp((∆H̄An − T∆S̄An )/RT)]/[exp((∆H̄Ab
− T∆S̄Ab )/RT) − exp((∆H̄An − T∆S̄An)/RT )].
(10.9)
When the temperature is equal to the melting temperature
of albite (that is, when T = TAb), we know that equation 10.5a
is equal to zero. In other words, the liquidus and solidus curves
coincide for pure albite. Thus:
∆H̄Ab = TAb∆S̄Ab.
With this relationship, we can recast equation 10.9 into:
(
)
∆H̄An
1 − exp ———–———
R(1/T − 1/TAn)
XAb, S = ——————————————————— . (10.10)
∆H̄Ab
∆H̄Ab
exp ———–———
− exp ———–———
R(1/T − 1/TAb)
R(1/T − 1/TAn)
(
) (
)
This equation contains the enthalpies of melting for albite
and anorthite, which have been determined calorimetrically.
Numerically, ∆H̄Ab and ∆H̄An are equal to 13.1 kcal mol−1 and
197
29.4 kcal mol−1, respectively. With these values in hand, we can
calculate the mole fraction of albite in the solid solution at any
temperature above the melting temperature of pure albite from
equation 10.10. The calculated set of compositions and temperatures defines the solidus. The liquidus curve can be determined from these values with the help of equation 10.8a. These
calculated values are tabulated below and plotted graphically in
figure 10.12. This phase diagram constructed from thermodynamic data is an almost perfect fit to the plagioclase phase
diagram determined by experiment.
T (°C)
Solidus (XAb )
Liquidus (XAb )
1100
1150
1200
1250
1300
1350
1400
1450
1500
1550
1.000
0.983
0.955
0.913
0.852
0.767
0.650
0.491
0.280
0.000
1.000
0.830
0.689
0.579
0.463
0.366
0.275
0.185
0.095
0.000
In the example above, we calculated a T-X2 phase
diagram that was already known from experimental
petrology. This is an interesting academic exercise, but it
could be considered a waste of time insofar as we have
learned nothing new about plagioclase crystallization.
However, it illustrates an important point. There are
other geochemically important systems for which experimentally derived phase equilibria are difficult to obtain
on any reasonable time scale. Calculation of equilibrium
relationships in these systems from thermodynamic data
may be the most efficient way of attacking such problems.
There is another cogent reason for calculating phase
diagrams. If the phase relationships in a system can be
modeled accurately under one set of physical conditions,
then these relationships can be calculated under other
physical conditions. In the example above, the phase diagram was determined at atmospheric pressure. Now
that we have demonstrated that we can model this system, we could calculate the phase diagram at elevated
pressures by using appropriate thermodynamic data and
the relationships (solution models, corrections for compressibility, etc.) derived in the previous chapter.
BINARY PHASE DIAGRAMS
INVOLVING FLUIDS
FIG. 10.12. The plagioclase solid solution loop calculated in
worked problem 10.5.
We might also wish to show on phase diagrams the
effects of parameters other than temperature, pressure,
and X2. The nature of the fluid phase in a system can be
198
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 10.13. Pfluid-T diagram for the system MgO-H2O, showing how
→ periclase + H O depends on P , even
the reaction brucite ←
2
fluid
though confining pressure remains constant at 1 kbar.
a particularly important variable. For example, fluid pressure may be decidedly less than confining (total) pressure
in some metamorphic systems because fluids have been
progressively driven off. In such a case, Pfluid can be considered an independent variable. Employing the same
technique we used earlier to derive T-X2 diagrams, we
could envision Ḡ-Pfluid diagrams at various values of T
and at constant Ptotal, and “stack” these to construct a
T-Pfluid phase diagram. However, we will spare you the
derivation and go directly to a real geochemical example.
Phase equilibria in the system MgO-H2O are illustrated
in figure 10.13. This particular diagram is for a fixed confining pressure of 1000 bars and a fixed system composition such that XH2O equals that in brucite. We can
readily see that the temperature at which the equilibrium
→ MgO + H O
Mg(OH)2 ←
2
brucite periclase
occurs is controlled by Pfluid , even though Ptotal remains
constant. Notice that the slope of this reaction curve, like
most involving a fluid phase, becomes markedly steeper
at high fluid pressures, as we learned from consideration
of the Clapeyron equation in chapter 9.
Even in systems in which Pfluid equals Ptotal, it is sometimes necessary to treat the fluid composition as a separate parameter. For example, the fluid in contact with a
marble undergoing metamorphism may be a mixture of
H2O and CO2. But only one of these fluid components
(CO2) exerts an influence on reactions in the marble,
such as:
FIG. 10.14. T-XCO2 diagram at constant confining pressure
(2 kbar), showing the effect of fluid composition on the reaction
→ wollastonite + CO . X in the fluid can be
calcite + quartz ←
2
CO2
altered by mixing with H2O.
CaCO3 + SiO2 → CaSiO3 + CO2.
calcite quartz wollastonite
(10.11)
In this case, we might want to devise a phase diagram
that shows the relationship of this equilibrium to the
composition of the fluid. The reaction as written above
is univariant. Adding H2O to the system increases the
number of components by one, but does not change the
number of phases, because CO2 and water are miscible.
Consequently, the variance is two when calcite, quartz,
and wollastonite coexist with a mixed CO2-H2O fluid.
But by specifying one of the variables (in this example,
we fix Ptotal at 2000 bars), we can represent this equilibrium by a univariant curve on a plot of T versus XCO .
2
Such a diagram is illustrated in figure 10.14. In this example, we have specified fluid composition in terms of
XCO2, but we could have used PCO2 as well, since PCO2 =
Ptotal XCO2.
Worked Problem 10.6
To illustrate the usefulness of phase diagrams involving fluid
components, compare the conditions under which wollastonite
forms in a closed and an open system. From figure 10.14, we
can see that siliceous limestone heated to approximately 740°C
will form wollastonite, but the same reaction occurs at much
lower temperatures if the resulting CO2 fluid is diluted with
H2O (thus lowering XCO ).
2
Picturing Equilibria: Phase Diagrams
199
First, we assume that the siliceous limestone contains a pore
fluid that is a mixture of water and CO2 with XCO = 0.3. What
2
happens during metamorphism under closed-system conditions?
Inspection of figure 10.14 indicates that wollastonite will begin
to form at 625°C. However, this reaction releases CO2, thereby
changing the composition of the fluid in the pore space. The
temperature must then increase for the reaction to continue.
Provided that the rocks are hot enough, the equilibrium follows
the reaction curve, and the fluid becomes progressively richer in
CO2. The fluid composition is controlled by the mineral assemblage in the rock itself; in other words, the system is internally
buffered. At some point (say XCO = 0.6, corresponding to T =
2
695°C), one of the reactants is finally consumed, and the reaction ceases.
Now let us consider the same reaction in an open system,
in which an infiltrating fluid with XCO = 0.3 is derived from
2
adjacent rocks (outside of the system as we have defined it).
We will stipulate that this fluid permeates through the limestone
in such large quantities that it can flush away any CO2 generated by local reactions. In this case, the fluid composition is externally buffered and does not change during metamorphism.
So, the wollastonite-producing reaction occurs at 625°C,
and the temperature remains constant during the course of the
reaction.
P-T DIAGRAMS
So far, we have been dealing mostly with isobaric slices
through P-T-X2 space, largely because we are constrained
by the two-dimensional nature of a sheet of paper. It is
possible to extend our results into three dimensions, however. As an example, let us consider the system Fe-O, a
sketch of which is illustrated in figure 10.15a.
What happens to the various univariant equilibria that
we have already defined in such a three-dimensional diagram? A eutectic (the equilibrium between three phases)
traces out a ruled surface in P-T-X2 space. The surface
is everywhere parallel to the X2 axis because we have the
equilibrium condition that P and T must be identical in
all phases. A peritectic traces out a three-dimensional
curve in P-T-X2 space. An equilibrium involving a pure
substance defines a two-dimensional curve lying completely in the plane of constant composition (X2). These
univariant curves and surfaces divide P-T-X2 space into
one- and two-phase volumes (with variances of two or
higher) as they intersect at invariant points or lines.
We have already examined some T-X2 slices (and
briefly mentioned the form of P-X2 slices) through such
a P-T-X2 diagram, so let us now consider the remaining
two-dimensional section, a P-T diagram with fixed com-
FIG. 10.15. (a) P-T-X2 diagram for the system Fe-O. (b) P-T
diagram for the system Fe-O obtained by projecting the threedimensional diagram in part (a) onto the P-T plane.
position. Of all the diagrams we have discussed, P-T
diagrams are perhaps the most commonly used and certainly the most difficult to understand.
Inspection of the Fe-O diagram in figure 10.15a suggests the key features of such diagrams. The P-T diagram
is obtained by collapsing the three-dimensional diagram
onto the P-T plane (that is, projecting parallel to the X2
axis), as illustrated in figure 10.15b. On this P-T projection, univariant equilibria appear as curves. Invariant equilibria appear as intersections of these curves,
although not all apparent intersections are invariant
points, because some curves “intersect” only in projection
and not in three-dimensional space.
200
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 10.16. (a) An example of one of the 14 ways in which univariant (three-phase) curves can
intersect on a P-T diagram. P1, P2, and P3 are three pressures for which representative T-X2 diagrams
are constructed. (b) Isobaric T-X2 diagrams at the three pressures identified in part (a). Figure out
for yourself where the two- and three-phase fields are in these diagrams.
There are many unique topological ways (14, to be
precise) in which univariant curves may intersect on a
P-T diagram. Derivation of these, or even presentation
of them, is beyond the scope of this book. Here we
consider only one such intersection, just to illustrate
how this relates to its corresponding T-X2 and Ḡ-X2
sections.
The intersection that we examine is illustrated in figure 10.16a. Each of the univariant lines on this diagram
describes an equilibrium between three of the possible
phases A, B, C, and D. Each might also be thought of as
a line describing a reaction. To see this, refer to the T-X2
diagrams in figure 10.16b. (It probably isn’t obvious
that these T-X2 sections are not topologically unique,
but that is the case. The topology of the set of T-X2 sections that can be drawn from this single invariant point
depends on its orientation relative to the P and T axes.
In fact, three distinct topologies can be constructed for
different arrangements of isotherms and isobars. We
consider only one.)
In the T-X2 diagram at pressure P1, there are two univariant three-phase equilibria shown. The first, encoun-
tered at low temperature, is the one in which phases B,
C, and D coexist stably. The second, at a somewhat higher
temperature, is the one in which phases A, B, and D are
stable. Inspection of the P-T diagram (fig. 10.16a) indicates the same sequence of equilibria with increasing
temperature along the P1 isobar. Similarly, the T-X2 diagram drawn at pressure P3 (fig. 10.16b) can easily be
correlated with the sequence of phase equilibria appearing with increasing temperature on the P3 isobar. The
T-X2 diagram at P2, however, contains only one line of
interest, the one at which phases A, B, C, and D are all
stable together. This corresponds to the unique point
where all four three-phase lines on the P-T diagram
intersect—the invariant point.
Now consider the significance of any of these univariant lines as a reaction. The three-phase line marked
ABD, for example, is a line along which A, B, and D coexist. It is also a line separating fields for phases A, B,
and D on the low-temperature side from a two-phase field
at A + D on the high-temperature side (fig. 10.16b). Therefore, the three-phase line ABD may also be viewed as a
description of the reaction:
Picturing Equilibria: Phase Diagrams
201
→ A + D,
B←
in which B is consumed in the formation of A and D with
increasing temperature. Similarly, the three-phase lines
ACD, ABC, and BCD may be seen as reactions in which
phases C, B, and D are consumed. The invariant (fourphase) point ABCD represents a reaction:
→ A + D,
B+C←
in which both B and C are consumed. These relations
should not come as any surprise, of course, because they
are what we have already been led to expect from our
earlier treatment of Ḡ-X2 diagrams.
SYSTEMS WITH THREE COMPONENTS
It becomes increasingly difficult to handle graphically
systems with three or more components. Although we
will not consider this subject in great detail, you should
at least be aware of how such systems are presented.
For systems containing a liquid phase, we commonly
combine three binary T-X2 sections into a single triangular diagram, as illustrated for the system diopsidealbite-anorthite in figure 10.17a. The three binaries fold
up to form a triangular prism, whose vertical axis represents temperature, as illustrated in figure 10.17b. The
upper lid of the prism is the liquidus, now a curved,
isobaric T-X surface. Three-dimensional perspective
sketches such as figure 10.17b are difficult to draw, and
it is almost impossible to derive quantitative information
from them, so details of the liquidus surface are commonly projected onto the compositional base of the prism,
as shown in figure 10.17c. The slopes of the liquidus
surfaces with temperature are illustrated by isothermal
contour lines, and the heavy line delineates the intersection of the liquidus surfaces for different phases—a
low-temperature valley called a boundary curve or cotectic. Slope on the boundary curve is indicated by the
arrowhead. Along such boundary curves, the liquid coexists with two solid phases, either crystallizing both
simultaneously (subtraction curve) or reacting with one
solid phase to produce the other (reaction curve). Thus,
subtraction and reaction boundary curves are equivalent
to eutectic and peritectic points, respectively, in binary
systems. We have included a more detailed analysis of
the various kinds of boundary curves and points in ternary phase diagrams in an accompanying box for interested readers.
FIG. 10.17. Representations for ternary systems. (a) The three
binary phase diagrams for the system albite-anorthite-diopside
(Ab-An-Di) oriented around a triangle. (b) These binary systems
can be stood upright, and the liquidus surface within the triangle
can be extrapolated from the binary systems and contoured, as
shown here. (c) A projection of the liquidus surface in (b) onto
the bottom triangle is shown in this figure. The arrowhead indicates the direction of slope for the boundary between the primary
phase fields of diopside and plagioclase.
SUMMARY
Phase diagrams provide a useful way of visualizing the
effects that intensive and extensive variables have on a
system. For binary systems, one of the most widely used
is the T-X2 diagram. We have seen how this diagram
can be derived from stacking Ḡ-X2 sections at various
temperatures, holding pressure constant. We have explored the characteristics of univariant equilibria in
these diagrams, including eutectic and peritectic points,
solid solution loops, and solvi. We also considered phase
diagrams in which the composition or partial pressure
of fluid phases were variables. In addition, we discussed
how univariant equilibria appear in P-T diagrams. The
various kinds of boundary curves and invariant points in
ternary liquidus phase diagrams have also been explored.
202
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
INTERPRETATION OF TERNARY LIQUIDUS PHASE DIAGRAMS
Ternary phase diagrams are divided into primary
phase fields by boundary curves. Each primary phase
field represents the first phase to crystallize upon intersecting the liquidus. As an example, we take the
phase diagram for the system anorthite-leucite-silica,
shown in figure 10.18a. Such phase diagrams can be
analyzed using Alkemade’s Theorem, which encompasses the following statements:
FIG. 10.18. (a) The phase diagram for the system anorthiteleucite-silica (CaAl2 Si2O8-KAl2Si2O8-SiO2 ) at P = 1 atm. Other
phases and their abbreviations are: orthoclase (Or), tridymite
(Tr), and cristobalite (Cr). Primary phase fields are identified
by abbreviations in italics. The boundary between the primary
phase fields for Tr and Cr is dashed to indicate that a phase
transition between these two polymorphs occurs at a temperature above the eutectic point. (b) The same diagram with the
addition of the Alkemade line between An and Or. Boundaries
between primary phase fields are identified as subtraction or
reaction curves by single- or double-headed arrows, respectively. (c) These curves may intersect to form three different
kinds of invariant points, as illustrated.
1. An Alkemade line is a straight line connecting the
compositions of two phases whose primary phase
fields share a common boundary curve. The first
step in the analysis of any ternary phase diagram
is to draw in all of the Alkemade lines. In figure
10.18a, most of the Alkemade lines already exist as
external boundaries to the triangle. For example,
the primary phase fields for leucite and anorthite
share a common boundary curve, and the appropriate Alkemade line joining their compositions is
the left side of the triangle. In this example, the
only additional Alkemade line we must draw is a
line joining the compositions of anorthite and orthoclase, as illustrated in figure 10.18b.
2. Alkemade lines divide a triangular diagram into
smaller triangles. The ultimate equilibrium crystallization product is defined by the corners of
the small Alkemade triangle, in which the original
liquid composition lies. For example, a liquid of
composition X in figure 10.18b will ultimately
crystallize to form leucite + orthoclase + anorthite.
If the original liquid lies on an Alkemade line, the
ultimate assemblage consists of a mixture of the
two phases at the ends of the line.
3. The intersection of an Alkemade line with its pertinent boundary curve represents the maximum
temperature along that boundary curve. For example, the intersection of the leucite-anorthite
boundary curve and Alkemade line occurs along
the left side of the triangle. This point represents the
highest temperature along that particular boundary curve, and temperature must fall off as one
follows the curve to the right.
4. If a boundary curve or its tangent intersects the
corresponding Alkemade line, that portion of the
boundary curve is a subtraction curve. In figure
10.18b, a tangent to the leucite-anorthite boundary
curve at any place along its length will intersect the
Alkemade line, so the entire curve is a subtraction
curve, along which the liquid crystallizes to form
two phases, leucite + anorthite. It is conventional
to identify subtraction curves with a single arrowhead pointing in the direction of decreasing temperature. If the boundary curve must be extended
Picturing Equilibria: Phase Diagrams
to intersect the pertinent Alkemade line, the curve
is still a subtraction curve and the point of intersection represents the thermal maximum. The
orthoclase-anorthite boundary curve in figure
10.18b must be extended to intersect its Alkemade
line, so it is a subtraction curve and temperature
decreases away from the Alkemade line, as shown
by the arrowhead.
5. If a boundary curve or its tangent does not intersect the appropriate Alkemade line, but merely an
extension of that line, then that portion of the
boundary curve is a reaction curve. For example,
a tangent drawn to any point along the leuciteorthoclase boundary curve in figure 10.18b does
not intersect the leucite-orthoclase Alkemade line,
but only an extension of that line. Thus, this boundary is a reaction curve along its entire length, as
indicated by double arrowheads. You might envision a hypothetical boundary curve that changes
from subtraction to reaction along its length as its
tangent swings outside the limits of its Alkemade
line. The intersection of the curve with the extended
Alkemade line is the maximum temperature along
the boundary curve.
Once the various kinds of boundary curves have
been identified using the rules above, we can recognize several types of invariant points formed by intersections of these curves, illustrated in figure 10.18c.
A eutectic point occurs where three boundary curves
form a “dead end,” and the liquid has no direction
of escape with further cooling. An example of a
eutectic in figure 10.18b is the intersection of the
orthoclase-anorthite, orthoclase-tridymite, and anorthite-tridymite subtraction curves. A tributary reaction point provides one outlet for liquid to follow once
the reaction that occurs at that point is completed. In
figure 10.18b, the intersection of the leucite-anorthite
and orthoclase-anorthite subtraction curves with the
leucite-orthoclase reaction curve produces this kind
of point. A distributary reaction point, shown in figure 10.18c, has two outlets and, after the reaction
that occurs at the point is complete, permits the liquid to follow either of two branches, depending on
what its final crystallization product must be (deter-
203
mined by the Alkemade triangle in which it resides).
There is no example of a distributary reaction point
in the anorthite-leucite-silica system.
We can now describe equilibrium crystallization
paths for any liquids in this system. We take two
examples, identified by X and Y in figure 10.18b. Liquid X lies in the Alkemade triangle leucite-anorthiteorthoclase, so those phases must be the final product.
These phases only coexist at one point in the diagram, the tributary reaction point, which must be
where crystallization will end. Upon intersecting the
liquidus, X first begins to precipitate leucite, as it lies
in the leucite primary phase field. The liquid composition moves directly away from the leucite corner as
leucite forms, finally intersecting the leucite-anorthite
subtraction curve. At this point, leucite + anorthite
crystallize simultaneously, and the liquid follows the
curve toward the tributary reaction point. The ratio
of leucite to anorthite in the crystallizing assemblage
can be found at any point along the subtraction curve
by noting the intersection of the tangent to the boundary curve with the Alkemade line and employing the
lever rule. When the liquid reaches the tributary reaction point, already crystallized leucite reacts with the
liquid to produce orthoclase. We know that the reaction cannot run to completion, because we must have
some leucite left in the final product. Therefore, we
must run out of liquid before all the leucite can react.
Liquid Y in figure 10.18b lies in the Alkemade
triangle anorthite-orthoclase-silica, so this must be
the final product. These phases only coexist at the eutectic point. Anorthite is the first phase to crystallize,
and the liquid composition moves from Y directly
away from the anorthite corner. Upon intersecting the
leucite-anorthite subtraction curve, the crystallization
sequence is joined by leucite. Both phases crystallize
as the liquid moves down the curve until it reaches
the tributary reaction point. Then leucite reacts with
the liquid to form orthoclase. In this case, the reaction runs to completion, and the liquid is then free
to move down the orthoclase-anorthite subtraction
curve, all the while crystallizing these two phases.
When it finally reaches the eutectic point, orthoclase
+ anorthite + tridymite continue to precipitate until
the liquid disappears.
204
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
At every opportunity, we have emphasized the relationships between phase diagrams and thermodynamics,
and we hope that this treatment will aid you in comprehending these relationships. So far, we have considered
only systems at equilibrium. The next chapter explores
nonequilibrium conditions.
suggested readings
Phase diagrams are usually introduced more rigorously in
igneous/metamorphic petrology texts than they are in geochemistry texts, although few relate these diagrams to thermodynamic functions. Some of the best explanations of phase
diagrams are noted below.
Edgar, A. D. 1973. Experimental Petrology, Basic Principles
and Techniques. London: Oxford University Press. (A techniques book that describes the equipment, procedures, and
pitfalls in experimentally determining phase diagrams.)
Ehlers, E. G. 1972. The Interpretation of Geological Phase
Diagrams. San Francisco: Freeman. (A complete and very
readable introduction to the subject that provides much
more detail than we do in this book; chapter 2 describes
binary phase diagrams and chapter 3 discusses ternary
systems.)
Ernst, W. G. 1976. Petrologic Phase Equilibria. San Francisco:
Freeman. (An exceptionally good reference on phase diagrams; chapter 2 describes the experimental determination
of phase equilibria, chapter 3 provides a look at the compu-
tational approach, and chapter 4 introduces unary, binary,
and ternary systems.)
Ferry, J. M., ed. 1982. Characterization of Metamorphism
through Mineral Equilibria. Reviews in Mineralogy 10.
Washington, D.C.: Mineralogical Society of America. (This
book contains numerous applications of phase diagrams to
metamorphic rocks; chapter 1 by J. B. Thompson describes
the relationship between thermodynamics and metamorphic
phase diagrams.)
Hess, P. C. 1989. Origins of Igneous Rocks. Cambridge: Harvard University Press. (Chapter 2 presents a thorough explanation of phase diagrams, although without reference to
thermodynamics.)
Morse, S. A. 1994. Basalts and Phase Diagrams. Malabar:
Krieger Publishing. (A thorough and very readable book
about phase diagrams. As in this book, Morse’s discussions
of binary diagrams end with Ḡ-X diagrams.)
Philpotts, A. R. 1990. Principles of Igneous and Metamorphic
Rocks. Englewood Cliffs: Prentice-Hall. (The relationship
between free energy and phase diagrams is introduced in
chapter 8 of this excellent text, and chapter 10 treats binary
and ternary phase diagrams.)
Ottonello, G. 1997. Principles of Geochemistry. New York:
Columbia University Press. (Chapter 2 of this thick book
provides one of the few systematic treatments of Ḡ-X diagrams that we are aware of in a geochemistry text.)
Yoder, H. S., Jr., ed. 1979. The Evolution of the Igneous Rocks,
Fiftieth Anniversary Perspectives. Princeton: Princeton University Press. (An updated version of Bowen’s treatise that
uses phase diagrams throughout. Chapter 4 by A. Muan gives
a thorough description of phase diagrams.)
Picturing Equilibria: Phase Diagrams
PROBLEMS
(10.1) Using figure 10.6, reconstruct the equilibrium crystallization sequence for a liquid with a composition that corresponds exactly to the eutectic in this system.
(10.2) Equilibrium melting is the reverse process of equilibrium crystallization. From figure 10.7, determine
the melting sequence for a mixture of 90 mol% orthoclase and 10% tridymite.
(10.3) The liquid in problem 10.2 is now cooled again from 1200°C to room temperature under equilibrium conditions. In this experiment, you continuously measure the temperature of the system over
the four hours it takes to reach room temperature. Sketch a qualitative plot of how temperature in
this system changes with elapsed time, and explain why temperature does not fall continuously with
time.
(10.4) Using figure 10.9, describe the melting under equilibrium conditions of plagioclase with composition
XAn = 0.2.
(10.5) Using the experimental results tabulated below, construct a T-XFa diagram for the system forsteritefayalite. The glass composition represents the liquid.
Bulk Composition (XFa)
T(°C)
Olivine Composition (XFa)
Glass Composition (XFa)
0.60
0.60
0.60
0.60
0.60
0.60
0.20
0.20
0.20
1800
1700
1600
1500
1400
1300
1800
1700
1600
—
—
0.34
0.42
0.58
0.60
0.08
0.17
0.20
0.60
0.60
0.60
0.75
0.91
—
0.25
0.47
—
(10.6) (a) Construct Ḡ-X2 diagrams for a system that is a regular solution with W = 15 kJ at 800, 900,
1000, 1100, and 1200 K. Then construct the corresponding T-X2 diagram for this system. (b) At
1000 K, what is the effect of changing W from 2 to 5 kcal? Show your work on a Ḡ-X2 diagram.
(10.7) Draw a series of topologically reasonable isothermal P-X2 sections that describe the phase relations
shown in figure 10.16a.
(10.8) Calculate a T-XCO2 diagram for the reaction in worked problem 10.2 at 1 atm for a CO2-H2O fluid
that behaves as an ideal gas.
(10.9) Using the T-XSiO2 diagram on the bottom of figure 10.7, determine the relative proportions of stable
phases at 900, 1100, and 1300°C for a bulk composition of XSiO2 = 0.3.
205
CHAPTER ELEVEN
KINETICS AND CRYSTALLIZATION
OVERVIEW
Not all geochemical processes make the final adjustment to equilibrium conditions. Even those processes
that eventually go to completion, such as crystallization
of magmas and recrystallization of metamorphic rocks,
are sometimes controlled by kinetic factors. In this
chapter, we discover what these factors are and how
they can be understood. We first examine the effect of
temperature on rate processes, which leads to the concept of activation energy. We then consider three specific rate processes: diffusion, nucleation, and growth.
Differences between volume diffusion and grain boundary diffusion are explained, and the importance of crystal defects is considered. We introduce nucleation by
discussing the extra free energy term related to surface
energy, and then consider rates for homogeneous and
heterogeneous nucleation. Crystal growth may be dominated by interface- or diffusion-controlled mechanisms,
so we discuss each briefly. Finally, we illustrate the
importance of determining the rate-limiting step with
several examples, as well as discuss several empirical
methods for estimating nucleation and growth rates without reference to kinetic theory.
206
EFFECT OF TEMPERATURE
ON KINETIC PROCESSES
You may not have been surprised to learn in chapter 5
that diagenetic processes are kinetically controlled. Most
of us are intuitively aware of how sluggish reactions are
at low temperature. It may not be so obvious that kinetics also plays a role in many geochemical processes
at high temperatures, such as crystallization of magma
or recrystallization of metamorphic rocks.
The rate r of a chemical reaction X → Y + Z can be
expressed as:
−dX/dt = k(X)n,
where X is the reactant’s concentration at any time t,
k is the rate constant, and the exponent n indicates the
order of the reaction. The minus sign denotes that the
rate decreases with time, as the reactant is used.
For rate-controlled processes in magmatic or metamorphic systems, it is critical that we take into account
the temperature dependence of rate constants. The relationship between any rate constant k and temperature
follows the classic equation proposed by Arrhenius more
than a century ago:
Kinetics and Crystallization
k = A exp(−∆Ḡ*/RT),
207
(11.1)
where A is the frequency factor, ∆Ḡ* is the activation
energy, R is the molar gas constant, and T is the absolute
temperature. Because a reaction requires the collision of
two molecules, the reaction rate is proportional to the
frequency of such collisions. The frequency factor, then,
indicates the number of times per second that atoms
are close enough to react. Not all collisions are comparable, though. A very gentle collision between molecules
may be insufficient to cause a reaction. The reaction rate,
therefore, is assumed to depend on the probability that
collisions will have energies greater than some threshold
value. The activation energy is an expression of this free
energy hurdle that must be overcome for the reaction
to proceed at an appreciable rate (fig. 11.1). The term
exp(−∆Ḡ*/RT ) expresses the fraction of reacting atoms
that have energy higher than the average energy of atoms
in the system and are thus most likely to participate in
chemical reactions. This probability term resembles the
Boltzmann distribution law of statistics, and is called
the Boltzmann factor. The negative exponential form of
the Boltzmann factor explains why this term must be considered in high-temperature systems: depending on the
size of ∆Ḡ*, rate constants may change by several orders
of magnitude over a temperature range of a few hundred
degrees.
FIG. 11.1. Representation of activation energy needed for the
reaction X → Y + Z. ∆G¯ is the free energy for the reaction under
equilibrium conditions, and ∆G¯* is the activation energy necessary to allow the reaction to proceed.
plot), we can graphically obtain values for the activation energy
and frequency factor.
A plot of Mathews’s data is shown in figure 11.2. The
intercept of the least squares regression line with the ln k axis
corresponds to a value for A of 1.2 × 1011. The slope of the
line corresponds to a ∆Ḡ* value of 44.5 kcal mol−1.
Unfortunately, there are too few experimental data
on reaction rates of geochemical interest, such as those
given in worked problem 11.1. One notable exception is
a 1963 study of the system MgO-SiO2-H2O by Hugh
Worked Problem 11.1
How can we determine activation energy and frequency factor
for a real geochemical system? As an example, consider the
reaction analcite + quartz → albite + water. Experimental data
for the rate of this reaction in NaCl at various temperatures
were reported by A. Mathews (1980):
k (kcal/RT)
10−4
9.02 ×
2.55 × 10−3
5.95 × 10−3
9.80 × 10−3
T (oC)
419
435
454
474
First, we recast equation 11.1 in a linear form by taking the
logarithm of both sides:
ln k = ln A − [∆Ḡ*R (1/T)].
(11.2)
This expression has the form of a straight line (y = b + mx), in
which the slope is −∆Ḡ* and the y intercept is ln A. Therefore,
by plotting ln k versus reciprocal temperature (an Arrhenius
FIG. 11.2. Arrhenius plot of ln k versus 1/T for the reaction
analcite + quartz → albite + water. The activation energy for the
reaction can be obtained from the slope of the regression line,
and the frequency factor is given by the y intercept.
208
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 11.3. Volume fractions of talc (Tc), anthophyllite (Ant),
enstatite (En), and quartz (Qz) as a function of time produced by
the talc dehydration reaction at 830°C and 1 kbar. (After Mueller
and Saxena 1977.)
Greenwood. Experiments show that at 830°C and
1 kbar, talc dehydrates to form enstatite plus quartz.
However, a multistep process involving the breakdown
of talc to form anthophyllite as a reactive intermediate is
kinetically more favorable than the direct transformation
of talc into enstatite plus quartz. Using Greenwood’s
measured rate constants, Robert Mueller and Surendra
Saxena (1977) calculated the proportional volumes of
talc, anthophyllite, enstatite, and quartz in a reacting
assemblage as functions of time. We summarize their
results in figure 11.3. Even though the stable assemblage
under these conditions is enstatite plus quartz, anthophyllite is the most abundant phase during the time interval from 1000 to 1500 minutes. Thus, experiments of
short duration might give the false impression that anthophyllite is the stable phase under these conditions.
DIFFUSION
Although diffusion may be easily thought of as a physical
phenomenon, it resembles a chemical reaction in being
governed by an activation energy. Components can migrate over free surfaces of grains (surface diffusion), along
the boundaries between grains (grain-boundary diffusion), and through the body of grains or liquids (volume
diffusion). The latter two styles of diffusion are of most
interest in high-temperature systems. When we consider
volume diffusion, it is also useful to distinguish between
self diffusion, by which we mean the one-directional
movement of similar atoms or ions (for example, when
oxygen diffuses through the oxygen framework of feldspar), and interdiffusion, in which the diffusion of one
atom or ion is dependent on the opposing diffusion of
another (for example, Mg-Fe exchange in olivine). Generally, the requirement of local charge balance implied
by interdiffusion results in a decrease in the rate of
diffusion.
In chapter 5, when we first discussed diffusion, we
presented a macroscopic or empirical view, in which
Fick’s Laws were used to describe the bulk transport of
material through a medium that was considered to be
continuous in all its properties. That treatment involved
the definition of diffusion coefficients (D), which come
in a bewildering array of forms. We can now add two
more. Self diffusion is commonly described by tracer
diffusion coefficients (D*), which are measured experimentally by following the progress of isotopically tagged
ions (for example, 18O moving through the oxygen framework of a silicate). Interdiffusion coefficients (D1–2), describing the interaction of ions of species 1 and species 2,
are measured empirically or can be calculated from the
tracer coefficients of each species, weighted by their relative abundance in the system:
D1–2 = (n1D1* + n2D2*)/(n1 + n2).
In geologic systems of even moderate complexity, it can
be difficult to calculate D1–2, because we first need to
collect a large amount of experimental data to estimate
tracer coefficients in systems that react sluggishly and
are hard to analyze.
Geochemists are adopting an alternate approach,
which considers diffusion from a mechanistic perspective at the atomic level and has been used with some
success by metallurgists and ceramists. It offers the
advantage of describing diffusive transport in terms of
forces between ions in a mineral lattice or other medium.
In theory, these can be predicted from conventional
bonding models, given detailed information about crystal structures. This approach holds great promise but
has been applied to very few materials of geologic significance so far. It is beyond the scope of this chapter,
although we can use the atomistic perspective to make
some broad generalizations to help visualize the behavior of diffusing ions.
Volume diffusion in solids usually involves migration
of atoms or ions through imperfections in the periodic
structures of crystals. Point defects can arise from the
absence of atoms at lattice sites (vacancies). Two kinds
Kinetics and Crystallization
FIG. 11.4. Volume diffusion is facilitated by defects in crystals.
(a) Frenkel defects result when ions leave normal crystal sites for
interstitial positions, leaving vacancies. (b) Schottky defects
occur when cation and anion vacancies preserve charge balance.
of intrinsic vacancies in ionic crystals maintain electrical
neutrality: Frenkel defects are caused by an ion abandoning a lattice site and occupying an interstitial position, and Schottky defects involve equal numbers of
vacancies in cation and anion positions. These imperfections are illustrated in figure 11.4. Impure crystals may
also contain extrinsic defects, which arise because foreign
ions may have a different charge from the native ions
they replace. The requirement for overall charge balance
leads to vacancies in the crystalline structure. For example, the incorporation of Fe3+ into olivine normally
requires a cation vacancy somewhere in the crystal.
Nonstoichiometric crystals, by definition, also have
vacancies. Real crystals commonly contain both intrinsic and extrinsic defects, each of which dominate the
overall diffusion at different temperatures. Intrinsic
mechanisms tend to be most effective at high temperatures, when thermal vibrations of atoms are rapid. As
temperatures are lowered, the number of vacancies induced by impurities becomes greater than those generated intrinsically. This can be seen in figure 11.5, an
Arrhenius plot of experimentally determined diffusion
coefficients for Fe-Mg along the c axis in olivine as functions of temperature. In the high-temperature region (left
side of fig. 11.5), diffusion occurs mostly by an intrinsic
mechanism, which has a steeper slope (larger ∆Ḡ*) than
for the extrinsic region. The transition between diffusion
mechanisms occurs at ∼1125°C. The difference in slope
tells us, qualitatively, that it takes more energy to move
ions between intrinsic defects than between extrinsic ones
(30 kcal mol−1 in the intrinsic region, 62 kcal mol−1 in the
extrinsic region, when extrapolated to pure forsterite).
209
Extrinsic diffusion mechanisms probably predominate in
natural systems, because most minerals contain many impurities, and temperatures for most geologic processes
are less than required for intrinsic mechanisms.
Extended defects also play important roles in volume
diffusion. These include dislocations, which are linear
displacements of lattice planes, and various kinds of
planar defects, such as twin boundaries and stacking
faults. Irregular microfractures also fall into this category. Extended defects provide very effective conduits
(sometimes called short circuits) for diffusion, and where
they are abundant, they may control this process. This
accounts in part for the high reactivity of strained versus
unstrained crystals. However, extended defects are difficult to model quantitatively.
Grain boundaries are sometimes considered to be
extended defects, but this blurs the distinction between
volume diffusion and grain-boundary diffusion. Most
geochemists believe that the most efficient path for diffusive mass transfer in metamorphic rocks is along grain
boundaries. This idea is grounded in the notion that
metamorphic fluids along such boundaries provide
faster diffusion paths than volume diffusion through
solid grains. Although this idea is probably correct in
many cases, the situation is not as straightfoward as it
may seem. The diffusion cross section offered by grain
FIG. 11.5. Arrhenius plot of ln D versus 1/T, summarizing data on
interdiffusion of Fe and Mg in olivines of different composition
(Buening and Buseck 1973). The kink in these lines at ∼1125°C
corresponds to a change from an intrinsic to an extrinsic diffusion
mechanism with different activation energy (and hence, slope).
210
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
boundaries is quite small, the tortuosity of these paths
may be appreciable, and the amount of fluid may decrease
considerably at high metamorphic grades. Attempts have
been made to place limits on grain-boundary diffusion
rates in metamorphic rocks, but there are few data with
which to test numerical models of this process.
Worked Problem 11.2
Can diffusion be used to constrain the rate at which a magma
cools? Larry Taylor and coworkers (1977) developed a “speedometer” applicable to basalts, using frozen diffusion profiles of
Fe/Mg in olivines. Let’s examine how this system works.
From the experimental data of Buening and Buseck (1973),
illustrated in figure 11.5, we can derive an expression for the
diffusion coefficient along the olivine c axis, Dc , for temperatures below 1125°C:
Dc = 102 fO1/6 exp(−0.0501 XFa − 14.03)
2
exp([−31.66 + 0.2191XFa]/RT),
(11.3)
where XFa is the mole fraction of Fe2SiO4 in olivine. Buening
and Buseck also showed that olivine diffusion is highly anisotropic, with transport being fastest in the c direction. The measured profiles of olivine grains, however, reflect reactions limited
by the slowest diffusive rate, which is along the a axis. Taylor
and coworkers estimated that Dc was greater than Da by a factor
of about four at 1050°C, so that:
Da = Dc /4.
One uncertainty in boundary conditions for this problem is
that the concentration profile in olivine as it solidified is unknown; that is, some compositional zoning may have been
present originally. Taylor and coworkers treated the forsterite
content of olivine as a step function, so that each grain originally had no compositional gradient other than a sharp step
between the magnesian core and iron-rich rim. Such a simplification means that the calculated cooling rate will be a minimum
rate. In a later paper, Taylor and coworkers (1978) corrected
this problem by estimating an “as-solidified” concentration profile. For simplicity, we ignore this later refinement.
In the case of olivine, the limiting diffusion coefficient Da is
a function of both temperature (which is, in turn, a function of
time t) and composition. We can use a one-dimensional diffusion equation to calculate the compositional profile:
{
}
∂C
∂
∂C
—– = —– Da [C(x), T(t)] —– ,
∂t
∂x
∂x
(11.4)
where C is the Fe concentration. Unlike the worked problem
at the beginning of chapter 5, which yielded easy, analytical
solutions, this equation is nonlinear and must be solved by numerical iteration to approach the correct solution.
FIG. 11.6. Calculated compositional profiles in olivine as a function of distance from the core-rim interface and cooling rate.
These profiles are produced by interdiffusion of Fe and Mg during
cooling. Circles are electron microprobe data for zoned olivines in
lunar basalt 15555. (After Taylor et al. 1977.)
Taylor and his collaborators applied their solution to determination of the cooling rates for lunar basalts. Oxygen fugacities
(which are temperature dependent) appropriate for lunar rocks
obey the relation:
log fO = 0.015T (K) − 34.6.
2
This relation provides the information we need to determine
Da as a function of fO for substitution into equation 11.4. The
2
measured compositional profile in an olivine grain in lunar
basalt 15555, obtained by electron microprobe traverse, is illustrated by circles in figure 11.6. Calculated diffusion profiles
for various cooling rates are shown for comparison. The minimum cooling rate for this basalt is ∼5°C/day.
NUCLEATION
It is a wonder that crystals ever form. The most difficult
step, from a thermodynamic perspective, is nucleation—
the initiation of a small volume of the product phase.
Crystal nuclei form because local thermal fluctuations
in the host phase result in temporary, generally unstable
clusters of atoms. This is an extremely difficult process
to study, because no experimental technique has yet been
devised that allows direct observation of the formation
of crystal nuclei.
The driving force for nucleation, like other rate processes, is the deviation of the system from equilibrium
conditions. A common expression for this driving force
is supersaturation, the difference between the concentration of the component of interest and its equilibrium
Kinetics and Crystallization
concentration. In studying nucleation in geochemical
systems, we often define deviation from equilibrium in
another way; that is, in terms of undercooling (∆T ). This
parameter is the difference between the equilibrium
temperature for the first appearance of a phase and the
temperature at which it actually appears. For example,
in magmas, crystallization commonly fails to occur at the
liquidus temperature of a melt, where it should be expected. The magma is therefore said to be undercooled
if it is at a temperature below its liquidus and crystallization is not yet initiated. The value of ∆T is equal to the
liquidus temperature minus the actual temperature.
( )
As we learned in earlier chapters, the appearance of
a new phase is associated with a decrease in the free energy of the system. This change can be expressed as a
total differential:
( )
∂G
dG = —–
∂T
P,ni
ticle is also minimized. Consequently, a growing particle
must perform work to stretch the interface. This work,
for a liquid droplet, is commonly defined in terms of a
tensional force, σ, applied perpendicular to any line on
the droplet surface. This concept of surface tension is only
strictly applicable to liquid surfaces, but no adequate alternative has been formulated for solids. It is common
practice to refer loosely to σ even when the nucleating
particle is a crystal.
Surface tension can be viewed as a measure of the
change in free energy due to an increase in particle surface area A; that is,
∂G
σ = —–
∂A
Nucleation in Melts
( )
∂G
dT + —–
∂P
T,ni
( )
∂G
dP + —–
∂ni
= dni ,
P,T,nj≠ni
in which the partial derivatives are written more familiarly as −S̄, V̄, and µi. In practice, we apply this equation
by declaring that each phase in the system consists of a
material whose thermodynamic properties are continuous. The boundary between one phase and another is
marked by a discontinuity in thermodynamic properties,
but we assume that it does not contribute any energy to
the system itself. Calculations performed on this basis are
generally consistent with observations of real systems.
We know, however, that the boundaries between
phases are not simple discontinuities. The atoms on the
surface of a droplet or a crystal are not surrounded by
the same uniform network of bonds found in its interior.
This irregular bonding environment leads to structural
distortions that must increase the free energy of the system. The surface free energy contribution is generally
quite small, because the number of atomic sites near the
boundary of a phase is much less than the number in its
interior. We can predict, however, that this contribution
increases in importance if the mean size of crystals or
droplets in the system is vanishingly small, as it is during
nucleation. Surface free energy, therefore, accounts for the
kinetic inhibition we have described as undercooling.
The extra energy associated with the formation of an
interface is minimized if the surface area of the new par-
211
T,P,n
.
The total differential of G thus can be modified to the
form:
( )
∂G
dG = (dG)vol + —–
∂A
T,P,n
dA
= (dG)vol + σdA,
in which (dG)vol is a shorthand notation for the familiar
temperature, pressure, and composition terms that apply
throughout the volume of the phase. For spherical nuclei
with radius r, this becomes:
4πr 3
∆Gtotal = —–— ∆Gv + 4πr 2σ,
3
(11.5)
where ∆Gv is the free energy change per unit volume.
The surface energy term dominates when the radius is
small, and the volume term dominates when it is large.
Because σ is always positive, equation 11.5 implies that
the total free energy of a nucleating particle increases
with r until some critical radius (rcrit) is reached, after
which the free energy decreases with further increase
in r. Small nuclei with r < rcrit are unstable and tend to
redissolve, whereas nuclei with r > rcrit persist and grow.
We can illustrate the importance of this observation with
the following problem.
Worked Problem 11.3
How does the free energy of a nucleus vary with radius? To
answer this, let’s consider the formation of spherical nuclei of
forsterite in Mg2SiO4 liquid. Assume an undercooling of 10°C
212
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
and a surface energy of 103 ergs cm−2 (= 2.39 × 10−5 cal cm−2).
Calculate the critical radius for olivine nuclei under these
conditions.
∆Gv for olivine at a temperature of 2063 K (10 K below the
forsterite liquidus) is −1.738 cal cm−3. Substitution of these
values into equation 11.5 gives:
4π– (−1.738)r 3 + 4π(2.39 × 10−5)r 2
∆Gtotal = —
3
the conditions necessary to form clusters of the critical radius
(equation 11.6) with equation 11.5 to give:
∆Ghomo = (16πσ3)/(3∆Gv2 ).
(11.7)
Also illustrated in figure 11.7 by dashed lines are energy
barriers to nucleation at ∆T = 0, 15, and 30°C. We can see that
this barrier, as well as the critical radius for nuclei, decreases as
undercooling increases.
= −7.276r 3 + 3.0 × 10−4r 2.
For r = 2 × 10−5 cm, ∆Gtotal is thus 6.18 × 10−14 cal. Substitution of other values for r gives a data set from which the solid
line in figure 11.7 is plotted. From this figure, we can readily see
that increasing r initially causes an increase in total free energy of
the system because the surface energy term dominates. At the
point where the volume energy term begins to dominate, there
is a downturn in total free energy, and further crystal growth
can take place spontaneously with a net decrease in free energy.
The peak of the free energy hump in this diagram corresponds to rcrit. We can calculate this value directly by recognizing that the reversal in slope corresponds to the point where the
first derivative of equation 11.5 with respect to r equals zero.
Solving for r under this condition yields:
rcrit = −2σ/∆Gv.
(11.6)
Substitution of the values given above for σ and ∆Gv gives rcrit
for olivine = 2.75 × 10−5 cm under these conditions.
The total energy barrier to nucleation ∆Ghomo (for homogeneous nucleation, defined below) can be found by combining
What we have presented so far is an apparent paradox. A particle is not stable until its radius is greater
than rcrit but it can never reach rcrit because it must always
begin as a nucleus too small to be stable. Worked problem 11.3, however, suggests one way to avoid this
paradox—by undercooling.
If nuclei persist once they reach critical radius, then
the nucleation rate can be expressed by an Arrheniustype equation as the product of the concentration of these
nuclei and the rate at which atoms attach themselves to
critical nuclei. The concentration of nuclei having the
critical radius (Nr ) at equilibrium is given by an expression of the Boltzmann distribution:
Nr = Nv exp(−∆Gtotal /RT ),
(11.8)
where Nv is the number of atoms per unit volume of
the reactant phase, and the exponential term is the
probability that they have sufficient energy to react. The
attachment frequency of atoms is the product of the
number of atoms (n) located next to the critical cluster,
the frequency (ν) with which atoms try to overcome the
barrier, and the probability (f ) of atoms having sufficient
energy to succeed. This is expressed by:
fattach = nν exp(−∆G*/RT),
(11.9)
where ∆G* is the activation energy needed for attachment. The forms of equations 11.8 and 11.9 are similar
to that of equation 11.1, the expression for the rate constant in terms of frequency factor and activation energy.
We will see equations of this form over and over again
in this chapter, because kinetic processes are governed by
statistical probability. Multiplying equations 11.8 and
11.9 gives the nucleation rate J:
J = nνNv exp(−∆Gtotal /RT )
exp(−∆G*/RT ).
FIG. 11.7. Calculated total free energy for nucleation of forsterite
in a liquid of its own composition, as a function of nucleus radius
r and undercooling ∆T. The critical nucleus at each undercooling
is rcrit.
(11.10)
For nuclei that form free in a liquid, ∆Gtotal in this equation is identical to ∆Ghomo in equation 11.7. We can
make the further approximation that ∆Gv = −∆S∆T, where
Kinetics and Crystallization
213
∆S is the entropy change associated with nucleation and
∆T is the undercooling. By combining equations 11.7
and 11.10, therefore, we find that:
(
)
16
J = nνNv exp —– πσ3∆S 2∆T 2RT exp(−∆G*/RT ).
3
(11.11)
We can use equation 11.11 to make some generalizations about nucleation rate. Not surprisingly, undercooling exerts a major influence, and there is a rapid rise
in nucleation rate as undercooling increases; that is, as the
system moves farther away from equilibrium. As ∆T increases and T decreases, the two exponential terms compete for dominance, with the result that the nucleation
rate reaches a maximum at some critical undercooling
and then decreases thereafter. In physical terms, the free
energy advantage gained by undercooling is finally overwhelmed by the increasing viscosity of the liquid. If we
were to plot J against ∆T, the result would be an asymmetrical bell-shaped curve.
Undercooling, then, reduces the kinetic inhibition to
nucleation and offers a solution to the paradox. Nature
provides another, more common solution as well. Equations 11.5–11.7 and equation 11.11 are written in terms
of a free-standing spherical particle. This geometrical condition is known as homogeneous nucleation. Observation tells us, however, that nucleation more commonly
takes place on a substrate. Frost forms on a windowpane,
dew on spiderwebs, and crystals on other crystals. There
must be an energetic advantage to this style of heterogeneous nucleation that makes it so widespread.
To see how nucleation on a substrate helps avoid the
problem of a critical radius, consider figure 11.8. Here,
a particle of a phase C with an exceedingly small volume
has been allowed to form from liquid phase L against a
foreign surface S. Note that the radius of this particle, if
it had condensed homogeneously, would have been <rcrit.
Here, however, the same volume of phase C forms a
spherical cap with an effective radius that is considerably larger than rcrit. Because the surface tension on the
particle is a function of its degree of curvature, this nucleus now has a low surface free energy and is stable.
If the interfaces between phases C, L, and S are all
mechanically stable, then the tensional forces perpendicular to any line along the phase interfaces must be perfectly balanced. This must also be true at the three-phase
contact (point P), where there are three force vectors to
resolve. Each is proportional to a surface tension in the
FIG. 11.8. Schematic cross section through a cluster of atoms
of phase C nucleating heterogeneously from phase L on a substrate S. The contact angle between the surface of the nucleus
and S is θ. The effective radius of this nucleus is much greater
than the critical radius of a spherical nucleus of the same volume
formed by homogeneous nucleation. Distances x and h are used in
problem 11.5 at the end of this chapter.
system. σLS and σCS oppose each other along the flat
boundaries L-S and C-S, respectively. The remaining vector, represented by σCL, is tangential to the curved C-L
interface at point P. Because the sum of these three must
be zero, we see that:
σLS = σCS + σCL cos θ,
(11.12)
where θ is the angle between the tangent and interface
C-S.
From equation 11.12, it is apparent that nucleation
of a new particle is particularly favorable if σCL is nearly
equal to σLS, and σCS is very small. This condition might
easily be met if the substrate were a seed crystal of the
nucleating phase itself. Under these conditions, θ is minimized, and the effective radius of the nucleus is at its
largest.
By combining equation 11.12 with appropriate expressions for the surface area and volume of the spherical cap in figure 11.8, we can express the energy barrier
to nucleation ∆Ghetero as:
(
)
4π 3
∆Ghetero = —– σ CL
∆Gv2 (2 − 3cos θ + cos3 θ).
3
(11.13)
Comparing this to equation 11.7, which describes the
total energy barrier to homogeneous nucleation, we see
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
that by nucleating against a substrate, the barrier has been
reduced by a factor of (2 − 3cos θ + cos3 θ)/4. Clearly,
then, nucleation against foreign objects is favored over
nucleation of unsupported particles.
The rate of heterogeneous nucleation can be expressed
by an equation similar to equation 11.11, with appropriate energy values substituted. We have not spent much
time on these rate equations, for the simple reason that
they do not appear to work very well for geologic systems. Theoretical models for nucleation qualitatively
reproduce the general shapes of nucleation curves, but
the temperatures for the maximum nucleation rate and
peak widths are commonly different from experimentally
determined curves. This difference between theoretical
and experimental curves might be due to heterogeneous
nucleation, but some researchers think that classic nucleation theory is flawed. Therefore, these equations may
require significant revision before they can be applied
successfully to understand nucleation in silicate melts.
Nucleation in Solids
The principles we have just discussed are also applicable to solid phases, but the process of nucleation in
solids is more complicated. A phase transformation in
the solid state may involve a significant volume change
that cannot be accommodated by flow of the reactant
phase. This produces a “room problem” that induces
strain energy into the system. This extra energy must
be added to the energy barrier to nucleation. The free
energy for nucleation thus becomes:
∆Gtotal = 4π/3r 3(∆Gv + ∆Gs ) + 4πr 2σ,
where ∆Gs is the strain energy.
Nucleation in metamorphic rocks is probably always heterogeneous. Grain boundaries, dislocations, and
strained areas provide favorable nucleation sites. The
formation of a nucleus at a boundary or dislocation involves the destruction of part of an existing surface,
thereby lowering the free energy of the system. Nucleation at grain corners and triple points is also energetically favored.
An interesting example of solid-state nucleation is provided by exsolution, which commonly occurs in slowly
cooled pyroxenes and feldspars. The upper part of figure 11.9 shows the experimentally determined solvus
in the system NaAlSi3O8-KAlSi3O8. In chapter 10, we
showed that subsolidus cooling of a homogeneous alkali
So
lvu
T
s
Spinod
al
214
An
Or
x3
G
b1
x1 x2
x4
b2
a1
a2
Ab
A1
B1
B2
A2
Or
XOr
FIG. 11.9. The top shows a T-XOr diagram of the subsolidus portion of the alkali feldspar system showing locations of the solvus
and spinodal, as determined by Waldbaum and Thompson (1969).
Below that is a schematic Ḡ -XOr diagram for the alkali feldspar
system representing some temperature below the crest of the
solvus. Tangent points a1 and a2 define the locations of solvus
limbs, and inflection points b1 and b2 define spinodal limbs at
this temperature.
feldspar below the solvus will result in its separation into
two phases with compositions defined by the limbs of
the solvus. The exsolved phases nucleate, homogeneously
or heterogeneously, in the solid state. The lower part of
figure 11.9 shows a schematic Ḡ−XOr diagram for some
temperature below the crest of the alkali feldspar solvus.
Points a1 and a2 define the locations of the limbs of the
solvus at this particular temperature, corresponding to
phase compositions A1 and A 2.
Another style of nucleation, which may be kinetically
more favorable, results in the formation of umixed feldspars with compositions that are metastable relative to
A1 and A2. To show how this occurs, we have illustrated
another curve, called the spinodal, in the upper part of
figure 11.9. Inside the limits of the solvus, the free energy curve at the fixed temperature in the lower diagram
has two inflections (∂2Ḡ/∂X 2 = 0), labeled b1 and b2,
which define the positions of the spinodal limbs. If any
feldspar having an overall composition between B1 and
B2 contains small local fluctuations in composition (as,
Kinetics and Crystallization
215
INTERNAL PRESSURE, BEER BUBBLES, AND PELE’S TEARS
If you have ever opened a shaken can of beer or soda
pop, you have experienced the fountain effect that
follows rapid pressure release. Why does this happen? If not shaken, the same can emits only a gentle
hiss when it is opened. To answer this question, we
need to explore the relationship between surface
tension and pressure.
The work required to transfer a small volume dV
of material from liquid to the inside of a bubble,
(Pint − Pext)dV, must equal the work required to extend
its surface, σdA:
(Pint − Pext)dV = σdA.
Differentiating the expressions for the volume and
area of a sphere gives:
dV = 4πr 2dr
and
dA = 8πrdr.
Substituting these values into the above equation results in:
(Pint − Pext) 4πr 2dr = σ8πrdr,
Pint = 1 + (2 × 350)/(10−3 × 106 ) = 1.7 bars.
or
Pint − Pext = 2σ/r.
Thus, the internal pressure of a vapor bubble must be
greater than external pressure by an amount 2σ/r.
In an unshaken can of beer, all of the vapor phase
comprises a large “bubble” at the top. Its interface
with the liquid is flat, so r in equation 11.14 is infinitely large. Under these circumstances, Pint = Pext. If
the can is shaken, however, the same volume of vapor
is distributed among a very large number of small
bubbles. The internal pressure on each is significantly
higher than the external pressure because r is a fraction of a millimeter. It is this vapor pressure that causes
the explosive fountain.
The actual pressure, of course, depends on the
magnitude of σ as well as the radii of the vapor bubbles. It can be particularly high in some systems of
geochemical interest. For example, the surface tension
of basaltic magma is ∼350 dynes cm−1 at 1200°C and
1 bar. Highly fluidized basaltic ejecta sometimes form
tiny glassy droplets and filaments called Pele’s tears.
The internal pressure of such droplets, which have
radii on the order of 10−3 cm, can be calculated using
equation 11.13:
(11.14)
for example, x1 and x2), its free energy will be less than
that for a homogeneous feldspar. This situation also results in unmixing, but in this case, the process is called
spinodal decomposition. Within the limits of the spinodal (b1 to b2), coexisting compositions such as x1 and
x2 are stable relative to a homogeneous phase. Outside
the spinodal, a homogeneous phase is stable with respect
to unmixed phases such as x3 and x4. The energy difference between the homogeneous and exsolved phases is
very small in either case, but unmixing is only favored
inside the spinodal. Because the compositional difference between x1 and x2 is small, the structural mismatch
between them is small and the strain energy is minimal.
Therefore, spinodal decomposition involves no major
nucleation hurdle and the interface between the unmixed
phases is diffuse. In this way, exsolution lamellae with
The internal pressure is nearly double the external
(atmospheric) pressure!
compositions between A1 and A2 can form metastably.
Feldspars with bulk compositions that lie between the
solvus and the spinodal (that is, between A1 and B1 or A2
and B2) can unmix only by solvus-controlled exsolution.
In this case, the unmixed compositions are relatively far
apart, requiring the accommodation of significant
structural mismatch, so there is a high kinetic barrier to
nucleation. For this reason, nucleation may be sluggish,
and exsolution phase boundaries are likely to be sharp.
The kinetics of exsolution processes can be rather complex, and it is convenient to summarize such behavior
with the aid of a time-temperature-transformation (sometimes called TTT) plot, as illustrated in figure 11.10.
These plots provide a synthesis of kinetic data obtained
from experiments carried out at different temperatures
and cooling rates. The two sets of curves in figure 11.10
216
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Slow
Moderate
Homogeneous
Nucleation
Fast
T
Spinodal
Decomposition
log Time
FIG. 11.10. Time-temperature-transformation (TTT) plot illustrating how various cooling rates might control whether homogeneous
nucleation or spinodal decomposition occurs. Spinodal decomposition takes place in more rapidly cooled systems because nucleation
is not required.
correspond approximately to the beginning (1% product)
and completion (99% product) of exsolution. Completion of exsolution occurs when equilibrium is reached.
The behavior of a mineral under various thermal conditions can be determined using cooling curves such as
those illustrated. For example, at slow cooling rates,
homogeneous nucleation is favored over spinodal decomposition, whereas the reverse is true for moderate
cooling rates. During fast cooling, even spinodal decomposition will be incomplete, and quenched samples show
no unmixing at all.
GROWTH
Crystal growth is a complex activity involving a number
of processes: (1) chemical reaction at the interface between a growing crystal and its surroundings; (2) diffusion of components to and from the interface; (3) removal
of any latent heat of crystallization generated at the
interface; and (4) flow of the surroundings to make
room for the growing crystal. For most silicate minerals,
only processes (1) and (2) are thought to be significant
factors. These lead, then, to two end-member situations.
When diffusion of components in the surroundings is
fast relative to their attachment to the crystal nucleus,
the rate of growth is controlled by the interface reaction.
Conversely, when attachment of components to the crystal nucleus is faster than transport of components to it,
the rate of growth is controlled by diffusion. For the
special case in which crystals and surroundings have the
FIG. 11.11. During diffusion-controlled growth, compositional
gradients are established in the medium surrounding a growing
crystal. Components rejected by the crystal cannot diffuse away
fast enough, and components used in making the crystal cannot
diffuse to the interface fast enough, so that gradients result.
same composition, no diffusion is necessary and growth
is interface-controlled, although removal of latent heat of
crystallization may be important in this situation. Buildup
of latent heat may also be important where two growing
crystals converge. Interface-controlled growth may dominate in igneous systems and diffusion-controlled growth
is likely to be more important in metamorphic systems,
but intermediate situations in which both processes play
a role are also common in both environments.
In diffusion-controlled growth, migration of components to and from the area immediately surrounding
the growing crystal cannot keep pace with their uptake
or rejection at the interface, so that concentration gradients are formed, as illustrated in figure 11.11. The
formation of these gradients obviously slows down the
rate of growth. Fluid advection, if it occurs, tends to destroy the gradients and causes growth rate to increase. In
interface-controlled growth, concentrations of components in the zone around the growing crystal are similar
to those of the bulk surroundings, because the limitation
on growth is not movement through the surroundings.
As a consequence, growth rate under these conditions is
unaffected by fluid movement.
Interface-Controlled Growth
The rate at which interface-controlled growth takes
place is the difference between the rates at which atoms
are attached to and detached from the crystal. The rate of
attachment, ra, can be expressed by a probability relationship similar in form to equations we have already seen:
Kinetics and Crystallization
ra = ν exp(−∆G*/RT ),
where ν is the frequency at which the atoms vibrate and
∆G* is the activation energy for attachment. Similarly,
the rate of detachment, rd, is given by:
rd = ν exp(−[∆G′ + ∆G*]/RT ),
where ∆G′ is the thermodynamic driving force. The
growth rate, Y, is (ra − rd) times the thickness per layer, a0,
and the fraction of surface sites to which atoms may successfully attach, f:
Y = a0 fν exp(−∆G*/RT)[1 − exp(−∆G′/RT)].
(11.15)
Equation 11.15 exhibits the same qualitative variation with undercooling as that for nucleation rate, equation 11.11. At zero undercooling, the rate is zero because
∆G′ is zero. As the system cools below the equilibrium
temperature, growth rate increases because ∆G′ increases.
However, at large degrees of undercooling, the growth
rate begins to decrease because ∆G* becomes significant,
reflecting the fact that atoms are less mobile as fluid viscosity increases. Therefore, with increasing undercooling, the growth rate first increases, then passes through
a maximum, and finally decreases, just as nucleation rate
does.
Chemists recognize several mechanisms for interfacecontrolled growth, each dominant in a different growth
environment with different proportions of surface sites
available for growth (f in equation 11.15). Continuous
models dominate where the growth interface is rough
and atoms can attach virtually anywhere. Consequently,
217
f = 1 for all undercoolings and growth is relatively fast.
In layer-spreading models, the interface is flat except at
steps, which provide the only locations where atoms
can attach. Thus, f < 1. Two commonly observed layerspreading mechanisms are illustrated in figure 11.12. Surface nucleation forms one-atom-thick layers that spread
laterally over the interface by adding atoms at the edges.
If screw dislocations occur on the interface, atoms are
attached to form spiral-shaped steps. One way to recognize these growth mechanisms is illustrated in worked
problem 11.4.
Worked Problem 11.4
Jim Kirkpatrick and coworkers (1976) measured the growth rate
of diopside from diopside melt by making movies of experiments
in which a seed crystal was introduced into melt at various
undercoolings. Using their data, presented in figure 11.13a, can
we deduce the growth mechanism for diopside?
Because diopside is growing from a melt of its own composition, we know that the growth rate will not be diffusioncontrolled. The interface-controlled growth rate expression is
difficult to solve in the form of equation 11.15 because we do
not have appropriate values for ∆G and ∆G′. A theoretical
alternative, known as the Stokes-Einstein relation, however,
allows us to replace the exponential terms in equation 11.15
with a viscosity (η) term. We do not produce the mathematical
derivation for this step, because it is rather complex, but the
step should be intuitively reasonable. For any fixed degree of
undercooling, the growth of crystals from a low-viscosity melt
(low ∆G and ∆G′) should be more rapid than from a highviscosity melt. Thus, we should be able to define a term Yη/T
which is, like f, a measure of available attachment sites on the
crystal surface. A plot of Yη/T versus undercooling can be used
to determine the growth mechanism. This yields a horizontal line
FIG. 11.12. Diagrams of layer-spreading mechanisms for crystal growth. (a) In surface nucleation, atoms are added only at the edges of a
spreading layer of atoms. (b) In screw dislocations, atoms attached to the crystal surface form spiral stairs.
218
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Worked Problem 11.5
Crystallization under plutonic conditions may occur at a very
small degree of undercooling relative to volcanic rocks. How
long will it take to grow a 1-cm diopside crystal at an undercooling of only 1°C?
The only growth rate data on diopside are those we presented in worked problem 11.3. It is not possible to measure the
growth rate at small degrees of undercooling accurately, but we
can extrapolate a rate from experimental data at larger undercoolings. The straight line passing through the point (0, 0) in
figure 11.13b allows us to estimate Y at ∆T = 1°. The equation
for this line is:
Yη/∆T = 6.6 × 10−5 T.
FIG. 11.13. Experimental data on the growth of diopside from
melt of the same composition. (a) Relationship between growth
rate Y and undercooling ∆T. (b) Calculated Yη/∆T versus undercooling for data in (a), using viscosity measurements of Kirkpatrick (1974). The linear trend with positive slope indicates a
screw dislocation growth mechanism. (After Kirkpatrick et al.
1976.)
for continuous growth, a straight line with positive slope for
growth by a screw dislocation mechanism, and a curve with positive curvature for growth by a surface nucleation mechanism.
The first step is to determine the parameter Yη/∆T for each
data point. Kirkpatrick (1974) tabulated viscosity data for liquid diopside at various degrees of undercooling. For example,
at an undercooling of 50°C, η equals 19 poise (g cm−1 sec−1).
Substituting this value and the measured growth rate of 9 × 10−3
cm sec−1 at this undercooling,
Yη/∆T = 3.4 × 10−3 g sec−2 deg−1.
A plot of similarly calculated values for all the data points in
figure 11.13a against ∆T is presented in figure 11.13b. These
data define a nearly straight line with positive slope, so diopside
must grow by a screw dislocation mechanism.
Therefore, at ∆T = 1°, Yη = 6.6 × 10−5 g sec−2. Because η =
15 poise at 1° undercooling (Kirkpatrick 1974), Y is on the
order of 10−6 cm sec−1. At this rate, it would take 106 sec, ∼10
days, to grow a 1-cm crystal.
For a number of reasons, this calculated rate must be higher
than the growth of diopside in a real plutonic system. The experimental temperature (1390°C) is considerably higher than
the liquidus temperatures of most magmas, and the viscosity of
the experimental system is lower. Also, diffusion might play a
role in a real, multicomponent magma.
Diffusion-Controlled Growth
Formulation of an equation for the rate of diffusioncontrolled growth must start with either Fick’s First
or Second Law, depending on whether a steady state is
reached. The growth rate is controlled by how rapidly
atoms can diffuse to the interface. Consequently, growth
rate decreases as time passes, because compositional gradients such as those shown in figure 11.11 extend farther
into the surroundings with time. The expressions for
growth rate under these conditions involve combining a
diffusion equation with a mass balance equation, because
the flux of atoms reaching the interface by diffusion
must equal the atoms attached to the interface. Derivation of such expressions is rather complex and is not
attempted here. Equations for growth rate Y under
diffusion-controlled conditions take the general form:
Y = k(D/t)1/2,
We can apply this formulation to other types of problems as well. For example, let’s consider the following
problem.
(11.16)
where k is a constant involving concentration terms, D
is the diffusion coefficient of the slowest diffusing species,
and t is time.
Kinetics and Crystallization
219
FIG. 11.14. Photomicrographs of (a) skeletal and (b) dendritic plagioclase crystals formed in rapidly cooled experimental melts. Sluggish
diffusion at large undercoolings causes breakdown of planar crystal surfaces. (Courtesy of G. Lofgren.)
structures of snowflakes, which grow rapidly at high
degrees of undercooling.
SOME APPLICATIONS OF KINETICS
Many geochemical systems reflect the interplay of several kinetic processes acting concurrently. In such cases,
understanding reaction rates requires identifying the
slowest step in the process, known as the rate-limiting
step. If we can do this, it may be possible to model the
1300
Liquidus
1100
T (oC)
The compositional or thermal gradients produced by
diffusion can affect growth by causing crystal faces to
break down. When a faceted crystal becomes unstable,
protuberances begin to develop. Heat and components
rejected during crystal growth flow away from these
protuberances, hindering the development of additional
structures in their vicinity. This can produce skeletal and
dendritic crystals, such as those in figure 11.14.
With increasing undercooling, D decreases and Y increases, so the ratio D/Y decreases. Gary Lofgren (1974)
illustrated by experiment how increasing undercooling
(that is, decreasing D/Y) controls crystal morphologies.
He studied the shapes of plagioclase crystals growing at
5 kbar in plagioclase melts at various degrees of undercooling, as shown in figure 11.15. At nearly constant
temperature intervals below the liquidus, various plagioclase compositions crystallized to form tabular, skeletal,
dendritic, and spherulitic morphologies. The onset of
diffusion-controlled growth in this system (that is, the
condition under which faceted crystals become unstable)
is defined approximately by the boundary separating
tabular and skeletal crystals in figure 11.15. The periodicity of protuberances increases as D/Y decreases. The
skeletal morphologies are instabilities with protuberances
spaced far apart. With increasing undercooling, these give
way to finer dendritic structures spaced closer together.
Similar morphological sequences are observed in olivine
crystals in the interior and exterior parts of pillow lavas.
Even more familiar, of course, are the fine, feathery
Solidus
ar
bul
Ta tal
ele
Sk itic
d
n r
De tic
uli
her
Sp
900
700
500
0
20
40
60
80
100
Wt.% An
FIG. 11.15. T-XAn diagram for the system albite-anorthite,
summarizing how plagioclase crystal morphologies change with
undercooling. Skeletal and dendritic morphologies are illustrated
in figure11.14. (After Lofgren 1974.)
220
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
kinetics of crystallization, as illustrated in two examples
below. However, the application of kinetic theory to experimental systems with controlled rates of nucleation
and growth has often produced disappointing results.
Impure crystals and other complexities can frustrate
even the most determined theorist. Empirical methods
that bypass theory but still provide information on nucleation and growth rates have been developed. We also
consider several of these.
→ Calcite: Growth as the
Aragonite ←
Rate-Limiting Step
→ calThe polymorphic transformation aragonite ←
cite has been studied extensively. Calcite is the thermodynamically stable phase under all conditions at the
Earth’s surface, as we demonstrated in worked problem 9.5. Although aragonite is only stable at high pressures, it is sometimes exposed on the surface, where
high-pressure metamorphic rocks have been uplifted. If
we understood the kinetic pathway by which aragonite
inverts to calcite, we could constrain the rates and conditions under which such terrains were uplifted.
Metastable aragonite in shells is readily replaced by
calcite during diagenesis, so the preservation of metamorphic aragonite must imply unusual conditions. Hydrothermal experiments indicate that at temperatures as
low as 200o C, there is significant replacement of aragonite in only a few days, although the reaction is slower
in the absence of water. In these experiments, however,
aragonite was dissolved and reprecipitated as new grains
of calcite rather than being replaced. In contrast, petrologists observe that metamorphic aragonite has often been
partly converted to calcite, indicating that the calcite replaced aragonite in situ. We can then conclude that metamorphic rocks containing aragonite cannot have had a
pore fluid present when they were uplifted.
Experiments to study the replacement reaction under
dry conditions were carried out by William Carlson and
John Rosenfeld (1981). They concluded that nucleation
of calcite was not the rate-limiting step, because partially
replaced aragonite is commonly surrounded by numerous
small calcite grains. Consequently, they set about to
measure calcite growth rates by observing the grains on
a heating stage attached to a microscope. An Arrhenius
plot of the calcite growth rate (in different crystallographic directions) versus temperature is shown in figure 11.16. Extrapolating these growth rates to lower
FIG. 11.16. Arrhenius plot showing the rate of growth of calcite
crystals replacing aragonite, as a function of crystallographic
orientation and temperature. (After Carlson and Rosenfeld 1981.)
temperatures allows inferences to be made about the
transformation of aragonite to calcite under natural conditions. At 250o C, calcite replaces aragonite at a rate
of 1 m/106 yr, so even large aragonite crystals would be
completely transformed. However, at 100o C, calcite
grows at only 10−3 mm/106 yr, allowing aragonite to
persist over geologic time periods. This work clearly
demonstrates that the uplifted rocks must have been cool
and dry, as well as illustrating the importance of temperature in controlling reaction rates.
Iron Meteorites: Diffusion as the
Rate-Limiting Step
Iron meteorites are thought to be samples of the differentiated cores of small bodies (asteroids). This interpretation derives from the rate of exsolution in metal
alloys, in which interdiffusion of Fe and Ni is the ratelimiting step.
The subsolidus phase diagram for Fe-Ni is illustrated
in figure 11.17, in which a two-phase region separates the
fields of kamacite and taenite (both iron-nickel phases).
The lines that bound this region are actually limbs of a
solvus, like that encountered in the alkali feldspar system in chapter 10, although this one has an unusual
shape. On cooling, an initially homogeneous alloy X
containing 10% Ni encounters the solvus at 680o C,
unmixing to form plates of kamacite within taenite.
Exsolution continues as temperature decreases, and the
Kinetics and Crystallization
221
FIG. 11.17. (a) Subsolidus phase diagram for the Fe-Ni system, showing the solvus for unmixing of
kamacite and taenite. (b) Calculated Ni diffusion profiles in taenite, as a function of cooling rate.
compositions of the coexisting kamacite and taenite at
any temperature are given by intersections of an isothermal line with the solvus limbs. Because both solvus limbs
slope in the same direction (fig. 11.17a), both phases become more nickel-rich with decreasing temperature. The
only way this is possible is if the proportion of kamacite
(the low-Ni phase) increases at the expense of high-Ni
taenite. Thus, the kamacite plates become thicker with
decreasing temperature.
Exsolution is made possible by diffusion of Fe and Ni
in the metal alloys. Measurements show that diffusion
is slower in taenite than in kamacite. Because nickel is
expelled from kamacite more rapidly than it can diffuse
into the interior of adjacent taenite grains, the nickel
distribution in taenite develops an “M”-shaped profile
(fig. 11.17b). The more rapidly the metal is cooled, the
more pronounced is the dip in the nickel diffusion profile. Calculations based on Fick’s Second Law, using diffusion rates determined from experiments, give nickel
profiles at various cooling rates (fig. 11.17b).
Iron meteorites commonly exhibit intergrowths of
kamacite and taenite (called a Widmanstätten pattern and
pictured in fig. 11.18). A nickel profile in meteoritic taenite analyzed by electron microprobe can be compared
with calculated nickel diffusion profiles for taenite plates
having the same width to estimate the cooling rate of
the iron meteorite. These estimates, generally a few tens
of degrees per million years, correspond to cooling rates
222
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Bypassing Theory: Controlled Cooling
Rate Experiments
order of appearance of plagioclase and ilmenite reverses.
Walker and coworkers inferred from the texture of the
rock itself that ilmenite began to form before plagioclase,
suggesting that the controlled cooling rate experiments
may be more relevant than equilibrium crystallization
experiments to understanding the cooling history of this
basalt. A second feature seen in figure 11.19 is that all
phases begin crystallizing below their liquidus temperatures in the programmed cooling experiments, and the
amount of temperature suppression increases with cooling rate. Another conclusion, not apparent from this
figure, is that the compositions of pyroxenes, in terms
of both major and minor elements, are distinct in the
equilibrium and controlled cooling rate experiments.
Based on order of phase appearance, texture, and mineral compositions, Walker and coworkers estimated the
actual cooling rate for this lunar rock decreased from
∼1°C/hr when olivine first crystallized to ∼0.2°C/hr during crystallization of pyroxene. Decreasing cooling rate
is consistent with its location near the margin of a lava
flow.
Programmed cooling experiments provide an empirical method by which to assess the mineralogical and textural features that characterize rapidly cooling igneous
rocks and to quantify the physical conditions that produced them.
The challenge of unraveling kinetic controls on
complex geologic systems may be more amenable to experiment than to calculation. Early in this chapter, we
summarized some experimental results on the rate of
dehydration of talc. There are very few kinetic data for
such metamorphic systems. Experiments in which magma
compositions are cooled and solidified at predetermined
rates are more common.
An example is a programmed cooling rate study of
a lunar basalt by David Walker and others (1976), the
results of which are summarized in figure 11.19. Experimental charges were cooled at different rates and
quenched from different temperatures within the crystallization range to construct the liquidus lines shown
in the diagram. Liquidus temperatures from equilibrium
experiments are illustrated by arrows on the left side of
the figure.
Several interesting features emerge from these data.
First, the crystallization sequence is different under equilibrium and dynamic cooling conditions; specifically, the
FIG. 11.19. Results of controlled cooling rate experiments on
lunar basalt 12002 by Walker et al. (1976). Equilibrium liquidi for
various phases are shown by arrows on the left-hand side. The
curves are contours of the first appearance of various phases at
different cooling rates.
FIG. 11.18. Etched surface of the Mount Edith (Australia) iron
meteorite, showing Widmanstätten texture formed by exsolution
of kamacite and taenite. Dark blobs are sulfide (troilite).
(Courtesy of the Smithsonian Institution.)
expected at the centers of asteroids with diameters
<200 km. The interiors of larger bodies would cool more
slowly, because rocks provide good insulation.
Kinetics and Crystallization
223
Bypassing Theory Again:
Crystal Size Distributions
Crystal size distribution (CSD) analysis provides a
way to assess the kinetics of crystallization in rocks. The
method depends on a population balance equation that
describes the change in numbers and sizes of crystals as
they nucleate and grow in either a liquid or solid environment. This relationship is expressed as:
n = n0 exp(−L/Yt).
(11.17)
In this equation, n is the crystal population density (the
number of crystals of a given size class per unit volume),
and n0 is a constant that represents the population density of crystal nuclei. The specific size range for crystals
is L, Y is the linear growth rate (the growth rate along
one crystallographic direction), and t is the average residence time for crystals in the system. A plot of ln n
versus L is a straight line with slope −1/Yt and intercept
n0. If either the growth time or the growth rate is known,
the other may be determined. The nucleation rate J can
be calculated by recognizing that:
J = n0 Y.
(11.18)
In practice, one must count the number density of crystals
as a function of size. This is expressed as a cumulative
distribution. The population density n is the number of
crystals in any size range divided by the size interval.
Katherine Cashman and Bruce Marsh (1988) applied
CSD analysis to drill core samples from Makaopuhi lava
lake, Kileauea volcano, Hawaii. The lengths of plagioclase crystals were measured in rock thin sections and
binned by size, and the population bins were converted
to volume proportions. A typical CSD plot (n versus L)
for plagioclase is illustrated in figure 11.20. The data
define a straight line, implying that crystal nucleation
and growth were continuous throughout the crystallization interval. Crystal fractionation or other magmatic
processes would have disturbed this pattern and produced more complex CSD plots. The nuclei population
density n0 for this basalt, determined from the y intercept, is 2.68 × 107 cm−4. The value for the slope, −1/Yt
= −544 cm−1, can be used to calculate the plagioclase
growth rate Y, since t has been estimated at various
depths for this cooling lava lake. Using a time t of
3.075 × 107 sec, Y = 6.0 × 10−11 cm/sec (∼0.02 mm/yr).
Applying equation 11.18, the nucleation rate J of plagio-
FIG. 11.20. Crystal size distribution (CSD) plot for plagioclase
grains in basalt sample from Makaopuhi lava lake. The slope of the
linear regression line is related to the growth rate Y and crystal
residence time t, and the y intercept gives the nuclei population
density n0. (After Cashman and Marsh 1988.)
clase is 1.6 × 10−3 nuclei/cm3 sec (∼50,000 nuclei/yr). This
example shows that the role of kinetics in the crystallization of rocks can be quantified from analysis of
completely solidified samples, even without recourse to
kinetic theory.
SUMMARY
In this chapter, we have seen that kinetic factors can be
very important to understanding some geologic systems,
even at high temperatures. The temperature dependence
of rate processes is commonly expressed in terms of an
activation energy barrier that must be overcome to initiate the process. This is reflected in supersaturation and
undercooling, both ways of overstepping equilibrium
crystallization conditions.
Three rate processes are particularly important in
high-temperature systems. The first of these, diffusion,
occurs in response to a gradient in chemical potential.
Migration of components may occur along grain boundaries or through phases. Volume diffusion uses naturally
occurring defects in crystal structures as pathways for
movement. Diffusion rates can be calculated from Fick’s
First and Second Laws.
Nucleation, the second major rate-controlling process, is inhibited by surface free energy; this extra energy
224
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
requirement is eliminated after the nucleus reaches a certain critical size. Nucleation may occur by homogenous
or heterogeneous mechanisms. This process is very difficult to study experimentally, and theoretical models provide only qualitative agreement with experimental data.
Exsolution requires nucleation and thus is rate sensitive,
but unmixing by spinodal decomposition occurs without
nucleation of a new phase.
Growth of crystals, the third process, may be controlled by the rate of diffusion of components to the
interface or their rate of attachment to the new phase.
Interface-controlled growth may occur by continuous,
surface nucleation, or screw dislocation mechanisms,
each of which is characterized by a different relationship
between growth rate and undercooling. The rate of
diffusion-controlled growth slows with time because of
chemical gradients that form in the medium surrounding the growing crystal. Compositional or thermal gradients near the growing interface also cause planar crystal surfaces to break down into skeletal or dendritic
morphologies.
Several examples (polymorphic transformation of
aragonite into calcite; unmixing of kamacite and taenite
in iron meteorites) illustrate the importance of determining the rate-limiting step in kinetic processes. Because
kinetic theory is often difficult to apply in understanding
natural systems, geochemists sometimes use other methods that bypass theory and directly determine nucleation
and growth rates. These include controlled cooling rate
experiments and analysis of crystal size distributions.
suggested readings
The literature on geochemical kinetics is not as voluminous as
that for equilbrium thermodynamics, and most references are
portions of books on other subjects.
Carmichael, I. S. E., F. J. Turner, and J. Verhoogen. 1974.
Igneous Petrology. New York: McGraw-Hill. (Chapter 4
provides a discussion of kinetic factors applicable to magmatic systems.)
Gill, R. 1996. Chemical Fundamentals of Geology, 2nd ed.
London: Chapman & Hall. (Chapter 3 provides an easily
understandable summary of kinetically controlled geochemical processes.)
Hoffmann, A. W., B. J. Gilletti, H. S. Yoder Jr., and R. A. Yund,
eds. 1974. Geochemical Transport and Kinetics. Washington, D.C.: Carnegie Institution of Washington. (A collection
of rather technical papers relating to diffusion.)
Kirkpatrick, J. R. 1975. Crystal growth from the melt: A review. American Mineralogist 60:798–814. (An excellent
review paper on the kinetics of crystal growth in silicate
melts.)
Lasaga, A. C., and R. J. Kirkpatrick, eds. 1981. Kinetics of
Geochemical Processes. Reviews in Mineralogy 8. Washington, D.C.: Mineralogical Society of America (Probably
the best available reference on geochemical kinetics; contains excellent chapters on rate laws and dynamic treatment
of geochemical cycles, as well as applications to metamorphic and igneous rocks.)
Lofgren, G. E. 1980. Experimental studies of the dynamic
crystallization of silicate melts. In Hargraves, R. B., ed.
Physics of Magmatic Processes. Princeton: Princeton University Press. (A comprehensive review of controlled cooling
rate experimentation.)
Marsh, B. D. 1988. Crystal size distribution (CSD) in rocks and
the kinetics and dynamics of crystallization. Contributions
to Mineralology and Petrology 99:277–291. (An excellent
review of the theory behind CSD analysis of rocks and the
means of extracting kinetic information from them.)
Thompson, A. B., and D. C. Rubie, eds. 1985. Metamorphic
Reactions, Kinetics, Textures, and Deformation. Advances
in Physical Chemistry 4. New York: Springer-Verlag. (This
reference includes chapters on the kinetics of metamorphic
reactions, as well as the importance of crystal defects and
deformation on rate processes.)
Other papers referenced in this chapter are:
Buening, D. K., and P. R. Buseck. 1973. Fe-Mg lattice diffusion in olivine. Journal of Geophysical Research 78:6852–
6862.
Carlson, W. D., and J. L. Rosenfeld. 1981. Optical determination of topotactic aragonite-calcite growth kinetics: Metamorphic implications. Journal of Geology 89:615–638.
Cashman, K. V., and B. D. Marsh. 1988. Crystal size distribution (CSD) in rocks and the kinetics and dynamics of
crystallization, II: Makaopuhi lava lake. Contributions to
Mineralogy and Petrology 99:292–305.
Greenwood, H. J. 1963. The synthesis and stability of anthophyllite. Journal of Petrology 4:317–351.
Kirkpatrick, R. J. 1974. Kinetics of crystal growth in the system
CaMgSi2O6-CaAl2SiO6. American Journal of Science 274:
215–242.
Kirkpatrick, R. J., G. R. Robinson, and J. F. Hays. 1976.
Kinetics of crystal growth from silicate melts: Anorthite and
diopside. Journal of Geophysical Research 81:5715–5720.
Lofgren, G. 1974. An experimental study of plagioclase morphology. American Journal of Science 273:243–273.
Mathews, A. 1980. Influences of kinetics and mechanism in
metamorphism: A study of albite crystallization. Geochimica et Cosmochimica Acta 44:387–402.
Kinetics and Crystallization
Mueller, R. F., and S. K. Saxena. 1977. Chemical Petrology. New
York: Springer-Verlag.
Taylor, L. A., P. I. K. Onorato, and D. R. Uhlmann. 1977. Cooling rate estimations based on kinetic modelling of Fe-Mg
diffusion in olivine. Proceedings of the Lunar Scientific
Conference 8:1581–1592.
Taylor, L .A., P. I. K. Onorato, D. R. Uhlmann, and R. A. Coish.
1978. Subophitic basalts from Mare Crisium: Cooling rates.
In R. B. Merrill and J. J. Papike, eds. Mare Crisium: The
View from Luna 24. New York: Pergamon, pp. 473–481.
225
Waldbaum, D. R., and J. B. Thompson Jr. 1969. Mixing properties of sanidine crystalline solutions, part IV: Phase diagrams from equations of state. American Mineralogist 54:
1274–1298.
Walker, D., R. J. Kirkpatrick, J. Longhi, and J. F. Hays. 1976.
Crystallization history of lunar picritic basalt sample 12002:
Phase equilibria and cooling rate studies. Bulletin of the
Geological Society of America 87:646–656.
226
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
PROBLEMS
(11.1) Given the following data, determine the activation energy for the reaction analcite + quartz → albite
+ water in sodium disilicate solution.
T (°C)
k
10−4
2.41 ×
2.68 × 10−3
1.84 × 10−2
345
379
404
(11.2) Sketch schematic plots of concentration of element x in the medium surrounding a growing crystal as
functions of distance from the interface for the following conditions: (a) diffusion-controlled growth,
(b) interface-controlled growth, (c) growth controlled by both mechanisms.
(11.3) Using equations 11.11 and 11.15, construct a plot that illustrates how the rates of nucleation and
growth vary with undercooling.
(11.4) (a) Given that ∆H = ∆G + T∆S, show that the surface enthalpy of a liquid droplet is correctly
described by ∆H = σ − T(∂σ/∂T )P. (b) The surface tension of water against air at 1 atm has the
following values at various temperatures:
T (°C)
σ (dynes cm−1)
20
22
25
28
30
72.75
72.44
71.97
71.50
71.18
What is the surface enthalpy (in cal cm−2) of water at 25°C?
(11.5) (a) The volume of the spherical cap shown in figure 11.8 can be calculated from:
V = 1–6 πh (3x2 + h2).
Using this formula and equations 11.6 and 11.1, derive expressions to calculate the volume (V) and
height (h) of the smallest stable nucleus that can be formed on a substrate S, assuming that the quantities σCL, ∆Gv, and (σLS − σCL) can be measured experimentally. (b) What is the volume of the smallest stable nucleus of forsterite for which (σLS − σCS) = 2.378 × 10−5 cal cm−2 at 2063°C? Use the values
for σ (=σCL) and ∆Gv for forsterite at 2063°C from worked problem 11.3. How does this minimal
volume compare to the volume of the smallest homogeneously nucleated forsterite particle?
CHAPTER 12
THE SOLID EARTH AS A
GEOCHEMICAL SYSTEM
OVERVIEW
In this chapter, we explore how the various parts of the
Earth’s interior can be integrated into a grand geochemical system. First, we estimate the compositions of the
reservoirs—that is, crust, mantle, and core—in terms of
both chemistry and phase assemblages. We then investigate how these reservoirs interact through the exchange
of heat and matter. The continental crust was extracted
from the upper mantle, which appears to be geochemically isolated from the lower mantle. Generally, convection within the upper mantle and crust is manifested in
plate tectonics. During certain episodes, however, wholemantle convection may occur, resulting in sinking of
lithospheric plates into the lower mantle and return flow
of matter to the surface via large mantle plumes. These
episodes are driven by thermal convection in the liquid
outer core, resulting from exothermic crystallization of
the solid inner core. Many geochemical interactions between mantle and crust involve basaltic magmatism, so
we discuss the thermodynamics of melting and the geochemical characteristics of partial melts of mantle rocks
generated under various conditions. Most magmas experience significant chemical changes en route to the
surface. Differentiation of magmas by fractional crystallization and liquid immiscibility is considered, and we
attempt to quantify their geochemical consequences. The
behavior of compatible and incompatible trace elements
during melting and differentiation is also examined. As
we shall see, trace elements can be used to quantify geochemical models of magma evolution. We then consider
the compositions, reservoirs, and cycling of volatile elements within the mantle and crust.
RESERVOIRS IN THE SOLID EARTH
Composition of the Crust
The crust is the outer shell of the Earth, in which
P-wave seismic velocities are <7.7 km sec−1. The lower
boundary of the crust is the Mohorovicic discontinuity
(the Moho), a layer within which seismic velocities increase rapidly or discontinuously from crustal values to
mantle values of >7.7 km sec−1. The average thickness
of the crust defined in this way is 6 km in ocean basins,
∼35 km in stable continental regions, and ranges up to
70 km under mountain chains. Even so, the crust is only
a minuscule portion of the mass of the planet, constituting only about 0.5% of it by weight.
Oceanic crust shows no major lateral and vertical
changes in P-wave velocity. As a first approximation, it
can be considered to have basaltic composition throughout, although studies of ophiolites reveal that it has a veneer of siliceous sediments and contains ultramafic rocks
227
228
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
at depth. In contrast, the lateral heterogeneity of continental crust, so evident in geologic maps, apparently
persists to great depth as revealed by variations in subsurface P-wave velocities. This heterogeneity complicates
any attempt to estimate the bulk composition of the crust
as a whole.
Continental crust is conventionally thought to be subdivided into a “granitic” upper layer and a “gabbroic”
lower layer, separated by a seismic discontinuity (the
Conrad discontinuity) that is not everywhere laterally
continuous. This view is certainly too simplistic and may
be wrong altogether. The common mineral assemblage
of gabbro, for example, is apparently not stable at the
temperatures and pressures that occur in the lower crust.
Rocks of this composition should be transformed to
eclogite, a mixture of garnet and clinopyroxene. The
density of eclogite, however, is too high to fit the measured seismic velocities of the lower crust. If there is
enough water in the lower crust, “gabbroic” rocks could
instead consist of amphibolites (amphibole + plagioclase
rocks), which do have the appropriate seismic properties.
Another idea consistent with seismic constraints is that
much of the lower crust under continents is composed of
granulites of intermediate composition.
There are various ways to approach the thorny problem of determining an average composition for the crust,
and all depend on assumptions that we make in describing crustal heterogeneity. We might choose to take the
weighted average of the compositions of various crustal
rocks in proportion to their occurrence, using, for example, geologic maps as a basis. Less direct approaches
utilize the analyses of clays derived from glaciers that
have sampled large continental areas, or estimate the
relative proportions of granitic and basaltic end members
by mixing their compositions in the proper ratio to reproduce the compositions of sediments derived by crustal
weathering. All of these methods have been tried and
generally give similar results.
The most widely accepted crustal compositions usually involve averaging the compositions of rocks assigned
to hypothetical crustal models. This approach was
adopted by A. B. Ronov and A. A. Yaroshevsky (1969),
using the crustal model illustrated in figure 12.1. Their
calculated crustal composition is given in table 12.1.
Ross Taylor and Scott McLennan (1985) estimated the
composition of the crust by assuming that it was composed of 75% average Archean crustal rocks and 25%
FIG. 12.1. A model for the Earth’s crust used in the calculation of
its average composition. (Adapted from Ronov and Yaroshevsky
1969.)
andesite. The Taylor and McLennan model (table 12.1)
is somewhat more mafic than that of Ronov and Yaroshevsky, with lower concentrations of SiO2 and alkalis
and higher concentrations of FeO, MgO, and CaO. However, both of these calculated crustal compositions are
remarkably similar to the composition of average andesite, also shown in table 12.1. The mineralogy of the
average crust is thought to be similar to that of andesite,
which is dominated by feldspars, quartz, pyroxenes, and
amphibole.
The crust is strongly enriched in incompatible elements (those that partition into magmas because they do
not fit easily into the crystal structures of the minerals
that remain unmelted). The incompatible elements in
the crust are lithophile, so-called because they prefer
TABLE 12.1. Estimates for the Average Composition
of the Earth’s Crust
Wt. %
Oxide
Ronov and
Yaroshevsky (1969)
Taylor and
McLennan (1985)
Average
Andesites1
SiO2
TiO2
Al2O3
Fe2O3
FeO
MnO
MgO
CaO
Na2O
K2O
P2O5
59.3
0.9
15.9
2.5
4.5
0.1
4.0
7.2
3.0
2.4
0.2
57.3
0.9
15.9
—
9.1
—
5.3
7.4
3.1
1.1
—
58.7
0.8
17.3
3.0
4.0
0.1
3.1
7.1
3.2
1.3
0.2
1
Average of 89 andesites from island arcs compiled by
McBirney (1969).
The Solid Earth as a Geochemical System
229
FIG. 12.2. Enrichment of lithophile elements in the continental crust relative to abundances in
the Earth’s primitive mantle. Ions of large size and/or charge are incompatible in common mineral
structures and so are concentrated in melts. Over time, these melts have enriched the crust in incompatible elements. Contours group elements with similar degrees of enrichment, from <2, 3–8,
9–20, and >20. (After Taylor and McLennan 1985.)
silicate phases to metal or sulfide, as noted in chapter 2.
Included in this category are elements with large ionic
size (K, Rb, Cs, U, Th, Sr, Ba, and some rare earth elements, such as La and Nd) and high field strength
elements that have high ionic charge (Nb, Ta, Zr, Hf,
Ti). Figure 12.2 summarizes the enrichment of lithophile
elements in the crust relative to abundances in the primitive mantle, plotted against ionic radius and valence.
Contours in this figure group elements with similar enrichment factors. Enrichments of lithophile elements in
the continental crust are so extreme that multiple periods
of melting are required, resulting in a refining process that
further sequesters these elements at each melting step.
Lithophile elements are thought to have been extracted
from the mantle during crust formation.
Composition of the Mantle
The mantle, that region of the Earth’s interior between the Moho and the core, is of course not directly
accessible for study. What we know about it comes from
indirect evidence, mostly geophysical. Seismic wave velocities as a function of depth can be used to calculate
elastic parameters, which depend on the chemical and
mineralogic composition of the mantle, as well as temperature and pressure. Comparison of measured seismic
data with the results of laboratory measurements of the
ultrasonic properties of various phases provides the information necessary for inferring the composition of
the upper mantle. For the lower mantle, we must use
theoretical relationships between velocity and density
and deduce compositions from shock wave experiments.
Despite the inherent uncertainties in interpreting mantle
seismic data, the mineralogic composition of the mantle
seems reasonably well determined.
P-wave seismic velocities of 7.8–8.2 km sec−1 just
below the Moho are consistent with an upper mantle
having the bulk density of peridotite (olivine + orthopyroxene + clinopyroxene) or eclogite (defined above as
clinopyroxene + garnet). Although both of these rock
types occur in the mantle, peridotite is thought to predominate. As we shall see, one of the constraints on petrologic modeling is that mantle rocks must be capable of
generating basaltic liquids on partial melting. Peridotite
can do this, but eclogite cannot. The simple reason is
that eclogite already has a basaltic composition, and
thus must melt completely to produce basaltic magma.
Clearly then, the mantle, consisting mostly of ultramafic
230
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
TABLE 12.2. Estimates for the Average Composition
of the Earth’s Mantle
Wt. %
White
(1967)1
Hutchison
(1974)2
Ringwood
(1975)3
Anderson
(1980)4
SiO2
TiO2
Al2O3
Fe2O3
Cr2O3
FeO
MnO
Ni
MgO
CaO
Na2O
K2O
44.5
0.15
2.6
1.5
—
7.3
0.14
0.2
41.7
2.3
0.25
0.02
45.0
0.09
3.5
—
0.41
8.0
0.11
0.25
39.0
3.25
0.28
0.04
45.7
0.09
3.4
—
0.4
8.0
0.14
0.27
38.4
3.1
0.4
—
47.3
0.2
4.1
—
0.2
6.8
0.2
0.2
37.9
2.8
0.5
0.2
1From
frequency histograms of 168 ultramafic rocks.
mantle peroditite nodules.
3Pyrolite.
480% garnet peridotite + 20% eclogite.
2From
rocks, has a profoundly different bulk composition from
the crust.
Some crustal occurrences of tectonically emplaced
ultramafic rocks may have been derived directly from
the mantle. The compositions of abundantly exposed
ultramafic rocks may be representative of the mantle. A
preferred mantle composition, calculated from frequency
histograms of ultramafic rocks, is presented in column 1
of table 12.2.
Another constraint on the composition of the mantle
comes from xenoliths of mantle material carried upward
by erupting magmas. Those that have not been altered
by extraction of basaltic magma consist of either peridotite or eclogite. We focus on the peridotites, which
we have already concluded are the best candidates for
the bulk of the mantle. There are essentially two types:
spinel peridotites, containing olivine, enstatite, diopside,
and Cr-Al spinel; and garnet peridotites, containing
olivine, enstatite, diopside, and pyrope garnet. In both
cases, olivine and pyroxenes predominate, but the presence of spinel or garnet is very important. Even though
the pyroxenes in peridotite xenoliths are commonly
aluminous, they do not contain enough aluminum to account for the plagioclase component of basalts produced
by partial melting. In rocks of ultramafic composition,
plagioclase itself is the Al-bearing phase at low pressures,
but this gives way to spinel and garnet at higher pressures. The spinel- and garnet-forming reactions can be
represented by:
2CaAl 2Si 2O8 + 2Mg2SiO4 →
anorthite
forsterite
CaMgSi 2O6⋅xCaAl 2SiO6
aluminous clinopyroxene
+ 2Mg2SiO3⋅xMgAl 2SiO6 + (1 − x)CaAl 2Si 2O8
aluminous orthopyroxene
unreacted anorthite
+ (1 − x)MgAl 2O4 → 2CaMg2Al 2Si 3O12.
spinel
garnet
One example of the use of peridotite xenoliths as a measure of the composition of the mantle is illustrated in
column 2 of table 12.2.
One of the problems we encounter in estimating the
mantle composition has to do with partial melting. A
peridotite assemblage without spinel or garnet (and often
without clinopyroxene as well) is said to be depleted,
because basaltic magma has already been extracted from
it. Conversely, peridotite that can still produce basaltic
melt is called undepleted. But how do we determine what
proportion of the mantle is depleted and what proportion is undepleted? This dilemma forms the basis for a
class of hypothetical mantle models known as pyrolite,
originally devised in 1962 by Ted Ringwood. He postulated a primitive mantle material that was defined by the
property that on fractional melting it would yield basaltic
magma and leave behind a residual refractory duniteperidotite. The composition of pyrolite (“pyroxene–
olivine” rock) can be derived, therefore, by combining
depleted peridotite and the complementary basalt in the
proper proportions. This is a rather flexible model for
the mantle composition. Ringwood and his collaborators at various times have presented pyrolite compositions with peridotite:basalt ratios varying from 1:1 to
4:1, although the 3:1 ratio is most often quoted as a reasonable value. A mantle composition based on pyrolite
is compared with others already discussed in column 3
of table 12.2.
Not only is basaltic magma removed from the mantle, but it is also recycled back into the mantle (in the
form of eclogite) by subduction. Consequently, we might
also envision a model composition for the mantle that
mixes average eclogite and undepleted peridotite in the
correct proportions. The model presented in column 4 of
table 12.2 does just that.
All of these methods give comparable results, and
the composition of the mantle seems reasonably well
constrained, at least for the major elements. We should
The Solid Earth as a Geochemical System
keep in mind, however, that these estimates are for the
upper mantle, although they are commonly used to describe the composition of the entire mantle. Is there any
basis for this extrapolation?
Ringwood (1975) provided a wealth of information
about the stability of phases at various depths in a mantle of pyrolite composition. Down to a depth of 600 km,
the assumed mineral assemblages are based on the results of high-pressure experiments, but below this depth,
Ringwood had to infer what phases would be stable from
indirect evidence, such as experimental data on germanate (GeO4−4) analog systems. If germanium is substituted
for silicon, phase transformations that were experimentally inaccessible for silicates occur within pressure ranges
that could be studied in the laboratory. Since that time,
multianvil high-pressure apparatus and shock wave
experiments have confirmed the proposed phase relationships for the equivalent silicates. These phase changes
(see the accompanying box) correspond very well with
observed seismic discontinuities in the mantle, so it seems
plausible that, to a first approximation, the mantle has a
pyrolite composition throughout. This would mean that
most discontinuities within the mantle result from phase
changes, in contrast to the Moho, which is due to a
change in chemical composition.
It is convenient to distinguish the upper and lower
mantle, which are separated by a seismic discontinuity
at 650–700 km depth. These two reservoirs apparently
have different abundances of incompatible trace elements.
The upper mantle is strongly depleted in incompatible
elements, which were presumably extracted during crust
formation, whereas the lower mantle has higher abundances of these elements. There is also some evidence
that the upper and lower mantle may have different
compositions in terms of major elements, expressed in
their inferred Mg/Si, Fe/Si, and Ca/Al ratios. Such variations might have resulted from crystallization in a largely
molten mantle during its formation. If this is correct, then
the 650-km discontinuity (sometimes called the 670-km
discontinuity) is also a compositional boundary, and not
just a phase change.
The lowermost 200 km of the mantle, sometimes
called region D, has peculiar seismic properties and may
have a distinct chemical composition from the rest of the
lower mantle. This region may be a mixture of mantle
material with subducted slabs of ancient crustal rocks.
The term piclogite is sometimes used to describe this mix-
231
ture. Piclogite differs from pyrolite in the upper mantle
in that its ultramafic component is undepleted and thus
rich in incompatible elements.
Composition of the Core
The mean density of the Earth is consistent with a
large, metallic iron core, and its moment of inertia and
free oscillations indicate that this mass is concentrated
at the planet’s center. A major seismic discontinuity at
2900 km depth marks the mantle-core transition, so the
core is approximately half the radius of the Earth. Variations in compressional wave velocities through the core
have led to the conclusion that the outer core is liquid
and the inner core is solid. In fact, the outer core is the
largest magma chamber inside the Earth, forming a layer
2260 km thick between the crystalline lower mantle
and the solid inner core. Most of what we know about
the core is derived from indirect seismic observations.
Despite the rigorous aspects of this discipline, one geophysicist has described the inversion of geophysical data
to yield geological information as “like trying to reconstruct the inside of a piano from the sound it makes
crashing down stairs.” Can geochemistry shed any light
on this complex problem?
Geophysical methods applied to study of the core are
capable of detecting elements with very different atomic
weights, but they cannot distinguish among elements
similar in weight to iron. One such element, nickel, is
thought to be an important constituent of the core, based
on its siderophile character (its geochemical affinity for
metal) and its high cosmic abundance (its abundance on
average in the solar system). To first order, the core can
thus be visualized as a mass of Fe-Ni alloy, part liquid
and part solid. We can also infer that other siderophile
elements, such as cobalt and phosphorus, might be important constituents.
The density of the outer core is about 8% less than
that of the inner core. This is interpreted as reflecting the
presence of an additional component, an element of low
atomic weight, in the outer core. Considerable support
has developed for the hypothesis that this element is
sulfur. The seismic velocities of sulfide and metal at high
pressures indicate that ∼10 wt % S would be necessary
to account for the lower density of the outer core. Iron
meteorites, which formed as cores in asteroids, also
contain significant amounts of sulfur as FeS (the mineral
232
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
PHASE CHANGES IN THE MANTLE
Ted Ringwood’s suggested mantle mineralogy as a
function of depth is summarized in figure 12.3. The
wavy diagonal line from upper left to lower right represents the density distribution with depth, inferred
from P-wave velocities and corrected for the effect of
compression due to the overlying rock. Starting near
the top, just below the Moho, we first encounter the
low-velocity zone, a region of inferred partial melting
in which seismic velocity decreases. The stable assemblage at this point is olivine + orthopyroxene +
clinopyroxene + garnet. Just below 300 km, pyroxenes transform to a garnet crystal structure. One of
the two major discontinuities in the mantle occurs at
400 km; this corresponds to the conversion of olivine,
volumetrically the most important phase to this point,
into a more tightly packed structure called the β phase
(also called wadsleyite). At this point, the mantle consists of β-olivine and a complex garnet solid solution.
Another less pronounced seismic wiggle at ∼500 km
may be due to the transformation of the β phase into
a spinel (γ) structure (also called ringwoodite), as
well as a second reaction in which the calcium component of garnet, possibly with some iron, separates
as (Ca,Fe)SiO3 in the dense perovskite structure.
The second major seismic discontinuity in the
mantle is seen at 650–700 km. At this point, γ-olivine
is thought to disproportionate into its constituent
oxides, MgO (periclase) + SiO2 (stishovite). The (Mg,
Fe)SiO3-Al2O3 components of the garnet solid solution
presumably also transform into the ilmenite structure. Below this transition, the mantle consists of
periclase + stishovite + MgSiO3 ⋅Al2O3 (ilmenite) +
(Ca,Fe)SiO3 (perovskite).
The density changes in mantle materials below
∼700 km could be caused by further phase transformations to more tightly packed structures with Mg
and Fe in higher coordination, but it is very difficult to
specify exactly what transformations may take place.
Some additional phases that have been suggested to be
stable in the lower mantle include (Ca, Fe, Mg)SiO3
in the perovskite structure, (Mg, Fe)O in the halite
structure, and (Mg, Fe)(Al, Cr, Fe)2O4 in the calcium
ferrite structure. From the generality of these formulas, we can see how speculative these phases are.
FIG. 12.3. Summary of possible phase transformations in the
mantle, as determined by Ringwood (1975). The densities
shown by the solid line are zero-pressure densities (corrected
for depth) calculated for pyrolite; density discontinuities
correspond to phase transformations inferred from seismic
studies.
Another possible explanation for the higher densities in the deep mantle is an increased Fe/Mg ratio.
Such a change in mantle composition would have important geochemical implications, but it is difficult to
confirm because the behavior of iron in the mantle is
so uncertain. It has been suggested that Fe2+ ions in
silicates may undergo a contraction in radius due to
spin-pairing of electrons below 1200 km depth. This
is potentially important because low-spin Fe2+ probably would not substitute for Mg in solid solutions.
Another hypothesis, with some experimental justification, is that Fe-bearing silicates might disproportionate to form some metallic iron, even at modest depths
of ∼350 km. This metal might then sink, stripping Fe
from some parts of the mantle and enriching others.
The Solid Earth as a Geochemical System
troilite). Other geochemists have argued that oxygen is
the light element in the outer core. Bonding in FeO is
expected to become metallic, rather than ionic, at very
high pressures, and an outer core with a composition of
about Fe2O (50 mol % oxygen) would satisfy the density constraint. Although oxygen is not alloyed with
metal in iron meteorites, the cores in asteroids would not
have formed at the high pressures necessary for this reaction. Another possibility is silicon; the density of Fe-Si
alloys at core conditions requires ∼17 wt % Si.
One way to estimate the composition of the core directly is to assume that siderophile elements are present in
cosmic relative abundances to iron. (We discuss cosmic
abundances in chapter 15.) An example of this approach
is shown in the following worked problem.
Worked Problem 12.1
What is the composition of the Earth’s core? One way to estimate core composition is to assume that siderophile elements
accompany iron in cosmic proportions. Of course, we must also
include major amounts of a light element to account for the
density of the outer core, and we have just seen that this element
may not be in cosmic proportion relative to iron.
Let’s assume a value of 9.0 wt % sulfur in the core, based on
sulfur abundance from density constraints. The first step, then,
is to determine how much iron must be combined with sulfur
to make FeS:
9.0/32.05 g mol−1 = x/55.85 g mol−1,
so that:
x = 15.68 wt % Fe.
The wt % FeS in the core is then 15.68 + 9.0 = 24.68. By difference, the core must consist of 75.32 wt % metal.
The metal phase also contains other siderophile elements in
addition to iron. Siderophile elements with the highest cosmic
abundances are nickel, cobalt, and phosphorus, and these are
prominent constituents of iron meteorites. A mass balance equation for metal is thus:
Femetal + Ni + Co + P = 75.32%,
or
Femetal = 75.32 − Ni − Co − P,
(12.1)
where the abundance of each element is expressed in wt %.
Nickel can be determined from the cosmic wt ratio of Ni and Fe:
Ni = (Nicosmic/Fecosmic) Fecore
= (4.93 × 10 4/58.77 g mol−1)/
(55.85 g mol−1/9 × 105)Fecor = 0.0521 Fecore,
233
where Fecore is the total abundance of iron in the core, equal to
Femetal + 15.68. Similarly, Co = 0.0024 Fecore and P = 0.0208
Fecore. Substitution of these values into equation 12.1 gives:
Femetal = 75.32 − 0.0521 Fecore − 0.0024 Fecore
− 0.0208 Fecore = 68.94 wt %.
Thus, the total amount of iron in the core is 68.94 + 15.68 =
84.62 wt %. Nickel, cobalt, and phosphorus wt percentages
are obtained by multiplying their cosmic wt ratios by 84.62;
the respective values are 4.41, 0.20, and 1.76 wt %.
FLUXES IN THE SOLID EARTH
Cycling between Crust and Mantle
The mantle and crust are coupled through plate tectonics, which provides a mechanism for adding mantle
material and heat to the crust (through volcanic activity) and recycling crust back into the mantle (through
subduction). Plate tectonics reflects mantle convection,
although the relationship is not as clear as was once
thought. At present, ocean floor is being created and
subducted at a rate of ∼3 km3 yr−1. If the average thickness of a plate is 125 km, then 375 km3 of rock is subducted each year. At that rate, the entire mantle (having
a volume of 9 × 1011 km3) could be processed through
the plate tectonic cycle in 2.4 billion years (approximately
half the age of the Earth). Mantle plumes, also called hot
spots, are sites of volcanism that also must relate to
convection. Mantle convection plays a dominant role in
cooling the Earth, but there is no consensus on how convective flow extends to depth.
The main controversy centers on whether the whole
mantle is involved in convective overturn, or the upper
and lower mantles convect separately. The most prominent convection is associated with the sinking of plates
at subduction zones. These plates are metamorphosed
to dense eclogite on subduction, which adds to the gravitational imbalance caused by their cooler temperatures.
On approaching 650 km depth, however, the high calcium and aluminum contents of eclogite stabilize garnet
and hinder transformation to the denser perovskite that
occurs in the surrounding mantle peridotite, so the density contrast decreases. Studies of earthquakes show that
the subducted plates descend to depths of 650 km,
below which seismicity is absent (fig. 12.4). Some highresolution seismic images indicate that subducted slabs
bounce off the 650-km discontinuity and accumulate in
234
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Spreading
Center
Subduction
Zone
Subduction
Zone
Plume
Plume
Two-Layer Mantle
Convection
Whole-Mantle
Convection
FIG. 12.4. Sketch of a cross-section of the Earth, showing convection in the core, mantle, and crust. Mantle plumes and subduction zones allow interaction of matter between mantle and
crust. The left side of the diagram illustrates two-layer mantle
convection and the right side shows whole-mantle convection.
the upper mantle. However, other seismic studies have
documented that subducted slabs sometimes sink below
this boundary, in some cases all the way to the interface
between the mantle and core (fig. 12.4). The subducted
crust is colder than the lower mantle into which it descends, so differences in seismic velocity allow the slabs
to be imaged using seismic tomography. If subducted
slabs are used as a tracer for convection, then the upper
and lower mantles are sometimes isolated from each
other and sometimes not.
If slabs sink to the core-mantle boundary, an equivalent mass of material must be transferred back into
the upper mantle or crust. Such material would be relatively hot and thus have lower seismic velocities. Seismic
tomography reveals that two regions of low velocity
occur within the lower mantle: one underlies the Pacific
basin and the other is situated beneath southern Africa
(fig. 12.5). Within both these areas lie the majority of
the Earth’s plumes, and both areas are highs on the geoid
(the geometrical form that describes the planet’s shape).
The return flow from the lower mantle thus appears to
consist of scattered upwellings that form clusters of hot
spots.
The existence of separate mantle reservoirs for various kinds of magmas supports the idea that the upper
and lower mantles do not convect as a single unit. The
midocean ridge basalts (MORB) that erupt at spreading
centers are strongly depleted in incompatible elements.
Melting experiments suggest that MORB is generated
at pressures corresponding to depths within the upper
mantle. The depletion of incompatible elements in this
source region constitutes evidence that the crust was extracted from the upper mantle, as illustrated in worked
problem 12.2. Conversely, magmas thought to be generated within the lower mantle and erupted at hot spots
have higher abundances of incompatible elements. The
lower mantle may thus be geochemically isolated from
+20
-10
+20
+10
0
-10
0
0
-10
-20
-20
FIG. 12.5. Contours of average lower mantle seismic velocity perturbations define two large regions of high temperature (low velocity, indicated by negative contours). These regions define the
locations of mantle plumes, and they correspond to clusters of hot spots (solid circles). (Redrawn
from Castillo 1988.)
The Solid Earth as a Geochemical System
the upper mantle, although plumes generated at the coremantle boundary obviously must pass through both the
lower and upper mantle on their way to the surface.
Worked Problem 12.2
The extraction of the incompatible-element-enriched crust
from the mantle has produced a complementary incompatibleelement-depleted reservoir in the mantle. Can we use incompatible element concentrations to estimate what fraction
of the mantle was involved in producing the crust?
An approximate upper limit for mantle involvement can be
calculated from mass balance considerations. Rubidium is a
highly incompatible element, strongly enriched in the crust
and depleted in MORB. Using the data in table 12.3, we can
calculate the effect of extracting the crustal complement of
incompatible Rb and relatively compatible Yb from varying
amounts of primitive mantle. The relative masses of crust and
mantle are 0.56% and 99.44%, respectively. The mass balance
equation is:
Rb in primitive mantle =
Rb in depleted mantle + Rb in crust,
or
mmantle (cpm) = nmmantle (cdm) + mcrust (cc),
(12.2)
where mmantle and mcrust are the mass fractions of the mantle
and crust, n is the fractional degree of mantle depletion, and
cpm, cdm, and cc are the concentrations of Rb in the primitive
mantle, depleted mantle, and crust, respectively. Substituting
the values in table 12.3 for 100% mantle depletion (n = 1) and
solving for depleted mantle abundance of Rb (cdm) gives:
0.9944(0.55) = 1(0.9944)(cdm) + 0.0056(32),
or
cdm = 0.37 ppm Rb at 100% mantle depletion.
Similarly, we can use equation 12.2 to calculate the concentration of Rb and Yb at other degrees of mantle depletion
(with n expressed as a fraction of total mantle), as shown in
table 12.3. Also shown in table 12.3 is the measured weight
ratio of Rb/Yb in MORB, which should reflect the ratio of these
elements in the depleted mantle source region. By comparing
the calculated Rb/Yb ratios at different degrees of mantle depletion with the measured MORB Rb/Yb ratio, we can estimate the
degree of mantle involvement in crust formation. These simple
calculations bracket the amount of mantle that has been affected
by extraction of the crust between 33% and 50%. A similar result can be obtained by doing the same calculation with other
incompatible elements, such as barium.
If we accept the base of the upper mantle as the 650 km discontinuity, then the upper mantle volume corresponds almost
exactly to one-third of the entire mantle. Thus, it seems plausible that the upper mantle constitutes the part that has been
partially melted to make the crust.
The contrasting ideas about mantle convection might
be reconciled by a model that allows the Earth to lose
heat by different mechanisms at different times. During
normal periods (sometimes called Wilson cycles), such
as we are experiencing now, upper and lower mantle
convection cells are distinct, with most heat loss occurring at spreading centers through plate tectonics. Periodically, however, the Earth experiences whole-mantle
convection, as large plumes of hot material reach the surface. During these MOMO (mantle overturn and major
orogeny) episodes, accumulated cold slabs descend from
the 650-km boundary into the lower mantle, and huge
plumes rise from the core-mantle boundary. Eruption
from the heads of such plumes produce large igneous
provinces characterized by vast outpourings of basaltic
magma (flood basalts). These contrasting styles are illustrated in figure 12.6. Flood basalt provinces form over
relatively short time periods and may have caused significant environmental catastrophies, including some mass
extinctions. After the plume head becomes depleted, the
plume tail continues to be a source of magmas on a
smaller scale. This usually produces a chain of volcanoes,
as in Hawaii, as the overlying plate rides over the stationary hot spot.
TABLE 12.3. Rubidium and Ytterbium Mass Balance in the Mantle and Crust
Rb (ppm)
Yb (ppm)
Rb/Yb
235
Predicted Depleted Mantle1
Primitive
Mantle
Crust
100%
50%
33%
25%
MORB
0.55
0.37
1.49
32
2.2
14.5
0.37
0.36
1.03
0.19
0.35
0.55
0.003
0.34
0.01
—
0.33
—
2.2
5.1
0.43
From Taylor and McLennan (1985).
1Predicted concentrations at various degrees of mantle involvement.
236
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Wilson Cycle
Subduction
Zone
MOMO Episode
Mid-Ocean
Ridge
Subduction
Zone
Plume
Upper
Mantle
Lower
Mantle
Mantle
Core
FIG. 12.6. Comparison of convection during Wilson cycles and MOMO episodes. During most periods of the Earth’s history, including the present day, plate tectonics prevails and convection cells
within the upper and lower mantles are largely isolated. During MOMO (mantle overturn and major
orogeny) episodes, subducted slabs that have accumulated at the base of the upper mantle sink
into the lower mantle, and multiple plumes rise from the core-mantle boundary.
Heat Exchange between Mantle and Core
The metallic liquid in the Earth’s outer core is constantly crystallizing, so that the solid inner core is growing larger with time. Variations in the velocity of seismic
waves traveling at different orientations within the inner
core have been interpreted as evidence that the inner core
may be one huge crystal. Exothermic crystallization yields
large amounts of heat that stirs the outer core. Convection of this metallic liquid is responsible in part for the
Earth’s magnetic field. Heat carried upward through
the core is eventually transferred to the bottom of the
mantle.
Although most of the flux across the core-mantle
boundary is heat, there is some suggestion of mass
transfer as well. Reactions between liquid metal and
perovskite in the lower mantle have been postulated. If
such reactions occur, this might provide a mechanism for
adding a light element to the outer core, so in this case,
mass transfer would be from the mantle downward into
the core. Small amounts of core material might also intrude into the lowermost mantle and eventually be entrained in mantle convection cells. The evidence for this,
too, is speculative, based on small anomalies in the isotopic composition of the siderophile element iridium.
Heat transfer through the core-mantle boundary, though,
is unambiguous. It is a major driving force in mantle
convection and must be responsible for mantle plumes.
Fluids and the Irreversible Formation
of Continental Crust
Production of continental crust occurs primarily
above subduction zones, where melting of the mantle
wedge above the subducted slab occurs. At the time it
is subducted, oceanic lithosphere has been hydrated by
reactions with seawater, so that water is also carried
downward into the upper mantle. When the slab reaches
a depth appropriate for transformation to eclogite, the
water is driven off and migrates into the overlying wedge
of mantle peridotite. Melting under hydrous conditions
can produce andesitic magmas with higher silica contents
than basalt. Andesitic magmas normally contain a few
percent of dissolved water, which is cycled back to the
surface when the magmas outgas on eruption. Magmas
that are erupted at subduction zones beneath continents
immediately become continental crust. Where subduction
occurs beneath oceanic lithosphere, magmatic arcs form
and may be eventually become accreted to continents.
The production of continental crust appears to be
unidirectional. The aggregate density of “granitic” continental crust is too low for this material to be subducted. Consequently, once continental crust forms, it is
likely to remain at the Earth’s surface. This is in marked
contrast to oceanic crust, which is consumed in subduction zones at virtually the same rate that it is generated
at midocean ridges. In fact, the only way that oceanic
The Solid Earth as a Geochemical System
crust can escape being reincorporated into the mantle
is by its accidental accretion onto continents by tectonic
processes (producing ophiolites).
As the continental crust accumulated over time, its
composition apparently changed. Scott McLennan (1982)
found that Archean clastic sedimentary rocks were depleted in silicon and potassium and enriched in sodium,
calcium, and magnesium relative to post-Archean rocks.
These data indicate that the sources of Archean sediments were more mafic than for Proterozoic sediments.
In an exhaustive survey of 45,000 analyses of common
crustal rocks, A. B. Ronov and his collaborators (1988)
determined that progressive changes occurred in the compositions of shales, sandstones, basalts, and granitic rocks.
Each of these lithologies showed decreases in magnesium, nickel, cobalt, and chromium and increases in
potassium, rubidium, and other lithophile elements with
increasing time. There is disagreement about whether
these secular changes are continuous or a sharp break
occurs at the Archean-Proterozoic boundary 2.5 billion
years ago.
Models for the growth of continental crust fall into
two basic categories: early formation of virtually the
entire crust with subsequent recycling, or continuous or
episodic increase in the amount of crust throughout
geologic time. Although this remains a contentious subject, continuous or quasicontinuous growth models are
most popular. Ross Taylor and Scott McLennan (1985)
argued that 90% of the continental crust was in place by
the end of the Archean. They also proposed that most
(60–75%) of the crust was produced episodically during
the period 3.2 to 2.5 billion years ago. This transition
might explain, in part, the compositional variations between Archean and post-Archean crust.
From our discussion of these interactions between
crust, mantle, and core, we can see that the interior parts
of the Earth constitute a grand geochemical system. The
fluxes of materials and energy between these reservoirs
are summarized in figure 12.7. Heat and possibly some
matter is exchanged between core and mantle. Heat and
matter in the form of ascending magmas or subducted
lithosphere are exchanged between mantle and crust. The
Earth is continuing to differentiate irreversibly in such a
way that the mantle is depleted in fusible components,
which are ultimately added to continental crust. Although
continental crust, once formed, cannot be recycled, the
fluids that play an important role in its formation are
cycled between mantle and crust at subduction zones.
Subduction
Crust
Lithophile
Elements,
Volatiles
237
Spreading
or Plumes
Lithophile
Elements,
Volatiles
Convecting
Mantle
Iron?
Convecting
Liquid
Outer Core
Light
Elements?
Transfer of
Mass
Crystallizing
Inner Core
Heat
FIG. 12.7. Cartoon summarizing the fluxes of heat and matter
between the core, mantle, and crust reservoirs.
All of these fluxes are driven by thermal convection. In
the following sections, we examine in more detail how
these geochemical interactions take place.
MELTING IN THE MANTLE
Because mass transfer from the mantle to the crust
commonly involves magmas (although hot, solid materials also migrate in rising plumes), we consider the
melting process in some detail. Melting is probably never
complete, and partial melts and residues usually exhibit
significant chemical differences.
The transformation from the crystalline to the liquid
state is not very well understood at the atomic or molecular level. One (largely chemical) description of the
melting process is that it results from increasing the vibrational stretching of bonds between atoms to the point
at which the weakest bonds are severed. Another (primarily structural) view of melting is that it parallels a
change from long-range crystallographic order to the
short-range order characteristic of melts. X-ray diffraction studies of minerals at low and high temperatures
have shown no significant distortions of crystal structures
238
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
as melting temperatures are approached, but unit cell
volumes increase. However, more work needs to be done
before the physics of the melting process can be detailed.
Thermodynamic Effects of Melting
Melting, like other phase transformations in the mantle, results in discontinuous changes in the thermodynamic properties of the system. A sudden increase in
entropy, reflecting the higher state of atomic disorder
in liquids relative to crystals, is probably the most obvious thermodynamic change. More important from the
geochemist’s point of view, however, is the increase in
enthalpy that accompanies melting. The thermal energy
necessary to convert rock at a particular temperature to
liquid at that same temperature is the enthalpy of melting, also called the latent heat of fusion. This is the thermal energy required beyond that necessary to raise the
temperature of the rock to its melting point.
The enthalpy of melting can be measured by dropping crystals or glass into a calorimeter at temperatures
just above and just below the melting point. (Calorimetry
was mentioned in chapter 3.) From the rise in temperature of the calorimeter bath, or change in the proportion
of crystals and liquid, the energy released by each state
of the material can be evaluated. It is possible to calculate the enthalpy of melting for more complex systems
that consist of several crystalline phases, provided that
the phase diagram relating these phases has been determined and the values of ∆Hm for the individual phases
are known, as illustrated in worked problem 12.3.
Worked Problem 12.3
How much thermal energy is required to melt a mixture of
diopside and anorthite, initially at 0°C? Let’s assume that the
proportions of these two minerals correspond to the eutectic
composition in the system diopside-anorthite (you may wish to
review this phase diagram, illlustrated at the bottom of fig. 10.6).
This calculation is performed in two steps. First, we must
determine the energy required to raise the temperature of this
material from 0°C to the melting point, which is 1274°C for a
eutectic composition in this system (see fig. 10.6). Recall that
the eutectic represents the composition of the last liquid to
crystallize or the first liquid to form during partial melting.
From the phase diagram, we find that the eutectic composition
(in wt %) is Di58An42. The heat capacity (∆C̄P) for diopside is
100.40 J mol−1 K−1 = 23.99 cal mol−1 K−1, and that for anorthite
is 146.30 J mol−1 K−1 = 34.97 cal mol−1 K−1 (Robie et al. 1978).
We convert these heat capacities into units of cal g−1 K−1 be-
cause the eutectic composition is given in units of weight. Division by the gram formula weights for diopside (216.55 g mol−1)
and anorthite (278.21 g mol−1) gives ∆C̄P = 0.1109 cal g−1 K−1
for diopside and 0.1257 cal g−1 K−1 for anorthite. The energy
required to raise the temperature of this mixture by 1274° is
therefore:
(0.1109 cal g−1 K−1)(1274 K)(0.58)
+ (0.1257 cal g−1 K−1)(1274 K)(0.42) = 149.21 cal g−1.
The second step is to calculate the enthalpy of melting of
this eutectic mixture. ∆H̄m for diopside is 77.40 kJ mol−1 =
18,345 cal mol−1 and for anorthite is 81.00 kJ mol−1 = 19,359
cal mol−1. Division by the appropriate gram formula weights
gives ∆H̄m for diopside of 84.71 cal g−1 and for anorthite of
69.58 cal g−1. Multiplying these values by the appropriate fractions for the eutectic composition yields:
(84.71 cal g−1)(0.58) + (69.58 cal g−1)(0.42) =
78.36 cal g−1.
The heats of mixing of these liquids are so small that they can
be neglected. Thus, the total thermal energy necessary for this
melting process is the sum of the energy required to raise the
temperature of a eutectic mixture of diopside and anorthite to
the melting point plus the energy required to transform this
mixture into a liquid at that same temperature:
149.21 cal g−1 + 78.36 cal g−1 = 227.51 cal g−1.
A realistic value for the enthalpy of melting of mantle
peridotite is difficult to obtain, because of the unknown
effects of high pressure and of other solid solution components in pyroxenes, olivine, and garnet or spinel. One
commonly quoted value for garnet peridotite at 40 kbar
pressure is 135 cal g−1. For melting to occur in the
mantle, this amount of thermal energy must be supplied
in excess of the heat necessary to bring mantle rock to
the melting temperature.
Types of Melting Behavior
If the temperature of mantle peridotite is increased to
the melting point and some extra thermal energy is added
for the enthalpy of melting, magma is produced. As in
the case of eutectic melting in the diopside-anorthite system discussed in worked problem 12.3, fusion begins at
some invariant point, so that several phases melt simultaneously. The melting process can be described in several ways. Equilibrium melting (also called batch melting)
is a relatively simple process in which the liquid remains
at the site of melting in chemical equilibrium with the
The Solid Earth as a Geochemical System
solid residue until mechanical conditions allow it to escape as a single “batch” of magma. Fractional melting
involves continuous extraction of melt from the system
as it forms, thereby preventing reaction with the solid
residue. Fractional melting can be visualized as a large
number of infinitely small equilibrium melting events.
Incremental batch melting lies between these two extremes, with melts extracted from the system at discrete
intervals.
Worked Problem 12.4
To illustrate the distinction between equilibrium and fractional
melting, let’s examine fusion processes in the two-component
system forsterite-silica, shown in figure 12.8. Describe melting
in this system under equilibrium and fractional conditions.
This system contains a peritectic point between the compositions of forsterite (Fo) and enstatite (En) and a eutectic between the compositions of enstatite and cristobalite (Cr), seen
more clearly in the schematic insets. If enough heat is available
to raise the temperature of the system to 1542°C and offset
the enthalpy of melting, any composition between En and Cr
239
begins melting at the eutectic, with both solid phases entering
the melt. For small degrees of partial melting under equilibrium
conditions, the melt will have the eutectic composition. With
further melting, one phase usually becomes exhausted from the
solid residue before the other does. Begin, for example, with a
composition x that plots between the peritectic and the eutectic,
shown in the upper inset in figure 12.8. As this mixture of En +
Cr melts, Cr will be exhausted from the residue first. As melting continues under equilibrium conditions, the composition of
the liquid leaves the eutectic and follows the liquidus upward,
all the while melting enstatite. Total melting occurs when the
liquid reaches the starting composition. This path is the reverse
of equilibrium crystallization.
Fractional melting produces a sequence of distinct melts
rather than one evolving melt composition. Again, consider
the same mixture x of enstatite and cristobalite as before,
shown in the upper inset. The initially formed liquid, which has
the eutectic composition, is removed from the system as it
forms, so the composition of the remaining solid shifts toward
En. After Cr is exhausted from the residue, the temperature
must increase to the peritectic temperature for any further melting to occur. When the temperature reaches 1557°C, En melts
incongruently to Fo plus liquid, and the liquid is immediately
removed. After all the En is converted to Fo, no further melting
FIG. 12.8. Phase diagram for the system forsterite (Mg2SiO4)-silica (SiO2). The insets illustrate the
eutectic and peritectic in this system; the upper inset (melting paths) is used in worked problem
12.4, and the lower inset (crystallization paths) in worked problem 12.5. The silica-rich side of the
diagram contains a region of silicate liquid immiscibility (“2 Liquids”).
240
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
can occur until the liquid reaches the melting point for pure Fo.
Fractional melting, if carried to completion, thus produces three
batches of melt having distinct compositions, whereas equilibrium melting results in one evolving but (at any time) homogeneous melt.
Some details of the melting relationships of mantle
peridotite are illustrated schematically in figure 12.9.
Phase changes for the stable aluminum-bearing mineral
in fertile peridotites are illustrated at temperatures below
the solidus. Melting of peridotite begins as its temperature is raised just above the solidus. As temperature
increases further, clinopyroxene, spinel, and garnet are
completely melted and thereby exhausted from the solid
residue (as indicated by lines labeled Cpx-out, Sp-out,
and Gar-out, respectively). This all occurs within ∼50°C
of the solidus, producing a wide P-T field for the coexis-
FIG. 12.9. Schematic phase diagram for mantle peridotite, illustrating the effects of partial melting. The solidus and liquidus are
indicated by heavy lines. The aluminous phase changes from plagioclase (Plg) to spinel (Sp) and then garnet (Gar) with increasing
pressure. With increasing temperature, first clinopyroxene (Cpx)
and then spinel or garnet are exhausted from the residue, leaving
olivine (Ol) + orthopyroxene (Opx) or olivine alone. The dashed
lines are contours of the percentage of dissolved olivine in the
melt; they show that with increasing pressure, magmas become
more olivine rich. Also illustrated are typical geothermal gradients
for oceanic and continental regions.
tence of olivine + orthopyroxene with liquid. The olivine
+ orthopyroxene residue is the depleted peridotite that
was discussed earlier.
The composition of the basaltic liquid produced by
partial melting of peridotite changes with pressure, becoming less silica rich at higher pressure. Lower silica
contents of liquids increase the proportion of olivine that
can ultimately crystallize from the melt on cooling; the
percentage of olivine in the melt is contoured as dashed
lines in figure 12.9. Variation in pressure (or depth) of
magma generation thus results in basaltic melts ranging
from tholeiite (the common lava composition at midocean ridges) at modest pressures to more olivine-rich
alkali basalt and basanite (which occur at hot spots) at
higher pressures.
Causes of Melting
The geothermal gradient (also called the geotherm)
is the rate of increase of temperature with pressure, or
depth. In an earlier section of this chapter, we noted that
the elastic properties of the low-velocity zones in the
mantle are consistent with partial melting. However,
from an examination of the relative positions of geotherms and the peridotite solidus in figure 12.9, there is
no apparent reason for incipient melting in this zone or,
for that matter, anywhere else in the mantle. How can
we then account for magma generation? There is no
simple answer to this question, but there are at least
three intriguing possibilities: localized temperature increase, decompression, and addition of volatiles. Let’s
see how each of these could cause melting.
Simply raising the temperature of a rock (the path
labeled +T in fig. 12.10) is the most obvious mechanism
by which melting can occur; this has sometimes been
called the hot plate model. However, in considering this
model, we should keep in mind the difference in magnitude between the enthalpy of melting and the heat capacity for silicate minerals. ∆C̄P values are generally several
hundred times lower than corresponding ∆H̄m values
(compare the values for diopside and for anorthite in
worked problem 12.3), so it is much easier to raise the
temperature of a rock than to produce a significant
amount of partial melt from it. Melting consumes large
quantities of thermal energy, moderating temperature
variations within the system in its melting range. Even
so, it may be possible to have localized mantle heating
that produces magma. Frictional heating of subducted
The Solid Earth as a Geochemical System
FIG. 12.10. P-T diagram illustrating three possible ways that
mantle rocks may melt. Increase in temperature (+T path) could
result from localized concentration of heat producing radionuclides or underplating by other magmas. Decrease in confining
pressure (−P path) may occur because of diapiric upwelling of
ductile mantle rock. Addition of a fluid phase (+H2O path) depresses the melting curve to lower temperature for the same
pressure.
lithospheric slabs or dissipation of tidal energy due to the
gravitation attraction of the Sun and Moon have been
suggested as possible causes of localized heating in the
mantle. More plausible is local heating of the lower crust
in response to underplating by mantle-derived magmas.
The most important mechanism for melting mantle
rocks is almost certainly the release of pressure, caused
by the convective rise of hot, plastic mantle rocks. The
effect of decreasing confining pressure is generally to
lower the solidus temperature. One example, in the
system diopside-anorthite, is illustrated in figure 12.11.
The eutectic temperature, which is the point at which
partial melting begins, is decreased by >100°C in going
from 10 kbar to 1 atm pressure. The position of the
eutectic point also shifts, because the melting point of
diopside increases much more for a given pressure change
than does that of anorthite. Therefore, the Di field is enlarged with increasing pressure at the expense of the An
field. The effect of decreasing pressure on the solidus of
mantle peridotite can be seen in figure 12.9. Because
of the positive slope of this melting curve in P-T space,
decompression of solid rock can induce melting, as illustrated by the path labeled −P in figure 12.10. Because
decompression melting occurs at constant temperature
(we used the term adiabatic to describe such a process in
241
chapter 3) over some interval of pressure, conventional
T-X phase diagrams (which are drawn for a fixed pressure) cannot really be used to understand this process. A
recent innovation in visualizing polybaric melting is
discussed in an accompanying box.
Changes in the composition of a rock system due to
gain or loss of volatiles, principally H2O and CO2, are
also potentially important in facilitating melting. Water
lowers the solidus of peridotite, as illustrated in figure 12.13 by the wet solidus curve. The addition of water
to a dry peridotite that is already above the wet solidus
will cause spontaneous melting, as illustrated by the
+H2O path in figure 12.10. We have already discussed
how an influx of water into the mantle can be caused by
dehydration reactions in a subducted slab of hydrated
lithosphere. The effect of the addition of CO2 to mantle
peridotite is less drastic than for H2O, but still significant. The solidus of peridotite in the presence of CO2 is
illustrated in figure 12.13 by the line labeled CO2. When
the fluid phase contains both H 2O and CO2, the effect
on the peridotite melting curve is complicated. A solidus
for a mixed fluid composition of XCO = 0.6 is illustrated
2
in figure 12.13. At modest pressures of 10 to 20 kbar,
water predominates, because it is most soluble in the
melt; but at higher pressures, the effect of CO2 becomes
marked as its solubility increases.
Both of these volatile components also strongly influence the compositions of the magmas that are generated.
Water selectively dissolves silica out of mantle phases
FIG. 12.11. Phase diagram for the system diopside (CaMgSi2O6)anorthite (CaAl2Si2O8), demonstrating the effect of increasing
confining pressure on melting relationships. Lowering pressure
from 10 kbar to 1 atm results in a decrease in the eutectic melting temperature of >100°. (Based on experiments by Presnall et
al. 1978.)
242
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
DECOMPRESSION MELTING OF THE MANTLE
T-X phase diagrams of the type considered in chapter 10 are appropriate for understanding how magmas
and crystals behave when temperature and composition are variables and pressure is constant. We have
already seen that the Gibbs free energy is minimized
at equilibrium under such conditions. However, decompression melting, which is the most important
process by which magmas are generated, cannot be
visualized using the conventional T-X phase diagrams.
What are the appropriate variables when melting
occurs over a range of pressures?
If decompression melting is adiabatic (q = 0) and
reversible (dq = TdS, where S is the entropy), the process is isentropic (dS = 0). The appropriate variables
for visualizing polybaric melting are thus P and S,
with composition (X) held constant. At equilibrium
under these conditions, enthalpy rather than free
energy is minimized.
Figure 12.12 illustrates a schematic P-S diagram
for a one-component system with two phases—solid
and liquid. The diagram resembles the familiar T-X
phase loop bounded by a solidus and a liquidus for a
binary system, and it can be read in much the same
way. Isentropic decompression is indicated on this diagram by a vertical (decreasing P, constant S) line. At
point 1, the system is entirely solid, but it begins to
melt at point 2, where it passes into the two-phase
field. At any given pressure within the two-phase field,
the relative proportions of solid and liquid can be
determined by applying the lever rule to a horizontal
line passing through the point. Under equilibrium
melting conditions, melting is complete at point 3.
The same diagram can also be used to model fractional melting. Strictly speaking, this process is neither
adiabatic nor isentropic, because melt leaving the
system carries entropy with it. However, each small
increment of melting can be considered isentropic.
A solid having a value of entropy given by point 1
undergoes pressure release and begins to melt when it
reaches point 2. As fractional melting proceeds, the
system constantly changes composition and the solid
residue follows the solidus, that is, the path from 2
to 3′. At the beginning of melting at point 2, the first
increment of melt is identical, whether melting is equilibrium or fractional. However, for fractional melting,
FIG. 12.12. Pressure-entropy diagram illustrating melting of
diopside under equilibrium (path 1-2-3) and fractional (path
1-2-3′) conditions. This diagram allows us to visualize how
melting takes place over a range of decreasing pressures typical of those experienced in an ascending mantle plume.
although the mass of the solid residue continuously
decreases, it is not possible to melt the solid completely
by this process. As a consequence, fractional melting
produces less melt than equilibrium melting for a given
decompression interval.
One subtlety of the P-S diagram is that the amount
of melt generated per decrement of pressure increases
with decreasing pressure. Using figure 12.12, you can
demonstrate this for yourself by comparing the pressure intervals required to produce 50% melting and
100% melting under equilibrium melting conditions.
This effect results from the curvature of the liquidus
and the solidus. The conclusion that melting accelerates with decompression cannot be worked out by
studying this process on a P-T diagram. The effect is
appreciable in natural rock systems and probably accounts for most melts generated within the mantle.
Complications result from phase changes in the
mantle, which produce nonuniform zones of melting
during upwelling, as discussed by Paul Asimow and
his collaborators (1995).
The Solid Earth as a Geochemical System
243
DIFFERENTIATION IN
MELT-CRYSTAL SYSTEMS
In the preceding section, we painted a broad picture of
melting processes in the mantle, gleaned mostly from
studies of volcanic rocks. In reality, the picture is much
fuzzier, because very few mantle-derived magmas arrive
at the Earth’s surface in pristine condition. Most have
been altered in composition (differentiated) to various
degrees. We will now see how that happens.
Fractional Crystallization
FIG. 12.13. Phase diagram showing the solidus in the system
peridotite-CO2-H2O. Melting in the presence of pure water occurs
at lower temperatures than for pure CO2. The dashed line labeled
XCO = 0.6 corresponds to peridotite melting with a mixed fluid of
2
that composition. Water is the most important volatile species
at pressures of less than ∼20 kbar, but CO2 becomes increasingly
important at higher pressures because of its increased solubility
in the melt.
before melting occurs; as soon as the melt forms, it effectively swallows this fluid and thereby becomes richer
in silica (sometimes producing andesites rather than
basalts). However, CO2 favors the production of alkalic
magmas poor in silica (nephelinites and kimberlites).
Concluding this section on melting, we summarize
some of these concepts by relating current ideas about
how mantle-derived magmas are formed in various tectonic settings: (1) Partial melting of relatively dry peridotite accounts for the vast outpourings of tholeiitic
basalt at midocean ridges. This style of melting is initiated by decompression resulting from the rise of mantle
under spreading centers. (2) Hydrated slabs of oceanic
lithosphere subducted at convergent plate boundaries
transform into eclogite at depths of ∼100 km, thereby releasing water into the overlying wedge of mantle peridotite. This addition of H2O may cause melting to produce
siliceous magmas, such as andesites. (3) At substantially
deeper levels in the mantle, localized heating (probably by
heat generated through crystallization of the inner core)
at hot spots generates alkalic magmas. Magmas underplating the crust can also induce melting in the overlying
crustal rocks by local increase in temperature.
The most common mechanism by which magmas are
differentiated is fractional crystallization: the physical
separation of crystals and liquid so that these phases can
no longer maintain equilibrium. Crystals may be isolated from their parent magma through gravity settling
(or rarely, floating), or they may be carried to the bottom
of a magma chamber and deposited by convection currents. A pile of accumulated crystals may be purified further by squeezing out of some of the interstitial liquid
as the overburden of additional crystals accumulates.
Rapid crystallization, in which the reaction of melt with
solid phases cannot keep up with falling temperature and
advancing solidification, can produce compositionally
zoned crystals. This kinetic isolation of the interiors of
crystals prevents their equilibration with the liquid and
thus can also be considered as fractional crystallization.
Unlike equilibrium crystallization and melting, fractional crystallization is not the reverse of fractional
melting. To see how fractional crystallization differs from
equilibrium crystallization, let’s consider two binary phase
diagrams that involve olivine.
Worked Problem 12.5
Fractional melting in the system forsterite-silica has already been
discussed in worked problem 12.4. Using the phase diagram for
this system in figure 12.8, trace the fractional crystallization
paths of melts having compositions x and y.
Cooling of any melt composition to the liquidus causes the
appropriate solid phase to appear. If we begin with a liquid of
composition x, removal of enstatite (En) crystals as they form
drives the residual liquid composition toward the eutectic. When
this point is reached, both En and crystobalite (Cr) separate
from the melt. The liquid trajectory is, in this case, no different
from that for equilibrium crystallization, but in fractional crystallization, the bulk composition of the magma changes because
244
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
so that this reaction cannot take place. Hence, the liquid crystallizes a range of olivines that are progressively more Fe-rich,
from composition b to d and beyond. Again, under conditions
of perfect fractional crystallization, the liquid should ultimately
reach the composition of pure fayalite, but in real systems we
cannot specify the exact point at which the last liquid will solidify, only that the liquid composition continues past point c.
FIG. 12.14. Phase diagram for the system forsterite (Mg2SiO4)fayalite (Fe2SiO4), illustrating how fractional crystallization
affects a solid solution series. Olivine crystals forming from a
melt with initial composition a will range in composition from
b to some point past d.
crystals are removed. Under equilibrium conditions, these crystals remain suspended in the magma and become part of the
resulting erupted lava and igneous rock.
Now consider a liquid having composition y, shown in the
lower inset of figure 12.8. The first solid phase to form is now
forsterite (Fo). Under equilibrium conditions, the liquid should
follow the Fo liquidus to intersect the peritectic, at which point
it reacts with already crystallized Fo to produce En. Under
fractional crystallization conditions, Fo is physically removed
from the system as it forms, so that there is none of this phase
to react with the melt at the peritectic temperature. Thus, there
is no Fo + L → En reaction to cause the liquid to delay at the
peritectic. The cooling liquid slides through the peritectic and
immediately begins to crystallize En rather than Fo. The melt
composition continues toward the eutectic. Under conditions of
perfect fractional crystallization, the liquid reaches the eutectic,
but such ideal conditions do not exist in nature, so that the
liquid may be exhausted at some point before the eutectic is
reached.
Olivine that crystallizes from real magmas is, of course, not
pure forsterite, but a solid solution of Mg- and Fe-rich end members. The phase diagram for the system forsterite (Fo)-fayalite
(Fa), shown in figure 12.14, is very similar to one for the plagioclase system that was used in chapter 10 to illustrate solid
solution behavior. Crystallization of a liquid of composition a
in figure 12.14 initially produces olivine of composition b. If
equilibrium were achieved, this olivine would then react continuously with the melt to form more Fe-rich compositions,
ultimately having composition d, as the last drop of liquid is
exhausted. Under fractional crystallization conditions, however,
the initially formed Mg-rich olivine is removed from the melt,
Fractional crystallization affects ternary systems in an
analogous manner. Reaction curves, the ternary equivalents of binary peritectic points, no longer define the compositions of fractionating liquids, because there are no
solid phases to react. Fractional crystallization in a more
complex system is described in worked problem 12.6.
Worked Problem 12.6
Phase relationships in the system forsterite-anorthite-silica are
shown in figure 12.15. This problem will not make sense to
you, unless you have studied the box on ternary phase diagrams
in chapter 10. (Fig 12.15 actually represents a pseudoternary
rather than a ternary system, because the primary phase field of
spinel appears in the diagram, but the composition of spinel lies
outside of this plane. The only complication that this situation
presents is that we cannot specify the crystallization path of
liquids within the spinel field, because we do not know where
FIG. 12.15. Phase diagram for the pseudoternary system
forsterite (Mg2SiO4)-anorthite (CaAl2Si2O8)-silica (SiO2). The
compositions of liquids derived by fractional crystallization of
melt compositions x and y are illustrated by dashed lines.
The Solid Earth as a Geochemical System
Consideration of these experimentally derived phase
diagrams provides some indication of the ways in which
magmas might differentiate, but what is the evidence
that fractional crystallization actually occurs as a widespread phenomenon in magmatic systems? To answer this
question, we must examine real igneous rocks for the
effects of this process. Experimental studies of most
basaltic rocks indicate that they are multiply saturated
at low pressures; that is, several minerals begin to crystallize at virtually the same temperature. We would expect that kind of behavior for partial melts formed at
low pressure, because melting begins at invariant points
(such as eutectics, the lowest temperature points at which
liquids are stable). However, basaltic magmas are mantle derived and could not have formed at low pressures,
so multiple saturation must have another explanation.
That most basalts are already multiply saturated at low
pressure suggests that they have experienced fractional
crystallization.
It has also been recognized that basalts cluster at
certain preferred compositions. This observation is difficult to interpret if these are primitive magmas, unless
the preferred compositions represent those of liquids in
equilibrium with mantle peridotite at invariant points.
An alternative (and more plausible) view is that these
2.9
MORB
Fractionation Path
2.8
Density (g cm-3)
the projection of spinel plots on this diagram.) We have already
applied Alkemade’s Theorem to identify the various subtraction
and reaction curves (single- and double-headed arrows, respectively) in this diagram. Describe fractional crystallization paths
for liquids of composition x and y in figure 12.15.
Liquid x cools to the Fo liquidus and then changes in composition away from the forsterite apex as this phase is removed.
When the liquid composition intersects the subtraction curve,
Fo + An crystallize and are separated out together. The liquid
moves to and through the tributary reaction point, because
there is no Fo left in contact with melt to react to form En. The
liquid continues toward the ternary eutectic point between En,
An, and Tr (tridymite). We cannot specify the exact point where
the liquid phase will be exhausted.
A liquid with composition y will fractionate Fo until it reaches
the curve for the reaction Fo + L → En. Because there is no Fo
left to react, the melt steps over this curve into the En primary
phase field. At this point, En begins to crystallize, and the liquid composition moves directly away from the enstatite composition. Continued En fractionation drives the melt to the
subtraction curve, and An joins the fractionating assemblage.
The liquid then follows the subtraction curve toward the ternary eutectic.
245
2.7
Window of
eruptibility
2.6
MORB
2.5
2.4
0
0.2
0.4
0.6
Fe/(Fe+Mg)mol
0.8
1.0
FIG. 12.16. The density of fractionating basaltic magma as a
function of composition. MORB compositions cluster at the minimum density, suggesting that only those compositions can ascend
through the crust and erupt. (After Stolper and Walker 1980.)
compositions represent points along a fractionation path
that have special properties that allow the magmas to
erupt. Edward Stolper and David Walker (1980) observed
that the density of basaltic liquids first decreases and
then increases during fractional crystallization, as illustrated in figure 12.16. In this diagram, fractionation is
indicated by the progressive increase in the Fe/(Fe + Mg)
ratio of the lavas. The commonly observed composition
for MORB corresponds to that which has the minimum
density. This fractionated composition would be most
likely to ascend through the less dense crust and reach
the planet’s surface.
Widespread fractional crystallization has obvious
ramifications for what we can hope to learn about the
mantle from igneous rocks. Crystallization experiments
on fractionated rocks serve no purpose, except to understand the fractionation path itself. Luckily, a few
unfractionated lavas somehow reach the surface, so
judicious selection of samples for experiments can still
provide useful information on mantle composition and
melting.
246
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Chemical Variation Diagrams
The chemical changes wrought by fractional crystallization can be modeled quantitatively by using mixing
calculations, which are most easily visualized from element variation diagrams. Figure 12.17 illustrates a
magma’s chemical evolution, called the liquid line-ofdescent, as it undergoes fractional crystallization. In figure 12.17a, crystals of composition A separate from
liquid L1, with the result that the liquid composition is
driven toward L2. In figure 12.17b,c, mixtures of two
or three minerals crystallize simultaneously. The bulk
composition of the extracted mineral assemblage is given
in each case by A, and the liquid line-of-descent follows
the path L1 → L2. In figure 12.17d, magma L1 first crys-
tallizes mineral D and then D plus E, with the bulk
composition of the fractionating assemblage represented
by A. In this case, the liquid line-of-descent contains a
kink, where the additional phase joins the crystallization
sequence. A more complex case in which the fractionating phase is a solid solution is illustrated in figure 12.17e.
Mineral AB is a solid solution of end members A and B,
and the proportions of these end members change as
temperature drops during crystallization. The trajectory
of the liquid at any point is defined by a tangent to the
liquid composition drawn from the composition of crystallizing AB. For example, the liquid path when A80B20
and A50B50 are extracted is shown by the tangents intersecting these compositions. In this case, the liquid line-ofdescent traces out a curve. In all of these diagrams, the
relative proportions of fractionating phases can be determined by application of the lever rule.
In actual practice, the problem must be worked in
reverse. A suite of volcanic rocks representing various
points along a liquid line-of-descent is analyzed, and
from these analyses, we must determine the fractionated
phases. To test possible solutions to this problem, we
require rock bulk compositions and the identities and
compositions of possible extracted minerals. The latter
may be obtained from petrographic observations and
electron microprobe analyses of phenocrysts in the rocks.
An example of such an exercise is given in worked problem 12.7.
Worked Problem 12.7
FIG. 12.17. Schematic chemical variation diagrams (axes can be
any two elements or oxides) showing the liquid line-of-descent
(L1 → L2) during fractional crystallization of one (a), two (b), or
three (c) phases. In (d), the line-of-descent has an inflection due
to late entry of a second phase. In (e), the line-of-descent is
curved because of continuous change in the composition of an
extracted solid solution phase.
How can we test for fractional crystallization in real rocks? Assume that we have collected a suite of volcanic rock samples
that we believe to be a fractionation sequence. The first step is
to obtain bulk chemical analyses of the rocks. Petrographic observations indicate that three phases—olivine (Ol), plagioclase
(Plg), and clinopyroxene (Cpx)—exist as phenocrysts in this suite
of these rocks, and we have also chemically analyzed these minerals. Because these three phases were obviously on the liquidus
at some point, we determine whether fractional crystallization
of any one or some combination of them could have produced
the observed liquid line of descent.
Figure 12.18 shows three different variation diagrams with
the liquid line-of-descent (L1 → L2) defined by the analyzed bulkrock compositions. Also shown are the compositions of the
three phenocryst minerals. The information gained from any
individual diagram is ambiguous, but we can compare diagrams
to obtain a unique answer. In figure 12.18a, we see that the liquid line-of-descent is consistent with fractionation of Ol + Plg,
Ol + Cpx, or Ol + Plg + Cpx. This diagram thus rules out Cpx
The Solid Earth as a Geochemical System
247
These programs compute the composition of the parent
magma by mixing the residual liquid with fractionating
phases in various proportions. The answer is a best fit to
the least-squares regression for the analyzed liquid lineof-descent.
Liquid Immiscibility
FIG. 12.18. Schematic chemical variation diagrams for a suite
of basalts containing phenocrysts of olivine (Ol), clinopyroxene
(Cpx), and plagioclase (Plg). The liquid line-of-descent defined by
bulk chemical compositions of these basalts can be derived by
fractional crystallization of Ol + Cpx.
+ Plg as a possibility. In figure 12.18b, Ol + Cpx or Ol + Plg
+ Cpx are permissible assemblages that could be extracted to
produce this sequence, but Ol + Plg (allowed in fig. 12.18a), is
not allowed. The final distinction between Ol + Cpx and Ol
+ Plg + Cpx is made using figure 12.18c. This diagram indicates
that fractionation of Ol + Cpx is the only internally consistent
answer. The graphical testing of all interrelationships between
elements is an important part of this kind of study; otherwise,
we could have incorrectly assumed that all of the observed
phenocrysts can fractionate to produce this liquid line-ofdescent. Notice also that the relative proportions of olivine and
clinopyroxene (approximately 1:1) in all the diagrams are the
same, as deduced from application of the lever rule to the point
at which the extrapolated liquid line-of-descent crosses the
Ol-Cpx tie-line. Not only the identity of the fractionating
phases but also their relative proportions should be consistent
in this set of variation diagrams.
As the number of possible fractionating phases or
analyzed elements increases, it becomes more difficult
to apply graphical methods. Numerical methods offer a
real advantage in such situations. Several computer
programs, commonly called petrologic mixing programs,
are available for performing calculations of this type.
Fractional crystallization is not the only mechanism
by which magmas can differentiate. Experimental petrologists have recognized for half a century that the binary
systems CaO-SiO2, FeO-SiO2, and MgO-SiO2 contain
fields in which silicate melts separate spontaneously into
two liquids. The phase diagram for the forsterite-silica
system is shown in figure 12.8; a stability field for two
liquids is shown near the SiO2 compositional end member. This region of unmixing behaves exactly like the
solvus we introduced earlier in chapter 10, except that
it occurs above the solidus. Liquid phase separation, or
immiscibility, can cause magmatic differentiation, provided that the two melts have different densities. In
such a case, globules of the more dense liquid tend to coalesce and sink to the bottom of the magma chamber, in
a manner similar to the segregation of crystals by gravity
fractionation.
Liquid immiscibility obviously indicates highly nonideal mixing of melt components. However, it may be
helpful to recognize that the separation of an immiscible
liquid fraction is really no different than the formation
of a crystal, at least in the thermodynamic sense. In both
cases, the original melt becomes saturated with respect
to a new phase as the temperature decreases. However,
there is a difference in the ease with which equilibrium
can be achieved in these two situations. Crystal growth
and separation of an additional liquid phase both proceed by diffusion of ions through the melt to the sites
of growth and across the interface. Diffusion in liquids
is generally many orders of magnitude faster than in
crystals, so immiscible melts are more likely to maintain
equilibrium with falling temperature. Consequently, a
cooling magma may form zoned plagioclase crystals
under nonequilibrium conditions while at the same time
segregating an immiscible liquid fraction under equilibrium conditions.
Although examples of silicate liquid immiscibility are
known from melting experiments on simple systems, most
of these experiments take place at very high temperatures and have geologically unreasonable compositions
248
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
(see, for example, the very high temperature and high SiO2
content of immiscible liquids in the system forsteritesilica in fig. 12.8). An experimentally known example of
silicate liquid immiscibility that may have a natural analog is in the system fayalite (Fe2SiO4)-leucite (KAlSi2O6)silica. This two-liquid field is a reasonable compositional
analog for immiscibility observed in some uncommon
dike rocks (lamprophyres) and some residual liquids
formed during the last stages of basalt crystallization
(mesostasis). It is unlikely, however, that silicate liquid
immiscibility produces large volumes of differentiated
magmas, because these compositions are only produced
near the end of fractional crystallization sequences, when
little liquid is left.
Liquid immiscibility seems to be most effective when
the structures of the two melts are radically different.
This can be achieved most readily if melt components
with nonsilicate compositions are considered. One important example involves the separation of carbonate
and silicate melts in CO2-bearing systems. A photomicrograph of carbonatite globules suspended in alkali basalt
is illustrated in figure 12.19. Immiscibility in such systems
has been documented experimentally and almost certainly accounts for the production of the carbonatite
magmas that are sometimes associated with alkalic silicate magmas. Another example is the separation of immiscible sulfide melts from basaltic magmas on cooling.
The sulfide melts are more dense than silicate magma and
readily sink. This may be an important mechanism for
producing some sulfide ore deposits.
THE BEHAVIOR OF TRACE ELEMENTS
Much of the gemstone industry depends on the chance
incorporation of minute quantities of certain elements
into otherwise “ho-hum” minerals. Small proportions
of foreign impurities can turn ordinary corundum into
valuable rubies or sapphires. For the geochemist, elements present in trace concentrations can serve another
purpose—to test and quantify geochemical models.
Trace elements are commonly defined as those occurring in rocks in concentrations of a few tenths of a
percent or less by weight. The mixing behavior of trace
components in crystals or melts is often highly nonideal;
consequently, different minerals may concentrate or
exclude trace elements much more selectively than they
do major elements. Trace element distributions can there-
FIG. 12.19. Photomicrograph of immiscible carbonatite globules
in alkali basalt. Two globules are coalescing. This sample is from
the Kaiserstuhl volcanic complex, Germany. Width of the photograph is ∼4 mm.
fore provide quantitative constraints on the processes
of partial melting and differentiation that cannot be
deduced from consideration of major elements.
Trace Element Fractionation during Melting
and Crystallization
In melt-crystal systems under equilibrium conditions,
elements are partitioned among phases according to their
activities in those phases. For trace elements, we normally assume Henry’s Law behavior, ai = hi Xi . The distribution coefficient, KD, derived in chapter 9, is given by
KD = concentration in mineral/concentration in liquid.
Trace element distribution coefficients are based on the
assumption that the trace component obeys Henry’s
Law in the phase of interest. Experimental studies suggest that the linear Henry’s Law region for many trace
elements extends at least to several wt %, so it is appli-
The Solid Earth as a Geochemical System
cable for the concentration ranges measured in natural
rocks. In principle, KD can be measured from partially
crystalline experimental charges or from natural lavas
containing phenocrysts and glass. Many of the available
distribution coefficients were determined from natural
rocks, but there is unfortunately no assurance that these
rocks reached equilibrium. There are also analytical problems in obtaining concentrates of pure phases (that is,
crystals with no adhering or included glass, and vice
versa), because trace elements are sometimes below the
detection limits for in situ measurements. Another problem with KD values is that they are dependent on the
compositions of the phases. In other words, their distributions are nonideal, and enough information to determine the appropriate activity coefficients is usually not
in hand. Despite these potential pitfalls, distribution coefficients are widely used to solve geochemical problems,
and their compositional dependence can be minimized
by selecting KD values for phases similar in composition
to those in the problem of interest.
The bulk distribution coefficient D is the KD value adjusted for systems containing more than one solid phase.
It is calculated from the weight proportions ω of each
mineral φ by:
D=
Σω K
φ
Dφ.
(12.3)
D must be calculated if we intend to handle models for
partial melting of mantle peridotite or for fractionation
of several minerals. Using the bulk distribution coefficient, we can predict trace element distributions during
various magmatic processes.
We now derive expressions that describe changes in
the concentrations of trace elements during solidification
of a magma. We consider two ideal cases: equilibrium
crystallization and perfect fractional crystallization (the
latter is sometimes called Rayleigh fractionation). We
assume that more than one phase is crystallizing.
In the case in which equilibrium between crystals and
liquid is maintained,
Xi solid /Xi melt = D.
(12.4)
If a fraction of magma crystallizes, the fraction of melt
remaining F is given by:
0 ,
F = nmelt /n melt
(12.5)
0
where n melt
is the number of moles of all components
in the original magma and nmelt is the total number of
249
moles remaining after crystallization commences. The
ratio of solid to melt remaining is therefore:
0
− nmelt)/nmelt = 1/F − 1.
(n melt
(12.6)
If we define yi as the number of moles of trace component
i in the system, equations 12.4 and 12.6 can be combined to give:
0
− nmelt)]/(yi melt /nmelt )
Xi solid /Xi melt = [yi solid /(n melt
= D.
Because yi solid = yi0melt − yi melt,
0
D = [(yi0melt /yi melt) − 1][(nmelt /(n melt
− nmelt)].
0 /n
gives:
Rearranging and dividing by n melt
melt
0
(yi0melt /n melt
)/(yi melt /nmelt ) =
0
{D[(n melt
− nmelt )/nmelt ] + 1}n/n0.
Substituting equation 12.5 into the expression above
yields:
Xi melt /Xi0melt = 1/(D − FD + F).
(12.7)
Expression 12.7 gives the concentration of trace element
i in the liquid as a function of the original concentration,
the bulk distribution coefficient, and the fraction of
liquid remaining for a system undergoing equilibrium
crystallization. The same expression also holds for equilibrium melting.
Now let’s see how this differs from fractional crystallization. Using the same symbols as above, the mole fraction of trace component i in the system is Xi = yi /n. If
fractionation of a mineral containing i occurs, then n becomes n − dn, and y becomes y − dy. The compositions
of solid and melt are therefore:
Xi solid = dy/dn
(12.8a)
and
Xi melt = (y − dy)/(n − dn).
(12.8b)
If trace component i obeys Henry’s Law, then:
Xi solid = dy/dn = D Xi melt.
(12.9)
However, dy/dn can also be expressed in terms of Xi melt.
If we neglect dy in comparison with much larger y
and dn in comparison with n in equation 12.8b, we see
that Xi melt can be approximated as y/n and y as n(Xi melt).
250
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Differentiating the latter expression with respect to n
gives:
dy/dn = n(dXi melt /dn) + Xi melt.
Substitution of this expression for dy/dn into equation 12.9 gives:
Xi melt /X i0solid = 1/D(1 − F)(1/D)−1.
which can be recast into the form:
(12.10)
Because Xi changes during crystallization, but not necessarily linearly, the way we obtain the total change in
concentration is to sum all the individual steps. In other
words, we integrate equation 12.10 between Xi0 and Xi
and between the initial and final moles of melt n0 and n:
ln(n/n0) = [1/(D − 1)] ln(Xi melt /Xi0melt),
or
Xi melt /Xi0melt = (n/n0 )D−1.
Because (n/n0 ) equals F, the fraction of liquid remaining:
Xi melt/Xi0melt = F D−1.
Xi melt /X i0solid = 1/(D − FD + F),
(12.12)
and for fractional melting:
DXi melt = n(dXi melt /dn) + Xi melt,
(1/Xi melt[D − 1]) dXi melt = (1/n) dn.
Using the same principles, analogous expressions
can be derived for partial melting models. These are given
here without derivation. For equilibrium melting:
(12.11)
Thus equations 12.7 and 12.11 describe the behavior of
trace elements under the extreme conditions of equilibrium and fractional crystallization, respectively.
(12.13)
Because fractional melting involves continuous removal
of melt as it is generated, this process probably does
not occur in nature, but it can be considered as an end
member for intermediate melting processes. Geochemical
models for magma generation have become very complex and are beyond the scope of this book; Shaw (1977)
gives an excellent account of this topic.
The equations we have just derived can be expressed
equally well in units of concentration other than mole
fractions, as long as the distribution coefficients are defined in the same terms. In fact, most trace element analyses are expressed in units of weight, for example, parts
per million (ppm), and we can substitute these concentration units for mole fractions in equations 12.7 and
12.11 if D is expressed as a weight ratio.
The changes in trace element concentration for
residual liquids under equilibrium and fractional crystallization are illustrated graphically in figure 12.20. Curves
FIG. 12.20. Changes in trace element concentrations in residual liquids for different values of KD
during (a) equilibrium crystallization and (b) Rayleigh fractionation. Element concentrations are
given in terms of mole fractions of each element i relative to its concentration in the parent
magma. F is the fraction of liquid remaining. (After Wood and Fraser 1977.)
The Solid Earth as a Geochemical System
labeled KD = 0 are limiting cases, and trace element
enrichments greater than these cannot be produced by
crystallization processes alone.
Compatible and Incompatible Elements
Earlier, we noted that it is useful to distinguish between
trace elements that are compatible in the crystallographic sites of minerals and those that are incompatible
with such minerals. Compatible elements have KD >>1,
whereas incompatible elements have KD << 1. Consequently, compatible elements are preferentially retained
in the solid residue on partial melting or extracted from
the liquid during fractional crystallization; incompatible
251
elements exhibit the opposite behavior. This can readily
be seen in figure 12.20. Of course, as new minerals begin
to melt or crystallize during the evolution of a magma, a
particular trace element may alter its behavior. For example, phosphorus may act as an incompatible element
during magmatic crystallization until the point at which
the phosphate mineral apatite starts to form, after which
it behaves compatibly. For applications involving mantlederived magmas, it is often convenient to define compatible and incompatible behavior in terms of the minerals
that make up mantle rocks—olivine, pyroxenes, garnet,
and spinel. Incompatible elements are generally those with
ionic radii too large to substitute for the more abundant
elements, or those with charges of +3 or higher.
PREDICTING TRACE ELEMENT BEHAVIOR: TRANSITION METALS AND RARE EARTH ELEMENTS
Chemical physics provides powerful ways to predict
trace element behavior. We illustrate this approach by
considering several types of transition elements, which
are characterized by inner d or f atomic orbitals that
are incompletely filled by electrons. Unpaired electrons
in transition elements are responsible for the distinctive colors of the minerals that contain them, as well
as the magnetism of minerals. First we see how chemical bonding models can predict relative distribution
coefficients.
Elements of the first transition series (Sc, Ti, V, Cr,
Mn, Fe, Co, Ni, and Cu) have incompletely filled 3d
orbitals, and cations are formed when the 4s and some
3d electrons are removed. Divalent ions in this series
are compatible in common ferromagnesian minerals
because of similarities to Mg2+ in ionic size and
charge. However, the degree of compatibility differs
markedly and can be explained by the effect of crystal structure on d orbitals. Crystal-field theory assumes
that electrostatic forces originate from the anions
surrounding the crystallographic site in which a transition metal ion may be located, and that these control its substitution in the structure.
A transition metal ion has five 3d orbitals, which
have different spatial geometries (illustrated in fig. 2.7).
In an isolated ion, the d electrons have equal probability of being located in any of these orbitals, but
they attempt to minimize electron repulsion by oc-
cupying orbitals one electron at a time. For ions with
more than five d electrons, this is obviously not possible, so a second electron can be added to each orbital by reversing its spin.
When a transition metal ion is placed in a crystal
(for example, in octahedral coordination with surrounding anions), the five 3d orbitals are no longer
degenerate; that is, they no longer have the same energy. Electrons in all five orbitals are repelled by the
negatively charged anions, but electrons in the dz 2
and dx 2−y 2 orbitals, which are oriented parallel to
bonds between the cation and the anions around it,
are affected to greater extent than those in the other
three. In octahedral coordination, therefore, these two
orbitals will have a higher energy level. The reverse is
true for ions in tetrahedral coordination.
The energy separation between the two groups of
orbitals is called crystal-field splitting. This is illustrated in figure 12.21. Ions with electrons in only the
lower-energy orbitals for a particular site geometry
are stable in that site, whereas electrons in the higherenergy orbitals destabilize the ion. Of course, the number of electrons depends on which transition metal
the cation is made from and on its oxidation state, so
some transition elements prefer octahedral sites more
than others do. For example, the predicted octahedral site preference energies for divalent cations decrease in the order Ni > Cu > Co > Fe > Mn, and for
252
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 12.21. Sketch of the energies of the five d orbitals in
transition metal ions (refer to fig. 2.7 for geometric representations of these orbitals). When the free ion is placed in an
octahedral mineral site, its total energy increases. Some orbitals have higher energy because they experience greater
electron repulsion; the magnitude of the energy difference between the two groups of orbitals is the crystal-field splitting.
trivalent cations, Cr > V > Fe. As olivine or spinel
(which offer octahedral sites) crystallizes from a
magma, we expect to see rapid uptake of Ni and Cr
in these minerals relative to other compatible transition metals. Tetrahedral sites in minerals produce a different but predictable order of cation site preferences.
Thus, crystal-field effects provide an explanation for
the differing KD values for transition metals. The geochemical and mineralogical applications of crystalfield theory were treated comprehensively by Roger
Burns (1970).
The lanthanide or rare earth elements, commonly
abbreviated as REE, include La, Ce, Pr, Nd, Pm, Sm,
Eu, Gd, Tb, Dy, Ho, Er, Tm, Yb, and Lu. The REEs
find wide application in metallurgy, the coloring of
glass and ceramics, and the production of magnets.
In this part of the periodic table, increasing atomic
number adds electrons to the inner 4f orbitals, so this
is a somewhat similar case to the transition metals just
discussed. The f electrons in REE, however, do not
participate in bonding. The valence electrons occur in
higher-level orbitals, and loss of three of these results
in a +3 charge for most REE cations. These are large
ions; their size, coupled with their high charges, makes
them incompatible.
Europium (Eu) is unique in that its third valence
electron does not enter the 5d orbital as in other REEs,
but a 4f orbital, which then becomes exactly half full.
This is a particularly stable configuration, and more
energy is required to remove the third electron. A
consequence of this is that under reducing conditions,
such as on the Moon or within the Earth’s mantle,
Eu may exist in both the divalent and trivalent states
(the ratio of Eu2+ and Eu3+ depends on the oxidation
state of the system). Eu2+ has about the same size and
charge as Ca2+, for which it readily substitutes, so
europium, at least in part, may behave as a compatible element.
Under oxidizing conditions, such as in the marine
environment, cerium (Ce) is oxidized to Ce4+. This
results in a significant contraction in its ionic radius;
thus, Ce4+ has a very different behavior from trivalent
ions of the other REEs. Cerium in the oceans precipitates in manganese nodules, consequently leading to a
dramatic Ce depletion in seawater. Marine carbonates
formed in this enviroment mimic this geochemical
characteristic. Metalliferous sediments deposited at
midocean ridges also exhibit Ce depletions, pointing
to a seawater source for the hydrothermal solutions
that produced them.
The natural abundances of REEs vary by a factor
of thirty. To simplify comparisons, REE concentrations
in rocks are normalized by dividing them by their
concentrations in another rock sample. Chondritic
meteorites, which are thought to contain unfractionated element abundances, are the most commonly used
standard for REE normalization. The geochemical
significance of these meteorites is considered in chapter 15. Shales also generally have uniform lanthanide
patterns, and several sets of shale REE abundances are
used for normalization in some applications. As we
have already discussed earlier in this chapter, shales
have geochemical significance because their compositions may represent that of the average crust. Normalization of REE data using chondritic meteorites or
shales produces relatively smooth patterns that enable
comparisons between samples, accentuating relative
degrees of enrichment or depletion and contrasting
behavior between neighboring elements in the lanthanide series. Normalized REE concentrations are
conventionally shown in diagrams such as that shown
in figure 12.22, plotted by increasing atomic number.
The elements on the left side of the diagram thus have
lower atomic weights and are called the light rare
The Solid Earth as a Geochemical System
earth elements (LREE) to distinguish them from their
heavier counterparts (HREE) on the right. What is
important from the standpoint of geochemical behavior, however, is that the lanthanides show a regular contraction in ionic radius from La3+ to Lu3+.
Except for Eu2+ in reduced systems and Ce4+ in
oxidized systems, the REEs are all incompatible but,
like the first series transition metals, their behavior
varies in degree. Some minerals can tolerate certain
REEs more easily than others, as shown in figure
12.22. For example, apatite can readily accommodate all rare earth elements. Some minerals discriminate between different rare earth elements, resulting
in their fractionation. Garnet has a phenomenal preference for HREEs because of their smaller ionic radii,
and pyroxenes show a similar but less pronounced
affinity. The extra Eu in plagioclase, accounting for
the spike or positive europium anomaly, is Eu2+ substituting for calcium. Thus KD values for the REEs in
various phases are controlled by size rather than by
bonding characteristics.
Studies of REEs have found many uses in geochemistry. They constitute important constraints on
models for partial melting in the mantle and magmatic
fractionation. The lanthanides are insoluble under
conditions at the Earth’s surface, so their abundances
in sediments reflect those of the average crust. The
short residence times for REEs in seawater also allow
them to be used to track oceanic mixing processes.
More exhaustive treatments of REE geochemistry
can be found in Henderson (1983) and Taylor and
McLennan (1987).
A knowledge of compatible and incompatible trace
element distribution coefficients can provide a powerful
test for geochemical models involving partial melting or
crystallization. Some representative KD values for trace
elements in minerals in equilibrium with basaltic magma
are presented in table 12.4. An example of such a test is
given in worked problem 12.8.
We are now in a position to evaluate how the crustmantle system has become stratified in terms of its heatproducing radioactive elements. Potassium, uranium, and
thorium are all very large ions and behave as incompat-
253
FIG. 12.22. Typical distribution coefficients for REEs between
various minerals and basaltic melt. REE concentrations are
normalized to values in chondritic meteorites. All of these
phases are intolerant of REE (that is, KD <1) except apatite
(and garnet for the HREEs). The positive Eu anomaly in plagioclase occurs because some of this element is in the divalent
state and can substitute for Ca2+ ions. (After Zielinski 1975.)
ible elements. Consequently, any partial melts produced
in the mantle scavenges these elements, and over time,
they become depleted from mantle residues and concentrated in the crust by ascending magmas. This differentiation of heat-producing elements has had a profound
effect on the Earth’s thermal history.
Worked Problem 12.8
How can trace element abundances in basalts be used to constrain models for their origin? A geochemical study of volcanic
254
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
TABLE 12.4. Representative Trace Element KD Values for Minerals in Equilibrium
with Basaltic Magma
Mineral
Olivine
Clinopyroxene
Orthopyroxene
Garnet
Amphibole
Plagioclase
Ni
Rb
Sr
Ba
Ce
Sm
Eu
Y
10
2
4
0.04
3
0.01
0.001
0.001
0.001
0.001
0.3
0.07
0.001
0.07
0.01
0.001
0.5
2.2
0.001
0.001
0.001
0.002
0.4
0.2
0.001
0.1
0.003
0.02
0.2
0.1
0.002
0.3
0.01
0.2
0.5
0.07
0.002
0.2
0.01
0.3
0.6
0.3
0.002
0.3
0.05
4.0
0.5
0.03
From Cox et al. (1979).
rocks from Reunion Island in the Indian Ocean by Robert
Zielinski (1975) provides an excellent example. Measured trace
element abundances in eight samples are shown in figure 12.23.
All of these samples contain phenocrysts of olivine, clinopyroxene, plagioclase, and magnetite in a groundmass of the same
minerals plus apatite. Zielinski used the major element compositions of whole rocks and phenocrysts to construct a fractionation model. Using a petrologic mixing program, he found that
lava samples 3–8 could be formed by fractionation of these
phenocrysts from assumed parental magma 2. Basalt 1 was not
a suitable parental magma for this suite, even though it has the
highest Mg/Fe ratio; this rock probably contains cumulus olivine
and pyroxene. Zielinski then tested this model using trace
elements.
The chondrite-normalized rare earth element patterns
(fig. 12.23a) are parallel and increase in the order sample 1 to
8. This is just as expected; incompatible REEs should have
higher concentrations in residual liquids, and none of the solid
phases fractionates LREEs from HREEs appreciably. Using the
KD values shown in figure 12.22 and the proportions of fractionating phases from the mixing program, it is possible to calculate the bulk distribution coefficient (equation 12.3) for each
sample. The mixing program also gives the fraction of liquid F.
Assuming equilibrium between the entire separated solid and
melt, substitution of these values into equation 12.7 gives the
ratio Ci melt/C i0 melt. Assume, as Zielinski did, that rock 2 is
the parental magma; its trace element concentrations are then
C i0 melt. Thus, we can calculate the expected REE pattern for
liquids derived from sample 2. When Zielinski did this exercise,
he was able to duplicate the measured REE patterns fairly accurately. The negative Eu anomaly in rock 8 is caused by the
removal of large quantities of plagioclase at the end of the
crystallization sequence.
The abundances of some other trace elements in these rocks
are shown in figure 12.23b. Uranium and thorium are even
more incompatible than REEs and increase progressively in the
sequence. Barium is also incompatible, but it and strontium can
substitute in sodic plagioclase, so both drop towards the end of
the sequence as feldspar fractionation becomes dominant. In
FIG. 12.23. Chondrite-normalized REE patterns (a) and other
trace element variations (b) for a suite of volcanic rocks analyzed
by Zielinski (1975). The patterns are consistent with a model
whereby samples 3–8 were formed by progressive fractional
crystallization of a liquid having the composition of sample 2.
Sample 1 appears to be a cumulate rock.
The Solid Earth as a Geochemical System
contrast, nickel and chromium are compatible and show marked
depletions in fractionated rocks. Their contents in rock 1 are
much higher than expected for a primary magma, and this observation lends support to the idea that this rock contains
cumulus phases. Using appropriate D and F values, we could
also predict the abundances of these elements in the hypothetical fractionation sequence. Zielinski was also able to model
these fairly well.
Even though Zielinski’s modeling may appear convincing,
we should note that this is not a unique solution to this problem. The same geochemical patterns might also be produced by
various degrees of partial melting of mantle peridotite. Rock 8
would represent a magma formed during a small amount of
melting. As melting advanced, the incompatible elements in the
melt phase would be progressively diluted, ultimately to produce magma 2. Actually, Zielinski recognized this possibility,
but argued against it on the basis of field evidence for shallow
differentiation. Like many other tools of geochemistry, trace
element modeling may give ambiguous answers, and is at its best
when used in combination with other investigative methods.
VOLATILE ELEMENTS
The crust and mantle also interact geochemically by the
exchange of volatile elements. Hydrogen, oxygen, carbon, sulfur, and other volatiles in the Earth’s interior may
be bound into crystal structures, dissolved in magmas, or
exist as a discrete fluid phase. We use the term fluid because P-T conditions in the mantle and crust are such that
this phase is generally above its critical point, where the
distinction between liquid and gas is meaningless. Before
255
we can investigate how fluids are cycled within the
planet’s interior, we should examine their compositions.
Crust and Mantle Fluid Compositions
The compositions of fluids in the crust and mantle can
be ascertained from several different lines of evidence.
The first is direct chemical analysis of waters from deep
wells in geothermal fields such as the Salton Sea, California. In chapter 7, we showed how computer programs
such as EQ6 can be used to model the chemical evolution of natural waters. Such sophisticated methods have
been used to verify that the measured compositions of
these fluids are in equilibrium with the metamorphic minerals in core samples, supporting the idea that these are
contemporary metamorphic fluids. It is often difficult,
however, to assess the extent to which such fluids may
have been contaminated by meteoric water or drilling
muds, and sampling of fluids from depths of more than
a few kilometers is out of the question.
Fluid inclusions trapped in the minerals of crustal and
mantle rocks (both igneous and metamorphic) provide
another method of direct observation. Although these
inclusions contain only minute quantities of fluids, they
can sometimes be extracted for chemical analysis. More
commonly, their compositions are inferred from heating
and freezing experiments in situ. The experimental manipulation of fluid inclusions in considered further in an
accompanying box.
EXTRACTING INFORMATION FROM FLUID INCLUSIONS
Fluids are commonly stranded in interstitial spaces as
minerals crystallize. They may be trapped along old
crystal boundaries, in fractures, or within pore spaces.
With time, these interstitial spaces gradually change
shape to minimize their surface energies. The result
is a population of more or less equally dimensioned
spherical fluid inclusions. Because the minimal surface
energy may be best attained by adapting inclusion
shape to the structure of the host mineral, some inclusions have negative crystal shapes. At the time of
trapping, the fluid is usually a single phase liquid or
supercritical fluid. The fluid shrinks on cooling, because the coefficient of thermal expansion for the host
mineral is much less than for the fluid. As the pressure in inclusions drops below the vapor pressure of
the fluid components, vapor bubbles nucleate and
grow. Most fluid inclusions, therefore, contain both a
liquid and a vapor phase. Solid phases are also commonly encountered in inclusions that have become
saturated with respect to one or more salts on cooling. These daughter crystals may be valuable clues
to the compositions of the parent fluid. It is also
common, particularly in sedimentary environments,
to find inclusions containing a second, immiscible
fluid (usually composed mostly of organic matter).
Many of these features are illustrated in the inclusion
256
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 12.24. Photomicrograph of a fluid inclusion in fluorite
from southern Illinois. The inclusion measures ∼200 mm in its
longest dimension, and consists of a NaCl-rich brine and a gas
bubble. Note that the inclusion has a shape that mimics the
host fluorite crystal.
in figure 12.24. The evolution of such an inclusion is
summarized in figure 12.25, which resembles the phase
diagram for water.
Experimental manipulation of fluid inclusions
can provide constraints on fluid composition and the
conditions of trapping. Phase transformations that
occur with changing temperature can be observed by
fitting a heating stage to a microscope, in effect observing the cooling process (fig. 12.25) in reverse. As
an inclusion is heated, daughter crystals redissolve
and the vapor bubble shrinks and finally disappears.
The temperature of homogenization is an approximate measurement of the temperature at which the
fluid was trapped, provided that the fluid was very
near its boiling point. If the trapped fluid was supercritical, its trapping temperature may be estimated
from the homogenization temperature and an independent estimate of confining pressure. Fluid inclusion
geothermometry is based on the assumption that the
original sealed fluid was a single, homogeneous phase;
an observational test of this assumption is that all inclusions trapped simultaneously should have the same
bulk composition and therefore the same volumetric
ratio of phases at room temperature.
The fluids trapped in large inclusions can be sampled by drilling into them and extracting them with a
micropipette. Their compositions can be determined
directly by various analytical techniques. Fluids in
small inclusions present a much greater challenge. If
only one generation of inclusions is present, the fluid
they contain may be released by crushing, but some
information can also be gleaned from nondestructive
observations. Like the salt used to melt ice on roads
in winter, the salinity of the fluid depresses its freezing point relative to pure water. Salinity can therefore
be determined by measuring the fluid’s freezing point.
This can be done on the same microscope stage used
for the heating experiments by passing liquid nitrogen
through a cooling jacket on the stage. Freezing point
depression is a function of total salinity and the identity of the salt in solution. Generally, it is assumed that
inclusions are primarily NaCl solutions, and salinity
FIG. 12.25. P-T diagram showing the evolution of a fluid inclusion. The fluid is trapped at A. Because it occupies a space
of constant volume, cooling causes the fluid pressure to fall
from A to B. At B, the fluid becomes saturated with vapor, and
a bubble nucleates and grows as the inclusion cools along the
liquid-vapor boundary. The inclusion becomes saturated with a
solute at C, and a daughter crystal (of, say, halite) nucleates
and grows with further cooling.
The Solid Earth as a Geochemical System
is calculated in terms of wt % NaCl equivalent. More
sophisticated studies involve spectroscopic measurements of inclusions to provide further information on
fluid compositions. Daughter crystals may be identified by using optical properties, such as color, pleochroism, and crystal habit, as well as melting points.
As with most geochemical techniques, there are
many potential pitfalls in fluid inclusion work. Leakage of fluid into or out of inclusions after trapping
is always of great concern. Care must be taken to
identify multiple generations of inclusions by optical
methods, because each generation will presumably
A third method of determining fluid compositions is
by studying metamorphic assemblages that depend on
fluid composition. Characterization of fluids in this case
is based on the thermodynamic principles that we considered in chapter 9. For example, for the equilibrium
reaction:
→ Al SiO + 3SiO + H O,
Al 2Si4O10(OH)2 ←
2
5
2
2
pyrophyllite
andalusite quartz fluid
we know that:
µAl 2Si 4O10(OH)2, pyroph = µAl 2SiO5, and + 3µ SiO2, qtz
+ µ H 2O, fluid .
Evaluation of the chemical potentials in the solid
phases allows calculation of µH O, fluid . This can be con2
verted into fluid composition (XH 2O, fluid) if the relationship between chemical potential and composition is
known from an appropriate equation of state for the
fluid. Worked problem 12.9 illustrates how this is done.
257
have its own specific composition and trapping temperature. Recrystallization of the host mineral may
cause several generations of inclusions to coalesce
into large inclusions, or separation of large inclusions
into smaller ones (necking). This latter process is
critical if inclusions consist of more than one phase,
because the phases may not be redistributed uniformly. These potential problems can be overcome in
most cases, however, so that fluid inclusions studies
offer a unique source of information on fluids from
the Earth’s interior. Shepherd et al. (1985) provide
practical information on how to study fluid inclusions.
At equilibrium,
0
+ RT ln aAn, plag) + ∆Ḡ0Cc + RT ln aCc, calcite
3(∆ḠAn
0
+ ∆ḠH
2O
+ RT ln aH O, fluid =
(12.12)
2
0
0
+ RT ln aCO , fluid .
+ RT ln a Zo, zoisite ) + ∆ḠCO
2(∆Ḡ Zo
2
2
In this equation, the ∆Ḡi0 terms refer to standard free energy
values at P and T. By recalling that:
(∆Ḡi0 )P,T = (∆Ḡi0 )1,T +
∫ ∆V̄
P
1
i
0 dP,
we can rewrite equation 12.12 as:
0 + ∆Ḡ 0 + ∆Ḡ 0
0
0
(3∆ḠAn
Cc
H O − 2∆Ḡ Zo − ∆Ḡ CO )
2
+
∫
P
1
2
0 + V̄ 0 − 2V̄ 0 )dP
(3V̄An
Cc
Zo
3
2
+ RT ln [(a An,
plag aCc, calcite)/a Zo, zoisite ]
(12.13)
− RT ln fCO + RT ln fH O = 0.
2
2
This is an equation of the general form:
0
∆Ḡ 0 + ∆V̄solids
(P − 1) + RT ln K − RT ln fCO
2
+ RT ln fH O = 0.
(12.14)
2
The first two terms in this equation can be evaluated from:
Worked Problem 12.9
John Ferry (1976) determined the composition of fluids during
regional metamorphism in south-central Maine. Let’s examine
his calculations for zoisite-bearing rocks, in which the following reaction occurs:
2 Ca2 Al3Si3O12(OH) + CO2 = 3 CaAl 2Si 2O8 + CaCO3
zoisite
anorthite
calcite
+ H 2O.
0
0
∆Ḡ 0 + ∆V̄solids
(P − 1) = ∆H 0 − T∆S̄ 0 + P∆V̄solids
,
because the difference between P and P − 1 is negligible in geologic problems. P and T can be estimated from mineral assemblage stability fields or from geothermometers and barometers.
The compositions of coexisting minerals can be used to determine K; however, in this case, we require two reactions to be
solved simultaneously for the two unknowns fCO and fH O.
2
2
For zoisite + calcite + plagioclase rocks in Maine, Ferry estimated metamorphic P and T values of 3500 bar and 798–710 K,
258
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
respectively. The ratio fCO /fH O can be determined from the
2
2
reaction discussed above. Using estimated thermodynamic data
and K values obtained from coexisting mineral compositions at
various temperatures, Ferry determined fCO /fH O ratios ranging
2
2
from 0.53 to 5.10 throughout his field area. These rocks coexist
with other assemblages that can be used to estimate either fCO
2
or fH O, and thus both variables can be determined by solving
2
simultaneous equations.
During prograde metamorphism, devolatilization reactions
commonly release enough fluid that we can safely assume the
condition:
Pfluid = PH
2O
+ PCO = Ptotal .
2
It follows that:
PH
2O
+ PCO = fH
2
2O
/γH
2O
+ fCO /γCO = 3500 bar. (12.15)
2
2
If we assume that the gas mixture was ideal,
fi = fi 0 X i = γi0 PXi = γi0 Pi .
(12.16)
From these relationships, Ferry determined values for XCO fluid
2
ranging from 0.06 to 0.32 for his assemblages. This range of
values indicates that gradients in fluid composition existed on
an outcrop scale, possibly due to mixing of several fluids.
The compositions of fluids determined from fluid
inclusions and calculated from metamorphic equilibria
can generally be represented by the system H-O-C-S,
although chloride complexes may also be present. The
major species in naturally occurring fluids are H 2O,
CO2 , CH 4, and H 2S. Although the species O2 and S2 actually occur in vanishingly small quantities, it is common
practice to characterize fluid compositions by defining
the fugacities of these species. These variables can serve
as useful monitors for oxidation and sulfidation processes
in such fluids. The relative proportions of these species
shift constantly as intensive variables for the system
change. Figure 12.26 illustrates how the species in a fluid
of fixed composition alter in response to changing
temperature.
Fluids in the upper crust are typically H 2O-rich except in carbonate rocks, where CO2 may predominate.
Granulites from the middle and lower crust have fluid
inclusions dominated by CO2 and, to a lesser extent, by
CO and CH4. The most abundant volatile species in
fluid inclusions from mantle xenoliths is CO2, although
some inclusions from the same specimens may be dominated by H 2O. Magmatic gases in tholeiitic basalts have
higher H 2O/CO2 ratios than those from alkaline lavas
that may have formed by melting at deeper levels.
FIG. 12.26. The effect of temperature on speciation of fluids in
the system H-C-O-S coexisting with graphite at 2 kbar pressure.
The fugacities of O2 and S2 are fixed by appropriate buffers.
(Based on calculations by Holloway 1981.)
Mantle and Crust Reservoirs for Fluids
The abundances of volatiles in the mantle are difficult
to specify, but analyses of dissolved fluids and fluid inclusions in basalts provide some constraints. For example,
the water content of the MORB source region has been
estimated at 140–350 ppm, whereas the deeper mantle
sources for ocean island basalts are slightly wetter, with
450–525 ppm water. The solubility of water and other
volatile species in most upper mantle minerals is low or
modest. However, it has been shown experimentally that
β-olivine (wadsleyite) and γ-olivine (ringwoodite) can
accommodate as much as 2 wt % water. Some hydrous
magnesian silicates, such as serpentine and chlorite, are
also stable at pressures appropriate for the lower reaches
of the upper mantle. The Mg-perovskite that dominates the lower mantle cannot contain much water. Thus,
it is possible that the mantle is stratified in terms of its
water storage capacity, with a relatively wet zone at
350–650 km depth sandwiched between drier upper and
lower mantle rocks.
In metamorphic rock systems within the crust, it is
common to express the proportion of fluid in terms of
The Solid Earth as a Geochemical System
a fluid pressure (Pfluid) relative to lithostatic pressure
(Plithostatic). The upper and lower crusts are characterized
by different hydrologic regimes that are thought to be
determined by rock strength. Strong upper crustal rocks
can maintain networks of open pores, so that Pfluid is
independent of Plithostatic and tends to follow the local
hydrostatic gradient. As a consequence, crustal fluids
can flow downward or upward between these pores
and thus will circulate vertically and horizontally. Lower
crustal rocks, however, tend to deform plastically, so
that porosity can be maintained only if the pores are
filled with fluids at or close to Plithostatic. Fluids in the
lower crust, then, tend to follow the lithostatic pressure
gradient and therefore always migrate upward. Following this logic, fluids in the lower crust are ephemeral.
Cycling of Fluids between Crust and Mantle
Volatile elements are readily transported as dissolved
fluid species in magmas. Significant quantities of mantlederived fluids are expelled from magmas erupting at
midocean ridges and subduction zones. The movement
of a fluid in this way depends critically on its solubility
in magma. H2O is very soluble for all magma compositions, but CO2 solubility depends more critically on
magma composition and pressure. Most magmas become
saturated with respect to volatiles during ascent, causing
the formation of a discrete fluid phase at some point.
Once a fluid phase has formed, all of the dissolved volatiles partition between it and the magma, and the fluid
phase is likely to escape into the surrounding crust or
atmosphere. Magmas provide transport for volatiles in
only one direction—upward.
Metamorphism occurs on the seafloor near midocean
ridges, as warm basaltic lavas react with seawater. Fluids trapped in these rocks are ultimately recycled into the
mantle within subducted slabs. At depths appropriate
for the eclogite transformation, dehydration reactions
occur and fluids are released. These may rise into the
mantle wedge above the subducted plate and promote
melting. The resultant magmas dissolve these volatiles
and carry them again into the crust. Alternatively, some
fluids in subducted slabs may endure in stable hydrous
minerals, which can be sequestered for a time near the
boundary between the upper and lower mantles.
Regional metamorphism is associated with collisions
between continents. The sedimentary rocks caught in this
predicament contain considerable quantities of crustal
259
fluids. High-grade metamorphism can drive off most of
the fluid component, and the effect of recycling these
volatiles into the atmosphere may be dramatic. For example, the collision of India and Asia ∼50 million years
ago was coincident with the warmest epoch in the Cenozoic. Kerrick and Caldeira (1993) suggested that metamorphism liberates huge quantities of CO2 (>1018 moles
per million years) by such reactions as CaCO3 + SiO2 →
CaSiO3 + CO2. Expulsion of CO2 into the atmosphere
then caused global warming by the greenhouse effect.
In contrast, Selverstone and Gutzler (1993) observed that
high-pressure metamorphic rocks in the Alps contain
carbonates and graphite that was buried to depths of
>50 km during the Alpine orogeny. In this case, carbon
remained sequestered in the lower crust, and they argued
that global cooling was the result. Although the net effect of metamorphism on fluid cycling is debated, it is
clear that metamorphic reactions at compressional plate
boundaries can cycle fluids into and out of the crust.
Near the beginning of this chapter, we noted that
granulites may constitute a major part of the lower crust.
The CO2-rich fluid inclusions that characterize granulites,
as well as the evidence for low aH 2O, suggest that these
rocks may form when the lower crust is invaded by CO2.
It is possible that such CO2-dominated fluids are due to
mantle degassing, possibly providing another mechanism
(streaming from mantle to crust, independently of plate
tectonics) for the cycling of fluids. During metamorphism,
continuous fluxing with CO2 must take place, because
dehydration reactions produce water that would otherwise dilute the fluid. Significant quantities of CO2 are
required over the metamorphic episode, but the actual
amount of CO2 in the rocks at any time may be small,
because the fluids occur as films at grain boundaries.
SUMMARY
The compositions of the crust, mantle, and core are so
different that it may be tempting to surmise that they
do not interact, but we have seen that geochemical exchange does occur on a global scale. Ascending magmas
formed by partial melting of mantle peridotite carry mantle components into the crust. Mantle melting is promoted by local temperature increases, depressurization,
or fluxing by fluids. Incompatible and volatile elements
are partitioned into these melts, so that they are ultimately
deposited in the crust. The geochemical record in igneous
rocks is commonly obscured by fractional crystallization
260
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
or other processes that alter magma compositions during ascent to the surface.
Crustal materials are carried downward in subducted
plates, where they may be stranded indefinitely. However, much of the volatile element content in these slabs
may be recycled because of dehydration or decarbonation reactions during metamorphism in the mantle. Only
oceanic crust is subducted; the amount of buoyant
continental crust appears to have increased episodically
with time.
With this background in the processes and pathways
by which materials in the Earth’s interior interact, we are
now in a position to explore another tool for formulating and testing geochemical interactions in these regions.
In the following chapters on isotopes, we build on the
information just presented.
suggested readings
The wealth of excellent books and papers on the subjects in this
chapter makes selection of a few difficult. We recommend the
following literature enthusiastically.
Anderson, D. L. 1989. Theory of the Earth. Oxford: Blackwell.
(This thoughtful book weaves together numerous constraints
from geophysics and geochemistry to provide much detail on
the composition of the Earth’s mantle.)
Carlson, R. W. 1994. Mechanisms of Earth differentiation:
Consequences for the chemical structure of the mantle.
Review of Geophysics 32:337–361. (A superb review of
constraints on mantle geochemistry and the processes that
account for its variations.)
Cox, K. G., J. D. Bell, and R. J. Pankhurst. 1979. The Interpretation of Igneous Rocks. London: Allen and Unwin. (Chapter 6 gives a superb account of element variation diagrams,
and chapter 14 discusses trace element fractionation.)
Ferry, J. M., ed. 1982. Characterization of Metamorphism
through Mineral Equilibria. American Reviews in Mineralogy 10. Washington, D.C.: American Mineralogical Society.
(An up-to-date manual of the quantitative geochemical
aspects of metamorphism; chapters 6 [by J. M. Ferry and
D. M. Burt] and 8 [by D. Rumble III] provide very readable
information on fluids in metamorphic systems.)
Fyfe, W. S., N. J. Price, and A. B. Thompson. 1978. Fluids in the
Earth’s Crust. Amsterdam: Elsevier. (Chapter 2 summarizes
the compositions of naturally occurring fluids, and other
chapters describe their generation and transport.)
Hargraves, R. B., ed. 1980. Physics of Magmatic Processes.
Princeton: Princeton University Press. (A high-level but very
informative book about the quantitative aspects of magmatism; chapter 4 [by S. R. Hart and C. J. Allegre] presents
trace element constraints on magma genesis, and chapter 5
[by E. R. Oxburgh] describes the relationship between heat
flow and melting.)
Henderson, P. 1983. Rare Earth Element Geochemistry. Developments in Geochemistry 2. Amsterdam: Elsevier. (A book
devoted to research developments in the field of REEs;
chapter 4 [by L. A. Haskin] provides an in-depth look at
petrogenetic modeling using these elements.)
Jeanloz, R. 1990. The nature of the Earth’s core. Annual Reviews
of Earth and Planetary Sciences 18:357–386. (An interesting review of geophysical research bearing on the properties
of the core; this paper contains some controversial ideas
about reactions at the core-mantle boundary).
Philpotts, A. R. 1990. Principles of Igneous and Metamorphic
Petrology. Englewood Cliffs: Prentice-Hall. (Chapter 22 of
this superb petrology text gives an up-to-date explanation of
how melting occurs in the mantle, and magmatic processes
are treated in chapter 13.)
Ringwood, A. E. 1975. Composition and Petrology of the
Earth’s Mantle. New York: McGraw-Hill. (An exhaustive,
although now somewhat dated, monograph on the nature
of the mantle; chapter 5 describes the pyrolite model, and
later chapters describe experimental constraints on mantle
mineralogy.)
Silver, P. G., R. W. Carlson, and P. Olson. 1988. Deep slabs,
geochemical heterogeneity, and the large-scale structure of
mantle convection: Investigation of an enduring paradox.
Annual Reviews of Earth and Planetary Sciences 16:477–
541. (A very thoughtful review of geophysical and geochemical constraints on the mantle.)
Taylor, S. R., and S. M. McClennan. 1985. The Continental
Crust: Its Composition and Evolution. Oxford: Blackwell
Scientific. (Everything you might want to know about the
composition and origin of the Earth’s crust, presented in a
very readable manner.)
Walther, J. V., and B. J. Wood, eds. 1986. Fluid-Rock Interactions during Metamorphism. Advances in Geochemistry 5.
New York: Springer-Verlag. (Chapter 1 [by M. L. Crawford
and L. S. Hollister] provides a summary of fluid inclusion
research.)
Wood, B. J., and D. G. Fraser. 1977. Elementary Thermodynamics for Geologists. Oxford: Oxford University Press.
(Chapter 6 presents derivations of equations for trace
element behavior.)
Yoder, H. S. Jr., ed. 1979. The Evolution of the Igneous Rocks.
50th Anniversary Perspectives. Princeton: Princeton University Press. (An updated version of the classic treatise by
N. L. Bowen; chapters 2 [by E. Roedder] and 3 [by D. C.
Presnall] describe liquid immiscibility and fractional crystallization and melting, and chapter 17 [by P. J. Wyllie] considers magma generation.)
The following publications are also cited in this chapter and
provide much more detail for the interested student:
The Solid Earth as a Geochemical System
Anderson, D. L. 1980. The temperature profile of the upper
mantle. Journal of Geophysical Research 85:7003–7010.
Asimow, P. D., M. M. Hirschmann, M. S. Ghiorso, M. J.
O’Hara, and E. M. Stolper. 1995. The effect of pressureinduced solid-solid phase transitions on decompression
melting of the mantle. Geochimica et Cosmochimica Acta
59:4489–4506.
Burns, R. G. 1970. Mineralogical Applications of Crystal Field
Theory. Cambridge: Cambridge University Press.
Castillo, P. 1988. The Dupal anomaly as a trace of the upwelling lower mantle. Nature 336:667–670.
Ferry, J. M. 1976. P, T, fCO , and fH O during metamorphism
2
2
of calcareous sediments in the Waterville-Vassalboro area,
south-central Maine. Contributions to Mineralogy and
Petrology 57:119–143.
Holloway, J. R. 1981. Compositions and volumes of supercritical fluids in the earth’s crust. In L. S. Hollister and M. L.
Crawford, eds. Fluid Inclusions: Applications to Petrology.
Calgary: Mineralogical Association of Canada, pp. 13–38.
Hutchison, R. 1974. The formation of the Earth. Nature 250:
556–568.
Kerrick, D., and K. Caldeira. 1993. Paleoatmospheric consequences of CO2 released during early Cenozoic regional
metamorphism in the Tethyan orogen. Chemical Geology
108:201–230.
McBirney, A. R. 1969. Compositional variations in Cenozoic
calc-alkaline suites of Central America. Oregon Department
of Geology Mineral Industries Bulletin 65:185–189.
McLennan, S. M. 1982. On the geochemical evolution of sedimentary rocks. Chemical Geology 37:335–350.
Presnall, D. C., S. A. Dixon, J. R. Dixon, T. H. O’Donnell, N. L.
Brenner, R. L. Schrock, and D. W. Dycus. 1978. Liquidus
phase relations on the join diopside-forsterite-anorthite from
1 atm to 20 kbar: Their bearing on the generation and crystallization of basaltic magma. Contributions to Mineralogy
and Petrology 66:203–220.
261
Ringwood, A. E. 1962. A model for the upper mantle. Journal
of Geophysical Research 67:857–866.
Robie, R. A., B. S. Hemingway, and J. R. Fisher. 1978. Thermodynamic properties of minerals and related substances at
298.15 K and 1 bar (105 pascals) pressure and at higher
temperatures. Geological Survey Bulletin 1452. Washington, D.C.: U.S. Geological Survey.
Ronov, A. B., and A. A. Yaroshevsky. 1969. Chemical composition of the earth’s crust. In P. J. Hart, ed. The Earth’s Crust
and Upper Mantle. Monograph 13. Washington, D.C.: American Geophysical Union.
Ronov, A. B., N. V. Bredanova, and A. A. Migdisov. 1988. General compositional-evolutionary trends in continental crust
sedimentary and magmatic rocks. Geochemistry International 25:27–42.
Selverstone, J., and D. S. Gutzler. 1993. Post-125 Ma carbon
storage associated with continent-continent collision. Geology 21:885–888.
Shaw, D. M. 1977. Trace element behavior during anatexis.
Oregon Department of Geology Mineral Industries Bulletin
96:189–213.
Shepherd, T. J., A. H. Rankin, and D.H.M. Alderton. 1985. A
Practical Guide to Fluid Inclusion Studies. London: Blackie.
Stolper, E., and D. Walker. 1980. Melt density and the average
composition of basalt. Contributions to Mineralogy and
Petrology 74:7–12.
Taylor, S. R., and S. M. McLennan. 1987. The significance of
the rare earths in geochemistry and cosmochemistry. Handbook on the Physics and Chemistry of Rare Earths 13.
Amsterdam: North-Holland.
White, I. G. 1967. Ultrabasic rocks and the composition of
the upper mantle. Earth and Planetary Science Letters 3:
11–18.
Zielinski, R. A. 1975. Trace element evaluation of a suite of
rocks from Reunion Island, Indian Ocean. Geochimica et
Cosmochimica Acta 39:713–734.
262
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
PROBLEMS
(12.1) Describe the fractional crystallization path for a liquid with 45 wt. % SiO2 in the system forsteritesilica (see fig. 12.8).
(12.2) Describe the equilibrium crystallization path for a liquid with 72 wt. % SiO2 in the system forsteritesilica (see fig. 12.8).
(12.3) Describe the phase assemblages in the melting sequence under equilibrium and fractional melting
conditions for a composition of Fo50 in the system forsterite-fayalite (fig. 12.14).
(12.4) (a) Sketch qualitative chemical variation diagrams for CaO versus MgO and Al2O3 versus SiO2 for
liquid lines-of-descent during fractionation of compositions x and y in figure 12.15. (b) How would
these variation diagrams be affected if the system also contained FeO? (Hint: what effect would this
have on olivine and orthopyroxene?) (c) Describe the sequence of rocks produced by perfect fractional crystallization of these two compositions.
(12.5) A basaltic magma has the following trace element abundances: Ni, 1100 ppm; Rb, 100 ppm;
Sr, 300 ppm; Ba, 220 ppm; Ce, 82 ppm; Sm, 9.4 ppm; Eu, 3.3 ppm; and Yb, 3.3 ppm. Fractional
crystallization of 20 wt. % olivine, 12 wt. % orthopyroxene, and 10 wt. % plagioclase from this
magma results in a residual liquid that erupts on the surface. Using the distribution coefficients in
table 12.4, calculate the trace element contents of this erupted melt.
(12.6) The following are REE abundances in chondritic meteorites: Ce, 0.616 ppm; Sm, 0.149 ppm;
Eu, 0.056 ppm; and Yb, 0.159 ppm. Construct a chondrite-normalized REE pattern for the parent
magma and residual liquid in problem 12.5, following the example in figure 12.23. How would you
explain the general shapes of these patterns?
(12.7) A spinel peridotite has a solidus temperature of 1500°C at 20 kbar. (a) Using figure 12.9, describe the
phases involved in its melting as a mantle diapir of this material rises adiabatically to a depth of 40
km. (b) How does the liquid composition, in terms of normative olivine, change over this interval?
CHAPTER 13
USING STABLE ISOTOPES
OVERVIEW
HISTORICAL PERSPECTIVE
This is the first of two chapters on isotope geochemistry. In it, we discuss how geochemists use stable isotopes of hydrogen, carbon, nitrogen, oxygen, and sulfur
to interpret geologic processes and environments. You
may want to review portions of chapter 2 to refresh your
familiarity with the language and some of the basic concepts in nuclear chemistry. With the aid of several case
studies, we show how those and other concepts can be
applied to problems of practical interest. In each case
study, the common principle is mass fractionation. We
explain how the different stable isotopes of a single
element can separate from each other in a variety of
thermal or biochemical processes, and how we can trace
the chemical history of a system by measuring the
abundance ratios of these isotopes in coexisting phases.
Geochemists interpret some fractionation processes by
using the familiar tools of equilibrium thermodynamics. Among these processes are simple isotopic exchange
reactions that have been used successfully as geothermometers. Other isotope fractionations are kinetically controlled. In many cases, abundance patterns
produced by geologic and biologic processes can yield
insights into the pathways along which isotopic changes
have occurred.
Some of the most dramatic advances in the modern geologic sciences have been made in the field of isotope
geochemistry. It is a discipline with a surprisingly short
history, having attracted talented geochemists in large
numbers since only about 1950. The roots of the field
lie barely half a century before that, in the work of Henri
Becquerel and the Curies, Marie and Pierre. Their studies
of radioactivity led British geologist Arthur Holmes and
others to create the new field of geochronology, which
occupies much of our attention in chapter 14. That work
was well underway, however, before chemists knew much
about the structure of the atom.
Frederick Soddy hypothesized the existence of isotopes
in 1913, attempting to explain why some atoms of a
single element weighed more than others. Even with that
theoretical step forward, it was 1932 before the neutron
was discovered and Soddy’s hypothesis was confirmed.
Not by coincidence, a dramatically new emphasis in geochemistry began that same year with Harold C. Urey’s
discovery of deuterium (2H). For that discovery and for
showing that the isotopes of hydrogen vary in relative
abundance from one geologic environment to another,
Urey won the 1934 Nobel Prize. Thus opened the field of
stable isotope chemistry.
263
264
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
As we noted in chapter 1, Alfred O. C. Nier’s refinement of the mass spectrometer led to an explosion of
knowledge about isotopes in the years before World
War II. By the late 1940s, Urey and his students at the
University of Chicago were pioneers in the use of oxygen
isotopes for geothermometry and in an increasing number of provenance studies. Samuel Epstein and his students at Cal Tech further refined that work and applied
it to the study of ore deposits. Canadian geochemist
Harry Thode, working at McMaster University in the
1950s, developed techniques for extracting sulfur isotopes from ore minerals, opening yet another fruitful class
of studies in stable isotope geochemistry. In the ensuing
half-century, geochemists have used isotopes not only
of oxygen and sulfur, but also carbon, nitrogen, silicon,
and boron to trace chemical pathways in the Earth. It is
difficult nowadays to find a study of ore genesis, crustal
evolution, or global geochemical cycles that does not
include consideration of stable isotopes.
4.
5.
6.
7.
WHAT MAKES STABLE
ISOTOPES USEFUL?
Interpretations based on studies of stable isotopes depend on the ability of various kinetic and equilibrium
processes to separate (fractionate) light from heavy isotopes. By understanding the mechanisms involved, we
can measure the relative abundances of light and heavy
isotopes of an element in natural materials and interpret
the pathways and environments they have undergone.
Isotopes of H, C, N, O, and S share a set of characteristics that make them particularly well suited for this sort
of work:
1. They all have low atomic mass. Nuclides with Z > 16
generally do not fractionate efficiently in nature.
2. The relative difference in mass between heavy and
light isotopes of each of these nuclides is fairly large.
The difference is largest for hydrogen, whose heavy
isotope, deuterium (2H, also given the symbol D) is
twice as massive as the light isotope, 1H. At the other
extreme on our short list of stable nuclides, 34S is still
6.25% heavier than 32S. Because isotopes fractionate
on the basis of relative mass, this is an important
criterion.
3. All five elements are abundant in nature and constitute a major portion of common Earth materials. Oxy-
gen, in particular, makes up almost 47% of the crust
by weight (nearly 92% by volume).
Carbon, nitrogen, and sulfur exist in more than one
oxidation state and may therefore participate in processes over a wide range of redox conditions.
Each of the five elements forms bonds with neighboring atoms that may range from ionic to highly covalent. As we show shortly, fractionation of heavy
and light isotopes is greatest between phases that have
markedly different bond types or bond strengths. For
this reason, most other light elements like magnesium
or aluminum, which form the same type of bonds in
almost all common Earth materials, do not show
marked isotopic fractionation.
For each of the five elements, the abundance of the
least common isotope still ranges from a few tenths
of a percent to a few percent of the most common isotope. From an analytical point of view, this means
that the precision of measurements is quite high.
All are biogenic elements, and thus can be fractionated by biologic processes.
Geochemists commonly use three different types of
notation to express the degree of isotopic fractionation.
First, they describe the distribution of stable isotopes between coexisting phases A and B in terms of a fractionation factor, αA-B, defined by:
αA-B = RA/RB ,
(13.1)
in which R is the ratio of the heavy to the light isotope in
the phase indicated by the subscript. The fractionation
of 18O and 16O between quartz and magnetite, for example, would be indicated by the magnitude of:
αqz-mt = (18O/16O)qz /(18O/16O)mt .
Concentrations, rather than activities, are used in calculating fractionation factors, because activity coefficients
for isotopes of the same element are nearly identical and
would cancel each other. Values for αA-B are commonly
between 1.0000 and 1.0040 for inorganic processes and
higher for biologic processes.
Because it is inconvenient to use absolute isotopic
ratios (RA/RB ) and because αA-B values commonly differ
only in the third or fourth decimal place, geochemists
commonly use a second notation, δ (delta), instead. With
this convention, the isotopic ratio in a sample is com-
Using Stable Isotopes
pared with the same ratio in a standard, using the
formula:
δ = 1000 × (Rsample − Rstandard)/Rstandard.
(13.2)
265
and
δB = 1000 × (RB − Rstandard )/Rstandard,
so:
αA-B − 1 = (δA − δB )/(1000 + δB ).
The numerical value that results from this procedure is a
measure of the deviation of R in parts per thousand, or
per mil (‰, by analogy with percent) between the
sample and the standard. Samples with positive values
of δ are said to be isotopically heavy (that is, they are
enriched in the heavy isotope relative to the standard);
those with negative values are isotopically light.
For each isotopic system, laboratories have agreed on
one or two readily available substances to serve as standards. For hydrogen and oxygen, the universal standard
is Standard Mean Ocean Water (SMOW). In older studies, δ18O values were commonly referred to the isotopic
ratio in a Cretaceous marine belemnite fossil from the
Pee Dee formation in South Carolina. Today, the PDB
scale for 18O is only used in paleoclimatology studies.
Carbon isotopic measurements, however, are almost
always compared with 13C /12 C in the Pee Dee Belemnite, so most people think of PDB as a standard for
carbon rather than oxygen. (In the literature since the
1980s, SMOW and PDB are called V-SMOW and V-PDB,
referring to reference standards available from the International Atomic Energy Agency [IAEA] in Vienna. Comparisons between these standards, with earlier samples
of SMOW and PDB, and with other commonly used
oxygen standards can be found in Coplen et al. [1983].)
Nitrogen measurements are all reported relative to the
15N/ 14N ratio in air, and sulfur measurements are reported relative to the 34S/ 32S ratio in troilite (FeS) from
the Canyon Diablo meteorite, whose impact produced
Meteor Crater in Arizona. In practice, each laboratory
develops its own standards, which are then calibrated
against these universal standards.
Finally, geochemists also use the symbol ∆ to compare
δ values for coexisting substances. We can derive this
third quantity from either of the other systems of notation. From the definition of αA-B, we see that:
αA-B − 1 = (RA − RB )/RB.
From the definition of δ, we also see that:
δA = 1000 × (RA − Rstandard )/Rstandard
Because δB is small compared with 1000, we can write
this last result as:
1000(αA-B − 1) ≈ δA − δ B.
For light stable isotope pairs with fractionation factors on the order of 1.000–1.004 (all except D-H), it is a
very good approximation to write:
ln αA-B ≈ αA-B − 1.
(The natural logarithm of 1.004, for example, is 0.00399.)
With this approximation, we finally see that:
1000 ln αA-B ≈ δA − δ B = ∆A-B.
(13.3)
In this way, fractionation factors between coexisting minerals can be calculated from their respective δ values,
measured relative to a universal laboratory standard.
Worked Problem 13.1
Atmospheric CO2 has a δ13C value of −7‰. If HCO3− in a
sample of river water assumed to be in equilibrium with the
atmosphere has a measured value of +1.24‰, what are the equivalent values of ∆ and the fractionation factor αHCO −-CO ?
3
2
We can easily substitute these values into equation 13.3 to
see that:
δHCO − − δCO = +1.24‰ − (−7‰) = +8.24‰,
3
2
and
αHCO −-CO = exp(∆ HCO −-CO /1000) = 1.00827.
3
2
3
2
Fractionation factors and delta values can be derived
in several different ways, generally yielding compatible
results. Experiments involving isotopic exchange can be
performed in a laboratory setting, and fractionation can
be determined by analyzing the run products. In some
cases, geochemists can also use isotopic analyses from
pairs of natural materials if they have independent information about the conditions of isotopic exchange
between them. For many situations, however, values of
α are calculated from principles of statistical mechanics,
266
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
using spectroscopic data. Our approach to thermodynamics in this textbook has been macroscopic rather
than statistical, so we have not laid the foundation for
an in-depth discussion of this third technique. Because
it helps explain the thermodynamic driving force for
isotopic fractionation, however, let’s take a brief detour
into processes at the molecular level.
MASS FRACTIONATION
AND BOND STRENGTH
Fractionation among isotopes occurs because some
thermodynamic properties of materials depend on the
masses of the atoms of which they are composed. In a
solid or liquid, the internal energy is related largely to
the stretching or vibrational frequency of bonds between
the atom and adjacent ligands. The vibrational frequency
of a molecule, in turn, is inversely proportional to the
square root of its mass (see accompanying box). Therefore, if two stable isotopes of a light element are distributed randomly in molecules of the same substance, the
vibrational frequency associated with bonds to the lighter
isotope will be greater than that with bonds to the heavier one. As a result, molecules formed with the lighter
isotope will have a higher internal energy and be more
easily disrupted than molecules with the heavier one.
The effect of isotopic composition on the physical
properties of a phase can be surprisingly large, as you
can see in table 13.1. Molecules containing the lighter
isotope are more readily extracted from a material during such processes as melting or evaporation. In general,
then, the lighter isotope of pairs such as D-H or 18O-16O
tends to fractionate into a vapor phase rather than a liquid or into a liquid rather than a solid. This is the basis
for many kinetic mass fractionation processes.
By inverting this logic, we can see that an isotope,
placed in an environment with coexisting phases, will
fractionate between the phases on the basis of bonding
TABLE 13.1. Physical Properties of Water as a
Function of Isotopic Composition
Melting point
Boiling point
Maximum density
Viscosity
H216O
D216O
0.0°C
100.0°C
at 3.98°C
8.9 mpoise
3.81°C
101.42°C
at 11.23°C
10 mpoise
characteristics. Bonds to small, highly charged ions have
higher vibrational frequencies than those to larger ions
with lower charge. (Note that we are now comparing
bonds that have higher vibrational frequencies because
they have different “spring” constants, not because they
connect atoms with different masses.) Substances with
such bonds tend to accept heavy isotopes preferentially,
because doing so reduces the vibrational component of
their internal energy. This, in turn, reduces the energy
available for chemical reactions—the free energy of the
system—and makes the phase more stable.
The effect of substituting heavy isotopes into a substance with fewer small, highly charged ions is generally
less pronounced. There is a tendency, consequently, for
substances such as quartz or calcite, in which oxygen is
bound to small Si4+ or C4+ ions by covalent bonds, to
incorporate a high proportion of 18O. These minerals,
therefore, have characteristically large positive values of
δ18O. In feldspars, this tendency is less dramatic, largely
because some of the Si4+ ions in the silicate framework
have been replaced by Al3+. The contrast is even greater
in magnetite, where oxygen forms ionic bonds to Fe2+
and Fe3+, which are much larger and less highly charged
than Si4+. These bonds, therefore, have a lower vibrational frequency. Magnetite, therefore, tends to prefer the
lighter isotope, 18O. At the extreme, uraninite (UO2) is
commonly among the most 18O-depleted minerals found
in nature, as we might predict from the large radius and
mass of the U4+ ion.
GEOLOGIC INTERPRETATIONS BASED
ON ISOTOPIC FRACTIONATION
The distribution of stable isotopes among minerals and
fluids has proven to be a highly fruitful area of research
in geochemistry. In this section, we examine some of the
more prominent problems to which isotopic interpretations have been applied. This survey is not intended to
be comprehensive, but rather to suggest the wide variety
of geologic questions that can be addressed from the
perspective of stable isotope fractionation.
Thermometry
We can describe fractionation of stable isotopes between minerals by writing simple chemical reactions in
which the only differences between the reactants and the
Using Stable Isotopes
267
CHEMICAL BONDS AS SPRINGS
Recall from chapter 2 that bond strength is primarily
a function of the electronic interaction between atoms.
A bond is in many ways analogous to a simple spring.
Hooke’s Law tells us that a mass (m) attached to a
spring and moved a distance (x) experiences a restoring force:
Ca
Ca
Xmax
F = −kx,
in which k is a spring constant that describes the elasticity of the spring itself. The elasticity of a chemical
bond, similarly, is an inherent property. From basic
physics, we also know that if we displace a mass on a
spring, it will oscillate with a frequency:
( )
1
f = — π √
(k/m).
2
Isotopes of the same element have the same electronic configuration, so the bonds they form with
other atoms should have the same “spring” constant, k. Electronic contributions to the strength of a
Ca-16O bond and a Ca-18O bond, for example, are
virtually identical, so the bonds are equally elastic.
But because 18O is heavier, the Ca-18O bond has a
lower vibrational frequency, f.
Finally, elementary physics allows us to calculate
the maximum velocity of an oscillating mass on a
spring from:
Vmax = 2πxmax f = 2πxmax√
(k/m).
Imagine, then, that we place two different masses
(16O and 18O) on the ends of two springs (Ca-16O
and Ca-18O) with the same spring constant, k, and
stretch the two springs to the same length, xmax, before releasing them (fig. 13.1). A little algebra shows
that the one with the greater mass and lower vibra-
products are their isotopic compositions. Water and calcite, for example, may each contain both 16O and 18O,
so the isotopic exchange between them is given by:
→ CaC18O + H 16O.
CaC16O3 + H218O ←
3
2
The equilibrium constant, Keq, for this reaction is equal to:
16O
18O
FIG. 13.1. The bond between atoms in a molecule can be
modeled as a spring. The elasticity of a Ca-O bond is analogous to the spring constant, k, for a mechanical spring. It
depends on the electronic interaction between atoms and is
therefore insensitive to the mass of either atom. The vibration
frequency ( f ) of a Ca-16O bond is higher than the vibration
frequency of a Ca-18O bond, because f is proportional to
√
(k/m).
tional frequency, 18O, will also have the lower maximum velocity:
Vmax(18O)/Vmax(16O)= f18O /f16O = √
(m16O/ m
18O).
When the velocity of an oscillating mass is maximized, the internal energy of the spring-mass system
2 ) rather than pois all in kinetic energy (K = 1–2 mVmax
tential energy. It is easy to see that the kinetic energy
associated with the heavier isotope will be lower than
with the lighter one:
2 18
KCa-18O = 1–2 m18OVmax
( O)
2 (16O)(m
= 1–2 m18OVmax
16 /m18 )
O
O
2 16
= 1–2Vmax
( O)(m16O /m218O)
= KCa-16O /m218O.
Keq = aCaC18O3aH 216O /aCaC16O3aH 218O
= (18O/16O)CaCO3 /(18O/16O)H 2O
= α CaCO 3-H 2O.
The free energy change for this reaction is due only
to the very slight difference between the vibrational
268
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
energy states of a 16O-C bond and a 18O-C bond in calcite and between a 18O-H bond and a 16O-H bond in
water. Unlike the mineral reactions we have discussed so
far, in which ∆Ḡro is on the order of tens of kilocalories
per mole, the free energy associated with this exchange is
therefore only a few calories per mole. As written, the exchange reaction has a small negative ∆Ḡro, so αCaCO -H 2O
3
has a positive value.
In 1947, Harold Urey recognized that the isotopic
exchange between 16O and 18O could be used as the basis
of a paleothermometer to estimate the temperature of
the ancient ocean. At 25°C, the fractionation factor for
the calcite-water system has a value of 1.0286, which
means that δ18O is equal to +28.6‰, if the water we
have been talking about is SMOW. The value of α is a
function of temperature, such that with decreasing temperature, calcite becomes more 18O-enriched relative to
seawater. Urey hypothesized that if accurate determinations of δ18O could be made on marine limestones, it
should be possible to calculate the ocean temperature
at their time of limestone formation.
Despite early promise, this particular paleothermometer has only proven reliable for Cenozoic sediments, and
even then with great difficulty. Its most serious limitation is that the oxygen isotopic composition of seawater
has varied considerably through time, so that the isotopic fluctuations in carbonate sediments do not reflect
changes in temperature alone. Later in this chapter, for
example, we discuss kinetic effects that cause polar ice
to be isotopically lighter than mean seawater. As glacial
epochs have come and gone and the balance of 16O and
18O has shifted between ice and the oceans, the δ18O of
the oceans may have varied by as much as 1.4‰. The
diagram in figure 13.2 attempts to correct for this complication by offering two temperature scales: one for
ice-free periods and a second for periods (like the present)
when there is glacial ice on the Earth. Sea surface temperatures within the past 50,000 years can be validated
independently by methods such as the alkenone U37K
index that we discussed in chapter 6, but relating these
temperatures to the isotopic record in bottom sediments
remains a challenge. Other practical problems related to
diagenetic changes in the isotopic composition of limestone and the biochemical fractionation of oxygen by
shell-forming organisms (vital effects) have contributed
further complications to carbonate paleothermometry.
Although Urey’s original idea for a marine carbonate thermometer has turned out to be of limited value in
FIG. 13.2. Oxygen isotopic data from benthic foraminifera collected in the North Atlantic Ocean. The process of forming glaciers fractionates 16O from the oceans, leaving them relatively
richer in 18O when ice is present on the Earth than in ice-free
times (see fig. 13.6 and the accompanying text). Therefore, an
isotopic temperature scale based on δ18O, shown at the bottom,
has to include a correction for this fractionation for ice-free
times. For large portions of the mid-Tertiary, it is not easy to
determine how great a correction to apply. (Modified from Miller
et al. 1987.)
practice, the basic concept is a good one. Especially at
metamorphic or igneous temperatures, kinetic complications are rarer than at 25°C, and a large number of
silicate and oxide mineral pairs have been shown to
yield valid temperature estimates. Oxygen isotope thermometry is particularly attractive in studies of deep
crustal processes, because isotopic fractionation takes
place almost independently of pressure. This is a substantial advantage over the geothermometers that we
examined in chapter 9.
Figure 13.3 shows how a number of fractionation factors for oxygen isotopes between quartz and other rockforming minerals vary with temperature. These have been
calculated from spectroscopic data, using a quantum mechanical model for lattice dynamics. More commonly,
plots of this type are constructed from experimental
data. At temperatures >1000 K, it has been shown that
the natural log of α is proportional to 1/T 2, so it is con-
Using Stable Isotopes
269
experiments between plagiociase and water. By performing leastsquares fits of equation 13.4 to their data, they determined that:
1000 ln αqz-H
2O
1000 ln αab-H
2O
= 3.13 × 106 T −2 − 2.94,
= 2.39 × 106T −2 − 2.51,
and
1000 ln αan-H
2O
= 1.49 × 106T −2 − 2.81,
where T is in kelvins. These equations are strictly valid only
over the range 673–773 K, where the experiments were performed. Considering the likely uncertainties encountered when
using these curves to determine temperatures from natural
samples, however, it is probably safe to apply them somewhat
outside this range.
If we assume, as Matsuhisa and his coworkers did, that the
fractionation factor for an intermediate plagioclase varies in a
linear fashion with its anorthite content, then we can combine
the two feldspar-water equations to yield:
FIG. 13.3. Temperature dependence of oxygen isotope fractionation factors between quartz and common rock-forming minerals,
calculated from spectroscopic data. Abbreviations: Ab = albite;
An = anorthite; And = andradite; Calc = calcite; Di = diopside;
En = enstatite; Fo = forsterite; Gros = grossular; Musc = muscovite; Rut = rutile; Pyrp = pyrope; Zrc = zircon. (Modified from
O’Neil 1986.)
1000 ln αpl-H
2O
= (2.39 − 0.9 An) × 106T −2
− (2.51 + 0.3An).
To describe the fractionation between quartz and plagioclase,
we subtract this result from the quartz-water equation above.
The final expression is:
1000 ln αqz-pl = (0.74 − 0.9An) × 106T −2
− (0.43 − 0.3An),
which we can solve for temperature to get:
venient to construct plots like figure 13.3 on which equations of the form:
1000 ln α = A (106 T −2) + B
(13.4)
plot as straight lines. A and B are empirical constants.
For many pairs of geologic materials, this proportionality can be used to extrapolate to temperatures much
lower than 1000 K. Deviations are most obvious below
500 K, but are often small enough to make extrapolation
reasonable. The following problem is typical of many in
isotope geothermometry.
Worked Problem 13.2
The Crandon Zn-Cu massive sulfide deposit in northeastern
Wisconsin is a volcanogenic ore body within a series of intermediate to felsic volcanic rocks. Munha and coworkers (1986)
have sampled andesitic flows and tuffs adjacent to the ore body
and determined the oxygen isotope compositions of quartz
and plagioclase in them. How can these be used to estimate the
temperature at which the ore body formed?
Matsuhisa and colleagues (1979) performed a series of isotopic exchange experiments between quartz and water, and other
T = [(0.74 − 0.9An) × 106/(δ18Oqz − δ18Opl)
+ 0.43 − 0.3An]1/2
(13.5)
Munha and his colleagues collected two samples from the
footwall rocks at the Crandon deposit. Both show petrographic
evidence of propylitic alteration (characterized by the appearance of chlorite, serpentine, and epidote in hydrothermally
altered andesitic rocks), which commonly causes marked albitization of plagioclase. The mole fraction of anorthite in feldspars from these rocks is ∼0.15. In one sample, the measured
δ18O for quartz is +9.38‰ and for plagioclase is +6.71‰. By inserting these values in equation 13.5, we find that T = 533 K
(260°C). In the other sample, δ18Oqz = +9.53‰ and δ18Opl =
+ 6.44‰, from which we calculate that T = 503 K (230°C).
These are compatible with temperatures from 240° to 310°C
measured by independent techniques.
Isotope geothermometry is perhaps most heavily used
by economic geologists, for problems like the one shown
above. Sulfur and oxygen isotopes are commonly measured to determine the temperature of ore formation. A
few of the most popular sulfur isotope thermometers are
indicated in figure 13.4 and table 13.2. Where sulfide
270
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Isotopic Evolution of the Oceans
minerals can be shown to have formed in isotopic equilibrium with a hydrothermal fluid, these can provide
valuable insights into the conditions of mineralization.
Unfortunately, however, sulfur isotopic systems are easily
upset by kinetic factors. The oxidation state and pH of
ore-forming fluids, for example, can seriously affect the
rates of isotopic exchange between aqueous species and
ore minerals, particularly at low temperature.
0
100
200
Age (m.y.)
FIG. 13.4. Temperature dependence of sulfur isotope fractionation factors between minerals or fluid species and H2S. Solid
curves are experimentally determined, and dashed curves are
estimated or extrapolated. (Modified from Ohmoto and Goldhaber
1997.)
In chapter 8, we found that SO42−, at an average concentration of 2700 ppm, is the second most abundant
anion in seawater. The total mass of sulfur in the oceans
is ∼3.7 × 1021 g. These values have probably not changed
significantly since the Proterozoic era. The isotopic composition of oceanic sulfate, however, has varied considerably. Today, marine SO42− is very nearly homogeneous,
with a δ34S value of almost +20‰. In the past 600 million
years, that value has ranged from ∼10 to almost +30‰.
As illustrated in figure 13.5, these systematic variations,
reflected in the isotopic composition of sulfate evaporites,
offer promise as a stratigraphic tool.
The specific causes of variation are not well understood, but they are widely interpreted to represent a dynamic balance between redox processes that supply and
remove sulfur from the ocean. Sulfur enters the ocean
primarily as sulfate derived from a variety of continental
sources and dissolved in river water. Among these sources
are sulfidic shales and carbonate rocks, volcanics and
other igneous rocks, and evaporites. Sulfur leaves the
300
400
500
TABLE 13.2. Selected Sulfur Isotope Thermometers
Isotope Thermometer
Pyrite-galena
600
Temperature (K)
[(1.01 ± 0.04) × 103]/∆1/2
Sphalerite-galena
or
Pyrrhotite-galena
[(0.85 ± 0.03) × 10 ]/∆
Pyrite-chalcopyrite
[(0.67 ± 0.04) × 103]/∆1/2
Pyrite-pyrrhotite
or
Pyrite-sphalerite
[(0.55 ± 0.04) × 103]/∆1/2
3
∆ = δ34SA– δ 34 SB = 1000 ln αA–B. Data from Ohmoto and
Goldhaber (1997).
1/2
700
5
10
15
20
δ34S/32S
25
30
35
(per mil)
FIG. 13.5. The variation of the sulfur isotopic composition of
sulfate evaporites through the Phanerozoic era has been used to
estimate the isotopic composition of the oceans through time.
Most measured values lie within the heavy lines. Thin lines indicate the range of extreme values. (Modified from Holser and
Kaplan 1966.)
Using Stable Isotopes
ocean as sulfate evaporites or in reduced form as pyrite
and other iron sulfides in anoxic muds.
In general, sulfides are isotopically light compared to
sulfates. This is due, in part, to equilibrium fractionation
(see fig. 13.4), but is also a product of kinetic fractionation by bacteria. Metabolic pathways in sulfate-reducing
bacteria involve enzyme-catalyzed reactions that favor
32
S over 34S, because 32S-O bonds are more easily broken
than 34S-O bonds. The degree of fractionation depends
on sulfate concentration as well as temperature. Some
sedimentary sulfide is >50‰ lighter than contemporaneous gypsum in marine evaporites.
The isotopic composition of sulfate entering the
oceans, therefore, may vary with time, because of fluctuations in global volcanic activity, or because different
proportions of evaporates and sulfidic shales are exposed
to continental weathering and erosion. In the same way,
changes in the rate of deposition of pyritic sulfur in
marine sediments may affect the δ34S value in seawater
sulfate. Berner (1987) has shown that such changes may
be the indirect result of fluctuations in the mass and diversity of plant life during its evolution on the continents,
or may reflect changes in the relative dominance of euxinic and “normal” marine depositional environments.
By studies of this type, the grand-scale processes that
have moderated the global cycle for sedimentary sulfur
(and carbon and oxygen, which mimic or mirror the history of sulfur) are becoming well understood. The events
that have given rise to specific fluctuations in the marine
δ34S record, however, remain largely unknown.
Worked Problem 13.3
What is the maximum change in δ18O that could have occurred
in ocean water due to the weathering of igneous and metamorphic rocks to form sediments during the history of the Earth?
To answer this question, we need a few basic pieces of information. The average δ18O value for igneous rocks is ∼+8‰,
and that for sedimentary and metamorphic rocks is ∼+14‰.
These numbers are not very well constrained, but they are
good enough for a rough calculation. The mass of the oceans
(appendix C) is 1.4 × 1024 g, and the mass of sedimentary and
metamorphic rocks in the crust is ∼2.1 × 1024 g.
Assume that all of the sedimentary and metamorphic rocks
were once igneous, and therefore once had a δ18O value of
+8‰ (8/1000 heavier than the present ocean). Their δ18O value,
then, has since increased by +6‰. The complementary change
in the composition of the hydrosphere must have been:
(+6‰)(2.1 × 1024 g)/(1.4 × 1024 g) = +9‰.
271
The oxygen isotopic mix in the oceans, in other words, is gradually becoming lighter through geologic time as a result of crustal
weathering reactions.
Fractionation in the Hydrologic Cycle
Because there are two stable isotopes of hydrogen
(1H and 2D) and three of oxygen (16O, 17O, and 18O),
there are nine different ways to build isotopically distinct water molecules. Masses of these molecules range
from a low of 18 for 1H216O to a high of 22 for 2D218O.
Earlier in this chapter, we found that the vibrational frequency associated with bonds to a light isotope is greater
than that for bonds to a heavy isotope. Consequently,
“light” water molecules escape more readily from a
body of water into the atmosphere than do molecules of
“heavy” water. When water evaporates from the ocean
surface, it has a δD value of about −8‰ and a δ18O value
around −9‰. The degree of fractionation is, of course, a
function of temperature. The result at any place on the
Earth, however, is that atmospheric water vapor always
has negative δD and δ18O values relative to SMOW.
The reverse process occurs when water condenses in
the atmosphere. Condensation is nearly an equilibrium
process, favoring heavy water molecules in rain or snow.
The first rain to fall from a “new” cloud over the ocean,
therefore, has values for both δD and δ18O that are ∼0‰.
Water vapor remaining in the atmosphere, however, is
systematically depleted in deuterium and 18O by this
process. Subsequent precipitation is derived from a vapor
reservoir that has delta values even more negative than
freshly evaporated seawater. The isotopic ratio (R) in the
remaining vapor is given by:
R = Ro f (α−1),
(13.6)
where Ro is the initial 18O/16O value in the vapor, f is the
fraction of vapor remaining, and α is the fractionation
factor. This is the Rayleigh distillation equation, which
also appears in chapter 14 when we consider element
fractionation during crystallization from a melt. Let’s use
it in an example.
Worked Problem 13.4
Suppose that rain begins to fall from an air mass whose initial
δ18O value is −9.0‰. If we assume that the fractionation factor
272
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
is 1.0092 at the condensation temperature, what should we
expect the isotopic composition of the air mass to be after 50%
of the vapor has recondensed? What happens when 75% or
90% has been recondensed? If we were to collect all of the rain
that fell as the air mass approached total dryness, what would
its bulk δ18O value be?
First, we convert the R values to equivalent delta notation:
(δ + 1000) = 1000(R/RSMOW).
The Rayleigh distillation equation (13.6), therefore, can be written in the form
R/Ro = (δ18O + 1000)/([δ18O] o + 1000) = f (α−1),
or
δ18O = [(δ18O)o + 1000]f (α−1) − 1000.
After 50% recondensation,
δ18O = [−9.0‰ + 1000](0.5)(0.0092) − 1000 = − 15.3‰.
After 75%,
δ18O = [−9.0‰ + 1000](0.25)(0.0092) − 1000 = −21.6‰.
After 90%,
δ18O = [−9.0‰ + 1000](0.1)(0.0092) − 1000 = −29.8‰.
What happens to the isotopic composition of rainwater
during this evolution? To find out, we write:
FIG. 13.6. Contours of δD and δ18O (values in parentheses) in
rainwater become increasingly negative as moist air moves inland,
to higher elevations, or to higher latitudes. For this reason, ice
sheets during glacial periods lock up a disproportionate amount
of the Earth’s 16O, leaving the oceans relatively 18O-enriched.
(After Sheppard et al. 1969.)
α = Rrain /Rvapor = (δ18Orain + 1000)/(δ18Ovapor + 1000).
By rearranging terms, we find that:
δ18Orain = α(δ18Ovapor + 1000) − 1000.
We use the results of the first calculation to find that after 50%
condensation, δ18O of rain is:
1.0092(−15.3 + 1000) − 1000 = −6.2‰.
After 75%, δ18Orain = −12.69‰. After 90%, it reaches
−20.9‰. With continued precipitation, therefore, rainwater
becomes lighter. Notice, however, that if we collected all of
the rain that falls, conservation of mass would require that its
bulk δ18O value be the same as the initial δ18O value of the
water vapor.
Because of this fractionation process, δ18O values for
rainwater are always negative compared with values for
seawater. A similar fractionation affects hydrogen isotopes. For both isotopic systems, the separation between
rainwater and seawater becomes more pronounced as
air masses move farther inland or are lifted to higher
elevations. Also, because the fractionation factors for
both oxygen and hydrogen isotopes become larger with
decreasing temperature, the difference between seawater
and precipitation increases toward the poles. The lightest natural waters on the Earth are in snow and ice at the
South Pole, where δ18O values less than −50‰ and δD
less than −45‰ have been measured. These various
trends are illustrated for the North American continent
in figure 13.6.
These observations make it possible to recognize the
isotopic signature of meteoric waters in a particular region and to use that information to study the evolution
of surface and subsurface waters. Analyses of δ18O and
δD are commonly displayed on a two-isotope plot like
figure 13.7, originally proposed by Cal Tech chemists
Sam Epstein and Toshiko Mayeda (1953) and refined
through the work of Harmon Craig (1961) and others.
Meteoric waters on such a plot lie along a straight line
given by:
δD = δ18O + 10.
Two-isotope plots have been used for many interesting purposes. Shallow groundwaters tend to retain the
Using Stable Isotopes
FIG. 13.7. Oxygen and hydrogen isotopic values in meteoric
water vary with temperature and hence latitude, but always lie
along the diagonal meteoric water line in this two-isotope plot.
(Modified from Craig 1961.)
273
isotopic pattern they inherit from rain. Water in deep
aquifers, however, can deviate from local rainwater in
many ways. Because of climatic changes, for example,
the isotopic composition of modern rainwater may be
significantly different from that of water that fell thousands of years ago and is now buried in deep aquifers,
even though both lie on the meteoric water line. Siegel
and Mandle (1984) have sampled well waters drawn
from the Cambrian-Ordovician aquifer system in Minnesota, Iowa, and Missouri. They find that δ18O values
in the north, where the aquifer is exposed, are close to
modern rainwater values. To the south, however, δ18O
becomes steadily more negative. In Missouri, well water
is as much as 11‰ lighter than present precipitation.
Siegel and Mandle conclude, therefore, that a contour
map of δ18O values (fig. 13.8) can serve in this case as a
qualitative guide to rates of groundwater movement
during the past 12,000 years.
FIG. 13.8. Variation of δ18O in well waters drawn from the Cambrian-Ordovician aquifer system in
the north-central midwestern United States. Values are more negative to the southwest, where the
aquifer is deepest. These waters may have been introduced from glacial meltwater sources during
the late Pleistocene epoch. (Modified from Siegel and Mandle 1984.)
274
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
MASS-INDEPENDENT FRACTIONATION
For studies of the type we are highlighting in this
chapter, geochemists use the ratio of the two most
abundant isotopes of an element. When three or more
stable isotopes exist, however, they could just as easily
choose to use a less abundant pair. One reason that
they usually do not is that the various isotopes of any
element tend to fractionate in direct proportion to
their masses, so that a measurement of δ17O, for
example, offers no more information than a measurement of δ18O. To see why, take a look at figure 13.9.
Because 18O is two mass units heavier than 16O and
17
O is only one unit heavier, any process that alters a
mixture containing both isotopes should change δ18O
twice as much as it does δ17O. For this reason, all
samples of oxygen-bearing materials on the Earth
should lie on a mass fractionation line through the
composition of SMOW (δ17O = δ18O = 0.0‰) for
which δ17O = 0.5δ18O.
Imagine the surprise when geochemists at the University of Chicago (Robert Clayton et al. 1973) discovered that some meteorites contain high-temperature
FIG. 13.9. On this “three-isotope plot,” values of δ17O and
δ18O for almost all Earth materials plot on a mass fraction line
with a slope of ∼0.5, passing through the isotopic composition of SMOW. Most natural processes fractionate isotopes
according to their masses, so that the resulting compositions
move up or down the line, but not off of it.
FIG. 13.10. Gypsum from the central Namib Desert and from
dry valleys in Antarctica is relatively enriched in 17O, so that
its composition plots above the terrestrial mass fractionation
line. This gypsum is deposited directly from the atmosphere,
where its sulfate is produced by reactions between biogenic
dimethyl sulfide and 17O-rich O3 and H2O2. (After Thiemens
et al. 2001.)
inclusions that do not plot on the mass fractionation
line. Because no mass fractionation process could
account for this anomaly, they concluded that some
unusual nucleosynthetic event—a supernova—must
have sprayed pure 16O into the early Solar System.
It was incorporated in the inclusions, enriching their
mixture of oxygen isotopes by as much as 40‰ in
16O without also adding 17O or 18O.
Although still attractive for other reasons we will
discuss at greater length in chapter 15, this interpretation lost some of its unique appeal following the early
1980s, when other examples of mass-independent
fractionation began to appear in the chemical literature. Mark Thiemens and coworkers at the University
of California in San Diego have found that atmospheric ozone, CO2, CO, H 2O2, aerosol sulfate, nitrous
oxide, and even O2 all possess mass-independent isotopic compositions. The fractionation mechanism is
still not understood, although it evidently involves
Using Stable Isotopes
molecular reactions triggered by ultraviolet radiation
in the upper atmosphere.
This discovery opens the possibility that the meteoritic anomalies may have had a chemical origin rather
than a nucleosynthetic one. As importantly, perhaps,
it has also given geochemists a new tool for tracing
atmospheric gases through terrestrial pathways. Bao
and colleagues (2000) and Thiemens and coworkers
(2001), for example, have reported relative δ17O enrichments between +0.5 and +3.4‰ in extensive gyp-
Fractionation in Geothermal
and Hydrothermal Systems
Where water and rocks combine chemically, shifts in
isotopic values can be much larger and more obvious.
Exchange reactions between groundwater and rocks
generally cause isotopic values to leave the meteoric
water line. This tendency is perhaps most easily seen in
analyses of geothermal brines. The amount of hydrogen
in most minerals is small, so most of the hydrogen in a
water-saturated rock is in the water. Exchange reactions,
therefore, produce only negligible shifts in the deuterium
content of geothermal brines. Particularly at elevated
temperatures, however, large changes in δ18O are possible. More than half of the oxygen in a typical waterrock system is in the rock. Reactions involving silicates
or carbonates commonly enrich geothermal waters in
18
O relative to rainwater. Four horizontal arrows on
figure 13.11 indicate the trends measured at Steamboat
Springs (Colorado), and Lassen Park, the Salton Sea,
and The Geysers (all in California) by Ellis and Mahon
(1977). Trends like the four shown here should converge
on the composition of juvenile (mantle-derived) water if
any appreciable mixing with deep fluids had taken place.
Because they do not, geochemists conclude that water in
geothermal systems is derived almost entirely from local
precipitation.
This conclusion appears reasonable for any waterdominated hydrothermal system, but what about those
in which water must circulate through plutonic rocks or
through country rock with a very low permeability? In
these, the water:rock ratio should be very low, and there
may be comparable amounts of exchangeable hydrogen
in the rock and in circulating fluid. In rocks with low
275
sum deposits in the central Namib Desert and in dry
valleys in Antarctica. In both cases, the sulfur source
is dimethyl sulfide (DMS), released as a gaseous product of microbial activity in adjacent oceans and then
oxidized to sulfate by interaction with atmospheric O3
and H2O2. The 17O-enriched isotopic signatures of
these two gases is retained as the sulfate is deposited
(see fig. 13.10), making it possible to distinguish the
Namibian and Antarctic deposits from more typical
sulfate deposits formed by chemical weathering.
permeabilities, therefore, we should observe that both
δ18O and δD shift in the fluid phase.
This is an exciting possibility, because it offers a way
to solve a problem that has interested economic geologists for a long time. Particularly where circulating
meteoric waters have mobilized and concentrated economically important metals, economic geologists want
to know how much water has been involved in exchange
reactions. Without making some assumptions about the
nature of the subsurface plumbing in an area, it is difficult to calculate the ratio of water to rock during a period
of hydrothermal alteration. Suppose, though, that we
observed shifts in δD and δ18O for the mineralizing fluid
in a given deposit. Maybe we could use those observations to estimate how much water had passed through.
FIG. 13.11. Isotopic compositions of geothermal waters. (Data
from Ellis and Mahon 1977.)
276
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
The following problem illustrates one attempt of this
type.
Worked Problem 13.5
The San Cristobal ore body in the Peruvian Andes is a
wolframite-quartz vein deposit associated with the intrusion of
Tertiary quartz monzonite into older metamorphic rocks.
Oxygen and hydrogen isotopic data were gathered in the area
by economic geologist Andrew Campbell and his colleagues in
1984. How can these data be used to estimate the water to rock
ratio during hydrothermal mineralization?
Campbell began by noting that isotopic mass balance for
either oxygen or hydrogen in a water-rock system can be expressed by:
WδH O(i) + Rδrock(i) = WδH O(f) + Rδrock(f),
2
2
(13.7)
in which W and R are the total atomic percents of exchangeable
oxygen or hydrogen in water and rock, respectively, and i and
f stand for initial and final states in the system. The δD or
δ18O values for fresh meteoric water (δH O(i)) and unaltered
2
rock (δrock(i)) are easy to measure in the lab. Determining the
final value of δH O(f), however, takes more work.
2
Assuming that altered rock and fluid were in equilibrium
at the temperature of mineralization, we can calculate 1000 ln
δrock-H O (= ∆ = δrock(f) − δH O(f)). If δrock(f) is known, we can
2
2
then use equation 13.3 to recast equation 13.7. With a little
rearranging, we find that:
δD = −70‰. They estimated values of δH O(i) for hydrogen and
2
oxygen by analyzing fluids trapped at low temperature in latestage barite. These values are δD = −140‰ and δ18O = −20‰.
They estimated the value of ∆ for hydrogen, calculated from
a fractionation equation for biotite and water (similar to the
method used in worked problem 13.2), to be −49.8‰. The
value of ∆ for oxygen, calculated from a fractionation equation
for plagioclase (An30) and water, was estimated to be +2.4‰.
Campbell and colleagues generated the plot in figure 13.12
by inserting these various parameter values into equation 13.9
and varying the relative magnitudes of w and r. Figure 13.12
shows that, in theory, hydrothermal fluids produced by exchange between meteoric water and rock at high w/r ratios
(>0.2) experience only a shift toward heavier δ18O without
any change in δD. As the water:rock ratio decreases below 0.2,
however, δ18O only changes slightly, whereas δD becomes increasingly positive.
To estimate the actual water to rock ratio at San Cristobal,
Campbell and coworkers measured δ18O and δD in trapped
fluids and plotted them on this diagram for comparison with
W/R = [∆ + δH O(f) − δrock(i)]/[δH O(i) − δH O(f)]. (13.8)
2
2
2
Now, a brief diversion: In this problem, we are interested in
finding out how final values of δDH O and δ18OH O in fluid are
2
2
affected by relative masses of water and rock. In writing equation 13.7, however, we defined W and R as the atomic proportions of oxygen or hydrogen in water and rock. To convert
from W and R to new variables w and r, which will refer to
mass quantities of water and rock, we introduce a new quantity
z (= wR/Wr) to adjust for the different amounts of hydrogen or
oxygen in water and rock, so that:
w/(w + zr) = W/(W + R).
So much for the diversion. To find the isotopic composition
of water in equilibrium with altered granite, then, we convert
from W and R to w and r and solve equation 13.8 for δH O(f):
2
δH O(f) = δH O(i)[w/(w + zr)]
2
2
+ (δrock(i)− ∆)[zr/(w + zr)].
(13.9)
Water is 89 wt % oxygen, and granite contains ∼45 wt % oxygen, so when we use equation 13.9 to calculate the δ18O for
water in equilibrium with granite, z will be 45/89, or ∼0.5. Bulk
rock analyses of granites typically contain ∼0.6 wt % water, so
when we calculate δD, z will be ∼0.006.
Campbell and his coworkers analyzed fresh granite at
San Cristobal and found that it has a δ18O value of +7‰ and
FIG. 13.12. Isotopic compositions of waters in the San Cristobal,
Peru, hydrothermal system. δD and δ18O values along the solid
curve are calculated for various ratios of water to rock (w/r), as
indicated at various points along the curve. By comparing these
calculated isotopic values to analyses of water in isotopic equilibrium with quartz and wolframite in the deposit (boxes), Campbell
and colleagues (1984) estimated that the water:rock ratio during
ore-formation was between 0.01 and 0.003. Solid circle indicates
the calculated value for magmatic water. Wolframite could also
have been deposited from a mixture of meteoric and magmatic
waters, as indicated by the dashed line. (Modified from Campbell
et al. 1984.)
Using Stable Isotopes
the theoretical values. These measurements, also indicated on
figure 13.12, tell us that w/r was probably between 0.1 and
0.003.
As this worked example indicates, the isotopic composition of hydrothermal fluids and coexisting rocks depends greatly on the relative amounts of water and rock
in a system. Because of this, it is not always clear that an
interpretation like the one by Campbell and colleagues
(1984) is appropriate. How can we distinguish between
fluids that are meteoric waters modified by exchange
with rock, on the one hand, and fluids that are simply
mixtures of meteoric and juvenile water, on the other?
In the San Cristobal study, Campbell and company used
equations for ∆pl-H2O and ∆ bio-H2O to calculate the composition of water in equilibrium with granite at 800°C.
This “magmatic” water, plotted as a filled circle on figure 13.12, is colinear with fresh meteoric water and the
calculated hydrothermal fluid in the box labeled “wolframite.” The fluid in equilibrium with quartz, however,
has a much lower δD value. Campbell concludes from this
that wolframite at San Cristobal may have been deposited
by a mixed meteoric-magmatic fluid, but that quartz
probably was not.
277
Another approach has been used to distinguish between meteoric and magmatic fluids in large-scale porphyry metal deposits. Mineralization in porphyry copper
and molybdenum deposits characteristically occurs at the
border between a porphyritic granite or granodiorite intrusion and country rock. Economically extractable sulfide minerals are disseminated in a broad, highly altered
zone and veins and fractures that cut across both the intrusion and its host. It has long been recognized that this
mineralization was promoted by the migration of heated
water. Before isotopic data began to accumulate on these
systems, however, the source of this water was widely
believed to be magmatic.
Several research teams during the 1970s (summarized
by Cal Tech geochemist Hugh Taylor [1997]) investigated
the oxygen and hydrogen systematics in sericite, pyrophyllite, and hypogene clay minerals from North American porphyry metal deposits. As shown in figure 13.13,
δ values for each of these minerals lie in a belt parallel
to the meteoric water line. As we compare the deposits
to each other, we see also that there is a clearly recognizable latitude trend among them. Both δD and δ18O
decrease consistently as we sample northward from Arizona and New Mexico (Santa Rita, Safford, Copper
Creek, and Mineral Park) through Utah, Colorado, and
FIG. 13.13. Isotopic compositions of OH-bearing minerals from a selection of North American
porphyry metal deposits. Values are more negative for northern deposits, indicating that the
hydrothermal fluid was probably circulating meteoric water. Compare with figure 13.7. (Data
from Sheppard et al. 1969, and Taylor 1997.)
278
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Nevada (Gilman, Bingham, Ely, and Climax) to Montana, Washington, and Idaho (Butte, San Poil, Ima, and
Wickes). This trend looks so much like what we see along
the meteoric water line itself (recall fig. 13.7) that it is
now commonly believed that porphyry metal deposits
are generated largely by recycling local meteoric water.
Fractionation in Sedimentary Basins
Basin brines present yet another class of trends on a
two-isotope plot for oxygen and hydrogen. It was once
widely believed that pore waters in deep sedimentary
basins were connate fluids (that is, samples of seawater
trapped at the time of sediment accumulation). Researchers working with Robert Clayton (1966) and others more recently have shown that waters within a given
basin generally have a distinct isotopic signature. Values
for both δ18O and δD show considerable scatter, but the
least saline brines in any basin tend to lie close to the
meteoric water line, whereas more saline fluids generally
contain heavier oxygen and hydrogen. Furthermore, as
with porphyry copper and molybdenum deposits, isotopic values display a latitudinal variation similar to
that for surface waters. As indicated on figure 13.14,
Gulf Coast basin brines are much heavier than those in
the Alberta Basin. The primary source for these brines,
therefore, seems to be local rainwater. The degree of
scatter and the disparate slopes of isotopic trends from
basin to basin, however, suggest that other sources (silicate or carbonate rocks, true connate water, clays, or
hydrocarbons) are also important.
Fractionation among Biogenic Compounds
Two stable isotopes of carbon, 12C and 13C, exist in
nature. On a global scale, δ13C values generally range
from about −30‰ to +25‰ relative to PDB, although
methane in some natural gas fields has been found with
values as low as −100‰. In general, carbon in the reduced compounds found in living and fossil organisms
is isotopically light, whereas carbonate minerals (particularly in marine environments) are isotopically heavy.
This twofold distinction itself is useful in characterizing
the sources of secondary carbonate in sediments. The
carbonate cap rocks on salt domes, for example, have
δ13C values that are very negative (about −36‰) compared with marine carbonates (near +0‰). This has been
taken as evidence that they are produced by oxidation of
FIG. 13.14. Isotopic compositions of basin brines are increasingly negative from south to north, indicating the influence of
meteoric waters. The scatter in values and the variable trends
within basins (indicated by dashed lines) are due to the influence
of other sources such as true connate waters or magmatic fluids,
and to isotopic exchange with host rocks. (Based on data from
Taylor 1997.)
methane in adjacent hydrocarbon-rich strata, rather than
from recrystallized marine limestones. In a similar way,
fresh-water carbonates and dripstone in caves are characteristically light, because they are formed from waters
that are charged with soil CO2. As we saw in chapter 7,
soil CO2 is largely the product of plant respiration and
decay, so it should be 12C-rich relative to atmospheric
CO2.
The complex biochemical pathways in living organisms offer many opportunities to fractionate carbon isotopes. Some of these have been investigated, with limited
success, for use as paleoenvironmental indicators or as
guides in provenance studies. Green plants fractionate
carbon isotopes in at least three metabolic steps during
photosynthesis, each of which favors the retention of 12C
rather than 13C. In broad terms, these involve (1) the
physical assimilation of CO2 across cell walls, (2) the
conversion of that CO2 into intermediate compounds
by enzymes, and (3) the synthesis of large organic molecules. The degree of fractionation in each step is kinetically controlled by such factors as the atmospheric or
intracellular partial pressure of CO2.
In addition, terrestrial plants can be divided into two
large groups on the basis of the enzyme that dominates
Using Stable Isotopes
in step 2 and the types of large organic molecules that
form in step 3. Those in one group, which includes trees,
bushes, and grains like wheat and rice, are known as C3
plants because the initial product of CO2 fixation is a
molecule with three carbon atoms. In these, lignin, cellulose, and waxy lipids constitute a large portion of the
total organic matter. By contrast, the C4 plants, which
include corn, sugar cane, many aquatic plants, and some
grasses, contain relatively high proportions of carbohydrate and protein. As summarized in table 13.3, C3
plants have δ13C values between about −22‰ and −33‰,
whereas C4 plants range from −9‰ to −16‰. Only a few
plants such as algae, lichens, and cacti occupy an isotopic middle ground between these two major plant
groups. Marine plants have 13C/12C ratios that are ∼7.5‰
less negative than C3 terrestrial plants, possibly because
they assimilate carbon from marine HCO3− (0‰) rather
than atmospheric CO2 (−7‰). Plankton vary from 17‰
to about −27‰.
Sackett and Thompson (1963) measured δ13C values
in Gulf Coast sediment and found that they decrease
systematically shoreward from about −21‰ to −26‰.
Using the patterns we introduced in the previous paragraph as a guide, they inferred that the isotopic trend reflects a near-shore mixing of terrestrial and marine plant
debris. In more recent studies, isotopic signatures of
individual species of phytoplankton have been used to
account for the spatial variability of 13C in estuarine and
shallow marine sediments. Without painstakingly detailed work, studies of this sort can yield highly ambiguous results because of the large number of hidden source
TABLE 13.3. Ranges of δ13C and δ15N Values of Inorganic
Nutrients, Higher Plants, Photoautotrophs, and Particulate
Organic Matter from Terrestrial and Marine Environments
Source
δ13C (‰)
Atmospheric CO2
Dissolved CO2
Bicarbonate (HCO3– )
Atmospheric N2
Terrestrial C3 plants
Marsh C3 plants
Marsh C4 plants
CAM plants
Seagrasses
Mangroves
Macroalgae
Temperate marine POM
Temperate estuarine POM
–7 to –8
–7 to –8
0
–30 to –23
–29 to –23
–15 to –6
–33 to –11
–16 to –4
–19 to –18
–27 to –10
–24 to –18
–30 to –15
δ15N (‰)
0
–7 to 6
3 to 5
1 to 8
–15 to –1
0 to 6
6 to 7
–1 to –10
–2 to 10
2 to 19
Data from Ostrom and Fry (1993) and Hoefs (1997).
279
materials and processes that may have contributed to
present isotopic compositions.
The transition from primary organic debris to kerogen causes an enrichment in 12C, which becomes more
pronounced as kerogen is converted to petroleum or coal.
Kerogen ranges from −17‰ to −34‰. Despite the degrading effects of diagenesis, source materials can still be
recognized isotopically in some kerogen and some Cenozoic coals and crude oils. Marine biogenic matter and C3
terrestrial plant remains retain their relatively light δ13C
values most effectively, presumably because the double
bonds associated with polycyclic hydrocarbons resist
thermal degradation well. The single bonds in hydrocarbon chains of C4 plant debris are more vulnerable; yet
they, too, resist mild diagenesis and retain their heavier
δ13C signatures well enough that geochemists can infer
some source characteristics. Ultimately, of course, diagenesis renders even the most obvious trends useless.
Pre-Tertiary coals all have δ13C values very near −25‰
and values in petroleum cluster very close to −28‰.
Isotopic Fractionation around
Marine Oil and Gas Seeps
In chapter 5, we discussed the bacterially mitigated
digenesis of organic matter in the presence of dissolved
sulfate, using the schematic reaction:
→
7CH 2O + 4SO42− ←
2S 2− + 7CO2 + 4OH − + 5H 2O.
Depending on whether we want to emphasize the end
production of sulfide or soluble bicarbonate, an equally
plausible overall reaction might be:
→ H S + 2HCO −.
2CH 2O + SO42− ←
2
3
“CH 2O,” in either context, is a generic placeholder for
the complex mix of hydrocarbons derived from C3 and
C4 plants. At places where crude oil and gas seep from
the cold ocean floor, however, the organic food for bacteria can instead be methane, the by-product of petroleum maturation deep below the seafloor. The process
of bacterial sulfate reduction using CH4 as the energy
source can be described by the overall reaction:
→ H S + CO 2− + H O.
CH4 + SO42− ←
2
3
2
Does it make a difference whether CH 2O or CH4 is
the dominant carbon source at a particular site?
280
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
AN ORGANIC ODYSSEY: FROM KANSAS WHEAT FIELDS TO THE MISSISSIPPI DELTA
Isotopic analysis is particularly useful for distinguishing among possible sources of organic matter in
aquatic sediments, each of which has a distinct isotopic composition. For example, vascular plants such
as trees and shrubs (δ13C = −25‰ to −28‰) that
utilize the C3 metabolic pathway can be distinguished
from corn, bamboo and many grasses (δ13C = −10‰
to −14‰) that employ the C4 pathway. In addition,
freshwater phytoplankton are typically depleted in
13C (δ13C = −30‰ to −40‰) relative to C land plants,
3
and marine phytoplankton are usually more enriched
(−15‰ to −20‰). As we discovered in chapter 6,
however, organic matter in aquatic and sedimentary
environments is a highly altered, complex mixture
of organic components from many sources. Isotopic
values of preserved organic matter also record the
isotopic history of all components from production
to bacterial alteration and diagenesis, further complicating things.
Fortunately, as we also found in chapter 6, some
biomolecules survive alteration during deposition and
burial. The δ13C values of biomarkers such as lipids
(alkanes and sterols) and lignin phenols provide a
more detailed picture of individual sources than the
isotopic compositions of bulk organic matter do. For
example, the δ13C values of lignin phenol derived from
C3 plants are typically between −28‰ and −34‰,
whereas carbon isotopic values of C4 lignin phenols
are roughly between −14‰ and −20‰. Compare each
of these to the relatively lower δ13C values for bulk C3
and C4 plants that we gave above. In contrast, other
biomarkers such as amino acids are enriched in 13C
relative to the whole organism. Using the δ13C values
of individual compounds, organic geochemists can determine the sources of multiple components in organic
matter in complex systems and trace the fate of these
compounds through complex reactions and processes.
Organic geochemists Miguel Goñi, Kathy Ruttenberg, and Tim Eglinton (1998) used the δ13C values of
bulk organic matter and lignin phenols from marine
sediments to investigate the input of terrestrial or-
ganic matter to the Gulf of Mexico. The δ13C values
of bulk organic matter sampled from depths between
100 and 2250 m ranged from −19.7‰ to −21.7‰. At
first glance, these bulk values seemed to reflect significant contribution from marine phytoplankton and
a minor input from C3 vascular plants. Goñi and his
colleagues were skeptical, however, because they knew
that a significant portion of Mississippi River sediment is derived from the central United States, which
contains extensive grasslands (C4 plants).
A more detailed picture of the potential sources
was needed, so Goñi’s research team investigated the
distribution of lignin phenols in sediments. In general, the S/V (1.6) and C/V (0.5) ratios (which we
introduced in chapter 6) suggested input from nonwoody angiosperms (see table 6.4). Fine-grained, deep
sediments (>100 meters), however, had much higher
S/V ratios than the coarser grained sediments in shallower samples. This suggested that organic matter in
the fine-grained, off-shore sediments was most likely
derived from nonwoody angiosperms, such as grasses,
which dominate the central United States. The organic matter in near-shore sediments presumably originated from nearby estuarine environments, where
nonwoody gymnosperms found in swamps, bays and
estuaries leave a lower S/V signature.
To confirm the S/V evidence that midcontinent
C4 plants are the dominant source of organic matter
in deep sediments, Goñi and his colleagues turned to
stable isotopes. The isotopic composition of the lignin
phenols provided unequivocal evidence. For the shallow sites, the δ13C values were between −21.1‰ and
−29.0‰, but in the deep sediments, signature was
much lighter, between −15.1‰ and −21.8‰. These
values are consistent with the δ13C signatures of lignin
phenol from C3 plants and C4 plants, respectively.
These data showed that terrestrial organic matter is
an important component of these deltaic sediments,
and also support some of the ideas about organic
matter preservation in marine sediments that we discussed in chapter 6.
Using Stable Isotopes
281
FIG. 13.15. Analyzed SO42− and H2S concentrations in pore waters in sediment associated with oil
and gas seeps in the Gulf of Mexico and with sediments at a reference site distant from any seeps.
(Modified from Aharon and Fu 2000.)
Geochemists Paul Aharon and Baoshun Fu (2000)
have analyzed pore fluids in sediment cores at oil and gas
seeps in the Gulf of Mexico. As illustrated in figure 13.15,
the decreasing abundance of SO42− with depth is accompanied by an increase in H 2S at all sites. Clearly, though,
the consumption of SO42− is much more complete near
seeps, where microbes live on CH4, than at a distance,
where CH 2O is their dominant carbon source. Several
previous studies have indicated that the type of source
organic matter, rather than the amount of it, controls the
rate of sulfate reduction. These data, then, confirm that
the microbial diet makes a difference.
Isotopic analyses add a dimension to this conclusion.
Generally, both δ18O and δ34S increase with depth in each
of their sample cores. This is to be expected, because the
rate of microbial SO42− reduction is faster for 32S16O42−
than for 34S18O42−, consistent with the principles we outlined in a box earlier in this chapter. Therefore, the more
SO42− is consumed in the closed system environment of
the sediment column, the heavier the isotopic composition of any remaining SO42− becomes. The surprise for
Aharon and Fu was that the degree of isotopic fractionation for both oxygen and sulfur varied dramatically
from one sampling site to another, apparently as a function of the mix of CH2O and CH4. At a distance from oil
or gas seeps (that is, in “normal” marine sediment), both
18O and 34S in residual SO 2− increase rapidly with depth,
4
even though only a small fraction of the SO42− in pore
fluid is consumed. Although a greater proportion of SO42−
is consumed, the relative isotopic enrichments are much
less pronounced at oil seep sites, and significantly less
at gas seep sites, as is evident in figure 13.16. That is,
microbes degrading CH2O discriminate between heavy
and light isotopes of sulfur and oxygen much more efficiently than do microbes consuming a CH4-rich diet.
Aharon and Fu infer that because temperature and other
compositional parameters are the same across all sites,
these isotopic differences reflect differences in the metabolic pathways—perhaps involving different enzymatic
intermediates—by which sulfate-reducing microbes
separate oxygen from sulfur, depending on what they
are eating. This may suggest a way to interpret the type
of bacterial activity that took place in ancient environments for which we have only an isotopic record in preserved sediments.
SUMMARY
In this chapter, we have discussed some of the basic
processes that lead to the fractionation of light stable
isotopes from one another in geologic environments.
Hydrogen, carbon, nitrogen, oxygen, and sulfur are all
abundant in the crust and atmosphere/hydrosphere, and
may be found in a wide variety of chemical compounds.
Variations in bond strength and site energy among these
compounds lead them to favor different proportions of
the stable isotopes.
The free energies of isotopic exchange reactions vary
with temperature, but not with pressure. This makes
them very useful as geothermometers. The most popular
282
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 13.16. Enrichments of δ34S (relative to a marine value of 20.3‰) and δ18O (relative to a marine value of 9.7‰) as a function of the residual fraction ( f ) of SO42− in pore waters after bacterial
sulfate reduction. (a, b) At a distance from seeps, (c, d) at oil seeps, and (e, f) at gas seeps. In
each case, the enrichment factor (E) is equal to 103(α − 1), or (δ − δo )/ln f, where δ is the measured δ34S or δ18O value, δo is the corresponding value in seawater, and f is the residual fraction of
SO42−. (Modified from Aharon and Fu 2000.)
geothermometers involve equilibrium fractionation of
oxygen or sulfur isotopes.
Kinetic fractionation effects are also commonly exploited by geochemists. The small difference in vapor
pressure between “light” water and “heavy” water leads
to a significant fractionation of oxygen and hydrogen
isotopes with increasing latitude, altitude, and distance
from the ocean. These systematic trends in meteoric
water have been useful in recognizing the source of
water in aquifers, sedimentary basins, and hydrothermal
systems.
The role of living organisms in fractionating light,
stable isotopes has made it possible to determine the
provenance and, to some degree, the depositional and
diagenetic history of organic matter. Isotopic biomarkers in kerogen and young fossil hydrocarbons
can be helpful stratigraphic guides, and biomarkers in
sedimentary environments can identify primary nutrient sources and biochemical pathways of deposited
organic matter.
suggested readings
The following books discuss principles of stable isotope geochemistry at a level appropriate for the beginning student. In
most cases, they emphasize the application of methods to
specific field problems.
Barnes, H. L., ed. 1997. Geochemistry of Hydrothermal Ore
Deposits, 3rd ed. New York: Wiley Interscience. (This collection of basic articles by geochemists at the leading edge
of research in ore-forming systems has two long chapters
devoted to light stable isotopes.)
Clayton, R. N. 1981. Isotopic thermometry. In R. C. Newton, A. Navrotsky, and B. J. Wood, eds. Thermodynamics
of Minerals and Melts. New York: Springer-Verlag,
pp. 85–109. (This concise review paper is a very good way
to learn the first principles of stable isotope geothermometry.)
Faure, G. 1986. Principles of Isotope Geology, 2nd ed. New
York: Wiley. (This is an easy-to-read text dealing largely
with radioactive isotopes. Chapters 18–21, however, give a
good introduction to light stable isotopes.)
Using Stable Isotopes
Hoefs, J. 1997. Stable Isotope Geochemistry, 4th ed. New
York: Springer-Verlag. (This is an excellent text, covering
all the basics, with a focus on field examples.)
Jager, E., and J. C. Hunziker. 1979. Lectures in Isotope Geology. Berlin: Springer-Verlag. (The lectures in this book were
given in 1977 as an introductory short course for geologists.
Most deal with radioactive isotopes, but the four final lectures are a useful overview of hydrogen, oxygen, carbon,
and sulfur.)
Valley, J. W., H. P. Taylor, and J. R. O’Neil, eds. 1986. Stable
Isotopes in High Temperature Geological Processes. Mineralogical Society of America Reviews in Mineralogy 16. Washington, D.C.: Mineralogical Society of America. (Papers in
this volume are of particular interest to igneous and metamorphic petrologists who use stable isotopes to study processes in the deep crust.)
We refer to the following papers in chapter 13. The interested
student may wish to consult them further.
Aharon, P., and B. Fu. 2000. Microbial sulfate reduction rates
and sulfur and oxygen isotope fractionations at oil and gas
seeps in deepwater Gulf of Mexico. Geochimica et Cosmochimica Acta 64:233–246.
Bao, H., M. H. Thiemens, J. Farquhar, D. A. Campbell, C. C.-W.
Lee, K. Heine, and D. B. Loope. 2000. Anomalous 17O compositions in massive sulphate deposits on the Earth. Nature
406:176–178.
Berner, R. A. 1987. Models for carbon and sulfur cycles and
atmospheric oxygen: Application to Paleozoic geologic history. American Journal of Science 287:177–196.
Campbell, A., D. Rye, and U. Petersen. 1984. A hydrogen
and oxygen isotope study of the San Cristobal mine, Peru:
Implications of the role of water to rock ratio for the genesis of wolframite deposits. Economic Geology 79:1818–
1832.
Clayton, R. N., I. Friedman, D. L. Graf, T. K. Mayeda, W. F.
Meents, and N. F. Schimp. 1966. The origin of saline formation waters, I: Isotopic composition. Journal of Geophysical Research 71:3869–3882.
Clayton, R. N., L. Grossman, and T. K. Mayeda. 1973. A component of primitive nuclear composition in carbonaceous
chondrites. Science 182:485–488.
Coplen, T. B., C Kendall, and J. Hopple. 1983. Comparison of
stable isotope reference samples. Nature 302:236–238.
Craig, H. 1961. Isotopic variations in meteoric waters. Science
133:1702–1703.
Deines, P., D. Langmuir, and R. S. Harmon. 1974. Stable carbon isotope ratios and the existence of a gas phase in the
evolution of carbonate ground water. Geochimica et Cosmochimica Acta 38:1147–1164.
Ellis, A. J., and W.A.J. Mahon. 1977. Geochemistry and Geothermal Systems. New York: Academic.
283
Epstein, S., and T. Mayeda. 1953. Variations of 18O content of
waters from natural sources. Geochimica et Cosmochimica
Acta 4:213–224.
Goñi, M. A., K. C. Ruttenberg, and T. I. Eglinton. 1998. A reassessment of the sources and importance of land-derived
organic matter in surface sediments from the Gulf of Mexico. Geochimica et Cosmochimica Acta 62:3055–3075.
Hattori, K., and S. Halas. 1982. Calculation of oxygen isotope
fractionation between uranium dioxide, uranium trioxide,
and water. Geochimica et Cosmochimica Acta 46:1863–
1868.
Holser, W. T., and I. R. Kaplan. 1966. Isotope geochemistry of
sedimentary sulfates. Chemical Geology 1:93–135.
Matsuhisa, Y., J. R. Goldsmith, and R. N. Clayton. 1979.
Oxygen isotopic fractionation in the system quartz-albiteanorthite-water. Geochimica et Cosmochimica Acta 43:
1131–1140.
Miller, K. G., R. G. Fairbanks, and G. S. Mountain. 1987. Tertiary oxygen isotope synthesis, sea level history, and continental margin erosion. Paleoceanography 2:1–19.
Munha, J., F.J.A.S. Barriga, and R. Kerrich. 1986. High δ18O
ore-forming fluids in volcanic-hosted base metal sulfide
deposits: Geologic, 18O/ 16O, and D/H evidence from the
Iberian Pyrite belt; Crandon, Wisconsin; and Blue Hill,
Maine. Economic Geology 81:530–552.
Ohmoto, H., and M. B. Goldhaber. 1997. Isotopes of sulfur
and carbon. In H. L. Barnes, ed. Geochemistry of Hydrothermal Ore Deposits, 3rd ed. New York: Wiley Interscience,
pp. 517–612.
O’Neil, J. R. 1986. Theoretical and experimental aspects of isotopic fractionation. In J. W. Valley, H. P. Taylor, and J. R.
O’Neil, eds. Stable Isotopes in High Temperature Geological Processes. Mineralogical Society of America Reviews in
Mineralogy 16. Washington, D.C.: Mineralogical Society of
America, pp. 1–40.
Ostrom, P. H., and B. Fry. 1993. Sources and cycling of organic
matter within modern and prehistoric food webs. In M. H.
Engel and S. A. Macko, eds. Organic Geochemistry: Principles and Applications. New York: Plenum, pp. 785–798.
Rye, R. O. 1974. A comparison of sphalerite-galena sulfur
isotope temperatures with filling temperatures of fluid inclusions. Economic Geology 69:26–32.
Sackett, W. M., and R. R. Thompson. 1963. Isotopic organic
carbon composition of recent continental derived clastic
sediments of the eastern Gulf Coast, Gulf of Mexico. Bulletin of the American Association of Petroleum Geologists
47:525–531.
Sheppard, S.F.M., R. L. Nielsen, and H. P. Taylor. 1969.
Oxygen and hydrogen isotope ratios of clay minerals from
porphyry copper deposits. Economic Geology 64:755–777.
Siegel, D. I., and R. J. Mandle. 1984. Isotopic evidence for
glacial melt-water recharge to the Cambrian-Ordovician
aquifer, North-Central United States. Quaternary Research
22:328–335.
284
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Taylor, H. P. 1997. Oxygen and hydrogen isotope relationships
in hydrothermal mineral deposits. In H. L. Barnes, ed. Geochemistry of Hydrothermal Ore Deposits, 3rd ed. New
York: Wiley Interscience, pp. 229–302.
Thiemens, M. H., J. Savarino, J. Farquhar, and H. Bao. 2001.
Mass-independent isotopic compositions in terrestrial and
extraterrestrial solids and their applications. Accounts of
Chemical Research 34(8):645–652.
Urey, H. C. 1947. The thermodynamics of isotopic substances.
Journal of the Chemical Society (London) 562–581.
PROBLEMS
(13.1) The relative atomic abundances of oxygen and carbon isotopes are:
16O:17O:18O
12C:13C
= 99.759:0.0374:0.2039,
= 98.9:1.1.
Calculate the relative abundances of (a) CO2 molecules with masses 44, 45, and 46; (b) CO molecules with masses 28, 29, and 30.
(13.2) The fractionation factor for hydrogen isotopes between liquid water and vapor at 20°C is 1.0800.
Make a graph to indicate how the value of δD changes in rainwater during continued precipitation
from an air mass that initially has a value of −80‰. Show how the δD value of the remaining water
vapor changes during this process.
(13.3) Analyses of foraminifera that formed at known temperatures have yielded the following isotopic
data:
T (°C)
δ18Oshell
30
25
20
15
10
5
0
−3.14
−1.97
−0.82
+0.35
+1.51
+2.68
+3.84
(a) Graph these data and determine the appropriate constants a and b for a linear equation of the
form T = a + bδ18Oshell.
(b) Seawater did not have its modern δ18O value of +0.0‰ until ∼11,000 years ago. If we know
what the isotopic composition of seawater (δ18Osw) was at any earlier time, we can adjust for it
by writing T = a + b (δ18Oshell − δ18Osw). Between 22,000 and 11,000 years ago, δ18Osw was
+2.0‰. If you find that 13,000-year-old foraminifera in a deep-ocean sample core have the same
δ18Oshell values as modern foraminifera in the same core, by how much would you infer that the
temperature of ocean water has changed?
(13.4) The following ∆34 S data were collected on samples from Providencia, Mexico, by Rye (1974). On the
basis of these data, what is your estimate for the temperature at which these ores formed?
Sample #
60-H-67
60-H-36-57
62-S-250
63-R-22
Galena
Sphalerite
−1.15‰
−1.11
−1.42
−1.71
+0.95‰
+0.87
+1.03
+0.01
Using Stable Isotopes
(13.5) Hattori and Halas (1982) have determined that the fractionation factor for oxygen exchange between UO2 and water varies with temperature according to the relationship:
1000 ln α UO2-H2O = 3.63 × 106T −2 − 13.29 × 103 T + 4.42.
Using the similar expression for quartz-water fractionation from worked problem 13.2, devise an
oxygen isotope geothermometer to estimate temperatures from coexisting quartz and uraninite.
(13.6) The fractionation of carbon isotopes between calcite and CO2 gas can be calculated from
1000 ln δcal-CO2 = 1.194 × 106T −2 − 3.63 (Deines et al. 1974). If the δ13 C value of atmospheric CO2
is −7.0‰, what should be the δ13C value of calcite precipitated from water in equilibrium with the
atmosphere at 20°C?
(13.7) Rain and snow from a large number of sampling sites have been analyzed and found to have the
following δ18O values:
δ18O (‰)
Mean Annual Air Temperature (°C)
−3.18
−8.04
−11.52
−20.55
−30.98
15
8
3
−10
−25
Using these data, write an equation to predict the oxygen isotopic composition of precipitation at
any locality, given its mean annual temperature. Write a second equation based on these data and the
meteoric water line, relating δD to mean annual temperature.
(13.8) Use the procedure by Campbell and colleagues (1984), described in worked problem 13.5, to determine what the final values of δ18O and δD would be in water that initially had values of −30‰ and
−230‰, but that had equilibrated in a closed system with granite at 400°C. Assume the same isotopic values for unaltered granite that were used for the study at San Cristobal, Peru, and a water to
rock ratio (w/r) of 0.05.
285
CHAPTER 14
USING RADIOACTIVE ISOTOPES
OVERVIEW
In the geochemist’s arsenal of techniques for unraveling
geological problems, the study of radioactive isotopes
and their decay products has become very prominent.
This chapter begins with a discussion of nuclide stability
and decay mechanisms. After equations that describe
radioactive decay are derived, we examine the utility of
certain naturally occurring radionuclide systems (K-Ar,
Rb-Sr, Sm-Nd, and U-Th-Pb) in geochronology. The
concept of extinct radionuclides that were present in the
early solar system is also discussed. The principles behind both mass spectrometry and fission track techniques
are introduced. We discuss how induced radioactivity
can be used to solve geochemical problems. Examples
include neutron activation analysis, 40Ar-39Ar dating,
and the use of cosmogenic nuclides (14C and 10Be) for
dating geological and archeological materials and for
detecting the subduction of sediments. We explore how
radionuclides can be used as geochemical tracers in such
global problems as determining when the Earth accreted
and differentiated, quantifying mantle heterogeneity,
assessing the cycling of material between crust and mantle, revealing global oceanic mixing patterns, and understanding degassing of the Earth’s interior to produce the
atmosphere.
286
PRINCIPLES OF RADIOACTIVITY
Nuclide Stability
In chapter 13, we learned how stable isotopes can be
used to solve geochemical problems. The study of unstable (that is, radioactive) nuclides, considered in this
chapter, is also an extremely powerful technique. In chapter 2, we saw that only ∼260 of the 1700 known nuclides
are stable, so we can infer that nuclide stability is the
exception rather than the rule. Most of the known radioactive isotopes do not occur in nature. Although some of
these may have occurred naturally in the distant past,
their decay rates were so rapid that they have long since
been transformed into other nuclides. In most cases, radioactive isotopes that are of interest to geochemists
require very long times for decay or are produced continually by naturally occurring nuclear reactions.
In chapter 13, we explored how variations in stable
isotopes are caused by mass fractionation during the
course of chemical reactions or physical processes. With
the exception of radioactive 14C and a few other nuclides,
the atomic masses of most of the unstable isotopes of geochemical interest are very large, so that mass differences
with other nuclides of the same element are minuscule.
Consequently, these isotopic systems can be considered
Using Radioactive Isotopes
to be immune to mass fractionation processes. Thus, all
of the measured variations in these nuclides are normally
ascribed to radioactive decay.
Decay Mechanisms
Radioactivity is the spontaneous transformation of
an unstable nuclide (the parent) into another nuclide (the
daughter). The transformation process, called radioactive
decay, results in changes in N (the number of neutrons)
and Z (the number of protons) of the parent atom, so
that another element is produced. Such processes occur
by emission or capture of a variety of nuclear particles.
Isotopes produced by the decay of other isotopes are
said to be radiogenic. The radiogenic daughter may be
stable or unstable; if it is unstable, the decay process
continues until a stable nuclide is produced.
Beta decay involves the emission of negatively charged
beta particles (electrons emitted by the nucleus), commonly accompanied by radiation in the form of gamma
rays. This is equivalent to the transformation of a neutron
into a proton and an electron. As illustrated in figure 14.1,
Z increases by one and N decreases by one for each beta
particle emitted.
FIG. 14.1. A portion of the nuclide chart illustrating the N-Z relationships for daughter nuclides formed from a hypothetical parent
by emission of beta particles, positrons, alpha particles, or by
electron capture.
287
Another type of radioactivity occurs as a result of
positron decay. A positron is a positively charged electron
expelled from the nucleus. This process can be regarded
as the conversion of a proton into a neutron, a positron,
and a neutrino (a particle with appreciable kinetic energy, but without mass). Positron decay produces a
nuclide with N increased by one and Z decreased by
one, as illustrated in figure 14.1.
Inspection of figure 14.1 allows us to predict which
nuclides tend to transform by beta decay or positron decay. Unstable atoms that lie below the band of stability
(as seen in fig. 2.3), and therefore have excess neutrons,
are likely to decay by emission of beta particles, so that
their N values are reduced. Similarly, nuclides that lie
above the band of stability, having excess protons, may
experience positron decay. In each case, these processes
occur in such a way that the resulting daughter nuclides
fall within the band of nuclear stability.
An alternate type of decay mechanism is electron
capture. In this process, the nuclide increases its N and
decreases its Z by addition of an electron from outside
the nucleus. A neutrino is also liberated during this process. The daughter nuclide produced during electron capture occupies exactly the same position relative to its
parent in figure 14.1 as that produced during positron
decay.
Alpha decay proceeds by the emission of heavy alpha
particles from the nucleus. An alpha particle is composed
of two neutrons and two protons. Therefore, the N and
Z values for the daughter nuclide both decrease by two,
as shown in figure 14.1.
To be complete, we should also add to this list of radioactive decay processes spontaneous fission, which is an
alternate mode of decay for some heavy atomic nuclei.
These atoms may break apart, because electrostatic repulsion between the Z positively charged protons overcomes the strong nuclear binding force. Fission products
generally have excess neutrons and tend to decay further
by beta emission. For most geochemical applications,
fission can be considered a relatively minor side effect.
The decay of natural radionuclides may be much more
complex than suggested by the simple decay mechanisms
just introduced. A particular transformation may employ several of these decay mechanisms simultaneously,
so that the parent atoms form more than one kind of
daughter. Such a process is called branched decay. As an
example, consider the decay of 40 K. Most atoms of the
288
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
parent nuclide decay by positron emission and electron
capture into 40Ar, but 40Ca is produced from the remainder by beta decay. Branched decay decreases the
yield of daughter atoms of a particular nuclide, requiring more sensitivity in measurement.
characteristic for a particular radionuclide and is expressed in units of reciprocal time. Rearranging the terms
of equation 14.1 and integrating gives:
∫
∫
− dN/N = λ dt,
or
Worked Problem 14.1
−ln N = λt + C,
How many atoms of 40Ar would be produced by the complete
decay of 40 K in a 1 cm3 crystal of orthoclase? To answer this
question, we must first calculate the number of atoms of parent
40 K in the sample. We can convert the volume of orthoclase to
weight by multiplying by its density, and weight to moles by
dividing by the gram formula weight of orthoclase:
where C is the constant of integration. If N = N0 at time
t = 0, then:
C = −ln N0.
By substituting this value into equation 14.2, we obtain:
1 cm3 (2.59 g cm−3)/ 278.34 g mol −1
−ln N = λt − ln N0,
= 9.30 × 10−3 mol.
Each mole of orthoclase contains one mole of potassium, so
there are 9.30 × 10−3 moles of K in this sample. Multiplication
of this result by Avogadro’s number gives the number of atoms
of potassium:
(14.2)
or
ln N/N0 = −λt.
This is more commonly expressed as:
9.30 × 10−3 mol (6.023 × 1023 atoms mol−1)
= 5.60 × 1019 atoms.
N/N0 = e−λt.
(14.3)
40 K
The natural abundance of
in potassium (before decay) is
0.01167%. Therefore, this orthoclase sample contains:
5.60 × 1019 atoms (1.167 × 10−4)
= 6.535 × 1015 atoms 40 K.
We have already mentioned that 40 K undergoes branched
decay, and we are concerned here with only one of the daughter nuclides. A total of 11.16% of the 40 K atoms decay into 40Ar
by electron capture and positron decay, the remainder transforming to 40Ca. Consequently, the number of atoms of 40Ar
produced by complete decay is:
6.535 × 1015 (0.1116) = 7.293 × 1014 atoms 40Ar.
Rate of Radioactive Decay
In each of the decay mechanisms just discussed, the
rate of disintegration of a parent nuclide is proportional
to the number of atoms present. In more quantitative
terms, the number of atoms (because of convention, we
use the symbol N, not to be confused with the symbol
for the number of neutrons used in earlier paragraphs)
remaining at any time t is:
−dN/dt = λN,
(14.1)
where λ is the constant of proportionality, usually called
the decay constant. The value of the decay constant is
Equation 14.3 is the basic relationship that describes
all radioactive decay processes. With it, we can calculate
the number of parent atoms (N) that remain at any time
t from an original number of atoms (N0) present at
time t = 0.
We can also express this relationship in terms of
atoms of the daughter nuclide rather than the parent. If
no daughter atoms are present at time t = 0 and none
are added to or lost from the system during decay, then
the number of radiogenic daughter atoms produced by
decay D* (not to be confused with the same symbol used
for tracer diffusion coefficients in chapter 11) at any
time t is:
D* = N0 − N.
(14.4)
By rearranging equation 14.3 and substituting it into
14.4, we see that:
D* = Ne λt − N,
or
D* = N(e λt − 1).
(14.5)
Equation 14.5 gives the number of daughter atoms
produced by decay at any time t as a function of parent
atoms remaining. If some atoms of the daughter nuclide
Using Radioactive Isotopes
were present initially (D0), then the total number of
daughter atoms (D) is:
D = D0 +
−11)(48.8
D = D0 + N(e λt − 1).
× 109)
,
or
D*.
N = 6.00 × 1019 atoms after one half-life.
Combining this equation with 14.5 produces the useful
result:
To calculate the number of atoms of the parent nuclide remaining after two half-lives, we substitute for N0 in the same
equation the value of N just calculated:
(14.6)
This important equation is the basis for geochronology.
Both D and N are measurable quantities, and D0 is a
constant whose value can be determined.
It is common practice to express radioactive decay
rates in terms of half-lives. The half-life (t1/2) is defined
as the time required for one-half of a given number of
atoms of a nuclide to decay. Therefore, when t = t1/2, it
follows that N = 1–2 N0. Substitution of these values into
equation 14.3 gives:
1
– N /N
2 0
0
N/(1.2 × 1020) = e −(1.42 × 10
289
N/6.00 × 1019 = e −(1.42 × 10
−11)(48.8
× 109)
,
or
N = 3.00 × 1019 atoms after two half-lives,
and so forth.
To calculate the number of atoms (D) of daughter 87Sr after
one half-life, we use equation 14.6:
D = 0.3 × 1020 + 6.00 × 1019 [e (1.42 × 10
−11)(48.8
× 109)
− 1]
or
D = 9.00 × 1019 atoms after one half-life.
We follow a similar procedure for calculating daughter atoms
after each successive half-life.
The easiest way to summarize these calculations is by means
of a graph, plotting the number of atoms of parent (N) or
daughter (D) versus time in units of half-life. This diagram is
shown in figure 14.2. The exponential character of the radio-
= e −λt1/2,
or
ln 1–2 = −λt1/2.
This is equivalent to:
ln 2 = λt1/2.
Solving for t1/2 gives:
t1/2 = ln 2/λ = 0.693/λ.
(14.7)
Equation 14.7 expresses the relationship between the
half-life of a nuclide and its decay constant.
The equations above show that the disintegration of
a parent radioactive isotope and the generation of daughter nuclides are both exponential functions of time. This
is illustrated in the following worked problem.
Worked Problem 14.2
Follow the decay of radioactive parent 87Rb and the growth of
daughter 87Sr in a granite sample over the course of six half-lives.
Assume that the granite sample initially contains 1.2 × 1020
atoms of 87Rb and 0.3 × 1020 atoms of 87Sr.
The half-life of 87Rb is 48.8 × 109 years. The decay constant
for this radionuclide can be calculated by using equation 14.7:
λ = 0.693/(48.8 × 109 yr) = 1.42 × 10−11 yr−1.
Using equation 14.3, we can determine the number of atoms
(N) of parent 87Rb remaining after one half-life:
FIG. 14.2. The decay of radioactive 87Sr (N) into its stable radiogenic daughter, 87Rb (D), as a function of time measured in halflives. N0 and D0 represent the initial number of atoms of parent
and daughter, respectively.
290
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
active decay is obvious. After six half-lives, N approaches zero
asymptotically and, for most practical purposes, the production
of daughter atoms ceases.
Decay Series and Secular Equilibrium
The calculations above presume that a radioactive
parent decays directly into a stable daughter. However,
many radionuclides decay to a stable nuclide by way of
transitory, unstable daughters; that is, a parent N1 may
decay to a radioactive daughter N2, which in turn decays
to a stable daughter N3. Naturally occurring decay series
arising from radioactive isotopes of uranium and thorium behave in this manner. For example, 238U produces
13 separate radioactive daughters before finally arriving
at stable 206Pb, as illustrated in figure 14.3. Equations
giving the number of atoms of any member of a decay
series at any time t are presented in a text by Gunter
Faure (1986), and will not be reproduced here. It is interesting, however, to examine the special case in which
the half-life of the parent nuclide is very much longer
than those of the radioactive daughters. In this situation,
it can be shown that after some initial time interval has
passed,
λ1N1 = λ2N2 = λ3N3 = λnNn,
where λ1, 2, . . . , n are the decay constants for each nuclide
(N) in the series. This condition, in which the rate of
decay of the daughters equals that of the parent, is
known as secular equilibrium. It allows an important
simplification of decay series calculations, because the
system can be treated as if the original parent decayed
directly to the stable daughter without intermediate
steps. Secular equilibrium is assumed in uranium
prospecting methods that utilize radiation measurements to calculate uranium abundances.
GEOCHRONOLOGY
One of the prime uses of radiogenic isotopes is, of course,
to determine the ages of rocks. All radiometric dating
systems assume that certain conditions are satisfied:
l. The system, which may be defined as the rock or as
an individual mineral, must have remained closed so
that neither parent nor daughter atoms were lost or
gained except as a result of radioactive decay.
2. If atoms of the daughter nuclide were present before
system closure, it must be possible to assign a value
to this initial amount of material.
FIG. 14.3. Representation of the radioactive decay of 238U to 206Pb through a series of daughter isotopes. Alpha and beta particles
produced at each step are labeled.
Using Radioactive Isotopes
291
PROSPECTING AND WELL-LOGGING TECHNIQUES THAT USE RADIOACTIVITY
Gamma Ray Log
Shale
Increasing Depth
Uranium prospecting tools rely principally on detection of gamma radiation, because alpha and beta particles cannot penetrate an overburden cover of even a
few cm thickness. The Geiger counter is commonly
used for field surveys, although it is a rather inefficient detector. The instrument consists of a sealed glass
tube containing a cathode and anode with a voltage
applied. The tube is filled with a gas that is normally
nonconducting, but when gamma radiation passes
through the gas, it is ionized and the ions accelerate
toward the electrodes. The resulting current pulses
are recorded on a meter or heard as “clicks.”
The scintillation counter is a more efficient tool
for gamma ray detection. Certain kinds of crystals
scintillate; that is, they emit tiny flashes of visible
light when they absorb gamma radiation. These
scintillations are detected by photomultiplier tubes
and converted to electrical pulses that can be counted.
This instrument can be used in either ground or airborne surveys.
A commonly used well-logging tool employs natural radioactivity to identify sedimentary lithologic
boundaries in drill holes. The gamma ray log depends
on scintillation detection of gamma rays produced by
decay of 40K and radioactive daughter products in the
uranium and thorium decay series. These large ions
are incompatible in most crystal structures, but are
accommodated in clay minerals. Therefore, increases
in gamma radiation are ascribed to clay concentrations (that is, shale formations) and, conversely, decreased radiation to cleaner sandstone or limestone
units. An example of a portion of a gamma ray well
log is shown in figure 14.4. Gamma rays from decay
Limestone
Shale
Limestone
Increasing Radiation
FIG. 14.4. An example of a gamma ray well log used in
identifying sedimentary lithologic variations. The horizontal
scale is radioactivity (increasing to the right) and the vertical
scale is depth in the bore hole. Limestone horizons have low
radioactivity relative to shale units because of lower contents
of potassium, uranium, and thorium.
of any one nuclide have a particular energy, and the
energy spectrum of this radiation is resolved in some
gamma ray logs. This permits estimates to be made
of the concentrations of potassium, uranium, and
thorium (assuming secular equilibrium), and the K/U
or K/Th ratios may serve as geochemical signatures
for certain shale horizons, which may be useful in
stratigraphic correlation.
3. The value of the decay constant must be known accurately. (This is particularly critical for nuclides with
long half-lives, because a small error in the decay constant will translate into a large uncertainty in the age.)
4. The isotopic compositions of analyzed samples must
be representative and must be measured accurately.
of these is employed for its special properties, such as rate
of decay, response to heating and cooling, or concentration range in certain rocks.
A number of naturally occurring radionuclides are in
current use for geochronology; some of the most common are listed in table 14.1 and discussed below. Each
In worked problem 14.1, we have seen that radioactive 40K undergoes branching decay to produce two
different daughter products. Its decay into 40Ca is not a
Potassium-Argon System
292
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
TABLE 14.1. Radionuclides Commonly Used in Geochronology
Parent
Stable Daughter
Long-lived radionuclides
40
40
K
Ar
87
87
Rb
Sr
147
143
Sm
Nd
176
176
Lu
Hf
187Re
187Os
232
208
Th
Pb
235
207
U
Pb
238
206
U
Pb
Half-life (yr)
Decay Constant (yr–1)
1.25 × 109
4.88 × 1010
1.06 × 1011
3.57 × 1010
4.56 × 1010
1.40 × 1010
7.04 × 108
4.47 × 109
5.54 × 10–10
1.42 × 10–11
6.54 × 10–13
1.94 × 10–11
1.52 × 10–11
4.95 × 10–11
9.85 × 10–10
1.55 × 10–10
Short-lived and extinct radionuclides
14
14
C
N
5.73 × 103
10
10
Be
Be
1.50 × 106
26
26
Al
Mg
7.20 × 105
129
129
I
Xe
1.60 × 107
182
182W
Hf
9.0 × 10–6
244Pu
131–136Xe1
8.20 × 107
1.21 × 10–4
4.62 × 10–7
9.60 × 10–7
4.30 × 10–8
7.7 × 10–8
8.50 × 10–9
1
Many other fission products are also produced.
useful chronometer, because 40Ca is also the most abundant stable isotope of calcium. As a result, the addition
of radiogenic 40Ca increases its abundance only slightly.
Even though 40Ar is the most abundant isotope of atmospheric argon, potassium-bearing minerals do not commonly contain argon. The radioactive decay of 40K can
therefore be monitored through time by measuring 40Ar.
Each branch of the decay scheme has its own separate
decay constant. The total decay constant λ is therefore
the sum of these:
The commonly adopted value for the total decay constant
and the corresponding half-life are given in table 14.1.
The proportion of 40 K atoms that decay to 40Ar is given
by the ratio of decay constants, λAr /(λAr + λ Ca), which
has a value of 0.105.
We can use equation 14.6 to specify the increase in
40Ar through time:
= 40Ar0 + 0.105 40 K(e λt − 1).
Notice that N, the number of atoms of 40 K, has been
multiplied by λAr /(λAr + λ Ca) to correct for the fact that
most of these atoms will decay to another daughter
nuclide. Because most minerals initially contain virtually
no argon, 40Ar0 = 0 and the equation reduces to:
40
Ar = 0.105 40 K(e λt − 1).
Rubidium-Strontium System
87
Rb produces 87Sr by beta decay, at a rate given in
table 14.1. Unlike the K-Ar system, the radiogenic daughter is added to a system that already contains some 87Sr.
Thus, the difficulty in applying this geochronometer
lies in determining how much of the 87Sr measured in
the sample was produced by decay of the parent isotope.
For the Rb-Sr system, equation 14.6 takes the form:
87Sr
λ = λ Ca + λAr.
40Ar
is calculated from total potassium using its relative isotopic abundance of 0.01167%.
The K-Ar system is useful for dating igneous and, in
a few cases, sedimentary rocks. Sediments containing
glauconite and certain clay minerals that formed during
diagenesis may be suitable for K-Ar chronology. However, because the daughter isotope is a gas, it may diffuse
out of many minerals, especially if they have been buried
deeply or have protracted thermal histories. Metamorphism appears to reset the system in most cases, and
K-Ar dates may record the time of peak metamorphism
in cases for which cooling was relatively rapid. During
slow cooling, argon diffusion continues until the system
reaches some critical temperature (the blocking temperature), at which it becomes closed to further diffusion.
Different minerals in the same rock may have different
blocking temperatures and thus yield slightly different
ages.
(14.8)
Application of equation 14.8 then requires only the
measurement of 40Ar and potassium in the sample; 40 K
= 87Sr0 + 87Rb(e λt − 1).
Another isotope of strontium, 86Sr, is stable and is not
produced by decay of any naturally occurring isotope.
If both sides of the equation are divided by the number
of atoms of 86Sr in the sample (a constant), we will not
affect the equality:
87
Sr/ 86Sr = 87Sr0 / 86Sr + 87Rb/ 86Sr(e λt − 1). (14.9)
Let’s see how the initial ratio of 87Sr0 / 86Sr in this equation, as well as the age, can be derived from graphical
analysis of a suite of rock samples. Equation 14.9 is the
expression for a straight line (of the form y = b + mx).
We will construct a plot of 87Sr/ 86Sr versus 87Rb/ 86Sr;
that is, y versus x, as shown in figure 14.5. Both of these
ratios are readily measurable in rocks. We can assume
that, at the time of crystallization (t = 0), all minerals in
a given rock will have the same 87Sr/ 86Sr value, because
they cannot discriminate between these relatively heavy
isotopes. The various minerals will have different con-
Using Radioactive Isotopes
293
cause it is easier to re-equilibrate adjacent minerals than
different portions of a rock body. Diagenetic minerals
in some sedimentary rocks may also have high enough
rubidium contents that the age of burial and diagenesis
can be determined using this system. The half-life of
87Rb is long, so that this geochronometer complements
the K-Ar system in terms of the accessible time range to
which it can be applied.
Worked Problem 14.3
tents of rubidium and strontium, however, and thus different values of 87Rb/ 86Sr.
This situation is illustrated schematically by the line
labeled t = 0 in figure 14.5. After some time has elapsed,
a fraction of the 87Rb atoms in each mineral have decayed to 87Sr. Obviously, the mineral that had the greatest initial concentration of 87Rb now has the greatest
concentration of radiogenic 87Sr. The position of each
mineral in the rock shifts along a line with a slope of −1,
as shown by arrows in figure 14.5. The slope of the
resulting straight line is (e λt − 1), and its intercept on the
y axis is the initial strontium isotopic ratio, 87Sr0 / 86Sr.
The diagonal line is called an isochron, and its slope increases with time. The fit of data points to a straight line
is never perfect, and a linear regression is required. Analytical errors that lead to dispersion of data about the
line give rise to corresponding uncertainties in ages.
Isochron diagrams such as figure 14.5 can be constructed by using data from coexisting minerals in the
same rock (producing a mineral isochron) or from fractionated comagmatic rock samples that have concentrated different minerals to varying degrees (producing a
whole-rock isochron). Metamorphism may rehomogenize rubidium and strontium isotopes, so that the timing
of peak metamorphism may be recorded by this geochronometer. A mineral isochron is more likely than a
whole-rock isochron to be reset by metamorphism, be-
Larry Nyquist and his coworkers (1979) measured the following isotopic data for a whole-rock (WR) sample and for plagioclase (Plg), pyroxene (Px), and ilmenite (Ilm) mineral separates
in an Apollo 12 lunar basalt (sample 12014).
Mineral
WR
Plg
Px
Ilm
Rb (ppm)
0.926
0.599
0.386
3.76
Sr (ppm)
87
Rb/86Sr
90.4
323
22.7
96.5
0.0296
0.00537
0.0492
0.1127
87
Sr/86Sr
0.70096
0.69989
0.70200
0.70490
What is the age of this rock?
We can determine the age and the initial Sr isotopic ratio,
87
Sr0 / 86Sr, by plotting these data on an isochron diagram (fig.
14.6). A least-squares regression through the data can then be
calculated, as illustrated by the diagonal line in the figure. The
y intercept of this regression line corresponds to an initial isotopic ratio of 0.69964.
87Sr/ 86Sr
FIG. 14.5. Schematic Rb-Sr isochron diagram, illustrating the
isotopic evolution of three samples with time. All samples have
the same initial 87Sr/ 86Sr ratio but different 87Rb/ 86Sr ratios at
time t = 0. The isotopic composition of each sample moves along
a line of slope −1 as 87Rb decays to 87Sr. After some time has
elapsed, the samples define a new isochron, the slope of which is
(e λt − 1). Extrapolation of this isochron to the abcissa gives the
initial Sr isotopic composition.
FIG. 14.6. Rb-Sr isochron for lunar basalt 12014. Isotopic data
for the whole-rock (WR), pyroxene (Px), plagioclase (Plg), and
ilmenite (Ilm) were obtained by Nyquist et al. (1979). The slope
of this isochron corresponds to an age of 3.09 b.y.
294
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Equation 14.9 can be solved for t:
t = 1/λ
ln([(87Sr/ 86Sr
−
87Sr
0
/ 86Sr)]/(87Rb/ 86Sr)
+ 1).
(14.10)
If we now insert the initial Sr isotopic ratio and the measured
isotopic data for one sample (we use WR) into this expression,
we find:
t = 1/(1.42 × 10−11) ln([(0.70096
− 0.69964)/(0.0296)] + 1),
or
t = 3.09 b.y.
This age corresponds to the time that the minerals in the basaltic
magma crystallized. It differs slightly from the age published
by Nyquist and coworkers, because they used a different decay
constant for 87Rb than that given in table 14.1. The estimated
uncertainty in age based on fitting the regression line is +0.11
billion years.
Samarium-Neodymium System
The rare earth elements samarium and neodymium
form the basis for another geochronometer. 147Sm decays
to 143Nd at a rate indicated in table 14.1. The complication in this system is the same as that in the Rb-Sr
system; that is, some 143Nd is already present in rocks
before radiogenic 143Nd is added, so we must find a way
to specify its initial abundance. This can be done by
graphical methods after normalizing parent and daughter to a stable, nonradiogenic isotope of neodymium
(144Nd), in the same way we normalized to 86Sr in the
Rb-Sr system. An expression for the Sm-Nd system analogous to equation 14.9 can then be written:
143
Nd/ 144Nd = 143Nd0 / 144Nd
+ 147Sm/ 144Nd(e λt − 1),
igneous rocks. This system may also be useful, in some
cases, to “see through” a metamorphic event to the
earlier igneous crystallization. Because both parent and
daughter are relatively immobile elements, they are less
likely to be disturbed by later thermal overprints.
Uranium-Thorium-Lead System
The U-Th-Pb system is actually a set of independent
geochronometers that depend on the establishment of
secular equilibrium. 238U decays to 206Pb, 235 U to 207 Pb,
and 232Th to 208Pb, all through independent series of
transient daughter isotopes. The half-lives of all three
parent isotopes are much longer than those of their respective daughters, however, so that the prerequisite
condition for secular equilibrium is present and the production rates of the stable daughters can be considered
to be equal to the rates of decay of the parents at the beginning of the series. The half-lives and decay constants
for the three parent isotopes are given in table 14.1. We
refer to the decay constant for 238U as λ1, that for 235U
as λ 2, and that for 232Th as λ3.
In addition to these three radiogenic isotopes, lead
also has a nonradiogenic isotope (204Pb) that can be used
for reference. (204Pb is actually weakly radioactive, but
it decays so slowly that it can be treated as a stable reference isotope.) We can express the isotopic composition of lead in minerals containing uranium and thorium
by the following equations, analogous to equations 14.9
and 14.11:
206Pb/ 204Pb
= 206Pb0 / 204Pb
+ 238U/ 204Pb (e λ1t − 1),
(14.12a)
207
Pb/ 204Pb = 207Pb0 / 204Pb
+ 235U/ 204Pb (e λ2t − 1),
(14.12b)
208
Pb/ 204Pb = 208Pb0 / 204Pb
+ 232Th/ 204Pb (e λ3t − 1).
(14.12c)
(14.11)
where the subscript 0 indicates the initial isotopic composition of neodymium at the time of system closure.
On an isochron plot of 143Nd/ 144Nd versus 147Sm/ 144Nd,
the y intercept is the initial isotopic ratio 143Nd0 / 144Nd
and the slope of the line becomes steeper with time.
The Sm-Nd radiometric clock is analogous in its
mathematical form to that of the Rb-Sr system, but the
two systems are applicable to different kinds of rocks.
Basalts and gabbros usually cannot be dated precisely
using the Rb-Sr system, because their contents of rubidium and strontium are so low. The Sm-Nd system is ideal
for determination of the crystallization ages of mafic
The subscript 0 in each case denotes an initial lead isotopic composition at the time the clock was set.
Each of these isotopic systems gives an independent
age. In an ideal situation, the dates should all be the same.
In many instances, though, these dates are not concordant, because the minerals do not remain completely
closed to the diffusion of uranium, thorium, lead, or
some of the intermediate daughter nuclides. To simplify
this complicated situation, let’s consider a system without thorium, so that we have to deal with only two radio-
Using Radioactive Isotopes
genic lead isotopes. From equation 14.5, we know that
the amount of radiogenic 206Pb at any time must be:
206Pb*
= 238 U(e λ1t − 1),
or
206Pb*/ 238U
= e λ1t − 1,
(14.13a)
where the superscript * denotes radiogenic lead; that is,
− 206Pb0. Similarly, the amount of radiogenic 207Pb
at any time can be expressed as:
206Pb
207Pb*/ 235U
= e λ2t − 1.
(14.13b)
If a uranium-bearing mineral behaves as a closed system, we know that equations 14.13a and 14.13b should
yield concordant ages (that is, the same value of t). Logically, then, we can reverse the procedure and calculate
compatible values for 206Pb*/ 238U and 207Pb*/ 235U at any
given time t. Figure 14.7 shows the results of such a
calculation. The curve labeled concordia in this figure
represents the locus of all concordant U-Pb systems.
At the time of system closure (t0), no decay has occurred yet, so there is no radiogenic lead present. The sys-
295
tem’s isotopic composition, therefore, plots at the origin
of figure 14.7. At any time thereafter, the system is represented by some point along the concordia curve; numbers along the curve indicate the elapsed time in billions
of years since formation of the system. If, instead, the
system experiences loss of lead at time t1, the system’s
composition shifts along a chord connecting t1 to the
origin. A complete loss of radiogenic lead would displace the isotopic composition all the way back to the
origin, but this rarely happens. If some fraction of lead
is lost, as is the more common situation, the system composition is represented by a point (such as x) along the
chord. A series of related samples with discordant ages
that have experienced varying degrees of lead loss will
define the chord, called a discordia curve, more precisely. After the disturbing event, each point x on the
discordia evolves along a path similar in form but not
coincident with the concordia curve.
The result is that the discordia curve retains its linear
form but changes slope, intersecting the concordia curve
at a new set of points. The time of lead loss is translated
along the concordia curve to t3, and t0 is similarly
FIG. 14.7. Concordia diagram comparing isotopic data for two U-Pb systems. An undisturbed system formed at time t0 will move along the concordia curve; the age at any time after formation is
indicated by numbers (in billions of years) along the curve. If episodic lead loss occurs at time t1,
the position of the system will be displaced toward the origin by some amount, as illustrated by
point x. Other samples that lose different amounts of lead are displaced along the same chord and
give discordant dates. With further elapsed time, the discordia line will subsequently be rotated
(dashed line) as t0 moves to t2 and t1 moves to t3.
296
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
translated to t2, the overall effect being a rotation of discordia about point x. This is illustrated by the dashed
discordia curve in figure 14.7.
If the discordia chord can be adequately defined, then
this diagram may yield two geologically useful dates:
the time of formation of the rock unit (t2), and the time
of a subsequent disturbance (possibly metamorphism
or weathering) that caused lead loss (t3). This interpretation assumes that loss of lead was instantaneous; if
lead was lost by continuous diffusion over a long period
of time, then t3 is a fictitious date without geologic significance. Unfortunately, it is not possible to distinguish between these alternatives without other geological
information.
It is possible to compensate for the effect of lead loss
on U-Pb ages by calculating ages based on the 207Pb/ 206Pb
ratio. This ratio is, of course, insensitive to lead mobilization, because it is practically impossible to fractionate
isotopes of the same heavy element. By combining equations 14.12a and 14.12b, we obtain:
(207Pb/ 204Pb − 207Pb0 / 204Pb)/(206Pb/ 204Pb −
206Pb / 204Pb) = (235U[e λ2t − 1])/(238U[e λ1t − 1]).
0
(14.14a)
The left side of this equation is equivalent to the ratio
of radiogenic 207Pb to radiogenic 206Pb; that is, to
(207Pb/ 206Pb)*. This is determined by subtracting initial
207Pb/ 204Pb and 206Pb/ 204Pb values from the measured
values for these ratios. Consequently, ages can be determined solely on the basis of the isotopic ratios in the
sample, without having to determine lead concentrations.
Moreover, because the ratio 235U/ 238U is constant (1/
137.8) for all natural materials at the present time, ages
can be calculated without measuring uranium concentrations. Equation 14.14a can therefore be expressed as:
(207Pb/ 206Pb)* = 1/137.8([e λ 2t − 1]/[e λ1t − 1]).
(14.14b)
This is the expression for the 207Pb/ 206Pb age, also called
the lead-lead age. The only difficulty in applying this
method is that equation 14.14b cannot be solved for t by
algebraic methods. However, the equation can be iteratively solved by a computer method of successive
approximations until a solution is obtained within an
acceptable level of precision.
The U-Th-Pb dating system is applicable to rocks containing such minerals as zircon, apatite, monazite, and
sphene. These phases provide large structural sites for
uranium and thorium. Igneous or metamorphic rocks of
granitic composition are most suitable for dating by this
method. Several generations of zircons, recognizable by
distinct morphologies, have been found to give distinct
ages in some rocks, and measurements of U-Th-Pb ages
in zoned zircons using ion microprobes have shown
overgrowths with younger ages. The possibility of minerals inherited from earlier events or affected by later
events further complicates the interpretation of U-Th-Pb
ages.
Worked Problem 14.4
John Aleinikoff and coworkers (1985) measured the following
lead isotopic data for a zircon from a granitic pluton in Maine
(units are atom %): 204Pb = 0.010, 206Pb = 90.6, 207Pb = 4.91,
208Pb = 4.43. The measured U concentration was 1396 ppm,
and that for Pb was 62.2 ppm. What are the 206Pb/238U,
207Pb/235U, and 207Pb/206Pb ages for this rock, and are they
concordant?
We begin by determining the proportions of 235U and 238U
in this sample. All natural uranium has 235U/238U = 1/137.8,
corresponding to 99.27% 238U. Aleinikoff and coworkers assumed that the nonradiogenic lead in this sample had isotopic
ratios of 204:206:207:208 = 1:18.2:15.6:38, corresponding to
an average atomic weight of 207.23. From these proportions,
we can calculate the following ratios:
238U/ 204Pb
= (1396/62.2)(207.23/238.03)(99.27/0.010)
= 193,970,
235U/ 204Pb
= 193,970/137.8 = 1407.6,
206Pb/ 204Pb
= (90.6/0.010) − 18.2 = 9041.8,
207Pb/ 204Pb
= (4.91/0.010) − 15.6 = 475.4,
206Pb
204Pb
= 18.2,
/ 204Pb
= 15.6.
207Pb
0/
0
We can now insert values into equation 14.12 and solve for t.
The expression for the 206Pb/238U age, corresponding to equation 14.12a, is:
t206 = 1/λ1 ln([(206Pb/ 204Pb)
− (206Pb0 / 204Pb)]/[238U/ 204Pb] + 1).
Substitution of the ratios above into this equation gives:
t206 = 1/(1.55 × 10−10) ln([9041.8 − 18.2]/[193,970]
+ 1) = 293 m.y.
The 207Pb/ 235U age, calculated from the analogous expression derived from equation 14.12b, is:
t207 = 1/(9.85 × 10−10) ln([475.4 − 15.6]/[1407.6] + 1)
= 287 m.y.
Using Radioactive Isotopes
Subtracting the initial ratios for the isotopic composition of nonradiogenic lead gives:
207
Pb/ 204Pb = 491 − 15.6 = 475.4,
and
206
Pb/ 204Pb = 9060 − 18.2 = 9041.8.
Dividing these results gives:
(207Pb/ 206Pb)* = 475.4/9041.8 = 0.0526.
We substitute this ratio into equation 14.14b and solve for t to
find that the 207Pb/206Pb age is 308 m.y.
These ages are similar and date the approximate time of
crystallization of the pluton. This sample would plot on the concordia curve, because ages determined by both U-Pb systems
are nearly identical.
297
this nuclide would have remained in the silicate mantle, and its
subsequent decay would have produced an excess of radiogenic
182
W relative to nonradiogenic 184W in the mantle. An anomalous 182W/ 184W ratio in the mantle would then be passed on to
basaltic magmas derived from it.
Geochemists Alex Halliday and Der-Chuen Lee (1999) compared the tungsten isotopic composition in terrestrial silicate
rocks with those measured in meteorites (fig. 14.8). In figure
14.8, tungsten isotopes are expressed as εW, defined as the deviation between 182W/184W in a sample relative to that ratio in
a chondrite of identical age. The 182W/184W ratio in the Earth
is indistinguishable from that of chondritic meteorites that have
never experienced metal fractionation. However, iron meteorites,
which are samples of the cores of differentiated asteroids, exhibit 182W deficits. The absence of excess 182W in the silicate
Extinct Radionuclides
All of the naturally occurring radionuclides that we
have considered so far have very long half-lives, so that
most of the parent isotopes still occur naturally on the
Earth. However, some terrestrial rock samples and meteorites contain evidence of the former presence of
now extinct radionuclides that had short half-lives. The
stable daughter products and rates of decay for 26Al,
129 182
I, Hf, and 244Pu are given in table 14.1. That these
isotopes occurred at all as “live” radionuclides demands
that the samples that contain them formed very soon
after the nuclides were created. In worked problem 14.5,
we see how decay of now extinct 182Hf can be used as
a geochronological tool. Because of rapid decay rates, a
few extinct radionuclides such as 26Al may have been
important sources of heat during the early stages of
planet formation. The implications of the former presence of 26Al and 182Hf in meteorites will be considered
in chapter 15.
Worked Problem 14.5
How can the now extinct radionuclide 182Hf be used to constrain the time of core formation in the Earth? 182Hf decays to
182
W with a half-life of only 9 million years, so any 182Hf present in the early Earth has long since decayed. Hafnium is partitioned into silicates (that is, it is lithophile), and tungsten is
concentrated in metal (it is siderophile). Consequently, tungsten
was partitioned into the Earth’s core when it separated from the
mantle. If core separation happened after 182Hf had decayed,
then all of the daughter 182W would have been sequestered in
the core. If the core formed while 182Hf was still alive, however,
FIG. 14.8. The isotopic composition of tungsten (182W/184W), expressed as εW values, of terrestrial samples compared with those
of meteorites. Terrestrial values are indistinguishable from those
of carbonaceous chondrites, which have never experienced metal
fractionation, but are distinct from iron meteorites, which formed
as differentiated cores in asteroids. These data suggest that core
formation in the Earth occurred after 182Hf had decayed to 182W,
so that no excess radiogenic tungsten was produced in the mantle.
(After Halliday and Lee 1999.)
298
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
portion of the Earth suggests that the Earth’s core formation
must have happened many 182W half-lives after core formation
in asteroids. Based on the elapsed time that would be necessary
to resolve a 182W/184W ratio higher than in chondrites, Halliday
and Lee estimated that core formation in the Earth occurred
>50 m.y after the formation of iron meteorites. A recent revision
in the initial ratio of 182Hf to stable 180Hf (Yin et al. 2002) indicates a value half that assumed by Halliday and Chuen, which
lowers the time of formation of the Earth’s core to 29 m.y.
Fission Tracks
If you have examined micas or cordierite under the
petrographic microscope, you may have noticed zones
of discoloration around tiny inclusions of zircon and
other uranium-bearing minerals. These pleochroic haloes
are manifestations of radiation damage produced by
alpha particles. The crystal lattices of minerals that contain dispersed uranium atoms also record the disruptive
passage of alpha particles in the form of fission fragments. The fragments are very small, but they nevertheless produce fission tracks as they pass through the host
mineral. Although the tracks are only ∼10 microns long,
they can be enlarged and made optically visible by etching with suitable solutions, because the damaged regions
are more soluble than the unaffected regions. A microscopic view of fission tracks can be seen in figure 14.9.
The density of fission tracks in a given area of sample can
be determined by counting the tracks under a microscope.
These tracks, except in rare instances, are produced
solely by spontaneous fission of 238U. The observed track
density is proportional to the concentration of 238U in
the sample and to the amount of time over which tracks
have been accumulating; that is, the age of the host mineral. If a means can be found to measure the uranium
concentration in the sample, fission tracks can then be
used for geochronology. The uranium abundance can be
determined by measuring the density of tracks induced
by irradiating the sample in a reactor, where bombardment with neutrons causes more rapid decay (induced
radioactivity is explained more fully in the next section).
P. B. Price and Robert Walker (1963) formulated a solution for the fission track age equation (presented here
without derivation):
t = ln(1 + [ρs /ρI][λ 2 φσU/λ F ])1/λ 2.
Explaining the terms in this equation illustrates how a
fission track age is determined. λ 2 and λ F are the decay
FIG. 14.9. Photomicrograph of fission tracks in mica. Radiating
tracks emanate from point sources, such as small zircon inclusions; isolated tracks probably formed from single U or Th atoms.
(Courtesy of G. Crozaz.)
constants for 238U and spontaneous fission (8.42 × 10−17
yr−1), respectively. The area density of fission tracks in
the sample, ρs, is a function of age and uranium content.
We noted earlier that the age can be determined uniquely
if the uranium content can be measured. Uranium concentration is measured by placing the sample (after ρs has
been visually counted) into a nuclear reactor, which induces rapid 238U decay and increases the number of fission tracks in the sample. The remaining terms in the
equation above are related to this irradiation process: ρi
is the area density of induced tracks, φ is the flux of neutrons passed through the sample, σ is the target cross
section for the induced fission reaction, and U is the
(constant) atomic ratio 235U/238U.
Annealing of the sample at elevated temperatures
causes fission tracks to fade, as the damage done to the
crystal structure is healed. The annealing temperatures
for most minerals used in fission track work are quite
modest (several hundred degrees Centigrade or less), so
the ages derived by this method must be interpreted as
times the host rocks cooled through temperatures at
which annealing ceased, commonly called cooling ages.
Such phases as apatite, sphene, and epidote are useful
for the interpretation of cooling histories, and they begin
Using Radioactive Isotopes
299
to retain fission tracks at temperatures that are different
from the blocking temperatures for argon isotopes. Fission track annealing temperatures, like radiogenic isotope blocking temperatures, can be extrapolated from
experimental data. The determination of cooling history
using a combination of fission tracks and other isotopic
methods is illustrated in the following worked problem.
path. The temperature-time curve in figure 14.10 implies that
this body cooled from >500o C through ∼100o C within a span
of only 4 million years. This rapid cooling could not have taken
place very deep in the crust and must indicate uplift and erosion. Nielson and coworkers calculated that uplift proceeded at
a rate between 0.6 and 1.1 mm yr−1 for this interval, based on
an assumed geothermal gradient of 30o km−1.
Worked Problem 14.6
GEOCHEMICAL APPLICATIONS
OF INDUCED RADIOACTIVITY
How can we determine cooling history from isotope blocking
temperatures and fission track retention ages? The elucidation
of cooling history usually requires integration of several different kinds of data, tempered with thoughtful geologic reasoning.
D. L. Nielson and coworkers (1986) characterized the temperature history of a Cenozoic pluton in Utah in the this way.
Critical observations for their interpretation of the cooling history are:
1. The K-Ar age of a hornblende sample from the pluton is
11.8 million years. The argon blocking temperature for hornblende is ∼525o C.
2. The K-Ar age of a biotite sample from the pluton is 10.8 million years. The argon blocking temperature for biotite is
∼275o C.
3. The fission track ages for zircons in these rocks range from
8.3 to 8.9 million years. The temperature at which zircon
begins to retain tracks depends on the cooling rate, but is
generally ∼175o C.
4. The fission track ages for apatites range from 8.1 to 9.1 million years. Apatite begins to retain tracks at ∼125o C.
Using these data, we can construct a temperature versus time
plot that describes the cooling history of this pluton (fig. 14.10).
We can also speculate about geologic controls on the cooling
In medieval times, alchemists toiled endlessly to turn various metals into gold. They were unsuccessful, but only
because they did not know about nuclear irradiation. It
is now possible to make gold, but it is still not economical, because the starting material must be platinum.
Nuclei that are bombarded with neutrons, protons,
or other charged particles are transformed into different
nuclides. In many cases, the nuclides produced by such
irradiation are radioactive. Monitoring their decay provides useful analytical information. Nuclear irradiation
can occur naturally or can be induced artificially. Here
we consider some geochemical applications of both
situations.
Neutron Activation Analysis
Nuclear reactions caused by neutron bombardment
are commonly used to perform quantitative analyses
of trace elements in geological samples. Fission of 235U
in a nuclear reactor produces large quantities of neutrons,
FIG. 14.10. The cooling history of an igneous pluton, inferred from argon blocking temperatures
for hornblende and biotite and fission track ages for zircon and apatite. (Data reported by Nielson
et al. 1986.)
300
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
which can be used to induce nuclear changes in mineral
or rock samples. Slow neutrons are readily absorbed by
the nuclei of most stable isotopes, transforming them
into heavier elements. Because many of the irradiation
products are radioactive, this procedure is called neutron
activation.
In earlier sections, we described radioactivity in terms
of N, the number of atoms remaining. In neutron activation analysis, radioactivity is monitored using activity
(A), defined as:
A = cλN,
(14.15)
where c is the detection coefficient (determined by calibrating the detector) and λN is the rate of decay. The
easiest way to determine the concentration of an element
by neutron activation is by irradiating a standard containing a known amount of that element at the same
time as the unknown. After irradiation, the activities of
both standard and unknown are measured at intervals
using a scintillation counter to detect emitted gamma
rays. (This step may be complicated for geological
samples because they contain so many elements. It is
necessary to screen out radiation at energies other than
that corresponding to the decay of interest. This can
normally be done by adjusting the detector to filter unwanted gamma radiation but, in some cases, chemical
separations may be required to eliminate interfering
radioactivities.) The activities are then plotted, as shown
in figure 14.11; the exponential decay curves are transformed into straight lines by plotting ln A. Extrapolation
of these lines back to zero time gives values of A0, the
activities when the samples were first removed from the
reactor. The amount of element X in the unknown can
then be determined from the following relationship:
A0 unkn /A0 stnd = amount of Xunkn /
amount of Xstnd.
The concentration of element X in the unknown can be
calculated by dividing its amount by the measured weight
of sample.
It is also possible to calculate A0 theoretically by using
the following relationship:
A0 = cN0 φσ(l − e −λti),
(14.16)
where c is the detection coefficient, N0 is the number of
target nuclei, φ is the neutron flux in the reactor, σ is the
neutron capture crosssection of target nuclei, and ti is
the time of irradiation. This computation, normally done
FIG. 14.11. Plot of the measured activity A of an unknown and a
standard after neutron irradiation, as a function of time. Extrapolation of these straight lines gives A0, the activity at the end of
irradiation. These data are then used to determine the concentration of the irradiated element in the unknown.
by computer, makes it easier to handle many elements
simultaneously in complex samples.
40Argon-39Argon
Geochronology
One of the disadvantages of the K-Ar geochronometer is that gaseous argon tends to diffuse out of many
minerals, even at modest temperatures. Loss of the radiogenic daughter can thus lead to an erroneously young
radiometric age. 40Ar-39Ar measurements provide a way
to get around the problem of argon loss. Let’s examine
how this technique works.
We might expect that the rims of crystals would lose
argon more readily than the crystal interiors because
of shorter diffusion distances. In this case, the ratio of
radiogenic 40Ar to 39K, a stable isotope of potassium,
will be highest in crystal centers and lower in crystal
rims. Irradiating the sample with neutrons in a reactor
causes some 39K to be converted to 39Ar. Consequently,
the ratio 40Ar/39K becomes 40Ar/39Ar, which is readily
analyzed using mass spectrometry.
The number of 39Ar atoms produced during the irradiation is given by:
39Ar
∫
= 39Kti φσ dε,
(14.17)
where ti is the irradiation time, φ is the neutron flux in
the reactor, σ is the neutron capture cross section, and
Using Radioactive Isotopes
the integration is carried out over the neutron energy
spectrum dε. The number of radiogenic 40Ar atoms produced by decay of 40 K is given by equation 14.8. Dividing equation 14.8 by 14.17 gives the 40Ar/39Ar ratio in
the sample after irradiation:
40Ar/ 39Ar
= (0.124/ti)(40 K/39 K)(e λt − 1)/
( ∫ φσ dε).
(14.18)
Because equation 14.18 can be solved for t, measurement
of the 40Ar/ 39Ar ratio of a sample defines its age.
In practice, the sample is heated incrementally, and
the isotopic composition of argon released at each temperature step is measured. Figure 14.12 illustrates a series of 40Ar/39Ar ages, calculated using equation 14.18,
for samples of mica and amphibole in a blueschist
sample from Alaska, as a function of total 39Ar released.
Each step represents a temperature increment of 50o,
starting at 550o C. If this sample had remained closed to
argon loss since its formation, all of the ages would have
been the same. The spectrum of ages that we see in figure 14.12 results from 40Ar leakage over time. Argon from
the cores of crystals is released at the higher temperature
steps, and the plateau of ages at ∼185 million years seen
at high temperatures represents the time of metamorphism for this sample, because the mica blocking temperature is approximately equal to the peak metamorphic
temperature.
Cosmic-Ray Exposure
In both of the examples we have just considered, the
samples were irradiated intentionally by humans. Natural irradiation can also have useful geochemical applications. Cosmic rays, consisting mostly of high energy
protons but also of neutrons and other charged particles
generated in the Sun and outside the Solar System offer
a natural irradiation source. Isotopes produced by interaction of matter with cosmic rays are called cosmogenic
nuclides.
Radioactive 14C is produced continuously in the atmosphere by the interaction of 14N and cosmic rays.
This isotope is then incorporated into carbon dioxide
molecules, which in turn enter plant tissue through photosynthesis. The concentration of cosmogenic 14C in living
plants (and in the animals that eat them) is maintained
at a constant level by atmospheric interaction and isotope decay. When the plant or animal dies, addition of
14C from the atmosphere stops, so its concentration decreases due to decay. Measurement of the activity of 14C
in plant or animal remains thus provides a way to determine the time elapsed since death.
The activity A of carbon from a plant or animal that
died t years ago is given by:
A = A0e−λt.
(14.19)
250
Age (m.y.)
200
150
Mica
100
Amphibole
50
0
0
0.2
0.4
301
0.6
0.8
1.0
Fraction of 39Ar
FIG. 14.12. 40Ar-39Ar age versus cumulative 39Ar released for mica and amphibole samples from
blueschist. The argon was released at successive temperature steps; each step in this figure corresponds to a 50° increment, starting at 550°C. The plateau of ages at 185 million years corresponds
to radiogenic argon from the interiors of crystals. (After Sisson and Onstott 1986.)
302
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
where A0 is the 14C activity during life = 13.56 dpm g−1
(expressed as disintegrations per minute per gram of
sample), and the decay constant λ is given in table 14.1.
(Compare equations 14.19 and 14.15. Both express activity in terms of λN times a constant, c or A0.) Radiocarbon ages have only a limited geologic range because
of the short half-life of 14C, but they are very useful for
archaeological dating and for some environmental geochemistry applications.
Cosmic rays sometimes fragment other nuclides, a
process called spallation. 10Be and 26Al are spallation
products formed from oxygen and nitrogen in the atmosphere. Both are rapidly transferred to the oceans
or to Earth’s surface in rain or snow, where they can be
incorporated into sediments on the ocean floor or into
continental ice sheets. After deposition, these cosmogenic nuclides decay at rates given in table 14.1, so they
can be used to date sediment or ice cores.
Rocks and soils at the Earth’s surface are also exposed
to cosmic rays, so the accumulation of cosmogenic nuclides in them can be used to measure the durations of
sample exposure. These time intervals, called cosmic ray
exposure ages, assess how long materials were situated
on the Earth’s surface, because a meter or more of overburden effectively shields out cosmic rays. Measured
concentrations of cosmogenic nuclides are converted to
exposure ages by knowing nuclide production rates,
which can be determined experimentally. Quartz is an
especially favorable target for measurement of cosmogenic 10Be and 26Al because of its simple chemistry, resistance to weathering, and common occurrence in many
kinds of rocks. By measuring exposure ages, it is possible
to quantify the rates of uplift and erosion. In a later section, we see how cosmogenic 10Be can be employed as a
tracer for the subduction of surface materials into the
mantle and as a way of constraining subduction rates.
RADIONUCLIDES AS TRACERS
OF GEOCHEMICAL PROCESSES
In biological research, certain organic molecules may be
“tagged” with a radioactive isotope so that their participation in reactions or progress through an organism
can be followed. Radionuclide tracers can be used in an
analogous way to understand geochemical processes
that have already taken place within the Earth. This usually requires a knowledge of the geochemical behavior
of parent and daughter elements. In this section, we consider several examples of the use of these tracers.
Heterogeneity of the Earth’s Mantle
We have already noted that most radioactive nuclides
and their decay products cannot be fractionated by mass.
However, radioactive isotopes generally have very different solid-liquid distribution coefficients from their
stable daughter isotopes, so that partial melting can
fractionate radioactive elements from their radiogenic
daughters. As melts have been progressively extracted
from the mantle to form the crust, the abundances of
radioactive isotopes in these reservoirs have changed,
and these differences have been magnified by subsequent
decay of the fractionated radioactive isotopes. The resulting isotopic heterogeneity within the mantle can be
monitored by analyzing magmas that are derived from
different mantle depths.
Samarium and neodymium are partitioned differently
between the continental crust and mantle. Although both
elements are incompatible and thus are fractionated into
the crust, neodymium is more incompatible and so is
∼25 times more abundant in the crust than in the mantle, whereas samarium is only ∼16 times more abundant.
This means that the Sm/Nd ratio is higher in the mantle
than in the crust. 147Sm decays to 143Nd over time, so
the ratio of radiogenic 143Nd to nonradiogenic 144Nd
changes in both crust and mantle. It is thought that the
mantle originally contained samarium and neodymium
in the same relative proportion as in chondritic meteorites. The deviation between 143Nd/144Nd measured in
a mantle-derived rock and the same ratio in a chondrite
of identical age is called ε Nd (analogous to ε W, defined in
worked problem 14.5) and can be used to estimate the
degree to which the mantle may have evolved from its
original pristine state.
Similar arguments can be used to trace the evolution
of the 87Sr/ 86Sr ratio, which is related to the Rb/Sr ratio
through the decay of 87Rb to 87Sr. In this case, rubidium
is more incompatible than strontium, so the Rb/Sr ratio
is higher in the crust and the crust becomes more radiogenic over time—the opposite of the Sm-Nd system.
The isotopic composition of strontium can also be expressed as εSr.
If basalts are derived from melting of mantle material,
their initial neodymium and strontium isotopic ratios
Using Radioactive Isotopes
303
NATURAL NUCLEAR REACTORS
Certain heavy nuclides (235U being the only naturally
occurring example) are fissile; that is, on absorbing a
neutron, they split into two nuclear fragments of
unequal mass. Because fission is exothermic, it is the
energy source for nuclear power reactors and for
the atomic bomb. When 235U fissions, it releases additional neutrons that collide with other 235U nuclei,
prompting further fission reactions. The nuclides
produced by fission have higher ratios of N to Z than
stable nuclides and are thus radioactive. This process
is used to advantage in the design of breeder reactors,
which produce plutonium and other transuranic elements. It also accounts for the radioactivity of nuclear
wastes.
In 1972, workers at a nuclear-fuel processing plant
in France found that some of the uranium ore shipped
from the Oklo mine in the Republic of Gabon, Africa,
was deficient in 235U relative to 238U. By exquisite analytical detective work, scientists at the French Atomic
Energy Commission eventually traced this uranium
isotopic anomaly to 17 individual pockets within the
Oklo deposit and the surrounding area. They had
discovered the intact remains of Precambrian natural
nuclear reactors. Besides the depletion in 235U in
the spent nuclear fuel, these pockets contain fissionproduced nuclides of Nd, Sm, Zr, Mo, Ru, Pd, Ag,
Cd, Sn, Cs, Ba, and Te. These isotopes are also part
of the radioactive waste from modern man-made
nuclear reactors.
The Oklo reactors are small (usually 10–50 cm
thick) seams of sandstone-hosted uraninite ore man-
should be the same as those of their mantle sources at the
time of melting. Measured εNd and εSr values for young
basalts from various tectonic settings are summarized
in figure 14.13. Midocean ridge basalts (MORB) have
positive εNd and negative εSr values; that is, they have
high 143Nd/144Nd ratios and low 87Sr/86Sr ratios that are
complementary to those of continental crust. From this,
we infer that the upper mantle source regions for MORB
have been depleted of materials used to form the continents. Ocean island basalts plot along an extension of
tled by clay. Today, that ore would be unable to initiate self-sustaining neutron chain reactions, because
the proportion of 235U in natural uranium is too low.
However, 1.9 billion years ago when these reactors
were active, the proportion of 235U was nearly five
times greater. The high uranium concentration in
quantities exceeding the critical mass was a necessary
condition for the Oklo reactors. These ores contained
few “pile poisons,” which absorb neutrons and prevent a chain reaction. In modern reactors, neutrons
are slowed (moderated) by water, which allows them
to be absorbed by other nuclei. Geological observations show that water was also present as a moderator in the Oklo natural reactors. Moreover, the
deposits also contain fission products of plutonium
and other transuranic elements, which indicates that
they functioned as fast breeder reactors. Breeding
fissile material would have allowed the reactors to
operate continuously for longer periods of time. You
can read more about the Oklo reactors in a paper published by F. Gauthier-Lafaye and coworkers (1989).
Besides providing information on the physics of a
fascinating natural phenomenon, these reactors are
important for understanding radioactive waste containment. Oklo is the only occurrence in the world
where actinides and fission products have been in a
near-surface geological environment for an extremely
long period of time. The fission products at these sites
allow characterization of the geochemical mobility
of various isotopes and thereby permit assessment of
the effectiveness of radionuclide waste repositories.
the MORB trend in figure 14.13, defining what is commonly called the mantle array. The position of ocean
island basalts in the diagram indicates that their source
is less depleted than that of MORB. These basalts also
show a wider range in isotopic composition. Ocean island basalts may be derived from the relatively undepleted lower mantle and are carried through the upper
mantle in plumes, where they mix to varying degrees
with depleted materials. Magmas in subduction zones
(not shown in fig. 14.13) often plot off the mantle array,
304
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 14.13. Neodymium and strontium isotopic compositions of basalts. MORBs and ocean island
basalts define a diagonal mantle array. MORBs are derived from a homogenized, depleted upper
mantle, whereas ocean island basalts represent mixing of magmas derived from a less depleted
lower mantle with upper mantle materials. (After Anderson 1994.)
reflecting the effects of mixing of seawater expelled
from the subducted slab and crustal materials during
ascent.
In terms of its radiogenic isotopes, the mantle can be
visualized as consisting of two layers: the lower portion
undifferentiated and the upper mantle highly fractionated.
The lower mantle has presumably remained relatively
undepleted since its formation, sampled only infrequently
at hot spots to produce flood basalts and oceanic islands.
However, periodic sinking of crustal plates into the lower
mantle provides a mechanism for enriching it in incompatible elements. The upper mantle has been continuously cycled by the formation and subduction of oceanic
crust, as well as the production of continental crust made
in magmatic arcs at subduction zones. Its isotopic composition has become homogenized but remains relatively
depleted in incompatible elements. This two-layer model
of the mantle was previously presented in chapter 12,
based on trace element abundances.
The estimated +12εNd value of the upper mantle is
counterbalanced by the −15εNd value for continental
crust. Knowing these values and the mass of the continental crust, it is possible to perform a mass balance
calculation to determine the proportion of the depleted
upper mantle. The mass of the continental crust is ∼2 ×
1025 g and its neodymium abundance is ∼50 times greater
than the upper mantle. The mass of the upper mantle is
thus 100 × 1025 g, which corresponds to a thickness on
the order of 600–700 km. This depth range coincides with
a prominent seismic discontinuity separating the upper
and lower mantles.
If this view of a layered mantle is correct, other isotopic systems should be correlated with neodymium and
strontium isotopic data. As an example, εPb, the deviation in 206Pb/204Pb from chondritic evolution, varies systematically, so we should actually consider a Nd-Sr-Pb
mantle array. Stan Hart and coworkers (1986) suggested
that inclusion of additional isotopic systems may require
that compositional models for the mantle be expanded
to as many as five components.
Magmatic Assimilation
In some cases, ascending magmas are contaminated
by assimilation of crustal materials, with the result that
their compositions are altered. Radiogenic isotopes pro-
Using Radioactive Isotopes
vide the least ambiguous way of recognizing this process.
In uncontaminated basalts, the initial 87Sr/86Sr ratio is
the same as that of the mantle source region. This ratio
always has a relatively low value (ranging from ∼0.699
to 0.704, depending on the age of the rock—the mantle
slowly accumulates radiogenic 87Sr over time), because
the amount of parent rubidium relative to strontium in
mantle rocks is low. Rb/Sr ratios in rocks of the continental crust are much higher; therefore, over time, crustal
rocks evolve to contain more radiogenic 87Sr. We can
infer then that magmas derived from mantle sources
should have low initial 87Sr/ 86Sr ratios. Conversely, those
derived from, or contaminated with, continental crust
should have high initial 87Sr/ 86Sr ratios. A corollary is
that rocks related only through fractional crystallization
of the same parental magma should have identical
87Sr / 86Sr ratios.
0
Other isotopic systems can also be used to test for
assimilation of crustal materials. Contaminated basalts
should have lower 143Nd/ 144Nd ratios than uncontaminated rocks. It is not possible to predict the exact isotopic composition of crustal lead, but if Pb isotopes can
be analyzed in plausible contaminating materials, its mixing effects can be assessed.
Assimilation is actually a rather complicated process,
because it requires a great deal of heat. In chapter 12,
we learned that melting requires enough heat to raise
the temperature of the wall rocks to their melting points
plus the heat necessary to fuse them. Most magmas are
already below their liquidus temperatures, so the only
plausible source of heat is exothermic crystallization of
the magmas themselves. Consequently, we might expect
assimilation and fractional crystallization to occur concurrently. An equation to model the change in isotopic
composition of a magma experiencing both assimilation
and fractionation is given below. Here we assume that a
magma body of mass Mo is affected by two rate processes: the rate at which the magma assimilates a mass of
rock, Ma; and the rate at which crystals separate during
fractionation of the magma, Mc. At any time during the
combined processes, the mass of magma remaining is M,
and the parameter F = M/Mo describes the mass of magma
relative to the mass of the original magma. Any isotopic
ratio in the magma Im resulting from the combined effects
of assimilation and fractional crystallization is given by:
o F −zI o }/
Im = {(r/[r − 1])(Ca/z)(1 − F−z)Ia + C m
m
−z
o
{(r/[r − 1])(Ca/z)(1 − F ) + C m F −z}. (14.20)
305
In this equation, r is defined as Ma/Mc, Ca is the concentration of the element of interest in the assimilated rock,
o is its initial concentration in the original magma, and
Cm
o and I are the initial isotopic ratios of the magma and
Im
a
assimilated rocks, respectively. The term z is defined as:
z = (r + D − 1)/(r − 1),
where D is the bulk distribution coefficient between the
fractionating crystals and the magma. Equation 14.20
predicts the isotopic composition of a magma (Im) resulting from combined assimilation and fractionation.
Im could be any isotope ratio, for example, 87Sr/ 86Sr, or
any normalized parameter describing such a ratio, such
as εSr.
Worked Problem 14.6
By considering the combined effects of assimilation and fractional crystallization, geochemist Don DePaolo (1981) showed
that elevated 87Sr/86Sr ratios in modern andesitic lavas in the
Andes resulted from crustal contamination. The inferred isotopic compositions of the initial magma and a plausible crustal
contaminant (wallrock) are illustrated in figure 14.14. Measured
data for most Andean volcanic rocks (open circles in fig. 14.14)
do not lie on a straight mixing line connecting these compositions, which might suggest that mixing was not important in
determining the isotopic compositions of these rocks. However,
the combined effects of assimilation and fractional crystallization
give a different result.
Models for the combined processes can be constructed using
equation 14.20. To illustrate, we calculate the magma isotopic
composition (Im) using the following parameters: the original
magma contained 400 ppm Sr and 10 ppm Rb and had an
initial strontium isotopic ratio 87Sr/ 86Sr = 0.7030. The contaminant had 500 ppm Sr and 40 ppm Rb, with an initial ratio
87
Sr/ 86Sr = 0.71025. We assume that r = 0.8, DSr = 0.5, and
F = 2. The corresponding value for z is then:
z = (0.8 + 0.5 − 1)/(0.8 − 1) = −1.5.
Substitution of the above parameters into equation 14.20 gives:
Im = {(0.8/−0.2)(500/−1.5)(1 − 21.5)(0.71025)
+ (400)(21.5)(0.7030)}/
{(0.8/−0.2)(500/−1.5)(1 − 21.5) + (400)(21.5)}
= 0.7078.
This calculated value is only one point on a curve like those in
figure 14.14, which illustrates how the strontium isotopic ratio
changes during the combined processes. The curves actually
shown in figure 14.14 are the values for Andean volcanics calculated by DePaolo. He assumed that DSr had a range of values
as shown in figure 14.14 (we might expect that DSr would vary
306
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 14.14. Strontium isotopic data for volcanic rocks of the
Peruvian Andes. The data points do not lie on a simple mixing line
connecting the isotopic compositions of initial magma and wallrock contaminant, but do lie near calculated mixing lines for an
assimilation-fractional crystallization model with various values of
DSr. It is expected that DSr increases in the latter stages of magma
evolution. This model appears to explain the observed isotopic
data fairly well. (After DePaolo 1981.)
during fractional crystallization and assimilation as different
phases crystallized and the liquid composition changed). The
measured strontium isotopic compositions and Rb/Sr ratios in
the lavas fall within the calculated curves, thus supporting the
model of assimilation and concurrent fractional crystallization.
Subduction of Sediments
Sediments deposited near oceanic trenches and subsequently subducted into the mantle could conceivably
be melted and incorporated into arc volcanics. However,
some geological and geophysical studies have stressed
the mechanical difficulty of subducting low density sediments into the mantle. Radiogenic isotopes that are
concentrated in such sediments may serve as tracers to
follow their fate. These sediments characteristically have
high 208Pb/ 204Pb and 207Pb/ 204Pb ratios. We have already
seen that uranium and thorium are incompatible elements
that are concentrated in the crust, and the high radiogenic lead contents of these sediments are the result of
their derivation from old U- and Th-enriched rocks on
continents. The isotopic compositions of lead in some
arc volcanics and in abyssal sediments are very similar,
but are so distinct from those of oceanic basalts that
recycling of crustal materials must occur. Mass balance
calculations suggest that 1–2% sediment in the source
could account for the observed lead isotope data.
This conclusion is supported by hafnium isotopic
data on arc volcanics. We have not discussed the Lu-Hf
system, but it is similar in most respects to the Sm-Nd
system, with specifics given in table 14.1. Hafnium is
partitioned strongly into the mineral zircon, which is
commonly concentrated in continent-derived sands. Thus,
there should be a major difference in Lu/Hf ratios between these sands and deep-sea muds. This behavior is
distinct from Sm/Nd, which shows no fractionation during sedimentation. Therefore, the isotopic variations in
hafnium and neodymium expected from mixing mantle
with various types of sediment can be predicted and compared with data from arc volcanics. Jonathan Patchett
and his collaborators (1984) estimated that <2% of nearly
equal proportions of continental sand and oceanic mud
could account for the measured Hf-Nd isotopic array.
The cosmogenic isotope 10Be can also be used to shed
light on this problem. The short half-life of this isotope,
only 1.5 million years (table 14.1), is comparable to the
time scale for subduction and magma generation. Young
pelagic sediments contain appreciable amounts of 10Be,
so it is possible that this isotope might make the round
trip from the Earth’s surface into the mantle and back
again if subduction and melting are fast enough. Analyses indicate that lavas that should not have incorporated
crustal materials (MORB, ocean island basalts) do not
contain significant amounts of 10Be, but this isotope is
enriched by factors of 10–100 in some arc volcanic rocks
(e.g., Aleutians, Andes, Kurile arcs). 10Be is absent in the
lavas of some arcs, but its absence does not necessarily
prove the absence of sediments, because the 10Be signal
depends on the rate of subduction and melting. Julie
Morris (1991) provided a comprehensive review of the
subject.
It might be argued that 10Be in lavas reflects cosmic
ray exposure on the surface after eruption, rather than
incorporation of subducted sediments. However, analyses of modern lavas show a constant ratio of 10Be to noncosmogenic 9Be in all minerals of the rocks, indicating
Using Radioactive Isotopes
that 10Be must have been incorporated prior to igneous
crystallization. Estimates of the amount of sediments incorporated into arc lavas are generally <3%. The time
scales for subduction to mantle depths, transfer of material from the slab, and subsequent melting, ascent, and
eruption are still debated. Constraints from decay of 10Be
can be coupled with U-Th disequilibrium systematics to
give times ranging from a few hundred thousand years
to as little as 10,000 years.
of 0.7067 during the Jurassic and increasing again to the
current high value of 0.7090. The isotopic composition
of marine carbonates mimics that of the ocean in which
they formed, so the oceans have varied in strontium
isotopic composition over time. In fact, changes in the
oceanic 87Sr/ 86Sr ratio have now been documented with
sufficient precision that this isotope ratio can be used to
date some fossil shells and carbonate sediments (the technique is most precise for intervals when the 87Sr/ 86Sr ratio
is rapidly changing due to climatic or tectonic events).
Different oceans drain continents with different carbonate sources, so the fact that oceans have the same strontium isotopic composition everywhere must relate to a
long residence time (∼4 million years) for this element
in seawater. Interoceanic circulation patterns are rapid
compared with strontium residence, so that a worldwide
isotopic composition results at any given time.
Continental crystalline rocks probably supply most
of the neodymium to the oceans. Differences in the ages
of continental rocks whose weathering products are
drained into the oceans produce distinct radiogenic
isotopic signatures in river waters. We might expect
that, as with strontium isotopes, these differences would
be erased by global circulation, but they are not. The
Atlantic is supplied by very old continental regions,
with high 143Nd/ 144Nd, whereas the Pacific is ringed by
younger regions with correspondingly lower values of
neodymium isotope ratios. Waters in the two oceans retain these isotopic signatures. This can be explained only
by assuming that the residence time for neodymium in
seawater is short; that is, that neodymium entering the
Isotopic Composition of the Oceans
The isotopic composition of ocean water depends on
inputs from various sources. These are primarily (1) submarine weathering of and magmatic interaction with
young basaltic rocks of the ocean floor, (2) recycling of
marine carbonate rocks that now reside on continents,
and (3) weathering of crystalline continental crust, which
consists mostly of granitic igneous and metamorphic
rocks. We briefly consider the behavior of strontium and
neodymium isotopes in the oceanic system.
At the present time, all oceans have a 87Sr/ 86Sr ratio
of 0.7090. In contrast, the measured 143Nd/ 144Nd ratios
of ocean water indicate that each ocean has a distinct and
characteristic compositional range, as summarized in figure 14.15 in terms of εNd. What causes this difference in
isotopic behavior?
It is thought that much of the strontium in seawater
is derived from weathering of marine carbonates on continents. The 87Sr/ 86Sr ratio in these carbonates has varied during Phanerozoic time, decreasing to a minimum
Antarctic
Ocean
Indian
Ocean Scotia
Sea
Number of Analyses
6
5
4
3
Pacific
Ocean
Island
Arcs
Atlantic
Ocean
MORB
2
1
-16
-14
-12 -10
-8
-6
-4
307
-2
0
+2
+4
+6
+8 +10 +12 +14
εNd
FIG. 14.15. The neodymium isotopic compositions of seawater and ferromagnesian nodules from
different oceans, expressed as εNd. Boxes represent individual analyses. These compositions are
distinct from each other and from oceanic volcanic rocks. (After Piepgras et al. 1979.)
308
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
oceans is lost to the seafloor before interoceanic circulation patterns can homogenize the isotopic composition
between oceans.
Radiogenic isotopic studies of seawater thus offer the
possibility of shedding light on the rates at which oceans
mix, provided that residence times can be fixed. Neodymium isotope measurements may even be able to
identify the location of currents responsible for such
mixing. Strontium isotopes also provide information on
the kinds of rocks exposed to weathering on continents
over geologic time.
by input from oceanic volcanics, marine carbonates, and crystalline continental crust. The 87Sr/ 86Sr ratio of seawater (sw) is
given by:
(87Sr/ 86Sr)sw = (87Sr/ 86Sr)vV + (87Sr/ 86Sr)mM
+ (87Sr/ 86Sr)cC,
where (87Sr/ 86Sr)v,m,c are strontium isotopic ratios contributed
by volcanic, marine carbonate, and continental crustal rocks,
respectively, and the coefficients V, M, and C are the fractions
of the total input contributed by these sources. By substituting
present-day estimates for the strontium isotopic compositions
of each of these sources, we obtain:
(87Sr/ 86Sr)sw = 0.704V + 0.708M + 0.720C.
Worked Problem 14.7
What are reasonable values for input into the modern oceans
from various geochemical reservoirs? Gunter Faure (1986) presented a mass balance model based on the premise that the
strontium isotopic composition of the oceans could be explained
(14.21)
If we then plot (87Sr/ 86Sr)sw versus V, as illustrated in figure
14.16, we can contour values of M and C from equation 14.21.
Plausible solutions are limited by the restriction that no source
can supply a negative amount of material; that is, V, M, and C
must be ≥0. The isotopic composition of modern seawater is illustrated in figure 14.16 by a horizontal dashed line at 87Sr/ 86Sr
= 0.7090. From this figure, we can see that V cannot be >0.68.
However, V cannot be very close to this upper bound, because
marine carbonates must make a substantial contribution. The
high strontium contents and susceptibility to chemical weathering demand an important role for the volcanic rocks in this
system. This model does not provide a unique solution to the
problem unless the input from any one of the sources can be
fixed unambiguously. However, it does allow us to assess the
relative contributions from these sources if a reasonable value
for one of them can be guessed. For example, if we assume that
M has the value 0.7, then the contributions from other sources
are V = 0.16 and C = 0.14.
Degassing of the Earth’s Interior
to Form the Atmosphere
FIG. 14.16. A model to explain the strontium isotopic composition of seawater. Mixtures of various proportions of volcanic rocks
(V), marine carbonates (M), and continental crustal rocks (C) can
account for the 87Sr/ 86Sr ratio of modern ocean water (dashed
horizontal line at 0.709). (After Faure 1986.)
The atmosphere is thought to have formed primarily
by degassing of the interior of the Earth, and isotopic
variations in gases dissolved in basalts provide information on this process. Noble gases are potentially very
useful in understanding atmospheric evolution, because
they are chemically inert and thus are not fractionated
by chemical reactions, and because all except helium
are too heavy to escape from the atmosphere to space.
Helium and neon cannot be subducted but are continuously released at midocean ridges, demonstrating that
the Earth’s interior has not been completely degassed.
Radioactive decay results in the production of many
noble gas isotopes. This is easily monitored by considering ratios of radiogenic and stable nuclides. For example,
variations observed at the present time in 3He/ 4He ratios
Using Radioactive Isotopes
reflect the balance between primordial (stable) 3He and
radiogenic 4He formed by decay of 232Th, 235U, and 238U.
(Why the radiogenic isotope is not the numerator in this
ratio, as is conventional for other radiogenic isotope
ratios, is a mystery.) Neon has one radiogenic isotope
(21Ne) and two nonradiogenic isotopes (20Ne and 22Ne),
as does argon. The 40Ar/ 36Ar ratio is controlled by the
progressive decay of 40 K to 40Ar. The isotopic composition of xenon reflects several processes. Radiogenic 129Xe
was produced by decay of now extinct 129I. Fission of
238U also generates 134Xe, 136Xe, and several other xenon
isotopes, creating variations in the ratios of these isotopes to nonradiogenic 130Xe.
Two different geological processes also affect these
isotopic ratios. Degassing occurs by volcanic and hydrothermal activity. Decreases in gas concentrations in the
mantle reservoir affect both radiogenic and stable nuclides, but not the nongaseous parent nuclides. For example, degassing decreases the amount of 36Ar (the
isotope to which radiogenic 40Ar is normalized) but
does not affect 40 K (which generates 40Ar). Fractionation of incompatible elements also affects these ratios.
We have already learned that K, U, and Th can be fractionated into the crust by magmatic processes. Because
some isotopes of these elements are parent nuclides for
noble gas isotopes, this process leads to depletion of
radiogenic gas isotopes in mantle source regions.
Degassing and crustal formation have occurred on different time scales, and the parent nuclides that produce
the isotopic variations in noble gases have very different
half-lives. Consequently, these isotopic systems may
permit us to trace these geologic processes and determine
their time dependence.
Claude Allegre and his coworkers (1986) modeled
degassing of the mantle using mass balance calculations
that employed argon and xenon isotopic ratios. Their
model employs four noble gas reservoirs: the atmosphere,
continental crust, upper mantle, and lower mantle. Their
calculations indicate that approximately half the Earth’s
mantle is 99% degassed. Based on other isotopic evidence, such as neodymium isotopes discussed earlier, this
is a larger mantle proportion than that corresponding
to the depleted upper mantle. The disagreement might
be rationalized by assuming that some of the primitive
lower mantle has also lost a portion of its noble gases.
Helium was not used in the mass balance calculations
because it has limited residence time in the atmosphere.
However, this becomes an advantage if we want to cal-
309
culate the gas flux from the mantle. Assuming a steady
state condition in which the mantle releases as much He
as is lost from the atmosphere to space, Allegre and coworkers calculated He fluxes in the mantle-atmosphere
system. Fluxes of Ar and Xe can also be estimated based
on correlations of their isotopic ratios with 3He/ 4He. The
model of Allegre and his collaborators is summarized
in figure 14.17. From this diagram, we can see that the
only recycling occurs at subduction zones, and that this
recycling is temporary because gases in the subducted
He Loss
Atmosphere
Arc
Volcanism
Plume Mid-Ocean
Volcanism Ridge
Volcanism
Continental
Crust
Upper
Mantle
Subduction
He Diffusion
Lower
Mantle
FIG. 14.17. Schematic representation of noble gas reservoirs and
fluxes in the mantle-crust-atmosphere system. Degassing of the
upper mantle occurs through volcanism at spreading centers, and
degassing of a portion of the lower mantle results from hot spot
volcanic activity. In this model, helium diffuses between mantle
reservoirs, but they behave as closed systems to other noble gases.
Subducted oceanic crust contains atmospheric gases, but these
are only temporarily recycled, because arc volcanism returns them
to the atmosphere. Continental crust is a significant reservoir for
radiogenic gases, because their parent isotopes are fractionated
into the crust and remain there for long periods of time. Helium is
the only noble gas that escapes from the atmosphere to space.
(Modified from Allegre et al. 1986.)
310
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
slabs are eventually returned to the atmosphere by arc
volcanism.
However, the ratio 20Ne/22Ne in the mantle differs
considerably from that in the atmosphere, and this difference cannot be attributed to nuclear processes (both
are nonradiogenic isotopes). This implies that mantle
degassing to form the atmosphere has not been a closed
system process. Either some dynamic process caused loss
of some early atmosphere (escape processes can result
in significant fractionations), or some of the atmosphere
has a different origin (perhaps carried in impacting
comets). Although the neon isotopic data appear to require a more complex evolution of the Earth’s atmosphere, simple degassing of the mantle as advocated by
Allegre and coworkers is probably sufficient to explain
the argon isotopes, because the mantle is such a huge
reservoir of potassium. Noble gas isotopes in the Earth’s
mantle were reviewed by Ken Farley and Elizabeth
Neroda (1998).
SUMMARY
In this chapter, we have seen that isotope stability is the
exception rather than the rule. Radionuclides may decay
by emission of beta particles, alpha particles, positrons,
capture of electrons, or spontaneous fission. Branched
decay involving several of these decay mechanisms occurs
in some cases. The rate of radioactive decay is exponential and is most easily expressed in terms of half-life. We
have derived equations that allow determination of the
isotopic evolution of a system as a function of time.
These equations permit the age dating of rocks, based on
the decay of certain naturally occurring radionuclides,
such as 40K, 147Sm, 87Rb, 232Th, 235U, and 238U. In such
systems, it is possible to determine the age of rocks even
though the daughter isotope may have been present in
the system initially. Uranium and thorium isotopes decay
through a series of intermediate radioactive daughters,
but the half-lives of transitory daughters are short enough
that each system can usually be treated as the decay
of parent directly to the stable daughter. Independent
U-Th-Pb geochronometers can give the same or different
results, but useful information can be gained even in cases
of discordancy. The interpretation of geochronological
data can be complicated by the failure of rock systems to
remain closed, as well as other factors. The earlier presence of now extinct radionuclides can, in some cases, be
determined, and their existence places constraints on
early processes such as planetary differentiation.
The abundances of radiogenic isotopes can be determined directly from mass spectrometry or indirectly from
fission track techniques. The production of radioactive
isotopes can also be induced artificially in a nuclear reactor, which forms the basis for elemental quantitative
analysis by neutron activation. The 40Ar-39Ar geochronometer also depends on induced radioactivity. Cosmogenic nuclides are produced naturally by cosmic ray
exposure, which forms 14C for dating recent events and
other radionuclides useful in quantifying surface processes such as erosion.
Radionuclides also provide powerful tracers for geochemical processes. Uses we have explored here include
understanding mantle heterogeneity, cycling of materials between crust and mantle reservoirs, global oceanic
mixing, and degassing of the Earth’s interior to form the
atmosphere. The importance of radiogenic isotopes is
considered further in the following chapter on the early
solar system and processes at planetary scales.
suggested readings
There are an increasing number of textbooks that treat isotope
geochemistry. The books by Faure and Dickin are required
reading for anyone seriously interested in this subject, and the
other references provide further elaboration on certain aspects
of radiogenic isotopes.
Bowen, R. 1988. Isotopes in the Earth Sciences. London: Elsevier Applied Science. (Chapter 2 provides a thorough discussion of mass spectrometry, and later chapters deal primarily
with radiometric dating techniques.)
Dickin, A. P. 1995. Radiogenic Isotope Geology. Cambridge:
Cambridge University Press. (An up-to-date treatment of this
field; everything you want to know about geochronology.)
Durrance, E. M. 1986. Radioactivity in Geology: Principles and
Applications. New York: Wiley. (This unconventional book
deals with many applications of radioactivity not treated
here; examples include its use in prospecting for radioactive
ores and petroleum, environmental applications, and the heat
generated by radioactive decay.)
Farley, K. A., and E. Neroda. 1998. Noble gases in the Earth’s
mantle. Annual Reviews of Earth and Planetary Sciences 26:
189–218. (A lucid review of a particularly complex subject;
noble gas geochemistry is complex and often understood
only by its practitioners, but this paper really helps.)
Faure, G. 1986. Principles of Isotope Geology, 2nd ed. New
York: Wiley. (One of the best available texts on isotope
Using Radioactive Isotopes
geochemistry. This is an excellent source for detailed information of radiogenic isotopes; geochronology is covered
more exhaustively than other aspects of this subject.)
Jager, E., and J. C. Hunziker, eds. 1979. Lectures in Isotope
Geology. Berlin: Springer-Verlag. (A series of lectures given
in Switzerland by respected isotope geochemists; the first
half of the book deals with geochronology.)
McLennan, S. M., S. Hemming, D. K. McDaniel, and G. N.
Hanson. 1993. Geochemical approaches to sedimentation,
provenance, and tectonics. Geological Society of America
Special Paper 284:21–40. (This review paper is a primer on
the use of radiogenic isotopes in deciphering the tectonic
setting of clastic sediments.)
Philpotts, A. R. 1990. Principles of Igneous and Metamorphic
Petrology. Englewood Cliffs: Prentice-Hall. (Chapter 21 gives
a particularly good overview of isotope geochemistry, and
especially of the use of radiogenic isotopes in understanding
magma sources.)
Vertes, A., S. Nagy, and K. Suvegh. 1998. Nuclear Methods in
Mineralogy and Geology. New York: Plenum. (This is an
excellence source for learning about neutron activation
analysis, and other subjects not covered here, such as dating
groundwater.)
Walker, F. W., D. G. Miller, and F. Feiner. 1983. Chart of the
Nuclides. San Jose: General Electric Company. (A real bargain that contains physical constants, conversion factors,
and periodic table; order from General Electric Nuclear
Energy Operations, 175 Curtner Ave., M/C 684, San Jose,
California 95125.)
The following references were cited in this chapter. These provide more thorough treatments of the applications of radionuclides in solving geochemical problems.
Aleinikoff, J. N., R. H. Moench, and J. B. Lyons. 1985. Carboniferous U-Pb age of the Sebago batholith, southwestern
Maine: Metamorphic and tectonic implications. Geological
Society of America Bulletin 96:990–996.
Allegre, C. J., T. Staudacher, and P. Sarda. 1986. Rare gas
systematics: Formation of the atmosphere, evolution and
structure of the Earth’s mantle. Earth and Planetary Science
Letters 81:127–150.
Anderson, D. L. 1994. Komatiites and picrites: Evidence that
the “plume” source is depleted. Earth and Planetary Science
Letters 128:303–311.
311
DePaolo, D. J. 1981. Trace element and isotopic effects of combined wallrock assimilation and fractional crystallization.
Earth and Planetary Science Letters 53:189–202.
Gauthier-Lafaye, F., F. Weber, and H. Ohomoto. 1989. Natural fission reactors of Oklo. Economic Geology 84:2286–
2295.
Halliday, A. N., and D.-C. Lee. 1999. Tungsten isotopes and
the early development of the Earth and Moon. Geochimica
et Cosmochimica Acta 63:4157–4179.
Hart, S. R., D. C. Gerlach, and W. M. White. 1986. A possible Sr-Nd-Pb mantle array and consequences for mantle
mixing. Geochimica et Cosmochimica Acta 50:1551–
1557.
Morris, J. D. 1991. Applications of cosmogenic 10Be to problems in the earth sciences. Annual Reviews of Earth and
Planetary Sciences 19:313–350.
Nielson, D. L., S. H. Evans, and B. S. Sibbett. 1986. Magmatic,
structural, and hydrothermal evolution of the Mineral Mountains intrusive complex, Utah. Geological Society of America
Bulletin 97:765–777.
Nyquist, L. E., C. –Y. Shih, J. L. Wooden, B. M. Bansal, and
H. Wiesmann. 1979. The Sr and Nd isotopic record of
Apollo 12 basalts: Implications for lunar geochemical evolution. Proceedings of the Lunar and Planetary Science
Conference 10:77–114.
Patchett, P. J., W. M. White, H. Feldman, S. Kielinszuk, and
A. W. Hoffman. 1984. Hafnium rare-earth element fractionation in the sedimentary system and crustal recycling
into the earth’s mantle. Earth and Planetary Science Letters
69:365–378.
Piepgras, D. J., G. J. Wasserburg, and E. J. Dasch. 1979. The
isotopic composition of Nd in different ocean masses. Earth
and Planetary Science Letters 45:223–236.
Price, P. B., and R. M. Walker. 1963. Fossil tracks of charged
particles in mica and the age of minerals. Journal of Geophysical Research 68:4847–4862.
Sisson, V. B., and T. C. Onstott. 1986. Dating blueschist metamorphism: A combined 40Ar/ 39Ar and electron microprobe
approach. Geochimica et Cosmochimica Acta 50:2111–
2117.
Yin, Q., S. B. Jacobsen, K. Yamashita, J. Blichert-Toft, P. Talouk,
and F. Albarede. 2002. A short timescale for terrestrial
planet formation from Hf-W chronometry of meteorites.
Nature 418:949–952.
312
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
PROBLEMS
(14.1) Given the following Rb-Sr isotopic data for whole-rock samples of a granitic pluton, determine the
initial strontium isotopic ratio and age of the pluton.
Sample No.
1
2
3
4
5
6
7
87Rb/86Sr
87Sr/86Sr
11.86
7.66
6.95
9.68
6.54
9.69
3.74
0.7718
0.7481
0.7436
0.7587
0.7413
0.7599
0.7259
(14.2) (a) Solve equation 14.8 for t. (b) Use the equation you just derived to determine the age of a biotite
sample, for which the following data have been obtained: K = 7.10 wt. %, 40Ar = 1.5 × 1012 atoms g−1.
(14.3) A zircon grain initially contained 1000 ppm 235U and 100 ppm 207Pb. How many atoms of 207Pb will
this zircon contain after 1 billion years?
(14.4) Using the data below and figure 14.7, construct a concordia diagram. Are these samples concordant?
Interpret your results, explaining any assumptions that you make.
Sample No.
1
2
3
206Pb/238U
207Pb/ 235U
0.500
0.543
0.565
15.93
18.77
20.00
(14.5) You obtain the following geochronological data for a portion of the Appalachians. A gneiss unit
gives whole-rock Rb-Sr and zircon U-Pb ages of 315 million years. Mineral geothermometers in this
unit give a peak metamorphic temperature of 530o C. 40Ar-39Ar gas retention ages are 295 million
years for hornblende and 283 million years for biotite. Fission track ages for apatites range from 130
to 150 million years. Using these data, construct a time-temperature path illustrating the late Paleozoic thermal history of this part of the Appalachian orogen.
(14.6) The measured activity of a sample of charcoal found at an archaeological site is 5.30 dpm g−1. What
is the age of the sample?
(14.7) How much heat energy would be generated in 100,000 years by a one cubic meter sample of rock
containing live 26Al? The half-life of 26Al is given in table 14.1. Assume that the sample initially contains 100 g of aluminum with a 26Al/ 27Al atomic ratio of 0.1. Also assume that the energy of decay
for 26Al is 10−14 cal atom−1.
(14.8) A magma containing 200 ppm strontium is contaminated by crustal rocks containing 1000 ppm Sr.
The magma simultaneously undergoes fractional crystallization of phases with a bulk distribution coefficient DSr= 0.75. The initial Sr isotopic ratios of the magma and crustal rocks are 0.704 and 0.713,
respectively. Assume that r = 0.5. Use equation 14.20 to calculate the isotopic composition of the resulting hybrid magma as a function of mass of magma relative to original mass (F), and illustrate the
result with a graph.
CHAPTER 15
STRETCHING OUR HORIZONS
Cosmochemistry
OVERVIEW
In this chapter, we explore the rapidly emerging field of
cosmochemistry. This subject involves the geochemical
aspects of systems of planetary or solar system scale. We
first consider nucleosynthesis processes in stars and use
this foundation to understand the abundances of elements
in the Sun and the solar system. Chondritic meteorites
are discussed next, as samples of average solar system
material stripped only of the lightest elements. Analyses
of these meteorites provide information on the behavior
of various elements in the early solar nebula; from these,
we learn that elements were fractionated according to
both geochemical affinity and volatility. Chondritic fractionation patterns are also important for understanding
planetary compositions. We then consider evidence for
proposed cosmochemical processes in the early solar nebula, such as condensation of solids from vapor and infusion of materials formed in and around other stars. The
characteristics and origin of extraterrestrial organic molecules and ices are also explored; these, too, were important planetary building blocks. Finally, we consider
how geochemical models for planets are formulated.
WHY STUDY COSMOCHEMISTRY?
In earlier chapters, we have sometimes used observations
from extraterrestrial materials or other planetary bodies
to illustrate processes or pathways in geochemistry. Perhaps these seemed exotic or even esoteric. From the geologist’s uniformitarian perspective, however, the planets
and other extraterrestrial bodies provide a larger laboratory in which to study geochemical behavior. Historically,
the study of cosmochemistry has paralleled (and, in some
cases, spurred) the development of geochemistry as a
discipline, and has involved many of the same scientists.
It is significant that the leading journal in geochemistry
is also the premier journal in cosmochemistry and incorporates both fields in its title (Geochimica et Cosmochimica Acta). In this chapter, we examine several
problems of huge (even astronomic) scale that attract
geochemists to study the solar system and beyond, and
we see how some of the answers improve our understanding of the world under our feet.
Our solar system consists of a single star surrounded
by nine planets and a large retinue of smaller bodies,
including their satellites, asteroids, comets, and a lot of
small rocks and dust. Samples of the small rocks and
dust are provided free of charge as meteorites and interplanetary dust particles, and in some cases, these can be
related to the parent planets or asteroids on which they
formed. Cosmochemists focus on chemical differences
and similarities among these bodies. To what extent have
they evolved separately, and what characteristics have
they inherited from a common origin? Which processes
are unique in a given body, and which processes are likely
313
314
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
to have affected all of them? Do they contain evidence for
how the solar system formed, and how has it evolved to
its present state?
Most (>99%) of the mass of our solar system is concentrated in the Sun, so many questions regarding the
bulk chemistry of the solar system and its early history
must first address the behavior and composition of stars.
Here, we rely on astrophysicists to construct models
that infer the structure and evolutionary development of
stars in general. These are based on observations of the
Sun and distant stars and on the principles of thermodynamics and nuclear chemistry we have discussed in
earlier chapters, although the language of astrophysics
is different from that of geochemistry.
From the Sun, we turn to the planets, which are usually divided into two groups. Those closest to the Sun
(Mercury, Venus, Earth, and Mars) are called the terrestrial planets, to distinguish them from the larger Jovian
planets beyond (Jupiter, Saturn, Uranus, and Neptune).
The Jovian worlds are part rock and part ice, now transformed into unusual forms by high pressures. Pluto and
its satellite Charon, small icy bodies at the outer reaches
of the solar system, have unusual orbits and may be related to comets. In this chapter, we focus on the terrestrial
planets, as they are more likely to provide geochemical
insights into the Earth. The presence and compositions
of the Jovian planets, however, place important restrictions on any models for the history of the solar system
as a whole.
The terrestrial bodies include not only the four inner
planets, but also the moons of Earth and Mars and several thousand asteroids. Studies of these, particularly
of our own Moon, have helped geochemists to determine how planetary size affects differentiation and other
global-scale processes. Lunar samples brought to Earth
by Apollo astronauts, by Soviet unmanned spacecraft,
and as meteorites are an important resource for cosmochemists. A small group of meteorites are also thought
to have come from Mars. Some of the most valuable
information about terrestrial bodies, however, is based
on meteorites that are fragments of asteroids. Some asteroids experienced differentiation and core formation,
producing igneous rocks (achondrites) and iron meteorites, whereas others (chondrites) remained almost
unchanged after their parent asteroids accreted. From
meteorites, we obtain data that can be used to model
the development of planetary cores, mantles, and crusts.
We also use the nearly pristine nature of chondrites to
deduce the chemistry of materials from which the larger
terrestrial planets were assembled.
We begin a survey of cosmochemistry by considering
the Sun as a star, to find chemical clues to its origin and
the origin of other solar system materials. As we proceed, our focus will be drawn more to the development
of planets, and finally back to the Earth itself. In this tour,
you will recognize many familiar geochemical concerns
and, we hope, see the place of cosmochemistry in understanding how the Earth works.
ORIGIN AND ABUNDANCE
OF THE ELEMENTS
Nucleosynthesis in Stars
The universe is believed to have begun in a cataclysmic
explosion—the Big Bang—and has been expanding ever
since. At this beginning stage, matter existed presumably
in the form of a stew of neutrons or as simple atoms
(hydrogen and deuterium). The Big Bang itself may have
produced some other nuclides, but only 4He was formed
in any abundance. In chemical terms, this was a pretty
dull universe. How then did all of the other elements
originate?
Local concentrations of matter periodically coalesce
to form stars. In 1957, Margaret Burbidge and her husband, Geoffrey Burbidge, along with William Fowler and
Fred Hoyle, teamed together to write a remarkable scientific paper that argued that other elements formed in
stellar interiors by nuclear reactions with hydrogen as
the sole starting material. (This paper is now rightfully
considered a classic, often referred to as B2FH from the
initials of the authors’ surnames.) When hydrogen atoms
are heated to sufficiently high temperatures and held together by enormous pressures—such as occur in the deep
interior of the Sun—fusion reactions occur. The protonproton chain, the dominant energy-producing reaction in
the Sun, is illustrated in figure 15.1. This fusion process,
commonly called hydrogen burning, produces helium.
When the hydrogen fuel in the interior begins to run low
and the stellar core becomes dense enough to sustain reactions at higher pressures, another nuclear reaction may
take place, in which helium atoms are fused to make carbon and oxygen. While helium burning proceeds in the
core of a star, a hydrogen-burning shell works its way
toward the surface; a star at this evolutionary stage expands into a red giant. This will be the fate of our Sun.
Stretching Our Horizons: Cosmochemistry
H
Pos
D
H
3He
H
H
H
3He
4He
FIG. 15.1. The proton-proton chain, in which hydrogen atoms are
fused into helium, is the predominant energy-producing reaction
in the Sun. Open circles are protons, filled circles are neutrons,
and Pos is a positron. For each gram of 4He produced, 175,000 kwh
of energy are released.
A star more massive than the Sun, however, can employ other fusion reactions, successively burning carbon,
neon, oxygen, and silicon. The ashes from one burning
stage provide the fuel for the next. The internal structure
of such a massive, highly evolved star would then consist of many concentric shells, each of which produces
fusion products that are then burned in the adjacent
inward shell. The ultimate products of such fusion reactions are elements near iron in the periodic table (V, Cr,
Mn, Fe, Co, and Ni). Fusion between nuclei cannot produce nuclides heavier than the iron group elements. For
elements lighter than iron, the energy yield is higher for
fusion reactions than for fission, but for elements heavier
than iron, the energy yield for fission is greater. When a
star reaches this evolutionary dead end, a series of both
disintegrative and constructive nuclear reactions occurs
among the iron group nuclei. After some time, a steady
state, called nuclear statistical equilibrium, is reached; the
relative abundances of iron group elements reflect this
process.
Elements heavier than the iron group can be formed
by addition of neutrons to iron group seed nuclei. We
315
have already seen in chapter 14 that nuclei with neutron/
proton ratios greater than the band of stability undergo
beta decay to form more stable nuclei. Helium burning
produces neutrons that are captured by iron group nuclei in a slow and rather orderly way, called the slow or
s process. This process is slow enough that nuclei, if they
are unstable, experience beta decay into stable nuclei before additional neutrons are added. This is illustrated in
figure 15.2, in which stable and unstable nuclei are shown
as white and gray boxes, respectively. A portion of the s
process track is illustrated by the upper set of arrows.
Neutrons are added (increasing N) until an unstable
nuclide is produced; then beta decay occurs, resulting in
a decrease of both N and Z by one. The resulting nuclide
can then capture more neutrons until it beta decays again,
and so forth. Notice that a helium burning star producing heavy elements by this process would have to be at
least a second generation star, having somehow inherited
iron group nuclei from an earlier star whose material
had been recycled.
Fusion, nuclear statistical equilibrium, and s process
neutron-capture reactions in massive stars thus provide
mechanisms by which elements heavier than hydrogen
can be produced. But how are these nuclides placed into
interstellar space for later use in making planets, oceans,
and people? Processed stellar matter is lost continuously from stars as fluxes of energetic ions, such as the
solar wind. Other elements are liberated by supernovae—stellar explosions that scatter matter over vast
distances. At the same time, supernovae generate other
nuclides by a different nucleosynthetic pathway, the rapid
or r process. Very rapid addition of neutrons to seed nuclei results in a chain of unstable nuclides.
One example of the r process is illustrated by the set
of arrows at the bottom of figure 15.2. Reaching this part
of the diagram, which is populated by unstable nuclei,
happens only when neutrons are added more rapidly than
the resulting nuclei can decay. The nuclides in crosshatched boxes decay much more rapidly than the seconds
or minutes required for the decay of nuclides occupying
gray boxes, so a nucleus shifted into a crosshatched box
by r-process neutron capture immediately transforms by
beta decay, as shown. When the supernova event ends,
the r process nuclides transform more slowly by successive beta decays into stable nuclides. Many of these stable
isotopes are the same as produced by the s process but
some, such as 86Kr, 87Rb, and 96Zr, can be reached only
by the r process. In an analogous manner, such nuclides
316
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Number of Protons (Z)
94Mo 95Mo 96Mo 97Mo 98Mo
92Mo
42
41
93Nb
40
90Zr 91Zr 92Zr
39
89Y
s
85Rb
87Rb
36
84Kr
86Kr
94Zr
oc
Pr
Unstable
ess
35
34
96Zr
ess
38 86Sr 87Sr 88Sr
37
Stable
r
82Se
oc
Pr
90Se 91Se
33
85As 86As 87As 88As 89As
32
84Ge
31
80Ga 81Ga 82Ga 83Ga
48
49
50
51
52
Highly
Unstable
54
55
56
57
Number of Neutrons (N)
FIG. 15.2. A portion of the nuclide chart, illustrating how the s process (upper set of arrows) and
the r process (lower set of arrows) produce new elements by neutron capture and subsequent decay.
Shaded boxes represent stable nuclides, whereas white boxes are unstable and undergo beta decay,
which shifts them up and to the left. Crosshatched boxes represent extremely unstable nuclides
that transform rapidly by neutron capture.
as 92Mo (fig. 15.2) on the proton-rich side of the s process
band can be produced during supernovae by addition
of protons to seed nuclei (the p process), followed by
positron emission or electron capture.
Cosmic Abundance Patterns
The relative abundances of the elements in a star,
then, are controlled by a combination of nucleosynthetic
processes. For the moment, let’s ignore how stellar or
solar abundances are determined (we return to this
problem shortly). Elemental abundances in the Sun, relative to 106 atoms of silicon, are given in table 15.1 and
illustrated graphically in figure 15.3. (Normalization to
silicon keeps these numbers from being astronomically
large.) This abundance pattern is commonly called the
cosmic abundance of the elements. This is a misnomer,
of course; the composition of our Sun is not necessarily
representative of that of the universe, but it does define
the composition of our own solar system. Other stars
have their own peculiar compositions because of different contributions from fusion, s and r processes, and
so forth.
Let’s inspect the cosmic abundance pattern in figure 15.3 to see why the Sun has this particular composition. Clearly, hydrogen and helium are the dominant
elements. This is because only these two emerged from the
Big Bang. Elements between helium and the iron group
formed predominantly by fusion reactions. These show
a rapid exponential decrease with increasing atomic number, reflecting decreasing production in the more advanced burning cycles. The Sun is presently burning only
hydrogen, so these elements must have been inherited
from an earlier generation of stars. Exceptions are the
elements lithium, beryllium, and boron, which have
abnormally low abundances. The abundances of these
Stretching Our Horizons: Cosmochemistry
317
TABLE 15.1. Cosmic Abundances of the Elements, Based on 106 Silicon Atoms
Atomic
Number
1
2
3
4
5
6
7
8
9
10
11
12
13
14
15
16
17
18
19
20
21
22
23
24
25
26
27
28
29
30
31
32
33
34
35
36
37
38
39
40
41
42
43
44
Element
Hydrogen
Helium
Lithium
Beryllium
Boron
Carbon
Nitrogen
Oxygen
Fluorine
Neon
Sodium
Magnesium
Aluminum
Silicon
Phosphorus
Sulfur
Chlorine
Argon
Potassium
Calcium
Scandium
Titanium
Vanadium
Chromium
Manganese
Iron
Cobalt
Nickel
Copper
Zinc
Gallium
Germanium
Arsenic
Selenium
Bromine
Krypton
Rubidium
Strontium
Yttrium
Zirconium
Niobium
Molybdenum
Technetium1
Ruthenium
Symbol
H
He
Li
Be
B
C
N
O
F
Ne
Na
Mg
Al
Si
P
S
Cl
Ar
K
Ca
Sc
Ti
V
Cr
Mn
Fe
Co
Ni
Cu
Zn
Ga
Ge
As
Se
Br
Kr
Rb
Sr
Y
Zr
Nb
Mo
Tc
Ru
Atomic
Number
Abundance
10
2.79 × 10
2.72 × 109
57.1
0.73
21.2
1.01 × 107
3.13 × 106
2.38 × 107
843
3.44 × 106
5.74 × 104
1.074 × 106
8.49 × 104
1.00 × 106
1.04 × 104
5.15 × 105
5240
1.01 × 105
3770
6.11 × 104
34.2
2400
293
1.35 × 104
9550
9.00 × 105
2250
4.93 × 104
522
1260
37.8
119
6.56
62.1
11.8
45
7.09
23.5
4.64
11.4
0.698
2.55
—
1.86
45
46
47
48
49
50
51
52
53
54
55
56
57
58
59
60
61
62
63
64
65
66
67
68
69
70
71
72
73
74
75
76
77
78
79
80
81
82
83
90
91
92
84–89
Element
Symbol
Rhodium
Palladium
Silver
Cadmium
Indium
Tin
Antimony
Tellurium
Iodine
Xenon
Cesium
Barium
Lanthanum
Cerium
Praseodymium
Neodymium
Promethium1
Samarium
Europium
Gadolinium
Terbium
Dysprosium
Holmium
Erbium
Thulium
Ytterbium
Lutetium
Hafnium
Tantalum
Tungsten
Rhenium
Osmium
Iridium
Platinum
Gold
Mercury
Thallium
Lead
Bismuth
Thorium
Protactinium1
Uranium
Unstable
elements1
Rh
Pd
Ag
Cd
In
Sn
Sb
Te
I
Xe
Cs
Ba
La
Ce
Pr
Nd
Pm
Sm
Eu
Gd
Tb
Dy
Ho
Er
Tm
Yb
Lu
Hf
Ta
W
Re
Os
Ir
Pt
Au
Hg
Tl
Pb
Bi
Th
Pa
U
Abundance
0.344
1.39
0.486
1.61
0.184
3.82
0.309
4.81
0.90
4.7
0.372
4.49
0.4460
1.136
0.1669
0.8279
—
0.2582
0.0973
0.3300
0.0603
0.3942
0.0889
0.2508
0.0378
0.2479
0.0367
0.154
0.0207
0.133
0.0517
0.675
0.661
1.34
0.187
0.34
0.184
3.15
0.144
0.0335
—
0.0090
—
Data from Anders and Grevesse (1989).
1
Unstable element.
three elements may have been reduced in the Sun by
bombardment with neutrons and protons, although the
production processes just discussed tend to bypass these
elements. The peak corresponding to the iron group elements represents nuclides formed by nuclear statistical
equilibrium, also an addition from earlier stars. The abundances in the Sun of elements heavier than iron reflect
additions of materials formed in supernovae.
Superimposed on the peaks and valleys in figure 15.3
is a peculiar sawtooth pattern, caused by higher abundances of elements with even rather than odd atomic
numbers (Z). The even-numbered elements form a lopsided 98% of cosmic abundances, because nuclides with
even atomic numbers are more likely to be stable, as
already noted in chapter 14. This explanation, however, begs the question: Why are even-numbered atomic
318
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 15.3. Cosmic abundance of the elements, relative to 106 silicon atoms.
numbers more stable? Whenever nuclear particles, either
protons or neutrons, can pair with spins in opposite
directions, they can come closer together and the forces
holding them together will be stronger. This stronger
nuclear binding force results in higher stability.
The chemical composition of the Sun, then, can be understood primarily in terms of nucleosysthesis reactions
from an earlier generation of stars whose materials have
been recycled. These earlier generation stars also provided heavy elements for the formation of planets like the
Earth. Some helium has been produced at the expense of
hydrogen during the Sun’s lifetime. Every second, 700 million tons of hydrogen is fused into 695 million tons of
helium, with the missing 5 million tons converted into
energy according to Einstein’s famous equation. This is,
of course, the source of the Sun’s luminosity.
CHONDRITES AS SOURCES
OF COSMOCHEMICAL DATA
Chondrites are the most common type of meteorite.
These objects are basically ultramafic rocks, and petrographic studies suggest that they have never been geologically processed by melting and differentiation (although
some components of chondrites have been melted prior
to incorporation into the meteorites). Radiometric age
determinations indicate that chondrites are the oldest
samples surviving from the early solar system, approximately 4.56 billion years old.
Chondrites occupy a singularly important niche in
cosmochemistry and geochemistry: they are the baseline
from which the effects of most chemical processes can be
gauged. The reason for this is quite simple. Of all known
materials that can be studied directly, chondrites most
closely match the composition of the Sun. This is illustrated in figure 15.5, a plot of the abundances of elements
in the solar atmosphere against those in a chondrite, all
relative to 106 atoms of silicon. A perfect correspondence
is given by the diagonal line in this figure. The correspondence may even be better than indicated, because of
some uncertainties in measuring the Sun’s composition.
To be honest, we should note that the scales in this figure are logarithmic but, with the exception of a few elements, the abundances are certainly within a factor of
two of each other. Hydrogen, helium, and some other
elements that occur commonly as gases—carbon, nitrogen, oxygen—are more abundant in the Sun, so it may
be reasonable to visualize chondrites as a solar sludge
from which gases have been distilled. In contrast, lithium,
boron, and possibly beryllium have higher abundances
in chondrites than in the Sun. Recall that these elements
are systematically destroyed in the Sun by interactions
with nuclear particles. In this regard, then, chondrites
may record the original composition of the solar system
even better than does the present day Sun.
It should now be obvious why in previous chapters
we have repeatedly used chondrites as a basis for normalization of geochemical data: their nearly cosmic compo-
Stretching Our Horizons: Cosmochemistry
319
MEASURING THE COMPOSITION OF A STAR
The spectrum of light emitted by a star’s photosphere
(the region of the stellar atmosphere from which radiation escapes) provides a means for determining its
composition. Superimposed on the continuous spectrum of a star are absorption bands, or missing wavelengths (fig. 15.4). These bands, produced by electron
transitions in various atoms, appear as dark lines
when photographed through a spectrograph. These
are sometimes called Fraunhofer lines, after the German physicist who discovered them in 1817.
Each Fraunhofer line corresponds to a particular
element, and that element’s abundance can be determined from the intensity of the absorption. What is
actually measured is the width of the absorption band,
because the Fraunhofer lines are broadened with increasing concentration. In practice, allowance must
be made for the prevailing conditions of temperature
and pressure in the star’s photosphere before converting line width to element abundance. This is because
the width of the absorption band depends not only
on the elemental abundance, but also on the fraction
of atoms that are in the right state of ionization and
excitation to produce the band, which is in turn a
function of temperature and pressure. From observations of the continuous spectrum, temperature
and pressure can be modeled theoretically for various
depths within the photosphere. It is then possible to
combine the contributions from atoms at all depths
throughout the photosphere to predict how line width
varies with abundance.
Absorption bands corresponding to any one element are not necessarily visible in all stars. This is
probably more a function of the varying temperatures
sitions provide the justification for this common practice.
As an example, let’s consider chondrite normalization of
rare earth element (REE) patterns, which we introduced
in chapter 13. REEs in both chondrites and basalts exhibit the sawtooth pattern of even-odd atomic numbers,
as shown in figure 15.6. However, division of the abundance of each REE in basalt by the corresponding value
in chondrites gives the smoothed, normalized values
indicated by the open boxes. Without this normalizing
FIG. 15.4. The solar spectrum contains dark absorption bands
(Fraunhofer lines) that can be used to estimate the Sun’s composition. This figure shows only a small part of the spectrum.
and pressures of stars, which control the ability of a
particular element to absorb radiation, than to the
absence of that element. Measurements at certain
wavelengths are also blocked by the Earth’s atmosphere, although this problem has now been circumvented by spacecraft observations. Finally, note also
that the composition of the photosphere may be quite
different from that of the bulk of the star, unless convective mixing takes place.
To illustrate the importance of stellar spectral measurements, it is worth remembering that the second
most abundant element was actually discovered in
the nearest star before it was recognized on Earth. In
1869, Norman Lockyer, who founded the scientific
journal Nature, noticed a line he could not identify in
the solar spectrum. Based on this spectral evidence,
Lockyer announced the existence of a new element
which he called helium, because it had been discovered
in the Sun—helios in Greek.
step, the geochemical interpretation of REE patterns
would be very cumbersome.
The ancient ages and nearly cosmic compositions of
chondrites suggest that they are nearly pristine samples
of early solar system matter (this is actually true only
to a degree; see the accompanying box). There is ample
evidence that chondrites are samples of asteroids—bodies
generally too small to have sustained geologic processes.
(However, we noted earlier that achondrites and iron
320
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 15.5. Comparison of element abundances in the Allende chondritic meteorite with those in
the solar photosphere. All elements are relative to 106 silicon atoms. This excellent correspondence
indicates that chondrites have approximately cosmic compositions, except for the gaseous elements
H, He, C, N, O, and noble gases.
meteorites are samples of melted and differentiated asteroids.) The terrestrial planets formed by accretion of
smaller building blocks having chondritic composition.
It is reasonable, therefore, to infer that the Earth itself
must have a chondritic bulk composition.
COSMOCHEMICAL BEHAVIOR
OF ELEMENTS
Controls on Cosmochemical Behavior
The cosmic abundance of an element is controlled by
its nuclear structure, but its geochemical character is a
function of its electron configuration. In chapter 2, we
learned that Victor Goldschmidt first coined the terms
lithophile, chalcophile, and siderophile to identify elements with affinity for silicate, sulfide, and metal, respectively. Geochemical affinity is actually a qualitative
assessment of the relative magnitudes of free energies of
formation for various compounds of the element under
consideration. For example, suppose that metallic calcium
were exposed to an atmosphere containing oxygen and
sulfur. Two potential reactions for calcium are:
Ca + 1–2 O2 → CaO,
and
Ca + S2 → CaS2.
Of the two, the first is dominant, because ∆Gf0 for the
oxide is more negative than ∆Gf0 of the sulfide. Therefore, calcium is lithophile. Of course, we must specify
the conditions on the system for this conclusion to have
meaning. For example, under extremely reducing conditions such as those under which enstatite chondrites were
formed, calcium partitions more strongly into sulfide and
thus is chalcophile.
There was little direct evidence on which to speculate about geochemical behavior when this idea was put
foward, but Goldschmidt recognized that chondrites
represent a natural experiment from which element behavior could be deduced. He and his coworkers measured the abundances of elements in coexisting silicate,
Stretching Our Horizons: Cosmochemistry
321
FIG. 15.6. Measured abundances of REEs in a basalt and average chondrites. Normalizing the
basalt data to chondrites removes the zigzag pattern of odd-even abundances, revealing the pattern imposed during the igneous history of this rock.
sulfide, and metal from chondrites, and this pioneering
work has been supplemented by studies of metal, sulfide,
and silicate slag from smelters. Geochemical affinity is
important in understanding element partitioning in any
systems containing more than one of these phases—for
example, the formation of magmatic sulfide ores or the
differentiation of core and mantle.
This is not, however, the only factor controlling
element behavior in cosmochemical systems. Another
important factor is volatility, by which we mean the temperature range in which an element will condense from
a gas of solar composition. Volatility depends on several
factors, such as an element’s vapor pressure, cosmic abundance (which controls partial pressure), total pressure
of the system, and element reactivity (which determines
whether the element in question occurs in pure form or
in a compound). Refractory elements are those that
condense at temperatures >1400 K; volatile elements
generally condense at temperatures <1200 K.
Chemical Fractionations Observed
in Chondrites
In chondrites, groups of elements appear to move in
unison. If one element of the group is enriched or depleted, all of the others are as well, and by nearly the
same factor. At least four types of chemical fractionations,
relative to cosmic (and C1 chondrite) composition, have
been observed in chondrites:
1. Fractionation of refractory elements (for example, Ca,
Al, Ti), resulting in high refractory element abundances
in carbonaceous chondrites and lower abundances in
other chondrite classes;
322
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
CHONDRITE PETROLOGY AND CLASSIFICATION
Chondrites take their name from chondrules: round,
millimeter-sized, quenched droplets of silicate melt
that are particularly abundant in these meteorites.
Chondrules formed by flash melting at temperatures
of 1700–2100 K and subsequent rapid cooling. They
are typically composed of olivine, pyroxenes, and glass
or feldspar. Irregularly shaped white objects called
Ca,Al-rich inclusions (CAI) also occur in most chondrites. These are composed mostly of melilite, spinel,
clinopyroxene, and anorthite. CAIs also formed at
high temperatures of 1700–2400 K, and many were
at least partly molten. Larger grains of iron-nickel
metal and sulfides are also scattered throughout these
meteorites. All of these objects are cemented together
by a fine-grained matrix consisting of mixtures of
olivine, pyroxene, feldspathoids, graphite, magnetite,
and other minor phases. These components are illustrated in figure 15.7. These diverse kinds of materials
have accreted together in space to form a kind of cosmic sediment. Despite extensive research, the origins
of most of these components are unclear.
There are actually many classes of chondrites,
differing slightly in petrography and chemical composition. The various classes of chondrites differ in
their proportions of CAIs, chondrules, and metal. The
FIG. 15.7. Photomicrograph of the Tieschitz (Czech Republic)
ordinary chondrite. The rounded objects are chondrules. The
image is ∼0.5 cm wide.
ordinary chondrites are by far the most common extraterrestrial material that falls to Earth. The H, L, and
LL subgroups of the ordinary chondrites contain varying proportions of iron, and differ in oxidation state
(H chondrites contain the highest concentration of
iron, mostly as metal; LL chondrites contain the least
iron, mostly in oxidized form in silicates). Another
class, the enstatite chondrites (E), are fully reduced,
containing iron only as metal. The enstatite chondrites
can also be subdivided by iron content into EH and EL
groups. Carbonaceous chondrites (C) consist of many
distinct chemical subgroups—CI, CM, CH, CO,
CV, and CK, distinguished by variations in petrography and subtle differences in chemistry. Carbonaceous
chondrites generally have high oxidation states relative to other chondrite types. The CI chondrites differ
from other groups in that they contain no chondrules,
but their chemical compositions are clearly chondritic.
The Rumaruti chondrites (R) are also oxidized, but in
other ways resemble ordinary chondrites.
Most ordinary and enstatite chondrites and a few
carbonaceous chondrites have experienced thermal
metamorphism. This has resulted in recrystallization,
which blurs the distinctive chondrule textures. The
major phases of chondrites (olivines and pyroxenes)
are preserved, although their original scattered compositions have been homogenized, and glasses have
been devitrified to form feldspars. The response of
chondrites to heating is very different from that of
terrestrial ultramafic rocks, because of the scarcity
of fluids in chondrite parent bodies. Judging from
the relative abundances of metamorphosed and unmetamorphosed chondrites in meteorite collections,
asteroids must consist mostly of thermally sintered
material. Many carbonaceous chondrites have been
affected by a different process: aqueous alteration by
fluids. The matrices of CM and CI chondrites have
been transformed to complex mixtures of phyllosilicates, and veins of carbonates and sulfates permeate some samples.
Randy Van Schmus and John Wood (1967) formulated a classification scheme for chondrites that takes
into account their compositions and thermal histories.
Increasing metamorphic grade is indicated by num-
Stretching Our Horizons: Cosmochemistry
bers from 1 to 6; these are appended to the chemical
group symbols, so that a particular chondrite may be
classified, for example, as H3 or EL6. Van Schmus
and Wood classified carbonaceous chondrites as
metamorphic grades 1 and 2, inferring that they had
experienced the least thermal effects. Grades 1 and 2
are now interpreted as reflecting varying degrees of
aqueous alteration. Thus, we are left with a somewhat confusing situation, in which type 3 chondrites
are the most primitive.
Many chondrites also show the effects of shock
metamorphism, resulting from the impacts that dis-
2. Fractionation of siderophile elements (for example,
Fe, Ir, Ni, Au) to produce depletions in some ordinary
and enstatite chondrites;
3. Fractionation of moderately volatile elements (for
example, K, Sb, Ga, Zn), leading to depletions to varying degrees in carbonaceous, ordinary, and enstatite
chondrites; and
4. Fractionation of highly volatile elements (for example,
Pb, In, Tl) to produce severe depletions in ordinary
and enstatite chondrites.
For three of the four of these fractionated groups, the
only common property is volatility, which must have been
a very important factor in controlling element behavior in
the early Solar System. This is illustrated graphically in
figure 15.8, in which elements are plotted in order of increasing volatility from left to right; degree of depletion
in each chondrite type clearly increases with increasing
volatility.
How could such element fractionations have been
accomplished? The separation of siderophile elements
may have resulted from the distinct densities of metal
and silicates, possibly leading to differences in the rates
at which grains of each accreted to form the meteorites.
Fractionation of elements by volatility requires thermal
processing in space. In the next section, we consider
volatile behavior in a more quantitative fashion.
CONDENSATION OF THE ELEMENTS
Stars are so hot that all of the matter they contain is in
the gaseous state, and any molecules are broken down
323
lodged them from the parent bodies. Despite the traumatizing effects of thermal and shock metamorphism
and aqueous alteration in chondrites, their bulk chemical compositions seem little affected. In fact, CI1
chondrites (often abbreviated to C1, a convention we
follow here), which have experienced pervasive mineralogic alteration, provide the best match for the
solar composition. Some elements, however, were
apparently mobilized by heating and were redistributed within ordinary chondrite parent bodies, and the
stable isotopic compositions of CM2 and C1 chondrites were altered by exchange with fluids.
into their constituent elements. However, as matter is
expelled from stars, it cools and many elements may
condense as solid metals or compounds. No liquids are
produced, at least for a gas of solar composition cooling
under equilibrium conditions, because the pressures in
space are so low.
When interstellar dust and gas subsequently aggregate
to form new stars, heating again occurs. Physical and
dynamic conditions in the collapsing disk-shaped cloud
that ultimately formed the Sun, called the solar nebula,
have been modeled. Most astrophysical models suggest
a hot nebula, with temperatures >1200 K near the center of the cloud but decreasing outward. These models
have led cosmochemists to conclude that most of the
presolar solid matter in the interior of the disk vaporized
and then subsequently condensed as the nebula cooled
after the Sun formed. Temperatures in the present-day
asteroid belt are never this high, but outflows from the
Sun’s poles could have lofted condensed solids into the
region where chondrites formed. Although the details of
this process are sketchy, condensation is thought to have
been an important process in establishing the chemical
characteristics of the early Solar System.
How Equilibrium Condensation Works
The condensation temperature of an element is the
temperature at which the most refractory solid phase
containing that element first becomes stable relative to a
gas of solar composition. The calculation of this temperature involves two steps. First, the distribution of a
particular element among possible gaseous molecules as
324
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
FIG. 15.8. The abundances of elements in two classes of chondrites show similar depletions related to element volatility, which increases from left to right. Serving as a volatility scale, 50%
condensation temperature represents the temperature at which half of the element is not gaseous.
All elements are normalized to cosmic abundances. (After Wasson 1985.)
a function of temperature must be determined. Next, the
condensation temperatures of all potential solid phases
that contain the element must be assessed. Comparison
of these two sets of calculations yields the most stable
phase—gas, solid, or some combination of the two. As
with other types of thermodynamic calculations, it is
necessary to consider all possible phases if the results
are to be meaningful. Because such phases normally consist of several elements, the condensation temperature of
any element may depend on the concentrations of other
elements in the system. As a consequence, previously
condensed phases may exert some control on the condensation temperatures of elements not contained in such
phases.
Worked Problem 15.1
How does one calculate the condensation temperature of corundum from a gas of solar composition? For this exercise, we adopt
a total pressure of 10−4 atm, which is appropriate for parts of
the solar nebula. We have already seen that hydrogen has by far
the highest cosmic abundance, and H 2 is much more abundant
than any other hydrogen-containing molecule below 2000 K.
Therefore,
PH = Ptotal,
2
where Ptotal is the pressure of some region of the nebula, in this
case 10−4 atm. The ideal gas law, which is applicable at these
low pressures, gives:
NH = PH /RT,
2
2
Stretching Our Horizons: Cosmochemistry
325
where NH refers to the number of moles per liter of H2. Because
2
2NH = NH, we can express the concentration of any element X
2
(in this case Al or O) in the gas as:
NX = (AX /AH)NH.
(15.1)
In this equation, AX and AH are the molar cosmic abundances
of element X and hydrogen, respectively.
A mass balance equation can now be written for each element X that describes its partitioning into various gaseous
molecules. For example, the total number of oxygen atoms can
be expressed as:
NO + NH
2O
+ NCO + 2NCO + NMgO + . . . = NO total.
2
(15.2)
All of these gaseous compounds are assumed to have formed by
reaction of monatomic elements. In the example above, water
formed by the reaction:
→ H O(g).
2H(g) + O(g) ←
2
From the free energies of the species in this reaction at the
temperature of interest, an equilibrium constant K can be calculated such that:
K = PH O /(PH PO ).
2
(15.3)
2
Equation 15.3 assumes, of course, that activities of these components are equal to partial pressures, probably a valid supposition at these low pressures. Substituting equation 15.3 into
the ideal gas law and rearranging gives:
NH
2O
= KNH NO(RT )2.
(15.4)
2
Equations of this form can be derived for each gaseous species
and substituted into the mass balance equation 15.2. Other analogous mass balance equations for all elements of interest are then
generated. The result is a series of n simultaneous equations in
n unknowns, the amounts of monatomic gaseous elements,
such as NO total. These equations can then be solved, using a
method of successive approximations, for the species composition of the gas phase at any temperature and a nebular pressure
of 10−4 atm.
For the second step, we write an equation representing equilibrium between solid corundum and gas:
→ 2Al(g) + 3O(g),
Al2O3(s) ←
for which:
log Keq = 2log PAl + 3log PO − log aAl
.
2O3
Because the activity of pure crystalline corundum is unity, this
reduces to:
log Keq = 2log PAl + 3log PO.
FIG. 15.9. Plot of log Keq versus temperature, illustrating how the
condensation temperature of corundum is obtained. Circles represent calculated conditions of equilibrium between corundum and
aluminum vapor, and squares represent the evolving Al(g) concentration in the nebula as temperature changes, determined by solving the system of mass balance equations. The intersection of
these lines marks the point at which corundum condenses from
the cooling vapor.
as the standard states. JANAF [1971] tables give free energy
data for the formation of gaseous monatomic species.) Log Keq
values for corundum calculated in this way are shown by circles
in figure 15.9. The line connecting these points divides the figure into regions in which solid corundum and gas are stable. We
then calculate log Keq for the same reaction by using the partial
pressures PAl and PO from the mass balance equations. These
values are shown by squares in figure 15.9. The temperature at
which the two values of log Keq are the same (that is, the point
at which the two lines in fig. 15.9 cross) is the temperature at
which corundum condensation occurs, in this case 1758 K. For
illustration purposes, we have treated this problem graphically,
but it could also be solved numerically by finding the temperature at which the two values of log Keq are equal.
Determining condensation temperatures for subsequently
condensed phases is a little tricky. Once the condensation temperature for any element X is reached, the equation analogous
to 15.5 must be solved again, this time adding terms for the
concentration of the crystalline phase in the appropriate mass
balance equations. If we require that the condensed phase be
in equilibrium with the gas as it cools, the gas composition will
have a lower PX than would have been the case had not the solid
phase appeared. Therefore, the affected mass balance equations
must be corrected for the appearance of each new condensed
phase.
(15.5)
Values for log Keq at various temperatures can be determined
from appropriate free energy data. (It is easy to make an error
at this step, because tabulated values of free energy of formation from the elements for corundum do not use Al(g) and O(g)
As elegant as condensation theory is, it seems likely
that kinetic factors may have been important in the
condensation process. Complexities such as barriers to
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Corundum then reacts with vapor to form spinel
(MgAl2O4) and melilite (Ca2[MgSi, Al2]SiO7), which in
turn react to produce diopside at a lower temperature.
Iron metal, containing nickel and cobalt, condenses by
1375 K. Pure magnesian forsterite appears shortly thereafter; it later reacts with vapor to form enstatite. Anorthite then forms by reaction of the previously condensed
phases with vapor. All of these refractory phases appear
above 1250 K.
Below this temperature, iron metal reacts with vapor
and becomes enriched in germanium, gallium, and copper. Similarly, anorthite reacts to form solid solutions with
the alkalis—potassium, sodium, and rubidium. Condensation of these moderately volatile elements occurs
between 1200 and 600 K. These elements are somewhat depleted in chondrites relative to cosmic values
(see fig. 15.8) and are sometimes called the normally depleted elements. Below 750 K, iron begins to oxidize,
and the iron contents of olivine and pyroxene increase
rapidly as the temperature drops further. This leads to
the interesting conclusion that the oxidation state changes
during condensation. Iron occurs only in metallic form
at high temperatures, and only in oxidized form at the
end of the condensation sequence. Metallic iron also
The Condensation Sequence
The equilibrium condensation sequence for a gas of
solar composition at 10−4 atm, as calculated by cosmochemists Larry Grossman and John Larimer (1974), is
illustrated schematically in figure 15.10. Some solid
phases condense directly from the vapor. Others form by
reaction of vapor with previously condensed phases. The
latter situation is the cosmic equivalent of a peritectic
reaction in solid-liquid systems.
Corundum (Al2O3) is the first solid phase to form
that contains a major element, although such trace elements as osmium, zirconium, and rhenium may condense
at even higher temperatures. Corundum is followed by
perovskite (CaTiO3), which may contain uranium, thorium, tantalum, niobium, and REEs in solid solution.
1.0
Ni
0.01
In
Ca
Mg
REE
Ta, Nb
U, Th
Ti
Si
Al
Diopside
Alkali Feldspar
Na
K
Rb
Forsterite
Enstatite
Ga
(Mg,Fe)SiO3
Troilite
0.1
Ag
Oxidation of Fe
Bi
Pb
Co
Ge Fe
Anorthite
S
Tl
Hydrated Silicates
Magnetite
Fraction of Element Condensed
Cu
Perovskite
Corundum
nucleation, the formation and persistence of metastable
condensates, and supersaturation by condensable species
in the gas phase are not currently understood and cannot be modeled. Treating the solar nebula as if it were in
thermodynamic equilibrium provides great insight, but
this may serve only as a boundary condition for the condensation process.
Melilite Spinel
326
Re
Zr
Os
Si
Mg
0.001
400
600
800
1000
1200
1400
1600
1800
Temperature (K)
FIG. 15.10. Summary of the calculated condensation sequence for a gas of solar composition at 10−4 atm as a function of temperature
and fraction of element condensed. The temperatures at which various minerals condense are indicated by the name of the mineral in
italics. (Modified from Grossman and Larimer 1974.)
Stretching Our Horizons: Cosmochemistry
reacts with sulfur in the vapor to form troilite (FeS) below
700 K.
The highly volatile elements, which are strongly depleted in most chondrites, condense in the interval 600
to 400 K. Examples are lead, bismuth, and thalium. Magnetite becomes a stable phase at 405 K, and olivine and
pyroxene react with vapor to form hydrated phyllosilicates at lower temperatures.
Evidence for Condensation in Chondrites
The mineralogy of chondrites is very similar to that
predicted from condensation theory. The most refractory
material occurs in the form of CAIs, one of which is
shown in figure 15.11. These occur most frequently in
carbonaceous chondrites and consist largely of melilite,
spinel, diopside, anorthite, and perovskite—all phases
predicted to condense at high temperatures. Hibonite
(CaO⋅9Al2O3), not included in the original calculations
because of the absence of appropriate thermodynamic
data, appears to take the place of corundum as an early
condensate. These inclusions sometimes contain tiny
nuggets of refractory platinum group metals as well.
Geochemical studies of CAIs demonstrate that they are
markedly enriched in other refractory trace elements.
The evidence that CAIs are condensates is compelling,
but they have not always survived intact. Many CAIs
have been melted, and some appear to be residues from
the evaporation of more volatile components.
327
Olivine and pyroxene are the major constituents of
chondrules and unaltered fine-grained matrix, and metallic iron occurs in abundance in most chondrites. Most
silicate grains in chondrules are clearly not condensates,
because they crystallized from melted droplets. However, the chondrules themselves may have formed by remelting of solid condensate material. A few relict olivine
grains in chondrules appear to have survived the melting
process; these have higher refractory trace element contents than other olivines and could be condensates. The
phyllosilicate matrix phases of many carbonaceous chondrites, commonly mixed with magnetite, may possibly
represent the low-temperature end of the condensation
sequence, although there is considerable evidence that
these phases formed by aqueous alteration in asteroids.
Thus, chondrites might be viewed as physical mixtures of material formed at various stages in the condensation sequence, although in many cases these materials
were reheated after they condensed. These were aggregated in various proportions, accounting for the differences in abundances of refractory and volatile elements
observed in ordinary, carbonaceous, and enstatite chondrite groups.
INFUSION OF MATTER FROM
OUTSIDE THE SOLAR SYSTEM
Among the billions of stars in our galaxy, only two or
three undergo supernova explosions each century. Yet
supernova nucleosynthesis accounts for the abundances
of many of the elements in the Solar System. Is supernovae debris flung so far and wide that it eventually gets
incorporated into all stars, or are there other factors at
work? Before we can answer this question, we must
consider some isotopic evidence for the infusion of supernova products into the solar nebula.
Isotopic Diversity in Meteorites
FIG. 15.11. CAI in a carbonaceous chondrite. The dark cores
consist of spinel, hibonite, and perovskite (all early condensing
phases), rimmed by later condensing melilite and diopside.
Variability in isotopic composition has been demonstrated for many elements in most geological samples. In
chapters 13 and 14, we learned that such diversity can
arise from (1) isotope fractionation processes, (2) mixing,
(3) radioactive decay, and (4) interaction with cosmic
rays. Given this variability, how could we use isotopes to
search for and recognize materials that formed outside
the Solar System? The similarity between minerals in
the condensation sequence and those that constitute CAIs
328
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
suggests that CAIs are some of the oldest surviving relicts of the early solar nebula. This conclusion is corroborated by radiometric dating of these inclusions. For
example, U-Th-Pb dating methods indicate ages of about
4.56 × 109 yr, and initial 87Sr/ 86Sr isotopic compositions
are the lowest known of any kind of materials. If we
wish to search for supernova products that have not been
completely blended with other Solar System matter, these
inclusions provide a good place to start.
The oxygen isotopic compositions of CAIs were first
analyzed by Robert Clayton and his coworkers (1973).
Figure 15.12 illustrates the relationship between δ17O
and δ18O in these inclusions and terrestrial rocks. The
terrestrial mass fractionation line has a slope of + 1–2 , because any separation of 17O from 16O (a difference of
1 atomic mass unit) will be half as effective as the separation of 18O from 16O (a difference of 2 atomic mass
units). The trend of the inclusion data, clearly distinct
from the mass fractionation line, must signify some other
process at work. Clayton and coworkers interpreted this
as a mixing line, joining “normal” solar system oxygen
on the terrestrial mass fractionation line to pure 16O (to
which the mixing line in this figure extrapolates). Pure
16
O is produced by explosive carbon burning in supernovae. Thus, it was argued that CAIs must contain an
admixture of interstellar grains containing supernovagenerated oxygen. This exotic component may have
served as a substrate for nucleation of other material
during condensation, or it may have survived evaporation in refractory mineral sites. Another possibility is that
oxygen isotope diversity occurred in the hottest part of
the nebula by non-mass-dependent gas phase reactions
that produced 16O enrichments.
The exciting discovery of oxygen isotopic diversity
initiated a flurry of research activity to find anomalies in
the isotopic compositions of other elements. Nucleosynthesis theory leads us to expect anomalies in magnesium and silicon in samples with large concentrations
of 16O, but the results obtained so far are puzzling. Although nuclear anomalies in magnesium and silicon, as
well as titanium, calcium, barium, and strontium, have
been found, these occur in only a very few CAIs and are
not generally correlated.
One of the most interesting isotopic anomalies in these
objects indicates the former existence of now-extinct
radionuclides. Excess amounts of 26Mg, attributed to
FIG. 15.12. The relationship between 16O, 17O, and 18O in CAIs and chondrules in carbonaceous
chondrites. Mass fractionation produces a line of slope + –21 in this diagram, as observed for terrestrial materials. CAIs and chondrules define an apparent mixing line between “normal” solar system
material, such as that along the terrestrial mass fractionation line, and pure 16O. δ values are per
mil relative to SMOW. (From Clayton et al. 1973.)
Stretching Our Horizons: Cosmochemistry
radioactive decay of 26Al, have been found in CAIs. The
half-life of 26Al is only 0.7 Ma, so its incorporation as a
“live” radionuclide indicates that formation of the solar
nebula occurred very soon after nucleosynthesis. In fact,
after just a few million years had elapsed, the abundance
of this nuclide should have decreased to the point at
which it would be undetectable. Like 16O, 26Al is produced by explosive carbon burning. The occurrence of
this isotope provides a time scale for the addition of supernova material to the solar nebula. Besides 26Al, other
short-lived radionuclides formed in supernovae and found
in chondrites include the following (half-life in parentheses): 41Ca (0.1 Ma), 60Fe (1.5 Ma), 53Mn (3.7 Ma),
107Pd (6.5 Ma), 182Hf (9 Ma), 129I (15.7 Ma), and 244Pu
(82 Ma).
Worked Problem 15.2
How would one demonstrate that 26Al was incorporated as a
“live” radionuclide? First let’s review the evidence that 26Al,
dead or alive, was added to CAIs. The decay product of this
radionuclide is 26Mg, so it is only necessary to look for an excess
of this isotope. This turns out to be a rather difficult measurement, however, because 26Mg is already an abundant isotope,
constituting of ∼11% of normal magnesium. Only in samples
with a high original proportion of 26Al would the addition of a
relatively small amount of extra radiogenic daughter 26Mg be
detectable.
Caltech graduate student Typhoon Lee, along with Dimitri
Papanastassiou and Gerry Wasserburg (1977), found enhancements of 26Mg that changed the relative abundance of this isotope in inclusions by up to ∼11.5%. Isotopic fractionation can
be ruled out as a cause for this anomaly, because there should
have been a similar but smaller effect in 25Mg relative to 24Mg,
which is not observed. Thus, it is clear that the 26Mg anomaly
must be nuclear in origin; but there are nuclear reactions other
than 26Al decay that could give rise to this product.
The origin of the excess 26Mg can be established by demonstrating that 26Mg is correlated with aluminum in individual
phases of the inclusions. Lee separated anorthite, melilite,
spinel, and diopside, each of which contains different amounts
of aluminum and magnesium. The results are illustrated in figure 15.13. (A word of caution is in order here. This diagram is
not analogous to the plots utilized for Rb-Sr and Nd-Sm chronology in chapter 14. 26Mg is not the radioactive parent, but a
stable isotope of magnesium. If the ordinate were 26Al/ 24Mg,
this would be an isochron diagram and the slope of the regression line would be a function of time. Such a diagram could only
be constructed during the first few million years after nucleosynthesis, when 26Al was still measurable.) Figure 15.13 shows
that phases that initially had high total aluminum and low
total magnesium contents incorporated more 26Al and there-
329
FIG. 15.13. Correlation between 26Mg and 27Al, both normalized
to 24Mg, in a CAI. This relationship suggests that the 26Mg excess
was produced by decay of the extinct radionuclide 26Al. (From Lee
et al. 1977.)
fore now have the highest 26Mg. Spinel and diopside, with very
low Al/Mg ratios, have virtually no excess 26Mg. Melilite,
which contains both elements, has a 4% 26Mg excess and anorthite, with a high Al/Mg ratio, has a 9% excess. Because the
radiogenic 26Mg now occurs in crystallographic sites occupied
by aluminum, it must have been incorporated as 26Al into these
phases.
From the slope of the correlation line in figure 15.13, it is
possible to calculate the initial isotopic ratio 26Al/ 27Al at the
time the inclusion formed. This value is ∼5 × 10−5, a value so
common in CAIs that it is normally taken as the canonical
26Al/ 27Al ratio for the early Solar System.
A Supernova Trigger?
The presence of freshly synthesized radionuclides
in chondrites demands that these meteoritic materials
formed soon after nucleosynthesis. Because of the long
travel times across vast interstellar distances, these data
further suggest that a supernova occurred in our cosmic
backyard at about the time that the Solar System formed.
The time from synthesis to injection into the collapsing
solar nebula must have been on the order of 300,000 yr,
consistent with the half-lives of the shortest lived radionuclides. Does this not seem highly unlikely? That would
indeed be the case if there were no connection between
the injection of short-lived radionuclides and solar
nebula formation. Astrophysicists Al Cameron and J. W.
Truran (1977) turned the argument around, proposing
that a nearby supernova caused the formation of the
solar nebula. They suggested that the advancing shock
wave generated by a supernova explosion triggered the
330
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
collapse of gas clouds surrounding it to form nebulae
and ultimately new stars. The isotopic anomalies found
in chondrites would be a natural consequence of the
blending of small amounts of supernova products with
matter indigenous to the gas-dust cloud. Any evidence
of this addition of superova products in materials other
than chondrites would have been subsequently obliterated by geological processing.
Astronomers have now observed the formation of
new stars at the leading edges of expanding supernova
remnants, providing independent evidence that such a
process is possible. The idea of a supernova trigger
neatly explains some of the most vexing questions in
cosmochemistry. It is also an interesting demonstration
of the utility of geochemistry in solving interdisciplinary
problems.
THE MOST VOLATILE MATERIALS:
ORGANIC COMPOUNDS AND ICES
The Discovery of Stardust in Chondrites
Many organic molecules have been discovered in space
by using microwave techniques. Detected organic compounds include most of the functional groups of biochemical compounds, except those involving phosphorus.
Much of the interstellar medium is an inhospitable place
for organic molecules, because any molecules that form
are promptly dissociated by ultraviolet radiation. Dense
interstellar clouds, however, are relatively opaque to such
radiation, and organic compounds form readily within
them. We can recognize these extraterrestrial organic molecules because they are enriched in deuterium. Under
the frigid conditions within these clouds, deuterium is
utilized in preference to hydrogen in chemical reactions.
It combines with carbon, oxygen, nitrogen, and sometimes sulfur to produce an array of interstellar organic
molecules.
When a molecular cloud collapses during star formation, the nebula inherits these elements already in some
state of molecular complexity, although simpler compounds such as CO and CH 4 may predominate. These
molecules probably occur as mantles on interstellar silicate dust grains.
When the solar nebula formed, organic compounds
in what is now the inner Solar System were heated and
partially decomposed into gases plus more refractory
residues, a process called pyrolysis. Complex molecules
could have survived in the outer portions of the nebula,
and some of these were probably incorporated into
comets prior to their ejection out of the vicinity of the
Jovian planets. During subsequent cooling of the nebula,
condensation of the vaporized fraction ensued. Some
Cosmochemist Edward Anders recognized that solid
grains formed from matter expelled from supernova
explosions would likely be tagged with exotic r-process
nuclides. In 1987, after working on and off for nearly
two decades, he and his coworkers finally succeeded in
isolating stardust from chondrites. Their method involved
the stepwise dissolution of meteorites in acids while monitoring the presence of carriers of exotic nuclides (in this
case, isotopes of neon and xenon) at each step in the
residue. After numerous steps, less than a thousandth of
the original of chondrite mass remained, and Anders finally arrived at his prize: a tiny amount of interstellar
powder that proved to be miniature diamonds!
Buoyed by this success, Anders’ University of Chicago group soon separated grains of another presolar
mineral, silicon carbide. By measuring the isotopic composition of silicon and carbon with the ion microprobe,
they demonstrated that these, too, were stardust. Subsequent work has isolated other interstellar phases, including graphite, corundum, and titanium carbide. The
careful chemistry necessary to separate and characterize
minute quantities of these grains tests the analytical skills
of cosmochemists, but the results are astounding. Isotopic measurements of stardust provide ground truth for
astrophysical models of nucleosynthesis in stars. With
presolar grains, we can see the nuclear fusion processes
at work. Who would have dreamed that the chemical
analysis of tiny grains in meteorites would help us understand how the stars shine?
It is sobering to realize that most of the carbon atoms in
our bodies were produced by nucleosynthesis in other
stars. But how did they arrive at their present molecular
complexity? Did all chemical evolution of organic compounds occur in the Earth’s atmosphere and oceans, or
was the process initiated at some other location? And
for that matter, where did the gases that make up the
atmosphere and oceans come from? We now turn to
the cosmochemistry of the most volatile elements—
hydrogen, carbon, nitrogen, and oxygen—which are
depleted in chondrites relative to cosmic abundances.
Extraterrestrial Organic Compounds
Stretching Our Horizons: Cosmochemistry
organic compounds are extremely volatile, so condensation occurred at low temperatures. As the temperature
dropped, it became possible to stabilize and preserve
organic compounds in regions where high temperatures
had previously prevented their condensation, so we can
envision these components as infiltrating the inner solar
system late in the evolutionary history of the nebula.
These compounds may have been relict interstellar materials transported from greater radial distances, local
condensates, or both.
Let’s first consider how organic compounds condense.
Under equilibrium conditions, the dominant carbon and
nitrogen species in the nebula should have been CO and
N2 at high temperatures, converting to CH 4 and NH3 at
low temperatures. Predicting the condensation of messy
organic compounds is difficult because of inadequate
thermodynamic data; in addition, kinetics must have
played a major role.
One condensation model involves the Fischer-Tropsch
type process, based on a commercial synthesis of hydrocarbon fuels from coal. As the nebula cooled below
600 K, adsorption of CO, NH3, and H 2 onto previously
condensed mineral grains led to the formation of a variety
of compounds. The Fischer-Tropsch synthesis depends
on catalytic activity on the mineral substrate. It has been
proposed that complex organic compounds, as well as
H 2O and CO2, formed in this way. This is a very difficult
model to test. Laboratory experiments indicate that the
expected compounds can be produced, but only under
very different conditions from those found in the nebula.
Because the dominant nitrogen species at nebular temperatures is N2 rather than NH3, the Fischer-Tropsch
systhesis would need to have been very efficient. The
formation of complex organic molecules may also have
been accomplished in the nebula by other processes,
such as photochemical reactions driven by ultraviolet
radiation.
Chondritic meteorites, especially carbonaceous chondrites, provide the only direct sampling of organic compounds from the nebula. Two categories of carbonaceous
materials have been isolated from these samples. One
consists of amorphous carbon mixed with macromolecular organic material similar to terrestrial kerogen.
At least some fraction of this acid-insoluble component
contains noble gases, whose elemental and isotopic compositions suggest that it may be relict interplanetary
material. The smaller category consists of organic compounds that are readily soluble and thus easier to char-
331
TABLE 15.2. Distribution of Carbon in the Murchison
Carbonaceous Chondrite
Species
Acid-insoluble carbonaceous phase
CO2 and carbonate
Hydrocarbons
Aliphatic
Aromatic
Acids
Monocarboxylic C2-C8)
Hydroxy (C2-C5)
Amino acids
Alcohols (C1-C4)
Aldehydes (C2-C4)
Ketones (C3-C5)
Ureas
Amines (C1-C4)
N-heterocycles
Abundance
1.3–1.8 wt. %
0.1–0.5 wt. %
12–35 ppm
15–28 ppm
170 ppm
6 ppm
10–20 ppm
6 ppm
6 ppm
10 ppm
2 ppm
2 ppm
1.1–1.5 ppm
Data from Wood and Chang (1985)
acterize. Among these are many of the compounds present
in living systems, such as carboxylic and amino acids.
At one time, it was argued that such compounds were
biologic in origin. However, the molecular structures of
compounds in chondrites do not usually show the preferred chiral forms that characterize those produced biologically. Most amino acids, for example, are racemic
mixtures of enantiomers, rather than being dominated by
l-forms as in the amino acids in the terrestrial biosphere.
The recent discovery of a slight excess of one enantiomer
in several nonbiologic amino acids in chondrites suggests
that organic synthesis in space may somehow introduce
slight variations in d- and l- abundances. Table 15.2
lists the carbonaceous materials identified in one wellcharacterized carbonaceous chondrite.
These two carbonaceous components have very different stable isotopic compositions. For the kerogen component, measured mean values of δD, δ13C, and δ15N
are +800, −12, and +20, respectively; corresponding
values for the soluble organic component are +250, +25,
and +75. These major differences appear to require either more than one source or more than one production
mechanism for meteoritic carbonaceous materials. Such
data are at least consistent with the idea that some organic components may be relict interstellar phases.
Ices—The Only Thing Left
What happened to the simple gaseous molecules, such
as CH 4, NH3, and H 2O, that were not used in the production of more complex organic compounds? Under
332
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
equilibrium nebular conditions, some of the water should
have been incorporated into hydrous silicates, but there
was no obvious sink for methane and ammonia. At temperatures below ∼200 K, these compounds finally condensed into ices. This probably did not occur in the inner
Solar System, but evidence of abundant ices is obvious
at greater radial distances. The Jovian planets and their
satellites consist of rocky cores surrounded by great
thicknesses of these ices or their high-pressure equivalents. Comets are probably dirty snowballs containing
considerable ice. It is generally thought that comets are
relatively pristine samples of condensed matter, containing even the most volatile elements in their cosmic
proportions.
It should come as no surprise that these ices are associated physically with organic compounds, because both
condensed at low temperatures. We know this from spectroscopic observations of comet comas. It is not possible
to reconstruct the primitive organic molecules that may
have occurred in comet nuclei, however, because solar
irradiation breaks down all but the simplest molecules
before they are blown away into the tail.
Worked Problem 15.3
What are the relative weight proportions of ices to rock in a
fully condensed solar nebula? Let’s restrict this calculation to
those elements with cosmic abundances >104 atoms per 106
atoms of silicon (table 15.1). Of these, we assume that the following elements condense as oxides to form rock: Na, Mg, Al,
Si, Ca, Fe, and Ni. (Under reducing conditions appropriate for
the nebula, some Fe and Ni condense as metals, but we can ignore this complication for now.) The following elements condense as ices: C, N, O, Ne, S, Ar, and Cl. We can model all of
these ices as hydrides, except for the noble gases. A correction
must be made to oxygen to adjust for the amount already partitioned into oxides. The element He does not condense, but
remains in the gas phase; some H is consumed in hydride ices,
and the remainder is gaseous.
Rock components are calculated in the following way. For
sodium, convert atomic abundance to moles by dividing by Avogadro’s number:
5.7 × 104 atoms/6.023 × 1023 atoms mol−1
= 9.46 × 10−20 mol.
Now multiply half this number of moles by the gram formula
weight of Na2O:
(9.46 × 10−20/2 mol)(61.98 g mol−1)
= 2.93 × 10−18 g Na2O.
If we follow the same procedure for the each of the other elements condensable as rock, these sum to 3.04 × 10−16 g of
oxides.
Now repeat the procedure for elements condensable as ices,
using the gram formula weight of the hydride for all but the
noble gases. For oxygen, we obtain:
2.01 × 107 atoms/6.023 × 1023 atoms mol−1
= 3.34 × 10−17 mol.
In the case of oxygen, an adjustment must be made to correct
for the amount already partitioned into oxides. Summing the
amounts of oxygen needed for Na2O, MgO, Al2O3, SiO2, CaO,
FeO, and NiO, we find that 7.03 × 10−18 mol of oxygen are required to combine with the condensable elements to make rock.
Subtracting this from the total moles of oxygen available leaves
2.63 × 10−17 mol of oxygen that can be condensed as ice. Multiplying this result by the gram formula weight of the hydride
gives:
(2.63 × 10−17 mol)(18.016 g mol−1)
= 4.73 × 10−16 g H2O.
Repeating for other elements condensable as ice, we obtain a
total of 1.02 × 10−15 g.
The ratio of these two numbers is:
ice/rock = 1.02 × 10−15/3.04 × 10−16 = 3.38.
Thus, potential ices are greater than three times more abundant
by weight than potential rock. Under reducing conditions, even
more oxygen would be available for the production of ice.
From this comparison, it is easy to see why the outer planets
consist mostly of gases derived from ices.
A TIME SCALE FOR CREATION
The short-lived radionuclides in chondrites are potentially useful as a chronometer, if their initial relative
abundances were uniform throughout the solar nebula.
These isotopic systems give only relative time, but with
very high resolution. To get absolute time scales, the
short-lived nuclide time scales must be coupled with
precise absolute ages, such as are determined from the
long-lived U-Pb system. Consistency among all the
chronometers would indicate an originally uniform
nebula composition. These measurements are very
challenging, but most evidence favors a homogeneous
nebula.
Unaltered CAIs have the oldest measured U-Pb ages
(determined very precisely to be 4.566 ± 0.002 b.y. by
Claude Allegre and coworkers [1995]). This is normally
Stretching Our Horizons: Cosmochemistry
taken to be the age of the solar system. CAIs also have
the highest initial abundances of short-lived 26Al, 41Ca,
and 53Mn, as appropriate for the earliest formed nebular materials. Chondrule ages based on 26Al decay are
usually several million years younger than CAIs. The
formation times of chondrites are impossible to date
directly, because these rocks are assortments of components that formed at different times and the accretion
process did not reset their radiometric clocks. After
chondritic asteroids accreted, however, many experienced
metamorphism (probably heated by 26Al decay). As
parts of these asteroids subsequently cooled through
appropriate blocking temperatures, their radiometric ages
were reset. The oldest measured chondrite, Ste. Marguerite (H4), has an absolute age of 4.563 ± 0.001 b.y.,
an 26Al age of 5.6 ± 0.4 m.y. after CAIs, and a 182Hf age
of 4 ± 2 m.y.
The parent asteroids of achondrites and iron meteorites obviously had their radiometric systems reset by
melting and differentiation. The oldest crystallization
age for an achondrite is 4.558 ± 0.001 b.y. 182Hf dating
of iron meteorites also points to rapid differentiation
within ∼4 m.y. after CAI formation. These data indicate
that asteroid-size objects (tens to hundreds of kilometers
in diameter) took less than a few million years to form and
differentiate. These planetesimals were rapidly heated to
varying degrees, resulting in metamorphism or melting.
Determining when the Earth and other terrestrial
planets formed is difficult because of their prolonged geologic evolution. In worked problem 14.5, we learned
how the 182Hf system can be used to constrain the timing of core formation. 182Hf data for Martian meteorites
indicate that Mars was differentiated within ∼13 m.y. of
Solar System formation (Yin et al. 2002). For the Earth,
182Hf chronology suggests formation within ∼29 m.y.
of CAI formation. Mars is smaller than the Earth and its
formation was probably completed before that of the
Earth. The abundances of radiogenic 40Ar and 129Xe in
the Earth’s atmosphere indicate atmospheric retention
beginning at 4.46 b.y. Thus, the terrestrial planets accreted after thermal processing of smaller asteroids.
Because planetesimals in the inner Solar System were
probably even more likely to have been melted, it is probable that they formed from already differentiated bodies.
However, the bulk compositions of these asteroids were
chondritic, so the Earth also has a bulk composition like
chondrites.
333
ESTIMATING THE BULK COMPOSITIONS
OF PLANETS
Some Constraints on Cosmochemical Models
Estimating the bulk compositions of planets is difficult, because differentiation ensures that no sample from
anywhere within the body has the composition of the
whole body. However, there are some constraints that
allow us to test cosmochemical models.
Knowing a planet’s mass (obtainable from its effects
on the orbits of moons and nearby spacecraft) and volume (calculated from its measured diameter), we can
calculate its mean density. The observed mean densities
of the planets range from 3.9–5.5 g cm−3 for the terrestrial planets to 0.7–1.6 g cm−3 for the Jovian planets.
Most of this difference can be explained by incorporation of varying amounts of rock and ice to form these
bodies, due to radial temperature differences within the
solar nebula. Within the terrestrial planet group, there
are density variations, less dramatic but still of sufficient
magnitude to suggest important chemical differences. To
compare these density differences, we must correct mean
densities for the effects of self-compression due to gravity.
The uncompressed mean densities (in g cm−3) for the terrestrial planets are: Mercury, 5.30; Venus, 4.00; Earth,
4.05; Mars 3.74; Moon, 3.34. Even after the effects of
self-compression have been compensated for, significant
density variations remain.
Several explanations have been put forward to explain these data. Cosmic abundance considerations indicate that planetary density variations must involve iron,
the only abundant heavy element that occurs in minerals
with different densities. Harold Urey (1952) proposed
that varying proportions of silicate and metal, with densities of 3.3 and 7.9 g cm−3, respectively, had been mixed
to produce the inner planets. Density would then be a
straightfoward function of metal/silicate ratio. Urey envisioned that metal-silicate fractionation occurred in the
nebula, due to differences in the physical properties of
these materials. A few years later, Ted Ringwood appealed to variations in oxidation state to explain density
differences. The effect of redox state on chondrites illustrates the principle: if all the iron and nickel in these
meteorites were converted to oxides, the density would be
3.78 g cm−3. If FeO were instead fully reduced to metallic iron, the density would increase to 3.99 g cm−3. This
334
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
density range is not sufficient to encompass the mean
densities of all the planets, so it is necessary to invoke
some metal-silicate fractionation as well. Both of these
proposals find some support in chondrites. The various
chondrite groups can be distinguished on the basis of
both Fe/Si and Fe0/FeO ratios.
The value of its moment of inertia constrains how
mass is distributed within a planet’s interior. This can
also be determined from seismic data, but aside from the
Earth, we have almost no seismic information on the
planets. The moment of inertia can be determined from
the observed degree of flattening of a planet as it rotates,
or from the precession of its rotational axis over time.
Although moment of inertia is most affected by the presence and size of a metallic core, it also depends on the
densities and thicknesses of mantle and crustal layers.
Once a planet’s bulk composition has been modeled and
the proportions and mineralogies of core, mantle, and
crust are estimated, its mean density and moment of inertia can be calculated. Comparisons between the calculated and observed mean density and moment of inertia
constrain acceptable geochemical models for planet bulk
compositions.
In the preceding section, we learned that the bulk composition of the Earth is chondritic, but what does that
really mean? To cosmochemists, the term chondritic indicates that the relative abundances of elements with
similar volatility are the same as in chondrites. We can
clarify this statement by considering figure 15.14, which
presents analyses of lanthanum, potassium, and uranium
100,000
10
Earth
Moon
4 Vesta
Mars
10,000
K (ppm)
U (ppm)
1
0.1
0.01
0.001
0.01
in a variety of samples. Included in these diagrams are
data from chondrites, terrestrial and lunar rocks, and
meteorites from Mars and from the differentiated asteroid Vesta. (Martian meteorites are basalts and ultramafic rocks classified as shergottites, nakhlites, and
chassignites, collectively called SNCs; Vesta meteorites
are basalts and ultramafic rocks classified as howardites,
eucrites, and diogenites, collectively called HED achondrites.) Lanthanum, potassium, and uranium all form
lithophile cations having large ionic sizes, so they are
incompatible elements. In other words, they are fractionated together during melting and crystallization, so
that La/U or K/U ratios do not change from their original values, even though the absolute concentrations may
be modified. Stated another way, these three elements
exhibit similar geochemical behavior. Planetary differentiation has not significantly modified La/U or K/U ratios
from their original values; thus, the measured ratios in
any samples should give the bulk planet ratios.
However, these elements do exhibit differences in
cosmochemical behavior. Uranium and lanthanum are
refractory elements, whereas potassium is volatile. In
figure 15.14a, we can see that all the bodies in the study,
including the Earth, have a chondritic U/La ratio, because both U and La are refractory elements. In contrast,
the K/La ratios vary from body to body (Fig. 15.14b),
because elements with different volatilities were fractionated in the solar nebula. All the bodies shown are
in fact depleted in volatile elements such as K, relative
to chondrites. Recall that the chondrites themselves are
(a)
Refractory
versus
Refractory
Chondrites
0.1
1
10
La (ppm)
100
1000
Earth
Moon
4 Vesta
Mars
Chondrites
(b)
Volatile
versus
Refractory
100
1000
10
0.01
0.1
1
10
100
1000
La (ppm)
FIG. 15.14. Fractionation of volatile and refractory elements is shown by comparison of the ratios of uranium and potassium to lanthanum in planetary samples. All the bodies shown have chondritic ratios of refractory elements (U/La) (a), but differ in the ratio of
volatile K to refractory La (b). (After Wänke and Dreibus 1988.)
Stretching Our Horizons: Cosmochemistry
depleted in volatile elements to varying degrees, relative
to the Sun. This depletion of volatiles, although not well
understood, is an important characteristic of planets that
must be accounted for in cosmochemical models. Equally
important, though, is the observation that planets have
chondritic bulk compositions, in terms of the relative proportions of elements having similar volatility.
The Equilibrium Condensation Model
The condensation sequence already described defines
a succession of equilibrium assemblages that form from
a cooling gas of solar composition. The basis for the
equilibrium condensation model for planet formation,
as originally devised by John Lewis (1972), is that solids
were thermally equilibrated with the surrounding nebula gas, and thus had compositions dictated by condensation theory. The temperature in the nebula is assumed
to have been highest near the Sun and to have declined
outward. Whatever solids had already condensed at
particular radial distances were accreted to form the various planets and thereby isolated from further reaction
with the gas. Any uncondensed materials were somehow
flushed from the system.
At the location of Mercury, the temperature (1400 K)
was such that refractory elements such as calcium and
aluminum had completely condensed and all iron was
metallic. Magnesium and silicon had only partly condensed at this point. The high uncompressed mean density of Mercury is thus explained by the occurrence of
iron only in its most dense, metallic state, as well as the
high iron abundance relative to magnesium and silicon.
The temperature (900 K) at the orbital distance of Venus
allowed complete condensation of magnesium and silicon, but no oxidation of iron. Earth formed in a somewhat cooler (600 K) region, in which some iron reacted
to form sulfide and ferrous silicate. The addition of sulfur, with an atomic weight greater than the mean atomic
weight of the other condensed elements, resulted in a
mean density for Earth that is higher than that of Venus.
The lower nebular temperature (450 K) for Mars permitted oxidation of all of the remaining iron and its incorporation into silicates. Its mean density was further
reduced by hydration of some olivine and pyroxene to
form phyllosilicates. Conditions in the asteroid belt were
appropriate for the formation of carbonaceous chondrites
and, further outward, ices were stable with silicates.
These combined to form the Jovian planets.
335
Worked Problem 15.4
How can we estimate a planetary bulk composition using the
equilibrium condensation model? This is a very long and complex problem, so we simply illustrate how it is done by examining how to set up the problem. It is first necessary to define
the dimensions of a “feeding zone,” from which each planet will
draw in material during accretion. These annular zones can be
the exclusive property of each planet or can be overlapping.
The condensation sequence represented in figure 15.10 was
calculated at one pressure. However, pressure and temperature
decreased outward in the solar nebula. Lewis (1972) generalized the condensation sequence by constructing a plot showing
how some of these reactions varied with temperature and pressure, illustrated in figure 15.15. Lewis adopted a P-T gradient,
also shown in this figure, that produced mineral assemblages
at the locations of planets whose densities corresponded with
those of the planets. The temperature at which a mineral first
becomes stable was taken from the intersection of its reaction
line with the P-T gradient in figure 15.15. The calculations of
Grossman (1972) were then used to estimate the degree of condensation below that temperature. The composition of each
FIG. 15.15. Equilibrium condensation in the solar nebula can
approximately account for the compositions of the terrestrial and
Jovian planets, if temperature and pressure decreased radially
away from the Sun, as shown by the heavy curved line. (Modified
from Barshay and Lewis 1976.)
336
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
planet was defined by integrating, over equal increments of distance, the compositions of solid material within its feeding zone.
This model satisfactorily accounts for planetary mean
densities, but a number of other problems are evident.
First, it is not possible to explain the large difference in
mean density of the Earth and Moon using equilibrium
condensation. Also, it is unrealistic to assume that a
large planet would accrete only condensates formed at a
single temperature. Mixing would have been unavoidable and, in fact, orbital calculations suggest that the
feeding zones for the terrestrial planets overlapped considerably. The kinetic barriers to gas-solid equilibrium
must have also played a role, which would have resulted
in further mixing of materials with different thermal histories. Finally, at the condensation temperatures for
Venus, Earth, and Mars, no condensed volatiles are possible. The atmospheres of these planets and the Earth’s
oceans are thus an embarrassment to equilibrium condensation models, as is the presence of Fe3+ in terrestrial
rocks. Despite these problems, the equilibrium condensation model is very useful as one of several end members
of idealized scenarios for planet formation.
The Heterogeneous Accretion Model
Another model for planet formation is based on the
premise that all solids were vaporized in a hot solar nebula. During cooling, condensation ensued. As each new
phase formed, it was immediately accreted. The equilibrium condensation sequence was, in a sense, beheaded,
because the removal of each solid phase prevented its
subsequent reaction with the vapor. This is somewhat
analogous to fractional crystallization, which we discussed in chapter 12, but in a cosmic setting. At some
point, the process was interrupted as the nebula dissipated. Because temperatures in the nebula decreased radially outward, the progress of condensation was not
everywhere the same. Mercury might have passed through
only the stages at which Ca,Al-rich minerals, metallic
iron, and some magnesium silicates condensed. Planets
farther out may have acquired veneers of phyllosilicateand organic-rich material before condensation was halted.
This model, originally developed by Karl Turekian
and Sydney Clark (1969), is called heterogeneous accretion, because it results in layered planets, with the most
refractory material in cores and more volatile phases at
the surfaces. In practice, the model cannot be applied exclusively, because it leads to bizarre results. For example,
the Earth should contain no FeO or FeS, because all of
the iron was buried as metal in the core and isolated from
further reaction through mantling by magnesian silicates.
The advocates of this model do not practice undeviating
loyalty, however; they ask only that equilibrium condensation be modified to allow some heterogeneous accretion to take place. The flexibility in this model makes
it impossible to derive a unique composition for a planet.
Heterogeneous accretion does, however, suggest a
source for volatile compounds in planetary atmospheres
and the Earth’s hydrosphere. Volatiles were accreted late
in the condensation sequence onto planetary surfaces.
The model also explains why rocks of the Earth’s crust
and mantle contain quantities of Fe2+ and Fe3+, which
are not stable in the presence of metallic iron. If reduced
and oxidized iron had ever been mingled, they should
have reacted to a uniform, intermediate oxidation state.
A major difficulty with this model is that the requirement for total vaporization of dust in the nebula and for
very rapid accretion are at odds with current conceptions of nebular conditions.
The Chondrite Mixing Model
If we accept the proposition that chondrites are leftover planetary building blocks, it is not necessary to rely
on abstract concepts of nebular condensate material to
construct models. In this case, we have actual samples
that can be used in various proportions in cosmochemical models. If planets formed by accretion of already
differentiated planestimals, the chondrite mixing model
may still be valid, because the bulk compositions of these
bodies were chondritic.
Our task would be simple if the compositions of the
planets matched those of individual chondrite groups.
As appealing a model as this is, it doesn’t quite work. We
have already seen that planets show the same kinds of
volatile element depletions that chondrite classes do, but
planetary depletions are more extreme. Mixing various
classes of chondrites provides better matches for planetary compositions, but even this is not enough to produce
a perfect fit.
One example of a chondrite mixing model is based
on the work of German cosmochemists Heinrich Wänke
and Gerlind Dreibus (1988). Their model mixes two
components, both having chondritic compositions, to
Stretching Our Horizons: Cosmochemistry
produce the planets. Component A is refractory, free of
all elements with equal or higher volatility than sodium,
but containing all other elements in C1 abundance ratios. It is also reduced, with all iron present as metal; its
degree of reduction is comparable to that of enstatite
chondrites. This component, with its complete absence
of highly volatile elements, does not correspond to any
particular chondrite class. However, an end member extremely depleted in volatiles is necessary to explain the
volatile depletions seen in planets. Component B is
oxidized and contains all elements, including volatiles,
in C1 abundances. In other words, it is C1 chondrite.
Wänke and Dreibus were able to model the compositions of Earth and Mars by mixing these two chondrite
components in differing amounts. The relative proportions of these components were constrained by element
ratios in terrestrial mantle rocks and in Martian meteorites. They also argued that heterogeneous accretion
(component A first, with only limited mixing of A and B
later) best explained the composition of the Earth’s
mantle, whereas homogeneous accretion (thorough mixing of components A and B during accretion) was neces-
337
sary for Mars. The elemental abundances in the silicate
(that is, noncore) parts of the Earth and Mars, based on
the Wänke-Dreibus chondrite mixing model, are compared in figure 15.16.
Several other recent chondrite mixing models for Mars
have been constructed in an attempt to duplicate the
oxygen isotopic composition of SNC meteorites. One
model mixes ordinary and carbonaceous chondrites,
whereas another mixes ordinary and enstatite chondrites
to arrive at the observed δ17O and δ18O. The admixed
materials obviously have very different oxidation states,
as do components A and B in the Wänke-Dreibus model,
so redox reactions involving metal, sulfide, carbon,
water, and the FeO in silicates must be used to adjust
these amounts to an equilibrium assemblage for the bulk
planet.
Other chondrite mixing models use the components
of chondrites, rather than bulk chondrites, to make
planets. One formulation, by John Morgan and Edward
Anders (1979), identifies components such as refractory material, metal, sulfide, silicates, and volatile material. Although these were actually defined as chemical
FIG. 15.16. Comparison of element abundances in the silicate portions of Earth and Mars, the latter calculated from SNC meteorites by
Wänke and Dreibus (1988). All elements are normalized to chondrites and 106 silicon atoms.
338
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
components, they have physical counterparts in chondrites. For example, CAIs qualify as refractory material,
chondrules are silicates, and matrix is volatile material.
The chondrite model is based on the postulate that the
materials that made the planets experienced the same
fractionations as recorded by chondrites. Because the
components can be mixed in different proportions to reproduce the chemistry of the various chondrite groups,
they clearly were fractionated in the nebula. Thus, these
same components can be blended to make planets, although more extreme proportions of some may be required. Each component carries its own suite of elements
in approximately cosmic proportions, but the components have been fractionated from each other to varying
degrees. This procedure seems to work reasonably well
for most elements, but breaks down for elements that do
not partition cleanly into one component. For example,
significant errors in volatile element abundances may
GEOCHEMICAL CONSTRAINTS ON THE ORIGIN OF THE MOON
At various times, five principal hypotheses have been
favored as possible explanations for the origin of the
Earth’s moon. These are:
1. Intact capture: the Moon formed elsewhere in the
solar system and was gravitationally captured by
the Earth.
2. Binary accretion: the Moon accreted as a companion to the Earth from materials captured in
geocentric orbit.
3. Rotational fission: the Moon separated from an
early Earth that was rotating so rapidly that it
was unstable.
4. Collisional ejection: off-center collision of a very
large projectile with the Earth ejected debris into
geocentric orbit, from which the Moon ultimately
accreted.
5. Disintegrative capture: a very large body that barely
missed the Earth was tidally disrupted. Debris was
retained in a geocentric orbit and accreted to form
the Moon.
Aside from dynamical arguments, how can we choose
among these possibilities?
Cosmochemical data do not provide a definitive
answer, but they at least serve as important constraints
for these hypotheses. Analyses of lunar rocks allow
some general conclusions about the geochemistry of
the bulk Moon. It is strongly depleted, by factors
of several hundred relative to cosmic abundances, in
highly volatile elements (for instance, Bi, Tl) and enriched in refractory elements (such as Ca, Al, Ti,
REEs, U, Th). Its bulk FeO content of 13% is significantly lower than that of C1 chondrites and higher
than that of the Earth. Trace siderophile elements are
depleted in order of the metal-silicate partition coefficients, consistent with their removal into a small
lunar core.
Let’s first list some cosmochemical predictions that
each hypothesis for its origin makes. Intact capture
is largely untestable, but it seems likely that any body
formed elsewhere in the Solar System might have a
distinct composition, both elemental and isotopic,
from the Earth. In contrast, binary accretion suggests
that the compositions of Earth and Moon should be
virtually identical, because they accreted from similar
infalling materials at the same time. Rotational fission
is normally envisioned as occurring after core formation in the Earth. Thus, in this model, the bulk Moon
should be depleted in siderophile elements to the same
degree as the Earth’s mantle. Dynamical simulations
of collisional ejection suggest that the incoming projectile provided most of the ejecta that was accreted
to form the Moon. During the violent collision, most
ejecta was vaporized and later recondensed, allowing
for volatile element depletions in the Moon. Disintegrative capture is a variant of the intact capture hypothesis, from a cosmochemical perspective.
The oxygen isotopic compositions of lunar rocks
lie along the same mass fractionation line as terrestrial rocks, as illustrated in figure 15.12. This is a very
significant observation, because most other extraterrestrial samples define oxygen isotope fractionation
lines displaced from the Earth-Moon line. The stable
isotopic data are most compatible with hypotheses
(2) and (3), although (4) is allowed if the Earth contributes a significant fraction of ejecta to the Moon.
Stretching Our Horizons: Cosmochemistry
The Mg/(Mg + Fe) ratio is an important indicator
of the extent of geochemical differentiation. A value
of 0.89 for this ratio for the Earth’s mantle seems
fairly well constrained, but that for the Moon is less
certain. A ratio of 0.80 is consistent with Fe and Mg
abundances in lunar basalts, which are more iron-rich
than terrestrial basalts. If the Moon really does have
a lower Mg/(Mg + Fe) ratio, this would argue against
its derivation from the Earth by hypothesis (3).
The low uncompressed mean density of the Moon
(3.34 g cm−3 ) is very different from that of the Earth
(4.05 g cm−3 ). Variations in oxidation state cannot
account for density differences of this magnitude, and
it is generally accepted that the Moon is impoverished
in iron metal relative to the Earth. Siderophile trace
elements are depleted in both the Earth’s mantle and
the Moon. The Earth’s situation is easy enough to
understand. Elements such as nickel and cobalt are
now sequestered in the core, but the Moon has only a
small core. Hypotheses (3) and (4) can readily explain
this constraint if fission or collision occurred after
core formation. Preferential accretion of the Moon
from the silicate parts of differentiated planetesimals
could allow hypotheses (2) and (5) to be consistent.
Hypothesis (1) must explain how two independent
result from this approach, because some of these may be
only partly condensed.
Planetary Models: Cores and Mantles
The techniques we have just discussed provide ways
to estimate the bulk chemical compositions of planets. It
is also of obvious value to know how these elements are
distributed within a planet; that is, which are partitioned
into crust, mantle, and core. Because the crust contains
such a minor portion of a planet’s mass, it is common
practice in constructing planetary models to combine
mantle plus crust as one differentiated component. The
problem then is reduced to that of how to estimate the
compositions of the silicate part of the planet and its
metallic core.
Worked problem 12.1 illustrated how the composition
of the Earth’s core can be estimated, based on chondritic
relative abundances.
339
bodies with such different bulk iron contents have such
similar siderophile trace element abundances.
Volatile trace elements are more depleted, and refractory elements more enriched, in the Moon than in
the Earth. This compositional distinction could be
an inherent property of the materials that accreted to
form these bodies, or it could result from some later,
high-temperature process. This observation is difficult to explain if hypothesis (2) is correct. Volatile
element differences are expected for hypothesis (1) and
(5). Most ejecta formed during collision would be vaporized, and the observed volatile element depletion
might result from recondensation in hypothesis (3).
Because a rapidly rotating Earth would probably be
partially molten, hypothesis (4) may also be consistent with this constraint.
These arguments obviously do not solve the problem, but they illustrate how cosmochemical data can
be used to test possible solutions for complex problems of vast scale. The chemical data, plus the observation that the Earth-Moon system has very high
angular momentum, suggest to many workers that a
collisional ejection model is likely. Cosmochemical evidence bearing on the question of the Moon’s origin
has been explained more fully by Ross Taylor (1987).
In chapter 12, we considered constraints on the composition of the Earth’s crust, mantle, and core. Because
the constraints are so meager, the problem is more difficult for other planets. Let’s now consider the mantle and
core of Mars, to illustrate how this problem is addressed
for other planets.
Core composition can be estimated indirectly if the
compositions of the bulk planet and mantle are known.
The core composition is presumably complementary to
that of the mantle from which it was extracted, so that
depletions of siderophile and chalcophile elements relative to the bulk planet abundance can provide insight
into the relative abundances of core-forming elements.
This approach was used by Wänke and Dreibus (1988)
to estimate that the core of Mars contains 61% FeNi
metal and 39% FeS.
Connie Bertka and Yingwei Fei (1998) considered a
variety of possible Martian core compositions, including iron as metal, sulfide, oxide, carbide, and hydride.
340
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
They also calculated the densities of these model cores,
as shown in figure 15.17. The previous year, they performed high-pressure experiments, using the WänkeDreibus Mars composition, to determine the mineralogy
of the Martian mantle as a function of depth. These data
allowed Bertka and Fei to estimate a density profile for
the mantle, also shown in figure 15.17. By assuming a
50 km-thick crust with the density of basalt, they then
calculated the mean density and moment of inertia.
Comparison with measured values suggested some inconsistencies with this Mars model. Although a firm estimate of the composition of Mars that meets all geophysical constraints is not yet available, this illustrates
how cosmochemical models for planetary mantles and
cores can be constructed and tested.
SUMMARY
Cosmochemistry is the application of geochemical principles to systems of vast scale. Fundamental to this discipline is the determination of the cosmic abundance of
the elements. At the beginning of this chapter, we saw
how elemental abundance patterns were controlled by
nucleosynthesis in stars. Mixture of the products of fusion, nuclear statistical equilibrium, neutron capture, and
proton capture reactions in earlier generations of stars
accounts for the composition of the Sun and thus the
cosmic abundance of the elements.
Chondritic meteorites are materials surviving from the
earliest stages of solar system history. Their chemical
compositions are essentially cosmic, although small fractionations occur in all chondrite classes. Element fractionations in the solar nebula, inferred from chondrites, were
governed by volatility and by geochemical affinity, the
latter reflected in siderophile, chalcophile, and lithophile
behavior.
The equilibrium condensation sequence provides a
conceptual framework, in which we can examine the
effects of volatility. We have also seen that the isotopic
compositions of the most refractory components of chondrites point to a supernova origin. The infusion of now
extinct radionuclides indicates that a stellar explosion
must have occurred in the vicinity of the solar nebula. It
has been proposed that the shock wave from such a supernova triggered the collapse of gas and dust to form the
solar system. Relict stardust, recognized by its isotopic
fingerprints, has now been isolated from chondrites.
FIG. 15.17. Density profile for the Martian mantle and core
(heavy line) and stability fields for mantle minerals, based on
experiments using the Wänke-Dreibus Mars model by Bertka and
Fei (1998). Also shown are density profiles and positions of the
core-mantle boundary for a range of model core compositions
(thin lines).
Some of the most volatile materials in chondrites are
organic compounds, synthesized in cold molecular clouds
and often modified by catalytic reactions in the nebula.
Among these are most of the raw materials for life,
suggesting that Earth may not have had to synthesize organic compounds from scratch. Ices were the last materials to condense in the nebula. Highly volatile materials
are concentrated in the outer solar system, primarily in
the Jovian planets and in comets.
The bulk compositions of the terrestrial planets can
be modeled based on equilibrium condensation, assuming that a radial temperature and pressure distribution
controlled mineral stability. The heterogeneous accretion
model assumes that materials were aggregated and isolated from the nebula as soon as they condensed. The
basis for a third planetary model is that fractionations in
protoplanetary materials were the same as those observed
in chondrites, so that chondritic comositions can be used
to construct planets. The compositions of planetary
mantles and cores can also be estimated from cosmochemical considerations. That these different techniques
lead to similar results suggests that cosmochemical behavior offers a valid mechanism for understanding problems of planetary scale.
Stretching Our Horizons: Cosmochemistry
suggested readings
Because chondrites provide most of the data for cosmochemistry, much of the literature in this field is embedded in books
and papers on meteorites. Many of the references below are
technically demanding, but they provide a thorough background
in this subject.
Anders, E., and N. Grevesse. 1989. Abundances of the elements:
Meteoritic and solar. Geochimica et Cosmochimica Acta
53:197–214. (An important paper in which cosmic abundances are derived and discussed.)
Grossman, L., and J. W. Larimer. 1974. Early chemical history
of the solar system. Review of Geophysics and Space Physics
12:71–101. (A comprehensive review of condensation and
other nebular processes.)
McSween, H. Y., Jr. 1999. Meteorites and Their Parent Planets,
2nd ed. New York: Cambridge University Press. (A nontechnical introduction to meteorites and what can be inferred
about their parent bodies.)
Newsom, H. E., and J. H. Jones, eds. 1990. Origin of the Earth.
New York: Oxford University Press. (A series of papers
dealing with geochemical and other kinds of constraints on
the origin and early evolution of our planet.)
Suess, H. E. 1987. Chemistry of the Solar System. New York:
Wiley. (A nice summary at a basic level. Chapter 1 gives the
isotopic composition of the elements, and chapter 2 discusses nucleosynthesis.)
Taylor, S. R. 1992. Solar System Evolution: A New Perspective.
Cambridge: Cambridge University Press. (A fascinating
introduction to the solar system, as constrained by cosmochemistry, and arguing for the importance of impacts in its
evolution.)
Wasson, J. T. 1985. Meteorites, Their Record of Early SolarSystem History. New York: Freeman. (A meteorite monograph that is rich in cosmochemical information. Chapters 7
and 8 describe nebula fractionation in detail.)
Additional papers referenced in this chapter are the following:
Allegre, C. J., G. Manhes, and C. Gopel. 1995. The age of the
Earth. Geochimica et Cosmochimica Acta 59:1445–1456.
Barshay, S. S., and J. S. Lewis. 1976. Chemistry of primitive
solar material. Annual Reviews of Astronomy and Astrophysics 14:81–94.
341
Bertka, C. M., and Y. Fei. 1998. Implications of Mars Pathfinder
data for the accretion history of the terrestrial planets. Science 281:1838–1840.
Burbidge, E. M., G. R. Burbidge, W. A. Fowler, and F. Hoyle.
1957. Synthesis of the elements in stars. Reviews of Modern
Physics 29:547–650.
Cameron, A. G. W., and J. W. Truran. 1977. The supernova
trigger for formation of the solar system. Icarus 30:447–
461.
Clayton, R. N., L. Grossman, and T. K. Mayeda. 1973. A component of primitive nuclear composition in carbonaceous
meteorites. Science 182:485–488.
Grossman, L. 1972. Condensation in the primitive solar nebula. Geochimica et Cosmochimica Acta 36:597–619.
JANAF 1971. Thermochemical Tables, 2nd ed. U.S. National
Standards Reference Data Series 37. Washington, D.C.:
National Bureau of Standards.
Lee, T., D. Papanastassiou, and G. J. Wasserburg. 1977. Aluminum-26 in the early solar system: Fossil or fuel? Astrophysical Journal 211:L107–110.
Lewis, J. S. 1972. Low temperature condensation in the solar
nebula. Icarus 16:241–252.
Morgan, J.W., and E. Anders. 1979. Chemical composition of
Mars. Geochimica et Cosmochimica Acta 43:1601–1610.
Taylor, S. R. 1987. The unique lunar composition and its bearing on the origin of the Moon. Geochimica et Cosmochimica Acta 31:1297–1306.
Turekian, K., and S. P. Clark. 1969. Inhomogeneous accretion
of the earth from the primitive solar nebula. Earth and Planetary Science Letters 6:346–348.
Urey, H. C. 1952. The Planets: Their Origin and Development.
New Haven: Yale University Press.
Van Schmus, W. R., and J. A. Wood. 1967. A chemical-petrologic
classification for the chondritic meteorites. Geochimica et
Cosmochimica Acta 31:747–765.
Wänke H., and G. Dreibus. 1988. Chemical composition and
accretion history of terrestrial planets. Philosophical Transactions of the Royal Society (London) A 325:545–557.
Wood, J. A., and S. Chang, eds. 1985. The Cosmic History of
the Biogenic Elements and Compounds. Washington, D.C.:
NASA.
Yin, Q., S. B. Jacobsen, K. Yamashita, J. Blichert-Toft, P. Telouk,
and F. Albarede. 2002. A short timescale for terrestrial
planet formation from Hf-W chronometry of meteorites.
Nature 418:949–952.
342
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
PROBLEMS
(15.1) Discuss why the comparison of terrestrial Nd-Sm isotopic data with a chondritic evolution curve
(εNd versus ε Sr) provides useful information.
(15.2) Using figure 15.10 and the cosmic abundances in table 15.1, estimate the bulk composition of the
refractory nebular condensate that forms in the temperature interval from 1900 to 1400 K.
(15.3) What are the relative weight proportions of condensable matter (rock plus ices) versus noncondensable gases (He and leftover H) in a gas of solar composition? (Part of the solution has
already been calculated in worked problem 15.3.)
(15.4) Calculate the composition of the Earth’s core in terms of Fe, Ni, Co, P, S, and O, assuming that half
of the light element is sulfur and half is oxygen. (See worked problem 12.1.)
(15.5) List the chemical fractionations observed in chondrites, and suggest a plausible physical mechanism
for each.
(15.6) Using figure 15.12, calculate the proportion of an exotic oxygen component consisting of pure 16O
that would have to be added to “normal” solar system oxygen (lying along the terrestrial mass fractionation line) to produce a CAI plotting along the mixing line at δ17O = −25 per mil.
APPENDICES
APPENDIX A: MATHEMATICAL METHODS
Most of the problems in this book require mathematics
that is no more advanced than standard algebra or firstyear calculus. We have found, however, that students
often need a refresher in some of these basics to boost
their confidence. We strongly urge you to use the problems in this text as an excuse to dust off your old math
books and get some practice. If you do not already own
a handbook of standard functions or a guide to mathematical methods in the sciences, you should probably
add one to your shelf. We have found Burington (1973),
Dence (1975), Boas (1984), and Potter and Goldberg
(1997) to be particularly useful.
This short appendix is intended to introduce a few
concepts not generally included in math courses required
for the geology curriculum, but that are very useful in
geochemistry. We do not pretend that it is an in-depth
presentation; our goal is purely pragmatic. Each of these
topics appears in one or more of the problems in this
text, and should become familiar as you advance in
geochemistry.
Partial Differentiation
If a function f has values that depend on only one
physical parameter, x, then you know from your first
calculus course that the derivative, df(x)/dx, represents
the rate of change of f(x) with respect to x. In graphical
terms, df(x)/dx (if it exists in the range of interest) is the
instantaneous slope of the curve defined by y = f(x).
Strictly, df(x)/dx is given by:
df(x)/dx = lim h→0 [(f(x + h) − f(x))/h].
Equations involving rates of change in the physical
world are extremely common, but it is rare to find one
in which the value of the function f depends on only one
parameter. More commonly, we deal with f(x, y, z, . . . ).
Still, it is useful to know how the function f varies if we
change the value of only one of its controlling parameters
at a time. For a function f(x, y), for example, we define
partial derivatives (∂f(x, y)/∂x)y and (∂f(x, y)/∂y)x by:
(∂f(x, y)/∂x)y = lim h→0 [(f(x + h, y) − f(x, y))/h],
and
(∂f(x, y)/∂y)x = lim h→0 [(f(x, y + h) − f(x, y))/h].
When we write partial derivatives, we use the Greek
symbol ∂, rather than d, to remind ourselves that f is a
function of more than one variable, and we indicate the
variables that are being held fixed by means of subscripts. In chapter 3, for example, we find that the Gibbs
free energy (G) of a phase is a function of temperature
(T), pressure (P), and the number of moles of each of the
components (n1, n2, . . . , ni) in the phase. To consider
how G changes as a function of T alone, we look for the
value of (∂G/∂T )P,n1,n2, . . . , ni.
Because there can be many interrelationships among
variables that describe a system, we can write many different partial derivatives. If pressure, temperature, and
volume all influence the state of a system, for example,
we may be interested in (∂P/∂T )V, (∂V/∂P)T, (∂T/∂V)P, or
any of their reciprocals. These various partial derivatives
are not independent of each other. This should be evident from our discussion of the Maxwell relations in
chapter 3, in which we investigated some of the relationships among partial derivatives of thermodynamic
functions. If that discussion was unfamiliar ground for
you, try learning four simple rules for working with partial derivatives. These follow from the standard Chain
Rule, which states that if w = f(x, y, z) and if x, y, and z
are each functions of u and v, then:
(∂w/∂u)v = (∂w/∂x)y,z(∂x/∂u)v + (∂w/∂y)x,z(∂y/∂u)v
+ (∂w/∂z)x,y(∂z/∂u)v ,
(A.1)
343
344
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
and
(∂w/∂v)u = (∂w/∂x)y,z(∂x/∂v)u + (∂w/∂y)x,z(∂y/∂v)u
+ (∂w/∂z)x,y(∂z/∂v)u.
(A.2)
The symmetry of the Chain Rule makes it easy to remember. Consider, now, what would happen if we applied the
Chain Rule to a function w = f(x, y), in which x was
itself a function of y and another variable, z; that is, x =
g(y, z). Then:
(∂w/∂z)y = (∂w/∂x)y (∂x/∂z)y + (∂w/∂y)x (∂y/∂z)y ,
but
(∂y/∂z)y = 0,
so that:
Rule 1: (∂w/∂x)y (∂x/∂z)y = (∂w/∂z)y.
(A.3)
It can also be shown that:
Rule 2: (∂w/∂x)y = 1/(∂x/∂w)y,
(A.4)
Rule 3: (∂x/∂y)z(∂y/∂z)x(∂z/∂x)y = −1,
(A.5)
and
Rule 4: (∂w/∂y)x = (∂w/∂y)z + (∂w/∂z)y (∂z/∂y)x .
(A.6)
We commonly describe changes in some property of
a system by writing differential equations, in which an
infinitesimal change df is related to infinitesimal changes
in one or more of the variables that control the property.
For a function of a single variable, f(x), we can write:
df = (df/dx)dx.
If more than one variable (x, y, z, . . . ) can potentially influence the value of f, the equivalent expression is:
df = (∂f/∂x)y,z, . . . dx + (∂f/∂y)x,z, . . . dy
+ (∂f/∂z)x,y . . . dz + . . . .
The expression on the right side of this equation is called
a total differential, because it accounts for the total
change df by summing the changes due to each independent variable separately. Consider land elevation, for
example, which varies as a function of both latitude and
longitude on the Earth’s surface. If you were to move a
short distance in some random direction, your total
change in elevation would be equal to the slope of the
land surface in a north-south direction times the distance
you actually moved in that direction, plus the slope in an
east-west direction times the distance you moved in that
direction.
In many geochemical applications, we find ourselves
working with a special type of total differential called an
exact or perfect differential. The differentials of each of
the energy functions (dE, dH, dF, dG) introduced in chapter 3, for example, belong in this category. To define what
we mean, suppose that we know of three properties of a
system, each of which is a function of the same set of
three independent variables. Identify these properties
as P(x, y, z), Q(x, y, z), and R(x, y, z). Can these three
properties be related by some function f(x, y, z) in such
a way that df = Pdx + Qdy + Rdz? If so, then the expression on the right side is not only the total differential of f, it is also an exact differential. We then say that
f is a function of state, with the following particularly
useful characteristics:
1. Not only are P, Q, and R the partial derivatives of f
with respect to x, y, and z, but their own partial derivatives (second derivatives of f ) are also interrelated.
The following cross-partial reciprocity expressions are
true:
(∂P/∂y)x,z = (∂Q/∂x)y,z ; (∂P/∂z)x,y = (∂R/∂x)y,z ;
(∂Q/∂z)x,y = (∂R/∂y)x,z.
This property is evident in the Maxwell relations,
which derive from the fact that each of the energy
functions (E, F, H, and G) is function of state. For
example, G = G(P, T, n) = −SdT + VdP + µdn. S(P, T,
n), V(P, T, n), and µ(P, T, n), therefore, are like the
functions P(x, y, z), Q(x, y, z), and R(x, y, z). Each
is a partial differential of G (that is, S = −(∂G/∂T )P,n ,
V = (∂G/∂P)T,n, and µ = (∂G/∂n)T,P) and the Maxwell
relations in this case,
−(∂S/∂P)T,n = (∂V/∂T )P,n = (∂µ/∂n)P,T ,
are the cross-partial reciprocity expressions we expect,
because G is an exact differential.
2. We can determine the total change in f between two
states of the system by integrating df between (x1, y1,
z1) and (x2, y2, z2) along some reaction pathway, C.
An integral of this type is known as a line integral.
Appendices
345
For an illustration of how this operation is performed,
see worked problem 3.1. When we do this integration
for a function of state, we discover that:
x2,y2,z2
养
x1,y1, z1
df =
x2,y2,z2
养
x1,y1, z1
(Pdx + Qdy + Rdz)
= f(x2, y2, z2) − f(x1, y1, z1).
Put simply, this means that we always get the same answer, regardless of how we get from state 1 to state 2.
The integral of df is said to be independent of path.
3. Path independence implies that if we integrate df
around a closed loop from state 1 to state 2 and back
again, the net change in f will be zero. Again, see
worked problem 3.1 to verify that the change in internal energy (an exact differential) around a closed
path is zero, whereas the change in either work or
heat is not.
Root Finding
It is easy to find the roots of sets of first-order equations by simple algebraic substitution. It is very common, however, to encounter problems in the physical
sciences that yield equations in which a key variable x is
raised to a higher power or appears in a transcendental
function like sin x. A familiar equation from which you
learned to extract x when you were in high school is the
polynomial:
0 = ax2 + bx + c.
In this example, x can be determined by applying the
quadratic formula:
)/2a.
x = (−b ± √
(b2 −4ac)
Unfortunately, there are no simple solutions of this type
for most equations you will encounter. Adding a cubic
term to this polynomial, for example, would make the
task of root finding considerably more difficult. There
are many ways around this problem, most of which involve a numerical approach instead of a closed-form
or analytical one. They are thus well suited to analysis
by computer or a pocket calculator. We briefly discuss
one of these approaches, the Newton-Raphson method,
which is mentioned as a means for solving several problems in chapters 4, 7, and 14.
Consider the curve for the function y = f(x) in figure A.1. Suppose that f(x) has a real root at x = a0 that
we want to find. One way to do this would be simply to
FIG. A.1. Geometric basis for the Newton-Raphson method for
finding a root of f(x). Line BP is the tangent to f(x) at point P.
guess at random successive values of x. If we are patient
and watch to see that f(x) is closer to zero with each guess,
we will eventually find the value for which f(x) = 0. By
examining figure A.1, however, we can see a way to make
guesses in a more sophisticated way. Suppose that our
first guess is that x = a1, reasonably close to a0 but not
correct. The error in this guess is |a1 − a0 |, shown by the
distance AC in the figure. Assuming that this error is unacceptably large, we now make a better guess by drawing the tangent BP to the curve at point P, and see that
the tangent crosses the x axis at x = a2. The error is now
only |a2 − a0 |, or the distance AB. We repeat the process,
each time estimating the improved value of ai+l by drawing a tangent to y = f(x) at the point corresponding to ai,
until the error |ai − a0 | is acceptably small.
We can write an algorithm to express this method in
a way that can be used in a computer program. Comparing the first two guesses geometrically, we see that:
AB = AC − BC,
or
(a2 − a0) = (a1 − a0) − PC/tan θ,
from which a2 = a1 − f(a1)/f ′(a1), where f ′(al) is the instantaneous slope of f(x) at a1 (that is, f ′(a1) = dy/dx at
a1). This simple rule gives us a powerful means for
making consecutive guesses, provided that (1) f(x) has a
346
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
derivative in the vicinity of a0, (2) a1 is reasonably close
to a0, so the method doesn’t converge on some other
root, and (3) f ′(ai+1) is not very close to zero, so that we
avoid introducing a large numerical error in ai+1. The
algorithm, then, is:
1. Make an initial guess, a1, of the root.
2. Calculate f(a1) and f ′(a1).
3. Calculate a better value for the root, a2 = a1 − f ′(a1)/
f(a1).
4. Calculate f(a2).
5. If f(a2) is acceptably close to zero, stop. Otherwise,
repeat steps (2)−(5) until the root is acceptably close.
You may define “acceptably close” with a specific numerical value, or by noting when f(ai) does not change much
from one iteration to the next.
Most texts that discuss the Newton-Raphson method
fail to point out how easily it can be applied to systems
of equations. Because geochemical problems commonly
involve functions of more than one variable, you may
find it useful to know how this is done. Suppose that
several variables (x1, x2, . . . , xn), related by a set of n
independent equations, describe a system. To make life
frustrating, each of the equations is a complicated function in which the variables appear in several terms, so
that finding the values of x1 − xn by substitution or
simple matrix methods is impossible. If each of the functions can be differentiated with respect to each of the
variables, however, there is hope for use of the NewtonRaphson method.
As in the simple case above, begin by guessing initial
values for x1 through xn. Using these, calculate the numerical values of each of the functions (fl, f2, . . . , fn) and
place them in a column vector F. Also, calculate each of
the partial derivatives (∂f/∂x) and build a matrix J that
looks like this:
(
(∂fl /∂x1)
(∂fl /∂x2)
...
(∂fl /∂xn)
(∂f2 /∂x1) (∂f2 /∂x2) . . . (∂f2 /∂xn)
J=
(∂f3 /∂x1) (∂f3 /∂x2) . . . (∂f3 /∂xn)
.
.
.
(∂fn /∂x1)
.
.
.
.
.
.
(∂fn /∂x2) . . .
.
.
.
(∂fn /∂xn)
)
.
The column vector X = (x1, x2, . . . , xn) can then be
found by successive approximations from the matrix ver-
sion of the Newton-Raphson formula, which we present
without proof:
Xk+1 = Xk − (Jk)−1 Fk.
This is too cumbersome for a pocket calculator, because it involves inverting the matrix J at each iteration,
but it is quite easily done with standard software, such
as Microsoft Excel™. An algorithm based on this matrix
method tests Fk at each step to see whether each of its
elements is acceptably close to zero.
Fitting a Function to Data
In a typical research problem, a geochemist gathers
observations on a natural or experimental system and
then tries to make sense of the results by looking for
functional relationships between the data and system
variables that control them. Sometimes the function that
best describes the data has a form based on theoretical
considerations. At other times, you may choose an empirical function—a polynomial or exponential equation,
for example. In any case, the task of fitting that function
to your data set involves finding the values of one or
more coefficients in the equation so that it gives results
that are as close to the observed data as possible.
We demonstrate a common approach to the problem
by considering the case in which measured values of some
property y appear to depend on some other property x
in such a way that a graph of y against x is a straight line.
That is, y = f(x) has the form:
y = a0 + a1 x.
What values of a0 and a1 are most appropriate for your
data? If all of the observed values of y lie precisely on a
line, the answer would be easy to find. Unfortunately,
however, there are always random errors in any data set,
due to sampling technique or undiagnosed complexities
in the system (fig. A.2.) The task, then, is to find values
of a0 and a1 such that the distance between f(x) and each
of the measured values of y is as small as possible. In
practice, we generally go one step further and look for a0
and a1 such that the sum of the squares of the deviations
from f(x) is minimized. This overall approach, therefore,
is known as the method of least squares.
Mathematically, the problem involves minimizing the
function:
Q(a0, a1) =
Σ (y − a
i
i
0
− a1xi)2,
Appendices
FIG. A.2. The measured values of property x (open circles), in
this example, are apparently related to values of property y by
some function y = f(x). Because of random errors in the data set,
however, the observed values are scattered around the most probable linear function (line).
or, if we have some reason to trust some observations
more than others,
Q(a0, a1) =
Σ w (y − a
i
i
i
0
− a1xi)2,
where wi is a weighting factor for observation i. The
summations are each taken over all values of i from 1 to
N, where N is the total number of observations. To
minimize Q, we first find expressions for the two partial
derivatives of Q with respect to a0 and a1 and set each
equal to zero:
(∂Q/∂a0)a1 = 2( wi yi − a0 wi − a1 wi xi) = 0,
Σ
Σ
Σ
and
(∂Q/∂a1)a0 = 2( wi xi yi − a0 wi xi − a1 wi xi2)
= 0.
Σ
Σ
Σ
These can be solved simultaneously to yield:
a0 = [( wi xi2)( wi yi) − ( wi xi)( wi xi yi)]/
Σ
Σ
Σ
[(Σ wi)(Σ wi xi2) − (Σ wi xi)2],
Σ
a1 = [( wi)( wi xi yi) − ( wi xi)( wi yi)]/
Σ Σ
Σ
Σ
w
w
x
−
w
x
.
[(Σ )(Σ ) (Σ ) ]
i
2
i i
2
i i
The same procedure can be followed for any polynomial
expression.
The Error Function
You are probably somewhat familiar with the func−t2 /2, where t = (x − x̄)/σ. This is the
tion f(x) = (1/√
[2π])e
347
standard form of the normal frequency function, which
describes the frequency distribution of events (x) around
a mean, x̄. The quantity σ is the standard deviation of x,
a measure of the spread of values around the mean. This
is the “curve” that teachers once used for calculating the
grade distribution in classes. It describes the expected
distribution of random variations (“errors”) in many
other natural situations. In chapter 5 and elsewhere, we
have used it indirectly as a means of finding the distribution of mobile ions diffusing across a boundary between
adjacent phases.
It can be shown that the area under this curve between
zero and some penetration distance x is equal to the
probability that an ion lies within that range. This, in
fact, is the context in which the normal distribution appears in chapter 5. The error function, erf(x), can be
defined as:
erf(x) = (1/√
[2π])
x
∫e
0
−t2/2 dt,
and is used to calculate this probability. This form of the
error function is commonly used by statisticians. Unfortunately, several other forms are also used by physical
scientists. The definition most commonly found in discussions of transport equations is the one we have applied
in chapter 5:
erf(x) = (2/√
π)
x
∫e
0
−t2
dt.
It is a simple matter to relate one form of the error function to another by making an appropriate substitution
for t and adjusting the integration limits accordingly.
Because several subtly different versions exist, you should
always check to see how erf(x) is defined for the particular application you have in mind.
Unfortunately, the integration of e −t2 cannot be performed analytically (that is, in neat, closed form). Because erf(x) is a widely used function, however, there
are many popular ways to evaluate it. The easiest is
simply to look it up in a table. The error function and
its complement, erfc(x) = 1 − erf(x), are tabulated in
most standard volumes of mathematical functions (for
example, Burington 1973). There are also many approximate solutions that are reliable for calculations
with pencil and paper. One that yields <0.7% error is
the function erf(x) = √[1 − exp(−4x2/π)]/2. In most
applications today, however, erf(x) is calculated by numerical integration.
348
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
references
Boas, M. L. 1984. Mathematical Methods in the Physical Sciences: Solutions of Selected Problems, 2nd ed. New York:
Wiley.
Burington, R. S. 1973. Handbook of Mathematical Tables and
Formulas, 5th ed. New York: McGraw-Hill.
Dence, J. B. 1975. Mathematical Techniques in Chemistry. New
York: Wiley.
Potter, M. C., and J. Goldberg. 1997. Mathematical Methods
for Engineers, 2nd ed. Wildwood: Great Lakes Press.
APPENDIX B: FINDING AND
EVALUATING GEOCHEMICAL DATA
Selected Data Sources
The data of geochemical interest are scattered throughout a large number of journals and research reports that
focus on the results of narrowly defined studies. In most
cases, a search for data on specific systems should begin
with the major abstracting journals, Chemical Abstracts
and Mineralogical Abstracts, which regularly review the
major journals and provide a topical summary of their
contents. Limited literature surveys, such as the excellent
summary of thermodynamic data sources in Nordstrom
and Muñoz (1986), are also useful guides to the primary
literature. Rather than provide another survey of this
type, here we have compiled a short list of general references that are widely used by geochemists. In almost all
cases, these are tabulations or databases built of information gathered from the primary research literature.
They offer the advantage of being a quick overview of
values that might otherwise be hard to find except in
obscure corners of the library. More importantly, a large
number of these (but not all!) are critical evaluations of
the data, checked for internal consistency.
Using the Data
Before you use the data in any tabulation, be sure to
read the text material that accompanies it, describing the
criteria for data selection, the reactions that were investigated experimentally to obtain the data, the choice of
standard and reference states, and the methods that were
used to calculate interpolated, or “derived,” values. Because data are gathered for many purposes, these criteria
may not be the same from one source to another. Also,
tests for the internal consistency of data compilations may
be applied in different ways by various teams of compilers. Therefore, you should not assume that you can
safely mix values garnered from different sources in a
single problem. For a very good discussion of the potential problems of dealing with multiple data sources,
consult chapter 12 of Nordstrom and Muñoz (1986).
Despite these problems, it is usually advisable and
often necessary to consult several sources as you compile
data for use in your calculations. You will find that teams
of researchers have made idiosyncratic choices that
emphasize different types of information for the same
substances. The JANAF tables (Stull and Prophet 1971
et seq.), for example, indicate the temperatures at which
phase changes take place in the substance being reported,
but do not indicate phase changes in the reference state
elements used to determine ∆H̄f0 and ∆Ḡf0. Robie and
coworkers (1978) report both. In addition, Robie and coworkers provide pairs of tables for silicate minerals that
report free energies and enthalpies of formation from
both the elements and from the oxides. The JANAF
tables, however, include a short summary of the data
sources consulted during compilation and, in most cases,
a critical justification for the particular choices that were
made in preparing their tables. This is a useful feature
rarely found in other compendia. Helgeson and colleagues (1978) carry this approach to an extreme by
producing a document in which evaluation of the data is
the major concern. The data tables are only a few pages
of the text and are in a compressed form. It is left to the
reader, for example, to calculate thermodynamic values
above 298 K by laboriously integrating heat capacity
power functions.
Most of the data in standard tabulations should be
familiar to students who have read chapters 3, 4, and 9
of this text. The lone exception is the free energy func0
)/T. Values of this function
tion, defined as (∆Ḡ0 − ∆H̄ 298
do not change very much with temperature, so that it is
usually very safe to interpolate linearly to find values
between tabulated temperatures. For reactions among
various substances, then, it can be shown that:
0 )/T = ∆Ḡ 0/T − ∆H̄ 0
(∆Ḡ0 − ∆H̄298
r
r,298 /T,
so that
0 = T([∆Ḡ 0 − ∆H̄ 0 ]/T) + ∆H̄ 0
∆Ḡr,T
298
r
r,298 ,
Appendices
where we have followed our convention that quantities
with the subscript r are the stoichiometric sum of values
for the product phases minus the reactant phases. When
the free energy function is available, it serves as a labor0 from heat casaving alternative to calculating ∆Ḡr,T
pacity data and reference state values for enthalpy and
entropy.
For those instances when it is preferable to use heat
capacity data, it is convenient to have them in the form
of a power law. A commonly used expression is:
CP = R(a + bT + cT 2 + dT 3 + eT 4 ),
in which R (= 1.98726 cal mole−1 K−1) is the gas constant.
Robie and coworkers (1978), for example, tabulate values
of the constants a, b, c, d, and e for each of the substances in their compendium.
general reference
Carmichael, I.S.E., and H. P. Eugster, eds. 1987. Thermodynamic Modeling of Geological Materials: Minerals, Fluids
and Melts. Reviews in Mineralogy 17. Washington, D.C.:
Mineralogical Society of America.
DeBievre, P., M Gallet, N. E. Holden, and I. L. Barnes. 1984.
Isotopic abundances and atomic weights of the elements.
Journal of Physical and Chemical Reference Data 13:809–
891.
Greenwood, N. N., and A. Earnshaw. 1997. Chemistry of the
Elements, 2nd ed. Oxford: Pergamon.
Li, Y.-H. 2000. A Compendium of Geochemistry. Princeton:
Princeton University Press.
Lide, D. R., ed. 2001. CRC Handbook of Chemistry and
Physics, 81st ed. Boca Raton: CRC.
Nordstrom, D. K., and J. L. Muñoz. 1986. Geochemical Thermodynamics. Malden: Blackwell.
Ronov, A. B., and A. A. Yaroshevsky. 1969. Chemical composition of the earth’s crust. In P. J. Hart, ed., The Earth’s
Crust and Upper Mantle. American Geophysical Union
Monograph 13. Washington, D.C.: American Geophysical
Union, pp. 35–57.
Wedepohl, K. H., ed. 1969. Handbook of Geochemistry, 2 vols.
New York: Springer-Verlag.
elements and inorganic compounds
Chase, M. W., Jr. 1998. NIST-JANAF Thermochemical Tables,
4th ed. Monograph 9. Journal of Physical and Chemical
Reference Data. Washington, D.C.: National Institute of
Standards and Technology.
349
CODATA. 1978. CODATA Recommended Key Values for
Thermodynamics 1977. CODATA Bulletin 28. Oxford:
Pergamon.
CODATA. 1987. 1986 Adjustment of the Fundamental Physical Constants. CODATA Bulletin 63. Oxford: Pergamon.
Cox, J. D., D. D. Wagman, and V. A. Medvedev, eds. 1989.
CODATA Key Values for Thermodynamics. New York:
Hemisphere.
Hultgren, R., P. D. Desai, D. T. Hawkins, M. Gleiser, K. K.
Kelly, and D. D. Wagman. 1973. Selected Values of the
Thermodynamic Properties of Binary Alloys. Metals Park:
American Society for Metals.
Merrill, L. 1982. Behavior of the AB2-type compounds at high
pressures and high temperatures. Journal of Physical and
Chemical Reference Data 11:1005–1064.
Naumov, G. B., B. N. Ryzhenko, and I. L. Khodakovskii. 1974.
Handbook of Thermodynamic Data. NTIS Document Pb226, 722/7CxA. Washington, D.C.: U.S. Department of
Commerce.
Parker, V. B., D. D Wagman, and W. H. Evans. 1971. Selected
Values of Chemical Thermodynamic Properties: Tables for
the Alkaline Earth Elements (Elements 92 through 97 in the
Standard Order of Arrangement). U.S. National Bureau of
Standards Technical Note 270-6. Washington, D.C.: U.S.
Department of Commerce.
Schumm, R. H., D. D. Wagman, S. Bailey, W. H. Evans, and
V. B. Parker 1973. Selected Values of Chemical Thermodynamic Properties: Tables for the Lanthanide (Rare Earth)
Elements (Elements 62 through 76 in the Standard Order
of Arrangement). U.S. National Bureau of Standards Technical Note 270-7. Washington, D.C.: U.S. Department of
Commerce.
Stull, D. R., and H. Prophet. 1971. JANAF Thermochemical
Tables. U.S. National Bureau of Standards NSRDS-NBS 37.
Washington, D.C.: U.S. Department of Commerce.
Supplement 1974 by Chase, M. W., Jr., J. L. Curnutt, A. T.
Hu, H. Prophet, and L. C. Walker. Journal of Physical
and Chemical Reference Data 3:311–480.
Supplement 1975 by Chase, M. W., Jr., J. L. Curnutt, H.
Prophet, R. A. McDonald, and A. N. Syverud. Journal
of Physical and Chemical Reference Data 4:1–175.
Supplement 1978 by Chase, M. W., Jr., J. L. Curnutt, R. A.
McDonald, and A. N. Syverud. Journal of Physical and
Chemical Reference Data 7:793–940.
Supplement 1982 by Chase, M. W., Jr., J. L. Curnutt, J. R.
Downey Jr., R. A. McDonald, A. N. Syverud, and E. A.
Valenzuela. Journal of Physical and Chemical Reference
Data 11:695–940.
Wagman, D. D., W. H. Evans, V. B. Parker, I. Halow, S. M. Bailey, and R. H. Schumm. 1968. Selected Values of Chemical
Thermodynamic Properties: Tables for the First Thirty-Four
Elements in the Standard Order of Arrangement. U.S. Na-
350
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
tional Bureau of Standards Technical Note 270-3. Washington, D.C.: U.S. Department of Commerce.
Wagman, D. D., W. H. Evans, V. B. Parker, I. Halow, S. M.
Bailey, and R. H. Schumm. 1969. Selected Values of Chemical Thermodynamic Properties: Tables for Elements 35
through 53 in the Standard Order of Arrangement. U.S.
National Bureau of Standards Technical Note 270-4. Washington, D.C.: U.S. Department of Commerce.
Wagman, D. D., W. H. Evans, V. B. Parker, I. Halow, S. M.
Bailey, R. H. Schumm, and K. L. Churney. 1971. Selected
Values of Chemical Thermodynamic Properties: Tables for
Elements 54 through 61 in the Standard Order of Arrangement. U.S. National Bureau of Standards Technical Note
270-5. Washington, D.C.: U.S. Department of Commerce.
Wagman, D. D., W. H. Evans, V. B. Parker, and R. H. Schumm.
1976. Chemical Thermodynamic Properties of Sodium,
Potassium, and Rubidium: An Interim Tabulation of Selected Material. U.S. National Bureau of Standards Interim
Report NBSIR 76-1034. Washington, D.C.: U.S. Department of Commerce.
Wagman, D. D., W. H. Evans, V. B. Parker, R. H. Schumm, and
R. L. Nuttall. 1981. Selected Values of Chemical Thermodynamic Properties: Compounds of Uranium, Protactinium,
Thorium, Actinium, and the Alkaline Metals. U.S. National
Bureau of Standards Technical Note 270-8. Washington,
D.C.: U.S. Department of Commerce.
Wagman, D. D., W. H. Evans, V. B. Parker, R. H. Schumm, I.
Halow, S. M. Bailey, K. L. Churney, and R. L. Nuttall. 1982.
The NBS tables of chemical thermodynamic properties:
Selected values for inorganic and C1 and C2 organic substances in SI units. Journal of Physical and Chemical Reference Data 11(Suppl. 2):1–392.
minerals
Haas, J. L., Jr., G. R. Robinson, and B. S. Hemingway. 1981.
Thermodynamic tabulations for selected phases in the system CaO-Al2O3-SiO2-H2O at 101.325 kPa (1 atm) between
273.15 and 1800 K. Journal of Physical and Chemical Reference Data 10:575–669.
Helgeson, H. C., J. M. Delany, H. W. Nesbitt, and D. K. Bird.
1978. Summary and critique of the thermodynamic properties of rock-forming minerals. American Journal of Science
278A:1–229.
Robie, R. A., B. S. Hemingway, and J. R. Fisher. 1978. Thermodynamic Properties of Minerals and Related Substances
at 298.15 K and One Atmosphere Pressure and at Higher
Temperatures. U.S. Geological Survey Bulletin 1259. Washington, D.C.: U.S. Geological Survey.
Robie, R. A., B. S. Hemingway, and H. T. Haselton. 1983.
Thermodynamic Properties of Minerals. U.S. Geological
Survey Professional Paper 1375. Washington, D.C.: U.S.
Geological Survey.
Robinson, G. R., J. L. Haas Jr., C. M. Schafer, and H. T.
Hazelton Jr. 1982. Thermodynamic and Thermophysical
Properties of Selected Phases in the MgO-SiO2-H2OCO2, CaO-Al2O3-SiO2-H2O-CO2, and Fe-FeO-Fe2O3-SiO2
Chemical Systems, with Special Emphasis on the Properties
of Basalts and Their Mineral Components. U.S. Geological
Survey Open-File Report 83-79. Washington, D.C.: U.S.
Geological Survey.
Woods, T. L., and R. M. Garrells. 1987. Thermodynamic Values
at Low Temperature for Natural Inorganic Materials. New
York: Oxford University Press.
aqueous species
Burnham, C. W., J. R. Holoway, and N. F. Davis. 1969. Thermodynamic Properties of Water to 1000°C and 10,000 bars.
Geological Society of America Special Paper 132. Boulder:
Geological Society of America.
Criss, C. M., and J. W. Cobble. 1964. The thermodynamic properties of high temperature aqueous solutions. IV. Entropies
of the ions up to 200° and the correspondence principle.
Journal of the American Chemical Society 86:5385–5390.
Criss, C. M., and J. W. Cobble. 1964. The thermodynamic
properties of high temperature aqueous solutions. V. The
calculation of ionic heat capacities up to 200°. Entropies
and heat capacities above 200°. Journal of the American
Chemical Society 86:5390–5393.
Hamer, W. J. 1968. Theoretical Mean Activity Coefficients of
Strong Electrolytes in Aqueous Solutions from 0 to 100°C.
U.S. National Bureau of Standards NSRDS-NBS 24. Washington, D.C.: U.S. Department of Commerce.
Helgeson, H. C. 1967. Thermodynamics of complex dissociation in aqueous solution at elevated temperatures. Journal
of Physical Chemistry 71:3121–3136.
Helgeson, H. C. 1969. Thermodynamics of hydrothermal systems at elevated temperatures and pressure. American Journal of Science 267:729–804.
Helgeson, H. C. 1982. Errata: Thermodynamics of minerals,
reactions, and aqueous solutions at high temperatures and
pressures. American Journal of Science 282:1144–1149.
Helgeson, H. C. 1985. Errata II: Thermodynamics of minerals,
reactions, and aqueous solutions at high pressures and temperatures. American Journal of Science 285:845–855.
Helgeson, H. C., and D. H. Kirkham. 1974. Theoretical prediction of the thermodynamic behavior of aqueous electrolytes
at high pressures and temperatures: I. Summary of the
thermodynamic/electrostatic properties of the solvent. American Journal of Science 274:1089–1198.
Helgeson, H. C., and D. H. Kirkham. 1974. Theoretical prediction of the thermodynamic behavior of aqueous electrolytes at high pressures and temperatures. II. DebyeHückel parameters for activity coefficients and relative
partial molal properties. American Journal of Science 274:
1199–1261.
Appendices
Helgeson, H. C., and D. H. Kirkham. 1976. Theoretical prediction of the thermodynamic behavior of aqueous electrolytes
at high pressures and temperatures. III. Equation of state for
aqueous species at infinite dilution. American Journal of Science 276:97–240.
Helgeson, H. C., D. H. Kirkham, and G. C. Flowers. 1982.
Theoretical prediction of the thermodynamic behavior of
aqueous electrolytes at high pressures and temperatures. IV.
Calculation of activity coefficients, osmotic coefficients, and
partial molal and standard and relative partial molal properties to 600°C and 5 kb. American Journal of Science
281:1249–1516.
Hogfeldt, E. 1982. Stability Constants of Metal Ion Complexes:
Part A. Inorganic Ligands. IUPAC Chemical Data Series
no. 21. Oxford: Pergamon.
Horne, R. A., ed. 1972. Water and Aqueous Solutions: Structure, Thermodynamics, and Transport Properties. New York:
Wiley Interscience.
Sillén, L. G., and A. E. Martell. 1964. Stability Constants of
Metal-Ion Complexes. The Chemical Society Special Publication no. 17. London: The Chemical Society. (Supplement 1971.)
reaction kinetics
Cohen, N., and K. R. Westberg. 1983. Chemical kinetic data
sheets for high-temperature chemical reactions. Journal of
Physical and Chemical Reference Data 12:531–590.
phase relations
9.7 × 1022 kg
6.378139 × 106 m
6.35675 × 106 m
5.1 × 1014 m2
3.62 × 1014 m2
1.48 × 1014 m2
1.37 × 1021 1
3.8 × 103 m
3.6 × 1016 1 yr−1
Physical Constants
Avogadro’s number
Gas constant
Faraday’s constant
Gravitational
constant
Boltzmann’s
constant
Planck’s constant
Base of natural
logarithms
N = 6.022094 × 1013 mol−1
R = 1.98717 cal mol−1 K−1
= 8.31433 J mol−1 K−1
= 82.06 cm3 atm mol−1 K−1
F = 96,487.0 coulomb equiv−1
= 23,060.9 cal volt−1 equiv−1
G = 6.6732 × 10−11 m3 kg−1 sec−2
= 6.6732 × 10−11 nt m2 kg−2
k = 1.380622 × 10−23 J K−1
h = 6.626176 × 10−34 J sec
e = 2.71828
Conversion Factors
Brookins, D. G. 1988. Eh-pH Diagrams for Geochemistry. New
York: Springer-Verlag.
Ehlers, E. G. 1987. The Interpretation of Geological Phase
Diagrams. Mineola: Dover.
Levin, E. M., C. R. Robbins, and H. F. McMurdie. 1964. Phase
Diagrams for Ceramists, vol. 1. Columbus: American Ceramic Society.
APPENDIX C: NUMERICAL VALUES
OF GEOCHEMICAL INTEREST
Dimensions of the Earth
Mass of the Earth
Mass of the atmosphere
Mass of the oceans
Mass of the crust
Mass of the mantle
Mass of the outer core
Mass of the inner core
Equatorial radius
Polar radius
Surface area of the Earth
Surface area of the oceans
Surface area of the continents
Volume of the oceans
Mean depth of the oceans
Continental runoff rate
351
5.973 × 1024 kg
5.1 × 1018 kg
1.4 × 1021 kg
2.6 × 1022 kg
4.0 × 1024 kg
1.85 × 1024 kg
Distance:
1 centimeter (cm)
Time:
1 year
Mass:
1 gram (g)
1 atomic mass unit
(amu)
Temperature:
Kelvins (K)
= 108 Ångström (Å)
= 0.3937 inch
= 3.154 × 107 sec
= 2.20462 × 10−3 lb
= 1.66054 × 10−24 g
= °C + 273.15
= 5(°F − 32)/9 + 273.15
Pressure (mass length−1 time−1):
1 atm
= 1.013250 × 106 dyne cm−2
= 1.013250 bar
= 1.013250 × 105 pascal (Pa)
= 1.013250 × 105 nt m2
= 14.696 lb in−2
352
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Energy (mass length2 time−1):
1 joule (J)
= 107 erg
= 2.389 × 10−1 cal
= 9.868 × 10−1 liter atm
= 6.242 × 1011 eV
= 2.778 × 10−7 kWh
= 9.482 × 10−4 Btu
2
Entropy (mass length time−2 deg−1):
1 Gibbs
= 1 cal K−1
−1
Viscosity (mass length time−1):
1 poise
= 1 g sec−1 cm−1
= 1 dyne sec cm−1
= 0.1 Pa sec
Radioactive decay:
1 Curie (Ci)
= 3.7 × 1010 sec−1
Logarithmic values: ln x = 2.303 log x
Units of concentration in solutions:
Molarity (M)
= moles liter−1 of solution
Normality (N)
= equiv. liter−1 of solution
Molality (m)
= moles kg−1 of H2O*
Parts per million
= g/106 g ( = mg/kg)
(ppm)
*Concentrations are commonly reported in the geochemical
literature as millimoles kg−1 of solution. To convert this measure to molal units requires that the density of the solution be
known. Except in highly concentrated brines, however, little
error is introduced if this difference is ignored.
GLOSSARY
Accretion. Accumulation of materials in space to form
larger objects. This process may occur homogeneously
(producing bodies with uniform composition) or
heterogeneously (producing layered bodies).
Activation energy. The amount of energy that must be
provided to overcome some kinetic barrier.
Activity. Concentration of a component, adjusted for any
effects of nonideality; also, a measurement of the
number of radioactive decay events per unit of time.
Activity coefficient. Ratio of the activity of a species to its
concentration; a measure of the degree of nonideal
behavior of a chemical species in solution.
Adiabatic process. A process that occurs without exchange
of heat with the surroundings.
Advection. Transport of ions or molecules within a moving medium.
Alkalinity. Charge deficit between the sum of dissolved
conservative cations and anions in an electrolyte
solution.
Alkane. Saturated aliphatic hydrocarbon with general
formula CnH2n+2; alkanes with at least 16 carbons are
solids, usually found in primary producers (photosynthesizing organisms).
Alloy. A nonstoichiometric combination of metals.
Amino acid. Organic compound containing an amino
(NH2) and a carboxyl (COOH) group. There are 20
amino acids used in protein synthesis.
Assimilation. Incorporation of solid material into a magma.
Atomic number. The number of protons in an atomic
nucleus.
Authigenesis. Mineral formation from dissolved or solid
constituents already present at the site, as opposed to
constituents transported from elsewhere.
Biomolecule. Organic compound, such as a protein or
carbohydrate, that is a component of an organism.
Biopolymer. Complex organic molecule synthesized by
plants or animals. Examples of these molecules include carbohydrates, proteins, lignin, and lipids.
Bitumin. Generic term applied to naturally occurring,
flammable organic substances consisting of hydrocarbons. Bitumin is soluble in organic solvents.
Buffer. Assemblage of chemical species whose coexistence allows a system to resist change in some intensive property, such as pH or oxygen fugacity.
Calorimetry. Experimental measurement of heat evolved
or absorbed during a specific reaction.
Carbohydrate. Organic compound present in living cells
with general formula CH2O. It is the most abundant
class of organic compounds and includes sugars,
starches, and cellulose.
Carbonate compensation depth (CCD). Depth in the oceans
at which the downward flux of carbonate minerals is
balanced by their rate of dissolution.
Chalcophile. An element with an affinity for sulfide phases.
Chelate. Large molecule or complex that can enclose
weakly bound atoms or ions.
Chemical potential (µ). Partial molar free energy, which
describes the way in which total free energy for a
phase responds to a change in the amount of component i in the phase [=(∂G/∂ni)P,T,nj≠i ].
Chondrite. A common type of meteorite, with nearly cosmic elemental abundances, thought to be a sample of
the earliest Solar System material.
Closed system. A system that can exchange energy, but
not matter, with its surroundings. Compare: Isolated
system, Open system.
Colloid. Stable electrostatic suspension of small particles
in a liquid.
Components. Abstract chemical entities, independently
variable within a system, which collectively describe
all of the potential compositional variations within it.
Compressibility. Measure of the relationship between phase
volume and lithostatic pressure, [β T = −1/V(∂V/∂P)T ;
βs = −1/V(∂V/∂P)S].
Condensation. The formation of solids or liquids from a
gas phase during cooling.
353
354
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Congruent reaction. Reaction in which one phase melts or
dissolves to form another phase of the same composition. Compare: Incongruent reaction.
Conservative species. Any dissolved species whose concentrations are not affected in solution in any way, other
than dilution, by variations in the abundance of other
species in solution.
Cosmic abundance of the elements. Relative elemental abundance pattern in the Sun.
Cosmogenic nuclide. Any isotope formed by interaction of
matter with cosmic rays.
Cotectic. Boundary curve on an n-component phase diagram, along which two or more phases crystallize
simultaneously. Compare: Peritectic.
Delta notation. Notation used for stable isotope data,
equal to [(Rsample − Rstandard)/Rstandard ] × 1000, where
R is the ratio of heavy to light isotope of the element
of interest.
Diagenesis. The set of processes, other than weathering
or metamorphism, that change the texture or mineral
composition of sediments; these include compaction,
cementation, recrystallization, and authigenesis.
Differentiation. Separation within a homogeneous planet
of crust, mantle, and core; also, any process by which
magma can give rise to rocks of contrasting composition.
Diffusion. Dispersion of ions or molecules through a
medium that is not moving, due to a gradient in
some intensive property across the system.
Diffusion coefficent. Proportionality constant relating flux
to gradient in a diffusion equation.
Distribution coefficient (KD ). Ratio of concentrations of a
component in two coexisting phases.
e-folding time. Time required for a compositional variable
to return to within a factor of 1/e of its steady state
value following some perturbing event, numerically
equal to the inverse of the kinetic rate constant.
Eh . Redox potential of a cell, expressed in terms of
voltage.
Electrolyte. Any substance that dissociates into ions when
dissolved in an appropriate medium.
Electron. Subatomic particle, found in a region surrounding the nucleus of an atom, with a unit negative electrical charge.
Electron affinity (EA). Amount of energy released when an
electron is added to an atom, creating an anion.
Elementary reaction. Any chemical reaction that occurs
at the molecular level among species as they appear
in the written representation. Compare: Overall
reaction.
Enantiomer. One of two compounds with the same composition but having structures that are mirror images
of each other.
Endothermic reaction. Reaction that absorbs heat, resulting in a decrease in enthalpy. Compare: Exothermic
reaction.
Enthalpy. Thermodynamic function of state, H; a measure
of the energy that is irretrievably converted to heat
during any natural process.
Entropy. Thermodynamic function, S, the change of which
is defined as the change in heat gained by a body at
a certain temperature in a reversible process [dS =
(dQ/T)rev].
Equation of state. Function that defines interrelationships
among intensive properties of a system.
Equilibrium. Condition in which the properties of a system do not change with time.
Equilibrium constant (Keq ). Constant relating the composition of a system at equilibrium to its Gibbs free energy.
Eutectic. Point on a phase diagram at which the liquidus
touches the solidus, where two or more solids crystallize simultaneously from a melt.
Exothermic reaction. Reaction that evolves heat, resulting
in an increase in enthalpy. Compare: Endothermic
reaction.
Exsolution. Physical unmixing of two phases.
Extended defects. Dislocations and planar defects in crystals that provide effective pathways for diffusion.
Extensive property. Variable whose value is a measure of
the size of the system. Compare: Intensive property.
Extinct radionuclide. Any unstable nuclide that decays so
rapidly that virtually none remains at the present time.
Fluid inclusion. Fluid trapped within crystals. These may
be primary (trapped as the crystal grew) or secondary
(introduced at some later time).
Flux. Amount of material moving from one location to
another per unit time.
Fractional crystallization. Physical separation of crystals
and liquid, preventing equilibrium between phases.
Fractionation. Separation of any two entities, such as
isotopes (by mass), elements (by geochemical or cosmochemical properties), or crystals and melt (by fractional crystallization or fractional melting).
Fractionation factor. Ratio of a heavy to a light isotope of
an element in one phase divided by the same ratio in
another coexisting phase.
Glossary
Fugacity. Gas partial pressure that has been adjusted for
nonideality.
Geobarometer. Calibrated mineral exchange reaction that
is a strong function of pressure but not temperature.
Geochronology. Determination of the ages of rocks using
radiometric dating techniques.
Geopolymer. Any of several complex organic compounds
of high molecular weight, formed from biopolymers
during diagenesis; these include kerogen and bitumen.
Geothermal gradient. Rate of change of temperature with
depth in the earth; also called a geotherm.
Geothermometer. Calibrated mineral exchange reaction that
is a strong function of temperature but not pressure.
Gibbs free energy. Thermodynamic function of state, G,
which is a measure of the energy available for changing chemical bonds in a system.
Half-life. Time required for one-half of a given number of
atoms of a radionuclide to decay.
Half-reaction. Hypothetical reaction that illustrates either
gain or loss of electrons, useful for understanding
redox equilibria.
Heat. Transfer of energy that results in an increase in
temperature.
Heat capacity. Functional relationship between enthalpy
and temperature [CP = (∂H/∂T )P ; CV = (∂H/∂T )V].
Helmholtz free energy. Thermodynamic function of state,
F, which, for an isothermal process, is a measure of
the energy transferred as work; also called the work
function.
Henry’s law. Expression of solution behavior in which the
activities of dissolved species are directly proportional
to their concentrations at high dilution.
Humic substance. Altered complex organic substance
derived from the remains of terrestrial plants and
phytoplankton.
Ideal solution. Any solution in which the end-member constituents behave as if they were independent.
Immiscibility. Condition that results in spontaneous phase
separation.
Incongruent reaction. Reaction in which one phase melts
or dissolves to form phases of different composition.
Compare: Congruent reaction.
Intensive property. Variable whose value is independent of
the size of the system. Compare: Extensive property.
Internal energy. Thermodynamic function of state, E, the
change of which is a measure of the energy transferred as heat or work between the system and its
surroundings.
355
Ionization potential (IP). Amount of energy needed to remove an electron from an atom, creating a cation.
Isochron. Line on an isotope plot whose slope determines
the age of a rock system. Isochrons may be defined by
the isotopic compositions of constituent minerals or
whole rocks.
Isolated system. A system that cannot exchange either matter or energy with the universe beyond the system’s
limits. Compare: Closed system, Open system.
Isotopes. Atoms of the same element but with different
numbers of neutrons and hence different atomic
weights. Isotopes may be stable or unstable; also called
nuclides.
Kerogen. Insoluble, high-molecular-weight organic matter
derived from algae and woody plant material and
found in sedimentary rocks. This material may yield
petroleum when heated.
Kinetics. Description of a system’s behavior in terms of its
rates of change. Compare: Thermodynamics.
Lignin. A complex polysaccharide that provides structural support for many plants and is one of the chief
organic substances in wood.
Lipid. Class of aliphatic hydrocarbons containing the fatty
acids, fats, waxes, and steroids. These compounds are
energy sources for the host organism.
Liquidus. Boundary curve on a phase diagram representing
the onset of crystallization of a liquid with lowering
temperature. Compare: Solidus.
Lithophile. Element with an affinity for silicate phases.
Local equilibrium. Equilibrium that is attained on a small
scale, but not on a larger scale; also called mosaic
equilibrium.
Lysocline. Depth in the oceans at which the effects of
carbonate dissolution are first discernable.
Mass number. The number of protons plus neutrons in a
nucleus.
Mean residence time. Average residence time for a chemical species in a system before it is removed by some
loss process.
Metamorphism. Collection of processes that change the
texture or phase composition of a rock through the
action of temperature, pressure, and fluids.
Metasomatism. Metamorphism with accompanying change
in chemical composition, usually as a result of fluid
migration.
Metastable system. A system that appears to be stable because it is observed over a short time compared to the
rates of reactions that alter it.
356
G E O C H E M I S T R Y : PAT H WAY S A N D P R O C E S S E S
Mobile component. A component free to migrate into or
out of a system.
Monolayer. An organic coating with a thickness of only
one molecule.
Monolayer equivalent. The mass of organic matter that is
theoretically present over a square meter if all surfaces of the sediment have a monolayer coating.
Neutron. Subatomic particle, found in the nucleus of an
atom, with no electrical charge.
Nuclear statistical equilibrium. The steady state in which
disintegrative and constructive nuclear reactions balance in an evolved star.
Nucleation. The initiation of a small volume of a new
phase. This process may occur homogeneously (without a substrate) or heterogeneously (on a substrate).
Nucleosynthesis. The production of nuclides, primarily
in stars.
Nuclides. See Isotopes.
Open system. A system that can exchange matter and energy with its surroundings. Compare: Closed system,
Isolated system.
Overall reaction. Chemical reaction that describes a net
change involving several intermediate steps and competing pathways. Compare: Elementary reaction.
Partial melting. Fusion of something less than an entire
rock, a process that may occur under equilibrium or
fractional (continuous separation of melt and crystals)
conditions.
Periodic table. A tabular arrangement of the elements, in
order of their increasing mean atomic weight, in such
a way that elements with similar chemical properties
related to their electronic configuration are grouped
together.
Peritectic. Point on a phase diagram at which an incongruent reaction occurs. Compare: Cotectic.
pH. Measure of acidity in solution [pH = −log10aH+ ].
Phase. Substance with continuous physical properties.
Phase diagram. Graphical summary of how a system
reacts to changing conditions or composition.
Point defects. Crystal imperfections caused by the absence of atoms at lattice sites; such defects may be
intrinsic or extrinsic.
Proton. Subatomic particle, found in the nucleus of an
atom, with a unit positive electrical charge.
Quantum. A discrete quantity or level of energy.
Racemic. An equal mixture of d- and l- enantiomers.
Racemization. Transformation of optically active compounds, such as amino acids, from the d-form to the
l-form or vice versa.
Radioactive decay. Spontaneous transformation of one
unstable nuclide into another isotope; this occurs by
alpha or beta decay, positron decay, electron capture,
or spontaneous fission.
Radiogenic nuclide. Isotope formed by decay of some
parent radionuclide.
Raoult’s law. Expression of solution behavior in which
the activities of dissolved species are equal to their
concentrations.
Refractory. Having a high melting temperature or condensing from a gas at a high temperature.
Salinity. The total dissolved salt content.
Saturated hydrocarbons. Organic molecules consisting of
carbon and hydrogen atoms linked only by single
bonds; these may form chain structures called aliphatic compounds, or cyclical structures called alicyclic
compounds.
Secular equilibrium. Condition in which the decay rate for
daughter radionuclides in a decay series equals that
of the parent.
Siderophile. An element with an affinity for metallic phases.
Solidus. Boundary curve on a phase diagram representing
complete solidification of a system with lowering temperature. Compare: Liquidus.
Solubility product constant (Ksp ). Equilibrium constant for
a dissolution reaction.
Solute. The dissolved species in a solution.
Solvent. The host species in a solution.
Solvus. Region of unmixing on a phase diagram.
Spinodal. Region of unmixing inside the solvus on a
phase diagram; such unmixing involves no nucleation
barrier.
Standard state. Arbitrarily selected state of a system used
as a reference against which changes in thermodynamic properties can be compared.
Steady state. Condition in which rates of change within
a system balance one another, so that there is no net
change in its appearance for an indefinite time period,
even though various parts of the system are not in
thermodynamic equilibrium.
Sterol. A group of solid, mainly unsaturated steroid alcohols, including cholesterol and ergosterol, present
in plant and animal tissues.
Stoichiometric. Summing to the whole, as in stoichiometric equations or coefficients.
Supernova. Massive stellar explosion, important for
nucleosynthesis.
Surface tension. Tensional force applied perpendicular to
any line on a droplet surface.
Glossary
System. That portion of the universe that is of interest for
a particular problem; a system can be open, closed, or
isolated.
Thermal expansion. Measure of the relationship between
phase volume and temperature [αP = −V(∂V/∂T )P].
Thermocline. Transitional range between the warm surface zone and the cold deeper zone in the oceans.
Thermodynamics. Set of laws that predict the equilibrium
configuration of a system and how it will change if
its environmental parameters are changed. Compare:
Kinetics.
Titration. Experiment in which a reaction is allowed to
proceed incrementally, so that proportions of reactants can be determined.
Trace elements. Elements that occur in rocks with concentrations of a few tenths of a percent or less by
weight; these may exhibit compatible or incompatible
behavior.
Transition elements. Elements with inner orbitals incompletely filled by electrons.
Troposphere. Lower portion of the atmosphere that contains the bulk of its mass.
357
Uncompressed mean density. Mean density of a planet, corrected for the effects of gravitational self-compression.
Undercooling. Difference between the equilibrium temperature for the appearance of a phase and the temperature at which it actually appears.
Unsaturated hydrocarbons. Organic molecules with double
or triple bonds between carbon atoms; these may form
chain structures, called aliphatic compounds, or cyclic
structures, called aromatic compounds.
Variance. Number of independent parameters that must
be defined in order to specify the state of a system at
equilibrium; also called degrees of freedom.
Volatile. Term indicating that a given element or compound commonly occurs in the gaseous state at the
temperature of interest, or indicating that a given element condenses from a gas at low temperature.
Weathering. Those processes occurring at the Earth’s surface that cause decomposition of rocks.
Work. Transfer of energy that causes a mechanical change
in a system or its surroundings; the integral of force
× displacement.
INDEX
Abundance, of elements,
cosmic, 231, 316–318
Accretion, 322, 336, 338
Achondrite, 334
Acid, 102, 112, 114, 122
Actinide, 22–23
Activation energy, 207
Activity, 68, 180
-activity diagrams, 132
coefficient, 66–71, 180
iron in garnet, 182
relation to concentration,
179
silica (aqueous), 112–113
Adiabatic process, 38, 241
Advection, 80, 85
Age of the Earth, 333
Albite, 179, 194–195, 201,
207, 214, 269
Alcohol, 101–102
Aldehyde, 102
Alkali, 22–23
Alkaline earth, 22–23
Alkalinity
carbonate, 147–148
titration, 147
Alkane, 101
Alkemade line, 202
Alkemade’s theorem,
202–203
Alkene, 101
Aliphatic structural units,
122
Alloy, 12, 221, 231
Alpha particle, 287
Amine, 102–103
Amino acid, 102–103, 107,
331
Aminostratigraphy, 108
Amorphous solid solutions,
60
Analcite, 179, 207
Andalusite, 176, 184, 257
Anders, Edward, 330
Anion, 22
Anorthite, 192, 194,
201–202, 241, 244, 269,
326
Aragonite, 150, 152, 175,
220
Arrhenius equation, 206
Arrhenius plot, 206, 209
Assimilation, 304–306
Atmophile, 23–24
Atmosphere
carbon dioxide, 121, 143
composition, 141–142
escape, 310
evolution with time,
163–166
outgassing, 159, 308–310
Atom, 13
Atomic mass unit, 15
Atomic number, 13, 253, 317
Atomic radius, 20–21
Avogadro’s number, 4
Bacteria, 97, 107
Banded iron formation, 164,
166
Benzene, 30, 101
Berner, Robert, 90, 157
Beta particle, 287
Biolimiting constituents of
seawater, 138
Biomarker, 105, 280
Biotite, 171, 173, 183
Bitumen, 95
Bjerrum plot, 144–145
Bohr, Niels, 18
Boltzmann factor, 207
Bond, 18, 24
conjugated, 101
covalent, 25, 29
double, 101
hydrogen, 30
ionic, 25–26
metallic, 30–31
polarity, 15
single, 101
strength and isotope
fractionation, 266–267
structural implications,
26–32
Van der Waals, 31
Bose-Einstein relation, 82
Boundary curve, 201–202
Bowen, Norman, 2
Brine, 278
Brucite, 113, 198
Buffer
capacity, 112
external, 199
fluid composition, 199
internal, 199
oxygen fugacity, 160
pH, 144–145
Bulk distribution coefficient,
249
C3, C4 plants, 279, 280
CAI. See Calcium-aluminum
inclusions
Calcite, 150, 152, 171, 174,
176, 198, 220, 266–269
Calcium-aluminum inclusions
(CAI), 327, 332
Calorimetry, 44, 48, 127,
238
Carbohydrate, 102–103,
279
Carbon
abundance, 95
in biologic cycle, 142–143
chiral, 102
cycle, 94–96, 155–157
dissolution in seawater, 95,
138–139
isotopes, 278
radioactive (carbon-14),
286, 301
reservoirs, 95, 155, 162
Carbon dioxide
atmospheric abundance,
95, 143
atmospheric variation with
time, 158
in fluids, 198
isotopic composition, 265
in mantle and crust, 258
and melting, 241
and metamorphism, 259
partial pressure, 121, 198
redox systems containing,
129
Carbonate. See also
Aragonite, Calcite,
Siderite
compensation depth,
150–151
equilibria, 143
isotopic composition, 278
marine, 307
precipitation, 87
solubility, 148–150
Carbonatite, 248
Catagenesis, 100
Cation, 21
Cellulose. See Carbohydrate
Chalcophile, 23–24, 320
Chelate, 75, 122, 132
Chemical potential, 51–52,
64, 169, 171, 179–181,
189
Chemical reaction potential,
64
Chemical variation diagram,
246–247
Chitin. See Carbohydrate
Chlorophyll, 94, 105
Chondrites
ages, 297, 332–333
classification, 322
composition, 318–321,
334
mixing model, 336–337
organic matter, 331
oxygen isotopes, 328
petrology/mineralogy, 322,
327
Clapeyron equation, 45, 177,
198
Clarke, F. W., 2
Clay, 117
Clayton, Robert, 274, 328
Coal
formation, 100
isotopic composition, 279
types, 100
Coalification, 100
359
360
Index
Colloid, 84, 97, 122
Comet, 332
Common ion effect, 74
Complex ions, 75
of aluminum, 114
in seawater, 151
of uranium, 132
Components, 48–49
chemical, 170, 188
fluid, 198
mixing, 179
mobile, 171
Compound, 12
Compressibility, 175
Concordia, 295
Condensation
calculation, 323–325
model, 335
sequence, 326, 335
temperature, 323–324, 335
Congruent reaction, 112–113
Conservative elements in
seawater, 137
Conservative ions in
seawater, 147
Constant
Avogadro’s number, 4
decay, 288, 292
equilibrium, 64, 111–115,
126, 267, 325
Faraday’s, 125, 353
fundamental, 10
gas, 353
Henry’s Law, 180
rate, 8, 206
solubility product, 69, 149
stability, 75
Continuity equation, 80–81
Controlled cooling rate
experiment, 222
Conversion factors, 10
Core, 231, 334
composition, 231, 233
Corundum, 326
Cosmic ray, 301
Cosmochemistry, 313
Cosmogenic radioactive
isotopes, 301
Cotectic, 201–202
Critical point, 170
Crust, 227
composition, 228
continental, 228, 236
oceanic, 227
Crystal
daughter, 255
field theory, 251–252
size distribution analysis,
223
Crystallization
on binary phase diagrams,
192–194
equilibrium, 244
fractional, 243
morphologies, 219
on ternary phase diagrams,
202–203
Dalton, John, 13
Debye-Hückel equation,
69–70
Decomposition, 114
Defect
extended, 209
extrinsic, 209
Frenkel, 209
Schottky, 209
Degrees of freedom, 170–171
Dehydration, 179
Deuterium, 263. See also
Stable isotopes
Diagenesis, 79, 99–100, 279
Diagenetic equation, 92
Differentiation, 243
Diffusion, 80
coefficient, 80, 84
and growth, 218
inter-, 208
self, 208
tracer coefficient, 208
Diopside, 174, 176, 192,
201, 241, 269, 326
Discordia, 295
Disorder, 42–43
Dissolution, 87, 113, 122
Distributary reaction point,
202–203
Distribution coefficient, 183,
248, 253
Divariant field, 170
Eclogite, 229–230
Eh, 125
Eh-pH diagram, 127–129,
131, 133
Electrochemical
measurements, 127
Electrode, 130
Electrolyte. See Solution
Electron, 13, 18
affinity, 22
capture, 287
exchange, 124
orbital, 18–20, 28, 252
shielding, 20
valence, 20
Electronegativity, 25
Element, 12
associations, 22–24
compatible, 251
incompatible, 228, 251
naming, 15
periodic properties, 14,
20–21
Enantiomer, 102
Endothermic, 44
Enstatite, 174, 176, 208,
269, 326
Enthalpy, 44, 172
of melting, 238
Entropy, 40, 172
and disorder, 42–43, 238
and melting, 241
Epimerization, 107
Epstein, Samuel, 264, 272
Equilibrium, 5, 42–43, 169,
189
condensation. See
Condensation
constant, 64, 111–115,
126, 267, 325
heterogeneous, 52–53
mechanical, 53
nuclear statistical, 315
secular, 290
tests for, 169–172
thermal, 36, 53
Euphotic zone, 97
Europium anomaly, 253
Eutectic, 192, 203, 241
Evaporite, 270
Exchange operator, 50
Exothermic, 44, 236
Exposure age, 302
Exsolution, 195–196,
214–216
Extensive properties, 4
Fats. See Lipids
Fayalite, 244, 248
Fick’s First Law, 80
Fick’s Second Law, 80–81
Fischer-Tropsch synthesis,
331
Fission
reactor, 303
spontaneous, 287
tracks, 298
Fluid
composition, 198–199,
255
cycling, 259
effect on phase diagrams,
197
effect on reactions,
177–179
inclusion, 255–256
and melting, 241
pressure, 259
speciation, 258
Flux, 7, 80
carbon, 155
equation, 7
fluid, 236
heat, 236
mass, 233–237
noble gases, 309
sodium, 153
Forsterite, 112, 175, 212,
239, 244, 269, 326
Fraunhofer lines, 319
Free energy. See also Gibbs
free energy
of reaction, 64
Frequency factor, 207
Fugacity, 66
Fulvic acid, 122
Functional group, 101–102
Fusion, 315
Ḡ-X2 diagram, 188–190
Galena, 270
Gamma-ray log, 291
Garnet, 171, 183–184, 229,
269
Gas seeps, 279
Geiger counter, 291
Geobarometer, 184
Geochemistry
conversion factors, 353
finding data, 349
historical overview, 1–3
problem solving, 9–10, 343
Geochronology, 290
Geopolymer, 100
Geotherm, 240
Geothermometer, 183,
266–270
Gibbs, Josiah Willard, 45, 170
Gibbs-Duhem equation, 53,
170
Gibbs free energy
and chemical potential, 189
defnition, 45–46, 169
effect of changing pressure,
174
effect of changing
temperature, 174
graphical representation,
188–190
isotopic exchange reaction,
268
Index
Gibbsite, 114–115
Glycerides. See Lipids
Goldschmidt, Victor, 2–3
element classification, 23,
320
Gradient
compositional, 80, 219
geothermal, 240
thermal, 219
Graphite, 31
Greenhouse effect, 154, 163
Growth, 216
crystal, 217
diffusion-controlled, 218
interface-controlled, 216
Half-life, 289, 292
Half-reactions, 124
conventions, 124–125
for iron oxides, 128–129
for manganese oxide, 127
Halogen, 22–23
Heat, 38
capacity at constant
pressure, 173
Helmholtz function, 45
Hematite, 128, 199
Henry’s Law, 179, 248
Heterogeneous accretion,
322, 336, 338
Holland, Heinrich, 153, 159,
166
Humic acid, 122–123
Hybrid orbital, 28
Hydrocarbon, 101
aliphatic, 101
aromatic, 101
primordial soup, 165
saturated, 101
unsaturated, 101
Hydrologic cycle, 271–273
Hydrothermal, 275–277
Ideal gas law, 175, 324
Ideal mixture, 62, 179
Immiscibility, 247–248
Incongruent reaction, 113,
115
Intensive properties, 4, 36
Interaction parameter, 182
Internal energy, 38
Invariant point, 171
Ion, 21
Ionic radius, 22, 229
Ionic strength, 70, 74, 151
Ionic structure, 26
Ionization potential, 21
Iron
condensation, 326
in core, 232–233
Iron meteorites, 221–222,
297, 333
Isochron, 293
Isothermals, 36
Isotopes, 14–17. See also
Radioactive isotopes,
Stable isotopes
Joly, James, 153
Joule, James, 37
Kaolinite, 115–118, 163
Kasting, James, 163–164,
166
Kerogen
definition, 95, 100
isotopic composition, 279
in meteorites, 331
types, 100, 102
K-feldspar, 116–118, 163
Kinetics, 5, 80, 87, 206
aragonite-calcite
transformation, 220
barrier, 6
biomarker, 106
condensation, 325–326
diffusion, 208
first-order, 8, 89, 143
isotope fractionation, 278
nickel diffusion in iron
meteorites, 220–222
nucleation, 210
rate constant, 206
temperature effects,
206–208
Kyanite, 176, 184
Lanthanide, 22–23, 252
Lasaga, Antonio, 87
Latent heat of fusion, 238
Lavoisier, Antoine, 1, 13
Lead-lead age, 296
Leucite, 193, 202, 248
Lever rule, 191
Ligand, 75, 122
Lignin, 105, 279–280
Limestone
equilibrium with water,
121, 149
metamorphic reaction, 199
Lipids, 279
characteristics, 104
types, 104–105
Liquid line-of-descent, 246
Liquidus, 192
Lithophile, 23–24, 228–229,
320
Lockyer, Norman, 319
Lysocline, 151
Maceral, 100
Magnesium problem, 154
Magnetite, 128, 160, 173,
199, 266
Manganese nodule, 127
Mantle, 229
array, 304–305
composition, 230, 337
convection, 233–234
degassing, 308–310
depleted, 230
discontinuity, 231
heterogeneity, 302–304
phase changes, 231–232
plume, 233–234
undepleted, 230
Mars
composition, 337
interior, 340
meteorites, 333–334
Mass
balance in oceans, 145,
152
fractionation, 278, 328
Maxwell relationships, 44–45
McLennan, Scott, 228, 237
Mean residence time, 153
Mean salt method, 69
Melting, 237
causes, 240
equilibrium (batch),
238–239
fractional, 239
incremental batch, 239
paths, on phase diagrams,
239
and tectonics, 243
thermodynamic effects,
238
Mendeleev, Dmitri, 13
Metastable, 6, 152, 175, 191
Meteorites. See Chondrites,
Iron meteorites
Methane, 162–163, 165,
258, 279, 281
Miller, Stanley, 165
Mixing
ideal, 179
molecular, 58
nonideal, 179
on sites, 59
361
Models
box, 153–158
continuum, 158
Moho, 227
Molecule, 13
Montmorillonite, 118
Moon
composition, 338–339
origin, 338
Muscovite, 116–118, 269
Nernst equation, 126
Neutrino, 287
Neutron, 13
activation, 299–300
capture, 315
decay, 263
irradiation, 299
Nier, Alfred, 2, 264
Nitrate, 140–141
Nitrogen
in atmosphere, 142
dissolved in oceans, 140
isotopes, 279
reservoirs, 162
Noble gas, 22–23
in atmosphere, 142,
309–310
isotopes, 309–310
Nonconservative elements in
seawater, 138
Nonconservative ions in
seawater, 147,
152–153
Nucleation, 208
barrier, 213
critical radius, 211
crystal, 223
growth, 216
heterogeneous, 212
homogeneous, 212
in melts, 211
in solids, 214
Nucleosynthesis, 314–316
Nuclides, 15
chart, 16, 316
radioactive. See
Radioactive isotopes
stable. See Stable isotopes
Oceans
chemical composition,
137–141
circulation, 141, 307
depth variations, 140
isotopic composition,
307–308
362
Index
Oceans (continued)
isotopic evolution,
270–271
paleotemperatures, 268
primary production
estimates, 96
salinity, 137, 140–141
sodium mean residence
time, 153
Oklo, 303
Olivine, 229. See also
Fayalite, Forsterite
Orbital. See also Electron
hybrid, 28
Organic acid, 102, 122
Organic matter, 94, 114
diagenetic alteration,
99–100
effect of redox state, 98
isotopic composition, 280
labile, 98
in meteorites, 330–331
in oceans, 96–97, 138
preservation, 97–100
refractory, 98
in sediments, 97–98
in soils, 96
Organic nomenclature,
101–102
Organic primordial
molecules, 165
Organic structure of
molecules, 101
Orthoclase, 193, 195, 214
Outgassing, 159, 308–310
Oxidation
buffer reactions, 160
graphical representation,
127–129, 131, 133
organic matter, 99
thermodynamic
conventions, 124–127
Oxygen
atmospheric, 142, 163
biologic cycle, 142–143
in core, 233
dissolved in oceans,
140–141
effect on organic matter, 98
fugacity, 160
isotopes, 266–269,
271–279, 328, 338
rise of, 164
Partial pressure, 64, 198
Pathways, 5, 118, 154. See
also Models
Pauli exclusion principle, 19
Pauling, Linus, 25
Peatification, 100
Periclase, 198
Peridotite, 229, 240
Periodic table, 13–14
Peritectic, 193
Perovskite, 326
pH
buffer, 144–145
definition, 112–115
measurement, 130
Phase
definition, 35
rule, 170
Phase diagram, 170
albite-anorthite, 194, 219
albite-anorthite-diopside,
201
albite-orthoclase, 195, 214
aluminosilicates, 178
analcite-quartz-albite, 179
anorthite-leucite-silica, 202
binary, 191–195
brucite-periclase, 198
calcite-quartz-wollastonite,
198
derivation from Ḡ-X2
diagram, 190–191
diopside-anorthite, 192,
241
entropy-pressure, 241
forsterite-anorthite-silica,
244
forsterite-fayalite, 244
forsterite-silica, 239
iron-oxygen, 199
leucite-silica, 193
peridotite, 240
peridotite–carbon
dioxide–water, 243
pressure-temperature, 199
ternary, 201–203
thermodynamic calculation
of, 196–197
water, 171
Phenol, 101–102, 122
Photochemical reactions, 162
Photosphere, 319
Photosynthesis, 96, 142, 164,
278
Phyllosilicate, 117
Phytoplankton, 96
Piclogite, 231
Plagioclase, 171, 194, 197
Planck, Max, 42
Planets, 314
bulk composition, 333, 339
cores, 339–340
mean density, 333
volatile element depletion,
334
Polymer, 60
Porosity, 83, 85
Porphyrin, 105
Precipitation, 87
carbonate, 87
ferric and manganese
oxides, 127
gibbsite, 115
uranium, 132
Primary phase field, 202
Primary production, 96
Protein, 103, 279
Proton, 13
Proton-proton chain, 315
Pseudoadvection, 85
Pyrite, 164, 270
Pyrolite, 230
Pyrolysis, 330
Pyrophyllite, 116–118, 257,
277
Pyroxene, 229
Pyrrhotite, 270
Quantum numbers, 18–19,
27
Quartz, 111, 160, 171, 174,
176, 184, 198, 266, 276,
302
Racemization, 107, 331
Radioactive isotopes, 16–17,
286
aluminum-26, 297, 302,
329, 333
argon-40/argon-39,
300–301
beryllium-10, 302, 306
cosmogenic, 301
daughter, 287
extinct, 29
as geochemical tracers, 302
hafnium-182/tungsten-182,
297, 333
helium-3, 308–309, 315
iodine-129/xenon-129,
309
neodymium-143/
samarium-147, 294,
302–304, 307
neon-21, 309
parent, 287
potassium-40/argon-40,
288, 291–292, 309
rubidium-87/strontium-87,
289, 292–293, 302–308,
328
uranium/lead, 290,
294–296, 332
Radioactivity, 287
branched decay, 287
decay constant, 288
decay mechanisms, 287
decay series, 290
fission reactor, 303
induced, 299
rate, 288
secular equilibrium, 290
spontaneous fission, 287
well-logging, 291
Radiogenic, 287
Radionuclides. See
Radioactive isotopes
Raleigh
distillation equation, 271
fractionation, 249
Raoult’s Law, 179
Rapid (r) process, 315–316
Rare earth element
normalization, 321
partitioning, 252–253
Rate-limiting step, 219
Reaction
congruent, 112–113
curve, 201, 203
incongruent, 113, 115
Redfield ratio, 138
Redox reaction. See
Oxidation
Reduction, 124–127
Refractory element, 321,
334. See also under
Organic
Reservoir, 5
carbon, 155
fluid, 258
noble gas, 309
sodium, 153
in solid Earth, 227
Residence time, mean, 153
Reversible, 40, 125
Ringwood, Ted, 230–232,
333
Rutile, 184, 269
Salinity, 137, 140, 256
Saturation, 150
Scintillation, 291
Seawater. See also Oceans
charge balance, 147
chemical model, 150–152
sodium budget, 153
Serpentine, 112
Shell, 18–20, 28, 252
Siderite, 129
Siderophile, 23–24, 231, 320,
323
Silica, 28, 111, 140, 193, 202
Index
Sillimanite, 171, 176, 184
Slow (s) process, 315–316
Smectite, 117
Soil, 120–121
Solar composition, 318–319
Solidus, 192, 241–243
Solubility, 71
aluminosilicates, 115–119
carbonates, 148–149
gibbsite, 114–115
magnesium silicates,
112–114
in seawater, 138–139
silica, 111–112
uranium, 132
Solute, 62
Solution, 56
electrolyte, 62, 68
gas, 62
ideal, 62, 179
melt, 60
models, 181–182
nonideal, 66, 72–73, 179,
247
solid, 57–60, 194
Solvent, 62
Solvus, 195
Sorption, 98–99
Spallation, 302
Sphalerite, 270
Spinodal, 214
Stability fields
aluminosilicates, 117
iron oxides, 128
kaolinite, 116
siderite, 131
uraninite, 133
Stable isotopes, 16–17, 263
carbon, 265, 278–279, 331
delta notation, 264
fractionation, 264
in biogenic systems,
278–279
and bond strength,
266–267
characteristics that
permit, 264
factor, 264–265, 269
in hydrologic cycle,
271–273
in hydrothermal systems,
275–277
in oceans, 270–271
in sedimentary basins,
278
hydrogen/deuterium, 263,
271–277, 331
mass-independent
fractionation, 274
nitrogen, 279, 331
oxygen, 264, 271–279, 281
standards, 264
sulfur, 264, 270–271, 281
thermometry, 266–270
Standard electrode potential,
126
Standard hydrogen electrode,
125
Standard state, 47, 180
Starch, 102
Stardust, 330
State
equations of, 35–36, 67
functions of, 44
of matter, 12
standard, 47, 180
steady, 6, 7, 94, 152
Stereoisomer, 102
Steroid (sterol), 104–105
Stoichiometry, 18
Stromatolite, 164
Structure
benzene, 30
and bonding, 26
clays, 117
graphite, 31
ice, 30
pyroxene, 58
sodium chloride, 26
Subduction, 306
Subtraction curve, 201–202
Sugar, 102
Sulfate, 279
Sulfide, 26
Supernova, 315, 329
Supersaturation, 114, 210
Surface
free energy, 211
tension, 211, 213
System, 4
closed, 5, 199
isolated, 4, 41
open, 5, 119, 171, 199
Talc, 113, 208
Taylor, Ross, 228, 237, 339
Temperature, 36
condensation, 323–324,
335
diagenetic (effective), 108
fluid species, 258
gradient in oceans, 140
potential, 140
sea surface, 106–107
Terpenoids. See Lipids
Thermal expansion, 175
Thermocline, 141
Thermodynamics, 5, 35, 169
First Law, 38
and phase diagrams,
196–197
Second Law, 39
Third Law, 48
Thermometer, 36
Thiemens, Mark, 274
Thode, Harry, 264
Titration curve, 145–146
Tortuosity, 83
Trace elements, 248
compatibility, 251
partitioning during
crystallization, 250
partitioning during
melting, 250
Transition metal, 20, 22–23,
124, 251–252
Treibs, Alfred, 3, 94
Tributary reaction point,
202–203
Triple point, 170
Troposphere, 142
TTT plot, 215–216
Undercooling, 211
Univariant curve, 170, 200
Unmixing. See Exsolution,
Immiscibility
Uraninite, 133, 164, 266
Uranium
isotopes, 290, 294–296
oxidation states, 132
as a refractory element,
334
transport, 132–133
Urey, Harold, 2, 263–264,
268, 333
Valence, 20, 229
Valence bond theory, 27
Van der Waals equation, 3, 6,
67
Van’t Hoff equation, 173
Vapor bubble, 256
Variables
extensive, 4
intensive, 4, 36
Variance, 170–171
Vernadsky, V. I., 3
Viscosity, 60–61
Volatiles
depletion, 334, 338
element, 255, 321
excess inventory, 159
fractionation, 321–322, 334
363
ices, 330–332
loss to space, 163
melting, 241
outgassing, 159–160
reservoir abundances, 162
weathering sinks, 164
Volatility, 321
Volcanic gases, 161
Volume, effect of temperature
and pressure, 175
Water.
condensation, 321
connate, 178
formation, 79
hydrologic cycle, 271–273
isotopic composition,
267–268
magmatic, 276
in mantle, 258
meteoric, 273, 276
miscibility with carbon
dioxide, 198
molecules, 29–30
natural, 72–73, 112
oceans. See Oceans
physical properties and
isotopic composition,
266
rain, 143, 272
reactions, 198
reservoirs, 162
rivers, 119, 153
as solute, 111
streams, 121
vapor, 142
Waxes. See Lipids
Weathering, 111
agents, 121
open-system, 119
oxygen and carbon
consumed, 142
rate, 121–122
sequence, 119–120
Well-logging techniques,
291
Widmanstätten pattern,
221–222
Wolframite, 276
Wollastonite, 171, 198
Work, 37, 125
Work function, 45
Xenolith, 230
Zooplankton, 96