Tips - Principles of Geochemistry PDF
Tips - Principles of Geochemistry PDF
Tips - Principles of Geochemistry PDF
Principles of
Geochemistry
GIULIO OT TONELLO
Ottonello, Giulio.
[Principi di geochimica. English]
Principles of geochemistry / Giulio Ottonello.
p. cm.
Includes bibliographical references (p. 817) and index.
ISBN 0-231-09984-3 (acid-free paper)
1. Geochemistry. I. Title.
QE515.08813 1997
551.9—dc20 96–23987
CIP
Preface xi
PA R T O N E
Geochemistry of Crystal Phases
CHAPTER ONE
Elemental Properties and Crystal Chemistry 3
1.1 General Information on Atomic Structure and the Nature
of Electrons 3
1.2 Periodic Properties of the Elements 10
1.3 Generalities on the Concept of Chemical Bond in Solids 15
1.4 Ionic Bond and Ionic Radius 16
1.5 Pauling’s Univalent Ionic Radii 19
1.6 Covalent Bond and Covalent Radius 20
1.7 Electronegativity and Fractional Ionic Character 21
1.8 Ionic Poliarizability and van der Waals Bond 26
1.9 Crystal Radius and Effective Distribution Radius (EDR) 29
1.10 Effective Ionic Radii 32
1.11 Forces Interacting Between Two Atoms at Defined
Distances 43
1.12 Lattice Energy of an Ionic Crystal 45
1.13 Born-Haber-Fayans Thermochemical Cycle 52
1.14 Internal Energy of an Ionic Crystal and Temperature 54
1.15 Internal Energy and Pressure 58
1.16 Crystal Field Theory 63
1.17 Ligand Field Theory 73
1.18 The Gordon-Kim Modified Electron Gas (MEG) Model 81
1.19 Polarization Energy 87
/ v
vi / Contents
CHAPTER TWO
Concepts of Chemical Thermodynamics 91
2.1 General Concepts and Definitions 91
2.2 Gibbs Free Energy and Chemical Potential 92
2.3 Gibbs Free Energy of a Phase as a Function of the
Chemical Potential of Its Components 93
2.4 Relationships Between Gibbs Free Energy and Other
Thermodynamic Magnitudes 97
2.5 Partial Molar Properties 100
2.6 Gibbs Phase Rule 101
2.7 Stability of the Polymorphs 103
2.8 Solid State Phase Transitions 107
2.9 Standard State and Thermodynamic Activity 113
2.10 Calculation of Gibbs Free Energy of a Pure Phase in
Various P-T Conditions 117
2.11 Gibbs-Duhem Equation 118
2.12 Local Equilibrium, Affinity to Equilibrium, and Detailed
Balancing 119
CHAPTER THREE
Thermochemistry of Crystalline Solids 122
3.1 Entropy, Heat Capacity, and Vibrational Motion of
Atoms in Crystals 122
3.2 Heat Capacity at Constant P and Maier-Kelley Functions 131
3.3 Entropy and Heat Capacity from Vibrational Spectra:
Kieffer’s Model 135
3.4 Empirical Estimates of Heat Capacity at Constant
Pressure 141
3.5 Empirical Estimates of Standard State Entropy 148
3.6 Estimates of Enthalpy and Gibbs Free Energy of
Formation from Elements by Additivity: the
⌬O2⫺ Method 151
3.7 Gibbs Free Energy of a Phase at High P and T, Based on
the Functional Forms of Heat Capacity, Thermal
Expansion, and Compressibility 155
3.8 Solid Mixture Models 157
3.9 General Equations of Excess Functions for Nonideal
Binary Mixtures 168
3.10 Generalizations for Ternary Mixtures 170
3.11 Solvus and Spinodal Decomposition in Binary Mixtures 173
Contents / vii
CHAPTER FOUR
Some Concepts of Defect Chemistry 185
4.1 Extended Defects 185
4.2 Point Defects 187
4.3 Intrinsic Disorder 187
4.4 Extrinsic Disorder 189
4.5 Point Impurities 192
4.6 Energy of Formation of Point Defects 192
4.7 Defect Concentration as a Function of Temperature
and Pressure 196
4.8 Associative Processes 198
4.9 Stability of a Crystalline Compound in the Presence of
Defect Equilibria: Fayalite as an Example 202
4.10 Diffusion in Crystals: Atomistic Approach 205
4.11 Diffusion and Interdiffusion 212
CHAPTER FIVE
Silicates 217
5.1 General Information 217
5.2 Olivines 223
5.3 Garnets 248
5.4 Pyroxenes 266
5.5 Amphiboles 299
5.6 Micas 321
5.7 Feldspars 345
5.8 Silica Minerals 371
5.9 Thermobarometric Properties of Silicates 375
PA R T T W O
Geochemistry of Silicate Melts
CHAPTER SIX
Thermochemistry of Silicate Melts 411
6.1 Structural and Reactive Properties 411
6.2 Structure of Natural Melts 419
viii / Contents
CHAPTER SEVEN
Introduction to Petrogenetic Systems 449
7.1 Phase Stability Relations in Binary Systems (Roozeboom
Diagrams) 449
7.2 Extension to Ternary Systems 466
7.3 Crystallization and Fusion in an Open System: Diopside-
Albite-Anorthite as an Example 471
PA R T T H R E E
Geochemistry of Fluids
CHAPTER EIGHT
Geochemistry of Aqueous Phases 479
8.1 General Information on Structure and Properties
of Water 479
8.2 Electrolytic Nature of Aqueous Solutions:
Some Definitions 491
8.3 Models of “Ionic Coupling-Complexation” and
“Specific Interations” 491
8.4 Activities and Activity Coefficients of Electrolytes in
Aqueous Solutions 492
8.5 Ionic Strength and Debye-Hückel Theory 494
8.6 The “Mean Salt” Method 497
8.7 Activity of Dissolved Neutral Species 499
8.8 Activity of Solvent in Aqueous Solutions 501
8.9 Speciation 502
8.10 Carbonate Equilibria 510
8.11 Thermodynamic Properties of Solutes Under High P and
T Conditions: the Helgeson-Kirkham-Flowers
Approach 520
8.12 Redox Conditions, pH, and Thermodynamic Stability of
Aqueous Solutions 539
Contents / ix
CHAPTER NINE
Geochemistry of Gaseous Phases 612
9.1 Perfect Gases and Gaseous Mixtures 612
9.2 Real Gases at High P and T Conditions 616
9.3 The Principle of Corresponding States and Other
Equations of State for Real Gases 620
9.4 Mixtures of Real Gases: Examples in the C-O-H System 623
9.5 Volcanic Gases 626
9.6 Solubilities of Gaseous Species in Silicate Melts 631
9.7 Fluid Equilibria in Geothermal Fields 643
PA R T F O U R
Methods
CHAPTER TEN
Trace Element Geochemistry 657
10.1 Assimilation of Trace Elements in Crystals 657
10.2 Raoult’s and Henry’s Laws 657
10.3 High Concentration Limits for Henry’s Law in Crystalline
Solutions 659
10.4 Lower Concentration Limits for Henry’s Law Behavior in
Silicate Crystals 666
x / Contents
CHAPTER ELEVEN
Isotope Geochemistry 707
11.1 Isotopes in Nature 707
11.2 Nuclear Energy 713
11.3 Nuclear Decay 715
11.4 Isotopic Decay 722
11.5 Growth of Radioactive Products 723
11.6 Isotope Fractionation 726
11.7 Application I: Radiometric Dating 741
11.8 Application II: Stable Isotope Geothermometry 767
APPENDIX ONE
Constants, Units, and Conversion Factors 801
APPENDIX TWO
Review of Mathematics 805
A2.1 Exact Differentials and State Functions 805
A2.2 Review of Implicit Functions 807
A2.3 Integration on Exact Differentials 808
A2.4 Operations on Vectors 809
A2.5 Lagrange Equations, Hamiltonian 811
A2.6 Error Function 814
A2.7 York’s Algorithm for Isochrons 815
References 817
Index 863
PREFACE
A quick look at the reference list of this textbook (initially conceived for Italian
university students) is sufficient to appreciate the leading role played by North
American scientists in the development of geochemistry, and it is a great honor
for me to present my appraisal of this fascinating discipline to readers whose
mother tongue is English.
The fascination of geochemistry rests primarily on its intermediate position
between exact sciences (chemistry, physics, mathematics) and natural sciences.
The molding of the quantitative approach taken in physical chemistry, thermody-
namics, mathematics, and analytical chemistry to natural observation offers enor-
mous advantages. These are counterbalanced, however, by the inevitable draw-
backs that have to be faced when writing a textbook on geochemistry: 1) the
need to summarize and apply, very often in a superficial and incomplete fashion,
concepts that would require an entire volume if they were to be described with
sufficient accuracy and completeness; 2) the difficulty of overcoming the diffi-
dence of nature-oriented scientists who consider the application of exact sciences
to natural observations no more than models (in the worst sense of that term).
To overcome these difficulties I gradually introduce the various concepts, first
presenting them in a preliminary way, and then, whenever possible, repeating
them more rigorously and in greater detail.
The organization of this book follows the various states of aggregation of the
earth’s materials, in an order that reflects their relative importance in geology.
Five chapters deal with the crystalline state. The first chapter is preparatory, the
second and third are operative. The fourth summarizes some concepts of defect
chemistry, the role of which in geochemistry is becoming more and more impor-
tant as studies on kinetics and trace element applications advance. The fifth chap-
ter is a (necessarily concise) state-of-the-art appraisal of the major silicate min-
erals.
Two chapters are devoted to silicate melts. One is an introduction to petro-
genetic diagrams, extensively treated in petrology. Aqueous solutions are covered
in a single chapter that basically deals with electrolyte solution theory and its
applications, since any further subdivision seemed unnecessary. A single chapter
was deemed sufficient to describe the up-to-date information about gases. The
decision not to treat chemistry and equilibria in the earth’s atmosphere was dic-
/ xi
xii / Preface
tated by the consideration that such treatment would involve a preliminary evalu-
ation of organic chemistry concepts that could not be adequately covered owing
to lack of space. The last two chapters concern trace element geochemistry and
isotope geochemistry.
As regards trace elements, whose widespread use is due to the simplicity of
operation, I have attempted to show that their simplicity is often only apparent
and that deductions on a planetary scale must be made with caution. Isotope
geochemists may find the last chapter rather concise, but again, space restrictions
did not allow more extended treatment, and I tried to favor the basic concepts,
which should always be borne in mind when dealing with global systematics.
This book is dedicated to my wife Elisabeth and my son Giulio Enrico.
Principles of Geochemistry
PART ONE
Geochemistry of Crystal Phases
CHAPTER ONE
Elemental Properties
and Crystal Chemistry
1. The principal quantum number (n) is related to the energy of the electron
levels. An increase in quantum number n corresponds to an increase in
the distance of the electron from the nucleus.
2. The orbital or “azimuthal” quantum number (l ) defines “form” (i.e., “ec-
centricity of elliptical orbit”; cf. Pauling, 1948) and indicates which sub-
level is occupied by the electron. It assumes integer values between 0 and
n ⫺ 1.
3. The magnetic quantum number (ml ) indicates which orbital of a given
sublevel is occupied by the electron. For a given orbital, quantum number
l assumes integer values between ⫺1 and ⫹1. An electron with a magnetic
quantum number ml subjected to a magnetic field B increases its energy
by a term E ⫽ bmlB, where b is Bohr’s magneton (Zeeman effect).
4. The spin quantum number (ms ) gives the projection of the intrinsic angular
momentum of the electron, with values ⫹12 and ⫺12. Two oppositely ori-
ented spins on the same orbital are graphically represented by arrows in
opposite directions (↑↓).
/ 3
4 / Geochemistry of Crystal Phases
Table 1.1 shows the quantum numbers for the first four levels. Table 1.2 lists
the complete electron configurations of all the elements.
h
λ= , (1.1)
mv
where h is Planck’s constant and v is velocity. Equation 1.1, which is one of the
fundamental results of quantum mechanics, establishes the dual nature of elec-
trons as particles and wavelike elements.
x
ψ = ψ 0 sin 2π − vt , (1.2)
λ
∂ 2ψ 4π
= − 2 ψ. (1.3)
∂ x2 λ
Elemental Properties and Crystal Chemistry / 5
n l ml ms
1 (level K) 0 (sublevel s) 0 ⫹12 ⫺12
2 (level L) 0 (sublevel s) 0 ⫹12 ⫺12
1 (sublevel p) ⫹1 ⫹12 ⫺12
0 ⫹12 ⫺12
⫺1 ⫹12 ⫺12
3 (level M) 0 (sublevel s) 0 ⫹12 ⫺12
1 (sublevel p) ⫹1 ⫹12 ⫺12
0 ⫹12 ⫺12
⫺1 ⫹12 ⫺12
2 (sublevel d ) ⫹2 ⫹12 ⫺12
⫹1 ⫹12 ⫺12
0 ⫹12 ⫺12
⫺1 ⫹12 ⫺12
⫺2 ⫹12 ⫺12
4 (level N) 0 (sublevel s) 0 ⫹12 ⫺12
1 (sublevel p) ⫹1 ⫹12 ⫺12
0 ⫹12 ⫺12
⫺1 ⫹12 ⫺12
2 (sublevel d ) ⫹2 ⫹12 ⫺12
⫹1 ⫹12 ⫺12
0 ⫹12 ⫺12
⫺1 ⫹12 ⫺12
⫺2 ⫹12 ⫺12
3 (sublevel f ) ⫹3 ⫹12 ⫺12
⫹2 ⫹12 ⫺12
⫹1 ⫹12 ⫺12
0 ⫹12 ⫺12
⫺1 ⫹12 ⫺12
⫺2 ⫹12 ⫺12
⫺3 ⫹12 ⫺12
( x2 + y2 + z 2 ) 1 2
ψ = ψ 0 sin 2π − vt , (1.4)
λ
and it can be shown that partial derivation of equation 1.4 over the three dimensions
results in
∂ 2ψ ∂ 2ψ ∂ 2ψ 4π
+ + 2 = − 2 ψ. (1.5)
∂x 2
∂y 2
∂z λ
Let us now consider a particle with mass m, total energy E, and potential
energy ⌽. The kinetic energy of particle Ek will be
6 / Geochemistry of Crystal Phases
1 2 3 4 5 6 7
Z Element s s p s p d s p d f s p d f s p d f s
1 H 1
2 He 2
3 Li 2 1
4 Be 2 2
5 B 2 2 1
6 C 2 2 2
7 N 2 2 3
8 O 2 2 4
9 F 2 2 5
10 Ne 2 2 6
11 Na 2 2 6 1
12 Mg 2 2 6 2
13 Al 2 2 6 2 1
14 Si 2 2 6 2 2
15 P 2 2 6 2 3
16 S 2 2 6 2 4
17 Cl 2 2 6 2 5
18 Ar 2 2 6 2 6
19 K 2 2 6 2 6 1
20 Ca 2 2 6 2 6 2
21 Sc 2 2 6 2 6 1 2
22 Ti 2 2 6 2 6 2 2
23 V 2 2 6 2 6 3 2
24 Cr 2 2 6 2 6 5 1
25 Mn 2 2 6 2 6 5 2
26 Fe 2 2 6 2 6 6 2
27 Co 2 2 6 2 6 7 2
28 Ni 2 2 6 2 6 8 2
29 Cu 2 2 6 2 6 10 1
30 Zn 2 2 6 2 6 10 2
31 Ga 2 2 6 2 6 10 2 1
32 Ge 2 2 6 2 6 10 2 2
33 As 2 2 6 2 6 10 2 3
34 Se 2 2 6 2 6 10 2 4
35 Br 2 2 6 2 6 10 2 5
36 Kr 2 2 6 2 6 10 2 6
37 Rb 2 2 6 2 6 10 2 6 1
38 Sr 2 2 6 2 6 10 2 6 2
39 Y 2 2 6 2 6 10 2 6 1 2
40 Zr 2 2 6 2 6 10 2 6 2 2
41 Nb 2 2 6 2 6 10 2 6 4 1
42 Mo 2 2 6 2 6 10 2 6 5 1
43 Tc 2 2 6 2 6 10 2 6 6 1
44 Ru 2 2 6 2 6 10 2 6 7 1
45 Rh 2 2 6 2 6 10 2 6 8 1
46 Pd 2 2 6 2 6 10 2 6 10
Elemental Properties and Crystal Chemistry / 7
1 2 3 4 5 6 7
Z Element s s p s p d s p d f s p d f s p d f s
47 Ag 2 2 6 2 6 10 2 6 10 1
48 Cd 2 2 6 2 6 10 2 6 10 2
49 In 2 2 6 2 6 10 2 6 10 2 1
50 Sn 2 2 6 2 6 10 2 6 10 2 2
51 Sb 2 2 6 2 6 10 2 6 10 2 3
52 Te 2 2 6 2 6 10 2 6 10 2 4
53 I 2 2 6 2 6 10 2 6 10 2 5
54 Xe 2 2 6 2 6 10 2 6 10 2 6
55 Cs 2 2 6 2 6 10 2 6 10 2 6 1
56 Ba 2 2 6 2 6 10 2 6 10 2 6 2
57 La 2 2 6 2 6 10 2 6 10 2 6 1 2
58 Ce 2 2 6 2 6 10 2 6 10 2 2 6 2
59 Pr 2 2 6 2 6 10 2 6 10 3 2 6 2
60 Nd 2 2 6 2 6 10 2 6 10 4 2 6 2
61 Pm 2 2 6 2 6 10 2 6 10 5 2 6 2
62 Sm 2 2 6 2 6 10 2 6 10 6 2 6 2
63 Eu 2 2 6 2 6 10 2 6 10 7 2 6 2
64 Gd 2 2 6 2 6 10 2 6 10 7 2 6 1 2
65 Tb 2 2 6 2 6 10 2 6 10 9 2 6 2
66 Dy 2 2 6 2 6 10 2 6 10 10 2 6 2
67 Ho 2 2 6 2 6 10 2 6 10 11 2 6 2
68 Er 2 2 6 2 6 10 2 6 10 12 2 6 2
69 Tm 2 2 6 2 6 10 2 6 10 13 2 6 2
70 Yb 2 2 6 2 6 10 2 6 10 14 2 6 2
71 Lu 2 2 6 2 6 10 2 6 10 14 2 6 1 2
72 Hf 2 2 6 2 6 10 2 6 10 14 2 6 2 2
73 Ta 2 2 6 2 6 10 2 6 10 14 2 6 3 2
74 W 2 2 6 2 6 10 2 6 10 14 2 6 4 2
75 Re 2 2 6 2 6 10 2 6 10 14 2 6 5 2
76 Os 2 2 6 2 6 10 2 6 10 14 2 6 6 2
77 Ir 2 2 6 2 6 10 2 6 10 14 2 6 7 2
78 Pt 2 2 6 2 6 10 2 6 10 14 2 6 9 1
79 Au 2 2 6 2 6 10 2 6 10 14 2 6 10 1
80 Hg 2 2 6 2 6 10 2 6 10 14 2 6 10 2
81 Tl 2 2 6 2 6 10 2 6 10 14 2 6 10 2 1
82 Pb 2 2 6 2 6 10 2 6 10 14 2 6 10 2 2
83 Bi 2 2 6 2 6 10 2 6 10 14 2 6 10 2 3
84 Po 2 2 6 2 6 10 2 6 10 14 2 6 10 2 4
85 At 2 2 6 2 6 10 2 6 10 14 2 6 10 2 5
86 Rn 2 2 6 2 6 10 2 6 10 14 2 6 10 2 6
87 Fr 2 2 6 2 6 10 2 6 10 14 2 6 10 2 6 1
88 Ra 2 2 6 2 6 10 2 6 10 14 2 6 10 2 6 2
89 Ac 2 2 6 2 6 10 2 6 10 14 2 6 10 2 6 1 2
90 Th 2 2 6 2 6 10 2 6 10 14 2 6 10 2 6 2 2
91 Pa 2 2 6 2 6 10 2 6 10 14 2 6 10 2 2 6 1 2
92 U 2 2 6 2 6 10 2 6 10 14 2 6 10 3 2 6 1 2
continued
8 / Geochemistry of Crystal Phases
1 2 3 4 5 6 7
Z Element s s p s p d s p d f s p d f s p d f s
93 Np 2 2 6 2 6 10 2 6 10 14 2 6 10 4 2 6 1 2
94 Pu 2 2 6 2 6 10 2 6 10 14 2 6 10 6 2 6 2
95 Am 2 2 6 2 6 10 2 6 10 14 2 6 10 7 2 6 2
96 Cm 2 2 6 2 6 10 2 6 10 14 2 6 10 7 2 6 1 2
97 Bk 2 2 6 2 6 10 2 6 10 14 2 6 10 8 2 6 1 2
98 Cf 2 2 6 2 6 10 2 6 10 14 2 6 10 10 2 6 2
99 Es 2 2 6 2 6 10 2 6 10 14 2 6 10 11 2 6 2
100 Fm 2 2 6 2 6 10 2 6 10 14 2 6 10 12 2 6 2
101 Md 2 2 6 2 6 10 2 6 10 14 2 6 10 13 2 6 2
102 No 2 2 6 2 6 10 2 6 10 14 2 6 10 14 2 6 2
103 Lw 2 2 6 2 6 10 2 6 10 14 2 6 10 14 2 6 1 2
1
EK = mv 2 = E − Φ. (1.6)
2
m2 v 2 = 2 m E − Φ ,( ) (1.7)
1
=
2m E − Φ( ) (1.8)
λ 2
h2
Substituting the right side of equation 1.8 into wave equations 1.3 and 1.5 leads,
respectively, to
d 2ψ 8π 2 m
dx 2
+
h2
E−Φψ =0 ( ) (1.9)
and
8π 2 m
∇2ψ +
h2
(E − Φ ψ = 0, ) (1.10)
∂2 ∂2 ∂2
∇2 = + + (1.11)
∂ x2 ∂ y2 ∂ z2
Elemental Properties and Crystal Chemistry / 9
(see appendix 2). Equation 1.9 is the Schrödinger equation for one-dimensional
stationary states, and equation 1.10 is the tridimensional form. Differential equa-
tions 1.9 and 1.10 possess finite single-valued solutions, within the domain of
squared integrable functions, only for definite values of energy (eigenvalues). For
vibrational systems, such values are given by
1
E ( n ) = n + hv , (1.12)
2
d 2ψ ( x ) 8π 2 m KF x2
+ E − ψ ( x ) = 0. (1.13)
dx 2
h2 2
If position x of the particle corresponds to its displacement from the origin, the restoring
force acting on the particle by virtue of the harmonic potential is given by Hooke’s law:
F = − KF x, (1.14)
dΦ( x )
F =− , (1.15)
dx
KF x2
Φ( x ) = , (1.16)
2
12
K h2 1
E( n ) = F 2 n + . (1.17)
4π m 2
We will later adopt function 1.17 when treating the vibrational energy of crystalline sub-
stances and its relationship to entropy and heat capacity.
10 / Geochemistry of Crystal Phases
The periodic law, discovered by Dmitri Mendeleev, establishes “an harmonic peri-
odicity of properties of chemical individuals, dependent on their masses” (Men-
deleev, 1905). To highlight the existence of periodic properties, the various ele-
ments are arranged, according to their increasing masses, in groups (8),
subgroups (8), and periods (7) (table 1.3). A given group comprises elements with
similar properties. The periodicity of elemental properties is easily understood if
we recall the laws governing the occupancy of atomic orbitals. The periodic table
shows the electron structures of the various elements. For the sake of brevity, the
structure of the inner shells is indicated by referring to that of the preceding inert
gas. An inert gas displays no chemical reactivity (i.e., no tendency to enter into
other states of combination). This “stable” configuration corresponds to comple-
tion of all sublevels of the principal quantum number n. As we can see, a period
begins with the occupancy of a new level; elements in this position are ready to
lose one electron, with the formation of a univalent positive ion. At the end of
the period, immediately preceding the inert gas, the atoms lack a single electron
to complete the stable configuration. Such atoms tend to be transformed into
univalent negative ions by electron capture. The number of electrons that com-
plete the period is readily calculated from the following summation:
n −1
∑ ( 4l + 2 ) = 2 n 2 (1.18)
l =0
—i.e., two electrons in the first period (n ⫽ 1), eight electrons in the second
(n ⫽ 2), 18 electrons in the third (n ⫽ 3), and so on.
The first period is completed by only two elements (H, He), because the first
level has a single sublevel that can be occupied by two electrons with opposite
spin (because n ⫽ 0, l is also zero; ml may vary from ⫺l to ⫹l and is also zero;
ms ⫽ ⫹12 and ⫺12 ). The second period begins with an electron occupying the lowest
energy orbital of sublevel 2s (Li) and terminates with the addition of the sixth
electron of sublevel 2p (Ne). The pile-up procedure is analogous for the third
period, but in the fourth period this regularity is perturbed by the formation of
transition elements (Sc-Zn) and by progressive pile-up of electrons in the 3d orbit-
als, whereas the fourth shell is temporarily occupied by one or two electrons. Pile-
up in the fourth level starts again with Ga, and the period terminates with the
Elemental Properties and Crystal Chemistry / 11
completion of the stable octet (Kr). The same sequence is observed in the fifth
period (Y-Cd transition elements). In the sixth period, the transition is observed
over sublevel f of the fourth shell (4 f ), generating a group of elements (rare
earths, La-Lu) of particular importance in geochemistry for the regularity of pro-
gression of reactive properties within the group (the most external shells are prac-
tically identical for all elements in the series; the progressive pile-up in the inner
fourth shell induces shrinkage of the atomic radius, known as “lanthanidic con-
traction”). In the seventh period, the transition begins with actinium (Ac) and
thorium (Th), whose external electrons occupy sublevel 6d, and continues with
the subsidiary series of protactinium (Pa-Lw), with electrons progressively occu-
pying sublevel f of the fifth shell (5 f ).
Ze 2
ΦC = − , (1.19)
r
where Z is the atomic number and r is the distance from the nucleus expressed in
Cartesian coordinates:
( )
12
r = x2 + y2 + z2 . (1.20)
8π 2 m Ze 2
∇2ψ + E + ψ = 0. (1.21)
h2 r
2π 2 me 4 Z 2
E( n ) = − , (1.22)
n2 h2
where E(n) is the energy of the various orbital levels in the hydrogen ion approxi-
mation.
Table 1.3 Mendeleev’s periodic chart and electron configuration of elements.
14 / Geochemistry of Crystal Phases
Figure 1.1 First ionization potentials as functions of atomic number. Adapted from Har-
vey and Porter (1976).
(Z − S )
2
e Ce 2
Z eff Ce
E = = (1.23)
n2 n2
* The “ionization potential” is the amount of energy that is necessary to subtract an electron
from an atom and to bring it at infinite distance at rest. The “first ionization energy” is defined as
the amount of energy necessary to subtract a valence electron from a neutral atom. The “second
ionization energy” is the amount of energy required to subtract a second electron from an ion from
which a first electron has already been subtracted, and so on. The “total electronic energy” is the
sum of all ionization potentials for a given atom. The “electron affinity” is the amount of energy
required to add an electron to a neutral atom or to an anion. Ionization energies for all atoms are
listed in table 1.6.
Elemental Properties and Crystal Chemistry / 15
serve a progressive increase in the first ionization potential, with minor disconti-
nuities for transition elements (due to the progressive filling of inner shells, which
modifies the value of screen constant Se in eq. 1.23). Recalling the definition of
“valence electron” and considering that bonding energies are related to external
orbitals, the periodicity of the reactive properties of the elements is readily under-
stood in light of equations 1.22 and 1.23.
“There is a chemical bond between two atoms or groups of atoms in the case that
the forces acting between them are such as to lead to the formation of an aggre-
gate with sufficient stability to make it convenient for the chemist to consider it
as an independent molecular species” (Pauling, 1948).
We will consider four distinct types of bonds:
1. ionic bonds
2. covalent bonds
3. van der Waals bonds
4. metallic bonds
Although in nature it is rare to see solid compounds whose bonds are completely
identifiable with one of the types listed above, it is customary to classify solids
according to the dominant type of bond—i.e., ionic, covalent, van der Waals, and
metallic solids.
In ionic solids, the reticular positions are occupied by ions (or anionic
groups) of opposite charge. In NaCl, for instance (figure 1.2A), the Cl ⫺ ions are
surrounded by six Na⫹ ions. The alternation of net charges gives rise to strong
coulombic interactions (see section 1.11.1).
In covalent solids, two bonded atoms share a pair of electrons. Covalent
bonds are particularly stable, because the sharing of two electrons obeys the
double duty of completing a stable electronic configuration for the two neigh-
boring atoms. In diamond, for instance (figure 1.2B), each carbon atom shares
an electron pair with four surrounding carbon atoms, generating a stable giant
molecule that has the dimension of the crystal.
In van der Waals solids (e.g., solid Ne; figure 1.2C), the reticular positions
are occupied by inert atoms held together by van der Waals forces. The van der
Waals bond is rather weak, and the bond does not survive a high thermal energy
(Fm3m solid Ne, for instance, is stable below 20 K). The van der Waals bond also
exists, however, in molecular compounds (e.g., solid CO2 ). The reticular positions
in this case are occupied by neutral molecules (CO2 ). Within each molecule, the
bond is mainly covalent (and interatomic distances are considerably shorter), but
16 / Geochemistry of Crystal Phases
Figure 1.2 Various types of solids. For each solid, a sketch of the bond is given.
the various molecules are held together by van der Waals forces whose nature will
be discussed later.
In metallic solids, the reticular positions are occupied by cations immersed in
a cloud of delocalized valence electrons. In solid Na, for instance (figure 1.2D),
the electron cloud or “electron gas” is composed of electrons of the s sublevel of
the third shell (cf. table 1.2). Note that the type of bond does not limit the crystal
structure of the solid within particular systems. For example, all solids shown in
figure 1.2 belong to the cubic system, and two of them (NaCl, Ne) belong to the
same structural class (Fm3m).
The concept of “ionic radius” was derived from the observation of crystal struc-
tures understood as “regular arrays of ions of spherical undeformable symmetry.”
The structure of ionic crystals obeys pure geometrical rules and corresponds to
Elemental Properties and Crystal Chemistry / 17
Figure 1.3 Interatomic distances in alkali halides, arranged in cationic and anionic series.
di = r+ + r− , (1.24)
where r⫹ and r⫺ are, respectively, the ionic radii of a cation and an anion at the
interionic distance observed in a purely ionic crystal.
The concept that the ionic radius is relatively independent of the structure of
the solid arose intuitively from experimental observations carried out on alkali
halides, which are ionic solids par excellence. Figure 1.3 shows the evolution of
interatomic distances in alkali halides as a function of the types of anion and
cation, respectively. Significant parallelism within each of the two families of
curves may be noted. This parallelism intuitively generates the concept of con-
stancy of the ionic radius.
The first attempt to derive empirical ionic radii by geometrical rules applied
to ionic compounds (and metals) was performed by Bragg (1920), who proposed
a set of ionic radii with a mean internal precision of 0.006 Å. Further tabulations
18 / Geochemistry of Crystal Phases
Figure 1.4 Electron density (in electrons per Å3) in the (100) plane of NaCl. Reprinted
from Witte and Wölfel (1955), with kind permission of R. Oldenbourg Verlag GmbH,
München, Germany.
NaCl thus correspond to the radii (radial extensions) of the two elements in the
condition of “ionic bond” when there is complete transfer of one electron from
Na to Cl.
The first attempt to derive ionic radii independently of the geometrical concept
of simple additivity was made by Pauling (1927a). His method was based on the
assumption that the dimension of an ion is determined by the distribution of its
outer electrons. As we have already seen (eq. 1.23), the “effective nuclear charge”
of an ion (Zeff ) is lower than its atomic number, due to the shielding effect of
inner electrons, expressed by screening constant Se . Values of Se may be obtained
theoretically (Pauling, 1927b), and partially by results of X-ray diffraction studies
(Pauling and Sherman, 1932). Based on such values, the univalent ionic radius
(r⫹⫺ ) can be expressed as
Cn C
r+ − = = n , (1.25)
Z − Se Z eff
rF − Z eff,Na + 6.48
= = , (1.26)
rNa + Z eff,F − 4.48
which gives 1.36 Å for F⫺ and 0.95 Å for Na⫹. If we now continue to apply
equation 1.25 to multivalent ions, assuming that Cn remains unchanged at the
value of the intermediate inert gas, we obtain values for all ions with He, Ne, Ar,
Kr, and Xe structures. These radii, known as “univalent ionic radii in VI-fold
co-ordination,” “do not have absolute values such that their sums are equal to
equilibrium inter ionic distances” (Pauling, 1960) but rather “are the radii the
20 / Geochemistry of Crystal Phases
multivalent ions would possess if they were to retain their electron distributions
but enter into Coulomb interaction as if they were univalent” (Pauling, 1960).
Univalent radii may be opportunely modified to obtain effective ionic radii in real
structures (Pauling, 1960; Huheey, 1975):
rZ = r+ − z (
−2 n −1 ), (1.28)
where n is the Born exponent (cf. section 1.11.2) and z is the ionic charge. Ionic
radii in VI-fold coordination derived by Huheey (1975) are listed in table 1.4.
both elements, regardless of the state of aggregation of the compound (i.e., solid,
liquid, or gaseous). Covalent radii are easily calculated by dividing by 2 the in-
teratomic distances observed in homopolar covalent compounds (e.g., P4, S8 ,
etc.). An internally consistent tabulation was derived by Sanderson (1960) from
interatomic distances in heteropolar covalent gaseous molecules (e.g., AlCl3 ,
CH4 , SiF4, etc.). The covalent radii of Sanderson (table 1.5) are defined as “non-
polar,” because they are obtained from compounds that do not have charge po-
larity.
χ A − χ B = γ 1 2 E AB −
1
2
(
E AA + E BB ,
) (1.29)
where A and B are the electronegativities of atoms A and B, EAB, EAA, and EBB
are the binding energies of compounds AB, AA, and BB, and operative constant
␥ is a term that constrains the adimensional A and B to change by 0.5 with each
unit valence change in the first row of the periodic table ( ␥1/2 ⫽ 0.208).
According to Mulliken (1934, 1935), electronegativity is the algebraic mean
of the first ionisation potential and of electron affinity.
According to Gordy (1946), electronegativity is represented by the value of
the potential resulting from the effect of the nuclear charge of an unshielded atom
on a valence electron located at a distance corresponding to the covalent radius
of the atom.
According to Sanderson (1960), electronegativity represents the mean elec-
tron density per unit volume of an inert element:
3Z
De = , (1.30)
4π rc3
∂W
χ( z) = . (1.31)
∂z
1
( )
2
f i = 1 − exp − χ A − χB . (1.32)
4
fi =
1
2
(
χ A − χB . ) (1.33)
Hinze et al. (1963) suggest that charge transfer in diatomic molecules obeys
the principle of equality of energies involved in the process. If we adopt equation
1.31 as a definition of electronegativity and imagine that the amount of energy
involved in the transfer of a fraction of charge dz from atom A to atom B is
quantifiable as
∂WA
dEA = dz , (1.34)
∂z z
A
∂WA ∂WB
dEA = dz = dz = dE B . (1.35)
∂z z ∂z z
A B
At present, the most widely used definition of fractional ionic character in solid state
chemistry is that of Phillips (1970), based on a spectroscopic approach. Phillips defined
fractional ionic character as
Ei2
fi = , (1.36)
Ei2 + E c2
where Ei (“ionic energy gap”) is obtained from the “total energy gap” (Eg ) between bond-
ing and antibonding orbitals (see section 1.17.1):
បωP
Eg =
(ε − 1) (1.37)
12
∞
( )
12
Ei = E g2 − E c2 , (1.38)
where បP is the plasma frequency for valence electrons, ∞ is the optic dielectric constant,
and Ec corresponds to Eg for the nonpolar system in the same row of the periodic table,
with a correction for interatomic spacing.
Table 1.6 Elemental properties: first four ionization potentials (I1, I2, I3, I4, from
Samsonov, 1968; values expressed in eV; values in parentheses are of doubtful reliability);
electron affinity (e.a., from Samsonov, 1968; eV); Pauling’s electronegativity (P, from
Samsonov, 1968; eV); Sanderson’s electronegativity (S, from Viellard, 1982; adimensional).
Z Atom I1 I2 I3 I4 e.a. P S
1 H 13.595 – – – 0.747 2.15 3.55
2 He 24.58 54.40 – – ⫺0.53 – –
3 Li 5.39 75.62 122.42 – 0.82 1.0 0.74
4 Be 9.32 18.21 153.85 217.66 ⫺0.19 1.5 2.39
5 B 8.296 25.15 37.92 259.30 0.33 2.0 2.84
6 C 11.264 24.376 47.86 64.48 1.24 2.5 3.79
7 N 14.54 29.60 47.426 77.45 0.05 3.0 4.49
8 O 13.614 35.15 54.93 77.39 1.465 3.5 5.21
9 F 17.418 34.98 62.65 87.23 3.58 4.0 5.75
10 Ne 21.559 41.07 63.5 97.16 ⫺0.57 – –
11 Na 5.138 47.29 71.8 98.88 0.84 0.9 0.70
12 Mg 7.644 15.03 78.2 109.3 ⫺0.32 1.2 1.99
13 Al 5.984 18.82 28.44 119.96 0.52 1.5 2.25
14 Si 8.149 16.34 33.46 45.13 1.46 1.8 2.62
15 P 10.55 19.65 30.16 51.35 0.77 2.1 3.34
16 S 10.357 23.4 34.8 47.29 2.07 2.5 4.11
17 Cl 13.01 23.80 39.9 53.3 3.76 3.0 4.93
18 Ar 15.755 27.6 40.90 59.79 ⫺1.0 – –
19 K 4.339 31.81 45.9 61.1 0.82 0.8 0.41
20 Ca 6.111 11.87 51.21 67.3 – 1.0 1.22
21 Sc 6.56 12.89 24.75 73.9 – 1.3 1.30
22 Ti 6.83 13.57 28.14 43.24 – 1.5 1.40
23 V 6.74 14.2 29.7 48.0 – 1.6 1.60
24 Cr 6.764 16.49 31 (51) – 1.6 1.88
25 Mn 7.432 15.64 33.69 (53) – 1.5 2.07
26 Fe 7.90 16.18 30.64 (56) – 1.8 2.10
27 Co 7.86 17.05 33.49 (53) – 1.7 2.10
28 Ni 7.633 18.15 36.16 (56) – 1.8 2.10
29 Cu 7.724 20.29 36.83 (59) 0.29 1.9 2.60
30 Zn 9.391 17.96 39.70 (62) – 1.6 2.84
31 Ga 6.00 20.51 30.70 64.2 – 1.6 3.23
32 Ge 7.88 15.93 34.21 45.7 – 2.0 3.59
33 As 9.81 18.7 28.3 50.1 – 2.0 3.91
34 Se 9.75 21.5 32.0 42.9 3.7 2.4 4.25
35 Br 11.84 21.6 35.9 47.3 3.54 2.9 4.53
36 Kr 13.996 24.56 36.9 52.5 – – –
37 Rb 4.176 27.56 40 52.6 – 0.8 0.33
38 Sr 5.692 11.026 43.6 57.1 – 1.0 1.00
39 Y 6.38 12.23 20.5 61.8 – 1.2 1.05
40 Zr 6.835 12.92 24.8 33.97 – 1.4 1.10
41 Nb 6.88 13.90 28.1 38.3 – 1.6 –
42 Mo 7.131 15.72 29.6 46.4 – 1.8 –
43 Tc 7.23 14.87 31.9 (43) – 1.9 –
44 Ru 7.36 16.60 30.3 (47) – 2.2 –
Table 1.6 (continued)
Z Atom I1 I2 I3 I4 e.a. P S
45 Rh 7.46 15.92 32.8 (46) – 2.2 –
46 Pd 8.33 19.42 (33) (49) – 2.2 –
47 Ag 7.574 21.48 36.10 (52) 2.5 1.9 2.57
48 Cd 8.991 16.904 44.5 (55) – 1.7 2.59
49 In 5.785 18.86 28.0 68 – 1.7 2.86
50 Sn 7.332 14.6 30.7 46.4 – 1.8 3.10
51 Sb 8.64 16.7 24.8 44.1 ⬎2.0 1.9 3.37
52 Te 9.01 18.8 31.0 38 3.6 2.1 3.62
53 I 10.44 19.0 33 (42) 3.29 2.5 3.84
54 Xe 12.127 21.2 32.1 (45) – – –
55 Cs 3.893 25.1 34.6 (46) – 0.7 0.29
56 Ba 5.810 10.00 37 (49) – 0.9 0.78
57 La 5.614 11.433 19.166 (52) – 1.1 0.88
58 Ce 6.54 12.31 19.870 36.7 – 1.1 –
59 Pr (5.76) (11.54) (20.96) – – 1.1 –
60 Nd (6.31) (12.09) (20.51) – – 1.1 –
62 Sm 5.56 11.4 (24.0) – – 1.1 –
63 Eu 5.61 11.24 (24.56) – – 1.1 –
64 Gd 5.98 (12) (23) – – 1.1 –
65 Tb (6.74) (12.52) (22.04) – – 1.2 –
66 Dy 5.80 (12.60) (21.83) – – 1.2 –
67 Ho 5.85 (12.7) (22.1) – – 1.2 –
68 Er 6.11 (12.5) (23.0) – – 1.2 –
69 Tm 5.85 (12.4) (24.1) – – 1.2 –
70 Yb 5.90 12.10 (25.61) – – 1.2 –
71 Lu 6.15 14.7 (21.83) – – 1.2 –
72 Hf 5.5 14.9 (21) (31) – 1.3 1.05
73 Ta 7.7 16.2 (22) (33) – 1.5 –
74 W 7.98 17.7 (24) (35) – 1.7 1.39
75 Re 7.87 16.6 (26) (38) – 1.9 –
76 Os 8.7 17 (25) (40) – 2.2 –
77 Ir 9.2 17.0 (27) (39) – 2.2 –
78 Pt 8.96 18.54 (29) (41) – 2.2 –
79 Au 9.223 20.5 (30) (44) 2.1 2.4 2.57
80 Hg 10.434 18.751 34.2 (46) 1.54 1.9 2.93
81 Tl 6.106 20.42 29.8 50 2.1 1.8 3.02
82 Pb 7.415 15.03 31.93 39.0 – 1.8 3.08
83 Bi 7.277 19.3 25.6 45.3 ⬎0.7 1.9 3.16
84 Po 8.2 19.4 27.3 (38) – 2.0 –
85 At 9.2 20.1 29.3 (41) – 2.2 –
86 Rn 10.745 21.4 29.4 (44) – – –
87 Fr 3.98 22.5 33.5 (43) – 0.7 –
88 Ra 5.277 10.144 (34) (46) – 0.9 –
89 Ac 6.89 11.5 – (49) – 1.1 –
90 Th 6.95 11.5 20.0 28.7 – 1.3 –
91 Pa – – – – – 1.5 –
92 U 5.65 14.63 25.13 – – 1.7 –
26 / Geochemistry of Crystal Phases
Catlow and Stoneham (1983) have shown that ionic term Ei corresponds to the
difference of the diagonal matrix elements of the Hamiltonian of an AB molecule in a
simple LCAO approximation (HAA ⫺ HBB; cf. section 1.18.1), whereas the covalent energy
gap corresponds to the double of the off-diagonal term HAB —i.e.,
(H − HBB ) + 4HAB .
2
Eg = AA (1.39)
µ d = Z e d = Fα . (1.40)
Let us consider two neighboring atoms A and B (figure 1.5) and imagine that
the electronic charge distribution is, for an infinitesimal period of time, asymmet-
rical. In this period of time, atom A becomes a dipole and induces a distortion
in the electron cloud of atom B by attracting electrons from its positive edge:
atom B is then transformed, in turn, into a dipole. All the neighboring atoms
undergo the same process, generating attractive forces among themselves. Due to
electron orbital motion, the induced forces fluctuate with a periodicity dictated
by the rotational velocity of electrons. The van der Waals bond is the result of the
induced attractions between fluctuating dipoles. The attractions diminish with
increased molecular distance and, when distances are equal, increase with the
number of electrons within the molecule.
Free-ion polarizability calculations have been proposed by various authors,
based on the values of molecular refraction of incident light (Rm ), according to
3Rm
α = , (1.41)
4πN 0
Figure 1.6 Free-ion polarizability as a function of atomic number. Curves are drawn for
isoelectronic series. Reprinted from Viellard (1982), Sciences Geologiques, Memoir n°69,
Université Louis Pasteur, with kind permission of the Director of Publication.
28 / Geochemistry of Crystal Phases
Table 1.7 Free–Ion polarizability (␣f ) arranged in isoelectronic series. Data in Å3.
N ⫽ number of electrons (adapted from Viellard, 1982).
Th4⫹ 86 1.75
U6⫹ 86 1.141
The first satisfactory definition of crystal radius was given by Tosi (1964): “In an
ideal ionic crystal where every valence electron is supposed to remain localised
on its parent ion, to each ion it can be associated a limit at which the wave func-
tion vanishes. The radial extension of the ion along the connection with its first
neighbour can be considered as a measure of its dimension in the crystal (crystal
radius).” This concept is clearly displayed in figure 1.7A, in which the radial elec-
tron density distribution curves are shown for Na⫹ and Cl ⫺ ions in NaCl. The
nucleus of Cl ⫺ is located at the origin on the abscissa axis and the nucleus of Na⫹
is positioned at the interionic distance experimentally observed for neighboring
ions in NaCl. The superimposed radial density functions define an electron den-
sity minimum that limits the dimensions or “crystal radii” of the two ions. We
also note that the radial distribution functions for the two ions in the crystal
(continuous lines) are not identical to the radial distribution functions for the
free ions (dashed lines).
The “crystal radius” thus has local validity in reference to a given crystal
structure. This fact gives rise to a certain amount of confusion in current nomen-
clature, and what it is commonly referred to as “crystal radius” in the various
tabulations is in fact a mean value, independent of the type of structure (see
section 1.11.1). The “crystal radius” in the sense of Tosi (1964) is commonly
defined as “effective distribution radius” (EDR). The example given in figure 1.7B
shows radial electron density distribution curves for Mg, Ni, Co, Fe, and Mn on
the M1 site in olivine (orthorhombic orthosilicate) and the corresponding EDR
radii located by Fujino et al. (1981) on the electron density minima.
Table 1.8 shows the differences among EDR radii, ionic radii, and crystal
radii for selected elements.
It also must be stressed that EDR may assume different values along various crystal-
lographic directions. This is clearly evident if we compare electron density maps obtained
30 / Geochemistry of Crystal Phases
Figure 1.7 (A) Radial electron density distribution curves for Na⫹ and Cl ⫺ in NaCl.
(B) Radial electron density distributions for Mg, Ni, Co, Fe, and Mn on M1 site in olivine.
Arrows indicate electron density minima defining EDR radii. Reprinted from Tosi (1964),
with kind permission of Academic Press Inc., Orlando, Florida (A); and from Fujino
et al. (1981), with kind permission of The International Union of Crystallography (B).
Table 1.8 Comparison of EDR radii (Fujino et al., 1981) with ionic radii
(IR) and crystal radii (CR) (Shannon, 1976) in various structures (adapted
from Fujino et al., 1981; values in Å).
Figure 1.8 Electron density plots along (110) plane of BeO: (A) effective total electron
density (pseudoatom approximation); (B) total electron density of IAM; (C) deformation
density (pseudoatom-IAM). From Downs (1992). Reprinted with permission of Springer-
Verlag, New York.
32 / Geochemistry of Crystal Phases
1
I ∝ (1.42)
r
and
1
r∝ (1.43)
I2
I n ∝ Zc , (1.44)
Figure 1.9 Ionic radii of rare earth ions (and In3⫹, Y3⫹, and Sc3⫹) as a function of coordi-
nation number with oxygen vc. Reprinted from Shannon (1976), with kind permission
from the International Union of Crystallography.
(table 1.4). Ahrens’s (1952) tabulation was a common tool for geochemists until
the 1970s, when the extended tabulations of Shannon and Prewitt (1969), Shan-
non (1976), and Whittaker and Muntus (1970) appeared.
∑ Sl = Zc .
(1.45)
P
This last principle is not strictly observed in natural compounds, and, particularly
in silicates (Baur, 1970), deviations of up to 40% over theoretical values have been
observed. These deviations must be attributed to partial covalence of the bonds.
34 / Geochemistry of Crystal Phases
Pauling (1947) has also shown that, in purely covalent bonds, modification of
bond distance within a given coordination polyhedron may be associated with
the “bond number” (nl ; number of shared electrons per bond) through arbitrary
constant K l —i.e.,
Nl
R
Sl = S 0,l l , (1.47)
Rl
where Sl is the strength of a bond with length R l , S0,l is the ideal bond strength
of a bond of length R l , coinciding with the mean of bond distances within the
coordination polyhedron, and Nl is a constant that has a single value for each
cation-anion pair and, in some instances, for each single coordination number
within a fixed cation-anion pair.
A family of empirical covariance curves between bond strength and bond
length was then derived by Brown and Shannon (1973) on the basis of equation
1.47 coupled with analysis of a large number of crystal structures. In solving the
system by means of statistical interpolation procedures, Brown and Shannon kept
the bond strength sum for a given polyhedron as near as possible to the theoreti-
cal value postulated by Pauling’s (1929) “electrostatic valence principle” (cf. eq.
1.45). They also showed that the degree of covalence for a given bond and its
relative bond strength are related by an empirical equation in the form
f cl = α l SlM l , (1.48)
where ␣l and Ml depend on the number of electrons of the cation and are constant
for isoelectronic series. The ␣l and Ml constants for cation-oxygen bonds are listed
in table 1.9.
Figure 1.10 shows the relationship between bond strength and degree of cova-
lence in cation-oxygen bonds for cations with 18, 36, and 54 electrons (note that
in table 1.9 the parameters of equation 1.48 are identical for the three isoelectronic
series). By combining equation 1.48 with a modified form of Donnay’s (1969)
equation:
Elemental Properties and Crystal Chemistry / 35
e ␣l Ml MlNl
0 0.67 1.8–2 4
2 0.60 1.73 7.04
10 0.54 1.64 7.04
18 0.49 1.57 7.04
28 0.60 1.50 9.08
36 0.49 1.57
46 0.67 1.43
54 0.49 1.57
Cation e R1,l Nl
H 0 0.86 2.17
Li, Be, B 2 1.378 4.065
Na, Mg, Al, Si, P, S 10 1.622 4.290
K, Ca, Sc, Ti, V, Cr 18 1.799 4.483
Mn, Fe 23 1.760 5.117
Zn, Ga, Ge, As 28 1.746 6.050
−Nl
R
S0 , l = l , (1.49)
R1,l
where R1,l and Nl are general parameters valid for isoelectronic series (with the
same anion; see table 1.10), Brown and Shannon obtained an empirical equation
relating bond length and degree of covalence:
−Ml Nl
R
f cl = αl l . (1.50)
R1,l
36 / Geochemistry of Crystal Phases
Figure 1.10 Degree of covalence ( f cl ) vs bond strength (So,l) in M-O bonds for cations
with 18, 36, and 54 electrons. Values are in valence units and scales are logarithmic. Re-
printed from Brown and Shannon (1973), with kind permission from the International
Union of Crystallography.
Ion CN SP CR IR Ion CN SP CR IR
Ac3⫹ VI 1.26 1.12 Bi5⫹ VI 0.90 0.76
Ag1⫹ II 0.81 0.67 Bk3⫹ VI 1.10 0.96
IV 1.14 1.00 Bk4⫹ VI 0.97 0.83
IVsp 1.16 1.02 VIII 1.07 0.93
V 1.23 1.09 Br1⫺ VI 1.82 1.96
VI 1.29 1.15 Br3⫹ IVsp 0.73 0.59
VII 1.36 1.22 Br5⫹ IIIpy 0.45 0.31
VIII 1.42 1.28 Br7⫹ IV 0.39 0.25
Ag2⫹ IVsp 0.93 0.79 VI 0.53 0.39
VI 1.08 0.94 C4⫹ III 0.06 ⫺0.08
3⫹
Ag IVsp 0.81 0.67 IV 0.29 0.15
VI 0.89 0.75 VI 0.30 0.16
Al3⫹ IV 0.53 0.39 Ca2⫹ VI 1.14 1.00
V 0.62 0.48 VII 1.20 1.06
VI 0.675 0.535 VIII 1.26 1.12
Am2⫹ VII 1.35 1.21 IX 1.32 1.18
VIII 1.40 1.26 X 1.37 1.23
IX 1.45 1.31 XII 1.48 1.34
Am3⫹ VI 1.115 0.975 Cd2⫹ IV 0.92 0.78
VIII 1.23 1.09 V 1.01 0.87
Am4⫹ VI 0.99 0.85 VI 1.09 0.95
VIII 1.09 0.95 VII 1.17 1.03
As3⫹ VI 0.72 0.58 VIII 1.24 1.10
As5⫹ IV 0.475 0.335 XII 1.45 1.31
VI 0.60 0.46 Ce3⫹ VI 1.15 1.01
At7⫹ VI 0.76 0.62 VII 1.21 1.07
Au1⫹ VI 1.51 1.37 VIII 1.283 1.143
Au3⫹ IVsp 0.82 0.68 IX 1.336 1.196
VI 0.99 0.85 X 1.39 1.25
Au5⫹ VI 0.71 0.57 XII 1.48 1.34
B3⫹ III 0.15 0.01 Ce4⫹ VI 1.01 0.87
IV 0.25 0.11 VIII 1.11 0.97
VI 0.41 0.27 X 1.21 1.07
Ba2⫹ VI 1.49 1.35 XII 1.28 1.14
VII 1.52 1.38 Cf3⫹ VI 1.09 0.95
VIII 1.56 1.42 Cf4⫹ VI 0.961 0.821
IX 1.61 1.47 VIII 1.06 0.92
X 1.66 1.52 Cl1⫺ VI 1.67 1.81
XI 1.71 1.57 Cl5⫹ IIIpy 0.26 0.12
XII 1.75 1.61 Cl7⫹ IV 0.22 0.08
Be2⫹ III 0.30 0.16 VI 0.41 0.27
IV 0.41 0.27 Cm3⫹ VI 1.11 0.97
VI 0.59 0.45 Cm4⫹ VI 0.99 0.85
Bi3⫹ V 1.10 0.96 VIII 1.09 0.95
VI 1.17 1.03 Co2⫹ IV HS 0.72 0.58
VIII 1.31 1.17 V 0.81 0.67
continued
Table 1.11 (continued)
Ion CN SP CR IR Ion CN SP CR IR
VI LS 0.79 0.65 Eu3⫹ VI 1.087 0.947
VI HS 0.885 0.745 VII 1.15 1.01
VIII 1.04 0.90 VIII 1.206 1.066
Co3⫹ VI LS 0.685 0.545 IX 1.260 1.120
HS 0.75 0.61 F1⫺ II 1.145 1.285
Co4⫹ IV 0.54 0.40 III 1.16 1.30
Cr2⫹ VI LS 0.87 0.73 IV 1.17 1.31
VI HS 0.94 0.80 VI 1.19 1.33
Cr3⫹ VI 0.755 0.615 F7⫹ VI 0.22 0.08
Cr4⫹ IV 0.55 0.41 Fe2⫹ IV HS 0.77 0.63
V 0.72 0.58 IVsp HS 0.78 0.64
VI LS 0.69 0.55 VI 0.75 0.61
HS 0.785 0.645 HS 0.920 0.780
VIII 0.92 0.78 VIII HS 1.06 0.92
VI LS 0.69 0.55 Fe3⫹ IV HS 0.63 0.49
5⫹
Cr IV 0.485 0.345 V 0.72 0.58
VI 0.63 0.49 VI LS 0.69 0.55
VIII 0.71 0.57 HS 0.785 0.645
Cr6⫹ IV 0.40 0.26 VIII HS 0.92 0.78
VI 0.58 0.44 Fe4⫹ VI 0.725 0.585
Cs1⫹ VI 1.81 1.67 Fe6⫹ IV 0.39 0.25
VIII 1.88 1.74 Fr1⫹ VI 1.94 1.80
IX 1.92 1.78 Ga3⫹ IV 0.61 0.47
X 1.95 1.81 V 0.69 0.55
XI 1.99 1.85 VI 0.760 0.620
XII 2.02 1.88 VII 1.14 1.00
Cu1⫹ II 0.60 0.46 Gd3⫹ VI 1.078 0.938
IV 0.74 0.60 VIII 1.193 1.053
VI 0.91 0.77 IX 1.247 1.107
Cu2⫹ IVsp 0.71 0.57 Ge2⫹ VI 0.87 0.73
V 0.79 0.65 Ge4⫹ IV 0.530 0.390
VI 0.87 0.73 VI 0.670 0.530
Cu3⫹ VI LS 0.68 0.54 H1⫹ I ⫺0.24 ⫺0.38
D1⫹ II 0.04 0.10 II ⫺0.04 ⫺0.08
Dy2⫹ VI 1.21 1.07 Hf 4⫹ IV 0.72 0.58
VII 1.27 1.13 VI 0.85 0.71
VIII 1.33 1.19 VII 0.90 0.76
Dy3⫹ VI 1.052 0.912 VIII 0.97 0.83
VII 1.11 0.97 Hg1⫹ III 1.11 0.97
VIII 1.167 1.027 VI 1.33 1.19
IX 1.223 1.083 Hg2⫹ II 0.83 0.69
3⫹
Er VI 1.030 0.890 IV 1.10 0.96
VII 1.085 0.945 VI 1.16 1.02
VIII 1.144 1.004 VIII 1.28 1.14
Eu2⫹ VI 1.31 1.17 Ho3⫹ VI 1.041 0.901
VII 1.34 1.20 VIII 1.155 1.015
VIII 1.39 1.25 IX 1.212 1.072
IX 1.44 1.30 X 1.26 1.12
X 1.49 1.35 I1⫺ VI 2.06 2.20
Table 1.11 (continued)
Ion CN SP CR IR Ion CN SP CR IR
I5⫹ IIIpy 0.58 0.44 Mo5⫹ IV 0.60 0.46
VI 1.09 0.95 VI 0.75 0.61
I7⫹ IV 0.56 0.42 Mo6⫹ IV 0.55 0.41
VI 0.67 0.53 V 0.64 0.50
In3⫹ IV 0.76 0.62 VI 0.73 0.59
VI 0.940 0.800 VII 0.87 0.73
VIII 1.06 0.92 N3⫺ IV 1.32 1.46
3⫹
Ir VI 0.82 0.68 N3⫹ VI 0.30 0.16
Ir4⫹ VI 0.765 0.625 N5⫹ III 0.044 0.104
Ir5⫹ VI 0.71 0.57 IV 0.27 0.13
K1⫹ IV 1.51 1.37 Na1⫹ IV 1.13 0.99
VII 1.60 1.46 V 1.14 1.00
VIII 1.65 1.51 VI 1.16 1.02
IX 1.69 1.55 VII 1.26 1.12
X 1.73 1.59 VIII 1.32 1.18
XII 1.78 1.64 IX 1.38 1.24
La3⫹ VI 1.172 1.032 XII 1.53 1.39
VII 1.24 1.10 Nb2⫹ VI 0.86 0.72
VIII 1.300 1.160 Nb4⫹ VI 0.82 0.68
IX 1.356 1.216 VIII 0.93 0.79
X 1.41 1.27 Nb5⫹ IV 0.62 0.48
XII 1.50 1.36 VI 0.78 0.64
Li1⫹ IV 0.730 0.590 VII 0.83 0.69
VI 0.90 0.76 VIII 0.88 0.74
VIII 1.06 0.91 Nd2⫹ VIII 1.43 1.29
3⫹
Lu VI 1.001 0.861 IX 1.49 1.35
VIII 1.117 0.977 Nd3⫹ VI 1.123 0.983
IX 1.172 1.032 VIII 1.249 1.109
Mg2⫹ IV 0.71 0.57 IX 1.303 1.163
V 0.80 0.66 XII 1.41 1.27
VI 0.860 0.720 Ni2⫹ IV 0.69 0.55
VIII 1.03 0.89 IVsp 0.63 0.49
Mn2⫹ IV HS 0.80 0.66 V 0.77 0.63
V HS 0.89 0.75 VI 0.830 0.690
VI LS 0.81 0.67 Ni3⫹ VI LS 0.70 0.56
HS 0.970 0.830 HS 0.74 0.60
VII HS 1.04 0.90 Np2⫹ VI 1.24 1.10
VIII 1.10 0.96 Np3⫹ VI 1.15 1.01
Mn3⫹ V 0.72 0.58 Np4⫹ VI 1.01 0.87
VI LS 0.72 0.58 VIII 1.12 0.98
HS 0.785 0.645 Np5⫹ VI 0.89 0.75
4⫹
Mn IV 0.53 0.39 Np6⫹ VI 0.86 0.72
VI 0.670 0.530 Np7⫹ VI 0.85 0.71
Mn5⫹ IV 0.47 0.33 O2⫺ II 1.21 1.35
Mn6⫹ IV 0.395 0.255 III 1.22 1.36
Mn7⫹ IV 0.39 0.25 IV 1.24 1.38
VI 0.60 0.46 VI 1.26 1.40
Mo3⫹ VI 0.83 0.69 VIII 1.28 1.42
Mo4⫹ VI 0.790 0.650 OH1⫺ II 1.18 1.32
continued
Table 1.11 (continued)
Ion CN SP CR IR Ion CN SP CR IR
III 1.20 1.34 Pu3⫹ VI 1.14 1.00
IV 1.21 1.35 Pu4⫹ VI 0.94 0.80
VI 1.23 1.37 VIII 1.10 0.96
Os4⫹ VI 0.770 0.630 Ra2⫹ VIII 1.62 1.48
Os5⫹ VI 0.715 0.575 XII 1.84 1.70
Os6⫹ V 0.63 0.49 Rb1⫹ VI 1.66 1.52
VI 0.685 0.545 VII 1.70 1.56
Os7⫹ VI 0.665 0.525 VIII 1.75 1.61
Os8⫹ IV 0.53 0.39 IX 1.77 1.63
P3⫹ VI 0.58 0.44 X 1.80 1.66
P5⫹ IV 0.31 0.17 XI 1.83 1.69
V 0.43 0.29 XII 1.86 1.72
VI 0.52 0.38 XIV 1.97 1.93
Pa3⫹ VI 1.18 1.04 Re4⫹ VI 0.77 0.63
Pa4⫹ VI 1.04 0.90 Re5⫹ VI 0.72 0.58
VIII 1.15 1.01 Re7⫹ IV 0.52 0.38
Pa5⫹ VI 0.92 0.78 VI 0.67 0.53
IX 1.09 0.95 Rh3⫹ VI 0.805 0.665
2⫹
Pb IVpy 1.12 0.98 Rh4⫹ VI 0.74 0.60
VI 1.33 1.19 Rh5⫹ VI 0.69 0.55
VII 1.37 1.23 Ru3⫹ VI 0.82 0.68
VIII 1.43 1.29 Ru4⫹ VI 0.760 0.620
IX 1.49 1.35 Ru5⫹ VI 0.705 0.565
X 1.54 1.40 Ru7⫹ IV 0.52 0.38
XI 1.59 1.45 Ru8⫹ IV 0.50 0.36
XII 1.63 1.49 S2⫺ VI 1.70 1.84
Pb4⫹ IV 0.79 0.65 S4⫹ VI 0.51 0.37
V 0.87 0.73 S6⫹ IV 0.26 0.12
VI 0.915 0.775 VI 0.43 0.29
VIII 1.08 0.94 Sb3⫹ IVpy 0.90 0.76
1⫹
Pd II 0.73 0.59 V 0.94 0.80
Pd2⫹ IVsp 0.78 0.64 VI 0.90 0.76
VI 1.00 0.86 Sb5⫹ VI 0.74 0.60
3⫹
Pd VI 0.90 0.76 Sc3⫹ VI 0.885 0.745
Pd4⫹ VI 0.755 0.615 VIII 1.010 0.870
Pm3⫹ VI 1.11 0.97 Se2⫺ VI 1.84 1.98
VIII 1.233 1.093 Se4⫹ VI 0.64 0.50
IX 1.284 1.144 Se6⫹ IV 0.42 0.28
Po4⫹ VI 1.08 0.94 VI 0.56 0.42
VIII 1.22 1.08 Si4⫹ IV 0.40 0.26
6⫹
Po VI 0.81 0.67 VI 0.540 0.400
Pr3⫹ VI 1.13 0.99 Sm2⫹ VII 1.36 1.22
VIII 1.266 1.126 VIII 1.41 1.27
IX 1.319 1.179 IX 1.46 1.32
Pr4⫹ VI 0.99 0.85 Sm3⫹ VI 1.098 0.958
VIII 1.10 0.96 VII 1.16 1.02
Pt2⫹ IVsp 0.74 0.60 VIII 1.219 1.079
VI 0.94 0.80 IX 1.272 1.132
Pt4⫹ VI 0.765 0.625 XII 1.38 1.24
Pt5⫹ VI 0.71 0.57 Sn4⫹ IV 0.69 0.55
Table 1.11 (continued)
Ion CN SP CR IR Ion CN SP CR IR
V 0.76 0.62 Tm3⫹ VI 1.020 0.880
VI 0.830 0.690 VIII 1.134 0.994
VII 0.89 0.75 IX 1.192 1.052
VIII 0.95 0.81 U3⫹ VI 1.165 1.025
Sr2⫹ VI 1.32 1.18 U4⫹ VI 1.03 0.89
VII 1.35 1.21 VII 1.09 0.95
VIII 1.40 1.26 IX 1.19 1.05
IX 1.45 1.31 XII 1.31 1.17
X 1.50 1.36 U5⫹ VI 0.90 0.76
XII 1.58 1.44 VII 0.98 0.84
Ta3⫹ VI 0.86 0.72 U6⫹ II 0.59 0.45
Ta4⫹ VI 0.82 0.68 IV 0.66 0.52
Ta5⫹ VI 0.78 0.64 VI 0.87 0.73
VIII 0.88 0.74 VII 0.95 0.81
Tb3⫹ VI 1.063 0.923 VIII 1.00 0.86
VII 1.12 0.98 V2⫹ VI 0.93 0.79
VIII 1.180 1.040 V3⫹ VI 0.780 0.640
Tb4⫹ VI 0.90 0.76 V4⫹ V 0.67 0.53
VIII 1.02 0.88 VI 0.72 0.58
Tc4⫹ VI 0.785 0.645 VIII 0.86 0.72
Tc5⫹ VI 0.74 0.60 V5⫹ IV 0.495 0.355
Tc7⫹ IV 0.51 0.37 V 0.60 0.46
VI 0.70 0.56 VI 0.68 0.54
Te2⫺ VI 2.07 2.21 W4⫹ VI 0.80 0.66
Te4⫹ III 0.66 0.52 W5⫹ VI 0.76 0.62
IV 0.80 0.66 W6⫹ IV 0.56 0.42
VI 1.11 0.97 V 0.65 0.51
Te6⫹ IV 0.57 0.43 VI 0.74 0.60
VI 0.70 0.56 Xe8⫹ IV 0.54 0.40
4⫹
Th VI 1.08 0.94 VI 0.62 0.48
VIII 1.19 1.05 Y3⫹ VI 1.040 0.900
IX 1.23 1.09 VII 1.10 0.96
X 1.27 1.13 VIII 1.159 1.019
XI 1.32 1.18 IX 1.215 1.075
XII 1.35 1.21 Yb3⫹ VI 1.008 0.868
Ti2⫹ VI 1.00 0.86 VII 1.065 0.925
Ti3⫹ VI 0.810 0.670 VIII 1.125 0.985
Ti4⫹ IV 0.56 0.42 IX 1.182 1.042
V 0.65 0.51 Zn2⫹ IV 0.74 0.60
VI 0.745 0.605 V 0.82 0.68
VIII 0.88 0.74 VI 0.880 0.740
Tl1⫹ VI 1.64 1.50 VIII 1.04 0.90
VIII 1.73 1.59 Zr4⫹ IV 0.73 0.59
XII 1.84 1.70 V 0.80 0.66
Tl3⫹ IV 0.89 0.75 VI 0.86 0.72
VI 1.025 0.885 VII 0.92 0.78
VIII 1.12 0.98 VIII 0.98 0.84
Tm2⫹ VI 1.17 1.03 IX 1.03 0.89
VII 1.23 1.09
42 / Geochemistry of Crystal Phases
than 1000 crystal structures, Shannon and Prewitt (1969) made two distinct as-
sumptions:
Shannon and Prewitt (1969) named the ionic radii derived from the first assump-
tion “effective ionic radii” (IRs) and those derived from the second assumption
“crystal radii,” thus stating that they are more appropriate to real structures.
Shannon’s (1976) tabulation, reported in the preceding pages, is a simple up-
dating of the previous tabulation of Shannon and Prewitt (1969), and obeys the
following assumptions:
We focus attention on the fact that the crystal radii (CRs) for the various
cations listed in table 1.11 are simply equivalent to the effective ionic radii (IRs)
augmented by 0.14 Å. Wittaker and Muntus (1970) observed that the CR radii
of Shannon and Prewitt (1969) conform better than IR radii to the “radius ratio
principle”* and proposed a tabulation with intermediate values, consistent with
the above principle (defined by the authors as “ionic radii for geochemistry”), as
particularly useful for silicates. It was not considered necessary to reproduce the
* The so-called “radius ratio principle” establishes that, for a cation/anion radius ratio lower
than 0.414, the coordination of the complex is 4. The coordination numbers rise to 6 for ratios
between 0.414 and 0.732 and to 8 for ratios higher than 0.732. Actually the various compounds
conform to this principle only qualitatively. Tossell (1980) has shown that, if Ahrens’s ionic radii are
adopted, only 60% of compounds conform to the “radius ratio principle.”
Elemental Properties and Crystal Chemistry / 43
tabulation here, because the actual incertitude of the interpolated values is ap-
proximately equal to the CR-IR gap.
Z+Z−e2
FC = . (1.51)
r2
Z+Z−e2
ΦC = ∫ − FC dr = r
. (1.52)
b
ΦR = , (1.53)
rn
where b is a constant and n (Born exponent) is an integer (between 5 and 12) that
depends on the nature of interacting ions. Equation 1.53 is then replaced by
44 / Geochemistry of Crystal Phases
−r
ΦR = b exp , (1.54)
ρ
F = − KF d (1.55)
1 1 1
− KF d 2 + KF d 2 = − KF d 2 − αF2 . (1.56)
2 2 2
The interaction of the two dipoles can thus be associated with a potential energy propor-
tional to the square of the electric field F. Moreover, F itself is proportional to r⫺3, so that
the functional dependency of the dispersive energy over the distance will be proportional
to r⫺6.
dd + − dq + −
ΦD = 6
+ . (1.57)
r r8
3 E+E−
dd + − = α +α − (1.58)
2 E+ + E−
and
27 α + α − E + E − α + E + α − E −
dq + − = +
n −
(1.59)
8 e 2 E + + E − n+
(Mayer, 1933; Boswara and Franklyn, 1968), where n⫹ and n⫺ are effective elec-
trons (see Tosi, 1964, for an extended discussion of the significance of the vari-
ous terms).
The lattice energy of an ionic crystal is the amount of energy required at absolute
zero temperature to convert one mole of crystalline component into constituent ions
in a gaseous state at infinite distance. It is composed of the various forms of ener-
gies, as shown above. The calculation is in fact somewhat more complex because
of the presence of various ions of alternating charges in a regular tridimensional
network.
Φ total = Φ C + Φ R . (1.60)
The potential at the (positive) central charge in figure 1.11 is that generated by
the presence of two opposite charges at distance r, by two homologous charges
at distance 2r, by two opposite charges at distance 3r, and so on. The coulombic
potential is then represented by a serial expansion of the type
2e 2 2e 2 2e 2 2e 2 2e 2 1 1 1
ΦC = − + − + −L= − 1 − + − + L . (1.61)
r 2r 3r 4r r 2 3 4
The term in parentheses in equation 1.61, which in this particular case corre-
sponds to the natural logarithm of 2 (ln 2 ⫽ 0.69), is known as the Madelung
constant (M ) and has a characteristic value for each structure.
46 / Geochemistry of Crystal Phases
Z +2 − e 2 12 8 6 24
Φ C , Na+ = − 6 − + − + − L . (1.62)
r0 2 3 4 5
where Z⫹⫺ is the largest common factor of the pointlike charges. If we now repeat
the calculation adopting Cl ⫺ as the central charge, we obtain a coulombic poten-
tial that (for this particular structure) is identical to that acting at the cationic
site. The bulk coulombic potential is given by the summation of the cationic and
anionic parts divided by 2:
ΦC , Na + + ΦC ,Cl − Z +2 − e 2 M
ΦC = =− . (1.63)
2 r0
Let us now go back to the monodimensional array of figure 1.11. The repul-
sive potential at the central charge is composed of
−r −2r −3r −r
ΦR = 2b exp + exp + exp + L . ≈ 2b exp (1.64)
ρ ρ ρ ρ
Elemental Properties and Crystal Chemistry / 47
(the effect of repulsive forces vanishes rapidly with increasing distance and, for
practical purposes, extended effects beyond the first neighbors at first sight can
be neglected). We then have
−1.38Z +2 − e 2 −r
Φtotal = + 2b exp . (1.65)
r ρ
The ions in the array arrange themselves in order to minimize the bulk energy of
the substance. The equilibrium condition is analytically expressed by equating to
zero the first derivative of the potential energy over distance:
∂ Φtotal
= 0. (1.66)
∂ r r =r
0
−1.38Z +2 − e 2 ρ
Φtotal( r = r0 ) = 1 − . (1.67)
r0 r0
− MN 0 Z +2 − e 2 ρ
Φtotal( r = r0 ) = 1 − . (1.68)
r0 r0
U = − Φtotal − E 0 . (1.69)
To understand the significance of this last term, we must go back to the monodi-
mensional harmonic oscillator and quantum mechanics.
Atoms in a crystal are in fact in vibrational motion even at zero temperature.
48 / Geochemistry of Crystal Phases
K F = 4 π 2 mv 02 , (1.70)
1
E ( n ) = n + hv 0 , (1.71)
2
which, with n ⫽ 0, gives the zero-point energy—i.e., the vibrational energy pres-
ent even at the zero point:
1
E0 = hv0 . (1.72)
2
However, for a real crystal at zero temperature it is impossible to group all vibra-
tional motions on the lowest single vibrational mode and, if the crystal behaves
as a Debye solid (see later), zero-point energy is expressed as:
9
E0 = N 0 hv max , (1.73)
4
M N 0 Z +2 − e 2 ρ N0 dd + − 6 ρ N 0 dq + − 8ρ
U = 1 − + 1 − + 1 −
r0 r0 r06 r0 r08 r0
(1.74)
9
− N 0 hv max .
4
Table 1.12 Lattice energies of olivine end-members; values in kJ/mole (Ottonello, 1987).
EBHF ⫽ energy of the thermochemical cycle; EC ⫽ coulombic energy; ER ⫽ repulsive energy;
EDD and EDQ ⫽ dispersive energies.
n n n n n n n n
U = ∑ ∑ Zi Z j CijEL + ∑ ∑ bij CijR + ∑ ∑ dd ij CijDD + ∑ ∑ dq ij CijDQ (1.75)
i =1 j =i i =1 j =i i =1 j =i i =1 j =i
Figure 1.12 Potential well for Ni2SiO4 olivine. Abscissa values are fractional values of
isotropic compression and/or expansion.
vF Z +2 − 0.345 Kcal
U = 287.2 1 − , (1.76)
r+′ + r−′ r+′ + r−′ mole
where vF is the number of ions per formula unit and r ⫹⬘ and r ⬘⫺ are the thermo-
chemical radii. Table 1.13 lists typical thermochemical radii for the most impor-
tant polyanionic groups. Coupling the thermochemical radius of the polyanionic
group SiO44⫺ with the ionic radius of Mg2⫹ or Fe2⫹ in VI-fold coordination with
oxygen, the lattice energy of Mg2SiO4 or Fe2SiO4 is readily derived by application
of equation 1.76.
An extended form of the Kapustinsky equation has been proposed by Saxena
(1977). Its formulation is a compromise between equation 1.76 and the Ladd and
Lee equation (eq. 1.74):
b′ 0.345 c′ 6 ⋅ 0.345
U = a′ + 1 − + 1 −
r+′ + r−′ r+′ + r−′
( ) r+′ + r−′
6
r+′ + r−′
(1.77)
d′ 8 ⋅ 0.345
+ 1 −
(r ′ + r ′ ) r+′ + r−′
8
+ −
The third and fourth terms on the right side of equation 1.77 take into account
the effect of dispersive potential. Constants a⬘, b⬘, c⬘, and d⬘ in equation 1.77 have
no direct physical meaning and are derived by interpolation procedures carried
out on the main families of crystalline compounds.
52 / Geochemistry of Crystal Phases
1. First we transform metal M into a gaseous atom M(gas) and the gaseous
molecule 12X2 into a gaseous atom X(gas):
E
M (metal)
S
→ M (gas) (1.78)
1
1 ED
X 2 (gas)
2
→ X (gas) . (1.79)
2
I
M (gas)
M
→M+ . (1.80)
(gas)
−E
X (gas)
→ X −(gas) .
X
(1.81)
The required energies are, respectively, those of first ionization (IM ) and
electron affinity (⫺EX ).
3. To transform the two gaseous ions into a crystalline molecule, we must
expend energy corresponding to the lattice energy of the crystalline sub-
stance (i.e., the lattice energy with its sign changed):
−U
+
M (gas) + X (gas)
−
→ MX (crystal) . (1.82)
Elemental Properties and Crystal Chemistry / 53
−H0 1
MX (crystal)
f
→ M (metal) + X 2 (gas) . (1.83)
2
1
1 Es + ED
M (metal) + X 2 (gas) → M (gas) + X (gas)
2
2
↑ 0
−H f ↓ IM − E x . (1.84)
MX (crystal) ←−
U
+
M (gas) + X (gas)
−
By applying the Hess additivity rule to the involved energy terms, we have:
1
H 0f = E S + ED + IM − EX − U . (1.85)
2
n
n
H 0f , MX = E S +
n 2
ED + ∑ IM − nE X − U + (1 + n )RT , (1.86)
1
cycle for alkali halides. We emphasize that the electron affinity of chlorine should
be constant in all compounds and that observed differences must be imputed to
experimental approximation.
U ( x ) = cx 2 − gx 3 − fx 4 , (1.87)
where c, g, and f are positive. The term ⫺gx3 reflects the asymmetricity of repul-
sive effects, and ⫺ fx4 accounts for the attenuation of vibrational motions at high
amplitude. With increasing temperature, gx3 and fx4 become more and more
significant. Thus, generally, thermal expansion is not a linear function in T.
CV n + 2
αl = , (1.88)
2U n
where ␣l is the linear thermal expansion coefficient and n is the Born exponent
(see Das et al., 1963 for an explanatory application of eq. 1.88).
VT = V0 + aT + bT 2 , (1.89)
1 ∂d
αl = (1.90)
d ∂T P ,X
1 ∂V
αV = . (1.91)
V ∂T P ,X
We have already seen that ␣l can be related to U by means of the Debye principle.
The usefulness of ␣V will be evident later, when we treat the thermodynamic prop-
erties of crystalline substances at high T.
Another coefficient generally used in thermal expansion calculations is
the mean polyhedral linear expansion coefficient between temperatures T1 and T2
( ␣ T1⫺T2 ):
2 d 2 − d1
α T1 −T2 = . (1.92)
d1 + d 2 T2 − T1
Table 1.15 Relationship among mean polyhedral linear thermal expansion (␣20–1000),
Pauling’s bond strength, and ionicity factor (adapted from Hazen and Finger, 1982).
S 2i Z⫹Z⫺ S 2i Z⫹Z⫺
Cation Anion c ␣20–1000 Cation Anion c ␣20–1000
c c
Re6⫹ O2⫺ 6 1 1 Li⫹ O2⫺ 4 1/4 22
Bi5⫹ O2⫺ 6 5/6 0 Li⫹ O2⫺ 6 1/6 20
Si4⫹ O2⫺ 4 1 0 K⫹ O2⫺ 6 1/6 21
Ti4⫹ O2⫺ 6 2/3 8 Na⫹ O2⫺ 6 1/6 17
Zr4⫹ O2⫺ 7 4/7 8 Na⫹ O2⫺ 7 1/7 35
Hf 4⫹ O2⫺ 7 4/7 8 Na⫹ O2⫺ 8 1/8 22
U4⫹ O2⫺ 8 1/2 10 Na⫹ O2⫺ 9 1/9 18
Th4⫹ O2⫺ 8 1/2 9 Li⫹ F⫺ 6 0.125 46
Ce4⫹ O2⫺ 8 1/2 13 Na⫹ Cl⫺ 6 0.125 51
Al3⫹ O2⫺ 4 3/4 1 K⫹ Cl⫺ 6 0.125 46
Al3⫹ O2⫺ 5 3/5 6 K⫹ Br⫺ 6 0.125 49
Al3⫹ O2⫺ 6 1/2 9 K⫹ I⫺ 6 0.125 47
V3⫹ O2⫺ 6 1/2 13 Na⫹ F⫺ 6 0.125 45
Fe3⫹ O2⫺ 6 1/2 9 Rb⫹ Br⫺ 6 0.125 44
Cr3⫹ O2⫺ 6 1/2 8 Ca2⫹ F⫺ 8 0.19 21
Ti3⫹ O2⫺ 6 1/2 8 Cs⫹ Br⫺ 8 0.094 68
Bi3⫹ O2⫺ 6 1/2 9 Mn2⫹ S2⫺ 6 0.27 18
Be2⫹ O2⫺ 4 1/2 9 Pb2⫹ S2⫺ 6 0.27 22
Zn2⫹ O2⫺ 4 1/2 7 Pb2⫹ Te2⫺ 6 0.27 20
Ni2⫹ O2⫺ 6 1/3 14 Pb2⫹ Se2⫺ 6 0.27 21
Co2⫹ O2⫺ 6 1/3 14 Zn2⫹ S2⫺ 4 0.40 9
Fe2⫹ O2⫺ 6 1/3 13 Zn2⫹ Te2⫺ 4 0.40 10
Mn2⫹ O2⫺ 6 1/3 15 Zn2⫹ Se2⫺ 4 0.40 9
Cd2⫹ O2⫺ 6 1/3 13 Al3⫹ As3⫺ 4 0.56 5
Mg2⫹ O2⫺ 6 1/3 14 Ga3⫹ As3⫺ 4 0.56 7
Mg2⫹ O2⫺ 8 1/4 13 B3⫹ N4⫺ 4 0.45 13
Ca2⫹ O2⫺ 6 1/3 15 Ti4⫹ N4⫺ 6 0.53 9
Ca2⫹ O2⫺ 7 2/7 16 U4⫹ N4⫺ 6 0.53 9
Ca2⫹ O2⫺ 8 1/4 14 Nb4⫹ C4⫺ 6 0.53 7
Sr2⫹ O2⫺ 6 1/3 14 Ta4⫹ C4⫺ 6 0.53 7
Ba2⫹ O2⫺ 6 1/3 15 Ti4⫹ C4⫺ 6 0.53 8
Ba2⫹ O2⫺ 9 2/9 15 Zr4⫹ C4⫺ 6 0.53 7
Pb2⫹ O2⫺ 12 1/6 23 C4⫹ C4⫺ 4 0.80 3.5
These generalizations led the authors to propose the following empirical relation,
valid for oxygen-based polyhedra:
Elemental Properties and Crystal Chemistry / 57
Figure 1.13 Expansion and/or compression effects on crystal structures (planar section).
A ⫽ shared edge; B ⫽ shared corner. From R. M. Hazen and L. W. Finger, Comparative
Crystal Chemistry, copyright 䉷 1982 by John Wiley and Sons. Reprinted by permission
of John Wiley & Sons.
Z
α 20 −1000 = 32.9 0.75 − + × 10 −6 ° C −1 . (1.93)
vc
Equation 1.93 was later superseded by a more general equation proposed by Ha-
zen and Finger (1979):
4v
α 20 −1000 = 2 C × 10 −6 ° C −1 . (1.94)
Si Z + Z −
The term S 2l is defined by the authors as an “ionicity factor” and assumes a value
of 0.5 for oxides and silicates, 0.75 for halides, 0.40 for calcogenides, 0.25 for
phosphides and arsenides, and 0.2 for nitrides and carbides (Z⫺ is the anion
charge). Equation 1.94 is based on the thermal expansion data listed in table 1.15.
It must be emphasized that the bulk thermal expansion of a crystal is not
simply related to the mean expansion of the various polyhedra within the struc-
ture, because progressive readjustments of polyhedral orientations are often dic-
tated by geometric constraints. As an example, figure 1.13 shows the different
behaviors of crystals with polyhedra sharing an edge (case A) or a corner (case
B). In case B, the thermal expansion is accompanied by a substantial modification
58 / Geochemistry of Crystal Phases
of the angular value of the shared corner. The same type of reasoning is valid for
compressional effects.
∂A
P = − , (1.95)
∂V T ,X
where A is the Helmholtz free energy, composed of internal energy U plus a vibra-
tional term (A vib ):
A( V ,T ,X ) = U ( V ,X ) + Avib(V,T,X ) . (1.96)
Adopting the harmonic approximation, the vibrational term is not explicitly de-
pendent on volume, and equation 1.95 may be reexpressed as follows:
∂U
P = − (1.97)
∂P T ,X
We now introduce two new parameters that describe the changes in interionic
distances and volume with pressure: isothermal linear compression coefficient l ,
and isothermal volumetric compression coefficient V :
1 ∂d
βl = − (1.98)
d ∂P T , X
1 ∂V
βV = − . (1.99)
V ∂P T ,X
The analogy with thermal expansion coefficients is evident when one compares
equations 1.98 and 1.99 with equations 1.90 and 1.91, respectively. For compress-
ibility, as for thermal expansion, a mean coefficient that defines the volumetric
variation (V2 ⫺ V1 ) for a finite pressure range (T2 ⫺ T1 ) may be introduced. This
coefficient, the “mean volumetric isothermal compressibility,” is given by
2 V2 − V1
βV = − . (1.100)
V1 + V2 P2 − P1
Elemental Properties and Crystal Chemistry / 59
The reciprocal of V, expressed in units of pressure, is called “bulk modulus” (K):
1
K = . (1.101)
βV
−1
1 ∂ 2U
βV = − . (1.102)
V ∂V 2 T ,X
As already noted, equation 1.102 is valid when vibrational energy does not depend explic-
itly on volume. In a more refined treatment of the problem, the vibrational energy of the
crystal can be treated as purely dependent on T. This assumption brings us to the Hilde-
brand equation of state and to its derivative on V:
∂U α
= −P + T V (1.103)
∂V βV
∂ 2U 1 T ∂β α ∂β
V = + 2 V2 + V V . (1.104)
∂V 2
βV βV ∂ T P, X βV ∂P T , X
∂S αV
= (1.105)
∂V T βV
∂α 1 ∂βV
V V = . (1.106)
∂V T βV ∂P T
An even more precise treatment, based on the assumption that the vibrational Helm-
holtz free energy of the crystal, divided by temperature, is a simple function of the ratio
between T and a characteristic temperature dependent on the volume of the crystal, leads
to the Mie-Gruneisen equation of state (see Tosi, 1964 for exhaustive treatment):
∂U A α
= −P + vib V . (1.107)
∂V CV βV
The ratio ␣v /CVV multiplied by the reciprocal of the density of the substance
(1/ ) is called the Gruneisen thermodynamic parameter ( ␥G ):
60 / Geochemistry of Crystal Phases
αV 1
γG = . (1.108)
CV βV ρ
In a more or less complex fashion, equations 1.97, 1.103, and 1.107 describe the re-
sponse of the crystal, in terms of energy, to modification of the intensive variable P. We
note that, in all cases, the differential forms combine internal energy and volume and
are thus analytical representations of the potential diagram seen above for Ni2SiO4 in
figure 1.12.
For simple compounds such as NaCl, where a single interionic distance (r) is sufficient
to describe the structure, if we adopt the simplified form of equation 1.97 and combine
the partial derivatives of U and V on r, we obtain
3
∂V
∂r
βV = . (1.109)
∂V ∂ 2U ∂U ∂ 2V
V −
∂ r ∂ r2 ∂ r ∂ r2
The error introduced by neglecting the term T ␣v / v (cf. eq. 1.97 and 1.103) is, in the
case of alkali halides, about 0.5 to 1.5% of the value assigned to U (Ladd, 1979).
For cubic structures with more than one interionic distance—as, for instance, spinels
(multiple oxides of type AB2O4 with A and B cations in tetrahedral and octahedral coordi-
nation with oxygen, respectively)—it is still possible to use equation 1.109, but the partial
derivatives must be operated on the cell edge, which is, in turn, a function of the various
interionic distances (Ottonello, 1986).
βV V = constant . (1.110)
Hazen and Finger (1979) extended equation 1.110 to mean polyhedral compress-
ibility  VP (mean compressibility of a given coordination polyhedron within a
crystal structure), suggesting that it is related to the charge of ions in the polyhe-
dron through an ionicity factor, analogous to what we have already seen for ther-
mal expansion—i.e.,
Elemental Properties and Crystal Chemistry / 61
Table 1.16 Polyhedral compressibility moduli in selected natural and synthetic compounds
(adapted from Hazen and Finger, 1982). Data are expressed in megabars (1Mbar ⫽ 106 bar).
Str ⫽ type of structure. K ⫽ 1/Vp.
0.044 d 3
βVP = ( Mbar −1 ). (1.111)
Si2 Z + Z −
S 2i in equation 1.111 has the same significance and values already seen in equation
1.94, and d is the bond distance.
Experimental values of mean polyhedral compressibility modulus in various
compounds are listed in table 1.16.
62 / Geochemistry of Crystal Phases
Figure 1.14 Mean polyhedral compressibility for different coordination states. Interpo-
lant is equation 1.111. From R. M. Hazen and L. W. Finger, Comparative Crystal Chemis-
try, copyright 䉷 1982 by John Wiley and Sons. Reprinted by permission of John Wi-
ley & Sons.
The reproducibility of equation 1.111 may be deduced from figure 1.14, based
on the tabulated values. Consistent deviations are noted only for some VIII-fold
coordinated complexes.
VP = V 0 + aP + bP 2 . (1.112)
An expansion of the series to the third term is not sufficiently accurate to repro-
duce the actual behavior of solids and it is preferable to use more physically sound
functional forms, such as the first-order equation of state of Murnaghan:
Elemental Properties and Crystal Chemistry / 63
K0 V K ′
P= 0 − 1 = P0 (1.113)
K′ V
V 7 3 V 5 3 V 2 3
3 3
P = K0 −
0 0
1 − ( 4 − K ′ )
0
− 1 + P0 , (1.114)
2 V V 4 V
where K 0 is the bulk modulus at zero pressure, V0 is the molar volume at zero
pressure, and K⬘ is the first derivative of the bulk modulus in dP.
Equations 1.113 and 1.114 are the most frequently used in geochemical stud-
ies. However, the most physically sound equation of state for solids is perhaps the
Born-Mayer form:
2 3 −1 3 V 4 3
V0 r0 V0
P = 3K 0 exp − − 1 − 0
V ρ V V
(1.115)
−2
r
× 0 − 2 + P0 .
ρ
Although not normally adopted in geochemistry, this form best takes into ac-
count the functional form of short-range repulsive forces (cf. eq. 1.54).
Second-order equations of state (i.e., equations involving the second deriva-
tive of the bulk modulus in P) have also been proposed (see for this purpose the
synoptic table in Kim et al., 1976). However these high-order equations are of
limited application in geochemistry because of the still large incertitude involved
in high-P compressional studies.
For better comprehension of the crystal field theory, we must reconsider to some
extent the wavelike behavior of electrons.
64 / Geochemistry of Crystal Phases
1
1 2 r
ψ1 = exp − , (1.116)
π a 03 a0
where
h2
a0 = . (1.117)
4π 2 me 2
In equation 1.116, a 0 is Bohr’s radius for the hydrogen ion in the ground state
(a 0 ⫽ 0.529 Å). The probability distribution relative to equation 1.116 is
1 −2r
ρ 1 = ψ 12 = exp (1.118)
π a 03 a0
We have already seen (section 1.1.2) that the wave equation (eq. 1.2, 1.4) is function
of both time and spatial coordinates. However, the wave function may be rewritten in
the form
ψ = ψ 0 exp( 2π i vt ), (1.119)
Then
ψ ψ = ψ 20 . (1.121)
The product in equation 1.121, which is independent of time, represents the density of
distribution—i.e., “the probability of finding a given particle in the region specified by the
spatial coordinates of .” The application of wave equation 1.10 to hypothetical atoms
Elemental Properties and Crystal Chemistry / 65
consisting of a nucleus of charge Z and a single electron progressively occupying the vari-
ous orbitals (hydrogenoid functions) requires transformation to polar coordinates (system
sketched in figure 1.15). The corresponding transformation of equation 1.10 is
1 ∂ ∂ψ 1 ∂ ∂ψ
r + 2 sinθ
2
r ∂r
2
∂r r sinθ ∂θ ∂θ
(1.122)
1 ∂ 2ψ 8π 2 m ( E − Φ )
+ + ψ = 0.
r 2 sin 2θ ∂φ 2 h2
As already shown (Section 1.2.1), the energy eigenvalues that satisfy wave equation 1.10
are given by equation 1.22:
2π 2 me 4 Z 2
E(n) = − . (1.123)
n2 h2
n = n′ + l + 1 (1.124)
(Hinshelwood, 1951), it is evident that l cannot be greater than n ⫺ 1. It is also clear that
E (n) does not depend on the total value of n but on the separate values of n⬘ and l. The
correspondence between n and the principal quantum number is thus obvious: the energy
levels follow the prediction of Bohr’s model, assuming distinct values according to quan-
tum numbers n, l, and ml , which lead to configurations s, p, and d. Equation 1.122 is then
satisfied by a product of separate functions, one dependent on l and a function of angular
coordinates and , and the other dependent on n, l, and a function of radial distance r:
ψ = ψ l (θ ,φ ) ψ n, l (r ) . (1.125)
Figure 1.15 1s, 2p, and 3d degenerate atomic orbitals (ADs) and corresponding wave
functions (cf. eq. 1.125). Upper left: spherical coordinate system. Adapted from Harvey
and Porter (1976).
Figure 1.16 Positions of anionic ligands for various configurational states: A ⫽ square-
planar coordination; B ⫽ octahedral coordination; C ⫽ tetrahedral coordination. Cation
nuclei are located in center of Cartesian coordinates.
Cartesian axes (dxy, dxz , and dyz ). The degeneracy (equal energy) of the five d
orbitals is thus partly eliminated. The energy gap between the first and second
groups of orbitals (defined by convention d␥ , d or eg , t2g , respectively) is known
as “crystal field splitting” and is usually denoted by the symbol ⌬ (or 10Dq).
According to electrostatic theory, the magnitude of crystal field splitting is
given by
5 e µa 4
∆= , (1.126)
d6
where e is the electronic charge, is the dipole moment of the ligand (cf. section
1.8), a is the mean radial distance of d orbitals, and d is the cation-to-anion dis-
tance. With respect to a spherical field, where the degeneracy of d orbitals is
maintained, the increase in energy of orbitals eg is 35⌬, and the decrease in energy
of orbitals t2g is 35⌬. Orbitals eg and t2g are therefore called, respectively, “destabi-
lized” and “stabilized.”*
* The crystal field effect is due primarily to repulsive effects between electron clouds. As we
have already seen, the repulsive energy is of opposite sign with respect to coulombic attraction and
the dispersive forces that maintain crystal cohesion. An increase in repulsive energy may thus be
interpreted as actual destabilization of the compound.
68 / Geochemistry of Crystal Phases
Figure 1.17 Absorption spectrum of a forsteritic olivine under polarized light. Ordinate
axis represents optical density (relative absorption intensity, I/I0). From R. G. Burns
(1970), Mineralogical Applications of Crystal Field Theory. Reprinted with the permission
of Cambridge University Press.
With the cation and the coordination number held constant, the magnitude
of ⌬ depends on the type of ligand and increases according to the spectrochemi-
cal series
∆
va = (1.128)
h
Figure 1.18 Crystal field splitting for d orbitals: A ⫽ square-planar field; B ⫽ octahedral
field; C ⫽ tetrahedral field.
70 / Geochemistry of Crystal Phases
Table 1.17 Electronic configuration and crystal field stabilization energies for 3d electrons in
transition elements. (Ar) ⫽ argon core: 1s22s22p63s23p6.
V3⫹ (Ar)3d 2 ↑ ↑ 4
5
24
45
12
45
Cr3⫹ (Ar)3d 3 ↑ ↑ ↑ 6
5
16
45
38
45
Co2⫹ (Ar)3d 7 ↑↓ ↑↓ ↑ ↑ ↑ 4
5
24
45
12
45
Ni2⫹ (Ar)3d 8 ↑↓ ↑↓ ↑↓ ↑ ↑ 6
5
16
45
38
45
Cu2⫹ (Ar)3d 9 ↑↓ ↑↓ ↑↓ ↑↓ ↑ 3
5
8
45
19
45
Zn2⫹ (Ar)3d 10 ↑↓ ↑↓ ↑↓ ↑↓ ↑↓ 0 0 0
energy with respect to a spherical field and compare the values for the tetrahedral
and octahedral fields, we see that, for all configurations (besides 3d 0 , 3d 5, and
3d 10 , for which crystal field stabilization is obviously zero), the octahedral field
has a higher stabilization effect, called “Octahedral Site Preference Energy”
(OSPE) (cf. table 1.17).
In 1957, almost contemporaneously, McClure (1957) and Dunitz and Orgel
(1957) proposed precise calculations of crystal field stabilization energies in tetra-
hedral and octahedral fields, and the corresponding OSPE values. Their data are
summarized in table 1.18. The ⌬ o values reported by McClure were actually de-
rived for hydrated ions (although the OH⫺ groups are near O2⫺ in spectrochemi-
cal series 1.127), whereas those listed by Dunitz and Orgel were obtained for
crystalline oxides and glasses from spectroscopic data on absorption of light. As
noted by Dunitz and Orgel, the ⌬ o values for O2⫺ and OH⫺ ligands are identical
for bivalent ions but very different for trivalent ions (8 to 12% lower for O2⫺ with
respect to OH⫺ ligands). In both works, the ⌬t value was not experimentally
obtained but was assumed to correspond to 49⌬ 0 , in agreement with theory. The
estimated OSPE values thus have a high level of incertitude. In some instances,
for trivalent ions, Dunitz and Orgel imposed the “strong field” condition (cf. sec-
tion 1.16.2).
Both McClure (1957) and Dunitz and Orgel (1957) derived their OSPE values
with the aim of explaining the intracrystalline distribution of transition elements
observed in multiple oxides of cubic structure (spinels). According to these au-
thors, a more or less high OSPE value was crucial in intracrystalline partitioning.
Their hypothesis was later taken up again by several authors. In particular, Burns
(1970) furnished several examples of natural compounds, calculating the crystal
field stabilization energy terms from adsorption spectra under polarized light (cf.
figure 1.17).
Nevertheless, the application of crystal field theory to natural compounds
Elemental Properties and Crystal Chemistry / 71
Table 1.18 Crystal field stabilization energies for 3d transition elements according to
McClure (1957) (1) and Dunitz and Orgel (1957) (2).
Octahedral Tetrahedral
⌬o ⌬t Stabilization Stabilization OSPE
Ion (cm⫺1) (cm⫺1) Energy (kJ/mole) Energy (kJ/mole) (kJ/mole)
Ti3⫹ (1) 20,300 9000 96.7 64.4 32.3
Ti3⫹ (2) 18,300 8130 87.4 58.6 28.8
V3⫹ (1) 18,000 8400 128.4 120.0 8.4
V3⫹ (2) 16,700 7400 160.2* 106.7 53.5*
V2⫹ (1) 11,800 5200 168.2 36.4 131.8
Cr3⫹ (1) 17,600 7800 251.0 55.6 195.4
Cr3⫹ (2) 15,700 6980 224.7 66.9* 157.8*
Cr2⫹ (1) 14,000 6200 100.4 29.3 71.1
Mn3⫹ (1) 21,000 9300 150.2 44.4 105.8
Mn3⫹ (2) 18,900 8400 135.6 40.2 95.4
Mn2⫹ (1) 7,500 3300 0 0 0
Fe3⫹ (1) 14,000 6200 0 0 0
Fe2⫹ (1) 10,000 4400 47.7 31.4 16.3
Fe2⫹ (2) 10,400 4620 49.8 33.1 16.7
Co3⫹ (1) – 7800 188.3 108.8 79.5
Co2⫹ (1) 10,000 4400 71.5 62.7 8.8
Co2⫹ (2) 9,700 4310 92.9* 61.9 31.0*
Ni2⫹ (1) 8,600 3800 122.6 27.2 95.4
Ni2⫹ (2) 8,500 3780 122.2 36.0* 86.2*
Cu2⫹ (1) 13,000 5800 92.9 27.6 65.3
Cu2⫹ (2) 12,600 5600 90.4 26.8 63.6
*Valid for strong field conditions.
must be carried out with extreme caution. OSPE is only a tiny fraction of the
cohesive energy of the crystal. For spinels in particular, it can be shown that the
transition from normal structure (type [A]T [BB]OO4 , where [A]T is a bivalent ion
in a tetrahedral site and [B]O is a trivalent ion in an octahedral site) to inverse
structure ( [B]T [AB]OO4 ) corresponds in terms of lattice energy to less than 200
kJ/mole and also depends on other energy effects. As an example, table 1.19 com-
pares lattice energy terms (⫺U ⫽ EC ⫹ ER ), the modulus of the bulk energy gap
between normal and inverse conditions ( | ⌬U | ), and the modulus of the crystal
field stabilization energy gap ( | ⌬CFSE | ) for some spinel compounds. In the
light of the contents of table 1.19 and of the above discussion, it is not surprising
that CFSE predictions (without detailed consideration of all forms of static ener-
gies) are denied in most cases by experimental observations.
The CFSE contribution to lattice energy is almost insignificant for meta- and
orthosilicates in which normal and distorted octahedral coordinations are present
(M1 and M2 sites, respectively). As shown in table 1.20, the CFSE gap between
normal and distorted octahedral fields is in fact only a few kJ/mole.
72 / Geochemistry of Crystal Phases
Table 1.19 Lattice energy and CFSE for selected spinel compounds.
Values of U, Ec , and Er from Ottonello (1986). CFSE values calculated
with values proposed by McClure (1957). Data in kJ/mole; x ⫽ degree of
inversion.
Compound x U EC ER | ⌬U | | ⌬CFSE |
CoFe2O4 1 18,712 ⫺21,730 3018 270 8.8
CoCr2O4 0 18,918 ⫺22,090 3172 300 186.6
CuFe2O4 1 18,805 ⫺21,697 2892 150 65.3
FeAl2O4 0 19,029 ⫺22,844 3815 400 16.3
FeFe2O4 1 18,655 ⫺21,664 3009 220 16.3
FeCr2O4 0 18,810 ⫺22,029 3219 260 179.1
NiCr2O4 0 18,948 ⫺22,116 3168 200 100.0
NiFe2O4 0 18,793 ⫺21,829 3036 220 95.4
Figure 1.19 Electron configurations under high spin (A) and low spin (B), corresponding,
respectively, to weak and strong field conditions.
leaving the destabilized eg orbitals free (figure 1.19B). The magnetic properties of
the compound (which depend on the electron spin condition) are thus influenced
by the crystal field. It should be recalled here that crystal field theory was origi-
nally developed mainly to account for the magnetic properties of crystalline com-
pounds (Bethe, 1929).
1 1 1 1 1 1
ΦC = − e 2 + + + +e
2
+ . (1.129)
dp e
A A
dp e dp e
B B
dp e
A B
dp p
B A
de e
A B A B
To obtain the wave function appropriate to the case examined in figure 1.20, we must
introduce ⌽C in the general equation (eq. 1.10). However, the complexity of the problem
does not allow an analytical solution, and the wave function is thus derived by application
74 / Geochemistry of Crystal Phases
of the “variational principle” (Hirschfelder et al., 1954). The general equation is rewritten
in the form
h2
Φ − ∇2 ψ = Eψ . (1.130)
8π 2 m
Hψ = Eψ . (1.131)
Multiplying both terms in equation 1.131 by , extracting E, and integrating over the
Cartesian coordinates, we obtain the energy of molecular orbitals (MOs):
E =
∫ ψ H ψ dτ ,
(1.132)
∫ ψ 2 dτ
where
dτ = dx ⋅ dy ⋅ dz. (1.133)
Equation 1.132 has the property of never giving an energy that is lower than the true
energy (resonance principle; cf. Pauling, 1960). This property allows us to assign to the
MO wave functions that are obtained by linear combination of the AO functions of the
separate atoms (Linear Combination of Atomic Orbitals, or LCAO, method), by progres-
sive adjustment of the combinatory parameters, up to achievement of the lowest energy.
The LCAO approximation may be expressed as:
Elemental Properties and Crystal Chemistry / 75
ΨMO = ψ A + λψ B , (1.134)
where A and B are wave functions of the AOs of atoms A and B, respectively. The
constant denotes the bond prevalence and is an expression of the polarity of MO. For
instance, with ⫽ 0, the MO of the AB molecule coincides with the AO of atom A; the
electron of B is completely displaced on A and the bond is purely ionic. With ⫽ 1, the
MO is a weighted mean of the AOs of A and B, the electrons rotate along a bicentric orbit
around the two nuclei, and the bond is perfectly covalent.
If we apply the variational procedure to hydrogen atoms at ground state, we find that
the lowest orbitals in order of increasing energy are
and
Figure 1.21 Bonding and antibonding conditions for a hydrogen molecule. Adapted from
Harvey and Porter (1976).
76 / Geochemistry of Crystal Phases
instance, *1s is the antibonding MO with rotational symmetry of the hydrogen molecule
(cf. figure 1.21).
The choice of the two AOs that can efficiently combine to form the MO may be based
on three principles:
Based on the above principles, only some combinations among the AO of the cation
and those of the ligands are possible. Plausible combinations are selected as follows:
1. Once the symmetry of the complex is known, the AO of the central cation that
best conforms to the studied case is selected—i.e., for an octahedral complex, we
select the AO directed along the Cartesian axes [ns, npx , npy , npz , (n ⫺ 1) d z2 , (n
⫺ 1)d x2 ⫺y 2 ].
2. For each ligand, the six AOs appropriate to the symmetry of the central cation
(denoted by symbols 1 , 2 , 3 , 4 , 5 , 6 ) are selected.
3. By applying the LCAO method and the principle of lowest energy, the selected
AOs are combined to form the ⌿ function of the corresponding MO. Figure
1.22A shows an example of combination of AO dxy with ⫺bonding orbitals, and
figure 1.22B shows a probable combination between d x2 ⫺y 2 orbitals of the central
cation with ⫺bonding orbitals. The function corresponding to the latter case is
12
1 − α2
Ψx 2 − y 2 = α d x 2 − y 2 + ( φ1 + φ2 − φ 3 − φ 4 ) (1.137)
4
α
Ψx*2 − y 2 = (1 − α 2 ) 1 2 d x 2 − y 2 − ( φ1 + φ2 − φ 3 − φ 4 ) (1.138)
2
The energy levels for the orbitals of the octahedral complex are sketched in figure
1.23A. For each orbital, both bonding (⌿) and antibonding (⌿*) conditions are repre-
sented. Also, the energies of the ⌿*z 2 and ⌿*x 2⫺y 2 MOs are identical so that, in analogy
with AO nomenclature, they are defined as “double degenerate” (the ⌿pz , ⌿py , and ⌿pz
MOs and their corresponding antibonding conditions are thus “triple degenerate”). Orbit-
als dxy , dxz , and dyz do not form any bonds and thus maintain their original energies.
Elemental Properties and Crystal Chemistry / 77
Figure 1.22 Examples of -bonding orbitals (A) and -bonding orbitals (B) for d elec-
trons.
Figure 1.23 General energy level diagram for an octahedral complex (A) and energy lev-
els of pyrite (B). Part (B) from Burns and Vaughan (1970). Reprinted with permission of
The Mineralogical Society of America.
The analogy between the two theories is only formal. Crystal field theory is
a purely electrostatic approach that does not take into consideration the forma-
tion of MOs and the nature of the bond. According to crystal field theory, optical
and magnetic properties are ascribed to crystal field splitting between two AOs,
whereas in ligand field theory energy splitting occurs between AOs (dxy, dxz , and
dyz ) and antibonding MOs ⌿* x 2 ⫺y 2 . Lastly, note that the condition
z 2 and ⌿*
implies combinations between AOs that are oriented in the same direction, and
crystal field theory assigns the highest energy to such configurations.
The fact that the ligand field theory can account for the fractional character
of the bond through the ␣ parameter (cf. eq. 1.137 and 1.138) leads to important
applications for covalent compounds, for which electrostatic theory is not ade-
quate.
The most classical example of application is that of sulfur compounds. Figure
1.23B shows the energy level scheme developed by Burns and Vaughan (1970) for
pyrite (FeS2 ).
The structure of FeS2 is similar to that of NaCl, with Fe2⫹ occupying the
position of Na⫹ and that of Cl ⫺ occupied by the mean point of a (S-S) group.
Each iron atom is surrounded by six sulfur atoms, pertaining to six distinct
(S-S) groups.
The energy level diagram of figure 1.23B is similar to that of figure 1.23A
(although the formal representation of degeneracy is different), but two distinct
conditions of ligand field separation, corresponding to and symmetry, are
Elemental Properties and Crystal Chemistry / 79
considered in this case. The orbitals of the ligands are composed of the six
hybrid orbitals 3s, 3p of sulfur (one for each ligand at the corners of the coordina-
tion polyhedron). The 3d orbitals of Fe2⫹ do not form MOs and remain un-
changed on the AO of the metal cation. Burns and Vaughan (1970) define ⌬cov⫺
as the crystal field separation between the 3d AO of Fe2⫹ and the t2g antibonding
MOs, implying that the bond is purely covalent (we recall for this purpose that
the condition of perfect covalence is reached when ␣2 ⫽ 12; see, for instance, eq.
1.138). These authors also suggest that, in FeS2 compounds, bonding and anti-
bonding MOs with symmetry may be formed between the 3d orbitals of the
(S-S)2⫺ group and those of Fe2⫹. In this case, energy splitting should be higher
than for the symmetry and therefore this should be the type of bonding present
in pyrite. When comparing figures 1.23A and B, it must be noted that although
the two representations of degenerate orbitals are different (on the same line in
part A and piled in part B), they have rigorously the same significance. Besides
the quoted work of Burns and Vaughan (1970) on transition element disulfides
(CuS2 , NiS2 , CoS2 , FeS2 ), ligand field theory has been successfully applied to
thiospinels (AB2S4 compounds with cubic structures) with a full account of sev-
eral of their physical properties (hardness, reflectivity, mutual solubility, density;
Vaughan et al., 1971).
In geochemistry, the Jahn-Teller effect is relevant for metals Fe2⫹ and Cu2⫹
in octahedral complexes and for Cr3⫹ and Ni2⫹ in tetrahedral complexes. Other
transition ions (e.g., Cr2⫹ and Co2⫹ ) require unusual oxidation or low-spin condi-
tions that can be reached only under extreme pressure.
Distortion of coordination polyhedra is commonly observed in silicates. Be-
cause distortion of the coordination field implies elimination of degeneracy, it is
important to evaluate to what extent the nondegenerate condition contributes to
the stability of the phase. Systematic MO calculations in this direction have been
carried out by Wood and Strens (1972) on the basis of the AOM (Angular Over-
lap Model) integrals of Ballhausen (1954).
The ionic models discussed in section 1.12 involve some sort of empiricism in
the evaluation of repulsive and dispersive potentials. They thus need accurate
parameterization based on experimental values. They are useful in predicting in-
teraction energies within a family of isostructural compounds, but cannot safely be
adopted for predictive purposes outside the parameterized chemical system or in
cases involving structural changes (i.e., phase transition studies).
However, the energies of crystalline substances can be evaluated from first
principles with reasonable precision. This goal, which is now in sight as a result
of impressive improvements in automated computing capabilities, will lead to
significant advances in theoretical computational geochemistry, as excellently
outlined by Tossell and Vaughan (1992).
We will limit ourselves here to introducing the simplest of the quantum me-
chanics procedures: the Modified Electron Gas (MEG) treatment of Gordon and
Kim (1971). This procedure is on the borderline between the classical atomistic
approach and quantum mechanics ab initio calculations that determine energy by
applying the variational principle. A short introduction to MEG treatment
should thus be of help in filling the conceptual gap between the two theories.
MEG treatment is based on the following main assumptions:
ρ = ρ A + ρB . (1.139)
82 / Geochemistry of Crystal Phases
4. Only the outer regions of the atoms in which atomic densities overlap
contribute significantly to the interaction.
5. The electron densities of the separate atoms are described by Hartree-
Fock wave functions approximated by analytic extended (Slater-type) ba-
sis sets.
Z AZB [ ρ A ( r1 ) + ρ B ( r1 )][ ρ A ( r2 ) + ρ B ( r2 )]
ΦC =
R
+ ∫∫ r12
dr1 dr2
(1.141)
[ ρ A ( r1 ) + ρ B (r1 )] ρ A ( r1 ) + ρ B ( r1 )]
−Z A ∫ r1 A
dr1 − Z B ∫ r1B
dr1 ,
where ZA and ZB are nuclear charges of atoms A and B, r12 is the distance between
two electrons, and r1A and r1B are the distances between electrons and nuclei.
The first term on the right in equation 1.141 is repulsion between the two
nuclei, the second term represents electron-electron repulsion, and the third and
fourth terms are attraction effects between electrons and nuclei. If we subtract the
coulombic energies of separate atoms from equation 1.141:
1 ρ A ( r1 ) ρ A ( r2 ) ρ A ( r1 )
ΦC , A =
2 ∫
r12
dr1 dr2 −Z A
∫
r1 A
dr1 (1.142)
1 ρ B ( r1 ) ρ B ( r2 ) ρ B ( r1 )
ΦC ,B =
2 ∫
r12
dr1 dr2 −Z B
∫
r1B
dr1 , (1.143)
Z AZB ρ A ( r1 ) ρ B ( r2 ) ρ A ( r1 )
VC =
R
+ ∫∫ r12
dr1 dr2 − Z B ∫ r1B
dr1
(1.144)
ρ B ( r1 )
−Z A ∫ r2 A
dr2 ,
Elemental Properties and Crystal Chemistry / 83
All integrals in equation 1.144 may be evaluated analytically but, in the case
of multielectron atoms, the calculations are quite complex and involve many-term
summations. Gordon and Kim (1971) thus suggest a solution by two-dimensional
(Gauss-Laguerre or Gauss-Legendre) numerical quadrature after reduction of
equation 1.144 into a single integrand that, in the case of two interacting univalent
ions, takes the form
VC = − R −1 + ∫∫ ρ A ( r1 ) ρ B ( r2 )
ZA Z B + 1 ZB ZA
R −1 + r12−1 − r1−B1 − r2−A1 dr1 dr2 . (1.145)
(
A Z − 1 )( Z B + 1) Z B + 1 ZA − 1
∫ ρ A ( r ) dr = ZA − 1 (1.146)
and
∫ ρB ( r ) dr = ZB − 1. (1.147)
Similar forms can be obtained for coulombic interaction between neutral atoms or be-
tween atoms and ions by appropriate consideration of the integrated electron densities of
the two species (in the case of interaction between two neutral atoms, the right sides of
equations 1.146 and 1.147 are, respectively, equivalent to the nuclear charges of the two
species: ZA and ZB ).
The energy of the electron gas is composed of two terms, one Hartree-Fock
term (⌽HF ) and one correlation term (⌽corr ). The Hartree-Fock term comprises
the zero-point kinetic energy density and the exchange contribution (first and
second terms on the right in equation 1.148, respectively):
13
3 4 3
ΦHF ( ρ ) = ( 3π 2 ) 2 3 ρ 2 3 − ρ 1 3. (1.148)
10 3 π
−1 −3 2 −2 −5 2
Φcorr = − 0.438r s + 1.325r s − 1.47r s − 0.4r s , (1.150)
84 / Geochemistry of Crystal Phases
3
rs = 3 , (1.151)
4π a 03 ρ
Veg = ∫ {[ ] }
dr ρ A ( rA ) + ρ B ( rB ) Φeg ( ρ A + ρ B ) − ρ A ( rA )Φeg ( ρ A ) − ρ B ( rB )Φeg ( ρ B )
(1.154 )
Table 1.21 MEG values of lattice energy and lattice parameter for
various oxides, compared with experimental values. Source of data:
Mackrod and Stewart (1979). UL is expressed in kJ/mole; a0 (in Å)
corresponds to the cell edge for cubic substances, whereas it is the lattice
parameter in the a plane for Al2O3, Fe2O3, and Ga2O3 and it is the lattice
parameter parallel to the sixfold axis of the hexagonal unit cell in rutiles
CaTiO3 and BaTiO3.
UL a0
Although the MEG model is essentially ionic in nature, it may also be used
to evaluate interactions in partially covalent compounds by appropriate choices
of the wave functions representing interacting species. This has been exemplified
by Tossell (1985) in a comparative study in which MEG treatment was coupled
with an ab initio Self Consistent Field-Molecular Orbital procedure. In this way,
Tossell (1985) evaluated the interaction of CO32⫺ with Mg2⫹ in magnesite
(MgCO3 ). Representing the CO32⫺ ion by a 4–31G wave function, holding its
geometry constant with a carbon-to-oxygen distance of 1.27 Å, and calculating
its interaction energy at various Mg-C distances, Tossell was able to reproduce
the experimental structure with reasonable precision. Calculations also showed
that the covalently bonded CO32⫺ anion has a smaller repulsive interaction with
Mg2⫹ than that of the overlapping free ions C4⫹(O2⫺ )3 . This results from the fact
that the electron densities along the C-O internuclear axis are higher near the
midregion of the C-O bond than the sum of overlapping spherical free-ion den-
86 / Geochemistry of Crystal Phases
Figure 1.25 Electron density in CO32⫺ minus sum of superimposed spherical free-ion
densities along C-O internuclear axis. Reprinted from Tossell (1985), with permission of
Elsevier Science B.V.
sities, whereas they are lower outside the oxygen nucleus (figure 1.25). Because
electron density falls in the peripheral zone of the CO32⫺ group, short-range re-
pulsion falls with approaching cations.*
* Incidentally it must be noted that the interaction energy of Mg2⫹ with CO32⫺ does not corre-
spond to the lattice energy of the crystalline substance, but is about 1⁄10 of its value. Most of the
cohesive energy is due to the covalent bonds within the CO32⫺ group.
Elemental Properties and Crystal Chemistry / 87
We have already introduced the concept of ionic polarizability (section 1.8) and
discussed to some extent the nature of dispersive potential as a function of the
individual ionic polarizability of interacting ions (section 1.11.3). We will now
treat another type of polarization effect that is important in evaluation of defect
energies (chapter 4).
Let us imagine subtracting a ⫺e charge at a given point in the crystal lattice:
at that point, effective charge ⫹e exerts polarization energy of the type
1
EP = − eΦe , (1.155)
2
1 µ 1 F
EP = − e ∑ 2i = − e ∑ α i 2i , (1.156)
2 i ri 2 i r
where Fi is the component directed along ri (toward effective charge ⫹e) of total
electric field ⌽e on the ith ion and i is the corresponding dipole moment.
Applying differential operators (cf. appendix 2), we can write
e γ
Fi = 2 + F j i , (1.157)
ri ri
Fj = ∑ ∇ i Φe , j (1.158)
j
and
The solution of the system shown in equations 1.156 to 1.159 leads to a large
number of equations (cf. Rittner et al., 1949; Hutner et al., 1949), and simplifying
88 / Geochemistry of Crystal Phases
Figure 1.26 Induced polarization effects arising from a cationic vacancy in a crystal (neg-
ative net charge). Arrows mark induced dipoles at various lattice positions. From Lasaga
(1981c). Reprinted with permission of The Mineralogical Society of America.
1
EP = − eΦe (1.160)
2
1 1
Φe = − ea 3 M i ∑ r4
+Mj ∑ r4 (1.161)
i j
2α i 1 1
Mi = 1 − (1.162)
α i + α j 4π εo
2α j 1 1
Mj = 1 − , (1.163)
αi + α j 4π εo
where a3 is the volume occupied by the pair of ions and 0 is the dielectric con-
stant of the medium.
Elemental Properties and Crystal Chemistry / 89
Table 1.22 Optical (␣e) and static (␣T) polarizabilities for silicates. Calculated values
obtained by additivity principle starting from polarizability of oxide constituents (adapted
from Lasaga and Cygan, 1982).
␣e ␣T
The Mott-Littleton approach has various degrees of approximation (up to 4). In the
zero-order approximation, polarization is regarded as the result of a series of induced
dipoles at the various lattice sites (cf. figure 1.26). In I-, II-, III-, and IV-order approxima-
tions, Mott and Littleton performed precise calculations, respectively, for the I, II, III,
and IV series of cells surrounding the origin, treating the remaining part of the crystal in
the same manner as the zero-order approximation (see Rittner et al., 1949a,b for a detailed
account of the method). Adopting the zero-order Mott-Littleton approximation, Lasaga
(1980) proposed polarization calculations for silicates. For the Mg2SiO4 component, the
system shown in equations 1.160 to 1.163 is rewritten as follows:
1
EP = − eΦe (1.164)
2
Vcell 1 1 1
Φe = −e M Mg ∑ r 4 + M Si ∑ 4
+ MO ∑ 4 (1.165)
4 Mg Si r O r
90 / Geochemistry of Crystal Phases
α iT 1 1
Mi = 1− T . (1.166)
2α Mg
T
+ α Si + 4α O 4π
T T
ε0
The ␣T terms in equation 1.166 represent total ionic polarizability, composed of electronic
polarizability ␣e plus an additional factor ␣d, defined as a “displacement term,” due to the
fact that the charges are not influenced by an oscillating electric field (as in the case of
experimental optical measurements) but are in a “static field” (Lasaga, 1980):
αT = α e + α d . (1.167)
Based on this concept, the dielectric constant is modified by application of the Clausius-
Mosotti relation:
ε 0T − 1 π
= ( 2α Mg
T
+ α TSi + 4α OT ). (1.168)
ε 0 + 2 3Vcell
T
Table 1.22 lists the values of optical polarizability ␣e and static polarizability
␣T, calculated by Lasaga and Cygan (1982) for various silicates on the basis of
the additivity principle
α Mg
T
= 2α MgO
T
+ α TSiO . (1.169)
2 SiO 4 2
Concepts of Chemical
Thermodynamics
/ 91
92 / Geochemistry of Crystal Phases
(MgO; FeO; SiO2 ), or of elements (Mg; Fe; O2 ). As we will see in section 2.6, the
choice of the various components is no longer arbitrary whenever the “phase
rule” is applied.
EQUILIBRIUM – This term is used to indicate a condition in which the various
components are subjected to dynamic and reversible exchanges among various
phases within the heterogeneous system. Stable equilibrium, dictated by the prin-
ciple of minimization of Gibbs free energy, must be distinguished from metastable
equilibrium or apparent equilibrium conditions, which are often encountered in
geological systems. The classic example of metastable equilibrium is the secretion
of carbonate shells by marine organisms, composed of the denser high-P stable
aragonite phase of the CaCO3 component (see table 2.1 and figure 2.6). In the P-T
conditions of synthesis, the calcite polymorph, which has lower Gibbs free energy,
should be formed instead of aragonite. Another example of metastable equilib-
rium is silica solubility in water: at room P-T conditions the stable phase is quartz;
however, because this phase does not nucleate easily, metastable equilibrium be-
tween amorphous silica and a saturated solution is observed. There is apparent
equilibrium in a system whenever two or more of its phases do not apparently
interact to any measurable extent and it is impossible to determine whether stable
equilibrium exists. The local equilibrium in a system represents a zone of complete
reversibility of exchanges (or “zero affinity”; see section 2.12) restricted to a re-
gion sufficiently large for compositional fluctuations to be negligible.
SOLUTION and MIXTURE – There is some confusion between these two terms
in geological literature. According to the I.U.P.A.C. (International Union for Pure
and Applied Chemistry), the term mixture must be adopted whenever “all compo-
nents are treated in the same manner”, whereas solution is reserved for cases in
which it is necessary to distinguish a “solute” from a “solvent.” This distinction
in terminology will be more evident after the introduction of the concept of “stan-
dard state.” It is nevertheless already evident that we cannot treat an aqueous
solution of NaCl as a mixture, because the solute (NaCl) in its stable (crystalline)
state has a completely different aggregation state from that of the solvent (H2O)
and, because NaCl is a strong electrolyte (see section 8.2), we cannot even imag-
ine pure aqueous NaCl.
To understand fully the chemical reaction processes that take place in rock assem-
blages, it is necessary to introduce the concept of chemical potential. Much the
same as in a gravitational potential, in which an object tends to fall from a high
to a low altitude, in a chemical potential field the reaction or “flow direction of
components” always tends to proceed from a high to a low chemical potential
region.
Concepts of Chemical Thermodynamics / 93
The Gibbs free energy of a system (Gtotal ) is composed of all the Gibbs free
energies of the various phases in the system, defined by the chemical potentials
of their components. Algebraically, this is expressed as
G total = ∑ µi n i , (2.1)
i
where ni is the number of moles of the ith component, i is its chemical potential,
and summation is extended to all components.
For a system at constant T and P,
∂ G total
= µi . (2.2)
∂ n i P ,T
The Gibbs free energy of a phase is a measurable property that depends on pres-
sure, temperature, structure, and composition. Because the chemical potential is
the partial derivative of the Gibbs free energy with respect to composition, it
graphically represents the tangent to the Gibbs free energy curve at any given
position of the compositional field. For instance, figure 2.1 shows the chemical
Figure 2.1 Relationships between chemical potentials of two components and Gibbs free
energy of a binary phase.
94 / Geochemistry of Crystal Phases
∂ Gα
µ 1,α = . (2.3)
∂ n1
∂ Gα
µ 2 ,α = . (2.4)
∂n2
Reasoning in terms of molar Gibbs free energy (G, the energy of one mole of the
given phase) and considering a homogeneous system composed of one phase with
two components, because
X 1 + X 2 = 1, (2.5)
where X1 and X2 are the mole fractions of the components in the system, we have
dG α µ 1,α − G α dG α µ 2 ,α − G α
= and = , (2.6)
dX 1 1 − X1 dX 2 1 − X2
µ1,α = Gα + 1 − X 1( ) dG
dX
α
. (2.7)
1
and
dGα
µ 2 ,α = Gα + X 1 .
(
d 1 − X1 ) (2.8)
It is also evident from figure 2.1 that the molar Gibbs free energy of a pure phase
composed of a single component is equivalent to the chemical potential of the
component itself.
Let us now consider two coexisting phases ␣ and , each having components
1 and 2 and the Gibbs free energy relationships outlined in figure 2.2. The follow-
ing four identities apply:
(
µ1,α = Gα + 1 − X 1,α ) dX
dG α
(2.9)
1, α
Concepts of Chemical Thermodynamics / 95
Figure 2.2 Relationships between chemical potential and Gibbs free energy in a two-
phase system with partial miscibility.
dGα
µ 2 ,α = Gα + X 1,α
(
d 1 − X 1,α ) (2.10)
dG β
(
µ1, β = G β + 1 − X 1, β ) dX (2.11)
1, β
dG β
µ 2 , β = G β + X 1, β .
(
d 1 − X 1, β ) (2.12)
It is clear from figure 2.2 that the chemical potentials of the two phases at equilib-
rium are equal:
µ 2 ,α = µ 2 , β . (2.14)
We thus write
dG β
(
Gα + 1 − X 1,α ) dX
dGα
(
= G β + 1 − X 1, β ) dX (2.15)
1, α 1, β
and
96 / Geochemistry of Crystal Phases
dGα dG β
Gα + X 1,α = G β + X 1, β .
(
d 1 − X 1,α ) (
d 1 − X 1, β ) (2.16)
Equations 2.15 and 2.16 are the general equilibrium relations for two coexisting
phases.
We can now examine phase stability in the compositional field: for amounts
of component 2 in ␣ higher than X␣ the tangent to the Gibbs free energy curve
of phase ␣ never intercepts the loop of phase  (in other words, the potentials of
the two components in the two phases are never identical); only phase ␣ is stable
in such conditions. Analogously for X1 ⬎ X , only phase  is stable, whereas
stable coexistence of the two phases is the case for X ⬍ X1 ⬍ X␣ and X␣ ⬍ X2
⬍ X . Although equations 2.15 and 2.16 imply that only two phases may form in
the binary system, this constraint is generally not valid, and the conformation of
the Gibbs free energy minimum curve in the G-X space becomes more complex
as the number of phases that may nucleate in the system increases. Although there
is no analytical solution to this problem, the graphical outline of the equilibrium
is quite simple. Let us consider, for instance, the isobaric-isothermal G-X plot
depicted in figure 2.3: four phases with partial miscibility of components may
form in the system (i.e., ␣, , ␥, and ␦ ), but the principle of minimization of the
Gibbs free energy and the equality condition of potentials must be obeyed over
the entire compositional range.
The conformation of the G ⫽ f (X) curve that satisfies these conditions can
be determined by draping a rope under the G -X loops of the single phases and
pulling it up on the ends, as shown in figure 2.3A. The resulting minimum Gibbs
Figure 2.3 Conformation of the minimum Gibbs free energy curve in the binary compo-
sitional field (modified from Connolly, 1992).
Concepts of Chemical Thermodynamics / 97
free energy curve may be conceived as composed of two kinds of regions (figure
2.3B): linear regions, in which the curve spans the G -X loops of the single phases,
and nonlinear regions, in which the curve overlaps the G -X loop of a single phase.
Figure 2.3 highlights three important facts (Connolly, 1992):
H = U + PV . (2.17)
G = H − TS. (2.18)
The concept of the entropy (S ) of a system is too often defined cryptically, for
the following reasons:
dU = T dS − P dV . (2.19)
dH = dU + P dV + V dP , (2.20)
dH = T dS + V dP. (2.21)
Differentiating equation 2.18 and combining it with equation 2.21, we also obtain
dG = − S dT + V dP. (2.22)
The differential equations above are valid for reversible processes taking place in
closed systems in which there is no flow of matter among the various phases.
The bulk differential of the Gibbs free energy for a compositionally variable
phase is
Concepts of Chemical Thermodynamics / 99
∂G ∂G ∂G
dG = dT + dP + dni ,
∂ T P, n ∂P T, n ∂ ni T, P, n (2.23)
j j j
which gives
dG = − S dT + V dP + ∑ µ i dni .
(2.24)
i
In equation 2.24, contrary to equation 2.22, the Gibbs free energy of the phase is
a function not only of the intensive variables T and P, but also of composition.
Equation 2.24 is thus of more general validity and can also be used in open sys-
tems or whenever there is flow of components among the various phases in the
system. Like the exact differential dG, we can reexpress exact differentials dH and
dU as
∂H
dH = T dS + V dP + dni (2.25)
∂ ni T, P, n
j
and
∂U
dU = T dS − P dV + dni . (2.26)
∂ ni T, P, n
j
The “Gibbs equations” 2.24, 2.25, and 2.26 (Gibbs, 1906) are exact differentials,
composed of the summations of the partial derivatives over all the variables of
which derived magnitude is a function (cf. appendix 2).
The partial derivatives not explicit in equations 2.25 and 2.26 are
∂G
= −S (2.27)
∂ T P, n
j
and
∂G
= V. (2.28)
∂P T, n
j
The need to make explicit the compositional derivations in equations 2.25 and
2.26 is dictated by the fact that, in the Gibbs (1906) sense, all partial derivatives
over composition are called “chemical potentials” and are symbolized by , re-
gardless of the kind of magnitude that undergoes partial derivation (i.e., not only
100 / Geochemistry of Crystal Phases
the partial derivative of Gibbs free energy G, as in equation 2.23, for instance, but
also ∂H/ ∂ ni and ∂U/ ∂ ni ). This fact gives rise to some confusion in several text-
books.
Deriving the G/T ratio in dT, we obtain
(
∂ G T ) =−
H
.
∂T T2 (2.29)
P,n j
Moreover, if we derive molar enthalpy H (the enthalpy of one mole of the sub-
stance) with respect to T, we have
∂H
= CP , (2.30)
∂T P , n
j
where CP is the specific heat of the substance at constant P—i.e., “the amount
of heat that is needed to increase by 1 kelvin one mole of the substance at the P
and T of interest.”
∂G
µi = . (2.31)
∂ n i P ,T
∂µ i ∂S
= − = − si (2.32)
∂ T P, n ∂ ni P , T
i
∂µ i ∂V
= = vi (2.33)
∂P T , n ∂ ni P ,T
i
Concepts of Chemical Thermodynamics / 101
(
∂ µi T ) =−
hi
,
∂T T2 (2.34)
P , ni
where magnitudes si , vi , and hi are defined as the “partial molar properties” of the
ith component in the phase. These magnitudes should not be confused with molar
properties S, V, and H , with which they nevertheless coincide in the condition of
a pure component in a pure phase (we have also seen that, in the same condition
i ⬅ G).
“Partial molar properties,” like molar properties, are intensive magnitudes
(i.e., they do not depend on the extent of the system), and the same relationships
that are valid for molar properties hold for them as well. For instance,
µ i = h i − Tsi . (2.35)
As we will see , partial molar properties are of general application in the thermo-
dynamics of mixtures and solutions.
The phase rule formulated by Gibbs establishes the degree of freedom (variance)
as a function of the number of phases and of the numbers of components and of
intensive (P, T ) and condition variables (gravitational field, magnetic field, elec-
tric field, etc.). This rule is intuitively assimilated to the degree of determination
of linear systems in elementary algebra: the variance of a system of variables
correlated by linear equations is equivalent to the number of independent vari-
ables less the number of correlation equations. If we consider an algebraic system
with two variables (x and y) that are not mutually dependent (i.e., there are no
correlation equations linking the two variables), the variance of the system is 2;
if y ⫽ f (x), the variance is 1 because a fixed value of y corresponds to a given
value of x. If there are two correlation equations between x and y, the variance is
0 (the system is completely determined).
Let us now consider a heterogeneous thermodynamic system at equilibrium.
If there are ⌽ phases in the system, it can easily be seen that ⌽ ⫺ 1 equations of
type 2.15 and 2.16 apply for each component in the system. Hence, if there are n
components, the number of equations will be n(⌽ ⫺ 1). Moreover, the following
mass-balance equation holds for each phase:
∑ X i ,α = 1,
(2.36)
i
( )
n Φ − 1 + Φ. (2.37)
The system variables are composed of n ⌽ compositional terms plus ambient vari-
ables that are usually two in number: temperature and pressure (hydrostatic and/
or lithostatic-isotropic pressure). The variance (V ) of the system is readily ob-
tained by subtracting the number of condition equations from the total number
of variables (n ⌽ ⫹ 2):
( ) [( )
V = n Φ + 2 − n Φ − 1 + Φ = n − Φ + 2. ] (2.38)
When applying the Gibbs phase rule, it must be remembered that the choice of
components is not arbitrary: the number of components is the minimum number
compatible with the compositional limits of the system.
Consider, for instance, the polymorphic reaction
CaCO3 ⇔ CaCO3 .
calcite aragonite (2.39)
There are two phases in reaction and, apparently, three components, correspond-
ing to atoms Ca, C, and O. However, the compositional limits of the system are
such that
Ca = C (2.40)
and
O = 3C. (2.41)
Because conditions 2.40 and 2.41 hold for all phases in the system, the minimum
number of components necessary to describe phase chemistry is 1.
Consider again the exchange reaction
O = Mg + Fe + 2 (2.43)
Concepts of Chemical Thermodynamics / 103
and
Si = 1, (2.44)
there are only two independently variable components (Fe and Mg).
Al2SiO5 occurs in nature in the solid state in three different aggregation forms:
kyanite, which is triclinic, dense, and stable at high P, and sillimanite and andalu-
site, which are orthorhombic, less dense, and stable at lower P (see table 2.1). If
we examine the G-T diagram for P ⫽ 10 kbar in figure 2.4, we note that the
kyanite polymorph has the lowest Gibbs free energy over a wide T range (T ⬍
780 °C). At T ⬎ 780 °C, sillimanite has the lowest energy. At P ⫽ 2 kbar, the T
range in which kyanite has the lowest energy is restricted to T ⬍ 320 °C, andalu-
site has the lowest energy for 320 °C ⬍ T ⬍ 600 °C, and sillimanite has the lowest
energy at T ⬎ 600 °C. At P ⫽ 3.75 kbar, we see that the Gibbs free energy is equal
for all three polymorphs at T ⬵ 500 °C (504 ⫾ 20 °C according to Holdaway and
Mukhopadhyay, 1993).
By combining the various observations obtained from the G-T diagrams in
different P conditions, we can build up a P-T diagram plotting the stability fields
of the various polymorphs, as shown in figure 2.5. The solid dots in figures 2.4
and 2.5 mark the phase transition limits and the triple point, and conform to the
experimental results of Richardson et al. (1969) (A, R, B⬘, C⬘) and Holdaway
(1971) (A, H, B, C). The dashed zone defines the uncertainty field in the
Density,
Compound Phase g/cm3 Symmetry Elemental Coordination
Figure 2.5 Polymorphs of Al2SiO5. Dashed area covers uncertainty range among the
various experiments. Marked dots B, B⬘, C, C⬘, and H are those plotted in figure 2.4.
* A recent study by Holdaway and Mukhopadhyay (1993) essentially confirms the stability
diagram of Holdaway (1971). However, it is of interest to show how even slight errors in the assigned
Gibbs free energy of a phase drastically affect the stability fields of polymorphs.
106 / Geochemistry of Crystal Phases
is then 1) would correspond to each dT variation. At the triple point R (or H),
all three polymorphs coexist stably, but any minimal change in the P-T conditions
results in the disappearance of one or two phases (variance is 0).
Figure 2.6 shows P-T stability diagrams for several components exhibiting
polymorphism in geology (the Co2SiO4 orthosilicate, which is not a major constit-
uent of rock-forming minerals, is nevertheless emblematic of phase transitions
observed in the earth’s mantle; cf. section 5.2.3).
Stability fields conform qualitatively to the following general rules:
The significance of the first two rules will be more evident in sections 2.10 and
3.7, when we discuss calculation of the Gibbs free energy of a phase at various
P-T conditions.
We saw in the preceding section that, at solid state in nature, a given component
may assume various structural states as a function of P and T variables. In given
P-T conditions, there is a given stable form. If one of the intensive variables is
changed, at a certain moment the univariant equilibrium of phase transition is
reached, and we then observe a solid state transformation leading to the new
aggregation state that is stable in the new P-T conditions. The Gibbs free energy
modification connected with the solid state phase transition is zero at equilibrium,
but the partial derivatives of the Gibbs free energy with respect to the P and T
variables are not zero.
We now distinguish solid state transformations as “first-order transitions” or
“lambda transitions.” The latter class groups all high-order solid state transforma-
tions (second-, third-, and fourth-order transformations; see Denbigh, 1971 for
exhaustive treatment). We define first-order transitions as all solid state transfor-
mations that involve discontinuities in enthalpy, entropy, volume, heat capacity,
compressibility, and thermal expansion at the transition point. These transitions
require substantial modifications in atomic bonding. An example of first-order
transition is the solid state transformation (see also figure 2.6)
C ⇔ C.
graphite diamond (2.45)
This reaction requires breaking of at least one bond and modification of the coor-
dination state of carbon (cf. table 2.1). During phase transformation, an interme-
diate configuration with high static potential is reached. In order to reach this
condition it is necessary to furnish the substance with a great amount of energy
(mainly in thermal form). First-order transitions, as a result of this energy re-
quirement, are difficult to complete, and metastable persistence of the high-energy
polymorph may often be observed. Diamond, for instance, at room P-T condi-
tions is a metastable form of carbon; aragonite, as already stated, at room P-T
conditions is a metastable form of CaCO3 , and so on.
In lambda transitions, no discontinuity in enthalpy or entropy as a function
of T and/or P at the transition zone is observed. However, heat capacity, thermal
expansion, and compressibility show typical perturbations in the lambda zone,
and T (or P) dependencies before and after transition are very different.
Alpha-beta ( ␣- ) transitions of the condensed forms of SiO2 quartz, try-
dimite, and cristobalite may all be regarded as lambda transformations. Their ki-
netics are higher than those of quartz-trydimite, quartz-cristobalite, and
quartz-coesite, which are first-order transformations. Figure 2.7 plots in detail the
evolution of enthalpy, entropy, heat capacity, and volume at the transition zone
108 / Geochemistry of Crystal Phases
Figure 2.7 Enthalpy, entropy, heat capacity, and volume modification in ␣-quartz/
-quartz transition region. Reprinted from H. C. Helgeson, J. Delany, and D. K. Bird,
American Journal of Science, 278A, 1–229, with permission.
between ␣-quartz and -quartz. The lambda point at 1 bar for this transition
occurs at T ⫽ 848 K. (The ␣-quartz/ -quartz transition in fact implies minimal
variations in enthalpy and entropy at the transition point and is therefore inter-
preted by many authors as an overlapping first-order-plus-lambda transition. We
will discuss this transition more extensively in section 5.8.)
As we have already stated, the Gibbs free energy modification connected with
a polymorphic transition is zero. The univariant equilibrium along the transition
curve is described by the Clapeyron equation:
dP ∆S transition
0
= = K transition , (2.46)
dT ∆Vtransition
0
where ⌬ S 0transition and ⌬ V 0transition are, respectively, the standard molar entropy and
standard molar volume of transition, and K transition is the Clapeyron slope of the
Concepts of Chemical Thermodynamics / 109
∂ ∆V transition
0 ∂ ∆V transition
0
∆Cp transition
0
= 2 K transition T + K 2
transition T . (2.47)
∂T P ∂P T
Tu
∆C P,excess,trans
∆S excess,trans = ∫ T
dT , (2.48)
Tl
Tu
( )
2
∆G excess,trans,T ,P , µ ,Q = G ∇Q − HQ + AQ 2 + BQ 4 + CQ 6 , (2.50)
110 / Geochemistry of Crystal Phases
(Landau and Lifshitz, 1980), where H is the conjugate field of the order parame-
ter, coefficient A is a function of temperature:
(
A = a T − Tc , ) (2.51)
∆G excess,trans =
1
2
( 1
) 1
a T − Tc Q 2 + BQ 4 + CQ 6
4 6
(2.52)
(Carpenter, 1988). Because the excess Gibbs free energy of transition must always
be at a minimum with respect to the macroscopic order parameter Q, i.e.:
∂ ∆G excess,trans
= 0, (2.53)
∂Q T ,P
the various types of transitions are evaluated on the basis of condition 2.53 by
setting Q ⫽ 0 at 0 K. The type of transition depends essentially on the signs of
B and C.
a (T − Tc )Q2 + BQ 4
1 1
∆Gexcess,trans = (2.54)
2 4
∂ ∆Gexcess,trans 1
∆Sexcess,trans = − = − aQ2 (2.55)
∂T 2
1 1
∆H excess,trans = ∆Gexcess,trans − T ∆Sexcess,trans = − aTcQ2 + BQ 4 (2.56)
2 4
a
∆C P, excess,trans = T, T ≤ Tc (2.57)
2B
Concepts of Chemical Thermodynamics / 111
B
Tc = . (2.58)
a
Ttrans = Tc . (2.59)
a (T − Tc )Q2 + CQ6
1 1
∆Gexcess,trans = (2.60)
2 6
1
∆Sexcess,trans = − aQ2 (2.61)
2
1 1
∆H excess,trans = − aTcQ2 + CQ6 (2.62)
2 6
aT
(T − T)
−1 2
∆C P, excess,trans = , T < Tc
4 Tc
c (2.63)
C
Tc = (2.64)
a
Ttrans = Tc . (2.65)
a (T − Tc )Q2 + BQ 4 + CQ6
1 1 1
∆Gexcess,trans = (2.66)
2 4 6
1 3 T − Tc
∆Sexcess,trans = − aQ2 1 + 1 −
4 Ttrans − Tc (2.67)
3
1 1 1
∆H excess,trans = − aTcQ2 + BQ 4 + CQ6 (2.68)
2 4 6
112 / Geochemistry of Crystal Phases
3 B2
Ttrans = Tc + . (2.69)
16aC
1
L= 2
aQtrans Ttrans, (2.70)
2
12
T
Q = 1 − , T < Tc . (2.71)
Tc
14
T
Q = 1 − , T < Tc . (2.72)
Tc
Figure 2.8 The order parameter Q as a function of T/Tc for second-order (a) and tricriti-
cal (b) transitions, and as a function of T/Ttrans for a first-order transition (c) with Tc ⫽
0.99Ttrans and Qtrans ⫽ 0.4. From Carpenter (1988). Reprinted by permission of Kluwer
Academic Publishers.
Concepts of Chemical Thermodynamics / 113
12
2 2 3 T − Tc
Q 2 = Q trans 1 + 1 − , (2.73)
3 4 T trans − T c
and at Ttrans,
12
4 a(Ttrans − Tc )
Q trans = ± . (2.74)
b
Curve c in figure 2.8, which is valid for a first-order transition, has been calcu-
lated by Carpenter (1988) for a relatively small difference between Ttrans and Tc
(Tc ⫽ 0.99Ttrans ) and a relatively high Qtrans (Qtrans ⫽ 0.4), resulting in a small
jump at Ttrans.
Obviously, if we know experimentally the behavior of the macroscopic order-
ing parameter with T, we may determine the corresponding coefficients of the
Landau expansion (eq. 2.52). However, things are not so easy when different
transitions are superimposed (such as, for instance, the displacive and order-
disorder transitions in feldspars). In these cases the Landau potential is a summa-
tion of terms corresponding to the different reactions plus a coupling factor asso-
ciated with the common elastic strain.
We will see detailed application of Landau theory to complex superimposed
transition phenomena when we treat the energetics of feldspars in chapter 5.
From what we have seen up to this point, we can perform “relative” evaluations
of the modification of Gibbs free energy with composition (section 2.2) and with
state conditions of the system (section 2.4). However, in order to calculate ther-
modynamic equilibrium, we need a rigorous energy scale—one that refers to a
firmly established “starting level” of the chemical potential. This level, or standard
state, which is kept constant as calculations develop, allows us to evaluate quanti-
tatively the change in chemical potential that takes place, for instance, in a solid
state transition such as those discussed in the preceding section. For the same
purpose, we introduce the concept of thermodynamic activity, which reflects the
difference in chemical potential for a given component in a given phase at fixed
P-T conditions with respect to chemical potential at standard state.
The algebraic formulation of the relationship among chemical potential i ,
standard state chemical potential 0i , and thermodynamic activity ai is
µ i ,α = µ i0,α + RT ln a i ,α , (2.75)
114 / Geochemistry of Crystal Phases
where i,␣ is the chemical potential of the ith component in phase ␣ and R is the
gas constant. The reversed expression for thermodynamic activity is therefore
µ i ,α − µ i0,α
ai ,α = exp . (2.76)
RT
µ i0,α = µ i ,α ( X i = 1) . (2.77)
ai , α = X i , α γ i , α , (2.78)
where ␥i,␣ is the rational activity coefficient, which is a function of P, T, and com-
position and embodies interaction effects among the component of interest and
all the other components in the mixture. It is obvious from equation 2.75 that, at
standard state, the activity of the component is equal to 1. If the adopted standard
state is that of “pure component,” we also have
X i , α = 1 ⇒ γ i , α = 1. (2.79)
Combining equation 2.78 with equation 2.75, we derive the general activity-
concentration relation that is valid over the entire compositional range:
µ i ,α = µ i0,α + RT ln X i ,α + RT ln γ i ,α . (2.80)
These fields are shown in a semilogarithmic plot in figure 2.9, relative to compo-
nent 1 in mixture ␣.
Chemical interactions among component 1 and all other components in the mix-
ture are virtually absent, because the mixture is composed essentially of compo-
nent 1 itself ( ␥1,␣ ⫽ 1). If we compare equations 2.81 and 2.75, we see that, in
this compositional field, we have
is straight, with slope RT, but, unlike the previous case, it also involves an inter-
cept term:
where h1,␣ is Henry’s constant for component 1 in mixture ␣ and depends on the
P, T, and composition of the mixture, but not on the amount of component 1 in
the mixture. In this case, the thermodynamic activity of component 1 is defined,
in terms of equation 2.78, by
Intermediate Field
Equation 2.78 is still valid, and the chemical potential of component 1 is defined
by equation 2.80, with ␥ varying with P, T, and composition (see figure 2.9).
where 1*,␣ is the chemical potential for Henry’s standard state. In this case, too,
as shown in figure 2.9, we observe in the field of Henry’s law a straight-line rela-
tion with slope RT between chemical potential and the natural logarithm of the
molar concentration, but the intercept is now * 1 , ␣ . Again, we introduce an activ-
ity coefficient varying with X1,␣ for the intermediate concentration field, and we
have
∂H ∂H
dH = dT + dP (2.90)
∂T P ∂P T
and
∂S ∂S
dS = dT + dP , (2.91)
∂T P ∂P T
where
∂H ∂V
= V −T (2.92)
∂P T ∂T P
and
∂S ∂V
= − , (2.93)
∂P T ∂T P
∂V
T P
H T, P = H T0 , P +
r r ∫ PC dT + ∫ V − T ∂T dP (2.94)
Tr Pr P
and
118 / Geochemistry of Crystal Phases
T P
dT ∂V
ST, P = ST0 , P +
r r ∫ CP
T
− ∫ dP.
∂T P (2.95)
Tr Pr
G T, P = H T, P − TST, P . (2.96)
The molar Gibbs free energy of the phase at P and T of interest is obtained by
combining equations 2.94, 2.95, and 2.96 for known values of molar enthalpy
and molar entropy at Tr and Pr reference conditions (H 0Tr ,Pr and S 0Tr ,Pr respec-
tively).
The general form of the Gibbs-Duhem equation for an n-component system can
be expressed as
X 1 d ln a1 + X 2 d ln a2 + L + X n d ln a n = 0. (2.97)
X 1 d ln a1 + X 2 d ln a2 = 0, (2.98)
which, by application of the properties of the exact differentials (see appendix 2),
may be expressed as
∂ ln a1 ∂ ln a2
X1 + X2 = 0. (2.99)
∂X 2 ∂X 2
∂ ln γ 1 ∂ ln γ 2
X1 + X2 = 0. (2.100)
∂X 2 ∂X 2
X2
X 2 ∂ ln γ 2
ln γ 1 = − ∫ dX 2 , (2.101)
0
1 − X 2 ∂X 2
X2
X2 ln γ 2
ln γ 1 = −
1 − X2
ln γ 2 + ∫ (1 − X 2 ) 2
dX 2 . (2.102)
0
∂G Qj
A j = − = −RT ln , (2.104)
∂ξ j P ,T Kj
120 / Geochemistry of Crystal Phases
where j is the progress variable for the jth reaction, Kj is the equilibrium constant,
and Qj is the observed activity product of the (macroscopic) reaction:
∏ ai , j
vi , j
Qj = , (2.105)
i
where ai, j is the activity of the ith term in the jth reaction, and vi, j is the stoichio-
metric coefficient of the ith term in the jth reaction (positive for products and
negative for reactants).
If the equilibrium condition of a system is perturbed by modifications in in-
tensive variables P and T, or by exchanges of energy or matter with the exterior,
the system will tendentially achieve a new Gibbs free energy minimum through
modification of the aggregation states in its various regions ( phases) or through
modification of phase compositions by flow of components. In a heterogeneous
system, several reactions will generally concur to restore equilibrium. The overall
chemical affinity to equilibrium of a heterogeneous system (Asystem ) is related to
the individual affinities of the various reactions through
Asystem = ∑ σ j Aj , (2.106)
j
where j is the relative rate of the jth reaction with respect to the overall reaction
progress variable ( ) (Temkin, 1963):
dξ j
σj = (2.107)
dξ
Asystem = ∑ Aj = 0,
(2.108)
j
and because activities can always be translated into compositional variables (mol-
alities in fluids, or molar concentrations in condensed phases), we may represent
local equilibria as discrete points in a compositionally continuous trend of locally
reversible but overall irreversible exchanges (figure 2.10); i.e., for a given compo-
nent i, in terms of total number of moles ni ,
ξ =1
∂n i , j
n i , end − n i , start = ∫ ∑j ∂ξ j dξ .
(2.109)
ξ=0
forward( k + ) →
← backward( k − )
k+
= K nj (2.111)
k−
F = −K F X. (3.1)
The indefinite integration in dX of equation 3.1 gives the potential energy of the
harmonic oscillator (cf. equation 1.16):
122 /
Thermochemistry of Crystalline Solids / 123
KF X 2
Φ( X ) = . (3.2)
2
Force constant KF thus represents the curvature of the potential with respect
to distance:
∂ 2 Φ( X )
KF = . (3.3)
∂X 2
Most diatomic molecules have a force constant in the range 102 to 103 N ⭈ m⫺1. A
common tool for the calculation of KF in diatomic molecules (often extended to couples
of atoms in polyatomic molecules) is “Badger’s rule”:
KF =
186
(N ⋅ m ) ,
−1
(X )
3 (3.4)
0 − dij
where dij is a constant relative to i and j atoms. The equation of motion for the harmonic
oscillator is, in the case of diatomic molecules,
d 2X
m* = − KF X , (3.5)
dt 2
m1m2
m* = . (3.6)
m1 + m2
The frequency of the harmonic oscillator (cycles per unit of time) is given by
1 KF
v0 = (3.7)
2π m*
Hψ = Eψ , (3.8)
124 / Geochemistry of Crystal Phases
h2
H =Φ− ∇2. (3.9)
8π 2 m
Introducing the potential of the harmonic oscillator (eq. 3.2) in the monodimen-
sional equivalent of equation 3.9 (i.e., the Schrödinger equation for one-
dimensional stationary states; see eq. 1.9), we obtain
d 2ψ ( X ) 8π 2 m * KF X 2
+ E − ψ ( X ) = 0.
2
(3.10)
dX 2 h2
The admitted energy levels correspond to the eigenvalues of equation 3.10 for
the various quantum numbers n (see section 1.1.2):
12
K h2 1
E ( n) = F n + (3.11)
4π 2 m * 2
1
E ( n ) = hv0 n + . (3.12)
2
1
E0 = hv0 . (3.13)
2
y2
ψ ( n ) = C( n )H( n, y ) exp − , (3.14)
X
Thermochemistry of Crystalline Solids / 125
where
4π 2 m * K
y= F
X . (3.15)
h 2
( )
−1 2
C ( n ) = 2 n n! π (3.16)
H ( 0, y) = 1
H (1 , y ) = 2 y
(3.17)
H ( 2 , y) = 4 y2 − 2
H( 3, y ) = 8 y 3 − 12 y.
[ ]
2
Φ ( X ) = De 1 − exp( − aX ) (3.18)
2 m*
a = π v0 . (3.19)
De
KF K
Φ( X ) = X 2 + F X 3 + L. (3.20)
2 3
126 / Geochemistry of Crystal Phases
Figure 3.1 Energy levels and wave functions of harmonic oscillator. Heavy line:
bounding potential (3.2). Light solid lines: quantum-mechanic probability density distri-
butions for various quantum vibrational numbers ( 2(n), see section 1.16.1). Dashed lines:
classical probability distribution; maximum classical probability is observed in the zone
of inversion of motion where velocity is zero. From McMillan (1985). Reprinted with
permission of The Mineralogical Society of America.
1 1
2
1
3
E ( n) = hve n + − X e n + + Ye n + + L , (3.21)
2 2 2
1 X Y
E 0 = hve − e + e (3.22)
2 4 8
and is lower with respect to the zero-point energy of the harmonic oscillator (Xe
and Ye are generally positive with 1 ⬎ ⬎ Xe ⬎ ⬎ Ye; cf. McMillan, 1985).
Thermochemistry of Crystalline Solids / 127
Figure 3.2 The experimental potential of H2 molecule (heavy line) compared with har-
monic fit (dashed line) and Morse potential (dotted line). From McMillan (1985). Re-
printed with permission of The Mineralogical Society of America.
Figure 3.2 shows the experimental potential well for the H2 molecule, com-
pared with the harmonic fit and the Morse potential. Horizontal lines represent
quantized energy levels. Note that, as vibrational quantum number n increases,
the energy gap between neighboring levels diminishes and the equilibrium dis-
tance increases, due to the anharmonicity of the potential well. The latter fact is
responsible for the thermal expansion of the substance.
At high T, when the spacing of vibrational energy levels is low with respect
to thermal energy, crystalline solids begin to show the classical behavior predicted
by kinetic theory, and the heat capacity of the substance at constant volume (CV )
approaches the theoretical limit imposed by free motion of all atoms along three
directions, in a compound with n moles of atoms per formula unit (limit of Dulong
and Petit):
CV = 3nR . (3.23)
Ni ∆E i
= exp − , (3.24)
Nv 0
kT
where ⌬Ei is the energy gap between excited state Ei and E 0 . The partition func-
tion is the summation of the relative statistical population over all accessible
quantum states:
∞
∆E i
Q= ∑ exp − kT
(3.25)
i =0
∞
U = ∑ Ni Ei . (3.26)
i =0
The constant-volume heat capacity of the substance (CV ) represents the vari-
ation of total energy U with T in the harmonic approximation (X is constant over
T), and the integration of CV over T gives the (harmonic) entropy of the sub-
stance:
T
∂U dT
CV = ;
∂T
S( T ) = ∫ CV T
. (3.27)
0
If a single energy level exists in the ground state, then clearly the partition func-
tion tends toward 1 as T decreases toward zero (nondegenerate ground state).
However, because several (n) configurations may occur at ground state (degener-
ate ground state), the system has a nonzero entropy even at the zero point:
vt = λ1t 2 . (3.29)
1 1 1
E = n1 + hv1 + n2 + hv2 + n3 + hv3 + L. (3.30)
2 2 2
The energy of the nondegenerate ground state is given by the summation of the
various oscillators over the characteristic frequency:
3N
1
E0 = h ∑ vt . (3.31)
2 t =1
Planck showed that the mean energy of a great number of oscillators, each with
a characteristic angular frequency t ⫽ 2vt , is given by
បωt
E = ,
ប ω t kT (3.32)
e −1
where
h
ប= . (3.33)
2π
Einstein (1907) applied equation 3.32 (originally conceived by Planck for the
quantization of electromagnetic energy) to the quantization of particle energy,
describing the internal energy of a solid composed of N 0 atoms as
3N 0
U = .
(e បω t kT
−1 ) (3.34)
* The three translational and three rotational degrees of freedom become normal modes when
the “unit cell molecule” is embedded in the crystal.
130 / Geochemistry of Crystal Phases
∂U 3N 0 kX 2 e X
CV = = = 3RE ( X )
∂T
( ) (3.35)
2
eX − 1
where
hvt បωt
X = = (3.36)
kT kT
and
X 2 eX X2
E( X ) = = .
(e ) 2X
2
X
−1 4 sin (3.37)
2
Equation 3.37, known as the Einstein function, is tabulated for various X-values
(see, for instance, Kieffer, 1985). In the Einstein function, the characteristic fre-
quency t (and the corresponding characteristic temperature E ; see, for instance,
eq. 3.40) has an arbitrary value that optimizes equation 3.35 on the basis of high-
T experimental data. Extrapolation of equation 3.35 at low temperature results
in notable discrepancies from experimental values. These discrepancies found a
reasonable explanation after the studies of Debye (1912) and Born and Von Kar-
man (1913).
In the Debye model, the Brillouin zone (see section 3.3) is replaced by a
sphere of the same volume in the reciprocal space (cf. eq. 3.57):
( )
3
4 2π
πK max
3
= . (3.38)
3 VL
ω D = VM K max , (3.39)
បωD
θD = . (3.40)
k
Thermochemistry of Crystalline Solids / 131
The Debye temperature of the solid defines the form of the vibrational spectrum
in the acoustic zone (low frequency) and is related to the molar volume of the
solid V and to the mean velocity of acoustic waves V M through
13
ប 6π 2 N 0
θD = VM ,
k ZV
(3.41)
3 θD T
T eX θ
C v = 9 nR ∫ X 4 dX = 3 nRD D , (3.42)
θD (e ) T
3
0 X
−1
As we have already seen, the heat capacity is the amount of heat that must be
furnished to raise the temperature of a given substance by 1 K at the T of interest.
If constant pressure is maintained during heat transfer, heat capacity is defined
as “heat capacity at constant P” (CP ). As already seen in chapter 2, CP is the
partial derivative of the enthalpy of the substance at constant P and com-
position—i.e.,
∂H
CP = . (3.43)
∂ T P, n
If heat transfer takes place at constant volume, the magnitude is defined as “heat
capacity at constant volume” (CV ) and is equivalent, as we have seen, to the
partial derivative of the internal energy of the substance at constant volume and
composition:
∂U
CV = . (3.44)
∂T V ,n
α V2
C P = CV + TV , (3.45)
βV
where ␣V and V are, respectively, the isobaric thermal expansion coefficient and
the isothermal compressibility coefficient of the substance (see also section 1.15).
We have seen in chapter 2 that the heat capacity at constant P is of fundamen-
tal importance in the calculation of the Gibbs free energy, performed by starting
from the standard state enthalpy and entropy values
T
H T ,Pr = H T0 ,P +
r r ∫ CP dT , (3.46)
Tr
T
dT
ST ,Pr = ST0 ,P +
r r ∫ CP T
, (3.47)
Tr
and
G T , Pr = H T , Pr − TST , Pr . (3.48)
With increasing T, CP, like CV, approaches a limit imposed by the Dulong and
Petit rule:
CV = 3nR (3.49)
and
α V2
C P = 3nR + TV , (3.50)
βV
C P = K 1 + K 2 T A + K 3T B + L + K nT N , (3.51)
C P = K 1 + K 2 T + K 3T −2 . (3.52)
Since then, several other polynomial expansions have been proposed. Haas and
Fisher (1976), for instance, proposed an interpolation with five terms in the form
C P = K 1 + K 2 T + K 3T −2 + K 4T 2 + K 5T −1 2 . (3.53.1)
The thermodynamic tables of Robie et al. (1978) use the Haas-Fisher polynomial
(eq. 3.53.1). Helgeson et al. (1978) use the Maier-Kelley expansion, changing sign at the
third term:
C P = K1 + K2T − K 3T −2 . (3.53.2)
Barin and Knacke (1973) and Barin et al. (1977) adopt a four-term polynomial:
C P = K1 + K2T + K 3T −2 + K 4T 2 . (3.53.3)
More recently, Berman and Brown (1985) proposed an interpolation of the type
C P = K1 + K2T − 1 2 + K 3T −2 + K 4T −3 , (3.54)
where K2 ⱕ 0 and K3 ⱕ 0. Polynomial 3.54 seems to obey the limits imposed by vibrational
theory better than the preceding forms (cf. eq. 3.50).
Even more recently, Fei and Saxena (1987) proposed the more complex functional
form
( ) ( )
CP = 3nR 1 + K 1T −1 + K 2 T −2 + K 3T −3 + A + BT + C P′ , (3.55)
where A and B are coefficients related, respectively, to thermal expansion and compress-
ibility, and C⬘P is the deviation from the theoretical Dulong and Petit limit due to anhar-
monicity, internal disorder, and electronic contributions. Figure 3.3 shows how the
Berman-Brown and Saxena-Fei equations conform with the experimental data of Robie
et al. (1978) for clinoenstatite (MgSiO3 ).
As has been repeatedly observed by various authors (see, for instance, Robie
et al., 1978; Haas et al., 1981; Robinson and Haas, 1983), Maier-Kelley and Haas-
Fisher polynomials can be used only within the T range for which they were
created. Extrapolation beyond the T limits of validity normally implies substan-
tial error progression in high-T entropy and enthalpy calculations. For instance,
figure 3.4 compares Maier-Kelley, Haas-Fisher, and Berman-Brown polynomials
for low albite. As can be seen, the first two interpolants, if extended to high T,
definitely exceed the Dulong and Petit limit. The Berman-Brown interpolant also
passes this limit, but the bias is less dramatic.
It has already been stated that, theoretically, N atoms in a crystal have 3N possible
vibrational modes. Obviously, if we knew the energy associated with each vibra-
tional mode at all T and could sum the energy terms in the manner discussed in
section 3.1, we could define the internal energy of the crystal as a function of T,
CV could then be obtained by application of equation 3.27, and (harmonic) en-
tropy could also be derived by integration of CV in dT/T.
Although the number of atoms in a crystal is extremely high, we can imagine
the crystal as generated by a spatial reproduction of the asymmetric unit by means
of symmetry operations. The calculation can thus be restricted to a particular
portion of space, defined as the Brillouin zone (Brillouin, 1953).
If we consider a primitive Bravais lattice with cell edges defined by vectors a1 , a2 , and
a3 , the corresponding reciprocal lattice is defined by reciprocal vectors b1 , b2 , and b3
so that
b1 =
( 2π )a 2 × a3
; b2 =
( 2π )a 3 × a1
; b3 =
( 2π )a 1 × a2
. (3.56)
a1 ⋅ a2 × a 3 a2 ⋅ a 3 × a1 a 3 ⋅ a1 × a2
The volume of Brillouin zone VB corresponds to (2 )3 times the reciprocal of cell volume
VL commonly adopted in crystallography, and the origin of the coordinates is at the center
of the cell (and not at one of the corners):
( 2π )
3
VB = . (3.57)
VL
K ( η ) = η1 b 1 + η 2 b 2 + η 3 b 3 , (3.58)
Figure 3.5 Phonon dispersion diagram for a complete unit cell with 3N degrees of free-
dom. From Kieffer (1985). Reprinted with permission of The Mineralogical Society of
America.
1. We can compute from first principles all possible vibrational modes for
3N oscillators in the cell unit, solving the Schrödinger equation with ap-
propriate atomic (and/or molecular) wave functions.
2. We can deduce the vibrational modes of atoms from experimental data
on the interaction of radiation of various frequencies (photons and pho-
nons) with the crystalline matter.
Figure 3.6 Diatomic chain with two force constants KF1 and KF2 and different atomic
masses (m1 ⬎ m2). Reprinted with permission from Kieffer (1979a), Review of Geophysics
and Space Physics, 17, 1–19, copyright 䉷 1979 by the American Geophysical Union.
of freedom, and the vectors are defined in the Brillouin zone from K ⫽ 0 to K ⫽
Kmax. In this phonon dispersion diagram, each vibrational mode, for a given value
of wave vector K, pertains to a branch represented by a curve in the figure. Three
acoustic branches (which approximate zero frequency at K ⫽ 0) and 3N ⫺ 3 opti-
cal branches (the three at the highest frequency being Einstein oscillators) can
be recognized.*
The nature of the wave vectors sketched in figure 3.5 is not dissimilar from
what can be deduced for a harmonic oscillator with two force constants K F1 and
K F2.
Consider a diatomic chain in which the atoms, of distinct masses m1 and m2 , are
positioned at distance a (figure 3.6). The repetition distance of the chain is 2a, and the
Brillouin zone falls between ⫺ /2a and /2a. If only the interactions between first neigh-
bors are significant, the equation of motion for atom r at position is given by
d 2 µr
Fr = KF1 ( µ r − 1 − µ r ) + K F2 ( µ r − 1 − µ r ) = m1 (3.59)
dt 2
d 2 µ r +1
Fr + 1 = KF1 ( µ r − µ r + 1 ) + KF2 ( µ r + 2 − µ r + 1 ) = m2 . (3.60)
dt 2
Because we are in a single dimension, we can adopt a scalar notation (i.e., K instead of K).
* The motion of atoms in the lattice can be depicted as a wave propagation (phonon). By
“dispersion” we mean the variation in the wave frequency as reciprocal space is traversed. The propa-
gation of sound waves is similar to the translation of all atoms of the unit cell in the same direction;
hence the set of translational modes is commonly defined as an “acoustic branch.” The remaining
vibrational modes are defined as “optical branches,” because they are capable of interaction with
light (see McMillan, 1985, and Tossell and Vaughan, 1992, for more exhaustive explanations).
138 / Geochemistry of Crystal Phases
The solution of equations 3.59 and 3.60 is a wave equation of amplitudes 1 and 2 :
µ r = ξ2 exp(irKa − iω t) (3.61)
[
µ r + 1 = ξ1 exp i (r + 1) Ka − iω t , ] (3.62)
ω = 2π ν (3.63)
The system comprised of equations 3.62 and 3.63 has solutions only when the determinant
of coefficients 1 and 2 (secular determinant) is zero. The coefficient matrix, or dynamic
matrix, is
12
KF + KF2 16KF1 KF2 m*2 2 Ka
ω2 = 1 1 ± 1 − sin , (3.65)
( ) 2
2
2m * KF1 + KF2 m1m2
where m* is the reduced mass (see section 3.1 and eq. 3.6).
For low values of K, sin2 Ka ⬵ K 2a2, and equation 3.65 has roots
(
ω 0 ≈ KF1 + KF2 )( m 1 + m2 ) K 2 a 2
−1
(3.66)
and
1 1
(
ω 1 ≈ KF1 + KF2 + ).
m1 m2
(3.67)
The first solution represents acoustic vibrational modes, and the second represents optical
vibrational modes.
diagram of the density of states: in this case the ordinate axis represents the num-
ber of frequencies between and ⫹ d. Acoustic modes generally prevail at
low frequencies, the optical continuum occurs at intermediate frequencies, and
Einstein’s oscillators are found at high frequencies. According to Kieffer (1985),
the density of states for acoustic branches may be represented by three sinusoi-
dal functions:
[ ( )]
2
2 1 arcsin ω / ω i
3 3
g (ω ) = ∑ 3N 0 + g 0(ω i ) , (3.68)
π Z
( )
12
i =1 ω i2 − ω
where i is the maximum frequency of each acoustic branch (see Kieffer, 1979c).
Posing X ⫽ ប /kT (cf. eq. 3.36), equation 3.68 (dispersed acoustic function; cf.
Kieffer, 1985) can be rewritten as
[arcsin (X X )] X
2
2
2
3 Xi
i e 2 dX
S ( Xi ) = ∫ . (3.69)
π
( ) ( )
12 2
0 Xi 2 − X 2 eX −1
Xu
X 2 e X dX
K ( X l ,X u ) = ∫ (3.71)
(X ) (e )
X 2
Xl u − Xl X
−1
(Kieffer, 1979c, 1985), where Xl and Xu correspond, respectively, to the lower and
upper bounds.
For the Einstein oscillators,
g (ω ) = g 0 + g E for ω = ωE , (3.72)
X E2 e XE
E( X ) =
(e ) (3.73)
E 2
XE
−1
(cf. eq. 3.37). The molar heat capacity at constant V, normalized to a monatomic
equivalent (CV*), based on vibrational functions 3.69, 3.71, and 3.73, is
3N 0 k 3 1
CV* = ∑
N i =1
S ( X ) + 3 N 0 k 1 −
i N
− q K ( Xl , Xu ) + 3N 0 kqE( XE ) , (3.74)
where N ⫽ nZ is the number of atoms in the primitive cell and q is the proportion
of vibrational modes due to Einstein oscillators.
Because N 0 k ⫽ R, the unnormalized heat capacity (CV ) is
3R 3 1
CV = ∑
Z i =1
S( X ) + 3nR1 −
i N
− q K ( Xl , Xu ) + 3nRqE ( XE ) .
(3.75)
Because the dispersed acoustic function 3.69, the optic continuum function 3.71,
and the Einstein function 3.73 may be tabulated for the limiting values of undi-
mensionalized frequencies (see tables 1, 2, 3 in Kieffer, 1979c), the evaluation of
CV reduces to the appropriate choice of lower and upper cutoff frequencies for
the optic continuum (i.e., Xl and Xu limits of integration in eq. 3.71), of the three
Thermochemistry of Crystalline Solids / 141
zone edges for the dispersed acoustic function (i.e., three values of the Xi upper
limit of integration in eq. 3.69), and of the frequency bands representing Einstein
oscillators (i.e., XE in eq. 3.73).
For many crystalline compounds, CP values are experimentally known and tabu-
lated at various T conditions. In several cases, the data are also presented in the
interpolated form of a heat capacity function through Maier-Kelley-type poly-
nomials.
Whenever the heat capacity CP or heat capacity function CP ⫽ f (T ) for a
given substance is not known, reasonable estimates can be afforded through addi-
tivity algorithms. The simplest procedure is decomposition of the chemical for-
142 / Geochemistry of Crystal Phases
Experimental
Mineral/T(K) Value E D K
Experimental
Mineral/T(K) Value E D K
continued
144 / Geochemistry of Crystal Phases
Experimental
Mineral/T(K) Value E D K
C P0
r ,Tr ,i
= ∑ v j ,iCP0r ,Tr , j , (3.76)
j
where C 0Pr ,Tr ,i is the standard state heat capacity for mineral i, C 0Pr ,Tr ,j is the stan-
dard state heat capacity for the jth constituent oxide, and vj,i is the stoichiometric
number of moles of the jth oxide in mineral i.
For instance, general equation 3.76 applied to forsterite gives
CP0r ,Tr ,Mg 2 SiO4 = 2CP0r ,Tr ,MgO + CP0r ,Tr , SiO2 . (3.77)
Thermochemistry of Crystalline Solids / 145
K 2 ,i = ∑ v j ,i K 2 , j (3.78.2)
j
K n,i = ∑ v j ,i K n, j . (3.78.3)
j
Table 3.2 lists the optimal values of the interpolation coefficients estimated
by Berman and Brown (1987) for the most common oxide constituents of rock-
forming minerals. These coefficients, through equations 3.78.1, 3.78.2, and 3.78.3,
allow the formulation of polynomials of the same type as equation 3.54, whose
precision is within 2% of experimental CP values in the T range of applicability.
However, the tabulated coefficients cannot be applied to phases with lambda tran-
sitions (see section 2.8).
Because the heat capacities of crystalline solids at various T are related to the
vibrational modes of the constituent atoms (cf. section 3.1), they may be expected
to show a functional relationship with the coordination states of the various
atoms in the crystal lattice. It was this kind of reasoning that led Robinson and
146 / Geochemistry of Crystal Phases
Table 3.3 Coefficients of “structural components” valid for polynomial 3.79. Coefficient K6
appears only at integration for calculation of entropy. The resulting CP is in J/(mole ⫻ K)
(from Robinson and Haas, 1983).
Structural
component K1 K2 K3 K6 K4 K5
C P = K 1 + 2 K 2 T + K 3T −2 + K 4T 2 + K 5T −1 2 . (3.79)
Using the structural coefficients in table 3.3 implies knowledge of the condi-
tions of internal disorder of the various cations in the structure. In most cases
(especially for pure components), this is not a problem. For instance, table 3.4
shows the decomposition into “structural components” of the acmite molecule,
in which all Na⫹ is in VIII-fold coordination with oxygen, Fe3⫹ is in VI-fold
coordination, and Si4⫹ is in tetrahedral coordination. In some circumstances,
however, the coordination states of the various elements are not known with
sufficient precision, or, even worse, vary with T. A typical example is spinel
(MgAl2O4 ), which, at T ⬍ 600 °C, has inversion X ⫽ 0.07 (i.e., 7% of the tetrahe-
drally coordinated sites are occupied by Al 3⫹ ) and, at T ⫽ 1300 °C, has inversion
X ⫽ 0.21 (21% of tetrahedral sites occupied by Al 3⫹ ). Clearly the calculation of
Thermochemistry of Crystalline Solids / 147
Table 3.4 Coefficients of modified Haas-Fisher polynomial 3.79 for acmite component of
pyroxene (NaFeSi2O6), after application of structural coefficients in table 3.3. The resulting
CP is in J/(mole ⫻ K) (adapted from Robinson and Haas, 1983).
Component Moles K1 K2 K3 K4 K5
The heat capacities of MgSiO3 , CaO, and MgO being known, we derive the heat
capacity of CaSiO3 by applying
According to Helgeson et al. (1978), the method is quite precise and leads to
estimates better than those provided by simple additivity procedures.
Table 3.5 compares estimates made by the different methods: the procedure
of Helgeson et al. (1978) effectively leads to the best estimates. According to
Robinson and Haas (1983), however, the result of this procedure is path-
dependent, because it is conditioned by the selected exchange components:
different exchange reactions (with different reference components) can lead to
appreciably different results for the same component.
148 / Geochemistry of Crystal Phases
M v X + v M i O ⇔ M iv X + v MO, (3.82)
and
where S 0 and V 0 are, respectively, standard molar entropies and standard molar
volumes, the standard molar entropy of the compound of interest is given by
S S0 VS0 + V 0
Miv X
S 0i = . (3.85)
Mv X 2VS0
S 0i
MvX
= ∑ v j ,i S j0 + K VM0 X − ∑ v j ,iV j 0 ,
i
v
(3.86)
j j
where vj,i is the number of moles of the jth oxide per formula unit of compound
i, V 0j is the standard molar volume of the jth oxide, and S 0j is the standard molar
entropy of the jth oxide. Constant K is related to the partial derivative of entropy
with respect to volume, corresponding to the ratio between isobaric thermal
expansion ␣ and isothermal compressibility :
∂S ∂P α
= = . (3.87)
∂V T ∂ T V β
∂S nRX 2
= ,
∂V 298 V 0
Miv X
( )(
e X − 1 1 − e −X ) (3.88)
where n is the number of atoms in the formula unit and X is the nondimension-
alized frequency (see eq. 3.36). As shown by Holland (1989), the mean value
of ( ∂S/ ∂V )298, obtained by application of equation 3.88 to a large number of
compounds, is 1.07 ⫾ 0.11 J ⭈ K⫺1 ⭈ cm⫺3. An analogous development using
Debye’s model (more appropriate to low-T conditions; see section 3.1) leads to a
mean partial derivative ( ∂S/ ∂V )298 of 0.93 ⫾ 0.10 J ⭈ K⫺1 ⭈ cm⫺3. Based on the
above evidence, Holland (1989) proposed a set of S 0j ⫺ V 0j finite differences for
Table 3.6 Comparison of predictive capacities of various equations in estimating standard
molar entropy (T ⫽ 298.15 K, P ⫽ 1 bar). Column I ⫽ simple summation of standard molar
entropies of constituent oxides. Column II ⫽ equation 3.86. Column III ⫽ equation 3.86
with procedure of Holland (1989). Column IV ⫽ equation 3.85. Values are in J/(mole ⫻ K).
Lower part of table: exchange reactions adopted with equation 3.85 (from Helgeson et al.,
1978) and S̄ 1j finite differences for structural oxides (Holland, 1989).
Standard molar entropy estimates obtained by means of equation 3.85 are generally
within 1% deviation of the calorimetric entropy for the majority of silicates, provided
that the application does not involve ferrous compounds. Estimates obtained by applying
equation 3.86 are generally less precise. When applied to ferrous compounds, either equa-
tion (3.85 or 3.86) leads to values that are generally higher than calorimetric ones. Ac-
cording to Helgeson et al. (1978), the bias is around 10 J per mole of FeO in the com-
pound. According to Burns (1970) and Burns and Fyfe (1967), the lower entropy of
ferrous compounds is due to the nonspherical symmetry of Fe2⫹ ions in oxides and sili-
cates. More precisely, the entropy differences are due to the different electronic contribu-
tions of the d orbitals in the various coordination states (see equations 4 to 7 in Wood,
1980). To account for such contributions, appropriate correction factors for nonspherical
ions can be introduced in equation 3.85. With the method of Holland (1989), equation
3.86 reduces to S M 0
i
vX
⫽ ∑ j v j,iS 1j ⫹ V 0MivX where the S j1 terms are finite differences
S j ⫺ V j for structural oxides. The precision of the method is comparable to that of equa-
0 0
tion 3.85 (average absolute deviation ⫽ 1.41 J ⭈ K⫺1 ⭈ mole⫺1 over 60 values).
Comparative evaluation of the predictive properties of equations 3.85 and 3.86 is
given in table 3.6. Column I lists values obtained by simple summation of the entropies
of the constituent oxides. This method (sometimes observed in the literature) should be
avoided, because it is a source of significant errors. The lower part of table 3.6 lists the
adopted exchange reactions and the S j1 terms of Holland’s model.
Both the enthalpy of formation from the elements and the Gibbs free energy of
formation from the elements of various crystalline compounds at the standard
state of 298.15 K, 1 bar, can be predicted with satisfactory approximation through
the linear additivity procedures developed by Yves Tardy and colleagues (Tardy
and Garrels, 1976, 1977; Tardy and Gartner, 1977; Tardy and Viellard, 1977;
Tardy, 1979; Gartner, 1979; Viellard, 1982). Tardy’s method is based on the defi-
nition of the ⌬O2⫺ parameter, corresponding to the enthalpy (⌬H O2⫺ ) or the
Gibbs free energy (⌬GO2⫺ ) of formation of a generic oxide M2/zO(crystal) from its
aqueous ion:
2
∆ H O 2 − M z + = ∆H 0f M 2 / z O ( crystal) − ∆H 0f M z( +aqueous) (3.89)
z
2
∆ G O 2 − M z + = ∆G 0f M 2 z O ( crystal) − ∆G 0f M z + . (3.90)
z ( aqueous)
152 / Geochemistry of Crystal Phases
The ⌬O2⫺ parameter is thus in linear relation with the solution energy of a given
crystalline oxide M2/zO(crystal) in water, through
2 z+
2 H + + M 2 z O ( crystal) ⇔ M (aqueous) + H 2 O ( liquid) (3.91)
z
∆G reaction
0
= − ∆G O 2 − + ∆G 0f H 2 O ( liquid) (3.92)
∆H reaction
0
= − ∆ H O 2 − + ∆H 0f H 2 O ( liquid) . (3.93)
Table 3.7 lists the ⌬GO2⫺ parameters derived by Tardy and Garrels (1976) on
the basis of the Gibbs free energy of formation of oxides, hydroxides, and aqueous
ions. The ⌬O2⫺ parameters allow us to establish linear proportionality between
the enthalpy (or Gibbs free energy) of formation of the compound from the con-
stituent oxides within a given class of solids. If we define ⌬Hcompound and ⌬Gcom-
pound as, respectively, the enthalpy of formation and the Gibbs free energy of for-
mation of a given double oxide, starting from the oxide components M 2 / z ⫹O and 1
N 2 / z O, the general proportionality relations between these two magnitudes and the
2
∆H compound = −α H
n1n2
n1 + n 2
(
∆ HO2− Mz + ∆ HO2− N z
+
1
+
2
) (3.94)
Thermochemistry of Crystalline Solids / 153
and
∆G compound = −α G
n1 n2
n1 + n2
(
∆GO2 − M z + ∆GO2 − N z
+
1
+
2
) (3.95)
where n1 and n2 are the numbers of oxygen ions linked, respectively, to the M z1⫹
and N z2⫹ cations, and ␣H and ␣G are the correlation parameters for a given class
of compounds (double oxides). Adding to the ⌬Hcompound (or ⌬Gcompound ) the
enthalpy (or Gibbs free energy) of formation from the elements of the oxides in
question, the enthalpy (or Gibbs free energy) of formation from the elements of
the compound of interest is readily derived.
For instance, let us consider the compound clinoenstatite. Its enthalpy of
formation from the constituent oxides is
n1 n2 2
With = ; α H = −1.3095 ( 3.96.2 )
n1 + n2 3
we obtain
The experimental value is ⫺1547.45 (kJ/mole). The estimate thus has an approxi-
mation of about 10 kJ/mole.
As figure 3.8 shows, equations 3.94 and 3.95 appear as straight lines of slope
␣H (or ␣G in the case of Gibbs free energy) in a binary plot ⌬compound ⫽ f (⌬O2⫺ ).
The intercept value on the abscissa, corresponding to zero on the ordinate, repre-
sents the ⌬HO2⫺ of the “reference cation” based on equation 3.94. Thus, in the
case of metasilicates (figure 3.8B), we have ⌬HO2⫺ Si4⫹ ⫽ ⫺193.93 kJ/mole (see
also equations 3.96.3 and 3.97), and, for orthosilicates (figure 3.8A),
⌬HO2⫺Si4⫹ ⫽ ⫺204.53 kJ/mole.
Table 3.8 lists values of ␣H and ⌬HO 2⫺
reference cation for silicates and aluminates.
The tabulated parameters, as simple interpolation factors, have good correlation
coefficients, confirming the quality of the regression.
Table 3.9 lists general regression parameters between ⌬GO2⫺ and the Gibbs
free energy of formation from constituent oxides, for various classes of sub-
stances. The values conform to the general equation
∆G compound = a × ∆ G O 2 − M z + + b. (3.98)
Figure 3.8 Relationship between ⌬Hcompound and ⌬H O2⫺ in orthosilicates (A) and meta-
silicates (B). The ⌬H O2⫺ value of the reference cation is found at a value of zero on the
ordinate. Reprinted from Viellard (1982), Sciences Geologiques, Memoir n°69, Université
Louis Pasteur, with kind permission of the Director of Publication.
Table 3.8 Regression parameters ⌬H O2⫺ for silicates and aluminates. Values in
kJ/mole (from Tardy and Garrels, 1976, 1977; Tardy and Viellard, 1977; Tardy
and Gartner, 1977).
Compound ␣H ⌬H Oref.
2⫺
cation Points R2 Precision
and
The final state of volume does not depend on the integration path (see appendix
2), and thus
r r [ (
VT ,P = VT0 ,P exp α V T − Tr )] exp[ − β (P − P )] .
V r (3.101)
The ⌬P values for which the change of volume is significant are quite high,
and reference pressure Pr (normally Pr ⫽ 1) can be ignored in the development of
calculations. The integrals on pressure in equations 2.94 and 2.95 take the forms
∂V
P
∫
Pr
V − T
[ (
1
)]
dP = PVTr, Pr exp α V T − Tr 1 − βV P 1 − α V T
∂T P
0
2
( )
( 3.102 )
156 / Geochemistry of Crystal Phases
and
Pr
∫
∂V
∂T P
0
1
2
[ (
dP = α V PVTrPr exp α V T − Tr 1 − βV P .
)] (3.103)
We have seen (section 3.2) that heat capacity at constant P can be expressed
in a functional form representing its T-dependency. Adopting, for instance, the
Haas-Fisher polynomial:
C P = K 1 + K 2 T + K 3T −2 + K 4T 2 + K 5T −1 2 , (3.104)
∫ CP dT (
= K 1 T − Tr + ) K2
2
(
T 2 − Tr2 − K 3 T −1 − Tr−1) ( )
Tr
(3.105)
K
( )
+ 4 T 3 − Tr3 + 2 K 5 T 1 2 − Tr1 2
3
( )
and
T
T
∫ CP
dT
T
K
(
= K 1 ln + K 2 T − Tr − 3 T −2 − Tr−2
Tr 2
) ( )
Tr
(3.106)
K
( )
+ 4 T 2 − Tr2 − 2 K 5 T −1 2 − Tr1 2 .
2
( )
The enthalpy of the phase at the P and T of interest is derived from the
standard state values of enthalpy and volume (H 0Tr ,Pr and V 0Tr ,Pr ) through
T
H T,P = H T0 ,P +
r r ∫ CP dT
Tr
(3.107)
+ PVT 0,P exp α V T − Tr
r r [ ( )] 1
1 − βV P 1 − α V T .
2
( )
T
ST,P = ST0 ,P +
r r ∫ CP
Tr
dT
T r r
1
2
[ (
+ αV PVT0,P exp αV T − Tr 1 − βV P
)] (3.108)
Thermochemistry of Crystalline Solids / 157
T T
GT,P = GT0 ,P − ST0 ,P T − Tr +
r r r r
( ) ∫C P dT − T ∫ CP
dT
T
Tr Tr
(3.109)
+ PVT ,0P
r r
(1
exp 1 − αV T − Tr − βV P .
2
)
We will now consider in some detail the mixture models that can be applied to
rock-forming minerals. These models are of fundamental importance for the cal-
culation of equilibria in heterogeneous systems. In the examples given here we
will refer generally to binary mixtures forming isomorphous compounds over the
entire compositional field. Each of these mixtures also has a well-defined crystal
structure, with two or more energetically distinguishable structural sites. Exten-
sion of the calculations to ternary mixtures will also be given. The discussion will
be developed in terms of molar properties (i.e., referring to one mole of sub-
stance). The term “molar” will be omitted throughout the section, for the sake
of simplicity.
H mixing = 0 (3.110)
and
(
S mixing = −R X A ln X A + X B ln X B , ) (3.111)
S configuration = k ln Q , (3.112)
R
k= (3.113)
N0
158 / Geochemistry of Crystal Phases
N0 !
Q= .
( N A ) ! ( N B) !
0 0
(3.114)
S configuration = k ln
( )(
N0 !
N0 A ! N0 B ) !
[ ( ) (
= k ln N0 ! − ln N0 A ! − ln N0 B ! . )] ( 3.115)
ln N0 ! = N0 ln N0 − N0 , (3.116)
as
N0 = N0 A + N0 B , (3.117)
we have
[ ( )
S mixing = k N0 ln N0 − N0A ln N0 A − N0B ln N0 B ( )]
N0 A N0 A N0B N0B
= − kN0 ln + ln .
N0 A + N0 B N0 A + N0 B N0 A + N0 B N0 A + N0 B
( 3.118 )
Because
N0 A N0 B
= XA and = XB , (3.119)
N0 A + N 0 B N0 A + N0 B
then
(
G mixture = X A µ A0 + X B µ B0 + RT X A ln X A + X B ln X B . ) (3.122)
Equations 3.121 and 3.122 distinguish the bulk Gibbs free energy of the mixture
(Gmixture ) from the Gibbs free energy term involved in the mixing procedure
(Gmixing ).
Let us now consider in detail the mixing process of two generic components
AN and BN, where A and B are cations and N represents common anionic radi-
cals (for instance, the anionic group SiO 44⫺ ):
AN + BN → ( A, B) N (3.123)
If mixture (A,B)N is ideal, mixing will take place without any heat loss or heat
production. Moreover, the two cations will be fully interchangeable: in other
words, if they occur in the same amounts in the mixture, we will have an equal
opportunity of finding A or B over the same structural position. The Gibbs free
energy term involved in the mixing process is
(
G mixing = ∆G123 = RT XA ln XA + XB ln XB )
(3.125)
[ (
= RT XA ln XA + 1 − XA ln 1 − XA) ( )] .
∂ ∆G 123
µ A − µ A0 = = RT ln X A (3.126)
∂X A P ,T
and
∂ ∆G 123
µ B − µ B0 =
∂X B P ,T
= RT ln X B = RT ln 1 − X A .( ) (3.127)
160 / Geochemistry of Crystal Phases
a A = XA and (
a B = XB = 1 − XA . ) (3.128)
G mixture = ∑ µ i0 Xi + G mixing .
(3.129)
i
G mixing = ∑ (µi )
− µ i0 Xi
(3.130)
i
Figure 3.9A shows that the Gibbs free energy of ideal mixing is symmetric-
concave in shape, with a maximum depth directly dependent on T. Note that the
ideal Gibbs free energy of mixing term is always present even in nonideal mixtures
(see figure 3.9B, C, and D, and section 3.11).
G mixing = RT ∑ Xi ln a i = RT ∑ Xi ln X i + RT ∑ Xi ln γ i
i i i
(3.133)
= G ideal mixing + G excess mixing ,
where Gexcess mixing is the excess Gibbs free energy of mixing term.
Figure 3.9 Conformation of Gibbs free energy curve in various types of binary mixtures.
(A) Ideal mixture of components A and B. Standard state adopted is that of pure compo-
nent at T and P of interest. (B) Regular mixture with complete configurational disorder:
W ⫽ 10 kJ/mole for 500 ⬍ T(K) ⬍ 1500. (C) Simple mixture: W ⫽ 10 ⫺ 0.01 ⫻ T(K) (kJ/
mole). (D) Subregular mixture: A0 ⫽ 10 ⫺ 0.01 ⫻ T (kJ/mole); A1 ⫽ 5 ⫺ 0.01 ⫻ T (kJ/
mole). Adopting corresponding Margules notation, an equivalent interaction is obtained
with WBA ⫽ 15 ⫺ 0.02 ⫻ T (kJ/mole); WAB ⫽ 5 (kJ/mole).
162 / Geochemistry of Crystal Phases
∂ G excess mixing
S excess mixing = − =
∂T
(3.134)
∂ ln γ A ∂ ln γ B
= − RT XA
∂T
+ XB
∂ T
(
− R XA ln γ A + XB ln γ B )
∂ G excess mixing T
H excess mixing = −T 2
∂T
(3.135)
∂ ln γ A ∂ ln γ B
= − RT XA 2
+ XB
∂T ∂ T
∂ G excess mixing ∂ ln γ A ∂ ln γ B
V excess mixing = = RT X A + XB , (3.136)
∂P ∂P ∂P
where ␥A and ␥B are the activity coefficients of components AN and BN, respec-
tively, in the (A,B)N mixture.
Let us now review the main types of nonideal mixtures.
W = N0 w (3.142)
2w AB − w AA − w BB
w= . (3.143)
2
The terms wAB, wAA, and wBB are, respectively, the ionic interactions between
A-B, A-A, and B-B atoms in the (A,B)N mixture. Note that, in this model, W is
not dependent on T and P (see also figure 3.9B). The condition of complete disor-
der is often defined as “approximation of the Zeroth principle.”
∂W I
− S excess mixing = X A XB (3.145)
∂T
∂W I
H excess mixing = X A XB W I − T . (3.146)
∂T
The activity coefficients may be related to the molar concentration through the ex-
pressions
WI 2
ln γ A = XB (3.147)
RT
and
WI 2
ln γ B = X . (3.148)
RT A
164 / Geochemistry of Crystal Phases
It can easily be deduced from equations 3.147 and 3.148 that the activity-
composition relationships for the two components are symmetric in the composi-
tional field (see also figure 3.9C).
XAq A
φ A = 1 − φB = (3.149)
X A q A + XB q B
[ ]
12
ω q = 1 + 4φ Aφ B ( λ − 1) , (3.150)
where
2W II
λ = exp . (3.151)
ZRT
( )
Zq 2
φ B ωq − 1 A
γA = 1 + (3.152)
( )
φ A ωq + 1
and
( )
Zq 2
φ A ωq − 1 B
γB = 1 + (3.153)
( )
φ B ωq − 1
Thermochemistry of Crystalline Solids / 165
G excess mixing =
Z
RT X A q A ln 1 +
φ B ωq − 1 ( )
+ XB q B ln 1 +
( )
φ A ωq − 1
.
2
φ A ωq + 1
( )
( )
φ B ωq + 1
( 3.154 )
AA + BB ⇔ AB + AB. (3.155)
It can easily be seen that, when W II ⫽ 0 in equation 3.151, then ⫽ 1 and also
q ⫽ 1 (cf. eq. 3.150); it follows that activity coefficients ␥A and ␥B are 1 at all
concentrations (cf. eq. 3.152 and 3.153). Nevertheless, if 2W II /RT ⬎ 0, then q is
greater than 1; in this condition configuration AA-BB is more stable than AB-
AB. There is then a tendency toward clustering of ions A-A and B-B, which is
phenomenologically preliminary to unmixing (the activity coefficients are corre-
spondingly greater than 1). If 2W II /RT ⬍ 0, then q is less than 1, and there is a
tendency toward mixing.
The quasi-chemical model was derived by Guggenheim for application to or-
ganic fluid mixtures. Applying it to crystalline solids is not immediate, because it
necessitates conceptual modifications of operative parameters, such as the above-
mentioned “contact factor.” Empirical methods of derivation of the above param-
eters, based on structural data, are available in the literature (Green, 1970; Sax-
ena, 1972). We will not treat this model, because it is of scanty application in
geochemistry. More exhaustive treatment can be found in Guggenheim (1952)
and Ganguly and Saxena (1987).
a Av M v 2Z
= X Av1 X Mv2 . (3.156)
1
Equation 3.156 follows the general treatment proposed by Temkin (1945) for
fused salts.
166 / Geochemistry of Crystal Phases
a Av M v 2Z
= X Av1 γ Av1 X Mv 2 γ M
v2
. (3.157)
1
If, however, site interactions of atoms on energetically equivalent sites are equal
and the standard molar volumes of mixing components are not dissimilar (i.e.,
within 5 to 10% difference), equation 3.157 may be simplified and the activity of
any component i in the mixture (ai ) may be expressed in a generalized fashion as
shown by Helgeson et al. (1978):
ai = ki ∏ ∏ X j ,ss, j , i ,
v
(3.158)
s j
where vs, j,i is the number of sth energetically equivalent sites occupied by the jth
atom in one mole of the ith component, Xj,s is the molar fraction of atom j on
site s, and ki is a proportionality constant imposed by the limit
lim ai = 1.
Xi → 1 (3.159)
− v s, j , i
v s , j ,i
ki = ∏ ∏ , (3.160)
s j v s ,i
v s ,i = ∑ v s , j ,i . (3.161)
j
Gmixing = RT ∑ X i ln ki + ∑ ∑ vs, j,i ln X j, s (3.162)
i s j
(Kerrick and Darken, 1975; Helgeson et al., 1978) and is strictly valid only for isovolu-
metric mixtures.
It must be noted that, when only one kind of atom occupies the s sites in the end-
Thermochemistry of Crystalline Solids / 167
member formula, as in our compound Av1M v 2Z, for instance, then vs, j,i ⬅ vs, i and ki
reduces to 1. In silicate end-members of geochemical interest, we often have two types
of atoms occupying the identical sublattice. For example, the end-member clintonite of
trioctahedral brittle micas has formula Ca(Mg2Al) (SiAl3 )O10 (OH)2 , with Al and Mg
occupying three energetically equivalent octahedral sites and Si and Al occupying four
energetically equivalent tetrahedral sites (cf. section 5.6). We thus have, for octahedral
sites, vs, i ⫽ 3, vs, j,i ⫽ 2 for Mg, and vs, j,i ⫽ 1 for Al. For tetrahedral sites, vs,i ⫽ 4, vs, j,i ⫽
1 for Si, and vs, j,i ⫽ 3 for Al. The resulting proportionality constant is
−2 −1 −1 −3
2 1 1 3
ki = = 64. (3.163)
3 3 4 4
In some cases, attribution of coefficients vs, i and vs, j,i is not so straightforward as it
may appear at first glance. This is the case, for instance, for K-feldspar (KAlSi3O8 ), with
Al and Si occupying T1 and T2 tetrahedral sites. Assuming, for the sake of simplicity,
that T1 and T2 sites are energetically equivalent (which is not the case; see section 5.7.2),
we have vs, i ⫽ 4, vs, j,i ⫽ 1 for Al, and vs, j,i ⫽ 3 for Si, so that
−1 −3
1 3
ki = = 9.481. (3.164)
4 4
Adopting topologic symmetry, with two T1 sites energetically different from the two T2
sites, and assuming Al and Si to be randomly distributed in T1 and T2, we obtain
−1 −1 −1 −1
1 1 1 1
ki = = 1. (3.166)
1 1 1 1
γ Av Mv 2 Z = γ Av1 γ M
v2
(3.167.1)
1
XAv Mv 2 Z = X Av1 X M
v2
(3.167.2)
1
168 / Geochemistry of Crystal Phases
a Av Mv 2 Z = a Av1 aM
v2
. (3.167.3)
1
Equations 3.167.1, 3.167.2, and 3.167.3 find their analogs in the theory of electro-
lyte solutions, as we will see in detail in section 8.4.
Let us now consider again the mixture (A, B, C, . . .) v1(M, N, O, . . .) v2Z.
We can identify four components, Av1M v2Z, Av1N v2Z, B v1M v2Z, and B v1N v2Z, ener-
getically related by the reciprocal reaction
According to equation 3.168, the energy properties of the four components are
not mutually independent but, once the Gibbs free energy is fixed for three of
them, the energy of the fourth is constrained by the reciprocal energy term ⌬G168
⬅ ⌬Greciprocal. Accepting this way of reasoning, the relationship between rational
activity coefficient and site-activity coefficients is no longer given by equation
3.167.1 but, according to Flood et al. (1954), involves a reciprocal term ( ␥reciprocal )
whose nature is analogous to the cooperative energy term expressed by Guggen-
heim (1952) in the quasi-chemical model (see preceding section):
aAv
1
Mv 2 Z (
= X Av1 γ Av1 X M
v2 v2
)
γ M γ reciprocal (3.169)
∆G reciprocal
0
γ reciprocal
(
= exp 1 − X A 1 − XM
RT
)( ,
) (3.170)
where ⌬ G 0reciprocal is the Gibbs free energy change of reciprocal reaction 3.168.
So far, we have seen that deviation from ideal behavior may affect one or more
thermodynamic magnitudes (e.g., enthalpy, entropy, volume). In some cases, we
are able to associate macroscopic interactions with real (microscopic) interactions
of the various ions in the mixture (for instance, coulombic and repulsive interac-
tions in the quasi-chemical approximation). In practice, it may happen that none
of the models discussed above is able to explain, with reasonable approximation,
the macroscopic behavior of mixtures, as experimentally observed. In such cases
(or whenever the numeric value of the energy term for a given substance is more
important than actual comprehension of the mixing process), we adopt general
(and more flexible) equations for the excess functions.
Let us consider a generic mixture of components 1 and 2 in the binary compo-
Thermochemistry of Crystalline Solids / 169
( )
G excess mixing = X 1X 2 A0 + A1 X 1 − X 2 + A2 X 1 − X 2 ( ) + L,
2
(3.171)
∂ G excess mixing
RT ln γ 1 = G excess mixing + X 2 =
∂X 1
(3.172)
= X 22 [A0 ( ) (
+ A1 3X 1 − X 2 + A2 X 1 − X 2 5X 1 − X 2 + L)( ) ]
[ ( ) (
RT ln γ 2 = X 12 A0 − A1 3X 2 − X 1 + A2 X 2 − X 1 5X 2 − X 1 + L . )( ) ] (3.173)
[
G excess mixing = X 1 X 2 A0 + A1 X 1 − X 2( )] (3.175)
[ (
RT ln γ 1 = X 22 A0 + A1 3X 1 − X 2 )] (3.176)
[ (
RT ln γ 2 = X 12 A0 − A1 3X 2 − X 1 . )] (3.177)
170 / Geochemistry of Crystal Phases
Van Laar parameters A 0 and A1 can be translated into the corresponding Mar-
gules subregular parameters W12 and W21 , posing
and
(
G excess mixing = X 1 X 2 W12 X 2 + W21 X 1 ) (3.180)
[ (
RT ln γ 1 = X 22 W12 + 2 X 1 W21 − W12 )] (3.181)
[ (
RT ln γ 2 = X 12 W21 + 2 X 2 W12 − W21 . )] (3.182)
Figure 3.9D shows the form of the curve of the excess Gibbs free energy of mixing
obtained with Van Laar parameters variable with T: the mixture is subregular—
i.e., asymmetric over the binary compositional field.
So far, we have seen several ways of calculating the Gibbs free energy of a two-
component mixture. To extend calculations to ternary and higher-order mixtures,
we use empirical combinatory extensions of the binary properties. We summarize
here only some of the most popular approaches. An extended comparative ap-
praisal of the properties of ternary and higher-order mixtures can be found in
Barron (1976), Grover (1977), Hillert (1980), Bertrand et al. (1983), Acree (1984),
and Fei et al. (1986).
Because the ideal Gibbs free energy of mixing contribution is readily general-
ized to n-component systems (cf. eq. 3.131), the discussion involves only excess
terms.
(
G excess mixing = X 1 X 2 W12 X 2 + W21 X 1 )
( )
+ X 2 X 3 W23 X 3 + W32 X 2 + X 1 X 3 W13 X 3 + W31 X 1 ( ) (3.183)
1
2
(
+ X 1 X 2 X 3 W12 + W21 + W13 + W31 + W23 + W32 ) − C ,
[ (
RT ln γ 1 = X 22 W12 + 2 X 1 W21 − W12 )] + X [W2
3 13 (
+ 2 X 1 W31 − W13 )] +
1
( )
X 2 X 3 W12 + W21 + W13 + W31 − W23 − W32 + X 1 W21 − W12 + W31 − W13 +
2
( )
(X 2 )( ) (
− X 3 W23 − W32 − 1 − 2 X 1 C . ) ] ( 3.184 )
(To obtain the corresponding function for component 2, substitute 2 for 1, 3 for
2, and 1 for 3, and for component 3, substitute 3 for 1, 1 for 2, and 2 for 3.)
X2
G excess mixing = [ (
X3 ×
X 1* X 2* A012 + A112 X 1* − X 2* +
1 − X1
1 − X1
)]
1 3 013
[
X * X * A + A X * − X *
113 1 3
( )] (3.185)
[
+ X 2 X 3 A023 + A123 V23 − V32 , ( )]
where
V23 =
1
2
(
1 + X2 − X3 ) (3.186)
and
172 / Geochemistry of Crystal Phases
V32 =
1
2
(
1 + X3 − X2 . ) (3.187)
The A 012, A 013, A 023, A 112, A 113, and A 123 parameters in equation 3.185 are anal-
ogous to the Van Laar parameters in eq. 3.175 and represent the binary interac-
tions between components 1 and 2, 1 and 3, and 2 and 3, respectively. X1 , X2 ,
and X3 are the molar fractions of components 1, 2, and 3 in the mixture, and
X 1* , X 2* , and X 3* are the corresponding scaled fractions in the binary field.
∑ ( Xi )
2
G excess mixing = + Xj ⋅ Gij , binary , (3.188)
i≠ j
where Gij,binary is the excess Gibbs free energy calculated for i and j components
as if they were in binary combination—i.e., for a two-component subregular Mar-
gules notation:
(
Gij , binary = Xi* X j* Wij X j* + Wji Xi* ) (3.189)
Xi
Xi* = (3.190)
Xi + X j
Xj
X *j = . (3.191)
Xi + X j
( ) ( )
2 2
G excess mixing = X 1* + X 2* G12 , binary + X 1* + X 3* G13, binary
(3.192)
( )
2
+ X 2* + X 3* G23, binary
(
G12 , binary = X 1* X 2* W12 X 2* + W21 X 1* ) (3.193)
Thermochemistry of Crystalline Solids / 173
X1 X2
X 1* = ; X 2* = , (3.194,195)
X1 + X 2 X1 + X 2
and so on.
The application of multicomponent mixing models to the crystalline state is
a dangerous exercise. For instance, Fei et al. (1986) have shown that, adopting
the same binary interaction parameters between components in a mixture, exten-
sions to the ternary field through the Wohl and Hillert models do not lead to the
same results. Moreover, from a microscopic point of view, we cannot expect ter-
nary interactions to be linear combinations of binary terms simply because the
interionic potential field in a crystal is not a linear expression of the single proper-
ties of the interacting ions. We have seen (in chapter 1) how the lattice energy of a
crystalline substance is a complex function of individual ionic properties (effective
charge, repulsive radius, hardness factor, polarizability, etc.) and of structure (co-
ordination, site dimension, etc.)—even in a simple pair-potential static approach
(however, see also relationships between electron density and coulombic, kinetic,
and correlation energy terms in section 1.18). Undoubtedly, multicomponent
mixing models cannot be expected to have heuristic properties when applied to
the crystalline state, but must be simply adopted as parametric models to describe
mixing behavior (experimentally observed or calculated from first principles) in
multicomponent space in an easy-to-handle fashion (see, as an example, Otto-
nello, 1992).
Let us now imagine that we are dealing with a regular mixture (A,B)N with an
interaction parameter W ⫽ ⫹20 kJ/mole. The Gibbs free energy of mixing at
various temperatures will be
The Gibbs free energy of mixing curves will have the form shown in figure
3.10A. By application of the above principles valid at equilibrium conditions, we
deduce that the minimum Gibbs free energy of the system, at low T, will be
174 / Geochemistry of Crystal Phases
Figure 3.10 Solvus and spinodal decomposition fields in regular (B) and subregular (D)
mixtures. Gibbs free energy of mixing curves are plotted at various T conditions in upper
part of figure (A and C, respectively). The critical temperature of unmixing (or “consolute
temperature”) is the highest T at which unmixing takes place and, in a regular mixture
(B), is reached at the point of symmetry.
reached by the coexistence of two phases: one rich in component (A)N, and the
other rich in (B)N. The loci of tangency points (or “binodes”) at the various
values of T will define a compositional field, in a T-X space, within which the two
phases will coexist stably at equilibrium—the “solvus field of the solid mixture”
(figure 3.10B). In this case, the solvus is symmetrical with respect to composition.
Thermochemistry of Crystalline Solids / 175
Instead, if we have a solid mixture whose excess Gibbs free energy of mixing
is approximated by a subregular Margules model with WAB ⫽ 5 kJ/mole and
WBA ⫽ 35 kJ/mole, the Gibbs free energy of mixing at the various values of T is
∂ 2 G mixing
=0 (3.198)
∂X
2
and
∂ 3G mixing
= 0. (3.199)
∂X
3
Recalling that
we can write
and
∂ 3G excess mixing
=−
∂ 3G ideal mixing
=−
(
RT 2 X − 1 ).
(3.202)
∂X ∂X
(1 − X )
3 3 2
X2
* The term “spinode” was proposed by van der Waals, who formulated it in analogy with the
shape of a thorn (“spine”) given by the intersection of the tangent plane to the ⫽ f (P) function
for gaseous phases at the critical point (see figure 1 in van der Waals, 1890; see also Cahn, 1968, for
an extended discussion of the etymology of the term).
176 / Geochemistry of Crystal Phases
The critical unmixing condition is obtained from equations 3.201 and 3.202
by substituting for the left-side terms the algebraic form of the excess function
appropriate to the case under consideration. For a simple mixture, we have, for
instance (cf. eq. 3.144),
∂ 2 G excess mixing
= −2W I (3.203)
∂X 2
and
∂ 3G excess mixing
= 0, (3.204)
∂X 3
RT
2W I =
(
X 1− X ) (3.205)
and
(
RT 2 X − 1 ) = 0.
(3.206)
(1 − X )
2
X2
Because the simple mixture has symmetric properties, the system defined by equa-
tions 3.205 and 3.206 can be easily solved. Setting X ⫽ 0.5, we have
WI
T = Tconsolute = . (3.207)
2R
∂ 2 G excess mixing
∂X 2
1
2
(
= − 4 W21 − W12 + 6 X W21 − W12 ) (3.208)
and
∂ 3G excess mixing
∂X 3
(
= 6 W21 − W12 ) (3.209)
Thermochemistry of Crystalline Solids / 177
∂ 2 G excess mixing
∂X 2
(
= −2 A0 + 6 A1 2 X − 1) (3.210)
and
∂ 3G excess mixing
= 12 A1 . (3.211)
∂X 3
The combination of equations 3.208, 3.209 or 3.210, and 3.211 with general
equations 3.201 and 3.202 gives rise to transcendental equations that must be
solved using iterative procedures. In the treated cases, we have seen a single solvus
field that occupies a limited portion of the compositional field (figure 3.10B), and
an asymmetric solvus field that extends to the condition of pure component (fig-
ure 3.10D). In nature, there are also cases in which two or more solvi may be
observed over the same binary range. One example is the CaCO3–MgCO3 system
(figure 3.11). Actually, within the compositional field, we observe second-order
phase transitions (see section 2.8): the spatial group dolomite (Ca0.5Mg0.5CO3 ;
group R3 ) is a subgroup of calcite (CaCO3 ; group R3 c), and the definition “mis-
cibility gap” or even “pseudo-solvus” instead of “solvus field” would be prefer-
able (Navrotsky and Loucks, 1977). We note that the consolute temperatures for
178 / Geochemistry of Crystal Phases
the two miscibility gaps are markedly different; moreover, the gap at low Mg
content is strongly asymmetric.
Let us again consider a solid mixture (A,B)N with a solvus field similar to the
one outlined in the T-X plot in figure 3.10B, and let us analyze in detail the “form”
of the Gibbs free energy of mixing curve in the zone between the two binodes
(shaded area in figure 3.10A).
We can identify the binodal tangency point that marks the equality of poten-
tials for the coexisting phases at a given composition (Xb in figure 3.12A). We
also note that the Gibbs free energy of mixing curve, after the binode, increases
its slope up to an inflection point called the spinode (Xs ). Beyond the spinode,
the slope of the curve decreases progressively. If we now analyze the geometrical
relationships between composition and Gibbs free energy of mixing of two ge-
I II
neric points intermediate between binode and spinode (X (1 ) and X (1) in figure
3.12A), we see that the algebraic summation of the Gibbs free energies at the two
compositional points is always higher than the Gibbs free energy of the mean com-
position X(1)m. Based on the Gibbs free energy minimization principle, a generic
mixture of composition X(1)m apparently will be unable to unmix in two phases,
because any compositional fluctuation in the neighborhood of X(1)m would result
in an increase in the Gibbs free energy of the system. This apparently contrasts
with the statements of the preceding paragraphs—i.e., within the solvus field, the
minimum Gibbs free energy condition is reached with splitting in two phases.
Indeed, if we analyze the bulk Gibbs free energy of the system when unmixing is
completed, we see that it is always lower than the corresponding energy of the
homogeneous phase. The increase in energy related to compositional fluctuations
in the zone between binode Xb and spinode Xs must be conceived of as an “energy
threshold of activation of the unmixing process” that must be overcome if the
process is to terminate. The spinodal field defined by the inflection points in the
Gibbs free energy of mixing curve thus represents a “kinetic limit” and not a
phase boundary in the proper sense of the term.
The Gibbs free energy variation connected with compositional fluctuations
in the vicinity of X(1)m (figure 3.12B) peaks near the so-called “critical nucleation
radius” (rc ) and then decreases progressively with the increase of (compositional)
distance. Let us analyze the significance of rc from a microscopic point of view.
A generic mixture (A,B)N in which AN is present at (macroscopic) concentration
X(1)m and BN is at a concentration of (1 ⫺ X(1)m ) may exhibit compositional
fluctuations around X(1)m, on the microscopic scale. If these fluctuations are such
that ⌬XAN at X(1)m is higher than ⌬X rc, the energy threshold of the process will
Thermochemistry of Crystalline Solids / 179
Figure 3.12 Energy relationships between solvus and spinodal decompositions. (A) Por-
tion of Gibbs free energy of mixing curve in zone between binodal (Xb) and spinodal (Xs)
points. (B) Gibbs free energy variation as a consequence of compositional fluctuations
around intermediate points X(1)m and X(2)m.
g(X ) + K ( ∇X )2 dV ,
G= ∫ (3.212)
V
where the first term, g(X ), represents the Gibbs free energy of an infinitesimal volume of
mixture with uniform composition, and the second term, K(∇X )2, defines a local energy
gradient that is always positive because K is inherently positive. The expansion of g(X ) in
the vicinity of a mean composition X 0 is given by the following expansion truncated at
the third term:
180 / Geochemistry of Crystal Phases
dg 2 d 2g
g ( X ) = g ( X 0 ) + ( X − X 0 ) + (X − X 0 )
1
+ L. (3.213)
dX X 2 dX 2 X
0 0
The Gibbs free energy difference between the system with compositional fluctuations
and an analogous homogeneous system of composition X 0 is given by
∆g(X ) + K ( ∇X )2 dV ,
∆G = ∫ (3.214)
V
where
X − X 0 = A sin βz , (3.216)
where z is the scalar component of the wave vector  along the decomposition direction
z, and A is a constant term. The Gibbs free energy variation due to periodic fluctuation is
A2 d 2 g
∆G = 2
+ 2 Kβ 2 . (3.217)
4 dX
Because K is inherently positive, the stability conditions of the mixture with compositional
fluctuations can be deduced by the value of the differential term d 2g/dX 2. The spinodal
locus is defined by d 2g/dX 2 ⫽ 0; if d 2g/dX 2 ⬎ 0, then ⌬G ⬎ 0, and the mixture is stable
with respect to infinitesimal fluctuations (i.e., region between binodes and spinodes, point
X(1) in figure 3.12A). However, if d 2g/dX2 ⬍ 0, then ⌬G ⬍ 0 for compositional fluctuations
of wavelength overcoming the critical value
12
2π 8π 2 K
λ critical = = − 2 .
β critical
(
d g dX 2 ) X0
(3.218)
d ( µ A − µB )
JA − JB = M , (3.219)
dX
Thermochemistry of Crystalline Solids / 181
where A and B are the chemical potentials of components A and B, JA and JB are their
fluxes, and M is the atomic mobility per unit volume. The diffusion equation along z is
dX dJ d 2g d 2X d 4X
=− = M − 2 MK 4 , (3.220)
dt dz dX 2 dz 2 dz
whereJ is the interdiffusional flux (J ⫽ JA ⫽ ⫺JB ) (note that eq. 3.220 can be generalized
to all directions by substituting the ∇ 2X operator for the second derivative in dz; see eq.
8 in Cahn, 1968).
Because mobility term M is always positive, the diffusion coefficient M(d 2g/dX 2 ) de-
pends on the sign of the second derivative of the Gibbs free energy. Within the spinodal
field, (d 2g/dX 2 ) is negative and the diffusion coefficient is thus also negative. This means
that there is “uphill” diffusion: the atoms of a given species migrate from low to high
concentration zones. The final result is a clustering of atoms of the same species. Neverthe-
less, if we are in the compositional zone between spinode and binode, (d 2g/dX 2 ) is posi-
tive, the diffusion coefficient is then also positive, and diffusion takes place from high to
low concentration zones (nucleation and growth). The two different accretion processes are
shown in figure 3.13 as a function of time (and distance).
Let us now consider the process from a thermal point of view (figure 3.14)
and imagine freezing the system from point p to point p⬘ (corresponding to tem-
perature T⬘): we begin to see unmixing phenomena in p⬘. Actually it may happen
that the mixture remains metastably homogeneous (i.e., the critical nucleation
radius is not overcome) down to a point p⬙ (temperature T⬙) at which we effec-
tively observe decomposition processes taking place (at p⬙ the system enters the
182 / Geochemistry of Crystal Phases
Figure 3.16 Solvus and spinodal decomposition fields as a function of elastic strain.
(a) “Strain-free” or “chemical” solvus; (b) strain-free spinodal; (c) “coherent” solvus;
(d) coherent spinodal. From Ganguly and Saxena (1992). Reprinted with permission of
Springer-Verlag, New York.
184 / Geochemistry of Crystal Phases
η2 E y
(X − X )
2
∆g elastic = ,
(1 − v ) 0 (3.221)
A2 d 2 g η2 E y
∆G = + 2 Kβ 2
+ .
4 dX 2 1− v( )
(3.222)
The energy of elastic strain modifies the Gibbs free energy curve of the mixture,
and the general result is that, in the presence of elastic strain, both solvus and
spinodal decomposition fields are translated, pressure and composition being
equal, to a lower temperature, as shown in figure 3.16.
CHAPTER FOUR
Extended defects are primarily composed of linear dislocations, shear planes, and
intergrowth phenomena. Figure 4.1A and B, for example, show two types of lin-
ear dislocation: an edge dislocation and a screw dislocation.
Extended defects interrupt the continuity of the crystal, generating crystal
subgrains whose dimensions depend, in a complex fashion, on the density of
extended defects per unit area. Table 4.1 gives examples of reported dislocation
densities and subgrain dimensions in olivine crystals from the San Carlos perido-
tite nodules (Australia). Assuming a mean dislocation density within 1.2 ⫻ 105
and 6 ⫻ 105 cm⫺2, Kirby and Wegner (1978) deduced that a directional strain
pressure of 35 to 75 bar acted on the crystals prior to their transport to the surface
by the enclosing lavas.
The presence of extended defects does not significantly affect the energetic
properties of crystalline compounds, because the concentration of extended de-
fects necessary to create significant energy modifications is too high. For instance,
Wriedt and Darken (1965) observed deviations in the solubility of nitrogen in
steels induced by an extremely high dislocation density: 1011 cm⫺2 was the limit
at which the solubility modification could be measured. Such densities are not
normally observed in natural phases, whose dislocation densities normally range
between 104 and 109. Kohlstedt and Vander Sande (1973) observed dislocation
/ 185
186 / Geochemistry of Crystal Phases
Figure 4.1 Examples of extended defects in crystals: edge dislocation (A) and screw dis-
location (B).
Table 4.1 Dislocation densities (cm⫺2) and mean subgrain dimensions (cm) in San Carlos
olivines (from Kirby and Wegner, 1978). n.d. ⫽ not determined.
Point defects exist in crystals at equilibrium conditions. They primarily affect some
chemical-physical features of crystalline substances, such as electronic and ionic
conductivity, diffusivity, light absorption and light emission, etc., and, in subordi-
nate fashion, their thermodynamic properties. As already noted, our idealized
picture of a crystal has little to do with reality. As extended defects interrupt the
symmetry of crystals, point defects modify their chemistry significantly, inducing
defect equilibria with coexisting phases and modifying stoichiometry. The disor-
der induced by the presence of point defects is grouped into two main categories:
intrinsic disorder and extrinsic disorder.
MO → VM
+ VO⋅⋅ (4.1)
or
×
MM + O O× → VM
+ VO⋅⋅ (4.2)
or
0 → VM
+ VO⋅⋅ , (4.3)
where the superscript ⫻ indicates neutrality relative to the ideal crystal, | stands
for negative excess charge taken by the cation vacancy, · stands for a positive
excess charge taken by the oxygen vacancy, and the subscripts indicate the lattice
position of the defect.
Let us now imagine moving a generic cation M from its normal position M
(M ⫻
M ) to an interstitial position i, leaving the previously occupied cationic site
empty:
×
MM → M ⋅⋅i + VM
. (4.4)
The equilibrium thus established is a “Frenkel defect.” In both the Schottky and
Frenkel equilibria, the stoichiometry of the crystal is unaltered (figure 4.2). As-
suming that the thermodynamic activity of the various species obeys Raoult’s law,
thus corresponding to their molar concentrations (denoted hereafter by square
brackets), the constant of the Schottky process is reduced to
[ ][ ]
K S = VM
VO⋅⋅ , (4.5)
Figure 4.2 Intrinsic Schottky (A) and Frenkel (B) defects in a MO crystal.
Some Concepts of Defect Chemistry / 189
[ ][ ]
K F = M ⋅⋅i VM
(4.6)
(because the defect concentrations are extremely low—i.e., generally lower than
10⫺4 in terms of molar fractions—the activity of ionic species at their normal
positions is assumed to be 1).
1
O O× → O 2 + VO⋅⋅ + 2e (4.7)
2 (gas)
and
1
×
MM + O O× → O 2 + M ⋅⋅i + 2e . (4.8)
2 (gas)
The first process produces doubly ionized positive oxygen vacancies and electrons
(e|) and the second produces doubly ionized positive cation interstitials and electrons.
The equilibrium constants of the two processes are given by
[ ][ ]
K 7 = VO⋅⋅ e PO1 2
2
2
(4.9)
and
[ ][ ]
K 8 = M ⋅⋅i e PO1 2 .
2
2
(4.10)
[ ] [ ]
2 VO⋅⋅ = e (4.11)
(i.e., the molar concentration of electrons must be twice the molar concentration
of doubly ionized positive vacancies). From equations 4.9 and 4.11 we obtain
[e ] = 2
13
K 71 3 PO−1 6
2 (4.12)
190 / Geochemistry of Crystal Phases
and
K 71 3 −1 6
[ ]
VO⋅⋅ = P .
41 3 O2
(4.13)
[ ] [ ]
2 M ⋅⋅i = e , (4.14)
[e ] = 2
13
K 81 3 PO−1 6
2 (4.15)
and
13
[M⋅⋅ ] = K4
i
8
13
PO−1 6 .
2
(4.16)
At PO2 higher than P*O 2 , we observe the formation of doubly ionized cationic
vacancies and positive electronic vacancies (usually defined as electron holes by
symbol h· ):
1
O2 → O O× + VM
+ 2h ⋅ . (4.17)
2 (gas)
[ ][ ]
h ⋅ PO−1 2
2
K 17 = VM
(4.18)
2
[h ⋅ ] = 2 13 13 16
K 17 PO
2 (4.19)
and
13
[ ]
VM =
K 17
P1 6 .
41 3 O2
(4.20)
Diagrams such as that presented in figure 4.3 are commonly used in the inter-
pretation of the defects in crystalline substances. Based on the equations above,
the defect concentration depends on PO2 in an exponential fashion (proportional
to PO⫺12 / 6 and PO1/26 in low and high PO2 ranges, respectively; because the plot is
logarithmic, the exponent becomes the slope of the function). Outside the intrin-
sic stability range, the stoichiometry of the crystal, as we have seen, is altered by
exchanges with the surrounding atmosphere. However, the above defect scheme
is an oversimplification of the complex equilibria actually taking place in hetero-
geneous systems. Indeed, defect processes are not simply ruled by the partial pres-
sure of oxygen, but by the chemistry of all phases present in the system. For
example, in the presence of a silica-rich phase, forsterite (Mg2SiO4 ) can incorpo-
rate SiO2 in excess with respect to its normal stoichiometry through the defect
192 / Geochemistry of Crystal Phases
equilibrium
2 SiO 2 ⇔ 2 VMg
+ Si ⋅⋅⋅⋅
i + Si Si + 4 O O
× ×
(4.21)
SiO 2 ⇔ 2 VMg
+ Si ×Si + 2 VO⋅⋅ + 2 O O× (4.22)
(cf. Smith and Stocker, 1975). Based on the experimental data of Pluschkell and
Engell (1968), the silica excess in this way reaches 2.4% at 1400 °C.
×
2 Fe Fe → VFe
+ 2 Fe ⋅Fe (4.23)
Sm 4 (SiO 4 ) 3 + 4 Mg Mg
×
+ 2 Si ×Si + 8 O O× ⇔ 4 Sm ⋅Mg + 2 VM
+ SiO 2 (4.24)
(Morlotti and Ottonello, 1984). In both cases, the crystal maintains its electroneu-
trality, but locally we observe the presence of cations whose charge (3⫹ in this
case) is not that of the normal constituent (2⫹) and which, in the case of equilib-
rium 4.24, do not correspond to a major component. These “point impurities,”
normally present in the crystal at equilibrium, significantly affect the chemical-
physical properties of the phase, acting, for instance, as activators in light
emission processes or as centers of heterogeneous nucleation (see section 3.12).
Processes such as 4.24 also regulate the solubility of trace components in rock-
forming minerals; the implications of these processes in trace element geo-
chemistry are discussed in chapter 10.
We have seen that defect concentrations in crystals may be described with the aid
of electroneutral equilibria involving species of differing charges. We have also
seen that defect concentrations are related to the partial pressure of oxygen in the
Some Concepts of Defect Chemistry / 193
0 → VM
+ VO⋅⋅ (4.25)
and
K S = VM
[ ][ ]
VO⋅⋅ , (4.26)
Gibbs free energy variation connected with the defect process is*
∆G S = − kT ln K S , (4.27)
×
Mg Mg + O O× → VM
+ VO⋅⋅ + MgO (surface) . (4.28)
To subtract the cation Mg2⫹ from its lattice position in the crystal and to bring
it to the surface, we must work against the static potentials (coulombic plus repul-
sive plus dispersive) at the Mg site. In terms of energy, this work corresponds to
half the lattice contribution of Mg2⫹ (in the Mg site of interest—i.e., M1 or M2;
see section 5.2) to the bulk static energy of the phase (see also section 1.12):
1
∆EU ,i = − ( EC ,i + E R,i + E D,i ). (4.29)
2
* Because the energy is defined at the atom scale, the standard state is implicit in the definition
and the zero superscript is hereafter omitted.
194 / Geochemistry of Crystal Phases
Subtracting a charge in a given lattice position also contributes to the defect en-
ergy with an induced polarization term whose significance was described in sec-
tion 1.19:
1
∆E P = − eΦe . (4.30)
2
Table 4.2 lists defect energies calculated with the method described above in fayal-
ite and forsterite crystals. Note that the energies obtained are on the magnitude
of some eV—i.e., substantially lower than the energies connected with extended
defects. Note also that, ionic species being equal, the defect energies depend on
Table 4.2 Defect energies in forsterite and fayalite based on lattice energy calculations. I ⫽
ionization potential; E ⫽ electron affinity; Ed ⫽ dissociation energy for O2; ⌬H ⫽ enthalpy of
defect process (adapted from Ottonello et al., 1990).
FAYALITE
(1) 22.50 ⫺4.89 0.89 0.31 ⫺15.31 - - - 3.50
(2) 24.60 ⫺4.81 0.87 0.28 ⫺14.84 - - - 6.10
(3) ⫺27.34 6.14 ⫺1.07 - 12.44 - 7.46 2.79 0.42
(4) ⫺27.32 7.11 ⫺1.15 - 12.65 - 7.46 2.79 1.54
(5) ⫺52.20 12.8 ⫺2.06 - 24.68 - 14.92 5.58 3.72
(6) 11.60 5.59 ⫺0.87 0.28 ⫺42.94 30.64 - - 4.30
(7) 15.80 5.75 ⫺0.72 0.22 ⫺42.00 30.64 - - 9.77
(8) ⫺18.67 ⫺6.70 1.01 0.62 ⫺7.66 30.64 - - ⫺0.76
(9) ⫺20.17 ⫺6.59 0.99 0.56 ⫺7.42 30.64 - - ⫺1.99
(10) ⫺1.07 - - - - - - - ⫺1.07
(11) ⫺1.03 - - - - - - - ⫺1.03
FORSTERITE
(12) 23.44 ⫺4.59 0.11 - ⫺17.14 - - - 1.82
(13) 23.44 ⫺6.29 2.08 - ⫺17.14 - - - 2.09
(14) 25.16 ⫺4.46 0.13 - ⫺16.46 - - - 4.37
(15) 55.39 ⫺12.71 1.48 - ⫺26.22 - ⫺7.46 ⫺2.79 7.69
(16) 54.73 ⫺14.51 1.54 - ⫺26.22 - ⫺7.46 ⫺2.79 4.89
(17) 52.73 ⫺13.10 1.50 - ⫺26.00 - ⫺7.46 ⫺2.79 4.89
⫻
(1) Fe M1 → V||M1 ⫹ Fe2⫹ (surface)
⭈ ⫹ Fe |Si → {FeM1
(10) Fe M1 ⭈ Fe|Si}⫻
(2) Fe M2 → V||M2 ⫹ Fe2⫹
⫻ ⭈ ⫹ Fe|Si → {Fe M2
(11) Fe M2 ⭈ Fe|Si}⫻
(surface)
Table 4.3 Energy of Schottky and Frenkel equilibria in halides, oxides, and sulfides. (1)
Kröger (1964); (2) Barr and Liliard (1971); (3) Greenwood (1970).
⌬HD , ⌬HD , K 0,
Compound Process Calculated (eV) Observed (eV) Observed
the lattice positions of the defects themselves. For instance, creation of a cationic
|| ||
vacancy on site M (V M1 ) requires less energy than that necessary to create V M2.
Also, the amount of energy required to create an oxygen vacancy on site O3
·· ) is lower than that necessary to create VO1
(VO3 ·· or VO2
·· .
Table 4.2 lists energy values assigned to point impurities such as FeM1 and
FeM2 . In the calculation of their energies, an ionization term necessary to subtract
one electron from Fe2⫹ has also been added (ionization potential). With defect
notation, this process can be expressed as
⋅ ⋅
1 + h → Fe M 1 .
×
Fe M (4.31)
Na ×Na → VNa
+ Na (gas)
+
( + 4.80 eV) (4.32)
196 / Geochemistry of Crystal Phases
×
Cl Cl ⋅ + Cl −
→ VCl ( +5.14 eV)
(gas) (4.33)
+
Na (gas) + Cl (gas)
−
→ Na ×Na + Cl Cl
×
( −7.82 eV) (4.34)
0 → VNa
⋅
+ VCl ( +2.12 eV) (4.35)
Note that in this case the energies of processes 4.32 and 4.33 are doubled with
respect to the energy necessary to bring the ion to the surface of the crystal;
moreover, the energy of process 4.34 is equal to the lattice energy of the crystal.
For comparative purposes, table 4.3 lists defect energies (enthalpies) of
Schottky and Frenkel processes in halides, oxides, and sulfides. The constant K 0
appearing in the table is the “preexponential factor” (see section 4.7) raised to a
power of 1/2.
As we have seen, a given enthalpy term can be associated with each defect process
(tables 4.2 and 4.3). The defect process generally also involves an entropic term,
which may be quantified through
∆S
exp D = g1 g 2 f , (4.36)
k
where g1 and g2 are the electronic degeneracies of the atoms participating in the
defect process and f is a vibrational term whose nature will be discussed later.
Because a defect entropy (⌬SD ), a defect enthalpy (⌬HD ), and a defect vol-
ume (⌬VD ) exist, the Gibbs free energy of the defect process is also defined:
The enthalpic term ⌬HD is the dominant one in equation 4.37 and determines the
variation of the defectual concentration with T. For instance, for a Schottky de-
fect we can write
∆HS
K S = K 0 exp − , (4.38)
kT
∆S
K 0 = exp S , (4.39)
k
where ⌬SS is the amount of entropy associated with the Schottky process.
If we assume, at first approximation, that ⌬HS does not change with T, it is
obvious that the increase in defect concentration with T is simply exponential.
Table 4.4, for instance, lists Schottky defect concentrations calculated in this way
at various T by Lasaga (1981c) for NaCl and MgO, assuming defect enthalpies
of 2.20 and 4.34 eV, respectively.
The approximation discussed above is in fact rather rough, and there are actually two
general cases.
1. A linear variation of the Gibbs free energy of the defect process with T, of type
in which ⬘ is the slope of the function. In this case, an Arrhenius plot of the
constant of the defect process results in a straight line:
1
ln K D ∝ . (4.41)
T
−β ′
K 0 = g1g2 f exp . (4.42)
k
NaCl MgO
| | ||
T (K) T (°C) [V Na] [V Na]/cm3 [V Mg] [V ||Mg]/cm3
0 ⫺273 0 0 0 0
25 298 2.5 ⫻ 10⫺19 5.6 ⫻ 103 2.1 ⫻ 10⫺37 1.1 ⫻ 10⫺14
200 473 1.9 ⫻ 10⫺12 4.2 ⫻ 1010 7.9 ⫻ 10⫺24 4.2 ⫻ 10⫺1
400 673 5.9 ⫻ 10⫺9 1.3 ⫻ 1014 5.8 ⫻ 10⫺17 3.1 ⫻ 106
600 873 4.5 ⫻ 10⫺7 1.0 ⫻ 1016 3.0 ⫻ 10⫺13 1.6 ⫻ 1010
800 1073 6.8 ⫻ 10⫺6 1.5 ⫻ 1017 6.5 ⫻ 10⫺11 3.5 ⫻ 1012
1000 1273 Melt 2.6 ⫻ 10⫺9 1.4 ⫻ 1014
1500 1773 Melt 6.9 ⫻ 10⫺7 3.7 ⫻ 1016
2000 2273 Melt 1.6 ⫻ 10⫺5 8.6 ⫻ 1017
198 / Geochemistry of Crystal Phases
∆H D = ∆H D0 + βT . (4.43)
In this case, the constant of the process does not lead to a straight function in an
Arrhenius plot and the preexponential factor varies with T according to
K 0 = g1 g2 f T β k . (4.44)
∆GS
ln K S = − (4.45)
kT
∂ ln K S 1 ∂∆GS ∆V
=− =− S , (4.46)
∂P T ,X kT ∂P T ,X kT
where ⌬VS is the volume variation associated with the formation of the defect.
The value of ⌬VS is generally positive for Schottky defects in ionic crystals, be-
cause the lattice around the vacant sites undergoes relaxation due to static effects.
The corresponding ⌬VF is generally lower for Frenkel defects. It follows that the
P effect on the defect population is more marked for Schottky disorder than for
Frenkel disorder. Because ⌬VS is inherently positive, the defect population gener-
ally decreases with increasing P.
A + B → AB .( ) (4.47)
If we imagine that both defects A and B and associate defect (AB) are distributed
in conditions of complete configurational disorder, and if we substitute thermody-
namic activity for the molar concentration, we have
Some Concepts of Defect Chemistry / 199
K A,B =
[( AB)] .
[ A ][B]
(4.48)
∆G AB = − kT ln K AB = ∆H AB − T∆SAB . (4.49)
q2
∆H AB,coulombic = − , (4.50)
ε sr
Z
∆S AB,conf = k ln , (4.51)
σ
where Z is the number of ways in which the associate may be formed and is a
symmetry factor. For an associate defect (AB), Z is the number of equivalent
sites that can be occupied by B at the shortest distance from A (or by A at the
shortest distance from B). For the same associate defect (AB), ⫽ 1, but becomes
2 for an associate dimer (AA), 3 for a trimer (AAA), and so on.
The vibrational entropy (⌬SAB,vibr ) contributing to the bulk entropy of defect
process arises from the variation of the vibrational spectrum of the crystal in the
neighborhood of the associate:
* Actually, as pointed out by Kröger (1964), at small distances the use of the static dielectric
constant based on a continuum approach to interaction energy is not justified, because one would
rather adopt the permittivity of a vacuum. However, using the static dielectric constant leads to
underestimation of the binding energy by 10 to 15%, counterbalancing the fact that short-range
repulsive forces are neglected.
200 / Geochemistry of Crystal Phases
∆S AB,vibr = k ln f (4.52)
v0, j
f = ∏ v ′j
, (4.53)
j
where v 0,j and v⬘j are vibrational frequencies related, respectively, to free and asso-
ciated defects (see Kröger, 1964, for a more detailed treatment).
For a pair of Schottky defects with, respectively, x and y first neighbors and v⬘ and v⬙
vibrational frequencies, vibrational term f has the form
x y
v v
f = 0 0 . (4.54)
v ′ v ′′
v0 v
= 0 = 2 (4.55)
v′ v ′′
x = y = 6, (4.56)
from which we obtain f ⫽ 4096. If the vibrational frequency of the defects increases as a
result of association, then f ⬎ 1, and the vibrational entropy also increases. The opposite
happens when f ⬍ 1.
In the case of a Frenkel defect pair, if vi is the frequency of interstitial atom i and v⬘i
is the vibrational frequency of the Z first neighbors surrounding it, we have
z x
v v v
f = 0 0 0 . (4.57)
vi vi′ v ′
Zf ∆H AB
K AB = exp − . (4.58)
σ kT
For the simple associate defect (AB), ⫽ 1 and the relative concentrations of
associate defect (AB) and free defects A and B are ruled by
and
Some Concepts of Defect Chemistry / 201
Introducing fractions
β AB =
[( AB)] ; βA =
[A] ; βB =
[B] ,
[A] [A] [B]
(4.61)
total total total
we have
β AB ∆H AB
βA βB
= A [ ] total
Zf exp −
kT
. (4.62)
c= A[ ] total
[ ]
= B
total
, (4.63)
β AB
=1 (4.64)
β A βB
and
∆H AB
= ln c + ln Zf . (4.65)
kT 0.5
Figure 4.4 plots relative concentrations of associate defect (AB) and free de-
fects A and B calculated as functions of ⌬HAB /kT for various values of c, with
Z ⫽ 4 and f ⫽ 1. Note that when the enthalpy change connected with the associa-
tion process is low (and/or T is high) all defects are practically free, whereas for
high (negative) values of ⌬HAB /kT all defects are associated. The dashed line in
figure 4.4 represents the equality condition outlined by equation 4.65.
Table 4.5 lists the energies of associative processes in halides and simple sul-
fides. As already mentioned, the enthalpy modification resulting from coulombic
interactions is generally higher with respect to the values resulting from a detailed
calculation. Table 4.5 also lists association energies involving point impurities. We
will see in chapter 10, when treating the stabilization of trace elements in crystals,
that association energy plays an important role in trace element distribution pro-
cesses.
202 / Geochemistry of Crystal Phases
Figure 4.4 Relative concentrations of free and associate defects for various values of total
defect concentration c, with Z ⫽ 4 and f ⫽ 1 (from Kröger, 1964; redrawn). Dashed line:
equality conditions in populations of free and associate defects.
n Si 1
ξ = − (4.66)
n Si + nFe 3
Some Concepts of Defect Chemistry / 203
Table 4.5 Energies of associative processes in halides and simple sulfides. Values in eV
(from Kröger, 1964; modified).
⫺⌬HAB , ⫺⌬HAB ,
Associative Coulombic Total ⫺⌬HAB ,
Compound Process Calculated Calculated Experimental K0
and
nO 4
η= − , (4.67)
n Si + nFe 3
where the n-terms are the moles of the various elements. Obviously, in the absence
of extrinsic disorder, parameters and take on a value of zero. Figure 4.5A
shows the Gibbs phase triangle for the Fe-Si-O system. The (stoichiometric) com-
pound Fe2SiO4 , coexisting with SiO2 and Fe, with Fe and FeO, or with FeO and
Fe3O4 , is represented by a dot. Actually this scheme is not of general validity, and,
in the presence of extrinsic disorder, the compositional field of intrinsic stability is
expanded, as shown in figure 4.5B.
Based on thermogravimetric experiments on the compound Fe2SiO4 at vari-
ous T and P O2 conditions, Nakamura and Schmalzried (1983) established that
the extrinsic disorder of fayalite is conveniently represented by the equilibrium
×
×
10 Fe Fe + Si Si
×
+ 2 O 2(gas) ⇔ 3VFe
+ 6 Fe⋅Fe + Fe Si
Fe⋅Fe + FeSiO 4 . (4.68)
The defect scheme shown in equation 4.68 was later confirmed by electromotive
force measurements with galvanic cells (Simons, 1986) and by diffusivity mea-
204 / Geochemistry of Crystal Phases
Figure 4.5 Nonstoichiometry of compound Fe2SiO4 (fayalite) in the Fe-Si-O system. (A)
Usual representation of Fe2SiO4 as a discrete point in Gibbs space. (B) Defect equilibria
expand intrinsic stability field of phase through variation in nonstoichiometry parameters.
From A. Nakamura and H. Schmalzried, On the nonstoichiometry and point defects in
olivine, Physics and Chemistry of Minerals, 10, 27–37, figure 1, copyright 䉷 1983 by
Springer Verlag. Reprinted with the permission of Springer-Verlag GmbH & Co. KG.
Table 4.6 shows the energy of extrinsic disorder calculated for the solid mixture
(Fe, Mg)2SiO4 at T ⫽ 1200 °C, based on the defect scheme of equation 4.68 and
on the defect energies of table 4.2.
We note that the defect energy contribution associated with extrinsic disorder
varies considerably as a function of the partial pressure of oxygen of the system.
These energy amounts may significantly affect the intracrystalline disorder, with
marked consequences on thermobarometric estimates based on intracrystalline
distribution. As we will see in detail in chapter 10, most of the apparent complexi-
ties affecting trace element distribution may also be solved by accurate evaluation
of the defect state of the phases.
(1) (2)
for vacancy migration and interstitial migration. Note that the required energies
are subordinate with respect to the energies of defect formation (cf. tables 4.3
and 4.7).
Let us now consider the crystal MO. If the diffusion takes place by migration
of cationic vacancies, the number of atoms that undergo the process depends on
||
the vacancy concentration [V M ] and the thermal state of single atoms M (the
jump takes place only whenever atom M in the neighborhood of the vacancy
has sufficient energy to perform it). The diffusion coefficient associated with the
vacancy migration process is given by
H
D=
d 2 v0
3
[ ]
VM exp −
kT
VM
(4.71)
(Jost, 1960; Lasaga, 1981c), where d is the interionic distance between vacant and
occupied sites, and v 0 is a specific vibrational frequency of the crystal (Einstein
frequency, or a fraction of the Debye frequency; see Lasaga, 1981c).
A similar relation holds for anionic vacancies and for diffusion through inter-
||
stitial migration (replace [V M ] and HV M
||
by [VO··] and HVO·· , and so on).
Because at high T the vacancy concentration depends exponentially on T, i.e.
Some Concepts of Defect Chemistry / 207
∆HS ∆SS
[V ] = K
M
12
S = exp − exp
2 kT 2k
, (4.72)
Q
D = D0 exp − D , (4.73)
kT
where QD represents the “bulk activation energy of the diffusion process,” com-
posed of the enthalpy of vacancy formation plus the enthalpy of vacancy mi-
gration:
∆H S
QD = + HVM
, (4.74)
2
d 2 v0 ∆S
D0 = exp S . (4.75)
3 2k
Equation 4.75 finds its application in the region of intrinsic disorder (a similar
equation can be developed for Frenkel defects), where Schottky and Frenkel de-
fects are dominant with respect to point impurities and nonstoichiometry.
With decreasing temperature, as we have seen, the intrinsic defect population
decreases exponentially and, at low T, extrinsic disorder becomes dominant.
Moreover, extrinsic disorder for oxygen-based minerals (such as silicates and ox-
ides) is significantly affected by the partial pressure of oxygen in the system (see
section 4.4) and, in the region of intrinsic pressure, by the concentration of point
impurities. In this new region, term QD does not embody the enthalpy of defect
formation, but simply the enthalpy of migration of the defect—i.e.,
QD = HVM
. (4.76)
Figure 4.7 Arrhenius plot of cationic diffusivity in olivines. Reprinted with permission
from Buening and Buseck (1973), Journal of Geophysical Research, 78, 6852–6862, copy-
right 䉷 1973 by the American Geophysical Union.
×
Fe M 1 → VM 1 + Fe (surface)
2+
( +2.09 eV ) . (4.77)
(To evaluate the energy of the Schottky process, based on the energies of table
4.2, the concomitant process of anion vacancy formation and the recombination
energies of cation and anion on the surface of the crystal must be added to pro-
cess 4.77.)
Not all crystalline compounds exhibit slope modifications similar to those
shown for olivine in figure 4.7. For instance, figure 4.8 shows diffusivity plots
for feldspars according to the revision of Yund (1983). The temperature ranges
investigated by various authors are sufficiently wide, but in any case we do not
observe any slope modifications imputable to modified defect regimes.
However, it must be emphasized that interpretation of elemental diffusion in
feldspars is complicated by the structural state of the polymorphs, which vary in a
complex fashion with temperature, chemistry and re-equilibration kinetics. These
complexities also account for the controversies existing in the literature regarding
diffusion energy in these phases (see also, incidentally, figure 4.8). Elemental
diffusivity data for rock-forming silicates are listed in table 4.8.
Whenever diffusivity rates are not experimentally known and cannot be esti-
mated by static potential calculations, approximate values can be obtained by
empirical methods. The most popular of these methods establishes a linear rela-
tionship between the enthalpy of the Schottky process (and enthalpy of migra-
tion) and the melting temperature of the substance (expressed in K):
210 / Geochemistry of Crystal Phases
Figure 4.8 Arrhenius plot of elemental diffusivity in feldspars. From Yund (1983). Re-
printed with permission of The Mineralogical Society of America.
∆H S ≈ 2.14 × 10 −3 Tf ( eV ) (4.78)
and
∆H VM ≈ 0.014Tf ( eV ) . (4.79)
Equations 4.78 and 4.79 are valid for ionic solids. The analogous relations for
metals are
and
∆H VM ≈ 6 × 10 − 4Tf ( eV ) . (4.81)
The significance of empirical equations 4.78, 4.79, 4.80, and 4.81 has not yet
been clarified in a satisfactory fashion. The enthalpy of Frenkel processes in met-
als can be related to the Debye temperature of the solid through
Some Concepts of Defect Chemistry / 211
Table 4.8 Elemental diffusivity in silicates (from Lasaga, 1981c; Morioka and Nagasawa,
1991).
T-Range, D0, QD ,
Compound Element (°C) (cm2s⫺1) (eV) Reference
Mg2SiO4 FeM1 1125–1200 5.6 ⫻ 10⫺2 2.73 Buening and Buseck (1973)
Mg2SiO4 FeM1 1100–1125 5.9 ⫻ 10⫺7 1.38 Buening and Buseck (1973)
Mg2SiO4 FeM2 900–1000 1.1 ⫻ 10⫺2 2.61 Misener (1974)
Mg2SiO4 Ni 1250–1450 2.2 ⫻ 102 4.29 Morioka (1981)
Mg2SiO4 Ni 1149–1234 1.1 ⫻ 10⫺5 2.00 Clark and Long (1971)
Mg2SiO4 Co 1150–1400 2.8 ⫻ 10⫺1 3.24 Morioka (1980, 1981)
Mg2SiO4 Ca 1200–1400 7.3 ⫻ 101 4.31 Morioka (1981)
Mg2SiO4 MgM1 1300–1400 1.5 ⫻ 103 4.60 Morioka (1981)
Mg2SiO4 MgM2 1000–1150 1.8 ⫻ 10⫺8 1.48 Sockel and Hallwig (1977)
Mg2SiO4 OO1 1150–1600 1 ⫻ 10⫺4 3.32 Jaoul et al. (1980)
Mg2SiO4 OO2 1275–1628 3.5 ⫻ 10⫺3 3.86 Reddy et al. (1980)
Mg2SiO4 OO3 1472–1734 2.9 ⫻ 10⫺2 4.31 Ando et al. (1981)
Mg2SiO4 Si 1300–1700 1.5 ⫻ 10⫺6 3.94 Jaoul et al. (1981)
Fe2SiO4 MgM1 900–1100 1.5 ⫻ 10⫺2 2.16 Misener (1974)
Fe2SiO4 MgM1 1125–1200 2.9 ⫻ 10⫺2 2.55 Buening and Buseck (1973)
Fe2SiO4 MgM1 1000–1125 7.6 ⫻ 10⫺7 1.28 Buening and Buseck (1973)
Co2SiO4 Si 1200–1300 8.9 ⫻ 10⫺3 3.44 Schmaizried (1978)
KAlSi3O8 O 350–700 4.5 ⫻ 10⫺8 1.11 Giletti et al. (1978)
Adularia O 400–700 5.3 ⫻ 10⫺7 1.28 Yund and Anderson (1974)
Orthoclase Na 500–800 8.9 2.29 Foland (1974)
Orthoclase K 500–800 16.1 2.96 Foland (1974)
Orthoclase Rb 500–800 38 3.17 Foland (1974)
Orthoclase Sr 800–870 6 ⫻ 10⫺4 1.79 Misra and Venkatasubramanian
(1977)
Microcline O 400–700 2.8 ⫻ 10⫺6 1.28 Yund and Anderson (1974)
Microcline K 600–800 133.8 3.04 Lin and Yund (1972)
Microcline Sr 800–870 5 ⫻ 10⫺4 1.67 Misra and Venkatasubramanian
(1977)
NaAlSi3O8 O 350–800 2.3 ⫻ 10⫺9 0.92 Giletti at al. (1978)
NaAlSi3O8 O 600–800 2.5 ⫻ 10⫺5 1.61 Anderson and Kasper (1975)
NaAlSi3O8 Na 600–940 5 ⫻ 10⫺4 1.52 Bailey (1971)
NaAlSi3O8 Na 200–600 2.3 ⫻ 10⫺6 0.82 Lin and Yund (1972)
Biotite K 550–700 2.7 ⫻ 10⫺10 0.91 Hofman and Giletti (1970)
Anorthite O 350–800 1.4 ⫻ 10⫺7 1.14 Giletti et al. (1978)
12
∆H F
θD = 34.3 , (4.82)
MV 2 3
where ⌬HF is in cal/mole. M is the molar weight and V is the molar volume of
the crystal (Mukherjee, 1965). Equation 4.82 is related to the preceding empirical
formulations through the Lindemann equation:
Tf ∝ θ D . (4.83)
212 / Geochemistry of Crystal Phases
∂ ln D ∆V +
=− , (4.84)
∂P kT
Q D( P ) = Q D( P
r)
(
+ ∆V + P − Pr , ) (4.85)
D( P )
( )
−Q D + P − 1 ∆V +
= D0 exp
. (4.86)
kT
Table 4.9 gives a summary of diffusion data for divalent cations in aluminous
garnets, with the relative activation volume. Although estimation of activation
volume is still largely uncertain, its evaluation is essential when dealing with the
wide baric regimes encountered in petrologic studies. As shown in the third col-
umn of table 4.9, the presence of ⌬V ⫹ implies substantial modifications of QD on
the kbar scale of pressure.
The diffusion coefficient for a given ion in a crystal is determined, as we have seen,
by the atomistic properties of the ion in the structural sites where the vacancy (or
interstitial) participating in the migration process is created (see eq. 4.71). The
units of diffusion (and/or “self-diffusion”) are usually cm2sec⫺1. Fick’s first law
relates the diffusion of a given ion A ( JA ) to the concentration gradient along a
given direction X:
∂C
JA = − DA A , (4.87)
∂X T ,P
Some Concepts of Defect Chemistry / 213
∂ CA ∂ ∂ CA
= DA . (4.88)
∂ t T ,P ∂X ∂X
∂C ∂C
JA = − D˜ A ; JB = − D˜ B (4.89)
∂X T ,P ∂X T ,P
and
∂ CA ∂ ∂ CA ∂ CB ∂ ∂ CB
= D˜ ; = D˜ . (4.90)
∂ t T ,P ∂X ∂X ∂ t T ,P ∂X ∂X
Let us now consider the diffusion of two neutral species A and B in two
portions of metal slag welded at one interface (figure 4.9A). If the diffusivity of
A is higher than that of B, the portion on the right in figure 4.9A will tendentially
grow, whereas the portion on the left will become smaller (figure 4.9B). However,
if we keep the metal slag in a fixed position, the interface will be displaced to
the left (Kirkendall effect, figure 4.9C). The interdiffusion coefficient is related to
individual diffusivities according to
214 / Geochemistry of Crystal Phases
Figure 4.9 Kirkendall effect in a slag for metals A and B, with DA ⬎ DB. (A) Initial
configuration of diffusion couple. Position of interface fixed (dashed line) during diffusion
(accretion of right part and shrinking of left part of slag block). (C) Position of slag block
fixed during diffusion (interface moves leftward). From P. Haasen, Physical Metallurgy,
copyright 䉷 1978 by Cambridge University Press. Reproduced with modifications by per-
mission of Cambridge University Press.
D˜ = XA DB + X B DA . (4.91)
If, however, the diffusing species are electrically charged, the net flux at the inter-
face is effectively zero, to maintain the charge neutrality. Particularly, if A and B
are of the same charge, the flux of A will equal that of B:
∑ Ji = 0.
(4.92)
i
DA DB
D˜ = (4.93)
X A DA + XB DB
(see Ganguly and Saxena, 1987 for a more detailed derivation of equations 4.91
and 4.93).
Equation 4.90 is of particular interest when treating diffusion in nonmetallic
solids, where diffusion is primarily limited to ionic species. Note that, based on
equations 4.91 and 4.93, D̃ → DB when XB → 0. Thus, in a diluted binary mixture,
Some Concepts of Defect Chemistry / 215
the interdiffusion coefficient assumes the same value as the diffusion coefficient
for the minor component.
In interdiffusion studies involving ions of the same charge, the dependence
of the single-ion diffusivity on concentration is obtained by means of annealing
experiments involving single oriented crystals of different compositions kept in
contact along a smooth surface. The interface coincides with the Boltzmann-
Matano plane—i.e., “a plane such that the gain of a diffusing species on one side
equals its loss on the other side” (Crank, 1975):
C0
∫ X dC = 0, (4.94)
0
where C 0 is the initial concentration in the diffusing couple and X is the distance
from the Boltzmann-Matano plane.
Figure 4.10 shows profiles of Mn2⫹ and Mg2⫹ distribution in the
Figure 4.10 Distribution profiles of Mn2⫹ and Mg2⫹ along axis c in the olivine mixture
(MnxMg1⫺x)2SiO4 after couple annealing at 1200 °C for 277 hours. From Morioka and
Nagasawa (1991). Reprinted with permission of Springer-Verlag, New York.
216 / Geochemistry of Crystal Phases
C1
1 ∂X
D˜ C =C 1 = −
2t ∂C ∫ X dC , (4.95)
0
Silicates
Qualitatively and quantitatively, compounds of silicon and oxygen are the class
of substances of greatest importance in the earth’s crust and mantle, in regard to
both mass and variety of structural forms.
In silicon and oxygen compounds, Si4⫹ has coordination IV or VI. The SiO44⫺
tetrahedron is energetically quite stable, with mixed ionic-covalent bonds.
The tetrahedral coordination of the [SiO4 ]4⫺ group could be regarded as a
direct consequence of the capability of the silicon atom to form four hybrid orbit-
als sp3, each directed toward one of the apexes of a regular tetrahedron, as shown
in figure 5.1. The energy gap between 3s and 3p orbitals at Z ⫽ 14 is limited, and
thus the combination is at first sight energetically plausible. Detailed quantum
mechanics calculations confirm that the tetrahedral symmetry corresponds to the
minimum energy for Si-O bonds.
Te1 =
1
2
(s + px + py + pz ) (5.1)
Te2 =
1
2
(
s + px − py − pz ) ( 5.2 )
Te3 =
1
2
(
s − px + py − pz ) ( 5.3 )
Te4 =
1
2
( )
s − px − py + pz . ( 5.4 )
/ 217
218 / Geochemistry of Crystal Phases
Figure 5.1 Tetrahedral hybrid orbitals sp3. (A) Section along development axis. (B) Spa-
tial form and orientation of hybrid orbital Te1 (from Harvey and Porter, 1976; redrawn
and reproduced with permission).
However, things are more complex when we look at the MO energy diagram (fig-
ure 5.2) developed by Tossell et al. (1973) for the [SiO4 ]4⫺ cluster on the basis of
a density functional molecular method (MS-SCF X␣: Multiple Scattering Self
Consistent Field). The orbital with the lowest binding energy is 1t1 , which is
essentially the nonbonding AO 2p of oxygen, followed in order of increasing
energy by the two 1e and 5t2 MOs with symmetry (⯝5 eV) formed by combina-
tion of the AOs 3s and 3p of silicon with the AO 2p of oxygen, and the 4t2 MO
(⯝7 eV) also formed by combination of the AO 3p of silicon with the AO 2p of
oxygen but with symmetry (high-energy orbitals in the inner valence region are
predominantly AO 2s in character).
Determination of electron density maps for the ␣-quartz polymorph estab-
lishes that the charge transfer between silicon and oxygen is not complete and
that a residual charge of ⫹1.0 (⫾0.1) electron units (e.u.) remains localized on
silicon, whereas a charge of ⫺0.5 (⫾0.1) e.u. is localized on each oxygen atom.
The interpretation of this fact in terms of the bond ionicity is not as univocal as
it may appear at first glance.
According to Pauling (1980), electron density maps confirm the fractional ionic char-
acter ( fi ⫽ 51%) deduced from electronegativity (see eq. 1.32 in section 1.7). Pauling’s
reasoning is as follows: the length of the Si-O bond in all compounds in which the [SiO4 ]4⫺
group shares the oxygen atoms with other tetrahedrally coordinated ions is 1.610 Å
(⫾0.003; cf. Baur, 1978). Because the observed Si-O distance is about 0.20 Å lower than
the sum of the single-bond radii and there are four Si-O bonds, the bond number 1.55 is
obtained and covalence is 6.20 e.u. The charge transfer from oxygen to silicon is thus 2.2
valence electrons (i.e., 6.2 to 4). Based on the net charge of ⫹1.0 e.u. on silicon, the frac-
tional ionic character is therefore
Silicates / 219
Figure 5.2 MO energy levels for the [SiO4]4⫺ cluster based on MS-SCF X␣ calculations.
Reprinted from Tossell et al. (1973), with permission of Elsevier Science B.V.
1 + 2.2
fi = = 0.5161 ≈ 51.6%, ( 5.5)
6.2
which is in good agreement with the value of 51% deduced from electronegativity.
However, starting from the same experimental evidence, Stewart et al. (1980) adopt
the following definition of fractional ionic character:
According to Stewart et al. (1980), the fractional ionic character thus defined is in
agreement with the value deduced from electronegativity theory for molecular orbitals.
The energy of a generic atom A is defined as
220 / Geochemistry of Crystal Phases
where aA, bA, and cA are constants valid for atom A, and nA is the number of electrons in
the orbital: 0, 1, or 2. Constants bA and cA are related to ionization potential I and to the
electron affinity E:
bA = I A ( 5.8 )
cA =
1
4
(EA − I A ) . ( 5.9 )
The definition of Iczkowski and Margrave (1961) for electronegativity (see section 1.7)
gives
∂WA
χA = = bA + 2cAnA . ( 5.10 )
∂n
According to the treatment of Hinze et al. (1963), the charge transfer obeys the equal-
ity principle outlined in equation 1.31 (section 1.7). The transfer of fractional charges nA
and nB for two atoms A and B results in a stable bond with a fractional ionic character:
1 χ A − χB
fi = , ( 5.11)
2 cA − cB
where A and B are the initial electronegativities of the AOs that form the bond. In the
case of ␣-quartz, applying the electronegativity values of orbitals A and B and coeffi-
cients cA and cB, both known for states sp and sp2 and evaluable for hybrid orbitals sp3 on
the basis of bond angles (for the Si-O-Si bond between neighboring tetrahedral groups,
the bond angle is 143.68°; cf. Le Page and Donnay, 1976; Hinze and Jaffe, 1962), Stewart
et al. (1980) obtain
1 2.25 − 5.80
fi = = 0.2536, ( 5.12 )
2 −2.26 − 4.74
Disregarding the type of bond, the tetrahedral group [SiO4 ]4⫺ is usually repre-
sented in the conventional ways outlined in figure 5.3, with the oxygen atoms at
the corners and silicon at the center of the tetrahedron, and the structure of sili-
cates is normally described as a function of the relative arrangements of the
Silicates / 221
Table 5.1 Structural arrangements of groups [SiO4 ]4⫺ and [SiO6 ]8⫺;
x ⫽ known, 0 ⫽ not known (from Liebau, 1982).
[SiO4 ]4⫺ groups. Besides coordination IV, silicon also assumes coordination VI
with oxygen. Although the number of phases containing [SiO6 ]8⫺ groups is lim-
ited (about 30), the importance of octahedral coordination becomes more and
more evident with the advance of studies on high-P compounds. Both groups,
[SiO4 ]4⫺ and [SiO6 ]8⫺, exhibit various structural arrangements (see table 5.1).
For [SiO4 ]4⫺ groups, corner sharing and presence in isolated groups are the
most common structural arrangements (only fibrous SiO2 seems to exhibit an
edge-sharing configuration, although this arrangement is rather suspect; S. Mer-
lino, personal communication). For [SiO6 ]8⫺ groups, the edge-sharing configura-
tion (as in stishovite, for instance) is as common as the corner-sharing ar-
rangement.
Several elements (generally those with high field strength) exhibit isomor-
phous substitution of silicon in the tetrahedron (Be2⫹, Al 3⫹, Fe3⫹, Ga3⫹, B3⫹,
Ge4⫹, Ti4⫹, P5⫹ ). The most important diadochy is that of aluminum, which forms
pentavalent tetrahedral groups [AlO4 ]5⫺. Tetrahedral groups with isomorphous
substitution at the center of the tetrahedron are usually designated [TO4 ].
The chemical classification of silicates is based on their “multiplicity” and
type of structural arrangement. It partly differs from the classification commonly
adopted in mineralogy (table 5.2).
In mineralogy, monosilicates are called “nesosilicates” and all other oligosili-
cates are called “sorosilicates”; moreover, the term “polysilicates” is replaced by
“inosilicates.”
The multiplicity of the groups establishes the number m of [TO4 ] groups that
222 / Geochemistry of Crystal Phases
Classes
Si + Al
X Si = , (5.13)
M′
where M⬘ ⫽ M⫹ 2⫹ 3⫹ 4⫹
2 ; M ; M 0.666; M0.5 .
The so-called degree of silicification is inversely proportional to the Si:O
atomic ratio and represents the degree of condensation of the [TO4 ] groups. Three
classes with distinct silicification degrees are thus distinguished, and each class is
further subdivided into groups, as a function of the geochemical affinity of the
elements, and into subclasses, as a function of the morphology of the substance
Silicates / 223
(table 5.3). We refer readers to Liebau (1982) for a more exhaustive discussion
on classification schemes for silicates.
5.2 Olivines
Figure 5.4 (A) Olivine structure: dashed line contours elementary cell (axes c and b). (B)
Details of coordination state and distortion of M1-M2 sites; O(1), O(2), O(3) oxygens
occupy nonequivalent positions; distortion from perfect octahedral symmetry is more
marked for M2 site.
224 / Geochemistry of Crystal Phases
Olivine is the main mineral, in terms of mass and volume, of the earth’s upper mantle
(peridotites), where it is present as a mixture of the main components forsterite and fayal-
ite, in molar proportions of 85 to 95% and 15 to 5%, respectively. The intrinsic stability
limit of the mixture is assumed to coincide with the seismic discontinuity observed at a
depth of about 400 km, where olivine undergoes transition to the - and ␥-spinel poly-
morphs (Ringwood, 1969). Olivine is also a primary phase of mafic rocks (basalts, gab-
bros, dolerites), where it occurs with amounts of forsterite ranging from Fo80 to Fo50 . Its
importance (in terms of relative abundance) decreases considerably in acidic rocks (gran-
ites, syenites, rhyolites, trachytes), where fayalite prevails. Monticellite (CaMgSi2O6 ) is
normally found as a constituent in Si-Mg marbles, where it is a marker of the sanidine
facies (and probably also of the granulitic facies) of thermal regional metamorphism
(Brown, 1970). Terms of the tephroite-fayalite mixture are commonly found in Fe-Mn ore
deposits associated with skarns and in metamorphosed manganiferous sediments.
The Ni2SiO4 content in natural olivines rarely exceeds 1% in mole (weight % of NiO
⬍ 0.5), and is usually directly correlated to the forsterite content (see, for instance, Davis
and Smith, 1993). The inverse correlation is observed for Mn2SiO4 (weight % of MnO ⬍
4). Figure 5.5 shows the relative distributions of Mn2SiO4 , Ni2SiO4 , and Ca2SiO4 in effu-
sive and plutonic rocks, based on the synthesis of Simkin and Smith (1970): marked con-
trol of the mineralization environment can be seen for Ca2SiO4 . This effect is less evident
for the manganous term and is virtually absent for Ni. The relationship between amount
of Ca2SiO4 and crystallization conditions may be imputed to the pressure effect on the
solvus field existing for the join Mg2SiO4-CaMgSiO4 (cf. Finnerty, 1977; Finnerty and
Boyd, 1978; see section 5.2.5).
Table 5.5 lists silicate components that are known to crystallize with the oliv-
ine structure, according to the compilation of Ganguly (1977). The pure compo-
nents Zn2SiO4 and Cr2SiO4 have never been synthesized in the laboratory, but are
known to form the mixtures (Mg,Zn)2SiO4 and (Mg,Cr)2SiO4 , in which they
reach maximum molar concentrations of 24% (Sarver and Hummel, 1962) and
5% (Matsui and Syono, 1968; Ghose and Wan, 1974), respectively. It must be
noted that the presence of chromium as a minor component of terrestrial olivines
is restricted to kimberlitic rocks, with a maximum content of 0.15 weight %
Cr2O3 , although greater abundances (⬎0.4% Cr2O3 ) have been observed in lunar
Silicates / 225
Table 5.4 Olivine major element compositions (in weight %). Samples occur in different
types of rocks: (1) ⫽ forsterite from a metamorphosed limestone: (2) ⫽ hortonolite from an
olivine gabbro: (3) ⫽ fayalite from a pantelleritic obsidian; (4) ⫽ fayalite from an Fe-gabbro;
(5) ⫽ forsterite from a cumulitic peridotite; (6) ⫽ forsterite from a tectonitic peridotite.
Samples (1) to (4) from Deer et al. (1983); sample (5) from Ottonello et al. (1979); sample (6)
from Piccardo and Ottonello (1978).
Samples
specimens. Because chromium is present in nature in the valence states Cr2⫹ and
Cr3⫹, but occurs in olivine only (or predominantly) in the Cr2⫹ form (Burns,
1975; Sutton et al. 1993), the paucity of Cr2SiO4 in terrestrial olivines must be
ascribed to the oxidation state of the system, which is generally too high to allow
the formation of an olivine compound with stoichiometry Cr2SiO4 . Lastly, it must
be noted that the synthesis of compounds of type LiLnSiO4 or NaLnSiO4 (in
which Ln is a generic rare earth with atomic number not lower than that of Ho)
apparently contrasts with natural evidence. Although microprobe analyses report
226 / Geochemistry of Crystal Phases
Figure 5.5 Relative distributions of minor components Mn2SiO4, Ni2SiO4, and Ca2SiO4
in fayalite-forsterite mixtures (weight %).
sufficient amounts of Na2O in natural olivines, the rare earth element (REE)
abundances are generally at ultratrace levels (i.e., REE ⬍ 1 ppm; see, for instance,
Ottonello, 1980; Frey, 1982) and the relative distribution of the various terms of
the series does not reveal any crystal chemical control.
Table 5.6 Structure of olivine compounds (from Smyth and Bish, 1988). Molar volume is
expressed in cm3/mole; molar weight in g/mole; density in g/cm3; cell edges and cell volume are in
Å and Å3, respectively.
ation of density between larnite (2.969 g/cm3 ) and liebembergite (4.921 g/cm3 ),
which results from the fact that a low molecular weight is associated to the highest
molar volume (larnite) and vice versa (liebembergite).
Figure 5.6 plots the molar volumes of various olivine mixtures against the
cubes of the ionic radii of the major cations in VI-fold coordination with oxygen,
according to the compilation of Brown (1982). The figure has been partially modi-
fied to correct the volume of Ca2SiO4 on the basis of the data of Czaya (1971).
Based on figure 5.6, olivine compounds seem to obey the simple proportionality
established by Vegard’s rule (with the exception of the pure component larnite):
Vcell ∝ r 3 . ( 5.14 )
Brown (1970) calculated the following equation, which is valid for olivine
compounds (ionic radii of Shannon and Prewitt, 1969):
× 10 −24
V 0
(
= 188.32r 3 + 220.17 )N 0
v
( cm 3
)
mole , ( 5.16 )
where N 0 is Avogadro’s number and v is the number of formula units per unit cell.
Equations 5.15 and 5.16 give only approximate values for the volumes of
olivine compounds. More accurate study of binary mixtures outlines important
deviations from ideal behavior, which result in slight curves in cell parameter vs.
composition plots, as shown in figure 5.7. The greatest deviations regard cell
edges and are particularly evident for the (Mg,Ni)2SiO4 mixture.
The volume properties of crystalline mixtures must be related to the crystal chemical
properties of the various cations that occupy the nonequivalent lattice sites in variable
proportions. This is particularly true for olivines, in which the relatively rigid [SiO4 ]4⫺
groups are isolated by M1 and M2 sites with distorted octahedral symmetry. To link the
various interionic distances to the properties of cations, the concept of ionic radius is in-
sufficient; it is preferable to adopt the concept of crystal radius (Tosi, 1964; see section
1.9). This concept, as we have already noted, is associated with the radial extension of the
ion in conjunction with its neighboring atoms. Experimental electron density maps for
olivines (Fujino et al., 1981) delineate well-defined minima (cf. figure 1.7) marking the
maximum radial extension (rmax ) of the neighboring ions:
Figure 5.6 Relationship between cell volume and cubed ionic radius of main cation in
VI-fold coordination with oxygen. From Brown (1982). Reprinted with permission of The
Mineralogical Society of America.
where (r) is electron density at distance r. Equation 5.17 may be operatively translated
to the form
C1 C1
rmax = = , ( 5.18 )
Zeff Zn − S e
where C1 is a constant determined by the electronic configuration of the inner shells, and
effective nuclear charge Zeff is the actual nuclear charge minus screening constant Se (cf.
section 1.5).
The energy of the atomic orbitals in the hydrogenoid approximation is proportional
to the square of the effective charge, and is (numerically) equivalent to the ionization po-
tential
Ry hcz 2 2
ZeffC2
E( n ) = − 2
= , ( 5.19 )
n n2
Figure 5.7 Molar volumes and cell edges in binary olivine mixtures (from Akimoto et al., 1976; modified). Note conspicuous deviations from linearity
in cell edges of (Mg,Ni)2SiO4 mixture, satisfactorily reproduced by site-occupancy calculations based on parameters in table 5.7.
Silicates / 231
where R y is Rydberg’s constant, h is Planck’s constant, c is the speed of light, and z is the
ionic charge.
In light of equations 5.18 and 5.19, a relationship can be derived between ionization
potential (I ) and crystal radius (Della Giusta et al., 1990):
C
rmax = − 112 I −1 2 . ( 5.20 )
(
C2 n )
Equation 5.20 is general and differs from the proportionality of Ahrens, which is valid for
ions of homologous charge:
r ∝ I −2 . ( 5.21)
Because the effective charge may also be related to free ion polarizability ␣ fi:
−4
α i f = 4n 4 a 30 Zeff , ( 5.22 )
where a 0 is Bohr’s radius (0.529 Å), the crystal radius may be related to polarizability
through
C1
rmax = α i f 1 4 . ( 5.23 )
(
1.4142n a 30 4
)
(Della Giusta et al., 1990). Experimental observations show that equations 5.20 and 5.23
both hold.
In olivine, 24 independent interatomic distances must be evaluated in order to define
the structure. For all 24 distances (Dij ) plus cell edges and volume, regressions of the
following type may be proposed:
In equation 5.24, ij,s and j are adjustable coefficients valid for a given distance j in a
given site s, and Xi,s are the atomic fractions of i ions in site s.
Equation 5.24 is quite precise. The maximum bias between observed and predicted
distances for a population of 55 samples is in fact 0.005 Å, practically coinciding with the
experimental error. The coefficients of equation 5.24 may be directly derived from the
ionization potential and ion polarizability, based on equations 5.20 and 5.23:
−1 2
z
∑ Im
1 4
ω ij , s = K 1 j , s + K 2 j , s α if + K 3 j ,s . ( 5.25)
m =1 i
Table 5.7 Constants of equation relating coefficients ij ,s to ionization potential and
polarizability of cations in M1 and M2 sites in olivine. R is the correlation coefficient
(from Della Giusta et al., 1990).
these constants for the principal interatomic distances (excluding the [SiO4 ]4⫺ tetrahe-
dron) are listed in table 5.7.
Table 5.8 lists cell edges and fractional atomic coordinates obtained for an olivine
mixture of composition Mg1.4Ni0.6SiO4 with 53.2% of the Ni distributed in M1 and 46.8%
in M2. The values obtained by calculation almost coincide with natural observation (see
also figure 5.7).
a0 b0 c0 V
m c m c m c m c
4.7458 4.7473 10.1986 10.1968 5.9563 5.9559 288.35 288.33
x y z
m c m c m c
α V = α 0 + α 1T + α 2T −2 . (5.26)
These coefficients are valid for a pressure of 1 bar, within the stability field of
the phase.
234 / Geochemistry of Crystal Phases
Figure 5.8 Relationship between cell volume and temperature for compounds of mixture
(Mg,Fe)2SiO4. Extrapolation to melting point (Tf ) identifies the same limiting volume
(VTf ⫽ 319 Å3) for the various terms of the mixture. From Hazen (1977). Reprinted with
permission of The Mineralogical Society of America.
(1979) (cf. section 1.15.1 and eq. 1.110 and 1.111): there is an inverse linear rela-
tion between polyhedral compressibility and polyhedral volume.
Table 5.11 lists experimental bulk moduli for some olivine end-members.
Combining thermal expansion and compressibility data for compounds in
the (Mg,Fe)2SiO4 mixture, Hazen (1977) proposed the following equation of state
for olivines of the series Mg2SiO4-Fe2SiO4 :
(
Vcell + 290 + 0.17XFa − 0.006T + 0.000006T 2 1 −
) P
, (5.27)
1350 − 0.16T
Figure 5.9 Phase stability relations in binary systems Mg2SiO4-Fe2SiO4 (A) and
Mg2SiO4-Co2SiO4 (B) at T ⫽ 1000 °C.
Table 5.12 Thermochemical data for various olivine end-members (Tr ⫽ 298.15 K; Pr ⫽ 1 bar).
Listed values in kJ/mole (from Ottonello, 1987). H 0f , Tr , Pr ⫽ standard state enthalpy of formation
from the elements; UL ⫽ lattice energy; EBHF ⫽ energy of Born-Haber-Fayans cycle; EC ⫽
coulombic energy; EDD ⫽ dipole-dipole interactions; EDQ ⫽ dipole-quadrupole interactions; ER ⫽
repulsive energy.
Table 5.13 Selected standard state entropy (S 0T r , P r ; Tr ⫽ 298.15 K, Pr ⫽ 1 bar) and Maier-Kelley
coefficients of heat capacity function. References as in table 5.12. Data in J/(mole ⫻ K).
C P = K 1 + K 2T + K 3T −2 + K 4T 2 + K 5T −1 2 . (5.28)
The polynomial expansion in equation 5.28 allows the calculation of the high-T enthalpy
and entropy of the compound of interest, from standard state values, through the equa-
tions
HT0r , Pr ≡ H f0 , Tr , Pr ( 5.29 )
T
HT , Pr = HT0r , Pr + ∫ C P dT ( 5.30 )
Tr
T
dT
ST , Pr = ST0r , Pr + ∫ CP T
( 5.31)
Tr
(see section 3.7 and eq. 3.105 and 3.106 for the form of integrals 5.30 and 5.31).
We recall also that the Gibbs free energy of the compound at the T and P of interest
may be obtained by applying
P
GT , P = HT , Pr − TST , Pr + ∫ VT , P dP. ( 5.32 )
Pr
To solve equation 5.32, we must first calculate the volume of the compound at the T of
interest and P ⫽ Pr. Adopting polynomial expansions of the type shown in equation 5.26,
we have
T
VT , Pr = VT0r , Pr 1 + ∫ α (T ) dT
Tr
(5.33)
α1 2
= VT0r , Pr (
1 + α 0 (T − Tr ) + 2 T − Tr − α 2 T − Tr
2 −1 −1
) ( ) .
Silicates / 239
Integration on P depends on the form of the equation of state. For instance, adopting the
Murnaghan form:
−1 K ′
K′
VT , P = VT , Pr 1 + K P , ( 5.34 )
0
we have
P P
∫ VT , P dP = ∫ VT , P dP
Pr 1
( K ′ − 1) K0
(5.35)
K 0 K′
= VT , Pr 1 + P − 1 .
K ′ − 1 K0
We have already seen that the binary mixture (Mg,Fe)2SiO4 undergoes poly-
morphic transition with increasing pressure. The univariant curves in figure 5.9
define the loci of equality of potential for the components in the two phases at
equilibrium. The Gibbs free energy of polymorph ␣⫺ (or polymorphs ␣⫺␥ and
⫺␥) is thus identical at each point on the univariant curve. On the basis of this
principle, Fei and Saxena (1986) proposed a list of thermochemical values that
(within the limits of experimental approximation) reproduce P/X stability fields
(T ⫽ 1000 °C) for the system Mg2SiO4-Fe2SiO4 . These values are reported in
table 5.14.
When adopting the values in table 5.14, note that they refer to half a mole of
CP J/(mole ⫻ K)
H 0f,T r ,Pr S 0T r ,Pr V To r ,Pr
Compound kJ/mole J/(mole ⫻ K) 3
cm /mole K1 K3 ⫻ 10⫺7 K5 ⫻ 10⫺3 K6 ⫻ 10⫺8
-MgSi0.5O2 ⫺1072.450 44.53 20.270 108.218 0 ⫺0.7362 ⫺2.7895
-FeSi0.5O2 ⫺736.473 70.93 21.575 136.332 0 ⫺1.3198 2.2412
␥-MgSi0.5O2 ⫺1069.300 42.75 19.675 100.096 ⫺0.066 ⫺0.5184 ⫺2.4582
␥-FeSi0.5O2 ⫺738.334 68.44 21.010 130.986 0 ⫺1.2041 1.3419
Compressibility
Thermal Expansion (K⫺1) (Mbar)
C P = K 1 + K 3T −2 + K 5T −1 2 + K 6T −3 (5.36)
(see section 3.2). It is also important to note that -Fe2SiO4 is fictitious because
it never occurs in nature as a pure compound (cf. figure 5.9A).
To reproduce the binary loops of the system Mg2SiO4-Fe2SiO4 accurately, Fei
and Saxena (1986) use a (symmetric) regular mixture model for ␣--␥ phases
with the following interaction parameters (cf. section 3.8.4): W␣ ⫽ ⫺8.314 kJ/
mole; W ⫽ ⫺9.977 kJ/mole; W␥ ⫽ ⫺11.210 kJ/mole. These parameters also
refer to one-half mole of substance and must thus be multiplied by 2 to be consis-
tent with the usual stoichiometry.
Alternative thermochemical tabulations to reproduce the P/X stability fields
of the system Mg2SiO4-Fe2SiO4 may be found in Bina and Wood (1987) and Na-
vrotsky and Akaogi (1984).
(Mg,Fe)2SiO4 Mixture
Mixing properties of the binary join Mg2SiO4-Fe2SiO4 have been extensively in-
vestigated by various authors. There is general consensus on the fact that the
mixture is almost ideal and that the excess Gibbs free energy of mixing is near
zero. Electrochemical measurements by Nafziger and Muan (1967) and Kitayama
and Katsura (1968) at T ⫽ 1200 °C and T ⫽ 1204 °C, respectively (P ⫽ 1 bar),
suggest slight deviation from ideality, which can be modeled with an interaction
parameter of about 1.5 to 2 kJ/mole (see also Williams, 1971; Driessens, 1968;
Sack, 1980). Calorimetric measurements by Sahama and Torgeson (1949) indi-
cate zero heat of mixing at 25 °C. However, more recent measurements by Wood
and Kleppa (1981) show the existence of positive enthalpy of mixing at T ⫽ 970
K and P ⫽ 1 bar, reproducible through a subregular mixing model of the form
( )
H mixing = 2 WH , FeMg X Fe X Mg + WH ,MgFe X Mg X Fe , (5.37)
Silicates / 241
where WH,FeMg ⫽ 1 kcal/mole; WH,MgFe ⫽ 2kcal/mole, and XMg and XFe are mole
fractions of magnesian and ferrous components in the mixture.
As regards intracrystalline disorder on M1 and M2 sites, there is agreement
on the fact that Fe2⫹ and Mg2⫹ are almost randomly distributed between the two
sites. In figure 5.10, the intracrystalline distribution coefficient
FeM1 Mg M 2
KD = , ( 5.38 )
FeM 2 Mg M1
potential calculations predict KD values lower than 1 at low T and higher than 1
at high T. Starting from a temperature of about 600 °C (which represents a prob-
able kinetic low-T limit in intracrystalline exchanges), we see progressive “in-
crease of order” with increasing T, which is also apparent in the experimental
study of Aikawa et al. (1985). However, because intracrystalline distribution is
virtually random at all P and T conditions of interest, the entropy of mixing,
within reasonable approximation, may approach the ideal case (i.e., Sexcess mixing ⫽
0). We may therefore adopt for the mixture of interest a regular (or subregular)
mixture model of the zeroth principle. For this purpose, Chatterjee (1987) sug-
gests a Margules model:
( )
G excess mixing = X Fe X Mg WFeMg X Mg + WMgFe X Fe , (5.39)
where WFeMg ⫽ WMgFe ⫽ ⫹9 kJ/mole. This model agrees with the calorimetric
data of Wood and Kleppa (1981) but contrasts with the interaction parameter
adopted by Fei and Saxena (1986) on the basis of the polymorphic transitions
observed in the same system (cf. section 5.2.4).
Recent experiments by Akamatsu et al. (1993) show that pressure has a significant
effect on Fe-Mg intracrystalline distribution ( ∂KD /∂P ⬵ 2 Mbar⫺1 ). At the upper pressure
limit of stability of olivine (P 艐 0.12 Mbar, corresponding to a depth of about 400 km;
p-wave seismic discontinuity, cf. section 5.2.3), the increase in KD with respect to the P ⫽
1 bar condition, T and composition being equal, is about 0.24. For a Fo80 olivine, this KD
variation is similar to that induced by a T increase of about 250 °C (cf. figure 5.11).
(Mg,Ca)2SiO4 Mixture
The formation of the intermediate compound CaMgSiO4 (monticellite), with lim-
ited solubility toward the end-member components Mg2SiO4 and Ca2SiO4 , is well
known in this system. Warner and Luth (1973) experimentally investigated the
miscibility between CaMgSiO4 and Mg2SiO4 in the T range 800 to 1300 °C at
P ⫽ 2.5 and 10 kbar. The miscibility gap was parameterized by the authors with
a simple mixture model in which the interaction parameter (kJ/mole) varies with
T (K) and P (kbar) according to
2
T T PT
WT , P = 85.094 − 56.735 + 15.175 − 1.487P + 1.292 . (5.40)
1000 1000 1000
The miscibility gap extends over practically the whole compositional field from
CaMgSiO4 to Mg2SiO4 , and the maximum amount of monticellite miscible in
Mg2SiO4 is lower than 5% in molar proportions. Based on equation 5.40, the
miscibility gap should expand with increasing pressure. Figure 5.11 shows the
Gibbs free energy of mixing curves calculated at T ⫽ 600 °C and P ⫽ 1 bar, with
a static interionic potential approach. Based on the principles of minimization of
Gibbs free energy at equilibrium and of equality of chemical potentials at equilib-
Silicates / 243
Figure 5.11 Gibbs free energy of mixing in binary join Mg2SiO4-Ca2SiO4 at T ⫽ 600 °C
and P ⫽ 1 bar, calculated with a static interionic potential approach. Reprinted from
G. Ottonello, Geochimica et Cosmochimica Acta, 3119–3135, copyright 䉷 1987, with kind
permission from Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington 0X5
1GB, UK.
rium, the plotted curves define two solvi (cf. sections 2.3 and 3.11): one extending
from (Mg0.02Ca0.98 )2SiO4 to (Mg0.49Ca0.51 )2SiO4 and the other extending from
(Mg0.51Ca0.49 )2SiO4 to pure forsterite. According to the model, monticellite is thus
virtually insoluble in forsterite.
(Mg,Ni)2SiO4 Mixture
According to the experiments of Campbell and Roeder (1968), at T ⫽ 1400 °C
the (Mg,Ni)2SiO4 mixture is virtually ideal. More recent measurements by Seifert
and O’Neill (1987) at T ⫽ 1000 °C seem to confirm this hypothesis. Because the
intracrystalline distribution of Mg2⫹ and Ni2⫹ on M1 and M2 sites is definitely
nonrandom (Ni2⫹ is preferentially stabilized in M1), the presumed ideality im-
plies that negative enthalpic terms of mixing counterbalance the positive excess
entropy arising from the configurational expression
S configuration = − R ∑ X i ,M1 ln X i ,M1 + ∑ X i ,M 2 ln X i ,M 2 ( 5.41)
i i
—i.e., Sconfiguration does not reach the maximum value of equipartition (11.526
J/(mole ⫻ K)) and consequently an excess term in S appears.
244 / Geochemistry of Crystal Phases
Figure 5.12 shows how Ni2⫹ and Mg2⫹ ions are distributed between M1 and
M2 sites in the (Mg, Ni)2SiO4 mixture at various T (P ⫽ 1 bar). Experimentally
observed distributions are compared with the results of static interionic potential
calculations carried out at two different values of the hardness factor (cf. sec-
tion 1.11.2).
Lastly, as far as volume is concerned, the mixture (Mg,Ni)2SiO4 exhibits
conspicuous deviations from ideality (see section 5.2.2): at high pressure, these
deviations presumably induce unmixing phenomena, not observed at P ⫽
1 bar.
For this purpose, it is of interest to recall that electrochemical galvanic cell measure-
ments indicate additional complexities of the system: Ottonello and Morlotti (1987) ob-
Silicates / 245
T ⫽ 1000 °C,
T ⫽ 1400 °C, P ⫽ 1 bar P ⫽ 1 bar
served two isoactivity regions in the binary join, one at approximately XNi ⫽ 0.2 and the
other between 0.4 and 0.6. Ottonello and Morlotti attributed this complex thermodynamic
feature to the formation of the intermediate compound (Ni0.25Mg0.75 )2SiO4 , with limited
solubility toward the end-member components Mg2SiO4 and Ni2SiO4 . Actually, as shown
by Campbell and Roeder (1968), in the presence of excess silica, an olivine mixture of
this limiting composition reacts with SiO2 to form pyroxene. Because the galvanic cell
compartment contains (Mg,Ni)2SiO4 ⫹ SiO2 ⫹ Ni, the two isoactivity regions indicate
the coexistence of a (reacted) pyroxene-type (Mg,Ni)SiO3 phase with a (Mg,Ni)2SiO4
mixture. The activity data derived from electromotive force (EMF) measurements cannot
thus be considered representative of the (Mg,Ni)2SiO4 mixture, and the most reliable ac-
tivity data are those of Campbell and Roeder (1968), which can be reproduced by appro-
priate parameterization of the repulsive potentials in the mixture, as shown in table 5.15.
(Mg,Mn)2SiO4 Mixture
According to Glasser (1960), the (Mg,Mn)2SiO4 mixture is almost ideal in the
temperature range T ⫽ 1300 to 1600 °C (P ⫽ 1 bar). The experiments of Maresch
et al. (1978) at T ⫽ 700 to 1100 °C (P ⫽ 2 kbar) indicate the presence of unmixed
phases in the composition range Fo40 to Fo75 . According to lattice energy calcula-
tions, there are two solvi at P ⫽ 1 bar, extending from intermediate composition
toward the two end-members (figure 6 in Ottonello, 1987). As regards intracrys-
talline disorder, Mn2⫹ appears preferentially stabilized in M2 at all investigated
T conditions (Francis and Ribbe, 1980; Urusov et al., 1984).
(Fe,Ca)2SiO4 Mixture
This mixture shows several analogies with the Mg2SiO4-Ca2SiO4 system. Also
in this case, an intermediate compound (kirschsteinite, FeCaSiO4 ) forms, with
246 / Geochemistry of Crystal Phases
(Fe,Mn)2SiO4 Mixture
According to Schwerdtfeger and Muan (1966), this mixture is essentially ideal at
T ⫽ 1150 °C, P ⫽ 1 bar, although the intracrystalline distribution measurements
of Annersten et al. (1984) and Brown (1970) show that Fe2⫹ is preferentially dis-
tributed on the M1 site.
(Co,Fe)2SiO4 Mixture
According to Masse et al. (1966), this mixture shows slight positive deviations
from ideality of activity values at T ⫽ 1180 °C, P ⫽ 1 bar. There are no experi-
mental data on intracrystalline distribution.
(Mg,Fe,Ca)2SiO4 Mixture
Only the quadrilateral portion of the Mg2SiO4-Ca2SiO4-Fe2SiO4 system with
XCa2SiO4 ⱕ 0.5 is of relevant interest in geochemistry. Mixing behavior in the oliv-
ine quadrilateral was investigated by Davidson and Mukhopadhyay (1984). As
figure 5.13 shows, the solvus between Ca-rich and Ca-poor compositional terms
extends within the quadrilateral, progressively shrinking toward the Fe-Ca join.
Multicomponent mixing behavior has been parameterized by Davidson
and Mukhopadhyay (1984) with a nonconvergent site-disorder model. The
interacting components were first expanded in a set of six site-ordered end-
members: FeM1MgM2SiO4 , MgM1FeM2SiO4 , FeM1FeM2SiO4 , MgM1MgM2SiO4 ,
FeM1CaM2SiO4 , and MgM1CaM2SiO4 . It was then assumed for simplicity that
mixing of Fe2⫹-Mg2⫹ is ideal in both M1 and M2 sites and that mixing of Fe2⫹-
Ca2⫹ and Mg2⫹-Ca2⫹ is nonideal in the M2 site. The authors then followed the
guidelines of Thompson (1969, 1970) and expanded the Gibbs free energy of mix-
ing of the mixture into two distinct contributions arising from configurational
entropy and enthalpic interaction terms:
+ µ MgMg
0
XMg,M 1 XMg,M 2 + µ FeFe
0
XFe,M1 XFe,M 2 (5.44)
+ µ MgFe
0
XMg,M 1 XFe,M 2 + µ FeMg
0
XFe,M 1 XMg,M 2
Silicates / 247
The mixing properties were then solved by minimizing the Gibbs free energy of
the mixture with respect to ordering parameter t:
∗
0
( )
+ ∆G 0.5X C a,M 2 + t + F − ∆ 0
G E0 (1 − 0.5X Ca,M 2 )
and by applying the concept of equality of chemical potential for components in
coexisting (unmixed) phases at equilibrium (tie lines in figure 5.13; cf. also sec-
tions 2.3 and 3.11).
Five excess terms appear in equation 5.48: two of them are the usual regular
248 / Geochemistry of Crystal Phases
interaction parameters (i.e., WMgCa ⫽ 32.9 kJ, WFeCa ⫽ 21.4 kJ); the other three
are related to standard-state potentials as follows:
∆G E0 = µ Fe
0
− µ Mg
0
= − 0.84 kJ ( 5.49 )
M 1Mg M 2 M 1FeM 2
∆G 0 = µ Mg
0
− µ Fe
0
= 7.0 ± 3.9 kJ ( 5.50 )
∗ M 1FeM 2 M 1Mg M 2
(
F 0 = 2 µ Mg
0
M1Ca M 2
− µ Fe
0
M1 Ca M 2
)+µ 0
Fe M1Fe M 2
(5.51)
− µ Mg
0
= 12.7 ± 1.6 kJ .
M1 Mg M 2
However, Davidson and Mukhopadhyay (1984) did not use the macroscopic fractions Xi
but thermodynamic fractions arising from pair probabilities and consistent with the six–
end-member identification (see section 3.8.7). Moreover, nonrandom configurational con-
tributions were accounted for in the entropy term (which is not ideal, because Ca is absent
in the M1 site). In Thompson’s (1969, 1970) notation, entropic effects are treated in the
same manner, and the standard state Gibbs free energies of the pure end-members and all
nonconfigurational effects are grouped in a vibrational Gibbs energy term (G*):
The word “vibrational,” is in this case somewhat misleading, because enthalpic interac-
tions are essentially static in nature and are purely static whenever the zeroth approxima-
tion is appropriate (see Della Giusta and Ottonello, 1993). Lastly, it must be noted that
the excess term F 0 represents a reciprocal energy arising from the noncoplanarity of the
four quadrilateral end-members (cf. section 3.8.7). The model of Davidson and Mukho-
padhyay (1984) was reconsidered by Hirschmann (1991), who proposed an analogous
treatment for (Ni,Mg,Fe)2SiO4 mixtures.
5.3 Garnets
The structure of garnet was determined for the first time by Menzer (1926). The
general formula unit for garnet may be expressed as X3Y2Z3O12 . Garnet has a
body-centered-cubic cell (spatial group Ia3d ) and 8 formula units per unit cell.
Cations X, Y, and Z occupy nonequivalent sites with, respectively, dodecahedral,
octahedral, and tetrahedral symmetry. The main IV-fold coordinated cation is
Si4⫹. Figure 5.14 shows that in garnet, too, the [TO4 ] groups are isolated from
Silicates / 249
Figure 5.14 Structure of garnet. Positional angle defines degree of rotation on 4 axis.
one other (nesosilicates). The octahedra are also isolated, and are linked by single
apexes to tetrahedra. Dodecahedra share edges either with tetrahedra and octahe-
dra or with other dodecahedra. With respect to other silicates, the structure of
garnet is very dense as regards oxygen packing. Assuming a spherical oxygen ion
with an ionic radius of 1.38 Å, in noncalcic garnets oxygen occupies 70% of the
volume of the unit cell—a value near the theoretical value of maximum packing
(eutaxis; 74%; cf. Meagher, 1982).
Ca X + Si Z ⇔ Na X + PZ . ( 5.54 )
250 / Geochemistry of Crystal Phases
Table 5.16 Elements present at major, minor, and trace levels in the garnet phase.
Table 5.17 lists analyses of several types of natural garnet, after Deer et al.
(1983). The calculation of ionic fractions on the basis of 24 oxygens shows effec-
tive diadochy of Al on Si in tetrahedral site Z, whereas Ti4⫹ and Fe3⫹ appear to
be stabilized in octahedral site Y. Deviation from ideal closure in sites Y and X
(6 and 4, respectively) may be attributed in some instances to erroneous estima-
tion of ferric iron (samples 4, 5, and 6), but in others it is difficult to interpret
(samples 1, 2, and 3).
As suggested by Winchell (1933), the main garnet components may be sub-
divided into two groups with complete miscibility of their members: the
calcic group of ugrandites, comprising uvarovite (Ca3Cr2Si3O12 ), grossular
(Ca3Al2Si3O12 ), and andradite (Ca3Fe2Si3O12 ); and the group of pyralspites,
including pyrope (Mg3Al2Si3O12 ), almandine (Fe3Al2Si3O12 ), and spessartine
(Mn3Al2Si3O12 ). Six other less common compositional terms (which, however,
become more and more important in phase stability studies at high P) may be
associated with these six end-members. The names and formula units of these six
terms are listed in table 5.18 (note that all the natural garnets listed in table 5.17
have compositions that may be reexpressed as molar proportions of the various
compositional terms in table 5.18).
Table 5.19 summarizes the main occurrences of the various compositional
terms. Because garnets are solid mixtures, the listed components merely indicate
the main compositional terms of the phase.
Using multiple regression techniques, Novak and Gibbs (1971) obtained linear
equations that give cell edge and oxygen positional parameters as functions of
the radii of cations occupying sites X, Y, and Z. The equation of Novak and
Gibbs for cell edge is
Equation 5.55 was later modified by Novak and Colville (1975) to include the
dimension of site Z:
Silicates / 251
Table 5.17 Compositions (in weight %) of some natural garnets (from Deer et al., 1983): (1)
almandine from quartz-biotite gneiss; (2) andradite from metamorphosed andesite; (3)
grossular from garnet-bearing gneiss; (4) pyrope from eclogite; (5) spessartine from a
cornubianitic rock; (6) uvarovite from a garnet-tremolite-pyrrothite vein.
Samples
Constituent
Oxide (1) (2) (3) (4) (5) (6)
冧 冧
H2O (⫹) - 0.48 0.13 -
0.16 0.18
(⫺)
H 2O - 0.16 0.06 -
Table 5.18 Chemistry and terminology of main components of the garnet phase.
Mineral Occurrence
Almandine Eclogites (alm-pir), gneisses, metamorphic schists, granites, pegmatites,
Fe3Al2Si3O12 granulites (mineral index)
Grossular Typical of contact-metamorphosed limestones (skarns, marbles, cornubianites)
Ca3Al2Si3O12
Andradite In contact-metamorphic zones, in association with magnetite and Fe-rich
Ca3Fe2Si3O12 silicates (feldspathoid-bearing igneous rocks)
Uvarovite Mainly in chromite deposits
Ca3Cr2Si3O12
Spessartine Pegmatites and other igneous acidic rocks, metamorphic rocks and Mn-rich
Mn3Al2Si3O12 clay deposits
Pyrope Basic and ultrabasic rocks (peridotites, serpentinites), kimberlitic xenolites,
Mg3Al2Si3O12 rodingites, eclogites
Both equations are based on the ionic radii of Shannon and Prewitt (1969). The
method of Novak and Gibbs (1971) was then reexamined by Hawthorne (1981b),
and finally extended by Basso (1985) to hydrogarnets. Basso’s (1985) equations
are shown in table 5.20. They determine cell edge, fractional atomic coordinates
of oxygen, and mean metal-to-oxygen bond distances with satisfactory precision.
For even better precision, it is good practice to submit the obtained parameters
to a further optimization procedure, based on crystallographic constraints (i.e.,
Distance-Least-Squares, or DLS, treatment).
Table 5.21 shows structural data on various garnet end-members, compared
Silicates / 253
⫺20冤冢X1 ⫺ O冣 ⫺ 冢Z ⫺ O冣 冥
2 2 2
⫺5冤冢X1 ⫺ O冣 ⫺ 2冢Y ⫺ O冣 ⫺ 冢Z ⫺ O冣 冥 冧 冧
2 2 2 2 1/2 1/2
with the values obtained by the method of Basso (1985), followed by the DLS
procedure of Baerlocher et al. (1977). The optimization procedure determinates
cell parameters for the six components not pertaining to the ugrandite and pyral-
spite series.
On the basis of equations of type 5.55 and of analogous expressions for posi-
tional parameters, Novak and Gibbs (1971) proposed several guidelines to estab-
lish whether a “garnet” compound may form and remain stable, according to the
dimensions of cations occupying sites X, Y, and Z. These guidelines consider the
following three factors.
On the basis of these principles, Novak and Gibbs (1971) proposed a map of
compatible rX and rY dimensions for the formation of a garnet phase. This map
(figure 5.15) shows that the commonly observed components plot in the low-
volume zone of the compatibility area. Theoretically, compounds with higher vol-
umes may form and remain stable at appropriate P and T conditions.
254 / Geochemistry of Crystal Phases
Figure 5.15 Dimensional compatibility field for stable garnets according to Novak and
Gibbs (1971). Points correspond to 12 main components listed in table 5.21. Molar vol-
umes are also indicated.
[MgO8 ] dodecahedron is more compressible than the [AlO6 ] octahedron, and that
the [SiO4 ] group, unlike that of olivine, is also compressible. Thermal expansion
and compressibility are thus not perfectly antipathetic in this case. Notwithstand-
ing experimental studies on the compressibilities of the various compositional
terms, complete agreement among authors does not exist (Isaak and Graham,
1976; Sato et al., 1978; Leitner et al., 1980; Babuska et al., 1978; Hazen and
Finger, 1978; Bass, 1986). Calculation of garnet compressibility and thermal
expansion based on structure simulations at various pressures and temperatures
gives bulk moduli that are within the range of experimental values and may be
extended to not previously investigated compounds. Values (listed in table 5.22)
are relative to the Birch-Murnaghan equation of state. The same table also com-
pares calculated thermal expansion values with the experimental values of Skin-
ner (1956), Suzuki and Anderson (1983), Isaak et al. (1992), and Armbruster and
Geiger (1993) reconducted to the usual polynomial expansion in T:
1 ∂V
α V ,T = = α 0 + α 1T + α 2 T .
2
( 5.57 )
V ∂T P
256 / Geochemistry of Crystal Phases
Table 5.22 Bulk modulus and thermal expansion for various compositional terms of the
garnet phase. The experimental range for bulk moduli is relative to a common value of
K⬘ ⫽ 4. The calculated values are from Ottonello et al. (1996). References for thermal
expansion are as follows: (1) Ottonello et al. (1996); (2) Skinner (1956); (3) Suzuki and
Anderson (1983); (4) Isaak et al. (1992); (5) Armbruster and Geiger (1993).
Experimental
Range
Calculated Values K0 (K⬘ ⫽ 4)
Compound Formula Unit K0 (Mbar) K⬘
Pyrope Mg3Al2Si3O12 1.776 4.783 1.66–1.79
Almandine Fe3Al2Si3O12 1.837 4.795 1.75–1.80
Spessartine Mn3Al2Si3O12 1.797 4.833 1.72–1.75
Grossular Ca3Al2Si3O12 1.673 4.887 1.39–1.69
Uvarovite Ca3Cr2Si3O12 1.743 4.977 1.43–1.65
Andradite Ca3Fe2Si3O12 1.726 4.919 1.38–1.59
Knorringite Mg3Cr2Si3O12 1.876 4.842 -
Calderite Mn3Fe2Si3O12 1.818 4.823 -
Skiagite Fe3Fe2Si3O12 1.867 4.762 -
Khoharite Mg3Fe2Si3O12 1.857 4.799 -
Mn-Cr garnet Mn3Cr2Si3O12 1.852 4.894 -
Fe-Cr garnet Fe3Cr2Si3O12 1.902 4.831 -
Thermal Expansion
Leitner et al. (1980) showed that the bulk moduli of terms of pyrope-almandine-
grossular and andradite-grossular mixtures vary with chemical composition in a
perfectly linear fashion (within the range of experimental incertitude). This be-
havior is also implicit in the generalization of Anderson (1972).
Table 5.23 Thermochemical data for various endmembers of the garnet series.
ECFS ⫽ crystal field stabilization energy. The other column heads are defined as in table 5.12.
Data in kJ/mole.
0
Compound EBHF Hf,Tr ,Pr
UL EC ER EDD EDQ ECFS
Pyrope 59,269.2 ⫺6290.8(1) 65,560.0 ⫺77,060.6 12,287.8 ⫺599.8 ⫺187.4 0.0
Almandine 60,487.6 ⫺5261.3(2) 65,611.0 ⫺76,916.7 12,328.7 ⫺769.6 ⫺253.4 ⫺138.0
Spessartine 59,822.8 ⫺5686.9(3) 65,509.8 ⫺76,704.9 12,097.9 ⫺684.2 ⫺218.6 0.0
Grossular 57,997.4 ⫺6632.8(4) 64,630.2 ⫺76,148.6 12,360.9 ⫺645.3 ⫺197.2 0.0
Uvarovite 58,356.4 ⫺6034.6(5) 63,919.0 ⫺75,446.1 12,396.7 ⫺660.6 ⫺209.1 ⫺472.0
Andradite 58,489.9 ⫺5758.9(6) 64,248.8 ⫺75,160.5 11,779.0 ⫺660.5 ⫺206.8 0.0
Knorringite 59,628.2 ⫺5576.1(7) 64,732.4 ⫺76,294.3 12,390.4 ⫺623.9 ⫺204.5 ⫺472.0
Calderite 60,315.3 ⫺4719.8(7) 65,035.1 ⫺75,661.1 11,556.8 ⫺701.7 ⫺229.1 0.0
Skiagite 60,980.1 ⫺4276.2(7) 65,118.4 ⫺75,856.7 11,783.8 ⫺783.5 ⫺262.0 ⫺138.0
Khoharite 59,761.7 ⫺5299.5(7) 65,061.2 ⫺75,981.6 11,750.0 ⫺626.5 ⫺203.1 0.0
Mn-Cr garnet 60,181.8 ⫺5007.1(7) 64,716.9 ⫺75,965.8 12,182.0 ⫺701.5 ⫺231.7 ⫺472.0
Fe-Cr garnet 60,846.7 ⫺4555.3(7) 64,792.0 ⫺76,162.5 12,420.4 ⫺784.6 ⫺265.3 ⫺610.0
References: (1) Charlu et al. (1975); (2)Anovitz et al. (1993); (3)Boeglin (1981); (4)Charlu et al. (1978); (5)
Kiseleva et
al. (1977); (6)Helgeson et al. (1978); (7)Ottonello et al. (1996)
258 / Geochemistry of Crystal Phases
Table 5.24 Entropy and heat capacity function values for the various garnet end-members compared
with calorimetric values.
(1) Ottonello et al. (1996); (2) Haselton and Westrum (1980); (3) Newton et al. (1977); (4) Tequı̀ et al. (1991);
(5) Anovitz et al. (1993); (6) Robie et al. (1987)
In the Kieffer model (Kieffer 1979a, b, and c; cf. section 3.3), the isochoric heat capac-
ity is given by
3R
3
1
CV =
Z
∑ S( X i ) + 3nR1 −
N
− q K ( X l ,X u ) + 3nRq E( XE ) ,
( 5.58)
i =1
where S(Xi ) is the dispersed acoustic function, K (Xl,Xu ) represents the optic continuum, and
E(XE ) is the Einstein function (see eq. 3.69, 3.71, and 3.73 in section 3.3). The expression
for third-law harmonic entropy is:
[arcsin(X X )] X dX
3 2
Xi
3R 2
3
∑∫
i
S =
Z π ( ) ( )
12
i =1 X −X 0 e −1 i
2 2 X
[arcsin(X X )] ln(1 − e ) dx
3 2
Xi
3R 2
3
∑∫
i
− −X
Z π ( )
12
i =1 X −X 0 i
2 2
(5.59)
3 3
3R1 − − q 3R1 − − q
Xu Xu
∫ ln(1 − e ) dX
3s X dX 3s
+
Xu − Xl
∫( eX − 1
−
) Xu − Xl
− XE
Xl Xl
+
(
3RqX E
e XE − 1 )
− 3Rq ln 1 − e − X E , ( )
Silicates / 259
Table 5.25 Parameters of the Kieffer model for calculation of CV and harmonic entropy for the 12
garnet end-members in the system (Mg, Fe, Ca, Mn)3(Al, Fe, Cr)2Si3O12. *l,Kmax is the lower cutoff
frequency of the optic continuum optimized on the basis of the experimental value of entropy at
T ⫽ 298.15 °C. D is the Debye temperature of the substance. Sm and San are, respectively, the
magnetic spin and anharmonicity contributions to third-law entropy expressed in J/(mole ⫻ K).
See section 3.3 for the significance of the various frequencies (cm⫺1) (from Ottonello et al., 1996).
where
បω
X = , ( 5.60 )
kT
−1 2
m
ω l , K max = ω l , K = 0 1 + 2 , ( 5.61)
m1
where m1 and m2 are reduced masses of intervening ions (oxygen and “heavy” cation,
respectively, in our case; see Kieffer, 1980). l,Kmax values for pyrope, almandine, spessar-
tine, grossular, and andradite are those of Kieffer (1980). For uvarovite, a value of 170
cm⫺1 for l,K⫽0 was estimated from the spectrum of Moore et al. (1971), and the corre-
sponding l,Kmax was obtained by application of equation 5.61.
The model parameters relative to the uncommon garnet end-members were obtained
by Bokreta (1992) through linear regression involving the atomistic properties of the in-
260 / Geochemistry of Crystal Phases
tervening ions. Equation 5.59 and the various vibrational function terms in eq. 5.58 were
calculated by gaussian integration. Values on nondimensionalized frequencies l,Kmax were
adjusted, stemming from the initial guess values with a trial-and-error procedure, match-
ing the calculated S with experimental third-law entropy values at T ⫽ 298.15 K and
P ⫽ 1 bar, after correction for a magnetic spin entropy contribution of the form
Sm = R ln( 2 q + 1) , ( 5.62 )
where q is 32 for Cr3⫹, 52 for Mn3⫹ and Fe2⫹, 2 for Fe3⫹, and correction for anharmonicity,
according to
T
San = ∫ Vα 2 K 0 dT . ( 5.63 )
0
The procedure of Ottonello et al. (1996a) is operationally different from that indicated by
Kieffer (1980), which suggested that the calculated CV be fitted to the calorimetric CP at
100 K after subtraction of the anharmonicity contribution
C P = CV + TVα 2 K 0 . ( 5.64 )
Figure 5.16A shows, for example, how calculated CP values for pyrope compare with
low-T calorimetric measurements of Haselton and Westrum (1980). Adopting the initial
guess value of l,Kmax (115.7 cm⫺1; see table 5.25) leads to an entropy estimate at Tr ⫽
298.15 of 272.15 J/(mol ⫻ K) (268.834 harmonic plus 3.316 anharmonic). The calorimet-
ric value being 266.27, the initial guess value of l,Kmax is too low and must be raised to
124.0 cm⫺1 to match the calorimetric evidence. As we see in the upper part of figure 5.16
also the resulting CP in the low-T region compares better with calorimetric values after
adjustment of the nondimensionalized frequency at the limit of the Brillouin zone. In
the lower part of figure 5.16B, model predictions for pyrope are compared with high-T
calorimetric data of Newton et al. (1977) and Tequı̀ et al. (1991), indicating a reasonable
agreement with the latter.
Tc =
W
=
3000
2 R 2 × 1.98726
( )
= 756 K = 483 ˚C . ( ) ( 5.65)
Silicates / 261
Figure 5.16 Heat capacity of pyrope at constant P, as determined from the Kieffer vibra-
tional model, compared with low-T (upper part of figure) and high-T (lower part of figure)
experimental evidences. From Ottonello et al. (1996). Reprinted with permission of The
Mineralogical Society of America.
Figure 5.17 Enthalpic interactions in the various binary joins of aluminiferous garnets.
䡺: calorimetric data; 䡵: results of interionic potential calculations. The corresponding
subregular Margules interaction parameters are listed in table 5.26 (from Ottonello et al.,
in prep.).
in tetrahedral coordination (cf. table 5.17). This fact should be carefully evaluated
in studies of the mixing properties of aluminiferous garnets. Unfortunately,
agreement among the various experimental sources is so poor that it prevents
appreciation of the effect of Al-Si substitutions in natural mixtures (cf. tables
5.26 and 5.27). Hence, the existence of excess entropy terms in mixtures must be
provisionally ascribed either to substantial short-range ordering on sites or to
vibrational effects.
Particularly, for the join (Ca,Fe)3Al2Si3O12 , the recent data of Koziol (1990)
confirm that excess entropy of mixing is virtually absent (cf. table 5.27). Existing
experimental data on the Mg3Al2Si3O12-Ca3Al2Si3O12 system show that the excess
entropy of mixing terms are probably symmetric and coupled to a slightly positive
excess volume. Newton and Wood (1980) suggested that the molar volume of the
Silicates / 263
Table 5.26 Subregular Margules parameters for enthalpic interactions in aluminiferous garnets
(from Ottonello et al., in prep.). Values in kJ/mole on a three-cation basis. The subregular Margules
model is Hmixing, ij ⫽ XiXj (Wij XiXj ⫹ WjiXjXi) with (1) Mg3Al2Si3O12, (2) Mn3Al2Si3O12,
(3) Ca3Al2Si3O12, and (4) Fe3Al2Si3O12.
WH12 WH21 WH13 WH31 WH14 WH41 WH23 WH32 WH24 WH42 WH34 WH43 Reference
- - 93.6 93.6 - - - - - - - - (1)
- - 25.1 47.9 - - - - - - - - (2)
- - 48.1 48.1 - - 0 0 0 0 - - (3)
4.5 4.5 - - - - - - - - - - (4)
- - - - 6.4 2.1 - - - - - - (5)
- - 12.6 50.8 31.4 2.5 - - - - ⫺7.9 58.0 (6)
- - 12.6 50.8 - - - - - - - - (7)
- - 15.0 56.5 - - - - - - - - (8)
- - - - 36.2 ⫺15.8 - - - - ⫺9.1 13.7 (9)
- - - - - - - - 0 0 - - (10)
12.6 12.6 - - - - - - - - - - (11)
- - - - - - 0 0 - - ⫺3.3 7.8 (12)
30.3 30.3 25.9 59.3 6.4 2.1 1.4 1.4 1.9 1.9 2.6 20.3 (13)
0 0 21.6 69.2 3.7 0.2 0 0 0 0 2.6 20.3 (14)
46.3 46.3 25.9 25.9 6.4 2.1 0 0 2.3 2.3 2.6 20.3 (15)
17.7 15.1 38.6 42.4 18.3 14.1 ⫺0.4 14.6 ⫺1.2 ⫺3.2 12.6 2.0 (16)
(1) Hensen et al., 1975; phase equilibration; (2) Newton et al., 1977; calorimetric; (3) Ganguly and Kennedy, 1974;
phase equilibration; (4) Wood et al., 1994; phase equilibriation; (5) Hackler and Wood, 1989; phase equilibration;
(6) Ganguly and Saxena, 1984; multivariate analysis; (7) Haselton and Newton, 1980; calorimetric; (8) Wood,
1988; phase equilibration; (9) Geiger et al., 1987; calorimetric; (10) Powenceby et al., 1987; phase equilibration;
(11) Chakraborty and Ganguly, 1992; diffusion experiments; (12) Koziol, 1990; phase equilibration; (13) Ganguly
and Cheng, 1994; multivariate analysis; (14) Berman, 1990; multivariate analysis; (15) Ganguly, 1995; personal
communication; (16) Ottonello et al. (in prep.); structure-energy calculations.
Y Cation
Figure 5.18 Enthalpy of mixing surface for aluminiferous garnets in the system
(Mg,Mn,Ca,Fe)3Al2Si3O12 based on the Wohl’s model parameters in table 5.28 (from Otto-
nello et al., in prep.).
Mattioli and Bishop (1984), the mixture is of regular type but has a positive bulk
interaction parameter W between 0.6 and 3 kcal/mole. Extension to multicompo-
nent mixtures was attempted by Berman (1990), Ganguly (personal communica-
tion), and Ottonello et al. (in prep.). Berman (1990) applied the model of Berman
and Brown (1984), originally conceived for silicate melts, to aluminiferous garnets
and deduced the magnitude of ternary interaction parameters by applying
Wi , j , k =
(W
i ,i , j + Wi , j , j + Wi ,i , k + Wi , k, k + W j , j , k + W j , k, k ) −C (5.66)
i, j,k
2
and neglecting ternary term Ci, j,k (note that eq. 5.66 corresponds to the part on
the right in square brackets in Wohl’s equation 3.183). Ganguly (personal com-
munication) applied the formulation of Cheng and Ganguly (1994) and found
the ternary interaction terms to be insignificant. Wohl’s model parameters for
266 / Geochemistry of Crystal Phases
5.4 Pyroxenes
Figure 5.19 (A) Schematic representation of [SiO3]n chains of pyroxenes: (a) projection
on plane (100): (b) projection along axis Z; (c) projection along axis Y; (d) perspective
representation. (B) Relative arrangement of tetrahedral chains in pyroxene structure (seen
in their terminal parts with intercalation of M1 and M2 positions. From Putnis and
McConnell (1980). Reproduced with modifications by permission of Blackwell Scientific
Publications, Oxford, Great Britain.
R2⫹ in table 5.29 represents divalent cations (Mn2⫹, Fe2⫹, Mg2⫹ ), and R3⫹ repre-
sents trivalent cations (Fe3⫹, Cr3⫹, Al 3⫹ ). Thus, from the crystal chemical point
of view, the terms “augite,” “omphacite,” and “augitic aegirine” do not identify
precise stoichiometry. Analogously, the general term “orthopyroxene,” according
to Deer et al. (1978), represents a mixture of magnesian and ferroan components,
and the term “pigeonite” identifies an Mg-Fe-Ca mixture.
In geochemistry it is necessary to describe the composition of pyroxene by
end-members that are compositionally simple and stoichiometrically well defined.
It is also opportune to distinguish the various structural classes, because P-T
stability and reactivity vary greatly with type of polymorph. This inevitably re-
quires the formulation of “fictitious” components (i.e., components that have
never been synthesized as pure phases in the structural form of interest, but that
are present as members of pyroxene mixtures). For example, table 5.30 gives the
list of pyroxene geochemical components proposed by Ganguly and Saxena
(1987) (partly modified here).
In compiling table 5.30, the “orthopyrope” component (Mg1.5Al1.0Si1.5O6 ) of
Ganguly and Saxena was discarded, because it may be obtained as a mixture of
Mg-Tschermak and enstatite. Moreover, Cr- and Li-bearing terms (ureyite and
spodumene) are added to the C2/c class and a Co-bearing term to the Pbca class.
The manganous term (Mn2Si2O6 ; Tokonami et al., 1979) has been added to the
P21 /c class. It must be noted that natural pigeonites are mixtures of compositional
Silicates / 269
Table 5.31 Occurrence of main compositional terms of pyroxenes in various types of rocks.
0.030.
Based on these four rules, Cameron and Papike (1982) selected 175 analyses out
of 405 reported by Deer et al. (1978) and discarded 230 compositions. This selec-
tion is extremely rigorous and does not take into account either the possible stabi-
lization of Fe3⫹ in tetrahedral sites or the existence of extrinsic disorder (cf. chap-
ter 4). Robinson (1982) showed that, by accepting a limited amount of cationic
vacancies in M2 sites and assuming possible stabilization of Fe3⫹ in tetrahedral
positions, 117 additional analyses out of the 230 discarded by Cameron and Pa-
pike (1982) may be selected.
Samples 6 and 7 in table 5.32 are from the Zabargad peridotite (Red Sea)
and are representative of the chemistry of upper mantle pyroxenes (Bonatti et al.,
1986). The absence of Fe2O3 in these samples is due to the fact that microprobe
analyses do not discriminate the oxidation state of iron, which is thus always
expressed as FeO. It must be noted here that the observed stoichiometry (based
on four oxygen ions) is quite consistent with the theoretical formula and that no
Fe3⫹ is required to balance the negative charges of oxygen.
Silicates / 271
Table 5.32 Compositions (in weight %) of natural pyroxenes (samples 1–5 from Deer et al.,
1983; samples 6 and 7 from Bonatti et al., 1986): (1) enstatite from a pyroxenite; (2) ferrosilite
from a thermometamorphic iron band; (3) hedembergite; (4) chromian augite from a
gabbroic rock of the Bushveld complex; (5) aegirine from a riebeckite-albite granitoid;
(6) diopside from a mantle peridotite; (7) enstatite from a mantle peridotite.
Samples
Table 5.33 Structure of pure pyroxene components in stable state. Molar volume in cm3/
mole; molar weight in g/mole; cell edges in Å; cell volume in Å3. Adapted from Smyth and
Bish (1988).
continued
274 / Geochemistry of Crystal Phases
The various lattice site dimensions in C2/c clinopyroxenes depend in a complex fash-
ion on the bulk chemistry of the phase. Nevertheless, the various interionic distances (and
cell edges and volumes) may be expressed as functions of cationic occupancies of lattice
Silicates / 275
Figure 5.21 Mean cation-to-oxygen distances in sites M1 (A) and M2 (B) plotted against
ionic radii of occupying cations (algebraic mean). Adopted radii are those of Shannon
(1976). Data from Cameron and Papike (1981).
Table 5.34 Multiple regression parameters relating site population to interionic distances in
C2/c pyroxenes.
where
The ionic radii in equations 5.69 to 5.71 (rAlIV, rAlVI, rSi, . . .) are those of Shannon (1976)
in the appropriate coordination state with oxygen and XAl,T , XSi,T , . . . are the atomic
occupancies in the various sites.
Table 5.35 Molar volume (cm3/mole), isobaric thermal expansion (K-1), and isothermal bulk
modulus (Mbar) of pyroxene end-members according to Saxena (1989)
C2/c at about 960 °C in pigeonite, and Smyth (1974) observed a first-order transi-
tion in the low-T clinopyroxene at T ⫽ 725 °C.
Pressure has an antithetical effect on the interionic distances of pyroxenes
with respect to temperature, and the observed compressibilities of M2, M1, and T
sites obey the generalizations of Anderson (1972) and Hazen and Prewitt (1977).
The properties above indicate that the responses of the various pyroxene poly-
morphs to modifications of P and T intensive variables are different. It is essen-
tially this fact that determines the opening of miscibility gaps during the cooling
stage from “magmatic” P-T regimes. To rationalize this phenomenon in thermo-
dynamic terms, we must distinguish the structural properties of the various com-
positional terms crystallizing in the various spatial groups with appropriate vol-
ume factors. Table 5.35 lists the values of molar volume, isobaric thermal
expansion, and isothermal bulk modulus for pyroxene polymorphs in the thermo-
dynamic database of Saxena (1989). Note that, for several components, Saxena
(1989) adopts identical thermal expansion parameters for the structural classes
C2/c and Pbca. This assumption renders the calculations straightforward in phase
transition studies but is not compulsory.
Let us examine in detail the phase stability relationships of the four end-members of
the pyroxene quadrilateral. Figure 5.24A shows the P and T stability ranges of the various
polymorphs of the Mg2Si2O6 compound, according to Lindsley (1982). The stable form
Silicates / 279
Figure 5.24 Phase stability relations for end-members of pyroxene quadrilateral. Melting
curves refer to anhydrous conditions. Solidus curves for CaMgSi2O6 in saturated vapor
phase conditions are also shown for various CO2/H2O ratios in the vapor phase. Dashed
lines are extrapolated. From Lindsley (1982). Reprinted with permission of The Mineral-
ogical Society of America.
P ⫽ 15 kbar. At higher temperatures, the Pbca form is thus stable at all T conditions, up
to congruent melting of the compound. At P ⫽ 1 bar, protoenstatite is stable up to T ⫽
1557 °C, at which incongruent melting occurs, with formation of forsterite plus liquid.
Incongruent melting has been observed only at low pressure, but the high-P limit of the
process is not known precisely.
At low P, the Fe2Si2O6 component is extrinsically unstable with respect to ␣-quartz
plus fayalite paragenesis:
This reaction proceeds spontaneously toward the right, up to pressures that are not known
precisely but that are estimated to be around 5 to 8 kbar for T in the range 200 to 400 °C.
At higher P and low T, a monoclinic phase is stable (clinoferrosilite; probably in P21 /c
form; CFs in figure 5.24B), and at high P and T the orthorhombic form Pbca is stable
(OFs in the same figure). According to Lindsley (1982), below the melting point the ortho-
rhombic phase has another transition to a pyroxenoid phase (ferrosilite III; Fs-III in figure
5.24B) or to the monoclinic C2/c polymorph.
At high P, CaMgSi2O6 undergoes congruent melting (figure 5.24C). The melting tem-
perature at P ⫽ 1 bar is about 1390 °C and (not shown in figure) the melting process at
low P is incongruent with the formation of forsterite:
The T stability of CaMgSi2O6 is greatly affected by fluid partial pressure. In the presence
of H2O-rich fluids, the thermobaric gradient of the melting reaction (which is positive in
anhydrous conditions, as observed for most crystalline silicates; cf. chapter 6) is reversed,
according to the experiments of Rosenhauer and Eggler (1975) (figure 5.24C).
The stable form of the CaFeSi2O6 compound is monoclinic C2/c in a wide
P-T range (hedembergite; Hd in figure 5.24D). At P ⬍ 13 kbar, hedembergite has a transi-
tion to the triclinic form bustamite (Bu in figure 5.24D). At low P, bustamite melts incon-
gruently to produce silica plus melt. Nevertheless, the melting reaction is congruent at P
higher than 2 to 3 kbar.
no apparent discontinuities are observable in the Gibbs free energy curve of the
crystalline mixtures as a function of their chemical composition (see, for instance,
Buseck et al., 1982). Table 5.36 lists the thermodynamic data of end-members at
standard state of pure component according to Saxena (1989). These data are
internally consistent and, combined with the appropriate mixing properties (see
section 5.4.5), quantitatively describe the phase relations observed within the
quadrilateral at the various P and T conditions of interest (the database of Sax-
ena, 1989 is given preference here with respect to the more recent revision of
Saxena et al., 1993). For comparative purposes, the same table also lists entropy,
enthalpy of formation from the elements, and the heat capacity function adopted
by Berman (1988) and by Holland and Powell (1990).
Table 5.37 lists the various terms of the lattice energy of some clinopyroxene
components. Data refer to the C2/c spatial group. Components Mg2Si2O6 and
Fe2Si2O6 , which crystallize in P21 /c, must thus be considered “fictitious.”
K3 K5 K6 K7
Phase Component Reference H 0f,Tr ,Pr S 0Tr ,Pr K1 K2 ⫻ 104 K4
106 108 102 104
Table 5.37 Lattice energy terms for C2/c pyroxenes. Values in kJ/mole. EBHF ⫽ energy of Born-
Haber-Fayans thermochemical cycle; UL ⫽ lattice energy; EC ⫽ coulombic energy; ER ⫽ repulsive
energy; EDD ⫽ dipole-dipole interactions; EDQ ⫽ dipole quadrupole interactions; H 0f,Tr ,Pr ⫽
enthalpy of formation from the elements at 298.15 K, 1bar reference conditions
lite facies), pigeonite disappears and a unique miscibility gap extends from the
orthorhombic Pbca polymorph to the C2/c monoclinic form (figure 5.25C). The
gap covers a wide compositional range, from the iron-free join up to Fe/(Fe ⫹
Mg) ⬵ 0.80. For higher relative amounts of Fe, pyroxene is extrinsically unstable
with respect to the olivine plus liquid paragenesis (Huebner, 1982).
In order to assess appropriately the above phase stability relations, we must
rely on experimental evidence concerning the limiting binary join and, particu-
larly, attention must focus on the Mg2Si2O6-CaMgSi2O6 and Fe2Si2O6-CaFeSi2O6
joins. Figure 5.26A shows phase relations in the system Mg2Si2O6-CaMgSi2O6 at
P ⫽ 15 kbar, based on the experiments of Lindsley and Dixon (1976) (LD in
figure 5.26A) and Schweitzer (1977) (S). A wide miscibility gap is evident between
Silicates / 285
Figure 5.25 Phase stability relations for natural pyroxenes in the quadrilateral. (A) Mag-
matic pyroxenes (not reequilibrated at low T). (B) Three-phase region for pyroxenes of
stratified complexes and magmatic series. (C) Pyroxenes reequilibrated in subsolidus con-
ditions. 䡺 ⫽ Pbca orthopyroxene; ∇ ⫽ augite (C2/c); ⌬ ⫽ pigeonite (P21/c); ◊ ⫽ olivine.
From Huebner (1982). Reprinted with permission of The Mineralogical Society of
America.
286 / Geochemistry of Crystal Phases
Figure 5.26 (A) Phase stability relations along binary join Mg2Si2O6-CaMgSi2O6 at P ⫽
15 kbar (anhydrous conditions) based on experiments of Lindsley and Dixon (1976) (LD)
and Schweitzer (1977) (S). En ⫽ enstatite solid mixture (Pbca); Di ⫽ diopside solid mix-
ture (C2/c); Pig ⫽ pigeonite solid mixture (P21/c). From Lindsley (1982). Reprinted with
permission of The Mineralogical Society of America. (B) Schematic representation of
Gibbs free energy properties responsible for opening of a miscibility gap between ortho-
rhombic (opx) and monoclinic (cpx) forms (Saxena, 1973). Coherence is assumed between
P21/c and C2/c polymorphs, which are thus represented by a single Gibbs free energy
curve. Note distinction between “solvus field” (intrinsic instability) and “miscibility gap”
(extrinsic instability).
the orthorhombic (opx) and monoclinic (cpx) polymorphs. This gap is compli-
cated by the appearance at T ⫽ 1450 °C of pigeonite. At T higher than 1450 °C,
two distinct miscibility gaps occur (Pbca-P21 /c and Pbca-C2/c). With increasing
P, the gap widens and the stability field of pigeonite progressively disappears (P ⫽
25 kbar). As shown in part B of figure 5.26, the opx-cpx gap may be conceived
as the result of the combined intrinsic instabilities of the two polymorphs: i.e.,
the “miscibility gap” between cpx and opx covers the two “solvi” valid, respec-
tively, for the opx and cpx phases. A third phase (pigeonite) appears at the T-X
zone of intersection of the two solvus limbs on the Ca-poor side of the binary
join. Lindsley et al. (1981), on the basis of all existing experimental data on the
Mg2Si2O6-CaMgSi2O6 join, proposed a model valid for the pressure range 1 bar
⬍ P ⬍ 40 kbar. This model does not distinguish spatial groups P21 /c and C2/c
and treats the equilibria as stability relations between a generic orthorhombic
phase and a generic monoclinic phase. In this model, the Gibbs free energy curves
for the monoclinic mixture thus imply the existence of a certain amount of elastic
strain energy (to maintain coherence) the nature of which was described in sec-
tion 3.13.
To solve the energy model, Lindsley et al. (1981) based their calculations on the prin-
ciple of equality of chemical potentials of components in mixture for phases at equilibrium
(cf. section 2.3):
Silicates / 287
applied at every point in the P-T-X space of interest. Application of the usual relations
between chemical potential and thermodynamic activity to identities 5.75 and 5.76 gives
µ Mg 2 Si 2 O6 , cpx = µ Mg
0
2 Si 2 O6 , cpx
+ RT ln a Mg 2 Si 2 O6 , cpx ( 5.77 )
µ Mg 2 Si 2 O6 , opx = µ Mg
0
2 Si 2 O6 , opx
+ RT ln a Mg 2 Si 2 O6 , opx ( 5.78 )
a Mg 2 Si 2 O6 , opx
∆µ Mg
0
2 Si 2 O6
= RT ln (5.81)
a Mg 2 Si 2 O6 , cpx
aCaMgSi2 O6 , opx
∆µCaMgSi
0
2 O6
= RT ln . (5.82)
aCaMgSi2 O6 , cpx
Adopting for both components the standard state of pure component at the Tr and Pr of
reference, as for a pure component:
µ 0 ≡ GTr , Pr , ( 5.83 )
Table 5.38 Model of Lindsley et al. (1981) for binary system Mg2Si2O6-CaMgSi2O6.
{
r
Figure 5.27 Phase stability relations along binary join CaFeSi2O6-Fe2Si2O6 under various
P conditions. From Lindsley (1982). Reprinted with permission of The Mineralogical So-
ciety of America.
we also have
∆µ 0 ≡ ∆GTr , Pr . ( 5.84 )
The ⌬GTr ,Pr of reaction between pure components may be split into enthalpic, entropic,
and volumetric parts:
Figure 5.28 Phase relations in pyroxene quadrilateral for P ⫽ 15 kbar at two subsolidus
temperatures. From Lindsley (1982). Reprinted with permission of The Mineralogical So-
ciety of America.
gules notation) for clinopyroxene. Moreover, for the monoclinic phase, the inter-
action is split into enthalpic and volumetric terms, with no excess entropic term
of mixing. (In the interpretation of Lindsley et al. (1981), because Ca never occurs
in M1 sites, the configurational entropy is that of ideal mixing of Ca and Mg in
M2 sites, with no excess terms.)
Figure 5.27 shows phase stability limits in the (pseudo)binary system
CaFeSi2O6-Fe2Si2O6 in various P conditions, as reported by Lindsley (1982). At
high P (i.e., above 11–12 kbar), the stability relations are similar to those pre-
viously seen for the CaMgSi2O6-Mg2Si2O6 join: there is a miscibility gap between
monoclinic (cpx) and orthorhombic (opx) phases. Moreover, within a limited P
range, pigeonite forms. In the iron-rich side of the join, the stability diagram
shows additional complexities resulting from the extrinsic instability of ferrosilite
with respect to the fayalite plus liquid paragenesis (see also figure 5.24D).
For the binary join CaFeSi2O6-Fe2Si2O6 , Lindsley (1981) proposed a thermo-
290 / Geochemistry of Crystal Phases
Table 5.40 Binary interaction parameters for pyroxenes. Parameters refer to an ionic mixing
model in which n is the number of sites over which permutability is calculated—i.e., Gmixing ⫽
nRT (X1 ln X1 ⫹ X2 ln X2 ). Data in J/mole (H ), J/(mole ⫻ K) (S), and J/(bar ⫻ mole) (V ),
respectively.
Margules Paramters
Mixture H S V n Reference
CLINOPYROXENES (C2/c; P21 /c)
Mg2Si2O6-Fe2Si2O6 W12 ⫽ ⫺960.5 0.041 0.0557 2 (1)
W21 ⫽ 1190.1 ⫺0.081 ⫺0.0180
W12 ⫽ ⫺1558.6 24.389 0.0282 1 (1)
W21 ⫽ 2518.6 24.651 0.0226
Mg2Si2O6-CaMgSi2O6 W12 ⫽ 25,484 0.0 0.0812 1 (2)
W21 ⫽ 31,216 0.0 ⫺0.0061
Mg2Si2O6-CaFeSi2O6 W12 ⫽ 52,971 15.987 0.0416 2 (1)
W21 ⫽ 29,085 3.538 0.0294
W12 ⫽ 93,300 45.0 0.0 1 (3)
W21 ⫽ 20,000 ⫺28.0 0.0
Fe2Si2O6-CaMgSi2O6 W12 ⫽ 46,604 14.996 0.019 2 (1)
W21 ⫽ 22,590 3.940 0.072
W12 ⫽ 15,000 0.0 0.0 1 (3)
W21 ⫽ 24,000 0.0 0.0
Fe2Si2O6-CaFeSi2O6 W12 ⫽ 16,941 0.0 0.0059 1 (4)
W21 ⫽ 20,697 0.0 ⫺0.023
CaMgSi2O6-CaFeSi2O6 W12 ⫽ 13,984 ⫺0.001 0.0068 1 (1)
W21 ⫽ 17,958 0.004 ⫺0.0084
W12 ⫽ 12,000 0.0 0.0 1 (3)
W21 ⫽ 12,000 0.0 0.0
Mg2Si2O6-NaAlSi2O6 W12 ⫽ 30,000 0.0 0.0 1 (5)
W21 ⫽ 30,000 0.0 0.0
CaFeSi2O6-NaAlSi2O6 W12 ⫽ 25,200 17.05 0.0 1 (5)
W21 ⫽ 0.0 1.85 0.0 1 (5)
CaMgSi2O6-CaAl2SiO6 W12 ⫽ 30,918 18.78 0.146 1 (6)
W21 ⫽ 35,060 ⫺44.72 0.469
Redlich-Kister Parameters
Mixture H S V n Reference
CLINOPYROXENES (C2/c; P21 /c)
CaMgSi2O6-NaAlSi2O6 A0 ⫽ 12,600 9.45 0.0 1 (7)
A1 ⫽ 12,600 7.60 0.0
A2 ⫽ ⫺21,400 ⫺16.2 0.0
A0 ⫽ 19,800 ⫺4.4 0.0 1 (8)
A1 ⫽ 9600 9.1 0.0
A2 ⫽ ⫺8200 ⫺16.4 0.0
Silicates / 291
Redlich-Kister Parameters
Mixture H S V n Reference
Margules Parameters
Mixture H S V n Reference
ORTHOPYROXENES (Pbca)
Mg2Si2O6-Fe2Si2O6 W12 ⫽ 13,100 15.0 0.0 1 (1) T ⬍ 873 K
W21 ⫽ 3370 5.0 0.0
Mg2Si2O6-CaMgSi2O6 W12 ⫽ 25,000 0.0 0.0 1 (2)
W21 ⫽ 25,000 0.0 0.0
Mg2Si2O6-CaFeSi2O6 W12 ⫽ 60,000 0.0 0.0 1 (3)
W21 ⫽ 30,000 0.0 0.0
Fe2Si2O6-CaMgSi2O6 W12 ⫽ 16,000 0.0 0.0 1 (3)
W21 ⫽ 20,000 0.0 0.0
Fe2Si2O6-CaFeSi2O6 W12 ⫽ 30,000 0.0 0.0 1 (4)
W21 ⫽ 33,000 0.0 0.0
CaMgSi2O6-CaFe2Si2O6 W12 ⫽ 5000 0.0 0.0 1 (3)
W21 ⫽ 7000 0.0 0.0
Ideal Mixtures
Mg2Si2O6-Fe2Si2O6 T ⬎ 873 K (3)
References:
(1) Ottonello et al. (1992); based on interionic potential calculations
(2) Lindsley et al. (1981); experimental
(3) Saxena et al. (1986); based on phase equilibria
(4) Lindsley et al. (1981); experimental
(5) Chatterjee (1989); based on phase equilibria
(6) Gasparik and Lindsley (1980); estimate on the basis of various experiments
(7) Gasparik (1985); estimate on the basis of various experiments
(8) Cohen (1986); estimate on the basis of various experiments
(9) Ganguly and Saxena (1987); estimate on the basis of various experiments
(10) Ganguly (1973); based on crystal-chemical arguments
292 / Geochemistry of Crystal Phases
Figure 5.29 Excess Gibbs free energy surface at T ⫽ 1000 °C and P ⫽ 1 bar obtained
with Wohl model parameters of table 5.41 and consistent with results of static interionic
potential calculations in C2/c pyroxene quadrilateral. From G. Ottonello, Contributions to
Mineralogy and Petrology, Interactions and mixing properties in the (C2/c) clinopyroxene
quadrilateral, 111, 53–60, figure 6, copyright 䉷 1992 by Springer Verlag. Reprinted with
permission of Springer-Verlag GmbH & Co. KG.
dynamic model similar to that of Lindsley et al. (1981) for the CaMgSi2O6-
Mg2Si2O6 join. The model parameters are listed in table 5.39.
Figure 5.28 shows stability relations in the pyroxene quadrilateral, based on
experimental evidence, at P ⫽ 15 kbar. The miscibility gap contours depend
markedly on the mixing properties observed along the binary joins CaMgSi2O6-
Mg2Si2O6 and CaFeSi2O6-Fe2Si2O6 (cf. figures 5.26 and 5.27). Figure 5.28 con-
firms petrological observations on natural pyroxenes (figure 5.25)—i.e., at high
T the miscibility gap between monoclinic (C2/c) and orthorhombic (Pbca)
phases is small, with the formation, at intermediate compositions, of pigeon-
ite (P21 /c), which in turn causes the formation of two separate miscibility gaps
(C2/c-P21 /c and P21 /c-Pbca); at low T, pigeonite disappears and the single misci-
bility gap Pbca-C2/c expands throughout the compositional field.
Table 5.40 lists some of the binary interaction parameters that may be used
to describe the mixing properties of monoclinic and orthorhombic pyroxenes.
Readers are referred to sections 3.8 and 3.9 for the meanings of these parameters.
Some of the interaction parameters in table 5.40 are deduced from experimental
Silicates / 293
( ) (
Gexcess mixing = X1X2 W12X2 + W21X1 + X1X 3 W13X 3 + W31X1 )
( )
+ X1X 4 W14X 4 + W41X1 + X2X 4 W24X 4 + W42X2 ( )
+ X X (W X
3 4 34 4 +W X )
43 3
1
+ X1X2X 3 (W12 + W21 + W13 + W31 + W23 + W32 ) − W123
2
1
+ X1X2X 4 (W12 + W21 + W14 + W41 + W24 + W42 ) − W124
2
(5.86)
1
+ X1X 3X 4 (W13 + W31 + W14 + W41 + W34 + W43 ) − W134
2
1
+ X2X 3X 4 (W23 + W32 + W24 + W42 + W34 + W43 ) − W234
2
1
(
+ X1X2X 3X 4 W12 + W21 + W13 + W31 + W14 + W41 + W23
2
+ W32 + W24 + W42 + W34 + W43 − W1234 . )
The excess terms refer to a strict macroscopic formulation in which the ideal Gibbs free
energy of mixing term is given by
where Xi represents molar fractions of the components in the mixture. Using a strictly
macroscopic formulation leads to negative entropic interactions along binary joins in
which intracrystalline disorder occurs in two separate sites (i.e., M1 and M2: cf. table
5.41). A nonnegligible quaternary interaction term results (W1234 in eq. 5.86), because the
system is treated as quaternary whereas it is in fact ternary. The complete set of interaction
parameters of the model is listed in table 5.41.
Figure 5.30 (A) Simplified Gibbs free energy curves for various polymorphs along
enstatite-diopside join at T ⫽ 1300 °C. (B) Resulting solvus, spinodal field, and miscibility
gap compared with experimental data of McCallister and Yund (1977) on pyroxene un-
mixing kinetics (part B from Ganguly and Saxena (1992). Reprinted with permission of
Springer-Verlag, New York).
Figure 5.31 TTT diagram for a pyroxene of composition similar to C2 in figure 5.30.
Curves a, b, c, and d: cooling rates of a lava flow at increasing depths from surface (a) to
1 m inside the body (d ), based on calculations of Provost and Bottinga (1974). From Bu-
seck et al. (1982). Reprinted with permission of The Mineralogical Society of America.
Exsolution kinetics are appropriately described with the aid of TTT (Time-
Temperature-Transformation) diagrams, in which the cooling rates of the system cross the
kinetic boundaries of the beginning and end of the process. For example, figure 5.31 is a
TTT diagram for a pyroxene with a composition like that of C2 in figure 5.30 (Buseck et
al., 1982). Curves a, b, c, and d are cooling rates calculated by Provost and Bottinga (1974)
for a basaltic lava: curve a is the cooling rate at the surface of the lava flow, and curves b,
c, and d are those at increasing depths beneath the surface. These curves encounter the
kinetic curves of the beginning and end of heterogeneous nucleation (cpx-opx miscibility
gap or chemical solvus in figure 5.30B), homogeneous nucleation (coherent solvus), and
spinodal decomposition (spinodal field) at various values of time and temperature. Note
that the three processes take place under different temperature conditions, times being
equal.
Let us examine the significance of figure 5.31 in some detail: cooling rate a does not
cross any transformation curve; a pyroxene of composition similar to C2 at the surface of
the lava flow thus remains metastably homogeneous throughout the entire cooling process.
Cooling rate b encounters the starting curve for spinodal decomposition but not the curve
of the end of the process; the cooling rate of the system is still too high to allow completion
of the exsolutive process. Cooling rate c encounters the starting curve for heterogeneous
nucleation but not its end curve; the process starts but does not reach completion. At
lower temperatures, cooling rate curve c crosses the TT field of homogeneous nucleation;
298 / Geochemistry of Crystal Phases
a second process of nucleation and growth develops until completion. Cooling rate d,
which corresponds to a depth of about 1 meter beneath the lava flow surface, brings the
process of heterogeneous nucleation to completion.
5.5 Amphiboles
Figure 5.32 Structural features of amphiboles. (A) Double chain [T4O11]n seen along axis
c (a) and in perspective (b). (B) Double chain seen from terminal part and various cationic
positions (compare with figure 5.19, for analogies with the pyroxene structure).
(
W0−1X 2Y5Z 8 O 22 OH, F, Cl , ) 2
(5.90)
W X Y Z
Cation A M4 M1 M2 M3 T1 T2
Si ⫻ ⫻
Al ⫻ ⫻ ⫻ ⫻ ⫻
Fe3⫹ ⫻ ⫻ ⫻ ⫻(?) ⫻(?)
Ti ⫻ ⫻ ⫻ ⫻(?) ⫻(?)
Fe2⫹ ⫻ ⫻ ⫻ ⫻
Mn ⫻ ⫻ ⫻ ⫻
Mg ⫻ ⫻ ⫻ ⫻
Ca ⫻
Li ⫻ ⫻ ⫻ ⫻ ⫻
Na ⫻ ⫻
K ⫻
Table 5.43 Amphibole terminology and chemical classification. After Hawthorne (1981a), with
modifications. Note that X and Y here do not correspond to identification 5.90.
Gedrite
冦 NaX (Mg,Mn,Fe)7⫺Y AlY (AlX⫹Y ,Si8⫺X⫺Y) O22 (OH,F,Cl)2
X ⫹ Y ⱖ 1.00
Pnma
Table 5.44 Compositions (in weight %) of some natural amphiboles (from Deer et al.,
1983): (1) ⫽ anthophyllite from a serpentinite; (2) gedrite from a gedrite-kyanite-garnet
paragenesis; (3) cummingtonite from a oligoclase-biotite schist; (4) common hornblende from
a tonalite; (5) pargasite from a metamorphic limestone; (6) basaltic hornblende from a latite;
(7) glaucophane from a glaucophane schist.
Samples
5.44 lists chemical compositions of natural amphiboles (Deer et al., 1983), to-
gether with the corresponding structural formulae on a 24-anion basis (O2⫺,
OH⫺, F⫺, Cl ⫺ ). Because of the complex chemistry of natural amphiboles, calcu-
lation of the structural formula from a complete set of chemical analyses should
begin with an evaluation of the relative amounts of H2O, F, and Cl. If the analysis
302 / Geochemistry of Crystal Phases
Table 5.45 Occurrence of amphiboles in rocks. Note that X and Y here do not correspond
to site identification 5.90.
ORTHORHOMBIC AMPHIBOLES
Name: anthophyllite
Formula: NaX (Mg,Mn,Fe)7⫺Y AlY(AlX⫹Y ,Si8⫺X⫺Y)(OH,F,Cl)2 X ⫹ Y ⬍ 1.00
Occurrence: crystalline schists, peridotites, serpentinites
Name: gedrite
Formula: NaX (Mg,Mn,Fe)7⫺Y AlY(AlX⫹Y ,Si8⫺X⫺Y)(OH,F,Cl)2 X ⫹ Y ⱖ 1.00
Occurrence: amphibolites, eclogites, cordierite-bearing and garnet-bearing gneisses
MONOCLINIC AMPHIBOLES
Name: cummingtonite
Formula: (Mg,Mn,Fe)7Si8O22(OH,F,Cl)2
Occurrences: cornubianites, low-metamorphic schists, micaschists, (iron ores)
Name: actinote (tremolite, actinolite)
Formula: Ca2(Mg,Fe)5Si8O22(OH)2
Occurrence: crystalline schists, calcic cornubianites, cipollines
Name: common hornblende
Formula: NaX Ca2(Mg,Fe)5⫺XAlY (AlX⫹Y ,Si8⫺X⫺Y)O22(OH,F,Cl)2
Occurrence: calc-alkaline granites, sienites, diorites and corresponding effusive terms,
lamprophires, amphibolites, amphibole-bearing hornfels.
Name: pargasite
Formula: NaCa2(Mg,Fe)4AlSi6Al2O22(OH,F,Cl)2
Occurrence: paragneisses, methamorphosed limestones, amphibole peridotites
Name: basaltic hornblende
Formula: NaX Ca2(Mg,Fe)5⫺X(Al,Fe3⫹)Y(AlX⫹Y ,Si8⫺X⫺Y)O22(OH,F,Cl)2
Occurrence: basic volcanites, gabbros, and ultramafic rocks
Name: glaucophane
Formula: Na2(Mg,Fe)3Al2Si8O22(OH,F,Cl)2
Occurrence: glaucophanites, eclogites, prasinites
Name: arfvedsonite
Formula: NaNa2(Mg,Fe)4Fe3⫹Si8O22(OH,F,Cl)2
Occurrence: hypersodic granites, nepheline sienites, metamorphic rocks
is very accurate, one can begin directly with a structural calculation based on 24
anions (O2⫺, OH⫺, F⫺, Cl ⫺ ). However, if there are any doubts about the internal
precision of the data set, it is preferable to assume that the bulk charges of OH⫺,
F⫺, and Cl ⫺ amount to 2 and that the cationic charges are balanced by these two
negative charges plus 22 oxygens (O2⫺ ).
According to Robinson et al. (1982), the correct procedure for assigning site occupan-
cies is as follows.
Site X (M4)
Add the remaining parts of Mg, Fe2⫹, Zn, Mn, Ca, Li, and Na in this precise order up to
2.0. If the sum is substantially lower than 2.0, there is an analytical problem or, alterna-
tively, a high number of cationic vacancies on the site.
Site W
Add the remaining part of Ca with Li, Na, and K up to 1.0. If some Ca still remains after
this operation, the analysis is suspect. There are also problems if the sum exceeds 1.0.
Crystal-chemical data indicate that Na ⫹ K occupancies generally vary between 0 and 1,
and it is precisely this variability that sometimes renders structural reconstruction prob-
lematic.
Table 5.45 lists the main occurrences of the various compositional terms. The
name of the mineral reflects the dominant component in the mixture.
C2/m Amphiboles
There are three sites with pseudo-octahedral coordination (M1, M2, and M3)
whose mean dimensions exhibit straight correlation with the ionic radii of cations
occupying these lattice sites. The relationship is complicated by the fact that, for
sites M1 and M3, two of the coordinating anions may be OH⫺, F⫺, or Cl ⫺ instead
of O2⫺. Figure 5.33A shows the relationship between mean 〈M-O〉 distance and
304 / Geochemistry of Crystal Phases
Table 5.46 Structural data on amphibole end-members (from Smyth and Bish, 1988). Molar
volume in cm3/mole; molar weight in g/mole; cell edges in Å; cell volume in Å3.
{
W - Na0.63K0.30 Na0.38K0.12
X Fe1.7Mg0.3 Ca2 Ca1.8Mg0.2
Formula
Y Fe0.6Mg4.4 Fe1.10Mg3.25Al0.55 Mg5
Z Si8 Al1.86Si6.14 Al0.23Si7.8
Molar weight 853.72 870.30 824.57
Density 3.142 3.165 3.010
Molar volume 271.68 274.94 273.93
Z 2 2 2
System Monoclinic Monoclinic Monoclinic
Class 2/m 2/m 2/m
Group C2/m C2/m C2/m
Cell parameters:
a 9.51 9.910 9.863
b 18.19 18.022 18.048
c 5.33 5.312 5.285
 101.92 105.78 104.79
Volume 902.14 912.96 909.60
Reference Fisher (1966) Robinson et al. Hawthorne and
(1973) Grundy (1976)
{
W Na0.04 - Na0.05Ca0.03
冧 冧
X Na1.80Ca0.20 X X
Formula Fe1.47Mg5.53 Fe1.1Mg4.5Al1.2
Y Fe1Mg2.38Al1.58Ti0.06 Y Y
Z Al0.08Si7.92 Si8 Al1.8Si6.3
Figure 5.33 Correlations between algebraic mean of ionic sizes of intervening ions in M
positions (rM) and mean dimension of M sites (M1, M2, M3) in C2/m (A) and Pnma (B)
amphiboles. From Hawthorn (1981a). Reprinted with permission of The Mineralogical
Society of America.
mean ionic radius of intervening cations. This may be described by the two equa-
tions
and
where rM is the mean radius (algebraic mean) of cations occupying sites M1,
M2, and M3, and equation 5.92 explicitly takes into account the dimensions of
coordinating O3 anions (rO3 ).
Concerning tetrahedral sites, several authors agree that a straight correlation
exists between the amount of Al in the tetrahedral site and the mean site dimen-
sion. From the equations proposed by Robinson et al. (1973) we may derive
Based on equations 5.91 to 5.96, we may expect C2/m amphibole mixtures to have nearly
ideal behavior in regard to volume properties.
Any deviations from ideal behavior may be attributable to a high site distortion, rep-
resented in linear terms by parameter ⌬:
n l − l 2
∆ = ∑ l × 10 4 ,
i m
(5.97)
i =1
m
n
where li is the individual length of a single bond in the polyhedron, lm is the mean length
of polyhedral bonds, and n is the number of bonds in the polyhedron. Site distortion may
also be described in angular terms by parameter 2:
(Θ − Θm )
2
n
∑
i
σ2 = , (5.98)
i =1 n −1
where ⌰i is the individual bond angle and ⌰m is the ideal bond angle.
As table 5.47 shows, the observed distortions of the various sites are in some cases
quite high and there is also considerable variability in the degrees of polyhedral distortion
of the various compositional terms.
Hawthorne (1976) showed that the distortion parameter of site M2 is a function of
the formal charge on site M4. Because site M2 shows the most severe distortion, the simple
relationships connecting site M4 occupancy and cell edges may be complicated to some
extent by the distortion induced on site M2.
NaCa2Mg4AlSi6Al2O22(OH)2-NaCa2Fe4AlSi6Al2O22(OH)2
(pargasite–ferro-pargasite; Charles, 1980)
Na2Mg3Al2Si8O22(OH)2-Na2Fe3Fe2Si8O22(OH)2
(glaucophane-riebeckite; Borg, 1967)
Na2CaMg5Si8O22(OH,F)2-KNaCaMg5Si8O22(OH,F)2
(sodium–magnesio-richterite–potassium-magnesio-richterite; Huebner and
Papike, 1970)
Silicates / 307
Na2CaMg5Si8O22(OH,F)2-Na2CaFe5Si8O22(OH,F)2
(magnesio-richterite–ferro-richterite; Charles, 1974)
Na2Mg3Fe2Si8O22(OH)2-Na2Fe3Al2Si8O22(OH)2
(magnesio-riebeckite–ferro-glaucophane; Borg, 1967)
For the first four systems, the experimental data cover the entire compositional
field, whereas for magnesio-riebeckite–ferro-glaucophane, the data are limited to
the field XNa2Mg3Fe2Si8O22(OH)2 ⫽ 1.0 → 0.5. Within the studied compositional
range and within experimental approximation, the volume of the mixture is ideal.
Pnma Amphiboles
In orthorhombic amphiboles, as in monoclinic C2/m amphiboles, site dimensions
are linearly dependent on the ionic radii of occupying cations (figure 5.33B).
However, no empirical equations exist to derive M-O mean dimensions from cat-
ionic occupancy, because of the paucity of experimental information. The same
is true for sites M4, T1, T2, and A.
Concerning cell volume, the experimental data of Popp et al. (1976) on the
system Mg7Si8O22(OH)2-Fe7Si8O22(OH)2 indicate substantial ideality of mixing.
P21 /m Amphiboles
Volume properties in the system Mg7Si8O22(OH)2-Fe7Si8O22(OH)2 (cumming-
tonite-grunerite) show well-defined linearity in the compositional range
XFe7Si8O22(OH)2 ⫽ 0.60 → 1.00 (Klein, 1964; Viswanathan and Ghose, 1965); Ve-
gard’s law probably holds over the entire compositional field.
1
Fe 2 + + OH − ⇔ Fe 3+ + O 2 − + H2 (5.99)
2
T1 T2 M1
Phase ⌬ 2 ⌬ 2 ⌬ 2
Grunerite 0.29 0.8 0.78 15.8 2.25 36.0
Glaucophane 0.05 0.6 1.88 17.4 0.21 69.5
Tremolite 0.54 5.0 4.14 19.9 0.15 35.6
Fluor-richterite 2.06 14.3 5.60 22.4 0.62 43.1
Fluor-tremolite 0.38 6.2 3.45 19.3 0.01 46.9
Potassic pargasite 0.59 6.6 1.42 19.1 1.25 50.5
Ferro-tschermakite 0.63 4.8 0.61 17.0 4.00 57.3
Potassic ossi-kaersutite 0.29 5.8 1.16 18.9 14.35 47.9
Potassic arfvedsonite 1.05 11.4 5.54 21.4 0.51 44.6
Potassic ferri-taramite 1.25 3.6 1.53 13.8 2.45 45.8
Fluor-riebeckite 0.02 4.4 2.02 12.7 0.29 47.9
Ferro-glaucophane 0.06 0.7 2.04 14.8 0.39 70.5
M2 M3
Phase ⌬ 2 ⌬ 2
Grunerite 2.79 43.2 0.11 60.9
Glaucophane 15.77 35.0 0.34 85.0
Tremolite 5.52 22.9 0.09 43.6
Fluor-richterite 12.54 33.1 0.97 43.3
Fluor-tremolite 5.76 21.9 1.03 47.9
Potassic pargasite 5.98 24.3 0.01 75.7
Ferro-tschermakite 5.00 19.6 0.33 106.0
Potassic ossi-kaersutite 7.06 31.2 0.05 56.8
Potassic arfvedsonite 17.27 44.5 0.60 67.6
Potassic ferri-taramite 8.53 28.7 0.08 82.3
Fluor-riebeckite 15.48 39.1 0.26 68.7
Ferro-glaucophane 14.36 31.9 1.14 97.4
Figure 5.35 Structure of tremolite projected on (001) plane. Dashed lines on [T4O11]n
tetrahedral chains: displacement (highly exaggerated for illustrative purposes) due to ther-
mal decoupling. From Sueno et al. (1973). Reprinted with permission of The Mineralogi-
cal Society of America.
310 / Geochemistry of Crystal Phases
Table 5.48 Thermal expansion and compressibility of some amphibole end-members according to
Saxena et al. (1993) (1) and Holland and Powell (1990) (2).
sure of fluids (fH2O ) and by redox state (fO2 ). An appropriate discussion of such
limits is beyond the scope of this text, and readers are referred to the work of
Gilbert et al. (1982).
Table 5.48 summarizes thermal expansion and compressibility data for am-
phibole end-members according to the databases of Holland and Powell (1990)
and Saxena et al. (1993). Isobaric thermal expansion ( ␣, K⫺1 ) and isothermal
compressibility ( , bar⫺1 ) may be retrieved from the listed coefficients by applying
the polynomial expansions
Silicates / 311
α (T ) = α 0 + α 1 × T + α 2 × T −1 + α 3 × T −2 (5.100)
and
β ( T ) = β 0 + β1 × T + β 2 × T 2 + β 3 × T 3 . (5.101)
The P-derivative of the bulk modulus (K⬘; cf. section 1.15.2) is 4.0 in all cases.
in kJ/mole; S 0Tr ,Pr in J/(mole ⫻ K); V 0Tr ,Pr in cm3/mole. (1) Helgeson et al. (1978); (2) Berman
(1988); (3) Saxena et al. (1993); (4) Holland and Powell (1990). Heat capacity equation: CP ⫽ K1 ⫹
K2T ⫹ K3T ⫺2 ⫹ K4T 2 ⫹ K5T ⫺3 ⫹ K6T ⫺1/2 ⫹ K7T ⫺1.
According to Ghose (1982), there are three general conditions (rules) leading
to immiscibility between two amphibole components:
Base Component
(Tremolite) Exchange Resulting Component
䡺 Ca2Fe5Si8O22(OH)2
䡺 Ca2Mg5Si8O22(OH)2 Fe⇔Mg Ferro-actinolite
䡺 NaCa2Mg5AlSi7O22(OH)2
䡺 Ca2Mg5Si8O22(OH)2 䡺 Si⇔NaAAlIV Edenite
䡺 Ca2Mg3Al2Si6O22(OH)2
䡺 Ca2Mg5Si8O22(OH)2 MgSi⇔AlVIAlIV Tschermak
䡺 Ca2Mg3Fe23⫹Si6Al2O22(OH)2
䡺 Ca2Mg5Si8O22(OH)2 MgSi⇔Fe 3⫹,VI
Al IV
Ferri-Tschermak
䡺 Na2Mg3Al2Si8O22(OH)2
䡺 Ca2Mg5Si8O22(OH)2 CaMg⇔NaM4AlVI Glaucophane
䡺 Ca2Mg5AlSi6TiAlO22(OH)2
䡺 Ca2Mg5Si8O22(OH)2 MgSi2⇔TiIVAlIVAlVI Ti-Tschermak
䡺 Mg7Si8O22(OH)2
䡺 Ca2Mg5Si8O22(OH)2 Ca⇔MgM4 Cummingtonite
(5.102)
2 Mg 5 Si 8 O22 (OH )2 + Fe 2 Fe 5 Si 8O22 (OH )2 ,
1 1 M4 Y
⇔ Ca M 4 Y
2 7
Mg-actinolite Fe-cummingtonite
316 / Geochemistry of Crystal Phases
where superscripts M4 and Y define the structural position of the element. According
to Mueller (1961), the observed intracrystalline partitioning of Fe2⫹ and Mg2⫹ may be
attributable to ideal mixing behavior in actinolite and to regular mixing behavior in cum-
mingtonite. Later, in order to assess the mixing properties of cummingtonite better,
Mueller (1962) proposed an intracrystalline exchange of the type
in which mixing on sites Y (M1, M2, M3) is ideal, but not on site M4. The constant of
equilibrium 5.103 may be expressed as
W
(1 − X i )
2
ln γ i = (5.106)
RT
(
X Fe,Y 1 − X Fe,M 4 )
(
exp 1 − 2X Fe,Y )W Y
RT
K103 = × .
(
X Fe,M 4 1 − X Fe,Y )
(
exp 1 − 2X Fe,M 4 ) WM 4 (5.107)
RT
Hafner and Ghose (1971) solved equation 5.107 on the basis of Mueller’s (1960) data
and obtained K103 ⫽ 0.19; WM4 /RT ⫽ 1.58, and WY /RT ⫽ 0.54. Their results thus imply
a certain nonideality of mixing on sites Y as well.
Ghose and Weildner (1971, 1972) experimentally studied the intracrystalline distribu-
tion of Fe2⫹-Mg2⫹ in natural cummingtonites annealed at various T conditions and Pto-
tal ⫽ 2 kbar in the presence of H2O (3 weight %). The observed distribution of Fe
2⫹
between M4 and Y sites may be due either to regular mixing (with K103 ⫽ 0.26, WM4 ⫽
1.09 kcal/mole, WY ⫽ 1.23 kcal/mole at T ⫽ 600 °C; cf. Ghose and Weidner, 1971; and
figure 5.39A) or to ideal mixing on both M4 and Y sites, as shown in figure 5.39B.
Silicates / 317
Figure 5.39 Intracrystalline distribution of Fe2+-Mg2+ between Y (M1, M2, M3) and M4
sites in natural cummingtonites. (A) Regular site-mixing model. (B) Ideal site-mixing
model. (C) Arrhenius plot of equilibrium constant.
In the example above, we did not make any distinction among M1, M2, and M3 site
populations, because these sites were grouped in the Y notation. Studying in detail the
structure of sodic amphiboles of the riebeckite-glaucophane series, Ungaretti et al. (1978)
observed that the Fe2⫹-Mg2⫹ population is not randomly distributed between M1 and M3
sites, being Fe2⫹ preferentially stabilized in M3. Considering the intracrystalline exchange
the resulting equilibrium constant, based on the simply hyperbolic distribution of cations
(figure 5.40), may be reconducted to an ideal site-mixing behavior:
K108 =
a Mg,M 3 aFe,M1
≡
(1 − X Fe,M 3 )X Fe,M1
= 0.403.
aFe,M 3 a Mg,M1 (
X Fe,M 3 1 − X Fe,M1 ) (5.109)
318 / Geochemistry of Crystal Phases
We recall to this purpose that “ideal site mixing” does not necessarily mean that the mix-
ture as a whole is ideal, but that the intracrystalline site distribution may be described by
a constant (without the necessity of introducing site-interaction parameters; cf. equation
5.109, for example). Generally speaking, if the adopted constant differs from 1 (as in the
case of sodic amphiboles in figure 5.40 and of cummingtonites in figure 5.39), the mixture
as a whole is obviously nonideal because site permutability does not reach the maximum
value attained by the ideal condition (cf. section 3.8.1).
The nature of the miscibility gap between orthorhombic and monoclinic am-
phiboles has been recently reconsidered by Will and Powell (1992) with an ex-
tended application of Darken’s Quadratic Formalism (Darken, 1967). This for-
malism assumes that, within a restricted compositional field (“simple regions”),
excess energy terms may be described by simple mixing models (regular mixture,
ideal mixture), even if the intermediate compositions necessitate more complex
treatment. Under this proviso, a general binary join may be considered to be
composed of two simple regions, each including one end-member, connected by
a transitional zone, as depicted schematically in figure 5.41. The complex activity-
composition relationship observed within the simple region is then reconducted
to a regular (or even ideal) mixture between the real end-member in the simple
region and a “fictive” end-member on the opposite side of the join, by appropriate
selection of the standard state properties of the fictive term.
Assuming ideal mixing within the simple regions, the principle of equality of
potentials between phases at equilibrium reduces to
Silicates / 319
where t,*␣ is the fictive standard state chemical potential of component i in phase
␣, and KD,i is the distribution coefficient KD,i ⫽ Xi,␣ /Xi, .
Consideration of the actual extension of miscibility gaps in natural systems
led Will and Powell (1992) to establish systematic relationships between the actual
free energy of pure components and their fictive potentials in the phases of inter-
est, as listed in table 5.51. Note that, amphiboles being multisite phases, their
ideal activity in a chemically complex phase is expressed in terms of multiple
product of site ionic fractions (see section 3.8.7). For anthophyllite
(䡺Mg2Mg3Mg2Si4Si4O22(OH)2 ), for instance, we have
(it is assumed that Al-Si mixing takes place only on T 2 sites and that M1 to M3
sites are energetically equivalent).
Table 5.51 Energy terms relating standard state chemical potentials of pure components (0i,␣) to fictive potentials in the
host phase (*
i,) in the Will and Powell (1992) application of Darken’s Quadratic Formalism to amphiboles.
*cumm,act ⫽ 0cumm,cumm ⫹ 51.5 (⫾4.5) *gl,act ⫽ 0gl,gl ⫹ 48.7 (⫾6.7) (low P) *anth,act ⫽ 0anth,oam ⫹ 35.4 (⫾4.0)
*tr,cumm ⫽ 0tr,act ⫹ 16.0 (⫾3.8) *gl,act ⫽ 0gl,gl ⫹ 31.8 (⫾2.9) (high P) *tr,oam ⫽ 0tr,act ⫹ 34.5 (⫾6.5)
*hb,cumm ⫽ 0hb,act ⫹ 62.6 (⫾8.6) *rich,act ⫽ 0rich,gl ⫹ 25.0 (⫾3.0) *hb,oam ⫽ 0hb,act ⫹ 40.1 (⫾4.8)
*ed,cumm ⫽ 0ed,act ⫹ 41.1 (⫾3.7) *tr,gl ⫽ 0tr,act ⫹ 46.3 (⫾3.4) *ed,oam ⫽ 0ed,act ⫹ 44.3 (⫾4.1)
*parg,cumm ⫽ 0parg,act ⫹ 12.6 (⫾1.3) *hb,gl ⫽ 0hb,act ⫹ 32.1 (⫾2.0) *parg,oam ⫽ 0parg,act ⫹ 54.5 (⫾4.4)
*ed,gl ⫽ 0ed,act ⫹ 59.8 (⫾3.8)
Orthoamphibole-actinolite series *parg,gl ⫽ 0parg,act ⫹ 46.0 (⫾2.5)
T ⬵ 650 °C
cumm ⫽ cummingtonite (P21/m) 䡺 Mg2Mg3Mg2Si4Si4O22(OH)2
tr ⫽ tremolite (C2/m) 䡺 Ca2Mg3Mg2Si4Si4O22(OH)2
hb ⫽ hornblende (C2/m) 䡺 Ca2Mg3(MgAl)(Si3Al)Si4O22(OH)2
ed ⫽ edenite (C2/m) NaCa2Mg3Mg2(Si3Al)Si4O22(OH)2
parg ⫽ pargasite (C2/m) NaCa2Mg3(MgAl)(Si2Al2)Si4O22(OH)2
gl ⫽ glaucophane (C2/m) 䡺 Na2Mg3(AlAl)Si4Si4O22(OH)2
rich ⫽ richterite (C2/m) Na(CaNa)Mg3Mg2Si4Si4O22(OH)2
anth ⫽ anthophyllite (Pnma) 䡺 Mg2Mg3Mg2Si4Si4O22(OH)2
act ⫽ actinolite; oam ⫽ orthoamphibole
Silicates / 321
5.6 Micas
Figure 5.42 Structural scheme of micas. (A) Tetrahedral sheet with tetrahedral apexes
directed upward. (B) Structure of mixed layers along axis Y; a, b, and c are edges of ele-
mentary cell unit.
322 / Geochemistry of Crystal Phases
Figure 5.43 Sliding between tetrahedral sheets of mixed 2:1 talc layer. (a) Set of positions
1 occupied. (b) Set of positions II occupied. From Bailey (1984a). Reprinted with permis-
sion of The Mineralogical Society of America.
(1956) have shown that six standard ideal polytypes may be formed, with period-
icity between one and six layers: 1M, space group C2/m,  ⫽ 100; 2M, space
group C2/c,  ⫽ 95; 3T, space group P3112; 2Or, space group Ccmm,  ⫽ 90;
2M2, space group C2/c,  ⫽ 98; 6H, space group P6122 (first number identifies
periodicity—i.e., polytype 3T ⇒ periodicity ⫽ 3; polytype 1M ⇒ periodicity ⫽
1, etc.). In trioctahedral micas, the most frequent polytype is 1M, followed by
3T and 2M1. In dioctahedral micas, the most frequent is 2M1, followed by
1M and 3T (Bailey, 1984a). The diadochy of Al (of Fe3⫹ ) on Si4⫹ normally
has the following ratios: 0:4, 0.5:3.5, 1:3, 2:2, and 3:1. Thus, polyanionic
groups composed of two tetrahedral sheets may be of these types:
[(OH)2Si4O10 ]6⫺; [(OH)2Al0.5Si3.5O10 ]6.5⫺; [(OH)2AlSi3O10 ]7⫺; [(OH)2Al2Si2O10 ]8⫺;
[(OH)2Al3SiO10 ]9⫺.
The intermediate octahedral sheet is normally made up of cations of charge
2 or 3 (Mg, Al, Fe2⫹, Fe3⫹, or, more rarely, V, Cr, Mn, Co, Ni, Cu, Zn), but in
some cases cations of charge 1 (Li) and 4 (Ti) are also found. In the infinite
octahedral sheet, formed by the sharing of six corners of each octahedron, there
may be full occupancy of all octahedral sites (“trioctahedral micas”); alterna-
tively, one site out of three may be vacant (“dioctahedral micas”). Nevertheless,
the primary classification of micas is based on the net charge of the mixed 2:1
layer. In “common micas” this charge is close to 1, whereas in “brittle micas” it
Silicates / 323
The main terms of trioctahedral common micas are phlogopite and biotite. Table 5.53
represents biotite as a mixture of Fe2⫹- and Mg2⫹-bearing terms, in variable proportions
on site Y, and phlogopite as a pure end-member. Current nomenclature is somewhat arbi-
trary; trioctahedral common micas with Mg:Fe2⫹ ratios less than 2:1 are generically de-
fined as “biotites,” trioctahedral common micas with Mg:Fe2⫹ ratios higher than 2:1 are
called “phlogopites.” Phlogopite is a marker of thermometamorphism in magnesian lime-
stones, where it is formed by reaction between dolomite and potash feldspar in hydrous
conditions, or by reaction between muscovite and dolomite according to
and
( 5.112.2)
Table 5.52 Structural parameters for some trioctahedral and dioctahedral micas
(from Smyth and Bish, 1988).
TRIOCTAHEDRAL MICAS
Name Phlogopite Polylithionite Zinnwaldite
Polytype 1M 1M 1M
Interlayer X cations K0.77Na0.16Ba0.05 K0.89Na0.06Rb0.05 K0.9Na0.05
Octahedral Y cations Mg3 Al1.3Li1.7 Al1.05Fe0.93Li0.67
Tetrahedral Z cations Al1.05Si2.95 Al0.6Si3.4 Al0.91Si3.09
Hydroxyls and fluorine (OH)0.7F1.3 (OH)0.5F1.5 (OH)0.8F1.2
Molar weight (g/mole) 421.831 396.935 428.995
Molar volume (cm3/mole) 146.869 140.5 144.55
Density (g/cm3) 2.872 2.825 2.986
Formula units per unit cell 2 2 2
System Monoclinic Monoclinic Monoclinic
Class 2/m 2/m 2
Group C2/m C2/m C2
Cell parameters (Å):
a 5.3078 5.20 5.296
b 9.1901 9.01 9.140
c 10.1547 10.09 10.096
 100.08 99.28 100.83
Volume (Å3) 487.69 466.6 480.00
Reference Hazen and Sartori (1976) Guggenheim and
Burnham (1973) Bailey (1977)
DIOCTAHEDRAL MICAS
Table 5.53 Classification and compositional terms of micas (from Bailey 1984a; modified).
COMMON MICAS
Trioctahedral Phlogopite KMg3(Si3Al)O10(OH,F)2
Na-phlogopite NaMg3(Si3Al)O10(OH,F)2
Biotite K(Mg,Fe)3(Si3Al)O10(OH,F)2
Annite KFe32⫹(Si3Al)O10(OH,F)2
Ferri-annite KFe32⫹(Si3Fe3⫹)O10(OH,F)2
Polylithionite K(Li2Al)Si4O10(OH,F)2
(ex-lepidolite)
Taeniolite K(Mg2Li)Si4O10(OH,F)2
Siderophyllite K(Fe2⫹Al)(Si2Al2)O10(OH,F)2
Zinnwaldite K(Li,Fe,Al)3(Si,Al)4O10(OH,F)2
Muscovite KAl2(Si3Al)O10(OH,F)2
Paragonite NaAl2(Si3Al)O10(OH,F)2
Phengite K[Al1.5(Mg,Fe2⫹)0.5](Si3.5Al0.5)O10(OH,F)2
2⫹
Dioctahedral Illite K0.75(Al1.75R0.25 )(Si3.5Al0.5)O10(OH,F)2
Celadonite K(Mg,Fe,Al)2Si4O10(OH,F)2
2⫹ 3⫹
Glauconite K(R1.33 R0.67 )(Si3.67Al0.33)O10(OH,F)2
BRITTLE MICAS
Trioctahedral Clintonite Ca(Mg2Al)(SiAl3)O10(OH,F)2
Bityite Ca(Al2Li)(Si2AlBe)O10(OH,F)2
Anandite BaFe32⫹(Si3Fe3⫹)O10(OH)S
Dioctahedral Margarite CaAl2(Si2Al2)O10(OH,F)2
biotites of effusive rocks are also generally richer in Fe3⫹ and Ti (and poorer in Fe2⫹ ) with
respect to biotites of intrusive counterparts.
In thermal metamorphism, biotite occurs in the clorite-sericite facies as a dispersed
phase in the argillitic matrix and is stable up to low-grade cornubianites. In regional meta-
morphism, biotite is typical of argillitic and pelitic rocks up to the staurolite-garnet facies
(biotite, biotite-sericite, biotite-chlorite, and albite-biotite schists, and garnet-staurolite
micaschists).
Muscovite is the main term of the common dioctahedral micas, and it is found in
acidic intrusive rocks (biotite and biotite-muscovite granites), although in subordinate
quantities with respect to biotite. It is a common mineral in aplitic rocks and is peculiar to
fluorine metasomatism in the contact zones between granites and slates (“greisenization”).
In metamorphic rocks, muscovite occurs in low-grade terrains of the regional meta-
morphism (albite-chlorite-sericite schists). It must be noted here that the term “sericite”
identifies fine-grained white micas (muscovite, paragonite).
Table 5.55 lists analyses of natural micas taken from the collection of Deer et
al. (1983).
Table 5.54 Occurrences of main compositional terms of mica in igneous, metamorphic and
sedimentary rocks.
sheets of the 2:1 mixed layer to accommodate the metal-to-oxygen interionic dis-
tances, with positional rotations of tetrahedra on the basal plane of oxygens, as
discussed by Hazen and Wones (1972).
Figure 5.44 shows the conformation of the hexagonal rings on the tetrahedral
sheet for limiting values of the rotational angle ( ␣r ⫽ 0° and 12°, respectively).
Due to structural constraints, the value of ␣r depends directly on the mean cation-
to-oxygen distances in [TO4 ] tetrahedra (dt ) and in octahedra (do ), according to
3 3 d o sin ψ
cos α r = , (5.113)
4 2 dt
where is the “octahedral flattening” angle (cf. Hazen and Finger, 1982), which
is also a function of do :
As already noted, the structural limits of the rotational angle are 0° and 12°.
This implies that the ratio of mean octahedral and tetrahedral distances is re-
stricted in the range
do
1.235 ≤ ≤ 1.275 (5.115)
dt
(Hazen and Finger, 1982). Mean do and dt distances vary as a function of the
ionic radii of intervening cations and, chemical composition being equal, vary
Silicates / 327
Table 5.55 Chemical analyses of natural micas (from Deer et al., 1983). Note that ionic
fractions are retrieved on a 24-anion basis—i.e. double the canonical formula. (1) Muscovite
from a low-grade metamorphic prasinite schist; (2) glauconite from a sandstone; (3)
phlogopite from a marble; (4) biotite from a quartz-bearing latite; (5) lepidolite from a
pegmatite.
Figure 5.44 Sketch of tetrahedral rotational angle ␣r for two limiting conditions, ␣r ⫽ 0°
and ␣r ⫽ 12°. From Hazen and Wones (1972). Reprinted with permission of The Mineral-
ogical Society of America.
Let us calculate the effects of intensive variables on the do distance, considering phlog-
opite [KMg3(Si3Al)O10 (OH)2 ] as an example. The octahedral cation is Mg2⫹; the mean
octahedral distance at P ⫽ 1 bar and T ⫽ 25 °C is about 2.06 Å (Hazen and Burnham,
1973). The mean polyhedral linear thermal expansion is ␣l ⫽ 1.4 ⫻ 10⫺5(°C⫺1 ), and the
mean polyhedral linear compressibility is l ⫽ 1.7 ⫻ 10⫺4 (kbar⫺1 ). We know that the
upper stability limit is do ⫽ 2.110 Å. We can solve the limiting P and T conditions by
applying the equality
( )
2.06 1 + 1.4 × 10 −5 T − 1.7 × 10 −4 P = 2.110 ( 5.118 )
and
Silicates / 329
2⫹
General Formula K(R2.5 Al0.5)(Si2.5Al1.5)O10(OH)2
Ni T ⫽ 11P ⫹ 2100 T ⫽ 11P ⫹ 3400
Mg T ⫽ 12P ⫹ 1000 T ⫽ 12P ⫹ 3400
Co T ⫽ 12P ⫹ 300 T ⫽ 12P ⫹ 2800
Fe T ⫽ 13P ⫺ 700 T ⫽ 13P ⫹ 1700
Mn T ⫽ 14P ⫺ 2400 T ⫽ 14P
( )
2.06 1 + 1.4 × 10 −5 T − 1.7 × 10 −4 P = 2.035 ( 5.120 )
and
( )
T o C ≈ 12P(kbar ) − 900. ( 5.121)
Table 5.56 lists intrinsic stability limits for common trioctahedral micas with gen-
eral formulas KR2⫹ 2⫹
3 (Si3Al)O10 (OH)2 and K(R 2.5 Al0.5 )(Si2.5Al1.5 )O10 (OH)2 , cal-
culated by Hazen and Finger (1982) with the method described above. Based on
these calculations, trioctahedral common micas with general formula
KR2⫹3 (Si3Al)O10 (OH)2 are structurally unstable for cations Fe
2⫹
and Mn2⫹,
which are too large and exceed the limiting value do /dt ⫽ 1.275. To keep the ratio
within limits, it is necessary to increase the value of dt , by means of a coupled
substitution involving Al (cf. table 5.56):
Si IV + R 2 + , VI ⇔ Al VI + Al IV. ( 5.122 )
Other elemental exchanges capable of reducing the do /dt ratio (Hazen and Finger,
1982) are
( )
R 2 + , VI + OH − ⇔ R 3 + , VI + O 2 − ( 5.123)
330 / Geochemistry of Crystal Phases
Al IV ⇔ Fe 3 + ,IV (5.124)
2R 2 + , VI ⇔ Ti 4 + + VY ( 5.125)
(
2KMg 3 Si 3 Al O 10 OH) ( ) 2
⇔ KAlSi 2 O 6 + KAlSiO 4 + 3Mg 2 SiO 4 + 2H 2 O.
(5.126)
phlogopite leucite kalsilite forsterite
Hewitt and Wones (1984) have shown that the equilibrium constant for the above reaction
may be expressed as a polynomial function of P (bar) and T (K):
log K 126 = 2 log fH2 O = −16633T −1 + 18.1 + 0.105(P − 1)T −1. ( 5.127 )
( )
KMg 3 Si 3 Al O 10 OH ( ) 2
+ 3 SiO 2 ⇔ KAlSi 3 O 8 + 3 MgSiO 3 + H 2 O .
( 5.128 )
phlogopite quartz sanidine enstatite
Silicates / 331
Figure 5.45 P-T stability curve of phlogopite compared with the incipient melting curves
of granite and basalt. Reprinted from H. S. Yoder and H. P. Eugster, Geochimica et Cos-
mochimica Acta, 6, 157–185, copyright 䉷 1954, with kind permission from Elsevier Sci-
ence Ltd., The Boulevard, Langford Lane, Kidlington 0X5 1GB, UK.
According to Hewitt and Wones (1984), the equilibrium constant may be expressed in
the form
log K 128 = log fH2 O = −2356T −1 + 4.76 + 0.077(P − 1)T −1. ( 5.129 )
With increasing P and T, the equilibrium curve of equation 5.129 reaches an invariant
condition, determined by the appearance of silicate melt (T ⫽ 835 °C, P ⫽ 0.45 kbar;
Wones and Dodge, 1977).
Concerning the ferriferous term annite [KFe32⫹(Si3Al)O10 (OH)2 ], the most important
experimentally studied reactions are as follows:
(
2 KFe 3 Si 3 Al O 10 OH ) ( ) 2
+ H 2 O ⇔ 2 KAlSi 3 O 8 + 3 Fe 2 O 3 + 3 H 2
( 5.130 )
annite sanidine hematite
with
(
2 KFe3 Si 3Al O10 OH ) ( ) 2
⇔ 3 Fe2SiO4 + KAlSi2O6 + KAlSiO4 + 2 H2O,
( 5.133)
annite fayalite leucite kalsilite
which is analogous to reaction 5.126 valid for the magnesian component, with
with
( ) ( )
KFe3 Si 3Al O10 OH
2
+ 3 H2 ⇔ KAlSi 3O8 + 3 Fe+ 4 H2O .
(5.138)
annite sanidine metal
Figure 5.46 shows the stability field of the annite end-member at 1 kbar total
pressure, based on equilibria 5.131, 5.133, and 5.135, as a function of temperature
and hydrogen fugacity in the system. The figure also shows the positions of the
iron-wuestite (IW) and wuestite-magnetite buffer (WM).
Muscovite-Paragonite
The P-T stability curve of muscovite intersects the incipient melting curve of
granite in hydrous conditions at about 2.3 kbar total pressure and T ⫽ 650 °C.
Thus, muscovite may crystallize as a primary phase from granitic melts above
these P and T conditions (interstitial poikilitic crystals) or may form by reaction
with pristine solid phases at lower P and T (muscovite as dispersed phase within
feldspars, for instance). In this second type of occurrence, the following two equi-
libria are of particular importance:
( )
KAl 2 Si 3 Al O 10 OH ( ) 2
⇔ KAlSi 3 O 8 + Al 2 O 3 + H 2 O
(5.139)
muscovite sanidine corundum
Silicates / 333
Figure 5.46 Stability field of annite at Ptotal ⫽ 1 kbar, as a function of T and fO2. From
Hewitt and Wones (1984). Reprinted with permission of the Mineralogical Society of
America.
( ) ( )
KAl2 Si3 Al O10 OH
2
+ SiO 2 ⇔ KAlSi3O 8 + Al2 SiO 5 + H2 O .
(5.140)
muscovite quartz K- feldspar
Equilibrium 5.139 has been investigated by several authors. According to Hewitt and
Wones (1984), the equilibrium constant may be expressed as a polynomial function of P
(bar) and T (K):
log K 139 = log fH2 O = 8.5738 − 5126T −1 + 0.0331(P − 1)T −1. ( 5.141)
The univariant equilibrium 5.139 for Ptotal ⫽ PH2O, based on the experimental data of
Chatterjee and Johannes (1974), is compared in figure 5.47A with the experimental results
of Ivanov et al. (1973) for XH2O ⫽ XCO2 ⫽ 0.5. Figure 5.47A also superimposes the melting
curves of granite in the presence of a fluid phase of identical composition. Note that the
composition of the fluid dramatically affects stability relations: if the amount of H2O in
the fluid is reduced by half (XH2O ⫽ XCO2 ⫽ 0.5), the stability field of muscovite is re-
stricted to approximately P ⬎ 5.5 kbar and T ⬎ 680 °C.
According to the experiments of Chatterjee and Johannes (1974), the equilibrium
constant for equation 5.140 may be expressed within the stability field of andalusite in
the form
334 / Geochemistry of Crystal Phases
Figure 5.47 Extrinsic stability curves for muscovite and paragonite, based on (A) equilib-
ria 5.139 and 5.144 (quartz absent) and (B) equilibria 5.140 and 5.145 (quartz present).
Incipient melting curves of granite for XH2O ⫽ 0.5 to 1, and stability curves of Al2SiO5
polymorphs according to Richardson et al. (1969), are superimposed.
( ) ( )
NaAl2 Si3 Al O10 OH
2
⇔ NaAlSi3O 8 + Al2 O 3 + H2 O
(5.144)
paragonite albite corundum
Silicates / 335
Figure 5.48 Extrinsic stability limits of muscovite, based on equilibrium 5.140 for vari-
able amounts of H2O component in fluid. Also shown are incipient melting curves of
granite for various XH2O isopleths and stability curves of Al2SiO5 polymorphs, according
to Holdaway (1971) From Kerrick (1972), American Journal of Science, 272, 946–58. Re-
printed with permission of American Journal of Science.
(
NaAl2 Si3Al ) O (OH)
10 2
+ SiO2 ⇔ NaAlSi3O8 + Al2SiO5 + H2O .
(5.145)
paragonite quartz albite
Figure 5.49 Stability field of margarite plus quartz in the CaO-Al2O3-SiO2-H2O system
for various values of aH2O, fluid. From Chatterjee (1974). Reprinted with permission of The
Mineralogical Society of America.
Margarite
Chatterjee (1974) studied in detail the equilibrium
( )
CaAl2 (Si2 Al2 ) O10 OH ⇔ CaAl2Si2O8 + Al2O3 + H2O
2 (5.146)
margarite anorthite corundum
in the pressure range 1 to 7 kbar. According to Hewitt and Wones (1984), the
equilibrium constant varies with P and T as
( ) ( )
4CaAl2 Si2 Al2 O10 OH + 3SiO2 ⇔ 2Ca2 Al3Si3O12 OH + 5Al2SiO5 + 3H2O
2
( )
margarite quartz zoisite
(5.148)
and
( ) ( )
CaAl2 Si2 Al2 O10 OH + 3SiO2 ⇔ CaAl2Si2O8 + Al2SiO5 + 3H2O.
2 (5.149)
margarite quartz anorthite
Figure 5.49 shows the P-T stability field of margarite plus quartz in the system
CaO-Al2O3-SiO2-H2O, defined by Chatterjee (1976) for various values of H2O
activity in the fluid.
All these phase stability relations may be expressed, as we have seen, as func-
Silicates / 337
Figure 5.50 Chemistry of natural biotites as a function of the redox state of the system.
Solid lines delineate the compositional trend imposed by the ruling buffer. From Speer
(1984). Reprinted with permission of The Mineralogical Society of America.
tions of the fugacity of a gaseous component in the fluid phase (i.e., fH2O, fH2;
see, however, section 9.1 for the significance of the term “fugacity”). It must be
emphasized here that the stability and chemical composition of natural micas are
also affected by the redox state of the system, which may itself be expressed as
the fugacity of molecular oxygen in fluid phases. Wones and Eugster (1965) have
shown that the oxidation state of biotites may be described as the result of mixing
of three compositional terms: KFe2⫹ 3 AlSi3O10 (OH)2 , KMg3AlSi3O10 (OH)2 , and
3⫹
KFe3 AlSi3O12H1 . In a ternary field of this type (figure 5.50), the compositional
trend of micas depends on the buffer capability of the system (for the meaning
of the term “buffer”, see section 5.9.5).
Figure 5.51A shows in greater detail how the stability field of biotite (phlogo-
pite-annite pseudobinary mixture) is affected by fO2-T conditions for a bulk pres-
338 / Geochemistry of Crystal Phases
Figure 5.51 fO2-T stability field of biotite. (A) Pseudobinary phlogopite-annite mixture.
Numbers at experimental points indicate observed Fe/(Fe ⫹ Mg) atom ratio. (B) Annite.
HM ⫽ hematite-magnetite buffer; NNO ⫽ Ni-NiO buffer; MW ⫽ magnetite-wuestite
buffer; QFM ⫽ quartz-fayalite-magnetite buffer. From Wones and Eugster (1965). Re-
printed with permission of The Mineralogical Society of America.
sure of 2.07 kbar. Numbers at the various experimental points indicate the com-
position of the mixture. Tie lines connect points of identical composition at
opposite sides of the stability region. Figure 5.51B shows the stability field of the
annite end-member for the same bulk pressure. We have already noted that, for
structural reasons, pure annite is intrinsically unstable at any P-T condition, and
that the presence of Fe3⫹ is required to constrain the angular distortion of [SiO4 ]
groups within structurally acceptable limits. The lower f O 2-T stability limit of
annite must thus be ascribed to this structural constraint (the amount of Fe3⫹ in
oxidized annite obviously diminishes with decreasing f O 2 in the system; see also
Wones and Eugster, 1965).
Silicates / 339
It is of interest here to recall the method suggested by Tardy and Garrels (1974) for
the empirical estimate of the Gibbs free energy of layer silicates. This method, which is
even simpler than the combinatory procedures mentioned in section 3.6, has an accuracy
comparable to the discrepancy of existing experimental data.
According to Tardy and Garrels (1974), the Gibbs free energy of formation of layered
silicates may be expressed as the simple sum of fictive energies ascribable to the constituent
oxides, after suitable “decomposition” of the chemistry of the phase. Consider, for in-
stance muscovite and paragonite: the Gibbs free energy of formation from the elements
at T ⫽ 298.15 K and P ⫽ 1 bar may be expressed as the sum of the fictive silicate compo-
nents (Gf,sil; see table 5.58):
1 0 3
G 0f ,KAl AlSi O = G f , sil,K O + G 0f , sil, Al O + 3G 0f , sil, SiO + G 0f , sil,H O
3 10(OH) 2
2
2 2
2 2 3 2 2
=
1
2
(
−786.6 +
3
2
) ( ) (
−1600.0 + 3 −856.0 + −247.7 ) ( ) (5.150)
= −5609.0
1 0
G 0f , NaAl2 AlSi3O10(OH)2 = G +L
2 f , sil,Na2O (5.151)
= −5556.3 (kJ/mole ).
Note that the Gibbs free energy of formation from the elements of the fictive constituent
oxides does not correspond to the actual thermodynamic value. It differs from it by an
empirical term ( ⌬G 0f,silicification ) that accounts for the structural difference between the ox-
ide in its stable standard form and as a formal entity present in the layered silicate (cf.
table 5.58).
Table 5.59 lists Gibbs free energies of formation from the elements of mica end-
members obtained with the procedure of Tardy and Garrels (1974). For comparative pur-
poses, the same table lists Gibbs free energies of formation from the elements derived from
the tabulated H 0f,Tr ,Pr and S 0Tr ,Pr values (same sources as in table 5.57) by application of
Table 5.59 Gibbs free energies of formation from the elements of micaceous components
according to method of Tardy and Garrels (1974), compared with tabulated values. References as
in table 5.57. Data in kJ/mole.
G 0ƒ,obs
G 0f , Tr , Pr = H 0f , Tr , Pr − Tr S 0f , Tr , Pr , ( 5.153 )
where ∑iviS 0i,Tr ,Pr is the sum of standard state entropies of the elements at stable state
(Robie et al., 1978), multiplied by their stoichiometric factors.
Table 5.60 Volumetric [WV ; J/(bar ⫻ mole)] enthalpic (WH ; kJ/mole), and entropic [WS ; J/(mole ⫻
K] terms of subregular Margules model for (Na,K)Al2Si3AlO10(OH)2 binary mixture, according to
various authors.
table 5.60. The same table also lists enthalpic and entropic subregular interaction
parameters for the same mixture according to Eugster et al. (1972) and Chatterjee
and Froese (1975).
Because excess volume terms are generally positive (with the exception of the
values indicated by Chatterjee and Froese, 1975 and Blencoe and Luth, 1973) and
because Margules parameters are also generally positive, the mixing model of
table 5.60 implies the existence of a solvus field in the binary range, whose exten-
sion should widen with increasing pressure [at T ⫽ 500 °C and Ptotal ⫽ P H2O ⫽
4 kbar, the solvus field should cover the range 0.1 ⬍ KAl2AlSi3O10 (OH)2 ⬍ 0.85;
Chatterjee and Froese, 1975].
According to Wones (1972) and Mueller (1972), the mixture KMg3AlSi3O10
(OH)2-KFe3AlSi3O10 (OH)2 (phlogopite-annite) is virtually ideal.
Munoz and Ludington (1974, 1977) experimentally measured the ionic ex-
change reaction
OH −mica + HF ⇔ Fmica
−
+ H2 O ( 5.154 )
( )
KMg 3 AlSi 3 O 10 OH, F( ) 2
→ log K 154 = 2100T −1 + 1.52 (5.155)
( ) (
KFe 3 AlSi 3 O 10 OH, F ) 2
→ log K 154 = 2100T −1 + 0.41 (5.156)
Silicates / 345
( )( )
K Fe 22 + Al Al 2 Si 2 O 10 OH, F ( ) 2
→ log K 154 = 2100T −1 + 0.20 (5.157)
( )
KAl 2 AlSi 3 O 10 OH, F ( ) 2
→ log K 154 = 2100T −1 − 0.11. (5.158)
(
KMg3 AlSi3 O10 OH) ( ) 2
+ ( ) ()
KFe 3 AlSi3 O10 F ⇔
2
(
KMg3 AlSi3 O10 F) ()
2
+ ( ) ( )
KFe3 AlSi3 O10 OH .
2
OH- phlogopite F-annite F- phlogopite OH-annite
(5.159)
The constant of equilibrium 5.159 is given by the ratio of the constant relative to
exchanges 5.155 and 5.156—i.e. in logarithmic notation,
Hinrichsen and Schurmann (1971) and Franz et al. (1977) detected the existence
of a solvus field between paragonite [NaAl2(Si3Al)O10 (OH)2 ] and margarite
[CaAl2(Si2Al2 )O10 (OH)2 ] by X-ray diffractometry and IR spectroscopy. Mea-
surements in the pressure range 1 ⱕ Ptotal (kbar) ⱕ 6 show a two-phase region
that covers the compositional field 0.20 ⬍ XCaAl2( Si2Al2 )O10( OH )2 ⬍ 0.55 at T ⫽
400 °C. The mixture appears to be homogeneous at T ⬎ 600 °C.
5.7 Feldspars
Figure 5.53 Idealized feldspar structure (topologic symmetry). (A) Projection from axis
a, showing conformation of four-member rings composed of two nonequivalent upward-
directed T1-T2 tetrahedra and two nonequivalent T1-T2 downward-directed tetrahedra.
(B) Double “gooseneck” chain seen along axis a. Periodicity along axis a is 4 (about 8.4
Å).
Mean
Element Concentration Notes
Table 5.62 Chemical analyses of some natural feldspars (Deer et al., 1983). (1) orthoclase
(Mogok, Burma); (2) Adularia (St. Gottard, Switzerland); (3) anorthoclase: inclusion in
augite (Euganean Hills NE Italy); (4) sanidine from a nepheline-leucitite; (5) albite from a
pegmatite; (6) anorthite from a calc-silicatic rock.
Sample
the energy of activation of the diffusive process (which embodies reticular distor-
tion effects) is quite high as a result of the rigidity of the structure. Low-T persis-
tency of metastable forms with intermediate intracrystalline disorder (adularia
and orthose) is thus often observed.
In NaAlSi3O8 , because the ionic size of Na⫹ is much smaller than that of
⫹
K , reticular distortion is implicit in the structure for simple topologic reasons,
and migration of tetrahedral ions is energetically easier, not implying symmetry
350 / Geochemistry of Crystal Phases
Figure 5.54 (A) Cationic occupancies in tetrahedral positions in case of complete disor-
der (monoclinic structure; upper drawing) and complete order (triclinic structure; lower
drawing). (B) Condition of complete order in microcline and low albite with Al:Si ⫽ 1:3,
compared with cationic ordering in anorthite (Al:Si ⫽ 2:2). Note doubling of edge c in an-
orthite.
modifications (in other words, the energy of activation of the migration process
is lower). “High albite” and “low albite,” both crystallizing in the triclinic system,
represent the two opposite arrangements of maximum disorder and maximum
order, respectively. The high-T monoclinic form “monalbite” (isostructural with
sanidine) is stable only in the vicinity of the melting point (i.e., 980 to 1118 °C at
P ⫽ 1 bar).
In tectosilicates, the Al-O-Al bonds are energetically unstable with respect to
the Al-O-Si arrangement, which better conforms to Pauling’s electrostatic valence
principle. It is essentially this fact that leads to the alternate ordering of silicon
and aluminum in Al:Si⫽1:1 feldspars (“aluminum avoidance principle”; Loew-
enstein, 1954). Thus, in anorthite (CaAl2Si2O8 ), because Al:Si⫽1:1, at all T con-
ditions we observe a natural invariant ordering of [AlO4 ] tetrahedral groups
alternating with [SiO4 ] groups. As shown in figure 5.54, this alternation implies
doubling of cell edge c.
2−
[L Si − O − Si] + [L Al − O − Al] [ ]
1−
⇔ 2 L Si − O − Al , ( 5.163 )
Silicates / 351
the gain in energy falls between ⫺36 and ⫺100 kJ per mole of substance, depending on
the type of nontetrahedral coordinating cation. This value agrees quite well with two-body
and three-body interionic potential lattice energy calculations for zeolites (⫺40 kJ/mole;
Bell et al., 1992) and with more recent ab initio SCF estimates (Tossell, 1993; ⫺40
kJ/mole)
2
a≈ + 3 l (5.164)
3
352 / Geochemistry of Crystal Phases
Table 5.63 Structural data of main feldspar polymorphs (from Smyth and Bish, 1988).
2
b = 3 + 2 l
3
(5.165)
(
c = 1+ 3 l ) (5.166)
(
β = arcsin 1 + 2 ) al (5.167)
(Ribbe, 1983a), where l varies between 2.62 Å ( [SiO4 ]4⫺ tetrahedra) and 2.88 Å
( [AlO4 ]5⫺ tetrahedra).
In feldspars with Al:Si ⫽ 1:3 stoichiometry, the length of the tetrahedral edge
is 2.8 Å. Introducing this value into the equations above gives a ⫽ 8.1, b ⫽ 13.0,
c ⫽ 7.62 Å, and  ⫽ 123°. Comparison with experimental values shows good
agreement for edge b, but the approximation is too rough for edges a and b and
the angle  to propose equations 5.164 to 5.167 as a determinative method.
Concerning the effect of P and T intensive variables on feldspar structure, we
must distinguish structural modifications associated with intracrystalline disorder
from pure displacive effects associated with compression and thermal expansion.
It may generally be stated that the triclinic structure of low-T feldspars has a
sufficient degree of freedom to modify interionic distances without incurring
polymorphic transitions—i.e., the (relatively inert) [TO4 ] groups are free to sat-
isfy the polyhedral thermal expansion and/or compressibility of M cations with-
out structural constraints. This fact results in the so-called “inverse relation” (i.e.,
structural modifications induced by cooling of the substance are similar to those
induced by compression; cf. Hazen and Finger, 1982).
Table 5.64 lists isobaric thermal expansion and isothermal compressibility
coefficients for feldspars. Due to the clear discrepancies existing among the vari-
ous sources, values have been arbitrarily rounded off to the first decimal place.
We may envisage the almost linear variation of thermal expansion and com-
pressibility with the amount of anorthite in the plagioclase mixture. We also see
that the three polymorphs of the KAlSi3O8 component have substantially similar
compressibilities, within uncertainties.
Table 5.64 Isobaric thermal expansion (K⫺1) and isothermal compressibility (bar⫺1)
of feldspars
variation between monalbite and low albite shows that the enthalpy change
associated with displacive phenomena is around 0.9 kcal/mole at T of transition
(Ttransition ⫽ 965 °C ⫽ 1338 K; Thompson et al., 1974). According to Thompson
et al. (1974), the intracrystalline disorder of the NaAlSi3O8 end-member in tri-
clinic form may be expressed by an ordering parameter Qod , directly obtainable
through
( ) (
Q od = X Al,T1o + X Al,T1m − X Al,T 2 o + X Al,T 2m ) (5.168)
= −12.523 − 3.4065 b + 7.9454c ,
where X Al,T1o is the atomic fraction of Al 3⫹ in T1o site and b and c are cell edges.
Equation 5.168 may be reconducted to equations 5.165 and 5.166. According to
Helgeson et al. (1978), displacive disorder is virtually absent at T ⬍ 350 °C; more-
over, the enthalpy variation connected with the displacive process ( ⌬ H 0di ) may be
described as a function of ordering parameter Qdi (it is assumed that Qdi ⫽ 0.06
at T ⫽ 623 K), according to
[ (
∆H di0 = 2.47 − 2.63 Q di − 0.06 )] (T − 623) (cal mole ) . (5.169)
Silicates / 355
∆H od
0
= H Q0
od
− H Q0
od =1
(
= 2630 1 − Q od ) ( cal mole) . (5.170)
∆H trans
0
= ∆H di0 + ∆H od
0
≈ 2.6 + 0.9 (kcal mole) . ( 5.171)
Moreover, because
∂ ∆H od
0
∆C P0r , od = ( 5.172 )
∂T P
r
and
∂ ∆H di0
∆C P0r , di = , ( 5.173 )
∂T P
r
equation 5.171 must be consistent with the similar equation for the heat capacity function:
Nevertheless, we have already noted that the transition of NaAlSi3O8 from monoclinic to
triclinic form may be regarded as the overlap of two high-order transitions, and that high-
order transition implies continuity in S, H, and G at the transition point and discontinuity
in the CP function (see section 2.8). The thermodynamic parameters for the two poly-
morphs on the opposite sides of the transition zone must thus be constrained to fulfill the
above requirements. The method followed by Helgeson et al. (1978) is as follows.
We thus obtain thermodynamic parameters of a generic albite phase that, with T, progres-
sively modifies its substitutional and displacive states from low-T to high-T forms.
Because for T ⬍ 473 K, ⌬C 0Pr ,trans ⫽ 0, Helgeson et al. (1978) adopt in the T range
298.15 to 473 K the original CP function tabulated by Kelley (1960) valid for low albite.
The albite CP function is described, starting from 473 K, by a second set of coefficients
(cf. table 5.65).
∆G trans =
1
2
( ) 1
4
1
( )
a di T − Tc , di Q di2 + B di Q di4 + aod T − Tc ,od Qod
2
2
(5.175)
1 1
+ Bod Qod4 + C od Qod
6
+ λQ di Qod .
4 6
Coefficients a, B, and C in equation 5.175 have the usual meanings in the Landau
expansion (see section 2.8.1) and for the (second-order) displacive transition of
albite assume the values adi ⫽ 1.309 cal/(mole ⫻ K) and Bdi ⫽ 1.638 kcal/mole
(Salje et al., 1985). Tc,di is the critical temperature of transition (Tc,di ⫽ B/a ⫽
1251 K). The corresponding coefficients of the ordering process are aod ⫽ 9.947
cal/(mole ⫻ K), Bod ⫽ ⫺2.233 kcal/mole, Cod ⫽ 10.42 kcal/(mole ⫻ K), and
Tc,od ⫽ 824.1 K. With all three coefficients being present in the Landau expansion
relative to substitutional disorder it is obvious that Salje et al. (1985) consider
this transition first-order. is a T-dependent coupling coefficient between displac-
ive and substitutional energy terms (Salje et al., 1985):
If ⭸ ⌬G/⭸x4 ⫽ 0 and ⭸ ⌬G/⭸x6 ⫽ 0, the strain components are linearly dependent on Qdi
and Qod (Salje, 1985):
Replacing the generalized strain with strain components x4 and x6 and adding the elastic
energy term in the Landau expansion results in equation 5.175.
As discussed by Salje et al. (1985) and also evident from equations 5.178.1 and
5.178.2, the two order parameters Qdi and Qod are not independent of each other (although
the displacive order parameter Qdi is largely determined by x4 whereas the substitutional
disorder Qod depends more markedly on x6 ). Their mutual dependence is outlined in fig-
ure 5.55.
Note that, displacive disorder being associated with a first-order transition, Qdi exhib-
its a jump at Ttrans whose magnitude decreases from fully ordered albite (i.e., Qod ⫽ 1) to
analbite (i.e., Qod ⫽ 0) (lower part of figure 5.5; see also section 2.8.1). Moreover, the
order-disorder phase transition is stepwise for Qdi ⫽ 0 (front curve in upper part of figure
5.5) and becomes smooth for higher Qdi values.
( )
Q od = 2 X Al,T1 − X Al,T 2 = 7.6344 − 4.3584 b + 6.8615c , (5.179)
where XAl,T1 and XAl,T2 are the atomic fractions of Al 3⫹ on T1 and T2 sites.
Note that the ordering parameter adopted by Hovis (1974) differs from the
form adopted for the triclinic phase of the NaAlSi3O8 end-member (eq. 5.168).
Helgeson et al. (1978) have shown that the enthalpy of disordering in sanidine
is similar to that observed in albite—i.e.,
∆H od
0
(
= 2650 1 − Q od ) (cal mole ) . (5.180)
However, the structural transition between the triclinic and monoclinic forms
(maximum microcline and high sanidine, respectively) “has no apparent effect on
its thermodynamic behavior” (cf. Helgeson et al., 1978); in other words, no gain
or loss of heat and no entropy variation is associated with the purely structural
modification. In this respect, the KAlSi3O8 end-member differs considerably from
NaAlSi3O8 , for which the contribution of structural disorder to bulk transforma-
tion (displacive plus substitutional) is about one-third in terms of energy. This
difference is evident if we compare the heat capacity functions of the two com-
pounds: a cusp is seen for albite in the triclinic-monoclinic transition zone (T ⫽
1238 K; cf. figure 5.56B), whereas no apparent discontinuity can be detected at
the transition point (T ⫽ 451 ⫾ 47 °C at P ⫽ 1 bar; Hovis, 1974) for KAlSi3O8
(figure 5.56A).
Also for KAlSi3O8 , it is convenient to account for phase transition effects
358 / Geochemistry of Crystal Phases
Figure 5.55 Mutual dependence of Qdi and Qod order parameters. In the upper part of
the figure is outlined the T dependence of substitutional disorder Qod for different values
of Qdi and, in the lower part, the T dependence of the displacive disorder parameter Qdi
for different values of Qod. The heavy lines on the surface of local curves represent the
solution for thermal equilibrium. From E. Salje and B. Kuscholke, Thermodynamics of
sodium feldspar II: experimental results and numerical calculations, Physics and Chemis-
try of Minerals, 12, 99–107, figures 5–8, copyright 䉷 1985 by Springer Verlag. Reprinted
with the permission of Springer-Verlag GmbH & Co. KG.
Figure 5.56 Maier-Kelley heat capacity functions for various structural forms of the
KAlSi3O8 (A) and NaAlSi3O8 (B) feldspar end-members (solid curves), compared with T
derivative of H 0T -H0Tr finite differences at various T (filled symbols) (from Helgeson et al.,
1978; redrawn and reproduced with permission).
second heat capacity function for the fictive form albite (valid for T ⬎ 473 K)
conforms quite well to experimental observations (H 0Pr ,T ⫺ H 0Pr ,473 finite differ-
ence derivatives) up to the T of triclinic-monoclinic transition (T ⫽ 1238 K).
It is generally assumed that CaAl2Si2O8 , based on the Al avoidance principle,
has complete ordering of Al and Si on tetrahedral sites at all temperatures. In-
deed, some authors suggest that ordering is not complete at high T (see, for in-
stance, Bruno et al., 1976). Moreover, anorthite at high T (⬎2000 K) in the disor-
dered state has I1 symmetry, unlike the P1 symmetry at low T (Bruno et al., 1976;
Chiari et al., 1978). According to Carpenter and MacConnell (1984), the P1 ⇔
I1 transition is of high order (as in the previous cases). Transition temperatures
between the P1 ⇔ I1 forms for pure anorthite at P ⫽ 1 bar are estimated by
Carpenter and Ferry (1984) to be around 2000 to 2250 K and (implicitly) the
authors consider structural transition as coincident with substitutional order ⇔
360 / Geochemistry of Crystal Phases
CP
NaAlSi3O8-KAlSi3O8 Mixture
Figure 5.58 shows subsolidus stability relationships in the NaAlSi3O8-KAlSi3O8
system at various Ptotal ⫽ Pfluid. At low P, the system is pseudobinary, because
KAlSi3O8 melts incongruently to form leucite (KAlSi2O6 ) plus liquid. However,
melting relations are not of interest here and will be treated to some extent later.
We note that, at subsolidus, an extended miscibility gap exists between Na-rich
(Absm ) and K-rich (Orsm ) terms. This gap also expands remarkably with increas-
ing P, due to the volume properties of NaAlSi3O8-KAlSi3O8 mixtures.
As a matter of fact, the volumes of NaAlSi3O8-KAlSi3O8 mixtures are not
ideal and have positive excess terms. Deviations from simple Vegard’s rule propor-
tionality are mainly due to edge b. Kroll and Ribbe (1983) provide three polyno-
mial expressions relating the volumes of mixtures to the molar fractions of the
potassic component (XOr ) valid, respectively, for monoclinic symmetry (complete
disorder), triclinic symmetry of perfect order (low albite and maximum micro-
cline), and the intermediate structural state:
( )
Ttrans oC = 978 − 19.2 X Or . ( 5.182 )
According to Merkel and Blencoe (1982), the relationship among the amount of
KAlSi3O8 , pressure, and temperature of transition is
Silicates / 363
Ttrans
XOr = 0.474 − 0.361 + 25.2Ptrans , ( 5.183)
1000
1. Macroperthites. When the unmixed K-rich and Na-rich zones may be observed
and characterized directly under the microscope;
2. Microperthites. When the existence of unmixing is anticipated under the micro-
scope by widespread turbidity or the presence of spots, but is clearly detectable
only by X-ray investigation;
3. Cryptoperthites. When unmixing is detected only by X-ray investigation.
Figure 5.59 Effects of structural and coherence state on extent of unmixing in the NaAl-
Si3O8-KAlSi3O8 binary join. (A) Maximum microcline–low albite, based on data from
Bachinski and Müller (1971) and Yund (1974). (B) Sanidine–high albite, based on data
from Thompson and Waldbaum (1969) and Sipling and Yund (1976).
taken entirely from Yund and Tullis (1983), lead us to a careful evaluation of the
real nature of perthitic unmixing in rocks and of the role of coherence energy in
determining feldspar phase relations. Figure 5.59A shows the approximate posi-
tion of the chemical solvus between maximum microcline and low albite ac-
cording to Bachinski and Müller (1971), compared with the coherent solvus de-
termined by Yund (1974). Figure 5.59B shows the miscibility gap between
sanidine and high albite and the corresponding coherent solvus, based on data
from Thompson and Waldbaum (1969) and Sipling and Yund (1976). Note that
the locations of the unmixing fields depend markedly on the structural state of
the components (compare parts A and B) and, structural state being equal, on
coherence energy. It is thus essential to determine both structure and coherence
of unmixed phases in order to assess the P and T conditions of unmixing.
Table 5.66 furnishes a synthesis of Margules subregular interaction parame-
ters according to various authors for the NaAlSi3O8-KAlSi3O8 join, collected by
Ganguly and Saxena (1987). The various sets of values show important differ-
ences that lead to consistent discrepancies in the calculated Gibbs free energy
values of the mixtures. This nonhomogeneity partly reflects the terminological
and operative confusion existing in the literature in the treatment of feldspar ther-
modynamics, and partly is the result of experimental artifacts.
NaAlSi3O8-CaAl2Si2O8 Mixture
Components NaAlSi3O8 and CaAl2Si2O8 show complete miscibility over a wide
range of P and T at subsolidus. Figure 5.60 shows the classical “lens diagram” of
phase stability of plagioclases. This conformation is indicative of ideality of mix-
ing in both liquid and solid states, and will be explained in detail in chapter 7. As
Silicates / 365
figure 5.60 shows, the structural form of mixtures at subsolidus is not unique
over the entire compositional range, but can be schematically subdivided into two
domains (P1 and I1) limited by a straight boundary between An59 at T ⫽ 1000
°C and An77 at T ⫽ 1400 °C (P ⫽ 1 bar; Carpenter and Ferry, 1984). If this
subdivision is adopted, the two-phase field (solid plus liquid) is also necessarily
subdivided into two domains: one constituted of P1 mixture plus liquid, and the
other of I1 mixture plus liquid, as shown in the figure.
Newton et al. (1980) calorimetrically measured the enthalpy of the NaAl-
Si3O8-CaAl2Si2O8 mixture at T ⫽ 970 K and P ⫽ 1 bar and found a positive
excess enthalpy reproduced by a subregular Margules model:
( )
H excess mixing = X 1X 2 WH ,12 X 2 + WH , 21X 1 , (5.184)
In anorthite, tetrahedral groups [TO4 ] lying parallel to the (123) plane exhibit perfect
alternation of Al-bearing and Si-bearing strata. Let us define NT as the total number of
tetrahedra occupied by Al and Si:
NT = NAl + N Si . ( 5.185)
Figure 5.60 X-T phase stability diagram for the NaAlSi3O8-CaAl2Si2O8 system at P ⫽ 1
bar, showing transition between forms P 1 and I 1 at subsolidus and within two-phase
field. Note that form of disorder phase is usually “mediated” into crystalline class C 1,
hence in the literature the phase transition is usually described as C 1 ⇔ I 1. From M. A.
Carpenter and J. M. Ferry, Constraints on the thermodynamic mixing properties of pla-
gioclase feldspars, Contributions to Mineralogy and Petrology, 87, 138–48, figure 3, copy-
right 䉷 1984 by Springer Verlag. Reprinted with the permission of Springer-Verlag
GmbH & Co. KG.
consider the exchange of one Si atom in a  site with one Al atom in an ␣ site: the exchange
is possible only when the resulting configuration does not result in Al-O-Al alternation.
Because the probability that an Si atom in a  site is surrounded by four other Si neighbors
is ( 12)4 ⫽ 16
1
, the total number of sites available for Al-Si mixing, based on the Al avoidance
principle, is
1
NAl ⇔ Si,α = NT ( 5.186 )
2
1 1
NAl ⇔ Si, β = × NT . ( 5.187 )
2 16
Therefore, in practice, mixing takes place over ␣ sites only. Applying this condition to the
stoichiometry of the binary mixture, we obtain
N Si,α = N Na ( 5.188 )
(cf. Kerrick and Darken, 1975). If we now recall the equation for permutability Q (section
3.8.1) and apply the calculation to the four atoms in the mixture (Si, Al, Na, Ca), because
Na and Ca mix over one M site and Si and Al mix essentially over ␣ sites, we obtain
Silicates / 367
Q=
(N Na + NCa ) !
×
(N Si,α + N Al ! )
N Na! NCa! N Si,α ! N Al !
[ 2( N ]
(5.191)
(N Na + NCa ) ! Ca + N Na ) !
= × .
N Na! NCa! N Na! ( 2 NCa + N Na )!
Application of Stirling’s formula to equation 5.191 and comparison with the configura-
tional state of pure components lead to the definition of a configurational entropy of
mixing term in the form
X (1 + X ) 2
. ( 5.192 )
S mixing conf [ 2
]
= − R X Ab ln X Ab (1 − X An ) + X An ln
An
4
An
In the absence of enthalpic contributions, the activities of the components in the mix-
ture become
a Ab = X Ab
2
( 2 − XAb ) ( 5.193 )
and
X An ( 1 + XAn ) .
1 2
a An = ( 5.194 )
4
(In eq. 5.193 and 5.194, Ab or An simply identifies the component without any implication
about its structural state.)
Figure 5.61B shows how the entropy of mixing derived by application of the
Al avoidance principle compares with the one-site ideal mixing contribution. Also
shown are entropy of mixing curves obtained by summing, up to the ideal one-
site contribution, additional terms related to phase transition effects [variable be-
tween 1 and 2.5 cal/(mole ⫻ K)] (Carpenter and Ferry, 1984). Figure 5.61C com-
pares the Gibbs free energy of mixing experimentally observed by Orville (1972)
at T ⫽ 700 °C and P ⫽ 1 bar with Gibbs free energy values obtained by combining
the entropy effect of the Al avoidance principle and the enthalpy of mixing experi-
mentally observed by Newton et al. (1980) (see also figure 5.61A):
( )
2
X2 1 + X2
G mixing = H mixing
[ (
+ RT X 1 ln X 12 2 − X 1 )] + X 2 ln
4
.
(5.195)
As figure 5.61C shows, the equation of Kerrick and Darken (1975), coupled with
the enthalpy of mixing terms of Newton et al. (1980), satisfactorily fits the experi-
368 / Geochemistry of Crystal Phases
Saxena and Ribbe (1972) have shown that the excess Gibbs free energy of
mixing of the mixture, based on the data of Orville (1972), may be reproduced
by a subregular Margules model:
( )
G excess mixing = X 1X 2 W12 X 2 + W21X 1 , (5.196)
KAlSi3O8-CaAl2Si2O8 Mixtures
The subsolidus properties of the KAlSi3O8-CaAl2Si2O8 system are essentially
those of a mechanical mixture (cf. section 7.1). According to Ghiorso (1984), the
miscibility gap is not influenced by pressure and may be described by a strongly
asymmetric Margules model:
( )
G excess mixing = X 1X 2 W12 X 2 + W21X 1 , (5.197)
where WH,12 ⫽ 16,125 cal/mole, WH,21 ⫽ 6688 cal/mole, WS,12 ⫽ 2.644 cal/(mole
⫻ K), and WS,21 ⫽ ⫺4.830 cal/(mole ⫻ K).
Figure 5.62 shows in detail the distribution of the NaAlSi3O8 component between
unmixed phases at T ⫽ 900 °C and P H2O ⫽ 0.5 kbar, as observed by Seck (1971)
and compared with Wohl’s and Kohler’s predictions. In both models, the binary
interaction parameters are those deduced by Saxena (1973) on the basis of Or-
ville’s (1972) data for the NaAlSi3O8-CaAl2Si2O8 system, coupled with the experi-
mental observations of Seck (1971).
A more recent model (Ghiorso, 1984) is based on the binary interaction pa-
rameters of Thompson and Hovis (1979) for the NaAlSi3O8-KAlSi3O8 join and
on the experimental results of Newton et al. (1980), coupled with the Al avoid-
ance principle of Kerrick and Darken (1975) extended to the ternary field. Ghi-
orso (1984) expressed the excess Gibbs free energy of mixing in the form
1 1
G excess mixing = W12 X 1X 2 X 2 + X 3 + W21X 1X 2 X 1 + X 3
2 2
1 1
+ W13X 1X 3 X 3 + X 2 + W31X 1X 3 X 1 + X 2 (5.199)
2 2
1 1
+ W32 X 3X 2 X 2 + X 1 + W23X 3X 2 X 3 + X 1 + W123X 1X 2 X 3 ,
2 2
nents in the unmixed phases of Seck (1971). Table 5.67 reports the resulting inter-
action parameters.
Silica has 22 polymorphs, although only some of them are of geochemical inter-
est—namely, the crystalline polymorphs quartz, tridymite, cristobalite, coesite,
and stishovite (in their structural modifications of low and high T, usually desig-
nated, respectively, as ␣ and  forms) and the amorphous phases chalcedony
and opal (hydrated amorphous silica). The crystalline polymorphs of silica are
tectosilicates (dimensionality ⫽ 3). Table 5.68 reports their structural properties,
after the synthesis of Smyth and Bish (1988). Note that the number of formula
units per unit cell varies conspicuously from phase to phase. Also noteworthy is
the high density of the stishovite polymorph.
Quartz is a primary phase in acidic igneous rocks (both intrusive and effusive) and
in metamorphic and sedimentary rocks. Tridymite is a high-T form found in porosities of
rapidly quenched effusive rocks (mainly andesites and trachites). Cristobalite is rarely
found in volcanic rocks and sometimes occurs in opal gems. Coesite is a marker of high-
P regimes. It has been found coexisting with garnet in the Dora-Maira Massif (Western
Alps) and in various other high-P metamorphic terrains (the former Soviet Union,
China). Stishovite is stable above about 80 kbar (cf. Liu and Bassett, 1986, and references
therein) and is found exclusively in “impactites” (i.e., rocks formed by meteorite impact).
Chalcedony and opal are amorphous phases generated by hydrothermal deposition from
SiO2-saturated solutions.
Figure 5.63A shows the relative positions of Si atoms in the ␣-quartz structure
projected along axis z. In this simplified representation, note the hexagonal ar-
rangement of Si atoms, which can be internally occupied by univalent cations
through coupled substitution involving Al 3⫹ in tetrahedral position:
372 / Geochemistry of Crystal Phases
Table 5.68 Structural properties of SiO2 polymorphs (from Smyth and Bish, 1988).
Figure 5.63 Quartz structure projected along axis z. Filled dots are Si atoms (oxygen
atoms not shown). ␣-quartz (A) - -quartz (B) transition takes place by simple rotation
of tetrahedra on ternary helicogyre (adapted from Gottardi, 1972).
(the hexagonal ring is actually a spiral with atoms positioned on separate planes).
Figure 5.63B shows the structure of high-T -quartz. The ␣-quartz–-quartz
transition, which may be described as the overlap of a first-order and a transi-
tion, takes place by simple rotation of tetrahedra on the ternary helicogyre.
Silicates / 373
Figure 5.64 Effect of cooling rate on SiO2 polymorphic transitions. From Putnis and
McConnell (1980). Reproduced with modifications by permission of Blackwell Scientific
Publications, Oxford, Great Britain.
We have already stated that the ␣- transition of quartz may be described as a
transition overlapping a first-order transition. The heat capacity function for the two poly-
morphs is thus different in the two stability fields, and discontinuities are observed in the
H and S values of the phase at transition temperature (Ttrans cf. section 2.8). For instance,
to calculate the thermodynamic properties of -quartz at T ⫽ 1000 K and P ⫽ 1 bar, we
374 / Geochemistry of Crystal Phases
Table 5.69 Thermal expansion at P ⫽ 1 bar (K-1) and bulk modulus at T ⫽ 25°C (Mbar) for some
SiO2 polymorphs in the databases of Saxena et al. (1993) (1) and Holland and Powell (1990) (2).
Thermal expansion at various T values of interest is obtained by applying
␣(T) ⫽ ␣0 ⫹ ␣1T ⫹ ␣2T ⫺1 ⫹ ␣3T ⫺2
Table 5.70 Thermodynamic data for silica polymorphs. Stishovite and tridymite
from Saxena et al. (1993); remaining polymorphs from Helgeson et al. (1978).
H 0f, Tr , P r in kJ/mole; S 0Tr , P r in J/(mole ⫻ K). Heat capacity function is
C P ⫽ K 1 ⫹ K 2 T ⫹ K 3T ⫺2 ⫹ K 4T 2 ⫹ K 5T ⫺3 ⫹ K 6 T ⫺1/2 ⫹ K 7T ⫺1.
Table 5.71 Transition energy terms for quartz and cristobalite: ⌬S 0trans ⫽ molar
apparent transition entropy at P ⫽ 1 bar [ J/(mole ⫻ K)]; ⌬H 0trans ⫽ molar
apparent transition enthalpy at P ⫽ 1 bar (kJ/mole); ⌬V 0trans ⫽ molar apparent
transition volume at P ⫽ 1 bar (cm3/mole).
proceed as follows: calculate enthalpy and entropy at Ttrans ⫽ 848 K with the aid of the
heat capacity expression valid for ␣-quartz (integral between Tr ⫽ 298.15 K and T ⫽
Ttrans ⫽ 848 K); to the obtained values add the apparent standard molar ␣- transition
enthalpy and entropy at Ttrans; calculate the integrals between Ttrans and T ⫽ 1000 K with
the CP function valid for the -phase and add the results to the previously obtained en-
thalpy and entropy values. Note that, at the transition point, the two polymorphs coexist
and thus have the same Gibbs free energy. Table 5.71 lists ␣- transition enthalpy, entropy,
and volume values for quartz and cristobalite. These values are consistent with the
␣-phase of Helgeson et al. (1978). An extensive discussion on the energetics of SiO2 poly-
morphs may be found in Helgeson et al. (1978) and Berman (1988).
A α + B β ⇔ A β + Bα . ( 5.201)
376 / Geochemistry of Crystal Phases
The Gibbs free energy change involved in reaction 5.201 is given by the sum of the
chemical potentials of components in reaction ( i ), multiplied by their respective
stoichiometric factors (i )—i.e.,
∆G 201 = ∑ ν i µi . (5.202)
i
In the case of equation 5.201, stoichiometric factors are 1 for all components;
equation 5.202 may thus be reduced to
∆G 201 = ∑ µi . (5.203)
i
µ = µ i0 + RT ln a i (5.204)
is valid, where 0i is the standard state chemical potential and ai is thermody-
namic activity.
Applying equation 5.203 to reaction 5.201 and recalling equation 5.204, we
obtain
∆G 201 = ∆G 201
0
+ RT ln K 201, P,T , (5.205)
where ⌬G 0201 is the Gibbs free energy change involved in reaction 5.201 at the
standard state of reference condition and K201,P,T is the equilibrium constant at
the P and T of interest. Because, at equilibrium, ⌬G of the reaction is zero, we
may rewrite equation 5.205 as
∆G 201
0
= − RT ln K 201, P,T . (5.206)
∂ ln K
= −
(
∂ ∆G 0 RT ) =−
∆V 0
. (5.207)
∂P T ∂P RT
T
∂ ln K
= 0. (5.208)
∂P T
∂ ln K ∂ ∆G 0 RT ( ) ∆H 0
= − = (5.209)
∂T P ∂T RT 2
P
∂ ln K ∆H 0
=− . (5.210)
( )
∂ 1 T
P
R
However, if the standard state is that of the pure component at P ⫽ 1 bar and T
of interest,
∂ ln K ∆H 0
= ( 5.211)
∂T RT 2
not dependent on P.
If we now consider the equilibrium constant expressed as activity product,
then
K 201, P,T = ∏ a iν i
. (5.212)
i
In the case of reaction 5.201, stoichiometric coefficients are 1 and equation 5.212
may be reduced to
The first term in parentheses on the right side of equation 5.213 is the distribution
coefficient (KD ), and the second groups activity coefficients related to the mixing
behavior of components in the two phases. The equilibrium constant is thus re-
lated to the interaction parameters of the two phases at equilibrium. For example,
the equilibrium between two regular mixtures is defined as
378 / Geochemistry of Crystal Phases
Wβ
ln K 201,P ,T = ln K D −
Wα
RT
(
1 − 2 X A ,α +
RT
)
1 − 2 X A ,β , ( ) (5.214 )
where W␣ and W are interaction parameters for phases ␣ and , respectively, or,
for two asymmetric Van Laar mixtures,
A0,α A1,α
ln K201,P ,T = ln K D +
RT
(X A, α − XB, α + ) RT
(6 XB, α )
X A,α − 1
(5.215)
A0, β A1, β
+
RT
(X
B, β − X A, β + ) RT
(6X A, β XB, β − 1 . )
where A0,␣, A1,␣, A0, , and A1, are Guggenheim’s parameters for phases ␣ and ,
respectively (Saxena, 1973).
Let us now examine two natural phases, ␣ and , coexisting at equilibrium,
for which we know in detail both mixing behavior (hence W␣ and W , or A0,␣,
A1,␣, A0, , and A1, ) and composition (molar fractions of AM, BM, AN, and
BN components, or the analogous ionic fractions XA,␣, XA,, XB,␣, and XB, ).
Application of equation 5.206 allows us to define the loci of P and T conditions
that satisfy equilibrium, provided that in the definitions of both ⌬G 0 and the
thermodynamic activity of components in the mixture we are consistent with the
adopted standard state. It is obvious that the more ⌬G 0 is affected by P and T
intensive variables, and the more the interaction parameters describing the mixing
behavior are accurate, the more efficient and precise is the thermobarometric ex-
pression. By stating “the loci of P and T conditions that satisfy equilibrium,” we
emphasize that we are dealing with univariant equilibria. In a P and T space, the
slope of the univariant curve is defined by equations 5.207 and 5.209; if the vol-
ume of reaction is negligible, then obviously the equilibrium constant is not
affected by P, and the elemental exchange furnishes only thermometric infor-
mation.
Exchange geothermometry generally considers couples of elements of identi-
cal charge showing the property of diadochy—for instance, Na⫹-K⫹, Mg2⫹-Fe2⫹,
or Fe3⫹-Cr3⫹. Intercrystalline exchange thermometers have been developed
mainly on the Mg2⫹-Fe2⫹ couple, and the investigated mineral phases comprise
olivine, orthopyroxene, clinopyroxene, garnet, spinel, ilmenite, cordierite, biotite,
hornblende, and cummingtonite. Table 5.72 lists bibliographic sources for these
geothermometers. More detailed references may be found in Essene (1982) and
Perchuk (1991).
Let us consider, for instance, the Fe2⫹-Mg2⫹ exchange reaction between garnet and
clinopyroxene:
1 1
Mg3Al2Si 3O12 + CaFeSi2O6 ⇔ Fe 3Al2Si 3O12 CaMgSi2O6 .
3 pyrope hedembergite
3 almandine diopside
(5.216)
Silicates / 379
Equilibrium References
Olivine-orthopyroxene-spinel Ramberg and Devore (1951); Nafziger and Muan (1967);
Speidel and Osborn (1967); Medaris (1969); Nishizawa and
Akimoto (1973); Matsui and Nishizawa (1974); Fujii (1977);
Engi (1978)
Ilmenite-orthopyroxene Bishop (1980)
Ilmenite-clinopyroxene Bishop (1980)
Orthopyroxene-clinopyroxene Mori (1977); Herzberg (1978b)
Garnet-Olivine Kawasaki and Matsui (1977); O’Neill and Wood (1979)
Garnet-Cordierite Currie (1971); Hensen and Green (1973); Thompson (1976);
Hensen (1977); Holdaway and Lee (1977); Perchuk (1991)
Garnet-clinopyroxene Saxena (1979); Banno (1970); Oka and Matsumoto (1974);
Irving (1974); Raheim and Green (1974, 1975); Raheim (1975,
1976); Mori and Green (1978); Slavinskiy (1976); Ellis and
Green (1979); Ganguly (1978); Saxena (1979); Dahl (1980);
Perchuk (1991); Krog (1988).
Garnet-biotite Saxena (1969); Thompson (1976); Perchuk (1977, 1991);
Goldman and Albee (1977); Ferry and Spear (1978).
Garnet-chlorite Ghent et al. (1987); Perchuk (1991)
Garnet-staurolite Perchuk (1991) and references therein
Garnet-amphibole Perchuk (1991) and references therein
Garnet-chloritoid Perchuk (1991) and references therein
Biotite-chloritoid Perchuk (1991) and references therein
Chlorite-chloritoid Perchuk (1991) and references therein
XFe,garnet XMg,cpx
KD = , (5.217)
XMg,garnet XFe,cpx
( )
3
aMg 3Al 2Si 3O12 ≡ X Mgγ Mg , ( 5.218)
garnet
where ␥Mg is the activity coefficient of ion Mg2⫹ in the structural site and 3 is the stoichio-
metric number of sites over which substitution takes place. Analogously, for clinopy-
roxene,
380 / Geochemistry of Crystal Phases
[( ]
13
) (X ) (X ) (γ )
3
X Feγ Fe Mg γ Mg Fe X Mg Fe γ Mg
garnet
K216 = × = × = KD K γ
cpx garnet garnet
(X ) (X ) (γ )
[( ] Feγ Fe γ Mg
13
) X Mg
3
X Mg γ Mg cpx Fe cpx Fe cpx
garnet
( 5.220 )
ln K216 = ln KD + ln Kγ . ( 5.221)
As figure 5.65 shows, the term ln K␥ definitely affects the equilibrium. This term is consis-
tent with a subregular Margules formulation for excess Gibbs free energy terms in both
phases (Ganguly, 1979; see also sections 5.3.5 and 5.4.5). Combining the various experi-
mental evidence, Ganguly (1979) calibrated the distributive function of equation 5.217
over T and P, and proposed the following thermometric expression:
4100 + 11.07P
ln KD = + c, ( 5.222 )
T
−Ea
D( T ) = A exp , ( 5.224 )
RT
382 / Geochemistry of Crystal Phases
Figure 5.66 Elemental concentration profiles in a thermometric couple. (A) Ideal case;
elemental concentration is constant from interface to nucleus of crystal. (B) Nonideal
case: slow diffusivity generates concentration gradients from interface to nucleus of
crystal.
where A is the preexponential factor and Ea is the energy of activation of the diffusive
process.
Diffusivity at time t may be related to initial diffusivity at time t0 by associating with
t the corresponding temperature T:
E 1 1
D( t ) = D(0t 0 ) exp a − 0 . ( 5.225)
R T( t ) T( t 0 )
T( t ) ≈ T( t0 ) − v t , ( 5.226 )
0
where
Ea v
γ = 2
. ( 5.228 )
RT( t0 )
0
Parameter ␥ is extremely important, because it defines the time within which elemen-
tal exchanges proceed, beginning at the initial time t 0 when temperature was T 0 . If, for
instance, the energy of activation of the diffusion process is 50 kcal/mole, the cooling rate
Silicates / 383
of the system is 2 °C/106 years, and T 0 ⫽ 1000 K, the resulting ␥ is (0.05 ⫻ 10⫺6 ) years⫺1.
The time elapsed from the initial condition (T 0 , t 0 ) to the (kinetic) closure of elemental
exchanges is
t′ =
1
γ
[
1 − exp( −γ t) ≤
1
γ
]
≤ 2 × 10 7 years (5.229)
(Lasaga, 1983).
As regards the effect of temperature on the distribution constant, if both phases are
ideal, the distribution constant is equivalent to the thermodynamic constant, and we can
write
∆H 0 1 1
K D = K D 0 exp − 0 , ( 5.230 )
R T T
where ⌬H 0 is the enthalpy of the exchange reaction at standard state and KD0 is the distri-
bution constant at the initial temperature T 0 . We can define KD as a function of time, as
was done previously for diffusivity (cf. equation 5.227):
KD = K D 0 exp( − ε ′t ) , ( 5.231)
(t)
where
∆H 0 v
ε′ = 2
. ( 5.232 )
RT 0
( t0 )
Lasaga (1979) has shown that the molar concentration of an exchanging element A
at the interface between two phases ␣ and  at time t (XA,␣,0,t ) obeys the law
X A, α , 0, t = X A0 , α exp( − ε t) . ( 5.233 )
Applying Fick’s law on diffusive fluxes at the interface between the two phases (cf. sec-
tion 4.11):
∂X A, α ∂X A, β
− D˜ α , t = − D˜ β , t ( 5.234 )
∂χ ∂χ
and
∂X B , α ∂X B , β
− D˜ α , t = − D˜ β , t , ( 5.235)
∂χ ∂χ
384 / Geochemistry of Crystal Phases
where is the distance from the interface, and D̃␣,t and D̃,t are interdiffusion coefficients
for elements A and B in phases ␣ and , Lasaga (1983) has shown that is related to
⬘ through
D˜ β0 D˜ α0
ε = ε′ ,
X A0 , α X A0 , α X A0 , α (5.236)
D˜ β0 D˜ α0 1 + X 0 + X 0
X 0
B, α A, β B, β
ε ≈ ε′ ( 5.237 )
and
ε ∆H 0
≈ . ( 5.238 )
γ Eα
ε = ε ′ D˜ β0 D˜ α0 ( 5.239 )
and
ε
≈ ∆ H 0. ( 5.240 )
γ
Analysis of existing data shows that ratio / ␥ in the main thermometric phases ranges
between 0.01 and 0.2. The highest / ␥ values correspond to minerals with high ionic
diffusivity. Knowledge of ratio / ␥ in phases used in thermometric studies allows us to
establish the reliability of the obtained temperature, as a function of the grain size of the
phases and the cooling rate of the system. Besides ratio / ␥, Lasaga (1983) also defined
quantity ␥⬘ as
E α va 2
γ′= , ( 5.241)
D˜ α0 RT (0t )
2
where D0␣ is the interdiffusion coefficient at temperature T 0 in phase ␣ and a is the length
of the crystal orthogonal to the interface. ␥⬘ allows us to establish whether or not the
nuclei of crystals undergoing equilibration maintain the original compositions they had at
crystallization (t 0 , T 0 ). Practical calculations show that this happens for ␥⬘ ⬎ 10. Figure
5.67 shows the combined effects of parameter ␥⬘ and ratio / ␥: for low values of ␥⬘, the
Silicates / 385
Figure 5.67 Concentration profiles in a crystal, for various values of ␥⬘ and /␥. Abscissa:
fractional distance from crystal edge ( ⫽ 0 at edge; ⫽ 1 at center). Reproduced with
modifications from Lasaga (1983), with permission of Springer-Verlag, New York Inc.
initial composition of the nucleus is not preserved even in phases with slow diffusivity
(low / ␥ ). We also see that, at constant ␥⬘, the concentration profiles are more or less
altered with respect to the initial condition, depending on diffusion rate. Lasaga (1983)
emphasized the fact that the core of the crystal may be homogeneous at both high (i.e.,
⬎ 100) and low (⬍ 0.01) values of parameter ␥⬘. Hence, contrary to what is commonly
assumed, the homogeneity of crystal cores does not necessarily mean preservation of the
initial composition acquired at crystallization.
T (°C)
Garnet
6
v (°C/10 y) Pyroxene (a) (b) Olivine
0.01 805 565 785 -
0.1 865 610 840 -
1 935 665 907 315
10.0 1010 720 980 370
100.0 1100 790 1050 430
1000.0 1200 860 1155 500
10000.0 1320 945 1260 585
This is the case, for instance, for the enstatite-ferrosilite solid mixture [orthopy-
roxene: (Fe,Mg)SiO3 ]:
ln K 243 = ln K D +
WM1
RT
( )
W
(
1 − 2 X Fe,M1 − M 2 1 − 2 X Fe,M 2 .
RT
) (5.244 )
K 245 =
aMg,M3 aFe2 +,M1
=
(1 − X Fe,M 3 )X Fe,M1
= 0.403. (5.246 )
aMg,M1aFe2 +,M 3 (1 − X Fe,M1 )XFe,M 3
In the case of reaction 5.243, Saxena and Ghose (1971) experimentally deter-
mined the distribution constant over a wide range of temperatures. According to
these authors, it has an exponential dependency on T (K) of the type
10 3
− RT ln K 243 = ∆G 243
0
= 4479 − 1948
T
(cal mole ) . (5.247)
K 243 = K D (5.248)
10 3
WM1 = 3525
T
− 1667 (cal mole ) (5.249)
and
10 3
WM2 = 2458
T
− 1261 ( cal mole ) . ( 5.250 )
The form of the distribution curves, based on equation 5.244 and application of
conditions 5.247, 5.249, and 5.250, is displayed in figure 5.68.
More recently, Ganguly and Saxena (1987) reconsidered equation 5.243 in
the light of all existing experimental data, proposing
1561.81 WM 1
ln K D = 0.1435 −
T
+
RT
1 − 2 X Fe,M 1 ( )
(5.251)
W
(
− M 2 1 − 2 X Fe,M 2 ,
RT
)
388 / Geochemistry of Crystal Phases
with WM1 ⫽ 1524 (cal/mole) and WM2 ⫽ ⫺1080 (cal/mole). Note that the first
term in parentheses in equation 5.251 corresponds to ln K243 in equation 5.244.
The diagrammatic form of figure 5.68 is that commonly adopted to display
intracrystalline distributions (see also figures 5.39 and 5.40). However, this sort
of plot has the disadvantage of losing definition as the compositional limits of
the system are approached. A different representation of intracrystalline disorder
is that seen for olivines (figures 5.10 and 5.12; section 5.2.5): the distribution
constant is plotted against the molar fraction of one of the components in the
mixture.
The effects of kinetics on intracrystalline disorder are less marked than those
affecting intercrystalline exchanges, essentially because intracrystalline exchanges
take place on the Å scale, whereas intercrystalline equilibration requires displace-
ments on the mm scale. As a result, temperatures registered by intracrystalline
exchange thermometers are generally lower than those derived from intercrystal-
line equilibria. Lastly, it must be noted that the effects of pressure on intracrystal-
line partitioning are not negligible, as generally assumed, because the volume
properties of phases are definitely affected by the state of internal disorder.
(and structures) of unmixed phases. The various solvi are located experimentally
by physical observation of the mixing-unmixing phenomenon induced on a fixed
composition by modifying the T and P. For the sake of accuracy, the experiment
is usually “reversed,” the miscibility gap limb being approached from opposite
directions in the T-P space.
The thermodynamic significance of unmixing was stressed in chapter 3 (sec-
tions 3.11, 3.12, and 3.13), together with the intrinsic differences among solvus
field, spinodal field, and miscibility gap, and the role of elastic strain on the T-X
extension of the coherent solvus and coherent spinodal field with respect to chem-
ical solvus and chemical spinodal fields, respectively. Although the significance
of the various compositional fields is clear in a thermodynamic sense, very often
the stability relations of unmixed phases do not seem as clear in microscopic
investigation of natural assemblages. The confusion in terms used to describe
unmixing phenomena in natural phases was discussed in some detail in section
5.7.4.
Feldspars
Feldspars, as we have already seen, are mixtures of albite (NaAlSi3O8 ),
K-feldspar (KAlSi3O8 ), and anorthite (CaAl2Si2O8 ) components (see section
5.7). Albite and anorthite are miscible to all extents (plagioclases), as are albite
and K-feldspar (alkali feldspars), but the reciprocal miscibility between the two
series is limited for T ⬍ 800 °C. Stormer (1975) calculated the distribution of the
NaAlSi3O8 component between plagioclases and alkali feldspars, accounting for
the nonideality of the NaAlSi3O8-KAlSi3O8 solid mixture (Parson, 1978) and for
P effects on the solvus field. Parson’s (1978) model reproduces quite well the ex-
perimental observations of Seck (1971) at T ⫽ 650 °C for anorthite amounts
between 5% and 45% in moles. Modifications of Stormer’s (1975) model were
suggested by Powell and Powell (1977a). Whitney and Stormer (1977) also pro-
posed a low-T geothermometer based on K-feldspar amounts in mixtures. The
isothermal distributions of Stormer’s (1975) model at various P conditions are
compared with the experimental evidence of Seck (1972) in figure 5.69. When
applying this sort of thermometer, we must bear clearly in mind the significance
of feldspar unmixes in the various assemblages, particularly with respect to the
degree of coherence of unmixed zones. Stormer’s geothermometer defines the loci
of chemical solvi (i.e., absence of strain energy) and so cannot be applied to coher-
ent exsolution or to spinodal decomposition. It is also worth noting that the alkali
feldspar series unmixes itself partially during cooling, so that the compositions of
unmixed grains must be reintegrated in order to evaluate the original composition
of the solid mixture on the solvus limb. Application of Stormer’s (1975) geother-
mometer may simply be graphical [plotting the molar amounts of NaAlSi3O8 in
the two series, for a selected value of P (figure 5.69) and comparing the results
with the isotherms] or analytical, through the appropriate thermometric
equation.
390 / Geochemistry of Crystal Phases
Figure 5.69 Feldspar geothermometry. From Stormer (1975). Reprinted with permission
of The Mineralogical Society of America.
Stormer’s (1975) model is based on the assumption that the potassic end-member has
no influence on the mixing behavior of the plagioclase series and that the calcic compo-
nent does not affect the mixing behavior of the K-feldspar series to any extent. With
these assumptions, the problem of equilibrium between two ternary feldspars, normally
represented by the equalities
µ NaAlSi = µ NaAlSi
0
+ RT ln a NaAlSi 3O 8, Pl ( 5.255)
3O 8, Pl 3O 8, Pl
and
µ NaAlSi = µ NaAlSi
0
+ RT ln a NaAlSi 3O 8 , Kf . ( 5.256 )
3O 8 , Kf 3O 8 , Kf
If we adopt for both phases the standard state of the pure component at the T and P of
interest (and if the two phases are strictly isostructural), because
µ NaAlSi
0
= µ NaAlSi
0
, ( 5.257 )
3O 8 , Pl 3O 8 , Kf
from equations 5.255 and 5.256, and by application of equation 5.252, we obtain
Note that the isoactivity condition of equation 5.258 is valid only for all the loci of the
solvus limb (see, for instance, figure 3.10) but not for a spinodal limit or for a miscibility
gap limb, because in these cases the structural state of unmixes is different, so that equa-
tion 5.257 does not hold.
If we call the distribution constant KD , as a ⫽ X ⭈ ␥, from equation 5.258 we derive
X NaAlSi 3O 8 , Kf γ NaAlSi 3O 8 , Pl
KD = = . ( 5.259 )
X NaAlSi 3O 8 , Pl γ NaAlSi 3O 8 , Kf
Moreover, if the plagioclase series is ideal (which is the case in the compositional range
Ab100 to Ab45 ; cf. Orville, 1972), equation 5.259 may be further reduced to
1
KD = . ( 5.260 )
γ NaAlSi 3O 8 , Kf
Adopting a subregular Margules model for the NaAlSi3O8-KAlSi3O8 (Ab-Or) binary mix-
ture and assuming that the activity coefficient of the albite component is not affected by
the presence of limited amounts of the third component in the mixture (i.e., CaAl2Si2O8 ),
equation 5.260 may be transformed into
ln K D = − ln γ Ab, Kf
(5.261)
1
(1 − X ) [W ( )] .
2
=− Ab, Kf Ab − Or + 2X Ab, Kf WOr − Ab − WAb − Or
RT
Using the interaction parameters of Thompson and Waldbaum (1969) (cf. second column
of parameters in table 5.66), i.e.,
and combining equations 5.262 and 5.263 with equation 5.260, we obtain the thermomet-
ric equation of Stormer (1975):
Ab , Kf
+ 0.0392 X
3
Ab , Kf
)P
+ .
X Ab , Kf
−1.9872 ln + 4.6321 − 10.815 X Ab , Kf + 7.7345 X Ab2 , Kf − 1.5512 X Ab3 , Kf
X Ab , Pl
Pyroxenes
As we have already seen in section 5.4.6, unmixing between monoclinic and or-
thorhombic pyroxenes is conveniently interpreted in the framework of the quadri-
lateral field diopside-hedembergite-enstatite-ferrosilite (CaMgSi2O6-CaFeSi2O6-
Mg2Si2O6-Fe2Si2O6 ). Indeed, the compositional field is ternary (CaSiO3-MgSiO3-
FeSiO3 ), but the molar amount of wollastonite (CaSiO3 ) never exceeds 50% in
the mixture. Initial studies on the clinopyroxene-orthopyroxene miscibility gap
considered only the magnesian end-members (enstatite and diopside; see, for in-
stance, Boyd and Schairer, 1964). Extension to the quadrilateral field was later
attempted by Ross and Huebner (1975). Figure 5.70 shows the pyroxene quadri-
lateral at 5, 10, and 15 kbar total pressure according to Lindsley (1983).
As figure 5.70 shows, with increasing T the miscibility gap shrinks toward
the iron-free join; moreover, on the Ca-poor side of the quadrilateral, miscibility
relations are complicated by the formation of pigeonite over a restricted P-T-X
range (cf. section 5.4.6). Once the compositions of coexisting clinopyroxene and
orthopyroxene are known, figure 5.70 is a practical geothermometer.
Clinopyroxene
If the analysis reports the amount of ferric iron, this value is accepted and aluminum is
partitioned between VI-fold and IV-fold coordinated sites, according to
Al VI + Fe VI
3+
+ CrVI + 2 Ti VI = Al IV + Na M 2 ( 5.265)
and
Al VI + Al IV = Al total. ( 5.266 )
Silicates / 393
If the analysis reports only total iron expressed as FeO (for example, microprobe data),
the amount of IV-fold coordinated aluminum is assumed to obey
Al IV = 2 − Si . ( 5.267 )
3⫹
FeVI is then defined by equations 5.265 and 5.266.
394 / Geochemistry of Crystal Phases
The resulting molar amounts are already normalized to 1 and may be plotted in figure
5.70 for thermometric deductions.
Orthopyroxene
(1) AlIV ⫽ 2 ⫺ Si
(2) AlVI ⫽ Altotal ⫺ AlIV
(3) Fe3⫹ calculated by applying equation 5.265
(4) R3⫹ ⫽ AlVI ⫹ Cr ⫹ Fe3⫹
(5) R2⫹ ⫽ (1 ⫺ X)Mg ⫹ X Fe2⫹
(6) NaR3⫹Si2O6 ⫽ smaller amount between Na and R3⫹
(7) NaTiAlSiO6 ⫽ smaller amount between Ti and AlIV or remaining Na
(8) R2⫹TiAl2O6 ⫽ smaller amount between Ti and AlIV /2
(9) R2⫹R3⫹AlSiO6 ⫽ smaller amount between remaining R3⫹ and AlIV
(10) Remaining Ca, Fe2⫹, and Mg normalized to form wollastonite ⫹ enstatite ⫹
ferrosilite.
On the basis of the experimental data of Lindsley and Dixon (1976) on the
CaMgSi2O6-Mg2Si2O6 binary join and accounting for the observed compositions
of natural pyroxenes coexisting at equilibrium, Kretz (1982) proposed two empiri-
cal thermometric equations applicable to clinopyroxene and valid, respectively,
for T ⬎ 1080 °C and T ⬍ 1080 °C:
( )
T K =
1000
( 5.268 )
(
0.468 + 0.246 X Fe,cpx − 0.123 ln 1 − 2 X Ca,cpx )
and
Silicates / 395
( )
T K =
1000
, ( 5.269 )
(
0.054 + 0.608 X Fe,cpx − 0.304 ln 1 − 2 X Ca,cpx )
where XCa,cpx ⫽ Ca/(Ca ⫹ Mg) and XFe,cpx ⫽ Fe/(Fe ⫹ Mg) in clinopyroxene.
An analogous formulation, also based on the experiments of Lindsley and
Dixon (1976), came from Wells (1977):
( )
T K =
7341
aMg 2 Si2 O6 , cpx
,
3.355 + 2.44 X Fe,opx − ln (5.270)
aMg 2 Si2 O6 ,opx
after attribution of all Ca to site M2 and all Al to site M1 (cf. Wood and Banno,
1973). This choice must be considered as purely operational, because it is incon-
sistent with the real nature of the mixture. It must also be emphasized that both
the formulations of Kretz (1982) and Wells (1977) are based on equilibrium distri-
bution of the magnesian component between the unmixed terms, although
Kretz’s (1982) formulation is explicitly derived for one of the two polymorphs. In
both formulations, the effect of adding iron to the system (i.e., moving from the
CaMg-Mg join into the quadrilateral) is regarded as simply linear (cf. eq. 5.268,
5.269, and 5.270), owing to the fact that, in natural clinopyroxenes, there is a
linear relationship between XFe,cpx and ln (1 ⫺ 2XCa,cpx ) (see Kretz, 1982).
Al 2 SiO 5 ⇔ Al 2 SiO 5 .
andalusite kyanite (5.274)
∆G P0,T = − RT ln K . (5.275)
If the heat capacity functions of the various terms in the reaction are known and
their molar enthalpy, molar entropy, and molar volume at the Tr and Pr of refer-
ence (and their isobaric thermal expansion and isothermal compressibility) are
also all known, it is possible to calculate ⌬G 0P,T at the various T and P conditions
of interest, applying to each term in the reaction the procedures outlined in sec-
tion 2.10, and thus defining the equilibrium constant (and hence the activity prod-
uct of terms in reactions; cf. eq. 5.272 and 5.273) or the locus of the P-T points
of univariant equilibrium (eq. 5.274). If the thermodynamic data are fragmentary
or incomplete—as, for instance, when thermal expansion and compressibility
data are missing (which is often the case)—we may assume, as a first approxima-
tion, that the molar volume of the reaction is independent of the P and T intensive
variables. Adopting as standard state for all terms the state of pure component
at the P and T of interest and applying
( )
∆G P0,T = ∆H P0,T − T∆S P0,T + P − Pr ∆VP0,T = − RT ln K , (5.276)
we may also identify the P-T locus of points where the equilibrium constant takes
on the identical value. It is obvious from equation 5.276 that equilibrium among
solids involving a conspicuous volume of reaction ( ⌬V 0 ) will be a good baromet-
ric function, and that one involving consistent ⌬H 0 (and limited ⌬V 0 ) will be a
good thermometric function.
Table 5.74 lists the main solid-solid reactions commonly adopted in geochem-
istry, with their respective ⌬ H 0Pr ,Tr, ⌬ S P0 r ,Tr, ⌬ V 0Pr ,Tr, and ⌬ G 0Pr ,Tr, calculated with
the INSP and THERMO computer packages (Saxena, 1989).
Some peculiar aspects of the various reactions are briefly summarized below.
The discussion is taken from Essene (1982), to whom we refer for more exhaus-
tive treatment.
Silicates / 397
Reaction
Paragenesis Reaction Number
Reaction
Number ⌬H 0Pr ,Tr ⌬S 0Pr ,Tr ⌬V 0Pr ,Tr ⌬G 0Pr ,Tr
1 ⫺4.1 ⫺9.1 ⫺7.4 ⫺1.4
2 8.2 13.5 5.7 4.1
3 4.1 4.4 ⫺1.7 2.7
4 1.6 3.4 3.0 ⫺2.9
5 7.5 3.1 3.7 6.5
6 41.56 7.69 4.61 39.27
7 ⫺11.0 ⫺44.1 ⫺17.3 2.1
8 19.3 ⫺18.8 ⫺13.9 24.9
continued
398 / Geochemistry of Crystal Phases
Reaction
Number ⌬H 0Pr ,Tr ⌬S 0Pr ,Tr ⌬V 0Pr ,Tr ⌬G 0Pr ,Tr
Reactions 1, 2, and 3. Al2SiO5 polymorphs are widely adopted as index minerals in meta-
morphic terranes and are easily recognizable even on the macroscopic scale. Their P-T
stability fields have long been subjected to some uncertainty because of disagreement
among authors on the exact location of the ternary invariant point (P ⫽ 6.5 kbar and
T ⫽ 595 °C according to Althaus, 1967; 5.5 kbar and 620 °C according to Richardson et
al., 1969; 3.8 kbar and 600 °C according to Holdaway, 1971). Essene (1982) reports the
substantial preference of petrologists for Holdaway’s (1971) triple point, which better con-
forms to thermobarometric indications arising from other phases. As already outlined in
section 2.7, the recent study of Holdaway and Mukhopadhyay (1993) confirms that the
triple point location of Holdaway (1971) is the most accurate.
Reactions 4 and 5. Bohlen and Boettcher (1981) studied the effect of the presence of
magnesian and manganoan terms in the mixture on the distribution of ferrous component
between orthopyroxene and olivine according to the reaction
Fe 2 Si 2 O6 ⇔ Fe 2 SiO 4 + SiO2 .
opx olivine quartz (α or β ) (5.277)
Figure 5.71A shows univariant equilibrium curves for various molar amounts of ferrous
component in the orthopyroxene mixture. The P-T field is split into two domains, corre-
sponding to the structural state of the coexisting quartz ( ␣ and  polymorphs, respec-
tively). If the temperature is known, the composition of phases furnishes a precise estimate
of the P of equilibrium for this paragenesis. Equation 5.277 is calibrated only for the
most ferriferous terms, and the geobarometer is applicable only to Fe-rich rocks such as
charnockites and fayalite-bearing granitoids.
Reaction 6. This reaction was investigated in detail by Finnerty and Boyd (1978) and
was more recently recalibrated by Köhler and Brey (1990). Because the Ca content in
olivine is P-dependent, due essentially to a large solvus between monticellite and forsterite
that expands with pressure (cf. section 5.2.5), this reaction should act as a sensitive baro-
metric function. However, the enthalpy of reaction is quite high and the effect of T on the
equilibrium is also marked. The calibrated P-T slope has an inflection, the origin of which
is not clear at first glance.
Silicates / 399
For practical applications, see Ghent (1976), Newton and Haselton (1981), and Perkins
(1983).
Reactions 10 and 11. The equilibrium
(Ca 0.333 )
Fe 0.666 Al2Si 3O12 ⇔ CaAl2Si2O8 + Fe2SiO4
3
( 5.278)
was calibrated by Green and Hibberson (1970). The reaction is clearer if we split the
garnet compound into the pure terms grossular and almandine (figure 5.71B):
3
a CaAl ⋅ a Fe
3
2 Si 2O 8 2 SiO 4
K 279 = . (5.280)
a Ca 3 Al 2 Si 3O12 ⋅ a 2
Fe 3 Al 2 Si 3O 12
Reaction 5.282 was calibrated by Bohlen et al. (1982) and is particularly useful in granu-
lites and garnet-bearing amphibolites. Figure 5.71C shows the loci of P-T points for which
equilibrium constant K282 assumes identical values (base-10 logarithm). The univariant
Silicates / 401
curve for pure end members (log K ⫽ 0) is in boldface. The corresponding reaction involv-
ing magnesian terms was calibrated by Newton and Perkins (1982).
Reactions 14 and 15. There are still several problems regarding the calibration of these
equilibria, which are discussed in detail by Essene (1982). Based on the energy terms listed
in table 5.74, both T and P markedly affect reaction 14, whereas reaction 15 furnishes
precise barometric indications.
Reactions 16 and 17. The reaction
was calibrated by Bohlen et al. (1983). The loci of identical value of equilibrium constant
K285 are plotted in the P-T space of figure 5.71D (base-10 logarithm), which also shows
the position of the univariant curve for pure end members (log K ⫽ 0) and, superimposed,
the kyanite-andalusite-sillimanite primary phase fields. As we can see, because the equilib-
rium is not markedly affected by temperature, it is a precise geobarometer.
often is) on the basis of a single component. This, however, is a dangerous ex-
ercise.
As shown in figure 5.72, the equality of potential is satisfied for component
a in mixtures ␣ and  at the measured compositions. Neglecting the energy infor-
mation arising from component b, one could deduce that the two phases, with
the measured composition (open circles), coexist stably at the estimated P and T
of equilibrium. However, as we can see, b,␣ ⬆ b, and equilibrium at the de-
duced P and T conditions would demand different compositions (solid circles).
Things may be even more dramatic with increasing chemical complexity of the
system: correct solution of equilibria of type 5.282 and 5.283 in terms of chemical
potentials involves, for example, five equalities at the univariant equilibrium (i.e.,
10 equations in six variables, SiO2 being a pure phase).
O2 , H2O, CO2 , H2 , and S2 . Of the reactions involving molecular oxygen (O2 ), the
following buffer equilibria are of great importance:
1
FeO ⇔ Fe + O 2 (IW) log a O 2 = 6.80 − 27568T −1
wuestite metal
2 gas (5.286)
1
NiO ⇔ Ni + O 2 (NNO) log aO2 = 9.31 − 24810T −1
bunsenite metal
2 gas (5.287)
1
Fe3O 4 ⇔ 3FeO + O2 (MW) log aO2 = 12.92 − 32638T −1
magnetite wuestite
2 gas (5.288)
1
3Fe2O3 ⇔ 2Fe3O 4 + O2 (HM) log aO2 = 14.41 − 24912T −1
2 gas (5.290)
hematite magnetite
The significance of the term “buffer” stems from the control operated by the
solid paragenesis on the activity of the O2 component in the gaseous phase. Con-
sider, for instance, the buffer magnetite-wuestite. Applying the usual relations
among chemical potential, standard state chemical potential, and thermody-
namic activity, we obtain
3
a FeO × a O1 2
∆G 288
0
= − RT ln 2
. (5.291)
a Fe 3O 4
If we imagine the solid phases to be composed of the pure terms FeO and Fe3O4 ,
equation 5.291 may be reduced to
∆G 288
0
= − RT ln a O1 2 (5.292)
2
and, by rearranging, to
2 ∆G288
0
aO2 = exp − . ( 5.293)
RT
404 / Geochemistry of Crystal Phases
1
Fe 2TiO 4 + Fe 3O 4 + O2 ⇔ FeTiO3 + 2 Fe2O3 .
2 (5.294)
spinel spinel gas hemo- ilmenite hemo- ilmenite
Equation 5.294 may be split into two partial equilibria, as shown by Powell and
Powell (1977b): the exchange reaction
which is independent of the gaseous phase, and the HM buffer (eq. 5.290).
Figure 5.73 shows the loci of aO2-T points obtained by intersecting the iso-
pleths of the titaniferous terms (XFeTiO3 and XFe2TiO4 ) in the coexisting spinel and
hemo-ilmenite phases. At T ⬍ 600 °C, the thermobarometer is inapplicable, be-
cause of the solvus field existing between magnetite and ulvospinel in the spinel
phase. Note in the same figure the position of the HM buffer, corresponding to
the absence of Ti in the system. The Buddington-Lindsley thermobarometer finds
application in both igneous (see Duchesne, 1972) and metamorphic terranes
(with the above low-T limits; see also Bohlen and Essene, 1977 for a discussion).
A multiple buffer of particular importance in volcanic gaseous equilibria (see
also section 9.5) is
Assuming unitary activity of components in solid phases (pure phases and stan-
dard state ⫽ pure component at P and T of interest), we have
aO2 ∆G296
0
K 296 = 2
= exp − . ( 5.297 )
a S3
2
RT
Let us now consider the effect of pressure on equilibrium. We assume for the
sake of simplicity that fluid pressure is equal to total pressure and that gaseous
components mix ideally. From
fO 2
a O2 = (5.298)
fO0
2
and
406 / Geochemistry of Crystal Phases
fS 2
a S2 = , (5.299)
fS0
2
0
where f O2 is the fugacity of O2 and f O 2
is the standard state fugacity of O2 :
fO 2 = PO 2 Γ O 2 (5.300)
fS 2 = PS 2 Γ S 2 , (5.301)
where PO2 is the partial pressure of O2 and ⌫O2 is the fugacity coefficient of O2 :
PO 2 = PX O 2 (5.302)
PS 2 = PX S 2 , (5.303)
3
f S0 Γ O2 X O2
PK 296 = 2 2 2
. ( 5.304 )
02
f O Γ 3S X S3
2 2 2
If we adopt as the standard state for gaseous components the state of pure perfect
gas at P ⫽ 1 bar and T ⫽ 298.15 K (f 0O2 ⫽ f 0S2 ⫽ 1) and neglect for simplicity
the fugacity coefficients, equation 5.304 combined with equation 5.297 gives
X O2 ∆G296
0
2
= P exp − . ( 5.305)
X S3
2
RT
There are three variables in equation 5.305: composition (and speciation) of the
fluid, P, and T. Once the chemistry of the fluid and P are known, we can deduce
T; if we know the chemistry of the fluid and T, we can deduce P; and if we know
P and T, we can deduce the chemistry of the fluid. This trivariance is a formidable
obstacle in the application of thermobarometry in the presence of fluid phases
and is often underevaluated in geology. Direct sampling of the fluid to constrain
the chemical factor is in most cases impossible, or, when it is possible (e.g., for
volcanic gases, geothermal gases, and fluid inclusions), the speciation state must
still be evaluated and, as we will see in detail in chapter 9, speciation in the fluid
phase is complex and quite variable as a function of P, T, and composition. Never-
theless, in the case of a simple oxygen buffer such as IW, NNO, WM, QFM, HM,
Silicates / 407
1
CaCO 3 + SiO 2 ⇔ Ca 2 Si 2 O 6 + CO 2 ,
2 (5.306)
calcite quartz wollastonite fluid
/ 411
412 / Geochemistry of Silicate Melts
If cations and anions are respectively of similar type, the ideal condition of maxi-
mum disorder will be reached in each matrix and the bulk entropy of mixing may
be expressed by
[( ) ( )]
S mixing = − R X Z ln X Z + X Y ln X Y + X A ln X A + X B ln X B , (6.2)
a AZ,melt = X A ⋅ X Z . (6.3)
2 O − ⇔ O 0 + O 2− .
melt melt melt
(6.4)
O− O− O− O−
O − − Si − O − + O − − Si − O − ⇔ O − − Si − O − Si − O − + O 2 − . (6.6)
− − − −
O O O O
Polymer chemistry shows that the larger the various polymers become, the
more their reactivity becomes independent of the length of the polymer chains.
This fact, known as the “principle of equal reactivity of cocondensing functional
groups,” has been verified in fused polyphosphate systems (which, for several
properties, may be considered as analogous to silicate melts; cf. Fraser, 1977) with
polymeric chains longer than 3 PO3⫺4 units (Meadowcroft and Richardson, 1965;
Cripps-Clark et al., 1974). Assuming this principle to be valid, the equilibrium
constant of reaction 6.4:
K4 =
(O )(O )
0 2−
(6.7)
(O )
2
−
(in which the terms in parentheses represent the number of moles in the melt) is
always representative of the polymerization process, independent of the effective
length of the polymer chains.
Toop and Samis (1962a,b) showed that, in a binary melt MO-SiO2 , in which
MO is the oxide of a basic cation completely dissociated in the melt, the total
number of bonds per mole of melt is given by
( ) ( )
2 O 0 + O − = 4 N SiO2 , (6.8)
where NSiO2 are the moles of SiO2 in the MO-SiO2 melt. The number of bridging
oxygens in the melt is thus
4 N SiO2 − O − ( ).
(O ) =
0
2
(6.9)
Mass balance gives the number of moles of free oxygen per mole of melt:
(O ) , −
(O ) = (1 − N )
2−
SiO 2 −
2
(6.10)
where (1 ⫺ NSiO2 ) are the moles of basic oxide in the melt. Equations 6.7 to
6.10 yield
414 / Geochemistry of Silicate Melts
K4 =
[4N SiO 2 ( )] [ 2 − 2 N
− O− SiO 2 − O− ( )] . (6.11)
4 (O )
2
−
(O ) ( 4 K ) ( )( ) ( )
2
−
4 − 1 + O − 2 + 2 N SiO2 + 8N SiO2 N SiO2 − 1 = 0, (6.12)
which may be solved for discrete values of K4 and NSiO2 (see table 6.1).
Figure 6.1 shows the solution of the system for K4 ⫽ 0.06: note that the distri-
bution of the three forms of oxygen is asymmetric over the compositional space.
Moreover, bridging oxygen is the only form present in the SiO2 monomer.
The Gibbs free energy change involved in equation 6.4 is
∆G 40 = − RT ln K 4 (6.13)
Because two moles of O⫺ produce one mole of O 0 and one of O2⫺, the Gibbs
free energy of mixing per mole of silicate melt is given by
∆G mixing =
(O ) RT ln K
−
4. ( 6.14 )
2
Figure 6.2 shows Gibbs free energy of mixing values obtained by application of
equation 6.14. They are remarkably similar to the Gibbs free energy of mixing
values experimentally observed on binary synthetic systems of appropriate com-
position. This means that the Gibbs free energy of mixing in an MO-SiO2 melt is
Thermochemistry of Silicate Melts / 415
Figure 6.1 Equilibrium distribution of (O0), (O⫺), and (O2⫺) in binary field MO-SiO2 for
K4 ⫽ 0.06. Reprinted from G. W. Toop and C. S. Samis, Canadian Metallurgist Quarterly,
1, 129–152, copyright 䉷 1962, with kind permission from Elsevier Science Ltd., The Bou-
levard, Langford Lane, Kidlington 0X5 1GB, UK.
Figure 6.2 Gibbs free energy of mixing in the quasi-chemical model of Toop and Samis
(1962a,b), compared with values experimentally observed in PbO-SiO2 and CaO-SiO2
melts. Reprinted from Toop and Samis (1962b), with kind permission of ASM Interna-
tional, Materials Park, Ohio.
416 / Geochemistry of Silicate Melts
∆G mixing
(1 − N ) log a
SiO 2 O2−
+ N SiO2 log a SiO2 =
2.303 RT
. (6.15)
In Masson’s model (Masson, 1965, 1968, 1972), polymerization reactions are vis-
ualized as a stepwise process of the type
Si 2 O 76 − + SiO 44− ⇔ Si 3O 10
8−
+ O 2− , (6.17)
The constants of equations 6.16 and 6.17 are identical, on the basis of the prin-
ciple of equal reactivity of cocondensing functional groups. Adopting Temkin’s
model for fused salts, and thus assuming that molar fractions represent activities
over the appropriate matrix, we obtain
X Si ⋅ X O2 − X Si ⋅ X O2 − X ⋅ X O2 −
6−
2O7
8−
3O10 Si n + 1O(32nn++44 ) −
= =K . ( 6.19 )
X SiO4 − ⋅ X SiO4 − X SiO4 − ⋅ X Si 6− X SiO4 − ⋅ X
4 4 4 2O7 4 Si nO(32nn++12 ) −
1 1 1
=2+ − .
X SiO2 1 − a MO 1 (6.20)
1 + a MO − 1
K 16
Equation 6.20 is valid only for open or “bifunctional” chains (cf. Fraser, 1977).
An extension of the model to open ramified polymer chains leads to the equation
Thermochemistry of Silicate Melts / 417
1 1 3
=2+ − .
X SiO2 1 − a MO 3 (6.21)
1 + a MO − 1
K 16
For instance,
—i.e.,
418 / Geochemistry of Silicate Melts
Figure 6.3 Experimental oxide activities in MO-SiO2 melts compared with estimates of
Masson’s model. Reprinted from Masson (1968), with kind permission of American Ce-
ramic Society Inc., Westerville, Ohio.
H2 SO 4 ⇔ HSO −4 + H + . ( 6.26 )
—i.e.,
MO ⇔ M 2 + + O 2 − (6.29)
ZMZ⫹
Cation K4
rMZ⫹ ⫹ rO2⫺
K⫹ 0.13 0.001
Na⫹ 0.18 0.001
Li⫹ 0.23 0.001
Pb2⫹ 0.30 0.002–0.01
Ca2⫹ 0.35 0.002
Mn2⫹ 0.41 0.04
Fe2⫹ 0.44 0.05
Co2⫹ 0.44 0.10
Ni2⫹ 0.46 0.35
MO + O 2 − ⇔ MO 22 − . (6.30)
Figure 6.4 Ternary plot of singly bonded oxygen O⫺, silicon SiIV, and bridging oxygen
O0 in each kind of polyanion. Shaded areas: compositional ranges 6.31 and 6.33. Re-
printed from Ottonello (1983), with kind permission of Theophrastus Publishing and Pro-
prietary Co.
In natural melts, the presence of high field strength ions such as Al 3⫹, Fe3⫹,
Ti , and P5⫹, which, like silicon, preferentially assume a tetrahedral coordination
4⫹
with oxygen, complicates the structure, and the constitution of the anion matrix
may not be deduced on the basis of the equations in section 6.1.2. Structural
parameters valid for compositionally complex melts were proposed by Mysen et
al. (1980) and Virgo et al. (1980) on the basis of the results of Raman spectros-
copy. These parameters are NBO/Si and NBO/T (NBO ⫽ Non-Bridging Oxygen;
T groups all tetrahedrally coordinated cations—i.e., Si4⫹, Al 3⫹, Fe3⫹, Ti4⫹, P5⫹,
. . .). Comparing the Raman spectra of silicate melts with various NBO/Si and
NBO/T ratios with the Raman spectra of crystals of the same chemical com-
position, the above authors subdivided the anion structure of silicate melts into
three compositional ranges:
T2 O 52 − ⇔ TO 23 − + TO 2 (6.37)
These equations do not imply any progression in the polymerization process and
reduce to identities in the Fincham-Richardson formalism (cf. table 6.3). Never-
theless, the experimental evidence of Mysen et al. (1980) and Virgo et al. (1980)
may be explained by the progressive polymerization steps listed in table 6.3. These
equilibria are consistent with the Fincham-Richardson formalism (of which they
constitute simple multiples) and obey the proportionality rules of figure 6.4.
Whiteway et al. (1970) derived an equation defining the fractional amount
NX of the various polymer units of extension X as a function of the ratio between
bridging oxygen (O 0 ) and singly bonded oxygen in an initially totally depolymer-
ized melt (O⫺*) composed entirely of SiO4⫺4 monomers (i.e., ␣ ⫽ O /O *):
0 ⫺
X −1 2 X +1
2α 2α
NX = ωX 1 − (6.38)
3 3
3X
ωX = . ( 6.39 )
( )
2X + 1 ! X!
Table 6.4 lists the molar fractions of the various polymer units calculated for ␣
values between 0.1 and 0.5.
Note the close connection between the values deduced by application of
equation 6.38 and the ternary representation of figure 6.4. The polymerization
field 0.1 ⬍ ␣ ⬍ 0.5 corresponds roughly to the range 4 ⬎ NBO/Si ⬎ 2. The
polymer units in this range are mainly monomers, dimers, and ramified tetramers
13 ). With increasing ␣, we progressively move along the polymerization
(Si4O10⫺
path, the molar fraction of monomers decreases, and the mean length of the
polymer units increases. For values of ␣ higher than 0.5, equation 6.38 is no
longer valid and is replaced by
422 / Geochemistry of Silicate Melts
Table 6.3 Reaction schemes for coexistence of various polymers in natural melts (from
Ottonello, 1983).
X −1 2 X +1
2α 2α
NX = ωX
3
1 −
3
(1 − N ) ,
O2−
(6.40)
Figure 6.5 Coordination state of Ti4+ in glasses of TiO2-SiO2 system. Reprinted from
R. B. Greegor, F. W. Lytle, D. R. Sanstrom, J. Wang, and P. Schultz. Investigation of
TiO2-SiO2 glasses by X-ray absorption spectroscopy, Journal of Non-Crystalline Solids, 55,
27–43, copyright 䉷 1983, with kind permission of Elsevier Science-NL, Sara Burgerhart-
straat 25, 1055 KV Amsterdam, The Netherlands.
ies show that Ce3⫹, Nd 3⫹, Eu3⫹, Er3⫹, and Yb3⫹ occupy mainly octahedral sites,
although coordinations VII and VIII may occasionally be observed (cf. Rob-
inson, 1974, and references therein).
Let us consider one mole of a pure crystalline phase at equilibrium with one mole
of melt of the same composition—e.g., pure forsterite at P ⫽ 1 bar and T ⫽ 2163
K. Because the two phases are at equilibrium, we have
dµ = dG = − S dT + V dP , (6.43)
we can write
− S Mg dT + V Mg dP = − S Mg dT + V Mg dP. (6.44)
2 SiO4 , olivine 2 SiO4 , olivine 2 SiO4 , melt 2 SiO4 , melt
(
dT S Mg SiO
2 4 , melt
− S Mg SiO
2 4 ,olivine
) = dP(V Mg 2SiO 4 , melt )
− VMg 2SiO4 ,olivine . (6.45)
Denoting the two terms in brackets, respectively, as ⌬S fusion and ⌬V fusion and
rearranging, we obtain the Clausius-Clapeyron equation:
dT ∆Vfusion
= . (6.46)
dP ∆S fusion
∆ H fusion
∆ S fusion = (6.47)
T
Thermochemistry of Silicate Melts / 425
Figure 6.6 Experimentally observed univariant dT/dP equilibria for several silicates of
petrogenetic interest. Melting curves are satisfactorily reproduced by Simon equation with
parameters listed in table 6.5.
dT Tfusion ⋅ ∆Vfusion
= . (6.48)
dP ∆ H fusion
Table 6.5 Parameters of Simon equation for some rock-forming silicates (after Bottinga, 1985).
Tmax and Pmax are maximum conditions of applicability.
Table 6.5 lists the parameters of the Simon equation simulating the dT/dP
evolution of the univariant curve. The Simon equation:
b
T
P = a − 1 + PSimon (6.49)
TSimon
is a purely empirical fitting of experimental evidence and must be used with cau-
tion (cf. Bottinga, 1985).
µ Mg
0
+ RT ln a Mg 2SiO4 ,olivine = µ Mg
0
+ RT ln a Mg 2SiO4 , melt . (6.50)
2SiO 4 ,olivine 2SiO 4 , melt
a Mg 2SiO4 , melt µ Mg
0
− µ Mg
0
2SiO 4 ,olivine 2SiO 4 melt
ln = . (6.51)
a Mg 2SiO4 ,olivine RT
Because
µ0 H0 S0
= − , (6.52)
RT RT R
∂ a Mg 2SiO4 , melt H Mg
0
− H Mg
0
2SiO 4 , melt 2SiO 4 ,olivine
ln = . (6.53)
∂ T a Mg 2SiO4 ,olivine RT 2
T T
if we assume ⌬CP to be constant within the integration limits, equation 6.53 be-
comes
a Mg 2SiO4 , melt 1
T
(
∆H fusion + ∆C P T − Tfusion ) dT ,
ln
a Mg 2SiO4 ,olivine
=
RT ∫ T 2
(6.55)
fusion
a Mg 2 SiO4 , melt 1 1 1 T
ln
a Mg 2 SiO4 ,olivine
= (
∆H fusion − ∆CP Tfusion
R Tfusion T
)
− + ∆CP ln .
Tfusion
( 6.56 )
(see Saxena, 1973 and references therein; Wood and Fraser, 1976).
Equation 6.56 is known as the “equation of lowering of freezing point” and
is valid for solid mixtures crystallizing from multicomponent melts. Like the
Clausius-Clapeyron equation, it tells us how the system behaves, with changing
T, to maintain equilibrium on the univariant curve. However, whereas in the
Clausius-Clapeyron equation equilibrium is maintained with concomitant
changes in P, here it is maintained by appropriately varying the activity of the
component of interest in the melt and in the solid mixture.
Recalling the relationship between thermodynamic activity and molar con-
centration,
ai = X i γ i , (6.57)
we see that, for practical purposes, the response of the system to modifications in
T consists of a flow of matter between the two phases at equilibrium.
Lastly, it must be noted that, whenever a pure single-component phase crys-
428 / Geochemistry of Silicate Melts
tallizes from the melt, the denominator of the term on the left side of equation
6.56 is reduced to 1 and the equation directly furnishes the activity of the dis-
solved component in the melt.
We have already seen that the degree of polymerization of the melt is controlled
by the amount of silica in the system (see, for instance, figure 6.4). If we mix two
fused salts with the same amount of silica and with cations of similar properties,
the anion matrix is not modified by the mixing process and the Gibbs free energy
of mixing arises entirely from mixing in the cation matrix—i.e.,
Table 6.6 Activity coefficients of jadeite component NaAlSi2O6 in molten (and/or over cooled)
mixture NaAlSi2O6-KAlSi2O6. Terms in parentheses indicate experimental incertitude at various T
(from Fraser et al., 1983).
␥NaAlSi2O6
0.4 1.032 (⫾0.03) 1.043 (⫾0.02) 1.043 (⫾0.02) 1.050 (⫾0.02) 0.998 (⫾0.04)
0.5 1.022 1.039 1.046 1.057 1.041
0.6 1.014 1.035 1.048 1.063 1.069
0.7 1.008 1.031 1.051 1.069 1.086
0.75 1.005 1.029 1.051 1.070 1.089 (⫾0.02)
0.8 1.004 1.018 1.032 1.045 1.056
0.9 1.001 1.005 1.006 1.011 1.014
1.0 1.000 1.000 1.000 1.000 1.000
Thermochemistry of Silicate Melts / 429
Table 6.7 Activity coefficients of albite (Ab) and orthoclase (Or) components in molten mixtures
and over cooled liquids in NaAlSi3O8-KAlSi3O8 system (after Rammensee and Fraser, 1982).
XAb Ab Or Ab Or Ab Or Ab Or
Figure 6.8 Partial molar enthalpies and integral enthalpy of mixing in pseudobinary
NaAlSi3O8-KAlSi3O8 system. Reprinted from W. Rammensee and D. G. Fraser, Geochim-
ica et Cosmochimica Acta, 46, 2269–2278, copyright 䉷 1982, with kind permission from
Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington 0X5 1GB, UK.
Fraser et al. (1983, 1985) deduced that, in several instances, the structure of the
silicate melt should mimic the structural arrangement of the solid phase at liq-
uidus. In incongruent melting processes, the structure of the liquid differs sub-
stantially from that of the solid phase of identical stoichiometry.
The quasi-ideality of several other molten systems with constant amounts of
silica may be deduced by applying the lowering of freezing point equation. A very
Thermochemistry of Silicate Melts / 431
Because the solid Fe2SiO4-Mg2SiO4 mixture is ideal (cf. section 5.2.5), the activi-
ties of components in the solid are given by
a Fo,olivine = X Mg,olivine
2
(6.61)
a Fa,olivine = X Fe,olivine
2
. (6.62)
Substituting equations 6.59, 6.60, 6.61, and 6.62 into the lowering of freezing
point equation, we obtain
X Mg, melt 1
T
( )
∆H fusion,Fo + ∆C P ,Fo T − Tfusion,Fo
2 ln = ∫
X Mg,olivine R Tfusion,Fo T2
dT ( 6.63)
X Fe, melt 1
T
( )
∆H fusion,Fa + ∆C P,Fa T − Tfusion,Fa
2 ln = ∫
X Fe,olivine R Tfusion,Fa T2
dT . (6.64)
Wood and Fraser (1976) assigned the following molar values to the various pa-
rameters: ⌬H fusion,Fo ⫽ 34,240 (cal); ⌬H fusion,Fa ⫽ 25,010 (cal); ⌬CP,Fo ⫽ ⌬CP,Fa ⫽
7.45 (cal/K); Tfusion,Fa ⫽ 1479.5 (K); Tfusion,Fo ⫽ 2164 (K).
Ratios XMg,melt /XMg,olivine and XFe,melt /XFe,olivine necessary to maintain equilib-
rium between the coexisting phases (solid plus melt) can be calculated by means
of equations 6.63 and 6.64. The results are shown in table 6.8.
If we let a represent the XFe,melt /XFe,olivine ratio and let b represent the XMg,melt /
XMg,olivine ratio, then, because
and
432 / Geochemistry of Silicate Melts
Table 6.8 Relative amounts of Fe and Mg in olivine and melt at equilibrium, based on equations
6.63 and 6.64 (after Wood and Fraser, 1976).
T (K )
XFe, melt / XFe, olivine (a) 1.0 1.21 1.39 1.70 1.93 2.06
XMg, melt / XMg, olivine (b) – 0.23 0.27 0.34 0.40 0.43
Figure 6.9 Melting relations in Mg2SiO4-Fe2SiO4 system. Experimental data confirm va-
lidity of Richardson’s (1956) ideal model for liquid. Reprinted from B. J. Wood and D. G.
Fraser, Elementary Thermodynamics for Geologists, 1976, by permission of Oxford Univer-
sity Press.
we obtain
X Fe, melt =
(
a 1− b ) (6.67)
a−b
and
1− b
X Fe,olivine = . (6.68)
a−b
Figure 6.9 shows how the calculated phase boundaries compare with the ex-
perimental observations of Bowen and Schairer (1935) on the same binary join.
The satisfactory reproduction of phase assemblages clearly indicates that the
Thermochemistry of Silicate Melts / 433
If we rapidly quench a silicate melt, we may obtain a glass without the formation
of any crystalline aggregate. The temperature of transition between melt and glass
may be operatively defined as “that temperature at which the differing thermody-
namic properties of melt and glass intersect” (Berman and Brown, 1987). The
supercooling phenomenon, which may easily be observed in melts with molar
amounts of silica higher than 50%, involves discontinuities in the values of heat
capacity, enthalpy, entropy, and Gibbs free energy. Moreover, the various thermo-
dynamic properties for the vitreous state (and for the transition temperature) vary
with the quenching rate. Figure 6.10 shows, for instance, the enthalpy and heat
capacity for a CaMgSi2O6 melt compared with the corresponding properties in
the crystalline and vitreous states at various T. Note in figure 6.10A that the
dH/dT slope is different for vitreous and molten states. Moreover, the glass-melt
transition is observed at two different temperatures, depending on quenching rate
(T ⫽ 1005 or 905 K), and the quenching rate also affects the enthalpy of the
vitreous state, which is not the same in the two cases. The enthalpy of the crystal-
line counterpart in all T conditions is lower than the molten and vitreous states.
At the temperature of fusion Tf , the enthalpy difference between crystal and melt
is the enthalpy of fusion of CaMgSi2O6 (⌬H fusion ), whereas the enthalpy differ-
ences at lower T are enthalpies of solution of pure diopside in a pure CaMgSi2O6
melt (⌬H solution ) and enthalpies of vitrification (⌬H vitrification ). Note also in figure
6.10B that the heat capacity of the vitreous state (CPv ) is very similar to the heat
capacity of the crystal (CPc ) and that both, T being equal, are notably lower than
the heat capacity of the molten state (CPf ).
Lastly, we observe that the glass-melt transition involves a marked discontinu-
ity at T of transition, reflecting the increase of vibrational freedom (rotational
components appear). This discontinuity involves some complexity in the extrapo-
lation of calorimetric data (usually obtained on glasses) from vitreous to molten
states, discussed in detail by Richet and Bottinga (1983, 1986).
434 / Geochemistry of Silicate Melts
Figure 6.10 Temperature dependence of enthalpy (A) and heat capacity (B) for CaMg-
Si2O6 component in crystalline, vitreous, and molten states at various T conditions. Re-
printed with permission from Richet and Bottinga (1986), Review of Geophysics and Space
Physics, 24, 1–25, copyright 䉷 1986 by the American Geophysical Union.
Table 6.9 Parameters for calculation of heat capacity of silicate melts: (1) after Carmichael et al.
(1977) and Stebbins et al. (1984); (2) model of Richet and Bottinga (1985)
Constituent Oxide
Model Parameter SiO2 TiO2 Al2O3 Fe2O3 FeO MgO CaO Na2O K2O
(1) K1 80.0 111.8 157.6 229.0 78.9 99.7 99.9 102.3 97.0
⫾0.9 ⫾5.1 ⫾3.4 ⫾18.4 ⫾4.9 ⫾7.3 ⫾7.2 ⫾1.9 ⫾5.1
(2) K1 81.37 75.21 27.21 199.7 78.94 85.78 86.05 100.6 50.13
K2 ⫻ 103 0.0 0.0 94.28 0.0 0.0 0.0 0.0 0.0 15.78
K3 ⫻ 10⫺5 0.0 875.3 0.0 0.0 0.0 0.0 0.0 0.0 0.0
(1) CP 冢moleJ⫻ K冣 ⫽ ∑ K
i
1,iX i ; (2) CP 冢moleJ⫻ K冣 ⫽ ∑ X (K i
i 1,i ⫹ K2,i T ⫹ K3,i T ⫺2) ⫹ CP,excess
where Xi is the molar fraction of component i in the molten mixture and C 0P,i is
its molar specific heat at the standard state of pure melt. On the basis of calori-
metric measurements carried out on natural and synthetic molten mixtures in the
T range 1000 to 1600 K, Carmichael et al. (1977) suggested that the ideal model
postulated for binary mixtures may be extended to multicomponent systems.
They also observed that the heat capacity of the various components in the mix-
ture may be considered independent of T, within the limits of experimental incer-
titude, through the suitable choice of standard state values (i.e., a single term in
Maier-Kelley-type functions; cf. section 3.2). However, further investigations by
Stebbins et al. (1984) and Richet and Bottinga (1985, 1986) showed significant
deviations from the ideal mixing model. For instance, in the pseudoternary system
CaAl2Si2O8-NaAlSi3O8-CaMgSi2O6 , positive deviations from ideality occur
along the NaAlSi3O8-CaMgSi2O6 join and negative ones in CaMgSi2O6-
CaAl2Si2O8 molten mixtures.
Table 6.9 lists the parameters of the model of Carmichael et al. (1977) for the
heat capacity of silicate melts, recalibrated by Stebbins et al. (1984) and valid for
the T range 1200 to 1850 K. To obtain the heat capacity of the melt at each T
condition, within the compositional limits of the system, it is sufficient to combine
linearly the molar proportions of the constituent oxides multiplied by their re-
spective CP values (cf. equation 6.69).
Table 6.9 also lists the parameters of Richet and Bottinga’s (1985) model, which,
devised primarily for Al-free melts, accounts for the nonideality of molten sili-
cate systems containing K by introducing an excess term that represents Si-K inter-
actions:
J
CP, excess = WSi − K X SiO2 X K2 O ; W Si − K = 151.7 . ( 6.71)
2 mol e × K
The model of Richet and Bottinga (1985) considers the heat capacity of the melt to
be variable with T and allows better reproducibility of experimental evidence in composi-
tionally simple systems. The Stebbins-Carmichael model seems to reproduce the experi-
mental observations on aluminosilicate melts better (Berman and Brown, 1987). Applica-
tion of both models to natural melts gives substantially identical results (i.e., differences
of about 1%, within the range of data uncertainty; cf. Berman and Brown, 1987).
Once the enthalpy of fusion is known, because solid and liquid are at equilibrium
at the melting point, the entropy of the molten component at Tf is also known
through
If the heat capacity of the molten component is known (cf. section 6.6.2), molar
enthalpy and molar entropy at the standard state of the pure melt at T ⫽ 298.15
and P ⫽ 1 bar may be readily derived by applying
Tf
0
H melt , 298.15 K ,1 bar = H melt,T f − ∫ C P , melt dT ( 6.74 )
298.15
Thermochemistry of Silicate Melts / 437
Table 6.10 Parameters for calculation of molar volume of silicate melts at various T
conditions (after Stebbins et al., 1984). Resulting volume is in cm3/mole.
and
Tf
dT
0
S melt , 298.15K ,1 bar = S melt,T f − ∫ C P, melt
T
. (6.75)
298.15
Vmelt = ∑ Xi Vi 0 . ( 6.76 )
i
Bottinga and Weill (1970) applied equation 6.76 extensively, and it was later recal-
ibrated by Stebbins et al. (1984). In the latter model, the molar volume of the
melt and its modification with T are calculated through the equation
whose coefficients are listed in Table 6.10. Equation 6.77 also implies that thermal
expansion is a simple additive function of the linear expansions of the various
oxide components.
Concerning melt compressibility, existing experimental data are rather scanty.
On the basis of solid/liquid univariant equilibria for six silicate components
(Mg2SiO4 , MgSiO3 , CaMgSi2O6 , Mg3Al2Si3O12 , NaAlSi2O6 , and NaAlSi3O8 )
438 / Geochemistry of Silicate Melts
silicate melts is lower than that of the corresponding solids, the calculations of
Bottinga (1985) imply that the density contrast decreases with increasing P. This
fact would tendencially obstaculate the separation of molten portions from the
residual solid, with increasing depth within the earth’s crust and mantle.
The existence of miscibility gaps in silicate melts has been experimentally known
for decades (Greig, 1927; Bowen, 1928; Bowen and Schairer, 1938; Schairer and
Bowen, 1938; Roedder, 1951, 1956; Holgate, 1954) and has been directly observed
both in terrestrial rocks (Anderson and Gottfried, 1971; Ferguson and Currie,
1971; De, 1974; Gelinas et al., 1976; Philpotts, 1979; Coltorti et al., 1987) and in
lunar basalts (Roedder and Wieblen, 1971).
Figure 6.12 combines experimental evidence and natural observations in the
form of a ternary SiO2 - (CaO⫹MgO⫹FeO⫹TiO2 ) - (Na2O⫹K2O⫹Al2O3 ) plot.
The glassy portions of lunar basalts have compositions that conform quite well
to the experimental miscibility gap of Gelinas et al. (1976), whereas terrestrial
volcanites have smaller gaps.
Immiscibility phenomena in silicate melts imply positive deviations from ide-
ality in the mixing process. Ghiorso et al. (1983) developed a mixing model appli-
cable to natural magmas adopting the components listed in table 6.12. Because
all components have the same standard state (i.e., pure melt component at the T
and P of interest) and the interaction parameters used do not vary with T, we are
dealing with a regular mixture of the Zeroth principle (cf. sections 2.1 and 3.8.4):
1
G melt, P,T = ∑ ni µ i0 + RT ∑ ni ln X i +
2
n total ∑ ∑ Wij X i X j , (6.78)
i i i j
where 0i is the standard state chemical potential of the pure ith melt component
at the P and T of interest, ni is the number of moles of the ith component, Xi is
the molar ratio of the ith component, and Wij is the interaction parameter between
components i and j.
In applying the model above (which is an extension of the previous version of Ghiorso
and Carmichael, 1980, integrated with new experimental observations), some caution
must be adopted, because of the following considerations:
Figure 6.12 Miscibility gap in natural and synthetic silicate melts. Dots: unmixing for a
tholeiitic magma, based on model of Ghiorso et al. (1983). ∑ FeO: total iron as FeO.
2. Calculation of the excess Gibbs free energy of mixing (third term on right side of
eq. 6.78) involves only binary interactions. Although there is no multiple interac-
tion model that can be reduced to the simple summation of binary interactions
used here (cf. Acree, 1984; see also section 3.10), this choice is more than ade-
quate for the state of the art, which does not allow precise location of the miscibil-
ity gap in the chemical space of interest.
3. Because the excess Gibbs free energy of mixing is independent of P, the volume,
compressibility, and thermal expansion of the melt are obviously regarded as lin-
ear combinations of the values of its components (see section 6.6.4). As table
6.12 shows, the model uses a set of ad hoc components. The choice of components
is dictated by simple convenience—i.e., the possibility of deriving realistic ther-
modynamic data for melt constituents through univariant equilibria involving
their crystalline counterparts or through additive procedures.
The first step in the model is a normative calculation (table 6.12) to establish
the molar amounts of the various melt components. For each component, the
molar Gibbs free energy at the standard state of pure component at T and P of
interest is given by
P
Gi 0,P ,T , melt = H i0,P
r ,T , melt
− TSi 0,P
r ,T , melt
+ ∫ Vi ,0P ,T , melt dP , ( 6.79 )
Pr
where
Tf
H i0,P ,T , melt
r
= H i0,P ,T , crystal
r r
+ ∫ CP ,i , crystal dT
Tr (6.80)
+ ∆H i , fusion + CP ,i , melt T − T f ( )
and
Tf
dT
Si 0,P
r ,T , melt
= Si 0,P
r ,T r , crystal
+ ∫ CP ,i , crystal T
Tr
(6.81)
T
+ ∆Si , fusion + CP ,i , melt ln .
Tf
Tables 6.13 and 6.14 list sets of internally consistent thermodynamic data for
crystalline and liquid components, for use in equations 6.79 to 81. Although equa-
tions 6.80 and 6.81 are simply based on the Clausius-Clapeyron approach, Ghi-
orso et al. (1983) use a semiempirical formulation for the volume of melt compo-
nents. Its development is
442 / Geochemistry of Silicate Melts
C P0
H i0, Pr , Tr , solid S i0, Pr , Tr , solid
Component (kcal/mole) (cal/(mole ⫻ K)) K1 K2 ⫻ 10⫺3 K3 ⫻ 105 K4
V i0, melt
(cal/(mole ⫻ bar))
Tf ⌬ S 0 fusion ⌬ H 0 fusion 0
C P, i , melt
Component (K) (cal/(mole ⫻ K)) (kcal/mole) (cal/(mole ⫻ K)) a V, i b V, i ⫻ 10 4
∫ Vi ,0P ,T , melt dP =
Pr
(6.83)
( 2
)
1 1
2
(
6
1
)
= Vi ,0P ,T , melt P − Pr − cV ,i P 2 − Pr2 − cV ,i P 2 P 3 − Pr3 ,
( )
r
Thermochemistry of Silicate Melts / 443
where aV,i and bV,i are the thermal expansion coefficients listed in table 6.14 and
cV,i is a compressibility factor of the form
1 × 10 −6
cV ,i = . ( 6.84 )
2.76 0
0.7551 + Vi ,P ,T , melt
ni r
Equation 6.83 is not valid for component Si4O8 , for which the following equation
is used:
We can then derive the calculated Gibbs free energy of mixing with respect to the
molar amount of the component of interest, thus obtaining the difference between
the chemical potential of the component in the mixture and its chemical potential
at standard state:
∂G mixing,melt
= RT ln a i , P,T , melt = µ i , P,T , melt − µ i0, melt . (6.87)
∂ X i P, T , X
j
444 / Geochemistry of Silicate Melts
1 2 3 4 5 6 7 8 9 10
2 ⫺29,364
3 ⫺78,563 ⫺67,350
4 2,638 ⫺6,821 1,240
5 ⫺9,630 ⫺4,595 ⫺59,529 4,524
6 5,525 ⫺2,043 ⫺1,918 212 ⫺703
7 ⫺30,354 12,674 ⫺48,675 ⫺1,277 ⫺57,926 ⫺2,810
8 ⫺64,068 ⫺102,442 ⫺98,428 1,520 ⫺59,356 699 ⫺78,924
9 ⫺73,758 ⫺101,074 ⫺135,615 ⫺3,717 ⫺36,966 780 ⫺92,611 ⫺62,780
10 ⫺87,596 ⫺40,701 ⫺175,326 284 ⫺84,580 ⫺61 ⫺45,163 ⫺27,908 ⫺18,130
11 ⫺412 ⫺196 ⫺71,216 ⫺103,024 7,931 310 ⫺20,260 ⫺38,502 ⫺49,213 ⫺23,296
When the partial derivative reaches zero, the two potentials coincide and compo-
nent i is present unmixed as a pure term. If unmixing takes place in solvus condi-
tions, the thermodynamic activity of component i remains constant for the entire
solvus field. However, in the case of spinodal decomposition, the activity of i
within the spinodal field plots within a maximum and minimum (cf. sections 3.11
and 3.12).
For instance, figure 6.13 shows the behavior of the partial derivative of the
Gibbs free energy of mixing with respect to the molar amount of Si4O8 com-
ponent:
∂ G mixing,melt
= RT ln a Si 4O 8 , P,T , melt
∂X Si 4O 8, melt P,T , X
j (6.88)
(
= 4 µ SiO2 , P,T , melt − µ SiO
0
2 , melt
).
On the SiO2-rich side of the diagram, spinodal decomposition (points b and d in
figure 6.13) is seen at 1700 K. Ghiorso et al. (1983) showed that the topology of
the Gibbs free energy of mixing surface in a multicomponent space is rather com-
plex, with a saddle-shaped locus of spinodal decomposition. The calculations of
Ghiorso et al. (1983) show that the immiscibility dome in a T-X space intersects
the liquidus field of a tholeiitic magma. Table 6.16 gives the composition of a
tholeiite (Rattlesnake Hill basalt) investigated by Philpotts (1979) because it was
evidently an unmixed magma. Philpotts (1979) remelted the rock and was able to
observe the unmixing phenomenon experimentally. During the cooling stage, two
liquids were produced by unmixing at 1313 °C. Analysis of the unmixed portions
indicated that one of the unmixed terms was richer in TiO2 , total iron (as FeO),
Thermochemistry of Silicate Melts / 445
Figure 6.13 Partial derivative of Gibbs free energy of mixing of melt with respect to Si4O8
component plotted against molar amount of silica in melt. Enlarged area: spinodal field
of critical immiscibility (spinodal decomposition). From M. S. Ghiorso and I. S. E. Car-
michael, A regular solution model for meta-aluminous silicate liquids: applications to
geothermometry, immiscibility, and the source regions of basic magma, Contributions to
Mineralogy and Petrology, 71, 323–342, figure 3, copyright 䉷 1980 by Springer Verlag.
Reproduced with modifications by permission of Springer-Verlag GmbH & Co. KG.
Figure 6.14 Gibbs free energy of mixing surface calculated for Archean basalts of Piumhi
(Minas Gerais, Brazil) in SiO2-Na2O compositional space.
MnO, MgO, CaO, and P2O5 , and the other was richer in SiO2 , Na2O, and K2O.
Table 6.16 lists the compositions of these unmixes according to the Ghiorso-
Carmichael model (also plotted in the ternary diagram of figure 6.12): the two
liquids have the same compositional features described in the experiment of Phil-
potts (1979).
Figure 6.14 shows in stereographic form the values of the Gibbs free energy
of mixing obtained for the Archean basalts of Piumhi (Minas Gerais, Brazil) on
the basis of the data of Coltorti et al. (1987): note the pronounced saddle defining
the spinodal decomposition field (unmix compositions also plotted in ternary
diagram of figure 6.12).
The capability of evaluating with sufficient accuracy the Gibbs free energy of
a silicate melt in various P-T-X conditions has obvious petrogenetic implications,
besides those already outlined. For instance, the P-T loci of equilibrium of a
given crystal with the melt can be determined with good approximation. Let us
consider, for example, the equilibrium
Mg 2 SiO 4 ⇔ Mg 2 SiO 4 .
olivine melt
(6.89)
Thermochemistry of Silicate Melts / 447
For the pure component, reaction 6.89 has a constant (K89 ) whose value under
various P-T conditions gives the activity ratio of the magnesian component in
the two phases:
For a given value of P, there is only one T condition at which equation 6.90
is satisfied. Hence, once the activity of the component in the two phases is known,
the P-T loci of equilibrium are also known. This concept can be generalized for
solid phases whose crystalline components (M) are stoichiometric combinations
of melt components (C) according to
n
M solid = ∑ viC i , (6.91)
i =1
n
RT ln K = ∑ vi RT ln a i , melt − RT ln a i ,solid . (6.92)
i =1
∂ Gα
µ1,α = ( 7.1)
∂ n1,α P ,T
∂ Gα
µ 2 ,α = ( 7.2 )
∂ n2 ,α P ,T
Gα = ∑ ni , α µ i , α ( 7.3)
i
/ 449
450 / Geochemistry of Silicate Melts
Figure 7.1 Relationships between chemical potential and composition in binary phases
with different miscibility behavior: ␣ ⫽ complete miscibility;  ⫽ partial miscibility; ␥ ⫽
lack of miscibility or “mechanical mixture.”
In equation 7.4, G ␣ is the molar Gibbs free energy of phase ␣ at the composi-
tional point defined by fractional amounts X1,␣ and X2,␣. The Gibbs free energy
curve of phase  is saddle-shaped, indicating the existence of a miscibility gap.
Applying the principles of minimization of Gibbs free energy and of equality of
potentials at equilibrium, we can deduce that the compositions of the two un-
mixed portions ⬘ and ⬙ are determined by the points of tangency (and identi-
cal slope):
G β ′ = X 1, β ′ µ1, β ′ + X 2, β ′ , µ 2, β ′ (7.6)
Introduction to Petrogenetic Systems / 451
G β ′′ = X 1, β ′′ µ1, β ′′ + X 2, β ′′ , µ 2, β ′′ . (7.7)
Denoting X⬘ and X ⬙ the molar fractions of the unmixed portions, we also have
G β = X β ′G β ′ + X β ′′G β ′′ . (7.8)
it is evident that phase ␣ has an excess Gibbs free energy sufficiently low (or
negative) to avoid the formation of a solvus. In the case of phase , Gexcess mixing
is positive and sufficiently high to induce a zone of immiscibility. In the case of
phase ␥, the terms Gideal mixing and Gexcess mixing are mutually compensated over the
whole compositional range.
Figure 7.2 G-X and T-X plots for a binary system with a molten phase with complete
miscibility of components at all T conditions and a solid phase in which components are
totally immiscible at all proportions (mechanical mixture, ␥ ⫽ ␥⬘ ⫹ ␥⬙).
the lowest temperature of melting of the system.) At T6, the Gibbs free energy of
the mechanical mixture is lower than that of phase ␣ over the entire binary range:
only the mechanical mixture ␥⬘ ⫹ ␥ ⬙ is stable at equilibrium.
Let us now consider the cooling process of a system of composition C in
figure 7.2. At high T, the system is completely molten (phase ␣, condition T1 ). As
T is progressively lowered, condition Tc is reached, at which ␥⬘ begins to crystal-
lize. The crystallization of ␥⬘ enriches the composition of the residual melt of
component 2. At T4, the melt has composition C⬙, while crystals ␥⬘ continue to
form. The molar proportions of crystals and melts may be deduced by application
of baricentric coordinates at temperature T4 (relative lengths of the two composi-
tional segments in the two-phase field ␣ ⫹ ␥⬘, or “lever rule”). We reach eutectic
point E at T5: phase ␥ ⬙ forms at equilibrium in stable coexistence with ␣ and ␥⬘;
the temperature now remains constant during the crystallization of ␥ ⬙, until all
the residual melt is exhausted.
At T6, the melt is absent and ␥⬘ and ␥ ⬙ coexist stably in a mechanical mixture
of bulk composition C. The crystallization process in a closed system is thus com-
pletely defined at all T (and P) conditions by the Gibbs free energy properties of
the various phases that may form in the compositional field of interest.
(i.e., cannot be defined simply on the basis of the two components CaMgSi2O6
and CaAl2Si2O8 ), but is rather “pseudobinary,” because a limited amount of
Al2O3 may be dissolved in CaMgSi2O6 . However, we will neglect this limited solu-
bility here and will hereafter treat the system as purely binary.
Pure components CaMgSi2O6 and CaAl2Si2O8 at P ⫽ 1 bar melt, respectively,
at Tf,Di ⫽ 1391.5 °C and Tf,An ⫽ 1553 °C. The eutectic temperature is 1274 °C.
The progressive decrease of incipient crystallization, from 1553 °C to 1274 °C on
the An-rich side of the join and from 1391.5 °C to 1274 °C on the Di-rich side, is
a result of the lowering of the freezing point expression (see chapter 6: eq. 6.56,
section 6.4). If we cool a system of initial composition C, we observe incipient
crystallization at T ⫽ 1445 °C (figure 7.3). At this temperature, the entire system
is still in the liquid form. At T ⫽ 1400 °C, some anorthite crystals are already
formed. We can determine their weight amounts by applying baricentric coordi-
nates (or the “lever rule”) to compositional segments An and L in figure 7.3—i.e.,
An 1
X An = = = 0.20 ( 7.11)
An + L 1 + 4
L 4
XL = = = 0.80. ( 7.12 )
An + L 1 + 4
Note that the resulting fractional amounts are in weight percent, because the
abscissa axis of the phase diagram reports the fractional weights of the two com-
ponents (similar application of baricentric coordinates to a molar plot of type
7.2 would have resulted in molar fractions of phases in the system). Applying the
lever rule at the various T, we may quantitatively follow the crystallization behav-
ior of the system (i.e., at T ⫽ 1350 °C, XAn 艐 0.333 and XL 艐 0.666; at T ⫽ 1300
°C, XAn 艐 0.44 and XL 艐 0.56; and so on). It is important to recall that the
baricentric coordinates lose their significance at the eutectic point, because solid/
melt proportions depend on time and on the dissipation rate of the heat content
of the system. Below the eutectic temperature, the lever rule can again be applied
to the couple Di-An, and the resulting proportions are those of the initial system
C (XAn ⫽ 0.70, XDi ⫽ 0.30). For a closed system (see section 2.1), the melting
behavior of an initial composition C exactly follows the reverse path of the crys-
tallization process (i.e., beginning of melting at T ⫽ 1274 °C/displacement along
univariant curve/abandonment of univariant curve at T ⫽ 1450 °C).
Figure 7.4 Gibbs free energy curves and phase stability relations for two binary mixtures
with complete miscibility of components (types I, II, and III of Roozeboom, 1899).
plete miscibility) may give rise to three distinct configurations of the phase dia-
gram, as outlined in figure 7.4.
Roozeboom Type I
Roozeboom type I is obtained whenever the Gibbs free energy curve of the melt
initially touches that of the mixture in the condition of pure component. This
takes place at the melting temperature of the more refractory component (T2 in
456 / Geochemistry of Silicate Melts
figure 7.4). With decreasing T, the intersections of the Gibbs free energy curves
of the solid and liquid mixtures define tangency zones spanning the entire binary
range. The two curves finally touch at the condition of pure component 1 (T5 ):
T5 is thus the melting temperature of the less refractory component in the mixture.
Projection of the two-phase zones in a T-X space gives the phase diagram shown
at the bottom of figure 7.4: liquidus and solidus curves limit a single lens-shaped
field of stable coexistence of solid and liquid ( ␣ ⫹ L).
Roozeboom Type II
The Gibbs free energy curves of both solid mixture and melt initially touch at an
intermediate composition of the binary join (T1 in the central column of figure
7.4). With decreasing T, the Gibbs free energy loop of crystalline mixture ␣ is
progressively lowered, crossing the Gibbs free energy loop of the melt in two
points. At T3, the chemical potential of pure component 2 is identical in the solid
and liquid phases. At this temperature (that of melting of pure component 2), the
Gibbs free energy curve of the solid phase still intersects the Gibbs free energy
loop of the melt in another part of the join. At T4 (melting temperature of pure
component 1), the two Gibbs free energy curves intersect at the condition of pure
component 1. At lower T, the entire Gibbs free energy loop of the solid lies below
the Gibbs free energy curve of the melt: the entire system is crystallized. The
resulting T-X phase diagram (figure 7.4, bottom center) shows two distinct two-
phase lenses ( ␣ ⫹ L) departing from a peak on the liquidus curve.
with a consolute temperature Tc ⫽ 660 °C, below which the homogeneous solid
mixture (Na, K)AlSi3O8 unmixes into an Na-rich (Abs.m. ) and a K-rich (Ors.m. )
composition. This further complexity introduces us to the third type of phase re-
lations.
Roozeboom Type IV
The Gibbs free energy curve of the solid phase with partial miscibility initially
touches the curve of the liquid in the condition of pure component 1. This hap-
pens at the melting temperature of that component (i.e., T1 ⫽ Tf,1 ). With decreas-
ing T, the Gibbs free energy loop of the solid is progressively displaced downward
(i.e., becomes more negative) and the equality conditions of potentials of compo-
nents 1 and 2 in the liquid and in unmixing lobe ⬙ are reached first (T2, T3 ). As
T further decreases, a particular condition (T4 ) is reached: the chemical potentials
of the two components are identical in the liquid and in the two lobes ⬘ and ⬙:
the three phases ⬘, ⬙, and L coexist stably at equilibrium. At T5, the solvus of
the solid expands (tangency segment ⬘-⬙) and there still exists a solid-liquid
zone of coexistence on the component 2–rich side of the join (segment L-⬘). At
T6, the chemical potential of pure component 2 is identical in the solid and liquid
phases: T6 is hence the melting temperature of the less refractory component
in the join (T6 ⫽ Tf,2 ). Below this temperature, only the solid phases are stable:
⬘ ⫹ ⬙ within the solvus field, and the homogeneous mixture  outside it. Projec-
tion of the various isothermal G-X geometries onto the T-X space gives the phase
diagram (figure 7.7, bottom right). The resulting phase stability relations may be
conceived as a combination of a Roozeboom type I diagram with a solvus field
intersecting it along the solidus curve.
Roozeboom Type V
Initially like Roozeboom type IV, Roozeboom type V differs from it at tempera-
ture T5, where the tangency condition of the two unmixing lobes ⬘ and ⬙ and
liquid L finds the Gibbs free energy curve of the liquid in an intermediate position
between the unmixing lobes, rather than in an external position, as in the previous
case. The resulting T-X stability relations may be conceived as a Roozeboom type
III diagram with a solvus field intersecting it along the solidus curve (cf. figures
7.6 and 7.7).
Lastly, let us consider the last column in figure 7.7, depicting a case in which
Figure 7.7 Gibbs free energy curves and T-X phase stability relations between a phase
with complete miscibility of components (silicate melt L) and a binary solid mixture with
partial miscibility of components (crystals ).
Introduction to Petrogenetic Systems / 461
the compounds are completely miscible in the solid phase at high T, but become
partially immiscible with decreasing T (this is the general case, from the thermo-
chemical point of view, for the crystalline solids of our interest). Because un-
mixing begins at T5, where the Gibbs free energy of the solid is entirely above
that of the liquid phase, the solvus does not intersect the solidus line in the re-
sulting T-X plot. We thus have a Roozeboom type III diagram with a solvus
opening at subsolidus conditions, which was the case in figure 7.6B for the
NaAlSi3O8-KAlSi3O8 system.
The unmixing fields shown in figure 7.7 are solvi (i.e., defined by binodes at
the various T), but in nature we may also observe the metastable persistency of a
homogeneous phase until a critical condition of unmixing is reached. The loca-
tion of a spinodal field on a T-X diagram follows the guidelines outlined in section
3.11—i.e., the spinodes at the various T plot on the inflection point on the Gibbs
free energy curve of the solid. Spinodal fields thus occupy an inner and more
restricted portion of the T-X space with respect to the corresponding solvi.
peritectic crystallization event. At T1, the Gibbs free energy of crystalline compo-
nent 1 is identical to that of the liquid (i.e., T1 ⫽ Tf,1 ). With decreasing T (T2, T3 ),
zones of tangency between the Gibbs free energy curves of the liquid and compo-
nent 1 are observed, but the chemical potential of intermediate compound C
decreases at the same time. Lastly, at T4 ⫽ Tf,C, we reach an invariant condition
with the coexistence of three phases: compound C, liquid L, and the pure crystals
of component 1. A peritectic reaction now takes place at constant T: the pure
crystals of component 1 react with the liquid, forming intermediate compound C.
Only when pure crystals 1 are completely exhausted does the temperature again
decrease. This decrease leads to another invariant point: the eutectic point of the
system, at which the liquid coexists stably with C and the pure crystals of compo-
nent 2 (T6 ). Below this T, the liquid is no longer stable and only the two solid
phases (crystals 2 ⫹ C) are stable at equilibrium. The resulting phase diagram is
shown at the bottom of figure 7.8.
NaAlSiO4 is immiscible with either SiO2 or NaAlSi3O8 (in other words, its stoichi-
ometry is fixed in the binary join of interest). When expressed in weight percent,
equation 7.13 involves 54% NaAlSiO4 and 46% SiO2 (the original diagram of
Schairer and Bowen, 1956, like most phase diagrams of that period, uses a weight
scale rather than a molar range, simply because mixtures were dosed in weight
amounts). Crystallization of any composition within NaAlSi3O8 and SiO2 follows
the classical behavior of a mechanical mixture, as discussed in section 7.1.1 (cf.
figure 7.2). For example, an initial composition C1 in figure 7.9 (corresponding
to 50% NaAlSiO4 and 50%SiO2 by weight) crystallizes at about 1110 °C, forming
NaAlSi3O8 (Ab). With decreasing T, the progressive crystallization of Ab dis-
places the composition of the residual liquid toward eutectic point E⬘, which is
Figure 7.9 NaAlSiO4-SiO2 (A) and KAlSiO4-SiO2 (B) system at P ⫽ 1 bar, after Schairer
and Bowen (1955, 1956).
Introduction to Petrogenetic Systems / 465
However, all the SiO2 component in the liquid is exhausted before reaction 7.14
is completed, and the resulting subsolidus assemblage is composed of Le ⫹ Sa
466 / Geochemistry of Silicate Melts
crystals. If the initial molar amount of SiO2 in the liquid is higher than the value
dictated by the molar proportion
peritectic reaction 7.14 achieves completion and the crystallization process con-
tinues with the production of sanidine at eutectic point E⬘. If, instead, the initial
molar amount of SiO2 in the liquid is lower than the value dictated by the mo-
lar proportion
(as, for instance, C5 in figure 7.9B), crystallization proceeds with the formation
of liquids increasingly impoverished in the SiO2 component until eutectic E⬙ is
reached.
Comparing figures 7.9 and 7.8, we can fully understand the nature of phase
stability relations in the examined systems: the binary KAlSiO4-SiO2 system is a
simple combination of the two configurations displayed in figure 7.8, whereas the
NaAlSiO4-SiO2 system shows some additional complexities on the Na-rich side
of the join. However, it is important to stress that the apparent complexity of T-
X phase stability diagrams always arises from the (quite simple) configuration of
the Gibbs free energy curves in the various phases forming in the system. This
indicates how important it is in geochemistry to establish accurately the mixing
properties of the various phases involved in petrogenetic equilibria.
We have seen that the crystallization of initial compositions respectively
poorer (C2) and richer (C1) in SiO2 than NaAlSi3O8 stoichiometry follows oppo-
site paths in the binary NaAlSiO4-SiO2 system. This occurs essentially because
compound NaAlSi3O8 melts at a local T maximum (or, in thermochemical terms,
with decreasing T the chemical potential of NaAlSi3O8 reaches the Gibbs free
energy curve of the liquid before the two eutectic configurations; cf. figure 7.8).
Analogously, in the KAlSiO4-SiO2 system, we have a local T maximum (T ⫽ 1686
°C) at the melting temperature of the stoichiometric compound KAlSi2O6 .
These apparently insignificant facts actually have profound consequences for
magma petrology, because they constitute thermal barriers below which magma
differentiation proceeds in opposite chemical directions (silica undersaturation or
silica oversaturation), depending on the initial composition of the system.
compositional plane. The latter is triangular, with each side representing a binary
join of the ternary system. The projection shows the liquidus surface and, at each
point of the diagram, highlights the condition of incipient crystallization of the
melt. To facilitate interpretation of the crystallization path followed by any initial
composition in the T-X field, the isotherms are drawn on the liquidus surface.
Figure 7.10 shows a ternary diagram generated by the projection from the T
axis of the phase limits for three mechanical mixtures of components 1, 2, and 3
(1–2, 1–3, and 2–3, respectively). Crystallization of an initial composition C1 will
follow the path shown in figure 7.10. Composition C1 is within “primary phase
field” ␥ ⬙. With decreasing T, once the value of the isotherm passing through C1
is reached, crystallization begins with the formation of ␥ ⬙ crystals of pure compo-
nent 2. As a result of crystallization of ␥ ⬙, the composition of the residual liquid
moves away from C1 along the direction obtained by connecting corner 2 with
the initial composition. During this step, the equilibria are divariant (because P
is fixed, the variance is given by: number of components ⫺ number of phases ⫹
1). Once the residual liquid reaches cotectic line E-EIII, determined by the inter-
section of primary phase field ␥ ⬙ with primary phase field ␥⬘, crystals ␥⬘ (com-
Figure 7.10 Phase stability relations in a ternary system in which components are totally
immiscible at solid state, and relationships with three binary joins.
468 / Geochemistry of Silicate Melts
Figure 7.11 Phase stability relations in a ternary system with components 1–3 and 2–3
totally immiscible at solid state and components 1–2 forming an intermediate compound
with fixed stoichiometry (Ci).
haustion of the residual melt, and at the end of crystallization the solid assem-
blage is composed of crystals ␥ ⬙, ␥ , and Ci.
The compositional paths of liquids for melting processes in closed systems
of initial composition C1 or C2 are exactly the opposite of those observed in
crystallization events.
Figure 7.12 Phase stability relations in a ternary system with complete solid state misci-
bility of components 1–2 and complete immiscibility at solid state for components 1–3
and 2–3.
compositional vectors at each crystallization step and has the curved shape shown
in figure 7.12. This curved trajectory brings the residual liquid to cotectic line E⬘-
E⬙ at point C3, and the crystals at equilibrium with the liquid have composition
X3. Crystallization continues along the cotectic line with simultaneous precipita-
tion of the 1–2 crystalline mixture and of crystals ␥ composed of pure compo-
nent 3. Crystallization terminates at point C4, where the crystalline mixture 1–2
has composition X4, coinciding (in terms of relative proportions of components
1 and 2) with initial liquid C1. If the initial composition of the liquid falls in
primary phase field ␥ (C5), the initial crystallization path is initially straight
until it intersects the cotectic line. Crystallization then proceeds as in the preced-
ing case.
In all the preceding cases, crystallization takes place in a closed system. The loss
of heat toward the exterior causes a decrease in temperature, total pressure and
bulk chemistry being equal, and this modifies the Gibbs free energy relationships
among the various phases of the system under the new P and T conditions. In
petrology, however, it is also of interest to investigate the effects of the combined
exchange of matter and energy toward the exterior. The processes taking place in
this case are usually defined as “fractional crystallization” (loss of energy and
matter) and “fractional melting” (gain of energy, loss of matter), whereas the
corresponding terms “equilibrium crystallization” and “equilibrium melting” are
applied when there is no exchange of matter with the exterior. The counterposi-
tion “fractionation-equilibrium” implicit in this terminology may give rise to con-
fusion, and it is worth stressing here that fractional melting and fractional crystal-
lization processes also take place at equilibrium, but that equilibrium is local.
We will therefore use the corresponding terms “melting in an open system” and
“crystallization in an open system” to avoid any ambiguity.
Let us now consider, as an example, crystallization in an open system in pla-
gioclases (figure 7.13). Let us imagine cooling the initial composition C until the
liquidus curve of the system is reached (C1); the crystals in equilibrium with
liquid C1 have composition X1. With decreasing T, we reach condition C2, where
crystals of composition X2 coexist with a liquid of composition L2. If the system
is further cooled by a simple heat loss, we reach point C3. The crystals that
formed during cooling at point C2 should now react with the liquid to reach
composition X3. Let us now imagine that the kinetic rate of reaction is lower
than the rate of heat loss toward the exterior: the inner portions of the crystals of
composition X2 do not react with the liquid, and the final composition of the
whole crystals is not X3 but intermediate between X3 and X2. The net result is
that part of the system (inner parts of crystals) plays no part in the equilibrium:
472 / Geochemistry of Silicate Melts
the new bulk chemistry is no longer that of the initial system C but is displaced
toward the less refractory component (i.e., more albitic). Proceeding in this way,
we define a compositional path of fractionation corresponding to progressive sub-
traction of the inner portions of the crystals as crystallization proceeds. Note that,
as a net result, crystallization is not complete at the solidus temperature of initial
system C (as in the case of crystallization in a closed system), but proceeds down
to the melting temperature of the low-melting component (i.e., in our case,
Tf,Ab ⫽ 118 °C). The resulting crystals exhibit wide compositional zoning, inde-
pendent of the initial composition of the system (in our case, crystal nuclei have
composition An80 and the surfaces of crystals have composition An 0 ).
The compositional relations discussed above are explained by Rayleigh’s dis-
tillation process (Rayleigh, 1896), choosing a suitable partition coefficient for the
component of interest—i.e., for component An, we can write
1 − F LK An
X An,solid = X An,initial , ( 7.17 )
1 − FL
whereX An,solid is the mean composition of the crystals, FL is the fractional amount
of residual liquid in the system, and KAn is the solid/liquid partition coefficient at
interface for component An (see section 10.9.1 for derivation of eq. 7.17).
Introduction to Petrogenetic Systems / 473
Figure 7.14 Crystallization and melting in an open system: ternary join CaMgSi2O6-
CaAl2Si2O8-NaAlSi3O8 (Di-An-Ab). Crystallization (B) and melting trends (C) are taken
from Morse (1980).
474 / Geochemistry of Silicate Melts
of composition C2 in the binary Ab-An join (T ⫽ 1408 °C). Once this tempera-
ture is reached, melting starts again, generating liquids that are increasingly richer
in An component, up to pure An. The composition of fractionally accumulated
liquids now moves along the dotted line in figure 7.14C. The amount of this liquid
may be determined by applying the baricentric coordinates to segments L2-C1-
C2, L3-C1-C3, and L4-C1-C4 (FL ⫽ 0.52, 0.59, and 0.79, respectively). Note that
the compositional paths of fractionally accumulated liquids and solids produced
during open-system melting and crystallization are not opposite, as in the case of
closed systems.
PART THREE
Geochemistry of Fluids
CHAPTER EIGHT
There are two main ways of representing the H2O molecule: the point-charge
model and molecular orbital representation.
Figure 8.1A shows the point-charge model of Verwey (1941). Point charges
are located at the equilibrium distance (0.99 Å) and form bond angles of 105°
between two H-O bond directions. There are six proton charges (⫹6e), located
on oxygen and half a proton charge (⫹12e) on each of the two hydrogen ions. A
charge of ⫺7e (resulting from the electroneutrality condition) is arbitrarily placed
at 0.024 Å from the oxygen nucleus along the bisector of the bond angle. The
resulting dipole moment ( ⫽ 1.87 ⫻ 10⫺18 esu ⫻ cm) thus agrees with experi-
mental observations.
Figure 8.1B shows an experimental contour map of electron density for the
H2O molecule in plane y-z, after Bader and Jones (1963). The electron density is
higher around the nuclei and along the bond directrix. The experimental electron
density map conforms quite well to the hybrid orbital model of Duncan and Pople
(1953) with the LCAO approximation.
/ 479
480 / Geochemistry of Fluids
Figure 8.1 Structural models for H2O molecule. (A) Electrostatic point-charge model
(from Eisemberg and Kauzmann, 1969; redrawn). (B) Electron density map (from Bader
and Jones, 1963). (C) Formation of MOs ⌿(b⬘) and ⌿(b⬙) starting from AOs 2p and 2s of
oxygen and 1s of hydrogen.
Based on the model of Duncan and Pople (1953), two molecular orbitals (b⬘ and b⬙)
may form as linear combinations of hybrid orbitals 2p and 2s of oxygen and orbital 1s
of hydrogen:
[ ]
Ψ( b′ ) = λ cos ε bψ O(2 s ) + sin ε bψ O(2 p, b′ ) + ξψ H′ (1s ) ( 8.1)
[ ]
Ψ( b′′ ) = λ cos ε bψ O(2 s ) + sin ε bψ O(2 p, b′′ ) + ξψ H′′ (1s ) ( 8.2)
and two “lone pair” orbitals (l⬘ and l⬙) may form as combinations of hybrid orbitals 2s
and 2p of oxygen:
Geochemistry of Aqueous Phases / 481
The significance of term was defined in section 1.17.1. Factor is analogous (ratio /
defines the polarity of the bond) and parameters b and l describe the hybridization state
of the orbitals. Because cos b ⫽ 0.093 (Duncan and Pople, 1953), molecular hybrid orbit-
als ⌿(b⬘) and ⌿(b ⬙) are composed essentially of the wave functions of atomic orbitals 2p of
oxygen and 1s of hydrogen. The value 0.578 obtained for cos l also indicates that orbitals
⌿(l⬘) and ⌿(l ⬙ ) are essentially of type sp3. Figure 8.1C shows the formation of hybrid MOs
⌿(b⬘) and ⌿(b ⬙ ) on plane y-z, starting from the AOs 2p and 1s of hydrogen.
(
∆U = K φ 1 − cos φ . ) ( 8.5)
In equation 8.5, K is the constant of distortion of the hydrogen bond. The degree
of distortion of the hydrogen bond in water varies with temperature [from 26° to
30° for 0 ⬍ T (°C) ⬍ 100] and is zero in ice, where H2O molecules are linked to
form perfect tetrahedra.
The “random network model” of Bernal (1964) is a development of Pople’s
model. In Bernal’s version, the water molecules are conceived as forming an irreg-
ular network of rings (not open chains, as in the preceding case). Most of the
rings have five members, because the O-H bond in a single H2O molecule is near
the angle (108°) characteristic of a five-membered ring, but four, six, seven, or
even more H2O molecules may be present in some rings.
482 / Geochemistry of Fluids
Figure 8.2 Models of distortion of hydrogen bond for water. Note that bond angle be-
tween two O-H bonds of a single molecule is 105°; formation of a five-membered ring
(part B) thus implies distortion of 3°.
(ε − 1)( 2ε + 1) = α +
µ 2 g 4πN 0 d
m × , ( 8.6 )
gε 3kT 3W
a a 4
g H2 O
= 1 + a1 d + a 2 d
3
5
− 1 , ( 8.7 )
T
Table 8.2 Dielectric constant of water at various T and P conditions, after Pitzer (1983).
Temperature (°C)
P (kbar) 400 450 500 550 600 650 700 750 800
1 – 12.49 9.43 7.04 5.34 4.24 3.52 3.04 2.69
2 – 16.46 13.87 11.69 9.87 8.37 7.14 6.16 5.37
3 21.78 18.72 16.20 14.10 12.31 10.80 9.51 8.42 7.49
4 23.46 20.35 17.83 15.73 13.95 12.44 11.14 10.02 9.05
5 24.82 21.63 19.63 16.94 15.17 13.65 12.35 11.21 10.22
aH + ⋅ aOH −
K H 2O = ( 8.9 )
a H 2O
varies directly with T and inversely with P, as shown in figure 8.4. Discrete values
of the ionic dissociation constant for T between 0 and 60 °C and P ⫽ 1 bar are
listed in table 8.3.
Eugster and Baumgartner (1987) suggested that the ion dissociation constant
can be numerically reproduced in a satisfactory way by the polynomial
B CP
log K H2O = A + + + D log d , (8.10)
T T
Figure 8.4 Ion dissociation constant of water as a function of P and T conditions. From
Eugster and Baumgartner (1987). Reprinted with permission of The Mineralogical Society
of America.
Within the PV region delimited by the two saturation boundary curves, liquid
and vapor phases coexist stably at equilibrium. To the right of the vapor satura-
tion curve, only vapor is present: to the left of the liquid saturation curve, vapor
is absent. Let us imagine inducing isothermal compression in a system composed
of pure H2O at T ⫽ 350 °C, starting from an initial pressure of 140 bar. The H2O
will initially be in the gaseous state up to P ⬍ 166 bar. At P ⫽ 166 bar, we
reach the vapor saturation curve and the liquid phase begins to form. Any further
486 / Geochemistry of Fluids
Figure 8.5 P-V projection of state diagram for H2O near critical point. Pc and Vc are
critical pressure and critical volume of compound.
compression on the system will not induce any increase in P, because the appear-
ance of a new phase has decreased the degree of freedom of the system, which is
now invariant. Compression thus simply induces an increase in the bulk density
of the vapor-plus-liquid assemblage, until the liquid saturation curve is reached.
At this point, the vapor phase disappears and the system is again univariant: any
further compression now induces a dramatic increase in P (because the liquid
phase is relatively incompressible with respect to the gaseous counterpart: the
slopes of the isotherms on the gas and liquid sides of the saturation boundaries
are very different).
If we now repeat compression along the critical isotherm Tc, we reach (at the
critical point) a flexus condition, analytically defined by the partial derivatives
∂P
=0 ( 8.11.1)
∂V T
c
∂ 2P
= 0. ( 8.11.2 )
∂V 2 T
c
Geochemistry of Aqueous Phases / 487
At T higher than Tc, a liquid H2O phase can never be formed, regardless of how
much pressure is exerted on the system.
Figure 8.6A is an enlarged portion of the P-T plane for H2O in the vicinity
of the critical point (Tc ⫽ 373.917 °C, Pc ⫽ 220.46 bar).
Five P-T regions of H2O stability are distinguished:
1. The “liquid” region, limited by the vaporization boundary (i.e., the liquid
saturation and vapor saturation boundary curves) and the critical isobar.
2. The “supercritical liquid” region, limited to supercritical pressures and
subcritical temperatures.
3. The “supercritical fluid” region, which encompasses all conditions of su-
percritical temperatures and pressures.
4. The “supercritical vapor” region, below the critical isobar and above the
critical isotherm.
Figure 8.6 (A) P-T regions of H2O stability. (B) Extended P-T field showing location of
the critical region. The critical isochore is ⫽ 0.322778 g/cm3. From Johnson and Norton
(1991), American Journal of Science, 291, 541–648. Reprinted with permission of Ameri-
can Journal of Science.
488 / Geochemistry of Fluids
5. The “vapor” region, limited by the vaporization boundary and the criti-
cal isotherm.
Figure 8.6B shows a wider P-T portion with the location of the “critical re-
gion” for H2O, bound by the 421.85 °C isotherm and the ⫽ 0.20 and 0.42 g/cm3
isochores. The PVT properties of H2O within the critical region are accurately
described by the nonclassical (asymptotic scaling) equation of state of Levelt
Sengers et al. (1983). Outside the critical region and up to 1000 °C and 15 kbar,
PVT properties of H2O are accurately reproduced by the classical equation of
state of Haar et al. (1984). An appropriate description of the two equations of
state is beyond the purposes of this textbook, and we refer readers to the excellent
revision of Johnson and Norton (1991) for an appropriate treatment.
The nonclassical equation of state of Levelt Sengers et al. (1983) for the critical region
of H2O may be expressed in terms of reduced parameters (cf. section 9.2) as follows:
∂P˜
ρr = , ( 8.12 )
∂µ˜ T˜
where r is the reduced density / c , with c ⫽ 0.322778 g/cm3 (critical density). The other
dimensionless parameters T̃, P̃, and ˜ are given by
T
T˜ = −Tr−1 = − c ( 8.13.1)
T
P T P
P˜ = r = c ( 8.13.2 )
Tr Pc T
ρ c Tc µ
µ˜ = , ( 8.13.3 )
Pc T
−1 2
∂A
( )
12
ρ = PCM ( 8.14 )
∂ρ T
(Johnson and Norton, 1991), where M is the molar weight of H2O (18.0152 g/mol), C is
a conversion factor (0.02390054 cal bar⫺1cm⫺3 ) and A is the molal Helmholtz free en-
ergy function:
( ) ( )
A( ρ , T ) = Abase ρ , T + Aresid ρ , T + Aideal ρ , T ( ) ( 8.15)
(see Johnson and Norton, 1991, for a detailed account on the significance of the various
terms in eq. 8.15). Note that the molar volume may be readily derived from equations
8.12 and 8.14 by applying V ⫽ M/ .
−1 −1
∂P ∂ 2P
=∞ and = ∞. ( 8.16 )
∂V T ∂V 2 T
c c
Moreover, recalling what we have already seen in section 1.15 (namely, eq.
1.95, 1.96, 1.97, and 1.103), we may also derive
−1 −1
∂ 2A ∂ 3A
= −∞ and = −∞, ( 8.17 )
∂V 2 T ∂V 3 T
c c
−1 −1
∂ 2U ∂ 3U
= −∞ and = −∞. ( 8.18 )
∂V 2 T ∂V 3 T
c c
Equations 8.16 and 8.17 or 8.16 and 8.18 show that, at the critical point, “the
specific first and second derivative properties of any representative equation of
state will be divergent” (Johnson and Norton, 1991). This inherent divergency
has profound consequences on the thermodynamic and transport properties of
H2O in the vicinity of the critical point. Figure 8.7 shows, for example, the behav-
490 / Geochemistry of Fluids
Figure 8.7 (A) Isobaric thermal expansion, (B) its first T-derivative, (C) isothermal com-
pressibility, and (D) isobaric heat capacity of H2O within the critical region, based on the
equation of state of Levelt Sengers et al. (1983). From Johnson and Norton (1991), Ameri-
can Journal of Science, 291, 541–648. Reprinted with permission of American Journal
of Science.
ior of the isobaric thermal expansion ( ␣ ) of H2O and of its first derivative on T
within the critical region, computed by Johnson and Norton (1991) on the basis
of the equation of state of Levelt Sengers et al. (1983). The divergent asymptotic
behavior of ( ∂␣ / ∂T ) in the vicinity of the critical point has been arbitrarily trun-
cated within the ⫺5 ⫻ 10⫺3 K⫺2 and 5 ⫻ 10⫺3 K⫺2 lower and upper limits for
illustrative purposes. Note also in parts C and D of figure 8.7 the huge domes in
the isothermal compressibility  and isobaric heat capacity CP of H2O within the
critical region.
Geochemistry of Aqueous Phases / 491
Now that we have examined to some extent the properties of water as a “solvent,”
let us discuss the nature of aqueous solutions.
From the point of view of chemical modeling, aqueous solutions are treated
as “electrolytic solutions”—i.e., solutions in which solutes are present partially
or totally in ionic form. “Speciation” is the name for the characteristic distribu-
tion of ion species in a given aqueous solution in the form of simple ions, ionic
couplings, and neutral molecules. Solutes in aqueous solutions are defined as
“electrolytes” and may be subdivided into “nonassociated” and “associated.”
Nonassociated electrolytes are also defined as “strong” and mainly occur in the
form of simple or simply hydrated ions. An example of a strong electrolyte is the
salt NaCl, which, in aqueous solution of low ionic strength, occurs in the form
of completely dissociated Na⫹ and Cl ⫺ ions.
Associated electrolytes are further subdivided into two groups: “weak electro-
lytes” and “ionic couplings.” A weak electrolyte occurs both in the form of cova-
lent molecules and in ionic form. Examples of weak electrolytes are various acids,
some bases, and a few inorganic salts. Aqueous solutions of carbonic acid H2CO3 ,
for instance, contain both undissociated neutral molecules H2CO03 and carbonate
and bicarbonate ions CO32⫺ and HCO3⫺. Solutions of strong inorganic acids such
as HF, when present in high molar concentration, also contain neutral molecules
HF 0 besides H⫹ and F⫺. The term “ionic coupling” is used when the association
of oppositely charged ions is due not to the formation of a strong covalent bond
(as in the case of weak electrolytes) but to electrostatic interactions. For instance,
in aqueous solutions of calcium sulfate CaSO4 , besides Ca2⫹ and SO42⫺ ions,
neutral particles CaSO4 0 are also present. In aqueous solutions of calcium hy-
droxide, some of the solutes are in the form of ionic couplings Ca(OH)⫹. The
distinction of weak electrolytes and ionic couplings in aqueous solutions is some-
what problematic, because in some cases both types of electrolytes have the same
stoichiometry. Lastly, the term “complex ion” identifies all nonmonatomic solute
ions, independent of the nature of the chemical bond. Examples of complex ions
are thus Ca(OH)⫹, CO32⫺, HCO3⫺, HS⫺, ZnCl42⫺, and B3O3(OH)4⫺.
siders the various aqueous ions as species physically present in solution, in molal
proportions determined by the coexistence of multiple homogeneous equilibria
of the type
a i = m iT γ T
i = miγ i , ( 8.20)
where mTi is the total molal concentration of the species and mi is the molal con-
centration of the free ion:
mi
γ T
=γi . ( 8.21)
i mT
i
The main advantage of the specific interactions model lies in the simplicity of its
calculations. Also, considering dissolved salts (in their neutral salt stoichiome-
try—e.g., NaCl) as “components,” activities and total activity coefficients are
experimentally observable magnitudes.
However, the ionic coupling–complexation model is more appropriate to the
actual complexity of aqueous solutions, as is evident from spectral absorption
and ionic conductivity studies.
When speaking of “solutions,” we implicitly state that the H2O “solvent” has a
different standard state of reference with respect to “solutes” (note that the term
“mixture” is used when all components are treated in the same way; see section
2.1). The standard state generally used for the solvent in aqueous solutions is that
of “pure solvent at P and T of interest” (or P ⫽ 1 bar and T ⫽ 298.15 K). For
solutes, the “hypothetical one-molal solution referred to infinite dilution, at P
and T of interest” (or P ⫽ 1 bar and T ⫽ 298.15 K) is generally used. This choice
is dictated by practical considerations.
Geochemistry of Aqueous Phases / 493
C v + Av − ⇔ v +C + v −A , ( 8.22 )
a e = K e m ev , ( 8.23)
where
v = v+ + v− . ( 8.24 )
( ) (a )
v+ v−
ae = a+ − ( 8.25 )
1v
( ) ( )
a± = a+
v+ v−
a− = a1e v . ( 8.26 )
1v
m± = m+ ( ) (m )
v+ v− 1v
= me , ( 8.27 )
−
a+ a−
γ+ = and γ− = , ( 8.28 )
m+ m−
1v
( ) (γ )
= γ+
a± v+ v−
γ = = γ 1e v . ( 8.29 )
m±
± −
The concept of ionic strength, which allows individual ionic activity coefficients
to be estimated, was developed by Lewis and Randall (1921). The ionic strength
of a solution is given by
1
I=
2
∑ m i Z i2 , ( 8.30.1)
i
where mi is the molality, Zi is the charge of the ith ion, and summation is extended
to all ions in solution.
Geochemistry of Aqueous Phases / 495
1 m
( ) = 3.
2
I= × 2 2 + m Cl − × −1 ( 8.30.2 )
2 Ca 2 +
1 m
( ) = 1.
2
I= × 1 2 + m Cl − × −1 ( 8.30.3)
2 Na −
The concept of ionic strength is very useful in quantifying the effects of electro-
static charges of all ions present in solution. Its relationship with the nature of
electrostatic interactions was stressed by Lewis and Randall: “In diluted solutions
the activity coefficient of a given strong electrolyte is the same in all solutions
having the same ionic strength.” In fact, in light of the Debye-Huckel theory, the
relationship between individual ionic activity coefficient ␥i of ion i and the ionic
strength of solution I is given by
AZ i2 I
log γ i = − . ( 8.31)
1 + åi B I
A and B in equation 8.31 are characteristic parameters of the solvent (in this case,
water), which can be derived in the various T and P conditions of interest by
application of
1.824829 × 10 6 ρ 1 2
A= (kg1 2 × mole −1 2 ) ( 8.32.1)
(εT )
32
and
50.291586 × 10 8 ρ 1 2
B= (kg1 2 × mole −1 2 × cm −1 ) ( 8.32.2 )
(εT )
12
(Helgeson and Kirkham, 1974). In equations 8.32.1 and 8.32.2, is solvent den-
sity, is the dielectric constant of the solvent, and åi is the “effective diameter”
of ion i in solution. The latter parameter must not be confused with the ionic
diameter of the ion in a condensed aggregation state, but represents the range of
electrostatic interaction with the solvent molecules (see section 8.11.3).
Table 8.4 lists discrete values of solvent parameters A and B for H2O in the
496 / Geochemistry of Fluids
T (°C) A B ⫻ 10⫺8
0 0.4911 0.3244
25 0.5092 0.3283
50 0.5336 0.3325
75 0.5639 0.3371
100 0.5998 0.3422
125 0.6416 0.3476
150 0.6898 0.3533
175 0.7454 0.3592
200 0.8099 0.3655
225 0.8860 0.3721
250 0.9785 0.3792
275 1.0960 0.3871
300 1.2555 0.3965
Table 8.5 Effective ion diameters (Å) for various ions and inorganic complexes in
aqueous solution (after Kielland, 1937).
4.5 2⫹ 2⫺ 2⫺ 2⫺ 2⫹
Pb , CO3 , SO3 , MoO4 , Co(NH3)5Cl , Fe(CN)5NO 2⫺
T range 0 to 300 °C (saturation P), and table 8.5 lists effective diameter values
for various ions in aqueous solution.
When the ionic strength of a solution is low, the denominator in equation
8.31 tends toward 1, and the individual ionic activity coefficient is well approxi-
mated by
log γ i ≈ − AZ i2 I . (8.33)
Geochemistry of Aqueous Phases / 497
The mean salt method derives implicitly from the concept of ionic strength, as
expressed by Lewis and Randall (1921). It is assumed that, within the range of
the ionic strength of interest, for a standard uni-univalent electrolyte such as KCl,
the following assumption is valid:
γ K+
=γ Cl −
( 8.34 )
From equation 8.34, with a reasonable approximation (see eq. 8.29), we can
derive
[( )(γ )]
1 2
γ ± KCl = γ K+ Cl −
=γ K+ =γ Cl −
. ( 8.35 )
Because the ␥⫾ values for the various salts in solution may be experimentally
obtained with a satisfactory precision, equation 8.35 is used to derive the corre-
sponding values of individual ionic activity coefficients from them. For instance,
from a generic univalent chloride MCl, the ␥⫾ of which is known, we may derive
the ␥⫹ of M⫹ by applying
1 2 1 2
γ ± MCl
=γ
( M+ )(γ ) Cl −
=γ
( M+ )(γ ± KCl )
( 8.36.1)
γ M+ = γ 2
± MCl γ ± KCl . (8.36.2)
13 13
( )(
γ ± MCl 2 = γ M 2 + γ Cl − )
(
= γ M 2 + γ ± KCl )( )
2 2
(8.37.1)
γ M 2+ = γ 3
± MCl γ ±2 KCl . (8.37.2)
Similarly, the same method and the same mean salt furnish values of individual
ionic activity coefficients for other ions—e.g., the sulfate group SO2⫺
4 :
(γ ) (γ )
13 13
(
γ ± K 2SO 4 = γ K + ) (
= γ ± KCl )
2 2
(8.38.1)
SO 24 − SO 24 −
γ SO2 − = γ 3
± K2 SO 4
γ ±2 KCl . ( 8.38.2 )
4
498 / Geochemistry of Fluids
For a salt with both cation and anion different from those of the mean salt (e.g.,
CuSO4 ), double substitution is required—i.e.,
( )
12
γ ± CuSO4 = γ Cu2 + γ SO 2 − ,
4
γ ±2 CuSO γ ±2 KCl
γ Cu2 + = 4
. (8.40)
γ ±3 K
2SO 4
Figure 8.9 shows mean activity values for various salts plotted according to
the ionic strength of the solution. Note how, at low I (I ⬍ 0.05), the values con-
verge for salts of the same stoichiometry (i.e., uni-univalent: NaCl, KCl: uni-
divalent: Na2SO4 , K2SO4 ).
Figure 8.10 shows the comparison between the mean salt method and the
Debye-Huckel model in the estimate of individual ionic activity coefficients. The
two methods give concordant results for low ionic strength values (I ⬍ 0.1). For
ions of the same net charge, convergence toward identical values occurs at zero
ionic strength. At high ionic strength, the individual ionic activity coefficients of
the cations increase exponentially with I. Estimates at I ⬎ 5 are affected by a high
degree of uncertainty.
Individual ionic activity coefficients may also be estimated more precisely
Figure 8.9 Mean activity coefficients for chlorides and sulfides, plotted following ionic
strength of solution. Reprinted from Garrels and Christ (1965), with kind permission from
Jones and Bartlett Publishers Inc., copyright 䉷 1990.
Geochemistry of Aqueous Phases / 499
Figure 8.10 Comparison between individual ionic activity coefficients obtained with
Debye-Hückel equation and with mean salt method for various ionic strength values. Re-
printed from Garrels and Christ (1965), with kind permission from Jones and Bartlett
Publishers Inc., copyright 䉷 1990.
AZ i2 I ˚ .
log γ i = − + BI ( 8.41)
1 + åi B I
Figure 8.11 shows activity coefficients for gaseous molecules dissolved in aqueous
NaCl solutions at various molalities (T ⫽ 15 and 25 °C; P ⫽ 1 bar). The coeffi-
cients converge on 1 at zero ionic strength. Each species also has distinct values,
at any I.
Figure 8.12 shows that, for a single neutral species (CO2 ), the activity coeffi-
cient varies according to type of dissolved electrolyte and increases with increas-
ing ionic strength of the solution. This phenomenon is known as “salting out,”
and generally takes place for nonpolar neutral species such as CO2 , O2 , H2 , and
N2 dissolved in aqueous solutions. For polar neutral species such as CaSO04 and
MgSO04, the inverse phenomenon, called “salting in,” is observed: in this case, the
500 / Geochemistry of Fluids
Figure 8.11 Activity coefficients for neutral gaseous molecules dissolved in aqueous solu-
tions of various ionic strengths. All values are for T ⫽ 25 °C and P ⫽ 1 bar, except for
hydrogen (T ⫽ 15 °C, P ⫽ 1 bar). Reprinted from Garrels and Christ (1965), with kind
permission from Jones and Bartlett Publishers Inc., copyright 䉷 1990.
Figure 8.12 Salting-out phenomenon for aqueous CO2. Activity coefficient of neutral
species increases with increasing salinity, determining decreased solubility of aqueous CO2
in water, T and P conditions being equal. Reprinted from Garrels and Christ (1965), with
kind permission from Jones and Bartlett Publishers Inc., copyright 䉷 1990.
Geochemistry of Aqueous Phases / 501
The activity of H2O solvent aH2O in aqueous solutions decreases with increasing
molality of solutes. Figure 8.13 shows the effects of progressive addition of dis-
solved chlorides in solution on aH2O.
In theory, once the activity of an electrolyte in solution is known, the activity
of the solvent can be determined by the Gibbs-Duhem integration (see section
2.11). In practice, the calculation is prohibitive, because of the chemical complex-
ity of most aqueous solutions of geochemical interest. Semiempirical approxima-
tions are therefore preferred, such as that proposed by Helgeson (1969), con-
sisting of a simulation of the properties of the H2O-NaCl system up to a solute
Figure 8.13 Activity of H2O solvent in aqueous solutions of various solute molalities.
Reprinted from Garrels and Christ (1965), with kind permission from Jones and Bartlett
Publishers Inc., copyright 䉷 1990.
502 / Geochemistry of Fluids
−2 Iϕ ′
log a H 2 O = , ( 8.43)
2.303ω H 2 O
where ⬘ is the osmotic coefficient of water and H2O is the number of moles of
H2O per 103 g of solvent ( H2O ⫽ 1000
18 ). The osmotic coefficient is then calculated
1 1 1
ϕ ′ = 1 − D′J ′ + W 2 I E + W 3 I E2 + W 4 I E3 , ( 8.44 )
2 3 4
where
2.303A
D′ = ( 8.45 )
W13 I E
1
J ′ = B ′ − 2 ln B ′ − ( 8.46 )
B′
B ′ = 1 + W1 I E1/2 , ( 8.47 )
8.9 Speciation
T
mCa = mCa2 + + mCaOH + + mCaCO0 + mCaHCO + + mCaHPO0 + L ( 8.48 )
3 3 4
∑ Z i m i = 0, ( 8.50)
i
CaOH+ ⇔ Ca 2 + + OH − ( 8.51)
HF 0 ⇔ H + + F − ( 8.52 )
AlF30 ⇔ Al 3 + + 3 F − . ( 8.53)
Because the thermodynamic constants of the various equilibria of the type exem-
plified by equations 8.51 to 8.53 can easily be calculated from the partial molal
504 / Geochemistry of Fluids
thermodynamic properties of the solutes, and the individual ionic activity coeffi-
cients may also be calculated, the unknowns in equations 8.54 to 8.56 are repre-
sented simply by partial molalities. When equations such as 8.48 and 8.49 are
combined with equations such as 8.54 to 8.56 and with equation 8.50, the system
is determined and can be solved by iterative procedures.
H H
/ /
→ M − O
M Z + O solvation
hydrolysis
→[M − O − H] ( Z −1) + + H + . (8.57)
\ \
H H
The stronger the polarizing power of the cation, the stronger the tendency of
reaction 8.57 to proceed rightward. Figure 8.14 shows the relationship between
polarizing power and the degree of hydrolysis of aqueous cations in a solution
with ionic strength I ⫽ 0.65, pH ⫽ 8.2, T ⫽ 25 °C, and P ⫽ 1 bar according to
Turner et al. (1981). In the logarithmic plot of figure 8.14, the ordinate axis is the
coefficient of partial reaction (i.e., ratio of hydrolyzed cation with respect to total
concentration). The abscissa axis does not properly indicate the polarizing power,
because authors have adopted Shannon and Prewitt’s (1969) ionic radii (see
Turner et al., 1981, for details).
Five elemental groups can be distinguished on the basis of a diagram such as
figure 8.14:
Observe the sharp transition from zero hydrolysis to complete hydrolysis in the
region 7 ⬍ Z2 /r ⬍ 25 Å⫺1. The fully hydrolyzed elements belong to groups IV
Geochemistry of Aqueous Phases / 505
Figure 8.14 Effect of polarizing power on degree of hydrolysis of cations in water. Re-
printed from D. R. Turner, M. Whitfield, and A. G. Dickson, Geochimica et Cosmochimica
Acta, 45, 855–881, copyright 䉷 1981, with kind permission from Elsevier Science Ltd.,
The Boulevard, Langford Lane, Kidlington 0X5 1GB, UK.
and V. For these elements (except for uranium, present as uranyl ion UO2⫹2 , whose
speciation is dominated by a strong interaction with carbonate ion CO2⫺ 3 ), com-
plexing with any other anionic ligands that may be present in aqueous solution is
negligible. Speciation of the remaining cations with the various anionic ligands is
related to their tendency to form partially covalent bondings. The arrangement
of the various groups on the periodic chart, according to the classification of
Turner et al. (1981), is shown in figure 8.15.
Ahrland’s (1973) classification distinguishes incompletely hydrolyzed ions
into types a and b. Cations of type a form stable complexes with electronic donors
of the first row of the periodic chart through electrostatic interactions; those of
506 / Geochemistry of Fluids
Figure 8.15 Arrangement of various ions from classification of Turner et al. (1981) on
periodic chart. Reprinted from D. R. Turner, M. Whitfield, and A. G. Dickson, Geochim-
ica et Cosmochimica Acta, 45, 855–881, copyright 䉷 1981, with kind permission from
Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington 0X5 1GB, UK.
type b form stable complexes with electronic donors of the second, third, and
fourth rows (P, S, Cl: As, Se, Br; Sb, Te, I) through essentially covalent bondings.
Turner et al. (1981) suggested a classification based on the different stabilities of
the chloride and fluoride complexes of the cations. Denoting the stability differ-
ence as ⌬K:
Type a cations: ⌬K ⬎ 2
Type a⬘ cations: 2 ⬎ ⌬K ⬎ 0
Type b cations: 0 ⬎ ⌬K ⬎ ⫺2
Type b⬘ cations: ⌬K ⬍ ⫺2.
Figure 8.16 shows how the complexing constants of ions vary with the polar-
izing power of the cation, ligand being equal. Note also here that the abscissa
Figure 8.16 Complexing constants of various cations with different anionic ligands ar-
ranged according to ratio between squared charge and ionic radius (plus a constant term),
where is the standard deviation on regression. Reprinted from D. R. Turner, M. Whit-
field, and A. G. Dickson, Geochimica et Cosmochimica Acta, 45, 855–881, copyright 䉷
1981, with kind permission from Elsevier Science Ltd., The Boulevard, Langford Lane,
Kidlington 0X5 1GB, UK.
508 / Geochemistry of Fluids
axis does not strictly represent polarizing power, because the effective radius is
replaced by the ionic radius plus a constant term.
We will now examine in detail the progression of hydrolysis reactions and
their influence on the stability in solution of hydrolyzed cations.
( ) ( )( )
Z+ ( Z −1 ) +
M OH 2 ⇔ M OH OH 2 + H+ ( 8.60)
n n −1
(Baes and Mesmer, 1981), where there is a proton loss by solvation of a water
molecule. In light of equation 8.60, it is not surprising to observe that the enthalpy
change associated with the process is close to the dissociation enthalpy of water
(13.3 kcal/mole at T ⫽ 25 °C and P ⫽ 1 bar).
If the metal concentration in solution is sufficiently low to remain below the
solubility product, increasing the pH, with molal concentration, T, and P re-
maining equal, hydrolysis proceeds by further solvation of H2O molecules, with
proton loss according to the scheme
( ) ( )
( Z −1 ) + (Z −2 )+
M Z + → M OH → M OH . ( 8.61)
2
( ) ( )
( Z −Y ) + ( Z −Y −1) +
M OH + H2 O → M OH + H+ , ( 8.62 )
Y (Y +1)
( ) (Y +1 ) [ ]
M OH ( Z −Y −1 ) + H +
K 62 = , ( 8.63)
( )
M OH ( Z −Y ) +
Y
Baes and Mesmer (1981) observed a simple relationship between the entropy change
involved in equilibrium 8.62 at reference T and P conditions ( ⌬S62,Tr ,Pr ) and the charge of
species in reaction (Z-Y)⫹:
cal
∆S62 , Tr , Pr = −17.8 + 12.2( Z − Y ) . ( 8.64 )
mole × K
Because
kcal
∆H 62 , Tr , Pr = −1.36 log K 62 , Tr , Pr − 5.3 + 3.64( Z − Y ) . ( 8.66 )
mole
( )
( XZ −Y ) +
X MZ + + Y H2O ⇔ MX OH + Y H+ , (8.67)
Y
K 67 = M X OH ( ) [ ] [M ]
H+
( XZ −Y ) + Y −X
Z+
. (8.68)
Y
( )
M OH
Z
+ Z H+ ⇔ MZ + + Z H2O (8.69)
[ ][ ]
−Z
K 69 = M Z + H + (8.70)
( ) + (XZ − Y ) H ( ) ( )
( XZ −Y ) +
X M OH +
⇔ MX OH + XZ − Y H2O (8.71)
Z Y
K 71 = M X OH ( ) [ ]
H+
( XZ −Y ) + − ( XZ −Y )
. (8.72)
Y
Figure 8.17 shows equilibrium relations between hydrolyzed cations and their
saturation limits as a function of pH. In this semilogarithmic plot, the saturation
510 / Geochemistry of Fluids
Figure 8.17 Equilibria in solution between various hydrolyzed species. From Baes and
Mesmer (1981), American Journal of Science, 281, 935–62. Reprinted with permission of
American Journal of Science.
limit defined by solubility product K71 is a straight line with slope XY-Z as a
function of pH (disregarding minor effects of the activity coefficients). This is
true for all species forming during the various hydrolysis steps with respect to
their saturation limits. The overall saturation curve (bold line) is given by the
envelope of the saturation limits valid for each single species. The predominance
limits of the hydrolyzed species (isoactivity condition of neighboring species) are
simply defined by pH (see eq. 8.59 and 8.62). At high metal concentrations, a
field of predominance of polynuclear species opens, whose boundaries are defined
by equilibria of the type exemplified by equation 8.67.
The above-discussed equilibria are of fundamental importance for an under-
standing of alteration processes, with special emphasis on the role of pH in water-
rock interaction events.
Besides controlling the redox conditions of seawater, as we will see in some detail
in section 8.10.1, carbonate equilibria are important when investigating sedimen-
tary rocks, of which carbonates are major components. The procedure normally
followed in geochemistry in establishing such equilibria is the classical approach
of Garrels and Christ (1965), later taken up by other authors (e.g., Helgeson,
1969; Holland, 1978; Stumm and Morgan, 1981). Five separate equilibria are
taken into consideration in this model, relating aqueous (a), gaseous ( g), and
condensed (c) phases:
Geochemistry of Aqueous Phases / 511
The first three equilibria relate the activities of the main aqueous species, the
fourth establishes the relationship between gaseous and aqueous phases, and the
fifth establishes the relationship between aqueous species and the condensed
phase.
The speciation state determined by the five equilibria depends on the type of
system (closed or open) and imposed conditions ( fCO2, pH). Two cases are partic-
ularly relevant in geochemistry:
Before we examine these two cases in detail, it is necessary to introduce some precau-
tionary considerations.
1. Equilibria are normally treated by considering calcite as the only carbonate poly-
morph present in the system. Actually, as already mentioned, CaCO3(c) crystal-
lizes in two forms: calcite (trigonal) and aragonite (orthorhombic). Aragonite,
which at low pressure is metastable with respect to the less dense calcite (cf. table
2.1), is in fact synthesized by several marine organisms that fix CaCO3(c) in this
form in their shells. Nevertheless, the difference in the Gibbs free energy of forma-
tion of the two polymorphs is so limited (0.2 to 0.5 kcal/mole) that the choice of
type of phase does not influence calculations.
2. Carbonates in sedimentary rocks are mixtures of several components: CaCO3 ,
CaMg(CO3)2 , FeCO3 , SrCO3 , and BaCO3 . Calculations are usually developed
assuming that the condensed phase is pure calcite and assigning unitary activity
to the CaCO3 component. For accurate calculations, we would introduce an ap-
propriate activity value, depending on the actual chemistry of the solid and bear-
ing in mind that carbonates are not at all ideal mixtures but have wide miscibility
gaps (cf. figure 3.11).
512 / Geochemistry of Fluids
3. Aqueous solutions of carbon dioxide contain both CO2(a) and H2CO3(a) neutral
molecules. The CO2(a) molecule has a linear structure (O⫽C⫽O), whereas
that of H2CO3(a) is planar trigonal because of the molecular orbitals constituted
of hybrid AO sp2. Conversion from the linear form CO2(a) to the trigonal-planar
form H2CO3(a) —i.e.,
—is relatively slow (about 3% of CO2(a) is converted into H2CO3(a) per second in
room conditions; cf. Kern, 1960). The dehydration process
is relatively faster (net rate constant Kr ⫽ 2.0 ⫻ 10⫺3 sec⫺1 at 0 °C, according to
Jones et al., 1964).
The equilibrium kinetics of reactions 8.78 and 8.79 show that, at room P and T
and in stationary conditions, the CO2(a) /H2CO3(a) ratio is near 600 (Kern, 1960).
Notwithstanding this fact, in geochemistry it is customary to represent aqueous
CO2 as if it were entirely in the form of carbonic acid H2CO3(a), not distinguishing
planar from trigonal forms (i.e., H2CO* 3(a) in eq. 8.74). This has little connection
with reality and may lead to errors unless all equilibria are explicitly defined ac-
cording to a single value of the Gibbs free energy of formation for aqueous neu-
tral species.
The amount of H2CO* 3(a) may then be determined through equation 8.74.
Table 8.6 lists equilibrium constants for reactions 8.73, 8.77, 8.80, and 8.81
along the water-vapor univariant curve. We use values of T ⫽ 25 °C and P ⫽ 1
bar in the following calculations.
Let us equilibrate some pure calcite crystals with an aqueous phase at T ⫽
25 °C and P ⫽ 1 bar. Calcite partially dissolves to reach the solubility product
(eq. 8.77).
Geochemistry of Aqueous Phases / 513
Table 8.6 Constants of carbonate equilibria at various P and T conditions along univariant
water-vapor boundary
P (bar) T (°C) log K77 log K77(*) log K73 log K80 log K81
1 25 ⫺8.616 ⫺8.451 10.329 ⫺7.818 ⫺18.147
0.006 0 ⫺8.485 ⫺8.317 10.553 ⫺7.683 ⫺18.236
0.123 50 ⫺8.771 ⫺8.608 10.123 ⫺7.988 ⫺18.111
1.013 100 ⫺9.072 ⫺8.912 9.759 ⫺8.385 ⫺18.144
4.758 150 ⫺9.359 ⫺9.198 9.434 ⫺8.831 ⫺18.265
15.537 200 ⫺9.652 ⫺9.487 9.115 ⫺9.323 ⫺18.438
39.728 250 ⫺9.987 ⫺9.817 8.751 ⫺9.890 ⫺18.641
85.805 300 ⫺10.459 ⫺10.282 8.235 ⫺10.626 ⫺18.861
165.125 350 ⫺11.480 ⫺11.297 7.117 ⫺11.973 ⫺19.090
( )
* With aragonite as condensed phase at equilibrium.
[Ca ][CO ] = K
2+ 2−
3 77 = 10 −8.62. (8.82)
Based on equations 8.80 and 8.81 and assuming unitary fugacity for CO2(g) at
standard state (see section 9.1), we can write
[ HCO ][ H ] ⋅ f
−
3
+ −1
CO 2( g )
= K 80 = 10 −7.82 (8.83)
[CO ][ H ]
2
2−
3
+
⋅ fCO
−1
= K 81 = 10 −18.15 (8.84)
2( g )
[ HCO ][ H ] [CO ]
−1 −1
−
3
+ 2−
3 = K 73 = 1010.33. (8.85)
We also recall that the ionic dissociation constant of pure water under room con-
ditions is
[ H ][OH ] = 10
+ − −14
. (8.86)
Let us now express the activities of the main ionic species in solution as a
function of the activity of hydrogen ions and f CO2(g). Combining equations 8.82
to 8.86, we obtain
[Ca ] = 10 [ H ]
2
2+ 9.53 +
⋅ fCO
−1
(8.87)
2( g )
514 / Geochemistry of Fluids
[ HCO ] = 10 [ H ]
−1
−
3
−7.82 +
⋅ fCO 2( g ) (8.88)
[CO ] = 10 [H ]
−2
2−
3
−18.15 +
⋅ fCO 2( g ) (8.89)
[OH ] = 10 [H ]
−1
− −14.00 +
. (8.90)
[ ] [ ] [ ] [
2 Ca 2 + + H + = 2 CO 23 − + HCO −3 + OH − . ] [ ] (8.92)
[ ] + [H ]
4 3
10 9.83 ⋅ f CO
−1
⋅ H+ +
2( g )
(8.93)
( )[ ]
− 10 −7.82 ⋅ f CO2 ( g ) + 10 −14 H + − 10 −17.85 ⋅ f CO2 ( g ) = 0.
Figure 8.18 shows how equation 8.93 behaves in the vicinity of zero. Equilibrium
pH corresponds to the zero value on the ordinate axis.
If we adopt the present-day atmospheric f CO2(g) value (10⫺3.48 ), the corre-
sponding pH, obtained through application of equation 8.93, is 8.2. It is well
known that the amount of atmospheric CO2 is steadily increasing, as a result
of intensive combustion of fossil fuels. It has been deduced, from present-day
exploitation rates and projections, that the amount of CO2(g) in the atmosphere
will double in the next 40 to 70 years (i.e., fCO2(g) will increase from 10⫺3.48 to
10⫺3.17 ). Correspondingly, the pH will fall by about 0.2 unit.
Equation 8.93 is asymptotic, with a flat flexus near zero (figure 8.18). Obviously, this
type of equation does not allow precise pH determinations, because even slight approxi-
Geochemistry of Aqueous Phases / 515
Figure 8.18 Relationship between pH of aqueous solution and fCO2 of gaseous phase in
equilibrium with water and calcite. Equilibrium pH is in correspondance with zero on
ordinate axis.
mations in the equilibrium constants generate large variations in the solving value (the
small pH differences between the procedure of Garrels and Christ, 1965, and the one
developed here at quasi-parity of fCO2(g) must be ascribed to this sort of problem). The
method therefore cannot be used for high ionic strength or chemically complex solutions,
in which the electroneutrality condition cannot be simplified in the form of equation 8.91.
In seawater, for instance, only 88% of calcium is in the form of Ca2⫹ and 12% is in the
aqueous sulfate form of CaSO04. Similarly, the bicarbonate ion constitutes only 73% of
total carbonate in solution: 13% is present as magnesium bicarbonate (MgHCO3⫹ ), 5.5%
as magnesium carbonate (MgCO03 ), 2.8% as calcium bicarbonate (CaHCO3⫹ ), and only
2.1% as carbonate ion. Clearly, for seawater the carbonate equilibria of equations
8.73→8.77 constitute only a particular subsystem of the n equilibria that actually deter-
mine the speciation state.
Nevertheless, it is of interest to note that precise speciation calculation conducted
with an automated routine clearly indicates that, at the pH observed in seawater (8.22),
the activities of Ca2⫹ and CO32⫺ in solution are near the solubility product of calcite and
fCO2(g) is near the atmospheric value (i.e., fCO2(g) ⫽ 10⫺3.35 ). We can thus deduce that car-
bonate equilibria are effectively determinant in seawater and that they are conditioned by
the atmospheric value of fCO2(g). Table 8.7 shows that the increase in CO2 in the atmo-
sphere determines a progressive decrease of pH in seawater, in line with what was pre-
viously deduced for an aqueous solution of low salinity. In light of this marked control by
atmospheric CO2 on the acidity of aqueous solutions, during geochemical samplings of
deep waters, direct contact with the atmosphere must be avoided before pH measurements
are taken, because otherwise the pH turns out to be intermediate between the in situ value
and the new equilibrium situation. Note here that, according to the calculation of Garrels
and Christ (1965), the pH of an aqueous solution in equilibrium with calcite at T ⫽ 25
516 / Geochemistry of Fluids
A B C
°C and P ⫽ 1 bar and in the absence of atmospheric CO2 is 9.9—i.e., much higher than
the value obtained for equilibrium with atmosphere (cf. table 8.7).
with
γH * ⋅ mH *
2 CO 3 2 CO 3
K 74 = = 10 6.4. (8.96)
γ H + ⋅ mH + ⋅ γ HCO− ⋅ mHCO−
3 3
With
we obtain
m HCO− ⋅ γ HCO−3
K 73 = 3
= 1010.33 . (8.98)
γ H+
⋅ m H+ ⋅ γ CO 23 −
⋅ m CO 2 −
3
Table 8.8 Compositions of superficial and deep waters of the Sarcidano region (Sardinia, Italy)
equilibriated with Mesozoic dolomite limestones (Bertorino et al., 1981). Values in mEq/l. CT total
inorganic carbon in mmol/l.
With ␥CO2⫺
3
⫽ 0.65, we obtain
From the equation of total inorganic carbon (eq. 8.94), combined with equa-
tions 8.97 and 8.99, we derive
Thus, mHCO⫺ 3
⫽ 10⫺2.139, mH2CO*3 ⫽ 10⫺3.079, and mCO2⫺
3
⫽ 10⫺5.027.
From the total molality of Ca, assuming all calcium to be present in solution
in the ionic form Ca2⫹, we obtain
γ Ca 2 +
⋅ m Ca 2 + = 0.66 × 0.00175 = 10 −2.937 . ( 8.101)
[ ][ ]
Q = Ca 2 + CO 23 − = 0.65 × 10 −5.027 × 10 −2.937 = 10 −8.151. (8.102)
Q kcal
AP = RT ln = 0.640 . (8.103)
K mole
Table 8.9 shows the results of detailed speciation calculations carried out with
the automated routine EQ3NR/EQ6 (Wolery, 1983). The molalities of the ions
in solution resulting from this calculation are tabulated in order of decreasing
abundance.
In the total inorganic carbon balance of equation 8.94, we neglected the presence of
some species. Table 8.9 shows that this approximation is reasonable although, for accurate
calculations, we should also have considered bicarbonates CaHCO3⫹ and MgHCO3⫹ and
the soluble carbonate CaCO30. Note also that sulfates occur in negligible molar propor-
tions and that carbonate alkalinity coincides with total alkalinity.
The precise speciation calculation indicates that four minerals are at almost perfect
saturation:
Our approximation in the calculation is 150 cal/mole, which is quite satisfactory. The fact
that dolomite is the mineral phase nearest to complete equilibrium is in agreement with
the geology of the area and was in fact already evident at first sight from the chemical
analyses, indicating similar molalities for Ca and Mg in solution. The resulting fCO2(g) is
10⫺1.67, which is much higher than the atmospheric value. The previous recommendation
about pH measurements is particularly obvious in this case.
reaches a carbonate layer: this is clearly a combination of the two cases discussed
in this section and in section 8.10.1. Calculations on a semiquantitative basis
in such a case have only academic interest (see Garrels and Christ, 1965, for a
detailed treatment).
Study of water-rock interactions occurring within the continental crust and along
mid-ocean ridges, where large masses of water are involved in convective trans-
port within the lithosphere, requires accurate estimates of the partial molal prop-
erties of solutes under high P and T conditions. In other words, in order to deter-
mine the reactivity of a given aqueous solution toward a given mineralogical
assemblage with which it comes into contact, a detailed speciation calculation at
the P and T of interest is necessary. As we have already seen in section 8.9, here
we need knowledge of the various equilibrium constants among solutes, and these
constants, to be determined, in turn require knowledge of the partial molal Gibbs
free energies of the solutes at the P and T of interest. The energy model developed
by Helgeson, Kirkham, and Flowers (1981) and later revised by Tanger and Hel-
geson (1988) allows accurate calculation of the partial molal properties of ions in
aqueous solutions up to a maximum P of 5 kbar and a maximum T of 1000 °C.
A brief outline of the procedure is given here (see the works of Helgeson and co-
authors for more exhaustive explanations).
molar property at any particular composition of the phase in the system. Partial molar
properties are related, in the terms outlined by equations 2.31 to 2.34 and 2.1, to the Gibbs
free energy of the phase (and of the system). The significance of the partial molal proper-
ties of solutes in aqueous solutions are obviously similar, although some distinction must
be made:
1. These properties are referred to the molality of the solute (i.e., moles of solute
per 103 g of solvent) and not to molarity.
2. “Absolute” properties are distinguished from “conventional” properties; i.e., a
generic standard partial molal property of an aqueous ion j ( ⌶ 0j ) is related to the
absolute property by
where Zj is the formal charge of the jth aqueous ion and ⌶ 0abs
H⫹ is the correspond-
ing absolute property of the hydrogen ion.
3. An “additivity principle” is adopted (Millero, 1972), according to which the con-
ventional standard partial molal property of a generic electrolyte K is related to
the absolute properties of its constituent ions through
Ξ K0 = ∑v j, K Ξ 0j abs ,
(8.105)
j
Ξ K0 = ∑v j, K Ξ j0 .
(8.106)
j
Because the conventional standard partial molal properties of a hydrogen ion is zero, it
follows that the conventional standard partial molal properties of a generic anion A are
identical to the experimental values of the corresponding acid electrolyte. Moreover, based
on equation 8.104, the standard partial molal properties of a generic cation C can be
calculated, once the experimental values for aqueous electrolytes Hv⫹A and Cv⫹Av⫺ are
known. Note here the close analogy between the additivity principle and the mean salt
method, previously discussed as a method of calculating ionic activity coefficients (sec-
tion 8.6).
The electroconstriction contribution derives from the structural collapse of the solvent in
the immediate neighborhood of the ion ( ⌬⌶ 0abs c, j ) and by the ion-solvation process
( ⌬⌶ 0abs
s, j ) (electrostatic solute-solvent interactions):
∆Ξ e0,abs
j = ∆Ξ c0,abs
j + ∆Ξ s0, abs
j . (8.108)
Equations 8.107 and 8.108, referring to absolute properties, are equally valid for conven-
tional properties based on equations 8.105 and 8.106. Moreover, the additivity principle
is applied to all partial contributions: thus, for a generic electrolyte K, we have
∆Ξ e0, K = ∑v j, K ∆Ξ e0, j ,
(8.109)
j
∆Ξ c0, K = ∑v j, K ∆Ξ c0, j ,
(8.110)
j
∆Ξ s0, K = ∑v j, K ∆Ξ s0, j ,
(8.111)
j
and so on.
The HKF model groups the intrinsic properties of the solute and the effects of struc-
tural collapse of the solvent into a single term, which is defined as the contribution of
“nonsolvation”:
∆Ξ n0,abs
j = Ξ i0, abs
j + ∆Ξ c0,abs
j . (8.112)
The partial molal property of the solute is thus composed of one nonsolvation term
( ⌬⌶ 0abs
n, j ) plus one solvation ( ⌬⌶ s, j ) term; i.e., for a generic ion j,
0abs
Ξ 0j abs = ∆Ξ n0,abs
j + ∆Ξ s0, abs
j (8.113)
structure of the solvent (section 8.5). We must now introduce the concept of effective
electrostatic radius. For this purpose, we recall that the Gibbs free energy change ⌬G 0abs
s, j
involved in the solvation of a generic ion j in solution is defined by the Born equation
1
∆Gs0, abs = ω abs
j − 1 , (8.116)
j
ε
(Born, 1920), where is the dielectric constant of the solvent and abs
j is the absolute Born
coefficient for the jth ion in solution:
N 0 e 2 Z 2j
ω abs
j = , (8.117)
2re , j
where N 0 is Avogadro’s number, e is the elementary charge, Zj is the formal charge of ion
j, and re, j is its electrostatic radius. Posing
1 Å ⋅ cal
η= N 0 e 2 = 1.66027 × 10 5 , (8.118)
2 mole
ηZ 2j
ω abs
j = . (8.119)
re , j
1
∆Gs0, j = ω j − 1 . (8.120)
ε
ω j = ω abs
j − Z j ω abs+ , (8.121)
H
where abs
H⫹ is the absolute Born coefficient of the hydrogen ion at the P and T of interest
( abs
H ⫹ ⫽ 0.5387 ⫻ 105 cal/mole at T ⫽ 25 °C and P ⫽ 1 bar; cf. Helgeson and Kirkham,
1974).
Stemming from the general relations developed in the preceding section, the solvation
energy of a generic electrolyte K can be related to that of its constituting ions j through
the Born coefficient K (Tanger and Helgeson, 1988):
v j , K ηZ 2j
ωK = ∑v j, K ω j
abs
= ∑ re , j
= ∑v j ,K
ωj (8.122)
j j j
524 / Geochemistry of Fluids
1
∆Gs0, K = ω K − 1 . (8.123)
ε
re , j = rx , j + Z j Γ ± , (8.124)
Γ ± = K Z + g. ( 8.125)
KZ in equation 8.125 represents a charge constant that is zero for anions and 0.94 Å for
cations, and g is a solvent parameter dependent on P and T (see equations 36A, B, and C
in Tanger and Helgeson, 1988).
Table 8.10 lists electrostatic radii of aqueous ions at T ⫽ 25 °C and P ⫽ 1 bar (Shock
and Helgeson, 1988).
V 0
= ∆Vn 0 + ∆Vs 0 . ( 8.126 )
Partial molal volumes can be related to the corresponding Gibbs free energy terms
through the partial derivatives on P (see equations 2.28 and 2.33).
∂∆Gn0
∆Vn 0 = (8.127)
∂P T
∂∆Gs 0
∆Vs 0 = . (8.128)
∂P T
ω ∂ ln ε 1 ∂ω
∆Vs0 = − + − 1 . (8.129)
ε ∂P T ε ∂P T
1
∆Vn 0 = σ + ξ , (8.130)
T − θ
where is a solvent constant that, for water, assumes the value ⫽ 228 K (singular tem-
perature of supercooled water; Angell, 1982), and and are coefficients not dependent
on T but variable with P, based on
1
σ = a1 + a2 (8.131)
ψ + P
1
ξ = a3 + a4 . (8.132)
ψ + P
In equations 8.131 and 8.132, is another solvent constant that, for water, has the value
⫽ 2600 bar, and a1 , a2 , a3 , and a4 are characteristic parameters of ions and electrolytes
in solution. These parameters, constant over P and T, are listed in table 8.11.
Combining equations 8.126 to 8.132, the standard partial molal volume can be ex-
pressed as
1 1 1
0
= a1 + a2 + a + a4
ψ + P 3 ψ + P T − θ
V
(8.133)
1 ∂ω
− ωQ + − 1 ,
ε ∂P T
1 ∂ ln ε
Q= . (8.134)
ε ∂P T
∂∆Vn 0 ∂∆Vs 0
−β 0 = + . (8.135)
∂P T ∂P T
Adopting the same parameters previously used for calculating the standard partial molal
volume, compressibility can be expressed as
Table 8.11 Volume parameters of HKF model (from Shock
and Helgeson, 1988). a1 ⫽ cal/(mole ⫻ bar); a2 ⫽ cal/mole;
a3 ⫽ (cal ⫻ K)/(mole ⫻ bar); a4 ⫽ (cal ⫻ K)/mole
2
1 1 ∂Q
β = a 2 + a 4
0
+ ω
T − θ ψ + P ∂P T
(8.136)
∂ω 1 ∂ 2ω
+ 2Q − − 1 2 .
∂P T ε ∂P T
The thermal expansion of a generic ion or electrolyte in solution is given by the varia-
tion with T of the nonsolvation and solvation terms:
∂∆Vn 0 ∂∆Vs 0
α0 = + . (8.137)
∂T P ∂T P
530 / Geochemistry of Fluids
Adopting the same parameters previously used for calculating the standard partial molal
volume (eq. 8.133) and compressibility (eq. 8.136), thermal expansion can be expressed as
2
1 1 ∂Q ∂ω
α 0 = − a 3 + a 4 − ω − Q
ψ + P T − θ ∂T P ∂T P
1 ∂ (∂ω ∂P ) T
(8.138)
∂ω
−Y + − 1 ,
∂P T ε ∂T P
1 ∂ ln ε
Y = . (8.139)
ε ∂T P
The variation of ⌬C 0P,n with T at reference pressure has an asymptotic form that can be
described by two coefficients, constant over T and P and characteristic for each ion and
electrolyte in solution (see table 8.12):
2
1
∆C P0, n = c1 + c2 . (8.141)
T − θ
2
1
∆C P0, n = c1 + c2
T − θ
(8.142)
1 ψ + P
3
− 2T a 3 (P − Pr ) + a 4 ln .
T − θ ψ + Pr
∂ω 1 ∂ 2ω
∆C P0, s = ωTX + 2TY − T − 1 2 , (8.143)
∂T P ε ∂T P
Table 8.12 Standard state heat capacity (Tr ⫽ 25 °C, Pr ⫽ 1 bar),
coefficients of heat capacity, and Born coefficient in HKF model
0
(from Shock and Helgeson, 1988). CP ⫽ cal/(mole ⫻ K); c1 ⫽ cal/
(mole ⫻ K); c2 ⫽ (cal ⫻ K)/mole; ⫽ cal/mole.
c2 ⫻ 10⫺4 ⫻ 10⫺5
0
Ion CP c1
H⫹ 0 0 0 0
Li⫹ 14.2 19.2 ⫺0.24 0.4862
Na⫹ 9.06 18.18 ⫺2.981 0.3306
K⫹ 1.98 7.40 ⫺1.791 0.1927
Rb⫹ ⫺3.0 5.7923 ⫺3.6457 0.1502
Cs⫹ ⫺6.29 6.27 ⫺5.736 0.0974
NH4⫹ 15.74 17.45 ⫺0.021 0.1502
Ag⫹ 7.9 12.7862 ⫺1.4254 0.2160
Tl⫹ ⫺4.2 5.0890 ⫺3.8901 0.1502
VO2⫹ 31.1 30.8449 3.3005 0.7003
Cu⫹ 13.7 17.2831 ⫺0.2439 0.3351
Au⫹ 0.6 8.1768 ⫺2.9124 0.1800
Mg2⫹ ⫺5.34 20.80 ⫺5.892 1.5372
Ca2⫹ ⫺7.53 9.00 ⫺2.522 1.2366
Sr2⫹ ⫺10.05 10.7452 ⫺5.0818 1.1363
Ba2⫹ ⫺12.30 3.80 ⫺3.450 0.9850
Pb2⫹ ⫺12.70 8.6624 ⫺5.6216 1.0788
Mn2⫹ ⫺4.1 16.6674 ⫺3.8698 1.4006
Co2⫹ ⫺7.8 15.2014 ⫺4.6235 1.4769
Ni2⫹ ⫺11.7 13.1905 ⫺5.4179 1.5067
Cu2⫹ ⫺5.7 20.3 ⫺4.39 1.4769
Zn2⫹ ⫺6.3 15.9009 ⫺4.3179 1.4574
Cd2⫹ ⫺3.5 15.6573 ⫺3.7476 1.2528
Be2⫹ ⫺1.3 22.9152 ⫺3.2994 1.9007
Ra2⫹ ⫺14.4 6.2858 ⫺5.9679 0.9290
Sn2⫹ ⫺11.2 11.4502 ⫺5.3160 1.2860
VO2⫹ ⫺12.0 13.3910 ⫺5.4790 1.5475
Fe2⫹ ⫺7.9 14.9632 ⫺4.6438 1.4574
Pd2⫹ ⫺6.3 15.3780 ⫺4.3179 1.4006
Ag2⫹ ⫺5.3 15.2224 ⫺4.1142 1.3201
Hg2⫹ 2.2 18.0613 ⫺2.5865 1.151
Hg22⫹ 17.1 23.7433 0.4487 0.8201
Sm2⫹ 3.7 19.6533 ⫺2.2809 1.2285
Eu2⫹ 6.0 19.7516 ⫺1.8124 1.0929
Yb2⫹ 0.7 17.8951 ⫺2.8920 1.2285
Al3⫹ ⫺32.5 10.7 ⫺8.06 2.8711
La3⫹ ⫺37.2 4.2394 ⫺10.6122 2.1572
Ce3⫹ ⫺38.6 4.0445 ⫺10.8974 2.2251
Pr3⫹ ⫺47.7 ⫺1.1975 ⫺12.7511 2.2350
Nd3⫹ ⫺43.2 1.6236 ⫺11.8344 2.2550
Sm3⫹ ⫺43.3 1.9385 ⫺11.8548 2.2955
Eu3⫹ ⫺36.6 6.0548 ⫺10.4900 2.3161
Gd3⫹ ⫺35.9 6.5606 ⫺10.3474 2.3265
Tb3⫹ ⫺40.5 4.2522 ⫺11.2844 2.3685
Dy3⫹ ⫺31.7 9.5076 ⫺9.4919 2.3792
Ho3⫹ ⫺33.3 8.6686 ⫺9.8178 2.3899
continued
Table 8.12 (continued)
c2 ⫻ 10⫺4 ⫻ 10⫺5
0
Ion CP c1
Er3⫹ ⫺34.3 8.2815 ⫺10.0215 2.4115
Tm3⫹ ⫺34.3 8.4826 ⫺10.0215 2.4333
Yb3⫹ ⫺36.4 7.3533 ⫺10.4493 2.4443
Lu3⫹ ⫺32.0 9.5650 ⫺9.7160 2.4554
Ga3⫹ ⫺30.8 13.2451 ⫺9.3086 2.7276
In3⫹ ⫺34.9 8.7476 ⫺10.1437 2.5003
Tl3⫹ ⫺39.3 4.7607 ⫺11.0400 2.3474
Fe3⫹ ⫺34.1 11.0798 ⫺9.9808 2.7025
Co3⫹ ⫺32.3 12.2500 ⫺9.6141 2.7150
Au3⫹ ⫺36.2 7.5724 ⫺10.4085 2.4554
Sc3⫹ ⫺35.4 8.4546 ⫺10.2456 2.5003
Y3⫹ ⫺35.7 7.1634 ⫺10.3067 2.3792
AlO2⫺ ⫺11.9 19.1 ⫺6.2 1.7595
HCO3⫺ ⫺8.46 12.9395 ⫺4.7579 1.2733
NO3⫺ ⫺16.4 7.70 ⫺6.725 1.0977
NO2⫺ ⫺23.3 3.4260 ⫺7.7808 1.1847
H2PO4⫺ ⫺7.0 14.0435 ⫺4.4605 1.3003
OH⫺ ⫺32.79 4.15 ⫺10.346 1.7246
HS⫺ ⫺22.17 3.42 ⫺6.27 1.4410
HSO3⫺ ⫺1.4 15.6949 ⫺3.3198 1.1233
HSO4⫺ 5.3 20.0961 ⫺1.9550 1.1748
F⫺ ⫺27.23 4.46 ⫺7.488 1.7870
Cl⫺ ⫺29.44 ⫺4.40 ⫺5.714 1.4560
ClO3⫺ ⫺12.3 8.5561 ⫺5.5401 1.0418
ClO4⫺ ⫺5.5 22.3 ⫺8.9 0.9699
Br⫺ ⫺30.42 ⫺3.80 ⫺6.811 1.3858
BrO3⫺ ⫺20.6 3.7059 ⫺7.2308 1.0433
I⫺ ⫺28.25 ⫺6.27 ⫺4.944 1.2934
IO3⫺ ⫺16.2 7.7293 ⫺6.3345 1.2002
MnO4⫺ ⫺1.8 13.7427 ⫺3.4013 0.9368
ReO4⫺ ⫺0.2 14.3448 ⫺3.0753 0.9004
BO2⫺ ⫺41.0 ⫺1.6521 ⫺11.3863 1.7595
BF4⫺ ⫺5.6 11.8941 ⫺4.1753 0.9779
CN⫺ ⫺32.7 ⫺1.1135 ⫺9.6956 1.2900
H2AsO4⫺ ⫺0.7 16.8622 ⫺3.1772 1.2055
H2AsO3⫺ ⫺2.9 15.8032 ⫺3.6253 1.2305
HO2⫺ ⫺30.2 2.7007 ⫺9.1863 1.5449
HSO5⫺ 29.2 31.2126 2.9134 0.8611
HSe⫺ ⫺12.6 11.1345 ⫺5.6012 1.3408
HSeO3⫺ 4.9 19.5432 ⫺2.0365 1.1402
HSeO4⫺ ⫺41.9 ⫺8.3616 ⫺11.5696 1.0885
HF2⫺ ⫺33.2 ⫺1.3751 ⫺9.7974 1.2934
ClO⫺ ⫺49.2 ⫺9.0630 ⫺13.0566 1.4767
ClO2⫺ ⫺30.5 ⫺0.0659 ⫺9.2474 1.2637
BrO⫺ ⫺49.2 ⫺9.0630 ⫺13.0566 1.4767
BrO4⫺ 0.6 14.8727 ⫺2.9124 0.9068
Br3⫺ 5.6 17.2705 ⫺1.8939 0.8490
IO⫺ ⫺64.1 ⫺16.2398 ⫺16.0918 1.6455
IO4⫺ 7.6 18.2345 ⫺1.4865 0.8264
I3⫺ 13.1 20.8712 ⫺0.3661 0.7628
Geochemistry of Aqueous Phases / 533
c2 ⫻ 10⫺4 ⫻ 10⫺5
0
Ion CP c1
H2VO4⫺ 0.6 17.4795 ⫺2.9124 1.1898
HCrO4⫺ 20.4 26.9872 1.1209 0.9622
H3P2O7⫺ 29.6 31.4072 2.9949 0.8568
CO32⫺ ⫺69.5 ⫺3.3206 ⫺17.1917 3.3914
HPO42⫺ ⫺58.3 2.7357 ⫺14.9103 3.3363
SO42⫺ ⫺64.38 1.64 ⫺17.998 3.1463
S2O32⫺ ⫺57.3 ⫺0.0577 ⫺14.7066 2.9694
S2O82⫺ ⫺25.0 12.9632 ⫺8.1271 2.3281
CrO42⫺ ⫺59.9 ⫺1.0175 ⫺15.2362 3.0307
MoO42⫺ ⫺47.5 7.0224 ⫺12.7103 3.1145
WO42⫺ ⫺44.5 8.3311 ⫺12.0992 3.0657
BeO22⫺ ⫺100.0 ⫺18.0684 ⫺23.5879 3.7880
SiF62⫺ ⫺47.1 4.0970 ⫺12.6289 2.7716
H2P2O72⫺ ⫺20.3 18.4241 ⫺7.1697 2.6218
HAsO42⫺ ⫺51.8 5.4710 ⫺13.5863 3.2197
SO32⫺ ⫺76.1 ⫺7.8368 ⫺18.5362 3.3210
S22⫺ ⫺65.1 ⫺3.3496 ⫺16.2955 3.1083
S2O42⫺ ⫺52.9 1.6707 ⫺13.8103 2.8772
S2O52⫺ ⫺50.5 2.6824 ⫺13.3215 2.8343
S2O62⫺ ⫺46.5 4.3301 ⫺12.5066 2.7587
S32⫺ ⫺57.9 ⫺0.3595 ⫺14.8288 2.9749
S3O62⫺ ⫺44.1 5.3169 ⫺12.0178 2.7131
S42⫺ ⫺50.7 2.6081 ⫺13.3622 2.8390
S4O62⫺ ⫺21.3 14.6933 ⫺7.3734 2.2805
S52⫺ ⫺43.6 5.5361 ⫺11.9159 2.7051
S5O62⫺ ⫺38.5 7.6266 ⫺10.8770 2.6076
SeO32⫺ ⫺68.1 ⫺4.5783 ⫺16.9066 3.1658
SeO42⫺ ⫺60.2 ⫺1.2986 ⫺15.2973 3.0192
HVO42⫺ ⫺48.3 6.9055 ⫺12.8733 3.1527
Cr2O72⫺ ⫺20.4 15.0820 ⫺7.1901 2.2654
MnO42⫺ ⫺59.3 ⫺0.9267 ⫺15.1140 3.0024
PO43⫺ ⫺114.9 ⫺9.4750 ⫺26.4397 5.6114
1 ∂ 2 ln ε ∂ ln ε
2
X = − (8.144)
ε ∂ T 2 P ∂ T P
(Helgeson and Kirkham, 1974). Summing equations 8.142 and 8.143, we obtain
c2
2T
C P0 = c1 + − ×
(T − θ ) (T − θ ) 3
2
(8.145)
ψ + P ∂ω 1 ∂ 2ω
( )
a3 P − Pr + a 4 ln + ωTX + 2TY
ψ + Pr
− T − 1
∂T P ε
.
∂T 2 P
534 / Geochemistry of Fluids
1 ∂ω
∆S s0 = ωY − − 1 , (8.146)
ε ∂T P
T
c2 1 1 1 Tr (T − θ )
∆S n0 = c1 ln − − + ln
Tr θ T − θ Tr − θ θ T (Tr − θ )
(8.147)
1 ψ + P
2
+ a (P − Pr ) + a 4 ln
T − θ 3 .
ψ + Pr
Combining the solvation and nonsolvation contributions, the partial molal entropy
change between reference state conditions Pr and Tr and the P and T of interest is given by
r r
(
S P0, T − S P0 , T = ∆S n0 − ∆S n0, P , T
r r
) + ( ∆S s
0
− ∆S s0, P
r , Tr
)
T c2 1 1 1 Tr (T − θ )
= c1 ln − + + ln
Tr θ T − θ Tr − θ θ T (Tr − θ )
(8.148)
1 ψ + P
2
+ a 3 (P − Pr ) + a 4 ln
T − θ ψ + Pr
1 ∂ω
+ ωY − − 1 − ω Pr , TrY Pr , Tr .
ε ∂T P
∂V 0
T P
H P0, T = H P0 , T +
r r ∫ C P0 dT +
r ∫ V
0
− T dP
∂T P (8.149)
Tr Pr T
and
GP0, T = GP0 , T − S P0 , T T − Tr
r r r r
( )
T T P
(8.150)
+ ∫ C P0 dT − T ∫ C P0 d ln T +
r r ∫V 0
dP.
Tr Tr Pr
Geochemistry of Aqueous Phases / 535
Note that thermodynamic tabulations do not normally report the standard partial
molar properties of solutes H 0Tr , Pr and G 0Tr , Pr, but rather the enthalpy of formation
H 0f, T r , Pr and the Gibbs free energy of formation from the elements at stable state under Tr
and Pr reference conditions G 0f, T r , Pr (note, however, that H 0T r , Pr and H 0f,T r ,Pr coincide,
because the enthalpy of elements at stable state and Tr ⫽ 25 °C and Pr ⫽ 1 bar is conven-
tionally set at zero). Because it is also true that
and
1 1
H f0, P , T = H f0, P , T + c1 T − Tr − c2 ( ) − + a 1 (P − Pr )
T − θ Tr − θ
r r
ψ + P 2T − θ
a 3 (P − Pr ) + a 4 ln ψ + P
+ a 2 ln +
ψ + Pr (T − θ ) 2 ψ + Pr
(8.153)
1 1 ∂ω 1
+ ω − 1 + ωTY − T − 1 − ω Pr , Tr − 1
ε ε ∂T P ε Pr , Tr
− ω Pr , Tr TrY Pr , Tr
and
T
(
G f0, P , T = G f0, P , T − S P0 , T T − Tr − c1 T ln ) − T + Tr
Tr
r r r r
ψ +P 1 1 θ − T
(
+ a 1 P − Pr + a 2 ln ) − c2 −
ψ + Pr T − θ Tr − θ θ
Tr (T − θ ) 1
(8.154)
ψ + P
a 3 (P − Pr ) + a 4 ln
T
− ln +
θ 2 T (Tr − θ ) T − θ ψ + Pr
1 1
+ ω − 1 − ω Pr , Tr − 1 + ω Pr , TrY Pr , Tr (T − Tr ).
ε ε Pr , Tr
Table 8.13 lists G 0f, Pr ,Tr , H 0f,Pr ,Tr , S 0Pr ,Tr, and V 0Pr ,Tr values for various aqueous ions,
after Shock and Helgeson (1988). These values, together with the coefficients in tables
Table 8.13 Standard partial molal Gibbs free energy of formation
from the elements (G 0f, P r , Tr ; cal/mole), standard partial molal
enthalpy of formation from the elements ( H 0f, P r , Tr ; cal/mole),
standard partial molal entropy ( S 0P r , T r ; cal/(mole ⫻ K)), and
standard partial molal volume ( V 0P r , T r ; cm3/mole) for aqueous ions
at Tr ⫽ 25 °C and Pr ⫽ 1 bar (Shock and Helgeson, 1988).
H⫹ 0 0 0 0
Li⫹ ⫺69,933 ⫺66,552 2.70 ⫺0.87
Na⫹ ⫺62,591 ⫺57,433 13.96 ⫺1.11
K⫹ ⫺67,510 ⫺60,270 24.15 9.06
Rb⫹ ⫺67,800 ⫺60,020 28.8 14.26
Cs⫹ ⫺69,710 ⫺61,670 31.75 21.42
NH4⫹ ⫺18,990 ⫺31,850 26.57 18.13
Ag⫹ 18,427 25,275 17.54 ⫺0.8
T1⫹ ⫺7,740 1,280 30.0 18.2
VO2⫹ ⫺140,300 ⫺155,300 ⫺10.1 ⫺33.5
Cu⫹ 11,950 17,132 9.7 ⫺8.0
Au⫹ 39,000 47,580 24.5 11.1
Mg2⫹ ⫺108,505 ⫺111,367 ⫺33.00 ⫺21.55
Ca2⫹ ⫺132,120 ⫺129,800 ⫺13.5 ⫺18.06
Sr2⫹ ⫺134,760 ⫺131,670 ⫺7.53 ⫺17.41
Ba2⫹ ⫺134,030 ⫺128,500 2.3 ⫺12.60
Pb2⫹ ⫺5,710 220 4.2 ⫺15.6
Mn2⫹ ⫺54,500 ⫺52,724 ⫺17.6 ⫺17.1
Co2⫹ ⫺13,000 ⫺13,900 ⫺27.0 ⫺24.4
Ni2⫹ ⫺10,900 ⫺12,900 ⫺30.8 ⫺29.0
Cu2⫹ 15,675 15,700 ⫺23.2 ⫺24.6
Zn2⫹ ⫺35,200 ⫺36,660 ⫺26.2 ⫺24.3
Cd2⫹ ⫺18,560 ⫺18,140 ⫺17.4 ⫺15.6
Be2⫹ ⫺83,500 ⫺91,500 ⫺55.7 ⫺25.4
Ra2⫹ ⫺134,200 ⫺126,100 13 ⫺10.6
Sn2⫹ ⫺6,630 ⫺2,100 ⫺3.8 ⫺15.5
VO2⫹ ⫺106,700 ⫺116,300 ⫺32 ⫺29.4
Fe2⫹ ⫺21,870 ⫺22,050 ⫺25.3 ⫺22.2
Pd2⫹ 42,200 42,080 ⫺22.6 ⫺20.8
Ag2⫹ 64,300 64,200 ⫺21.0 ⫺19.3
Hg2⫹ 39,360 40,670 ⫺8.68 ⫺19.6
Hg22⫹ 36,710 39,870 ⫺15.66 14.4
Sm2⫹ ⫺123,000 ⫺120,500 ⫺6.2 ⫺5.7
Eu2⫹ ⫺129,100 ⫺126,100 ⫺2.4 ⫺2.2
Yb2⫹ ⫺128,500 ⫺126,800 ⫺11.2 ⫺10.3
Al3⫹ ⫺115,377 ⫺126,012 ⫺75.6 ⫺44.4
La3⫹ ⫺164,000 ⫺169,600 ⫺52.0 ⫺38.6
Ce3⫹ ⫺161,600 ⫺167,400 ⫺49 ⫺39.8
Pr 3⫹ ⫺162,600 ⫺168,800 ⫺50 ⫺42.1
Nd3⫹ ⫺160,600 ⫺166,500 ⫺49.5 ⫺43.1
Sm3⫹ ⫺159,100 ⫺165,200 ⫺50.7 ⫺42.0
Eu3⫹ ⫺137,300 ⫺144,700 ⫺53.0 ⫺41.3
Table 8.13 (continued)
8.11 and 8.12, allow the calculation of the corresponding partial molar properties in high
P and T conditions.* Practical calculations are performed by the various releases of the
* Actually, the various equations listed in this section are insufficient to perform the complete
calculation since one would first calculate the density of H2O through eq. 8.12 or 8.14. Equation
8.14 in its turn involves the partial derivative of the Helmholtz free energy function 8.15. Moreover,
the evaluation of electrostatic properties of the solvent and of the Born functions , Q, Y, X involve
additional equations and variables not given here for the sake of brevity (eqs. 36, 40 to 44, 49 to 52
and tables 1 to 3 in Johnson et al., 1991). In spite of this fact, the decision to outline here briefly
the HKF model rests on its paramount importance in geochemistry. Moreover, most of the listed
thermodynamic parameters have an intrinsic validity that transcends the model itself.
Geochemistry of Aqueous Phases / 539
reduction→
Fe 3 + + e − ⇔ Fe 2 + . (8.155)
←oxidation
In a “reduction process” the flow of electrons is from left to right (i.e., from
oxidizing to reducing media). A Gibbs free energy change ⌬G 0155 is associated
with equilibrium 8.155, so that
a Fe 2 +
∆G155
0
= − RT ln K 155 = − RT ln . (8.156)
a Fe 3+ a e −
Term ae⫺ in equation 8.156 is the “activity of the hypothetical electron in solu-
tion,” which, by convention, is always assumed to be unitary.
Let us now consider a redox equilibrium involving metallic iron Fe and fer-
rous iron Fe2⫹ (one-molal solution at Tr ⫽ 25 °C and Pr ⫽ 1 bar):
Fe ⇔ Fe 2 + + 2 e − .
metal aqueous (8.157)
The Gibbs free energy change involved in equation 8.157 corresponds to the par-
tial molal Gibbs free energy of formation of Fe2⫹ from the element at stable state
(cf. table 8.13):
∆G157
0
=G 0
. (8.158)
f , Tr , Pr , Fe 2 +
In the case of oxidation of hydrogen, starting from the diatomic gaseous molecule,
we have
1
H2 ⇔ H+ + e − .
2 (8.159)
gaseous aqueous
540 / Geochemistry of Fluids
⌬G 0159 is the Gibbs free energy change associated with equation 8.159, corre-
sponding to the formation of the aqueous hydrogen ion H⫹ from the gaseous
molecule H2 . This amount of energy is, by convention (see, for instance, table
8.13), fixed at zero.
To the reactions of equations 8.155, 8.157, and 8.159, which imply electron
transfer from the oxidized to the reduced state, the concept of electrical potential
is associated through Faraday’s law:
∆G = − nFE , (8.160)
H2 , H + Zn, Z 2n + ,
(8.161)
gas aq metal aq
f H 2 gas f H 2 , gas
a H 2 , gas = = = 1. (8.162)
f H0 gas 1
2
Table 8.14 Standard potentials for various redox couples (aqueous cation–metal; element–aqueous
anion) arranged in order of decreasing E 0. Values, expressed in volts, are consistent with standard
partial molal Gibbs free energy values listed in table 8.13. E 0 value for sulfur is from Nylen and
Wigren (1971).
Equilibrium E 0 (V ) Equilibrium E 0 (V )
Li⫹ ⫹ e⫺ → Li ⫺3.034 Al3⫹ ⫹ 3e⫺ → Al ⫺1.668
Cs⫹ ⫹ e⫺ → Cs ⫺3.023 Mn2⫹ ⫹ 2e⫺ → Mn ⫺1.182
Rb⫹ ⫹ e⫺ → Rb ⫺2.940 Zn2⫹ ⫹ 2e⫺ → Zn ⫺0.763
K⫹ ⫹ e⫺ → K ⫺2.928 Ga2⫹ ⫹ 2e⫺ → Ga ⫺0.549
Sr 2⫹ ⫹ 2e⫺ → Sr ⫺2.922 S ⫹ 2e⫺ → S2⫺ ⫺0.480
Ra2⫹ ⫹ 2e⫺ → Ra ⫺2.910 Fe2⫹ ⫹ 2e⫺ → Fe ⫺0.474
Ba2⫹ ⫹ 2e⫺ → Ba ⫺2.906 Cd2⫹ ⫹ 2e⫺ → Cd ⫺0.402
Ca2⫹ ⫹ 2e⫺ → Ca ⫺2.865 In3⫹ ⫹ 3e⫺ → In ⫺0.338
Eu2⫹ ⫹ 2e⫺ → Eu ⫺2.799 Tl⫹ ⫹ e⫺ → Tl ⫺0.336
Yb2⫹ ⫹ 2e⫺ → Yb ⫺2.786 Co2⫹ ⫹ 2e⫺ → Co ⫺0.282
Na⫹ ⫹ e⫺ → Na ⫺2.714 Ni2⫹ ⫹ 2e⫺ → Ni ⫺0.236
Sm2⫹ ⫹ 2e⫺ → Sm ⫺2.667 Sn2⫹ ⫹ 2e⫺ → Sn ⫺0.144
La3⫹ ⫹ 3e⫺ → La ⫺2.371 Pb2⫹ ⫹ 2e⫺ → Pb ⫺0.124
Y3⫹ ⫹ 3e⫺ → Y ⫺2.368 Fe3⫹ ⫹ 3e⫺ → Fe ⫺0.059
Mg2⫹ ⫹ 2e⫺ → Mg ⫺2.353 2H⫹ ⫹ 2e⫺ → H2 0.000
Pr3⫹ ⫹ 3e⫺ → Pr ⫺2.350 Cu2⫹ ⫹ 2e⫺ → Cu 0.340
Ce3⫹ ⫹ 3e⫺ → Ce ⫺2.336 Co3⫹ ⫹ 3e⫺ → Co 0.462
Ho3⫹ ⫹ 3e⫺ → Ho ⫺2.333 Cu⫹ ⫹ e⫺ → Cu 0.518
Nd3⫹ ⫹ 3e⫺ → Nd ⫺2.321 I2 ⫹ 2e⫺ → 2I⫺ 0.538
Er3⫹ ⫹ 3e⫺ → Er ⫺2.311 Tl3⫹ ⫹ 3e⫺ → Tl 0.742
Tm3⫹ ⫹ 3e⫺ → Tm ⫺2.311 Ag⫹ ⫹ e⫺ → Ag 0.799
Tb3⫹ ⫹ 3e⫺ → Tb ⫺2.306 Hg2⫹ ⫹ 2e⫺ → Hg 0.853
Lu3⫹ ⫹ 3e⫺ → Lu ⫺2.304 Pd2⫹ ⫹ 2e⫺ → Pd 0.915
Dy3⫹ ⫹ 3e⫺ → Dy ⫺2.294 Br2 ⫹ 2e⫺ → 2Br⫺ 1.078
Gd3⫹ ⫹ 3e⫺ → Gd ⫺2.293 Cl2 ⫹ 2e⫺ → 2Cl⫺ 1.361
Yb3⫹ ⫹ 3e⫺ → Yb ⫺2.212 Ag2⫹ ⫹ 2e⫺ → Ag 1.394
Sc3⫹ ⫹ 3e⫺ → Sc ⫺2.027 Au3⫹ ⫹ 3e⫺ → Au 1.498
Eu3⫹ ⫹ 3e⫺ → Eu ⫺1.985 Au⫹ ⫹ e⫺ → Au 1.691
Be2⫹ ⫹ 2e⫺ → Be ⫺1.810 F2 ⫹ 2e⫺ → 2F⫺ 2.920
542 / Geochemistry of Fluids
Fe + Cu 2 + ⇔ Fe 2 + + Cu
metal aqueous aqueous metal
(8.163)
or even
Zn + Cu 2 + ⇔ Zn 2 + + Cu . (8.164)
metal aqueous aqueous metal
We can visualize the reactions 8.163 and 8.164 as composed of partial equilib-
ria—i.e., for equation 8.163,
Fe ⇔ Fe 2 + + 2 e − (0.474 V ) (8.165)
metal aqueous
Cu 2 + + 2 e − ⇔ Cu (0.340 V ) . (8.166)
aqueous metal
Zn ⇔ Zn 2 + + 2 e −
metal aqueous
(0.763 V ) (8.167)
Cu 2 + + 2 e − ⇔ Cu
aqueous metal
(0.340 V ) . (8.168)
0
E163 = 0.474 + 0.340 = 0.814 (V) , (8.169)
0
E164 = 0.763 + 0.340 = 1.103 (V). (8.170)
Based on Faraday’s law (equation 8.160), the ⌬G of the reaction is in both cases
negative: reactions then proceed spontaneously toward the right, with the dissolu-
tion of Fe and Zn electrodes, respectively, and deposition of metallic Cu at the
cathode.
Geochemistry of Aqueous Phases / 543
Let us now consider a galvanic cell with the redox couples of equation 8.164.
This cell may be composed of a Cu electrode immersed in a one-molal solution
of CuSO4 and a Zn electrode immersed in a one-molal solution of ZnSO4 (“Dan-
iell cell” or “Daniell element”). Equation 8.170 shows that the galvanic potential
is positive: the ⌬G of the reaction is negative and the reaction proceeds toward
the right. If we short-circuit the cell to annul the potential, we observe dissolution
of the Zn electrode and deposition of metallic Cu at the opposite electrode. The
flow of electrons is from left to right: thus, the Zn electrode is the anode (metallic
Zn is oxidized to Zn2⫹; cf. eq. 8.167), and the Cu electrode is the cathode (Cu2⫹
ions are reduced to metallic Cu; eq. 8.168):
e−
→
Zn, Zn 2 + , SO 24 − Cu, Cu 2 + , SO 24 − . (8.171)
䊞 䊝
As we have already seen, the standard potentials are relative to standard refer-
ence conditions—i.e., one-molal solutions at Tr ⫽ 25 °C and Pr ⫽ 1 bar, in equilib-
rium with pure metals or pure gases. Applying the Nernst relation to a redox
equilibrium such as reaction 8.163 and assuming unitary activity for the con-
densed phases (i.e., pure metals), we have
a Fe 2 + ,aqueous
∆G163 = ∆G163
0
+ RT ln . (8.172)
a Cu2 + ,aqueous
Combining equation 8.172 with the Faraday relation (eq. 8.160), we obtain
RT aFe2 + ,aqueous
E163 = E163
0
− ln
nF aCu2 + ,aqueous
(8.173)
RT m 2+ RT γ Fe2 +
= E163
0
− ln Fe − ln .
nF mCu2 + nF γ Cu2 +
the activity ratio conforms to the difference between the imposed Eh and the
standard state potential:
RT a Fe 2 + ,aqueous
Eh − E 163
0
=− ln . (8.175)
nF a Cu2 + ,aqueous
O2 + 4 H+ + 4 e − ⇔ 2 H2O .
gas aqueous
(8.176)
2.303RT 2.303RT
Eh = 1.228 + 4 log aH + + log f O2
nF nF (8.177)
= 1.228 + 0.0591 log aH + + 0.0148 log f O2 .
2 H + + 2 e − ⇔ H2 .
aqueous gas
(8.178)
The definition of pH is
pH = − log a H + , (8.183)
hence
2 H2O ⇔ 2 H2 + O2 . (8.186)
f H2 = 2 fO . (8.187)
(gas) 2 (gas)
Combining equation 8.187 with equations 8.176 and 8.178 and with the acid-
base neutrality concept (eq. 8.184 and 8.185), we obtain the absolute neutrality
condition for a diluted aqueous solution at T ⫽ 25 °C and P ⫽ 1 bar—i.e.,
pH = 7.0; log fH2 = −27.61; log fO2 = −27.91; E 0 = 0.4 ( V). (8.188)
546 / Geochemistry of Fluids
Figure 8.19 shows the thermodynamic stability field for diluted aqueous solu-
tions at T ⫽ 25 °C and P ⫽ 1 bar. The pH-Eh range is bounded upward by the
maximum oxidation limit, with a slope of ⫺0.0591 and an intercept on the ordi-
nate axis depending on the partial pressure of oxygen: several limiting curves are
drawn for various values of f O2. We recall here that the partial pressure of oxygen
in the earth’s atmosphere at sea level is about 0.2 bar, so that, assuming a fugacity
coefficient of 1 ( f O2 ⫽ PO2 ⫽ 0.2), the intercept on the ordinate axis at pH ⫽ 0 is
Eh ⫽ 1.22. The lower stability limit is defined, as we have already seen, by equa-
tion 8.179 for the various values of f H2. The absolute neutrality point subdivides
the Eh-pH field of thermodynamic stability into four parts:
We have already seen that the molality ratio of species involved in redox equilibria
is buffered by the redox state of the system with respect to the standard potential
of the redox couple. To understand this concept better, let us examine the various
Geochemistry of Aqueous Phases / 547
( )
3+
Ce OH + H+ + e − ⇔ Ce3 + + H2O (8.189)
( )
2+
Ce OH + 2 H+ + e − ⇔ Ce3 + + 2 H2O (8.190)
2
( ) ( )
3+ 2+
Ce OH + H2O ⇔ Ce OH + H+ . (8.191)
2
The ruling equilibria between aqueous species and condensed forms are
Ce3 + + 3 e − ⇔ Ce
aqueous metal
(8.192)
( ) ( )
2+
Ce OH + H2O+ e − ⇔ Ce OH + H+
2 3 (8.194)
aqueous crystal
( )
2+
CeO2 + 2 H+ ⇔ Ce OH ,
2 (8.195)
crystal aqueous
( )
Ce OH + 3 H+ + 3 e − ⇔ Ce+ 3 H2O
3 (8.196)
crystal metal
If we apply the Nernst and Faraday relations to equations 8.189 and 8.190 as
before (cf. eq. 8.177 and 8.179), we obtain
Ce OH
( )
3+
Eh = 0
E189 − 0.0591 pH + 0.0591 log (8.198)
Ce 3+ [ ]
548 / Geochemistry of Fluids
and
Ce OH
( )
2+
2
Eh = E190
0
− 0.1182 pH + 0.0591 log , (8.199)
Ce 3+ [ ]
where terms in square brackets denote thermodynamic activities of
solutes, E 0189 ⫽ 1.715 V, and E 0190 ⫽ 1.731 V.
On an Eh-pH plot (figure 8.20), the isoactivity conditions for solute species:
(
Ce OH
) [
= Ce 3+
]
3+
(8.200)
Figure 8.20 Eh-pH diagram for the Ce-H2O system (modified from Pourbaix, 1966). Fig-
ures on limiting curves are base 10 logarithms of solute activity: unitary activity (i.e., one-
molal solution) is identified by zero.
Geochemistry of Aqueous Phases / 549
Ce OH
( ) [
= Ce 3+
]
2+
(8.201)
2
( )
Ce OH
2+
2
log = −0.29 + pH . (8.204)
Ce (OH )
3+
The limit (equation 8.204) is thus parallel to the ordinate axis (pH ⫽ 0.29; cf.
figure 8.20).
As we did for equilibria between solute species, we can also define the bound-
aries between solute species and condensed phases. Assuming the condensed
forms to be pure phases (i.e., assuming unitary activity), in the presence of metal-
lic cerium we have
Eh = E 192
0
+ 0.0197 log Ce 3 + , [ ] ( 8.205)
[ ]
log Ce 3+ = 22.15 + 3 pH, (8.207)
(
log Ce OH ) = 19.22 − 2 pH .
2+
2 (8.208)
550 / Geochemistry of Fluids
As figure 8.20 shows, the boundary between condensed phases and solute species
is dictated by the activity of solutes, T being equal.
Lastly, the boundary between Ce and Ce(OH)3 (eq. 8.196) is given by
Eh = E 196
0
− 0.0591 pH, ( 8.209)
with E 0196 ⫽ 2.046, and the boundary between Ce(OH)3 and CeO2 by
Eh = E197
0
− 0.0591 pH, (8.210)
As repeatedly noted, standard potentials are linked to the standard molal Gibbs
free energy of formation from the elements through Faraday’s equation. Let us
Geochemistry of Aqueous Phases / 551
now reconsider the equilibrium between metallic iron and aqueous ion at T ⫽
25 °C and P ⫽ 1 bar:
Fe ⇔ Fe2 + + 2 e − .
metal aqueous
(8.211)
At standard state, equation 8.211 obviously concerns standard state stable com-
ponents—i.e., pure Fe metal at T ⫽ 25 °C, P ⫽ 1 bar, and a hypothetical
one-molal Fe2⫹ solution referred to infinite dilution, at the same P and T condi-
tions (cf. section 8.4). The chemical potentials of components in reactions
(aFe(metal), aFe2⫹(aqueous), ae⫺(aqueous) ) are those of standard state; hence, by defini-
tion, the activity of all components in reaction is 1—i.e.,
a Fe (metal) = 1; a Fe 2 + = 1; a e− = 1, ( 8.212)
(aqueous) (aqueous)
G f e − = G f0 e − = 0 ( 8.213)
—it follows that the activity of the hypothetical electron in solution ae⫺ is always
1 under all P, T, and molality conditions. We have already seen that, based on
this consideration, ae⫺ is never explicit in Nernst equations (as in eq. 8.172 and
8.175, for example).
However, in the literature we often encounter the redox parameter “pe,”
which, like pH, represents the cologarithm of the activity of the hypothetical elec-
tron in solution:
F
pe = Eh (8.215)
2.303RT
(cf. Stumm and Morgan, 1981). As outlined by Wolery (1983), equation 8.215
implies that the Gibbs free energy of formation of the hypothetical electron in
solution is not always zero, as stated in equation 8.213, but is given (cf. eq.
8.159) by
552 / Geochemistry of Fluids
1
−
G f e (aqueous) = G H 2(gas) − G H + . (8.216)
2 (aqueous)
The incongruency between equations 8.216 and 8.213 leads us to discourage the
use of parameter pe, which is formally wrong from the thermodynamic point of
view. We also note that, based on equation 8.215, pe is simply a transposition in
scale and magnitude of Eh (pe is adimensional, Eh is expressed in volts).
For the sake of completeness, we recall that analogies with the concept of pH
do not concern solely the pe factor. In biochemistry, for instance, O2 and H2
fugacities in gaseous phases are often described by the rO and rH parameters, re-
spectively:
Table 8.15 stresses the existence of marked redox disequilibria among the
various couples. In this case (and also for the previously quoted cases), the con-
cept of “system Eh” is inappropriate. Nevertheless, table 8.15 shows that the mea-
sured Eh is near the value indicated by the redox couple [NO3 ]/[NO2 ]. We can
thus assume that the redox equilibrium between [NO3 ] and [NO2 ] determines the
ruling Eh.
Of the various factors that cause redox disequilibria, the most effective are
biologic activity (photosynthesis) and the metastable persistence of covalent com-
plexes of light elements (C, H, O, N, S), whose bonds are particularly stable and
difficult to break (Wolery, 1983). For the sake of completeness, we can also note
that the apparent redox disequilibrium is sometimes actually attributable to ana-
lytical error or uncertainty (i.e., difficult determination of partial molalities of
species, often extremely diluted) or even to error in speciation calculations (when
using, for instance, the redox couple Fe3⫹ /Fe2⫹, one must account for the fact
that both Fe3⫹ and Fe2⫹ are partly bonded to anionic ligands so that their free
ion partial molalities do not coincide with the bulk molality of the species).
Table 8.16 Standard molal Gibbs free energies of formation from the elements for aqueous ions
and complexes and condensed phases, partly adopted in constructing the Eh-pH diagrams in figure
8.21. Data in kcal/mole. Values in parentheses: Shock and Helgeson’s (1988) tabulation. Sources of
data: (1) Wagman et al. (1982); (2) Garrels and Christ (1965); (3) Pourbaix (1966); (4) Berner (1971)
Ion or Ion or
Complex G f,0 T r , P r Reference Complex G f,0 T r , P r Reference
BO2 ⫺162.26 (⫺162.24) (1) S2⫺ 20.51 (19.00) (1)
B2O3(cr) ⫺285.29 (1) HS⫺ 2.89 (2.86) (1)
B4O72⫺ ⫺622.56 (1) H 2S ⫺6.65 (1)
BH4⫺ 27.33 (1) SO32⫺ ⫺116.28 (⫺116.30) (1)
B(OH)4⫺ ⫺275.61 (1) HSO3⫺ ⫺126.13 (⫺126.13) (1)
H3BO3 ⫺231.54 (1) SO42⫺ ⫺177.95 (⫺177.93) (1)
H2BO3⫺ ⫺217.60 (3) HSO4⫺ ⫺180.67 (⫺180.63) (1)
HBO32⫺ ⫺200.29 (3) F⫺ ⫺66.63 (⫺67.34) (1)
BO33⫺ ⫺181.48 (3) Cl⫺ ⫺31.36 (⫺31.38) (1)
CH4 ⫺8.28 (1) ClO3⫺ ⫺1.90 (1.90) (1)
CH2O ⫺31.00 (2) ClO4⫺ ⫺2.04 (⫺2.04) (1)
H2CO3 ⫺149.00 (1) Br⫺ ⫺24.85 (24.87) (1)
HCO3⫺ ⫺140.24 (⫺140.28) (1) BrO3⫺ 4.46 (4.45) (1)
CO32⫺ ⫺126.15 (⫺126.19) (1) BrO4⫺ 28.23 (28.20) (1)
NO3 ⫺25.99 (⫺26.51) (1) I⫺ ⫺12.33 (⫺12.41) (1)
NH4⫹ ⫺18.96 (⫺18.99) (1) IO3⫺ ⫺30.59 (⫺30.60) (1)
NH3(gas) ⫺3.98 (4) IO4⫺ ⫺13.98 (⫺14.00) (1)
H3PO4 ⫺273.07 (1) H2PO4⫺ ⫺270.14 (1)
HPO42⫺ ⫺260.31 (⫺260.41) (1) PO42⫺ ⫺243.48 (⫺243.50) (1)
Figure 8.21B shows the Eh-pH diagram for nitrogen, calculated for fN2 ⫽ 0.8
bar (total molality of dissolved N species ⫽ 10⫺3.3 ). The gaseous molecule N2 is
stable over most of the Eh-pH field. The ammonium ion NH⫹ 4 predominates un-
⫺
der reducing conditions, and the nitrate ion NO3 predominates under oxidizing
conditions. Gaseous ammonia forms only under extremely reducing and alkaline
conditions. The fields of ammonium and nitrate ions expand metastably within
the stability field of gaseous nitrogen (the limit of metastable predominance of
the two species is drawn as a dashed line within the N2 field in figure 8.21B).
Figure 8.21C shows the Eh-pH diagram for phosphorus at a solute total mol-
ality of 10⫺4. Within the stability field of water, phosphorus occurs as orthophos-
phoric acid H3PO4 and its ionization products. The predominance limits are dic-
tated by the acidity of the solution and do not depend on redox conditions.
Sulfur (figure 8.21D) is present in aqueous solutions in three oxidation states
(2⫺, 0, and 6⫹). The field of native S, at a solute total molality of 10⫺3, is very
limited and is comparable to that of carbon (for both extension and Eh-pH
range). Sulfide complexes occupy the lower part of the diagram. The sulfide-
sulfate transition involves a significant amount of energy and defines the limit of
predominance above which sulfates occur.
The extension of the field of native elements (S, C) depends on the adopted
Geochemistry of Aqueous Phases / 555
Figure 8.21 Eh-pH diagrams for main anionic ligands. From Brookins (1988). Reprinted
with permission of Springer-Verlag, New York.
molality values of solutes (i.e., it expands with increasing, and shrinks with de-
creasing, molal concentrations). The molal concentrations used for constructing
figure 8.21 are mean values for natural waters, but they may vary considerably.
The above-discussed anionic ligands are the most important ones from the
viewpoint of redox properties. For the sake of completeness, we can also consider
the redox behavior of halides and boron. Halides stable in water are the simple
556 / Geochemistry of Fluids
For all the other halides, Eh-pH conditions have no influence. Boron occurs in
water mainly as boric acid H3BO3 and its progressive ionization products at in-
creasing pH. Redox conditions do not affect the speciation state of boron.
Information on the speciation states of solutes and their equilibria with con-
densed phases furnished by Eh-pH diagrams is often simply qualitative and
should be used only in the initial stages of investigations. The chemical complexity
of natural aqueous solutions and the persistent metastability and redox disequi-
librium induced by organic activity are often obstacles to rigorous interpretation
of aqueous equilibria.
Let us now consider the effects of chemical complexity on the predominance
limits of solutes at environmental P and T conditions. The elements in question
are emblematic for their particular importance in ore formation.
Figure 8.22A shows the Eh-pH diagram of iron in the Fe-O-H system at T ⫽
25 °C and P ⫽ 1 bar. The diagram is relatively simple: the limits of predominance
are drawn for a solute total molality of 10⫺6. Within the stability field of water,
iron is present in the valence states 2⫹ and 3⫹. In figure 8.22A, it is assumed that
the condensed forms are simply hematite Fe2O3 and magnetite Fe3O4 . Actually,
in the 3⫹ valence state, metastable ferric hydroxide Fe(OH)3 and metastable goe-
thite FeOOH may also form, and, in the 2⫹ valence state, ferrous hydroxide
Fe(OH)2 may form. It is also assumed that the trivalent solute ion is simply Fe3⫹,
whereas, in fact, various aqueous ferric complexes may nucleate [i.e., Fe(OH)2⫹,
Fe(OH)⫹ 2 , etc.].
Figure 8.22B shows the effect of adding carbon to the system (molality of
solute carbonates ⫽ 10⫺2 ): in reducing alkaline states, the field of siderite FeCO3
opens; in the other Eh-pH conditions, the limits are unchanged.
Adding sulfur to the system (molality of solutes ⫽ 10⫺6; figure 8.22C), a wide
stability field opens for pyrite FeS2 in reducing conditions and, almost at the
lower stability limit of water, a limited field of pyrrhotite FeS is observed.
Figure 8.22 shows the effect of silica in the system (aqueous solution saturated
with respect to amorphous silica): under reducing alkaline conditions, the meta-
silicate FeSiO3 forms at the expense of magnetite through the equilibrium
Figure 8.22 Eh-pH diagrams for iron-bearing aqueous solutions. Chemical complexity
increases from A to F.
Reaction 8.220 takes place in the silicate bands present in iron ores: iron silicates
occur in such bands in coexistence with opaline amorphous silica (Klein and
Bricker, 1972). A complex Eh-pH diagram for the Fe-C-S-O-H system is shown
in figure 8.22E. The stability field of pyrrhotite FeS disappears and is replaced by
558 / Geochemistry of Fluids
siderite FeCO3 , which also limits the field of pyrite. When silicon is also present
in the system (figure 8.22F), magnetite does not nucleate as a stable phase.
The predominance limits shown in figure 8.22 are analytically summarized in
table 8.17. Compare figures 8.22 and 8.21 to better visualize the redox state of the
anionic ligands at the various Eh-pH conditions of interest (particularly the
sulfide-sulfate transition and carbonate limits). We remand to Garrels and Christ
(1965) for a more detailed account on the development of complex Eh-pH dia-
grams.
Let us now consider the effects of Eh-pH conditions on the speciation state
and solubility of manganese in aqueous solutions. Manganese complexes have
been carefully studied in the last decade, owing to the discovery on ocean floors
of economically important metalliferous deposits (nodules and crusts) in which
Mn compounds are dominant.
Figure 8.23A shows a simplified Eh-pH diagram for the Mn-O-H system.
Within the stability field of water, manganese occurs in three valence states (2⫹,
3⫹, and 4⫹). Figure 8.23A shows the condensed phases relative to the three va-
lence states as the hydroxide pyrochroite Mn(OH)2 (2⫹), multiple oxide haus-
mannite Mn3O4 (2⫹, 3⫹), sesquioxide Mn2O3 (3⫹), and oxide pyrolusite
MnO2 (4⫹).
A wide field is occupied by the Mn2⫹ ion. The condensed phases nucleate
only in alkaline solutions, and the stability fields of the various solids expand with
increasing Eh. The precipitation of Mn oxides and hydroxides is actually a com-
plex phenomenon involving the initial nucleation of metastable compounds. Hem
and Lind (1983) showed that, at fO2 near the upper stability limit of water and at
pH in the range 8.5 to 9.5, the type of nucleating solid depends on temperature:
at T ⫽ 25 °C metastable hausmannite Mn3O4 forms, whereas at T near 0 °C the
mixed hydroxides feitknechtite, manganite, and groutite [all with formula unit
MnO(OH)] precipitate. All these compounds are metastable in experimental Eh-
pH conditions and transform progressively into the tetravalent compound MnO2 .
Figure 8.23 Eh-pH diagrams for manganese. From Brookins (1988). Reprinted with per-
mission of Springer-Verlag, New York.
Table 8.17 Main predominance limits of aqueous complexes and saturation limits between
solutes and condensed phases in iron-bearing aqueous solutions (see figure 8.22). Standard
state Gibbs free energies of formation of species are listed in table 8.18. (c) ⫽ crystalline;
(a) ⫽ aqueous; (g) ⫽ gaseous; (l) ⫽ liquid. [ ] ⫽ thermodynamic activity.
Eh ⫽ 0.78
2⫹
2Fe(a) ⫹ 3H2O(l) → Fe2O3(c) ⫹ 6H⫹
(a) ⫹ 2e
⫺
Eh ⫽ 0.221 ⫺ 0.0591pH
FeS(c) ⫹ HS⫺ ⫹
(a) → FeS2(c) ⫹ H(a) ⫹ 2e
⫺
Figure 8.24 Eh-pH diagrams for copper. From Brookins (1988). Reprinted with permis-
sion of Springer-Verlag, New York.
Figure 8.23B shows the effect of adding sulfur and carbon in solution (solute
molality ⫽ 10⫺3 ). In reducing-alkaline conditions, both rhodochrosite MnCO3
and alabandite MnS are stable. However, the field of alabandite is quite restricted,
which explains why this mineral is rarely found in nature (Brookins, 1988).
Figure 8.24A shows the Eh-pH diagram for copper in the Cu-O-H system
(solute molality ⫽ 10⫺6 ). Within the stability field of water, copper assumes the
valence states 0, ⫹1, and ⫹2. Native copper occupies much of the diagram, en-
compassing all acidity states and extending to Eh conditions higher than the
sulfide-sulfate transition (cf. figure 8.21). In acidic solutions, native copper oxi-
dizes to cupric ion Cu2⫹. In alkaline solutions, with increasing Eh, cuprite Cu2O
and tenorite CuO form. Only in extremely alkaline solutions is the soluble com-
pound CuO2⫺ 2 stable. If carbon and sulfur are added to the system (figure 8.24B;
solute molality ⫽ 10⫺3 ), the field of native copper is reduced, owing to the pres-
ence of the sulfides calcocite Cu2S and covellite CuS. At high Eh, we find the
hydrated carbonate malachite Cu2(OH)2CO3 replacing tenorite. In the presence
of iron, the diagram becomes much more complex, with the formation of impor-
tant ore-forming phases, such as chalcopyrite CuFeS2 and bornite Cu5FeS4 (see,
for instance, figures 7.24 and 7.25 in Garrels and Christ, 1965).
The sequence of diagrams shown in figures 8.22, 8.23, and 8.24 is only an
example of an approach to the actual complexity of natural solutions. The infor-
mation given by such diagrams, although qualitative, furnishes a basis for more
rigorous calculations and is therefore extremely precious in the comprehension
of natural phenomena. Table 8.18 lists the standard state Gibbs free energy values
used in the construction of these diagrams.
Although the predominant species forming in the C-H-O system (besides H2O)
are methane CH4 , carbon dioxide CO2 and its aqueous complexes HCO3 and
Table 8.18 Standard state (T ⫽ 25 °C; P ⫽ 1
bar) Gibbs free energy of formation for iron,
copper, and manganese compounds, partly
adopted in constructing Eh-pH diagrams of
figures 8.22, 8.23, and 8.24. References: (1)
Wagman et al. (1982); (2) Garrels and Christ
(1965); (3) Robie et al. (1978); (4) Hem et al.
(1982); (5) Shock and Helgeson (1988); (6)
Bricker (1965).
Fe ⫽ 0.0
Fe2⫹ ⫽ ⫺20.30 (2)
Fe3⫹ ⫽ ⫺2.52 (2)
Fe(OH)2⫹ ⫽ ⫺55.91 (2)
Fe(OH)⫹2 ⫽ ⫺106.2 (2)
FeO2H⫺ ⫽ ⫺90.6 (2)
Fe0.95O (wustite) ⫽ ⫺58.4 (2)
Fe2O3 (hematite) ⫽ ⫺177.1 (2)
Fe3O4 (magnetite) ⫽ ⫺242.4 (2)
Fe(OH)2 ⫽ ⫺115.57 (2)
Fe(OH)3 ⫽ ⫺166.0 (2)
FeS (pyrrhotite) ⫽ ⫺23.32 (2)
FeS2 (pyrite) ⫽ ⫺36.00 (2)
FeCO3 (siderite) ⫽ ⫺161.06 (2)
Cu ⫽ 0.00
Cu⫹ ⫽ 11.94 (1)
Cu2⫹ ⫽ 15.65 (1)
Cu2S ⫽ ⫺20.60 (chalcocite) (1)
CuS ⫽ ⫺12.81 (covellite) (1)
Cu2O ⫽ ⫺34.98 (cuprite) (2)
CuO ⫽ ⫺31.00 (tenorite) (1)
CuO22⫺ ⫽ ⫺43.88 (1)
Cu2(CO3)(OH)2 ⫽ ⫺213.58 (malachite) (1)
Cu3(CO3)2(OH)2 ⫽ ⫺314.29 (azurite) (1)
Mn2⫹ ⫽ ⫺54.52 (1)
Mn(OH)⫹ ⫽ ⫺96.80 (1)
Mn(OH)3⫺ ⫽ ⫺177.87 (1)
MnSO4 ⫽ ⫺232.5 (1)
MnO4⫺ ⫽ ⫺106.9 (5)
MnO42⫺ ⫽ ⫺119.7 (5)
MnO ⫽ ⫺86.74 (1)
Mn(OH)2 ⫽ ⫺146.99 (1); ⫺147.14 (6) (pyrochroite)
MnCO3 ⫽ ⫺195.20 (1); ⫺195.04 (3) (rhodochrosite)
MnS ⫽ ⫺52.20 (1)
Mn3O4 ⫽ ⫺306.69 (1) (hausmannite)
Mn2O3 ⫽ ⫺210.59 (1)
Mn(OH)3 ⫽ ⫺181.0 (2)
 MnO(OH) ⫽ ⫺129.8 (4) (feitknechtite)
␥ MnO(OH) ⫽ ⫺133.3 (6) (manganite)
␣ MnO2 ⫽ ⫺108.3 (6) (birnessite)
 MnO2 ⫽ ⫺111.17 (3) (pyrolusite)
␥ MnO2 ⫽ ⫺109.1 (6) (␥-nsutite)
562 / Geochemistry of Fluids
CO2⫺3 , and graphite C, at low P and T conditions, several organic complexes form
metastably, mainly as the result of bacterial activity, and persist at P and T condi-
tions that may exceed those achieved at a burial depth on the order of 5 km,
which is generally considered the limit of complete degradation.
Organic matter suspended in solution and present in sediments is an efficient
complexing agent for most divalent and trivalent cations. Most organic matter in
aquatic and sedimentary environments is in the form of humic and fulvic acids.
About 80% of soluble organic matter in lakes and rivers is composed of fulvic
acids, whereas the abundance of humic acids varies markedly from region to re-
gion (from 0.1 to about 10 mg/l of solution; cf. Reuter and Perdue, 1977). In
classical terminology, “humic acids” are substances extracted from soils and sedi-
ments by alkaline solutions and precipitated through acidification, whereas “ful-
vic acids” refer to organic matter remaining in solution after acidification. These
substances are practically composed of polyelectrolytes varying in colour from
yellow to black, with molecular weights ranging from few a hundred to several
hundred thousand grams per mole. The relationships between macroscopic and
reactive properties for these classes of substances are outlined in figure 8.25.
The capacity of complexing of humic substances is ascribed to their oxygen-
based functional group (table 8.19).
The most common apical structure complexing metal cations in solution is
Figure 8.25 Classification and properties of humic substances. Reproduced with modifi-
cations from Stevenson (1983), with kind permission of Theophrastus Publishing and Pro-
prietary Co.
Geochemistry of Aqueous Phases / 563
Table 8.19 Oxygen-based functional groups in humic and fulvic acids (from Stevenson,
1982). Concentrations expressed in mEq/100 g.
Figure 8.26 (A) Functional groups of humic substances. (B) Complexing schemes for
fulvic substances. Reprinted from Stevenson (1983), with kind permission of Theophras-
tus Publishing and Proprietary Co.
shown in figure 8.26A. Figure 8.26B shows the most important complexing path-
ways followed by fulvic substances on copper, through the OH-phenolic and
COOH functional groups.
As shown in figure 8.25, the acidic-exchange capacity of humic substances
varies widely; a representative mean value is in the range 1.5 to 5 mEq/g. Assum-
ing a mean carbon content of 56% in polymer units, Stevenson (1983) deduced
that the ratio between complexed divalent cations and carbon atoms was around
1:20 to 1:60. Complexing capability is also greatly affected by the pH and ionic
strength of the solution (Guy et al., 1975). Generally, moreover, for elements
present in trace concentration levels, complexing by humic substances is more
efficient in pure water than in saline solutions, where organic matter is partly
bonded to the major dissolved cations (Ca2⫹, Mg2⫹; see Mantoura et al., 1978).
564 / Geochemistry of Fluids
Table 8.20 Metal-organic complexing (log 0) for some trace elements in
aqueous solution at pH ⫽ 8.2, T ⫽ 25 °C. I: ionic strength of solution.
References: (1) Turner et al. (1981); (2) Mantoura et al. (1978); (3)
Stevenson (1976).
Table 8.20 lists metal-organic complexing constants for some trace metals in wa-
ters of various ionic strengths, according to Stevenson (1976), Mantoura et al.
(1978), and Turner et al. (1981). The last column in table 8.20 lists the percentage
of chelate metal in seawater (I ⫽ 0.65), which is significant for Cu2⫹. According
to Turner et al. (1981), the metal-organic complexing in seawater is also signifi-
cant for rare earths and fully hydrolyzed elements forming stable complexes with
carbonate ions (cf. figure 8.16B).
In addition to their complexing capability, humic substances can also reduce
oxidized forms of metal cations and polyanions (Fe3⫹ → Fe2⫹; MoO2⫺ 4 → Mo ;
5⫹
VO3 → VO ; see Szilagyi, 1971). The reduced forms are then embodied in the
2⫹
HKF model (see section 8.11). The main observations on which the generaliza-
tions were based are as follows.
T(°C) T(°C)
Figure 8.27 Experimental values of standard partial molal heat capacity of sodium ace-
tate (A) and ethylene (B) in water as a function of T(°C) at pressures corresponding to
water-vapor equilibrium. Interpolating curves generated by the HKF model equations.
Reprinted from E. L. Shock and H. C. Helgeson, Geochimica et Cosmochimica Acta, 54,
915–946, copyright 䉷 1990, with kind permission from Elsevier Science Ltd., The Boule-
vard, Langford Lane, Kidlington 0X5 1GB, UK.
Figure 8.28 Correlation of standard partial molal properties of aqueous n-polymers with
the number of moles of carbon atoms in the alkyl chain. Reprinted from E. L. Shock and
H. C. Helgeson, Geochimica et Cosmochimica Acta, 54, 915–946, copyright 䉷 1990, with
kind permission from Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington
0X5 1GB, UK.
Geochemistry of Aqueous Phases / 567
The various polymerization steps along the series of alcohols may be visual-
ized in stereochemical representation as follows:
H H H H H H H H
H− C − OH+ C ⇔ H− C − C − OH+ C ⇔ H− C− C− C− OH.
(8.222)
H H H H H H H H
methanol ethanol propanol
The linearity in the standard partial molal properties of the aqueous polymers
as a function of the length of the chain (in terms of C atoms) observed by Shock
and Helgeson (1990) is the result of the principle of equal reactivity of co-
condensing functional groups (which, as we have already seen in section 6.1.2,
also holds for silica polymers). This principle is, however, strictly valid only when
the length of the chain is sufficiently elevated and small departures are observed
for chains with one or two carbon atoms (cf. figure 8.28).
The generalized behavior is consistent with an equation of the form
Ξ 0j = mn + p ( 8.223)
(Shock and Helgeson, 1990), where ⌶0j is the regressed standard state partial
molal property of the jth aqueous polymer, n is the number of moles of carbon
atoms in the alkyl chain, and m and p are, respectively, slope and intercept of the
regression lines shown in figure 8.28. Values of the regression constants for vari-
ous aqueous polymers according to Shock and Helgeson (1990) are listed in
table 8.21.
Table 8.21 Correlation parameters between the standard partial molal properties of
aqueous n-polymers and number of moles of carbon in the alkyl chain ( T ⫽ 25 °C, P ⫽ 1
bar; Shock and Helgeson, 1990). See section 8.11 and tables 8.11, 8.12, and 8.13 for
identification of symbols and units.
Property S P0 r , T r C P0 V P0 r , T r G ƒ0 , P r , T r H ƒ0 , P r , T r
n-hexylbenzene 40,390. ⫺25,590. 76.2 208.4 177. 25.7106 43.2732 13.8894 ⫺4.5678 180.5115 10.0338 ⫺0.8140
n-heptylbenzene 42,640. ⫺31,090. 82.9 229.6 192.8 27.8384 47.5238 14.2453 ⫺4.7435 197.8642 11.4670 ⫺0.9154
n-octylbenzene 44,690. ⫺36,760. 89.6 250.8 208.6 29.9662 51.7744 14.6016 ⫺4.9193 215.2173 12.9001 ⫺1.0169
Methanol ⫺42,050. ⫺58,870. 32.2 37.8 38.17 6.9383 5.5146 11.4018 ⫺3.0069 39.4852 ⫺1.4986 ⫺0.1476
Ethanol ⫺43,330. ⫺68,650. 35.9 62.2 55.08 9.2333 9.9581 12.1445 ⫺3.1906 60.0175 0.1507 ⫺0.2037
1-propanol ⫺41,910. ⫺75,320. 41.4 84.3 70.70 11.3426 14.3252 12.1079 ⫺3.3711 78.3142 1.6447 ⫺0.2869
1-butanol ⫺38,840. ⫺80,320. 46.9 104.4 86.60 13.4902 17.9400 14.1816 ⫺3.5205 94.8858 3.0034 ⫺0.3702
1-pentanol ⫺38,470. ⫺87,730. 53.4 125.2 102.68 15.6573 22.6230 13.6464 ⫺3.7141 111.9216 4.4095 ⫺0.4687
1-hexanol ⫺35,490. ⫺92,690. 64.8 144.4 118.65 17.7843 27.4289 12.5890 ⫺3.9128 126.8936 5.7074 ⫺0.6413
1-heptanol ⫺32,000. ⫺97,270. 72.1 174.2 133.43 19.7695 31.3977 12.9129 ⫺4.0769 151.5815 7.7219 ⫺0.7519
1-octanol ⫺30,250. ⫺103,060. 72.3 195. 149.2 21.9264 35.7093 13.2760 ⫺4.2551 169.4962 9.1280 ⫺0.7549
Phenol ⫺12,585. ⫺36,640. 45.8 75.3 86.17 13.4370 18.7373 11.8793 ⫺3.5535 69.9366 1.0363 ⫺0.3536
Acetone ⫺38,500. ⫺61,720. 44.4 57.7 66.92 10.8100 13.4912 11.4350 ⫺3.3366 54.9496 ⫺0.1534 ⫺0.3324
2-butanone ⫺36,730. ⫺67,880. 50.3 80.4 82.52 12.9145 17.6967 11.7839 ⫺3.5105 73.7083 1.3810 ⫺0.4217
2-pentanone ⫺34,390. ⫺73,460. 56.3 101.6 98.0 15.0021 21.8653 12.1372 ⫺3.6828 91.1588 2.8142 ⫺0.5126
2-hexanone ⫺32,480. ⫺79,220. 63.4 122.8 113.7 17.1141 26.0886 12.4806 ⫺3.8574 108.4560 4.2473 ⫺0.6201
2-heptanone ⫺30,430. ⫺84,890. 70.1 144.0 129.5 19.2419 30.3392 12.8366 ⫺4.0331 125.8091 5.6804 ⫺0.7216
2-octanone ⫺28,380. ⫺90,560. 76.8 165.2 145.3 21.3697 34.5960 13.1769 ⫺4.2091 143.1620 7.1135 ⫺0.8231
Formic acid ⫺88,982. ⫺101,680. 38.9 19.0 34.69 6.3957 4.6630 10.7209 ⫺2.9717 22.1924 ⫺3.1196 ⫺0.3442
Acetic acid ⫺94,760. ⫺116,100. 42.7 40.56 52.01 8.8031 12.4572 3.5477 ⫺3.2939 40.8037 ⫺0.9218 ⫺0.2337
Propanoic acid ⫺93,430. ⫺122,470. 49.4 56.0 67.9 10.9213 13.7115 11.4582 ⫺3.3457 54.313 ⫺0.246 ⫺0.4
Butanoic acid ⫺91,190. ⫺127,950. 56.1 80.5 84.61 13.1708 18.2050 11.8362 ⫺3.5315 72.9853 1.3878 ⫺0.5096
Pentanoic acid ⫺89,210. ⫺133,690. 62.8 103.3 100.5 15.3109 22.4858 12.1803 ⫺3.7085 91.7185 2.9291 ⫺0.6110
Hexanoic acid ⫺87,080. ⫺139,290. 69.5 125.2 116.55 17.4729 26.8090 12.5312 ⫺3.8872 109.6753 4.4095 ⫺0.7125
Heptanoic acid ⫺85,150. ⫺145,080. 76.2 146.4 132.3 19.5938 31.0465 12.8851 ⫺4.0624 127.0284 5.8426 ⫺0.8140
Octanoic acid ⫺83,400. ⫺151,050. 82.9 167.6 148.1 21.7216 35.2971 13.2410 ⫺4.2381 144.3812 7.2758 ⫺0.9154
Formate ⫺83,862. ⫺101,680. 21.7 ⫺22.0 26.16 5.7842 6.3405 3.2606 ⫺3.0410 17.0 ⫺12.4 1.3003
Acetate ⫺88,270. ⫺116,180. 20.6 6.2 40.5 7.7525 8.6996 7.5825 ⫺3.1385 26.3 ⫺3.86 1.3182
continued
Table 8.22 (continued )
Discrete values of the HKF model parameters for various organic aqueous
species are listed in table 8.22. Table 8.23 lists standard partial molal thermody-
namic properties and HKF model parameters for aqueous metal complexes of
monovalent organic acid ligands, after Shock and Koretsky (1995).
1 1 1
CO 2(gas) + H + + e − ⇔ C 6 H12 O 6 + H 2 O (8.224)
4 24 4
Figure 8.29 Sequence of redox equilibria mediated by biologic activity. From W. Stumm
and J. J. Morgan (1981), Aquatic Chemistry, copyright 䉷 1981 by John Wiley and Sons.
Reprinted by permission of John Wiley & Sons. The various equilibrium constants are
listed in table 8.24.
Table 8.24 Redox equilibria mediated by biological activity in eutrophic systems (from
Stumm and Morgan, 1981). pH ⫽ 7; [HCO3⫺] ⫽ 10⫺3; (g) ⫽ gaseous, (s) ⫽ solid.
Reduction log K
(A) 1/4 O2(g) ⫹ H⫹ ⫹ e⫺ ⇔ 1/2 H2O ⫹13.75
(B) 1/5 NO3⫺ ⫹ 6/5 H⫹ ⫹ e⫺ ⇔ 1/10 N2(g) ⫹ 3/5 H2O ⫹12.65
(C) 1/2 MnO2(s) ⫹ 1/2 HCO3⫺ ⫹ 3/2 H⫹ ⫹ e⫺ ⇔ 1/2 MnCO3(s) ⫹ H2O ⫹8.9
(D) 1/8 NO3⫺ ⫹ 5/4 H⫹ ⫹ e⫺ ⇔ 1/8 NH4⫹ ⫹ 3/8 H2O ⫹6.15
(E) FeOOH(s) ⫹ HCO3⫺ ⫹ 2H⫹ ⫹ e⫺ ⇔ FeCO3(s) ⫹ 2 H2O ⫺0.8
(F) 1/2 CH2O ⫹ H⫹ ⫹ e⫺ ⇔ 1/2 CH3OH ⫺3.01
(G) 1/8 SO42⫺ ⫹ 9/8 H⫹ ⫹ e⫺ ⇔ 1/8 HS⫺ ⫹ 1/2 H2O ⫺3.75
(H) 1/8 CO2(g) ⫹ H⫹ ⫹ e⫺ ⇔ 1/8 CH4(g) ⫹ 1/4 H2O ⫺4.13
(J) 1/6 N2(g) ⫹ 4/3 H⫹ ⫹ e⫺ ⇔ 1/3 NH4 ⫺4.68
Oxidation log K
⫹ ⫺
(L) 1/4 CH2O ⫹ 1/4 H2O ⇔ 1/4 CO2(g) ⫹ H ⫹ e ⫹8.20
(L-1) 1/2 HCOO⫺ ⇔ 1/2 CO2(g) ⫹ 1/2 H⫹ ⫹ e⫺ ⫹8.33
(L-2) 1/2CH2O ⫹ 1/2H2O ⇔ 1/2 HCOO⫺ ⫹ 3/2H⫹ ⫹ e⫺ ⫹7.68
(L-3) 1/2 CH3OH ⇔ 1/2 CH2O ⫹ H⫹ ⫹ e⫺ ⫹3.01
(L-4) 1/2 CH4(g) ⫹ 1/2 H2O ⇔ 1/2 CH3OH ⫹ H⫹ ⫹ e⫺ ⫺2.88
(M) 1/8 HS⫺ ⫹ 1/2H2O ⇔ 1/8 SO42⫺ ⫹ 9/8 H⫹ ⫹ e⫺ ⫹3.75
(N) FeCO3(s) ⫹ 2 H2O ⇔ FeOOH(s) ⫹ HCO3⫺ ⫹ 2H⫹ ⫹ e⫺ ⫹0.8
(O) 1/8 NH4+ ⫹ 3/8 H2O ⇔ 1/8 NO3⫺ ⫹ 5/4 H⫹ ⫹ e⫺ ⫺6.15
(P) 1/2 MnCO3(s) ⫹ H2O ⇔ 1/2 MnO2(s) ⫹ 1/2 HCO3⫺ ⫹ 3/2 H⫹ ⫹ e⫺ ⫺8.9
kJ
CH 3 COOH ⇔ 2 H 2 O + 2 C (solid) ∆G = −103 (8.225)
mole
kJ
CH 3 COOH ⇔ CH 4(gas) + CO 2 (gas) ∆G = −51 (8.226)
mole
kJ
CH 2 O ⇔ C (solid) + H 2 O ∆G = −107 . (8.227)
mole
4 H 2(gas) + CO 2(gas) → CH 4 + 2 H 2 O
Complex organic matter → organic acids
CH 3 COOH → CH 4 + CO 2(gas)
( 8.228 )
578 / Geochemistry of Fluids
kJ
SO24 − + 2 H+ + 3 H 2 S ⇔ 4 S(crystal) + 4 H2O ∆G = −27.7 . (8.229)
mole
As already mentioned, the Eh-pH extension of the field of native sulfur depends
on the total sulfur concentration in the system. With solute molality lower than
approximately 10⫺6 moles/kg, the field of native sulfur disappears.
We can calculate the equilibrium constant of reaction 8.230 under any given P-T
condition by computing the standard molal Gibbs free energy values of the vari-
ous components at the P-T of interest—i.e.,
− ∆G230
0
K 230,P ,T = exp . (8.231)
RT
Geochemistry of Aqueous Phases / 579
Applying the mass action law and assuming unitary activity for the water solvent
and the condensed phase, we obtain
K 230,P ,T = aK + ⋅ a Al 3 + ⋅ a 3 ⋅ a − 4+ . ( 8.232 )
H 4 SiO 04 H
( ) ( )
3 −4
Q = a K′ + ⋅ a Al
′ 3+ ⋅ a H′ 0 ⋅ a H′ + (8.233)
4SiO 4
will differ from equilibrium constant K, and the Q/K ratio will define the distance
from equilibrium of the new fluid with respect to K-feldspar at the P and T of
interest. In a generalized form, we can write
Q= ∏ ai′ v i
,
(8.234)
i
Q
log = log ∏ ai′ vi − log K . (8.235)
K i
Ratio Q/K is adimensional: if it is higher than 1 (i.e., log Q/K ⬎ 0), the aqueous
solution is oversaturated with respect to the solid phase; if it is zero, we have
perfect equilibrium; and if it is lower than 1 (i.e., log Q/K ⬍ 0), the aqueous phase
is undersaturated with respect to the solid phase.
Recalling now the concept of thermodynamic affinity (section 2.12), we can,
in a similar way, define an “affinity to equilibrium” as
Q
A = −2.303 RT log . (8.236)
K
This parameter has the magnitude of an energy (i.e., kJ/mole, kcal/mole) and
represents the energy driving toward equilibrium in a chemical potential field—
i.e., the higher A becomes, the more the solid phase of interest is unstable in
solution. In a heterogeneous system at equilibrium, based on the principle of
580 / Geochemistry of Fluids
equality of chemical potentials, all phases in the system are in mutual equilibrium.
It follows that equation 8.236 is of general applicability to all condensed phases,
and the value of affinity must be zero for all solids in equilibrium with the fluid.
It is immediately evident that equations 8.235 and 8.236 furnish a precise measure
of geochemical control of the degree of equilibrium attained by the fluid with
respect to a given mineralogical paragenesis. However, because the equilibrium
constant is a function of intensive variables T and P, plotting the affinity value
for various minerals under different T (or P) conditions against the T (or P) of
calculation provides precise thermometric (or barometric) information concern-
ing the equilibrium state of the fluid with respect to the reservoir. Figure 8.30A
shows a log Q/K ⫽ f (T ) diagram in the case of complete equilibrium between
mineralogical paragenesis and fluid at a given T condition (in this case, T ⫽
250 °C): note how the saturation curves for all solid phases converge to log
Q/K ⫽ 0 at the temperature of equilibrium.
Detailed speciation calculations for the fluid coupled with mineralogical in-
vestigation of the solid paragenesis generally allow a sufficiently precise estimate
of the T of equilibrium. As figure 8.30A shows, the chemistry of the fluid is
buffered by wall-rock minerals, so that the saturation curves of other phases not
pertaining to the system of interest are scattered (figure 8.28B).
The interpretation of log Q/K ⫽ f (T ) plots may be complicated by additional
processes affecting the chemistry of the fluid. The main complications result from
two processes that take place during the ascent of the fluid toward the surface:
adiabatic flashing and dilution by external aquifers.
Figure 8.30C shows the boiling effect on the fluid in equilibrium with the
paragenesis of figure 8.30A, calculated by Reed and Spycher (1984) assuming
subtraction of a gaseous phase amounting to 0.2% by weight of the initial fluid
and composed of 99.525% H2O, 0.451% CO2 , 0.013% H2S, and 0.011% H2 : the
increased molal concentration of dissolved silica displaces the intersection of the
saturation curve of quartz with the equilibrium condition log Q/K ⫽ 0 to a higher
T. Moreover, the change in pH, altering the speciation state of aluminum, iron
sulfides, and carbonates, modifies the shapes of the saturation curves of alkali
feldspars, muscovite, pyrite, and calcite.
The effect of dilution by shallow aquifers can also be quantified. It invariably
results in scattered saturation curves, proportional to the extent of contamination.
To the above considerations must be added the fact that equations of type
8.235 apply to pure compounds, whereas natural solids are mixtures of several
components. It follows that the activity of the solid component in hydrolysis is
never 1. Equation 8.235 must therefore be modified to take account of this fact:
Q
log = log ∏ ai′ vi − log K − log a K . (8.237)
K i
Geochemistry of Aqueous Phases / 581
aK in eq. 8.237 is the activity of the solid component in the mixture for which
hydrolytic equilibrium has been calculated. It must be emphasized that the effect
of dilution in a solid mixture is often quite limited: a change from activity ⫽ 1.0
to activity ⫽ 0.5 involves a displacement of the log Q/K factor of 0.3 (T, P, and
chemistry of the fluid being equal).
582 / Geochemistry of Fluids
The dissociation of a generic oxide Mv⫹Ov⫺ (where v+ and v⫺ are, respectively, the
cationic and anionic stoichiometric factors in the formula unit) can be expressed
by the general equilibrium
( )( )
v+ Z +
Mv +Ov − + v + Z + H+ ⇔
( ) ( ) H O + (v + ) M Z+
,
2 (8.241)
2
a2 ⋅ a H 2O
Na +
K 244 = . (8.242)
a Na 2O ⋅ a 2 +
H
a v +Z + ⋅ a
[ (v + )(Z + ) 2 ]
M H2 O
K = . (8.243)
(v + )(Z + )
aMv + Ov − ⋅ a
H+
Let us now rewrite equation 8.242 in logarithmic notation, assigning unitary ac-
tivity to the solvent:
Geochemistry of Aqueous Phases / 583
a +
ln a Na2 O = 2 ln Na − ln K 238 . (8.244)
aH +
a
( ) Z+
ln aMv + Ov − = v + ln M − ln K .
a (Z + )
(8.245)
H+
µ Na2 O = µ Na
0
+ RT ln a Na2 O . ( 8.246 )
2O
For two generic oxides i and j, the partial derivatives of equation 8.246 can be
combined (Helgeson, 1967) to give
a Z+
d ln M( i )
v(+i )
Z +( i )
aH
dµ i d ln ai +
= = + . (8.247)
dµ j d ln a j v( j ) a M Z +
d ln Z + ( j )
( j)
aH+
aM Z +
d ln Z + ( i )
(i )
aH+
where the last term on the right is the fugacity ratio of gaseous components.
Equation 8.248 is the basis for the construction of activity diagrams.
Let us now imagine a process of hydrothermal alteration of arkose sandstones
composed of Mg-chlorite, K-feldspar, K-mica, and quartz. Because precipitating
SiO2 during alteration is amorphous, we will assume the presence of amorphous
silica instead of quartz, and we will consider MgO as the generic oxide i and K2O
as the generic oxide j.
The alteration of muscovite to kaolinite is described by the reaction
584 / Geochemistry of Fluids
( )
2 KAl 3 Si3O10 OH + 2 H+ + 3 H2O ⇔ 3 Al 2 Si2O5 OH + 2K+.
2
( ) 4 (8.249)
muscovite kaolinite
Assigning unitary activity to the condensed phases and to solvent H2O, equilib-
rium constant K249 takes the form
a2 +
K 249 = K
, (8.250)
a2 +
H
1
log K 249 = log aK + − log aH + = log aK + + pH. (8.251)
2
( )
3 KAlSi3O8 + 2 H+ ⇔ KAl3Si3O10 OH + 2K+ + 6 SiO2 ,
2 (8.252)
K- feldspar muscovite
which yields
1
log K 252 = log aK + − log aH + = log aK + + pH. (8.253)
2
Mg 5 Al 2 Si 3 O 10 OH( ) 8
+ 10 H + ⇔ Al 2 Si 2 O 5 OH( ) 4
+ 5 Mg 2 + + SiO 2 + 7 H 2 O,
Mg- chlorite kaolinite
( 8.254 )
1
log K 254 = log a Mg 2 + + 2 pH . (8.255)
5
( )
3Mg5 Al 2 Si 3O10 OH + 2K + + 28H+ ⇔ 2KAl 3Si 3O10 OH
8
( ) 2
+ 15Mg2 + + 3SiO2 + 24H2O.
Mg- chlorite muscovite
(8.256)
Geochemistry of Aqueous Phases / 585
a 15
Mg 2 +
K 256 = . (8.257)
a2+ ⋅ a 28+
K H
a + aMg 2 +
2 log K + log K 256 = 15 log 2 . (8.258)
aH + a
H+
( )
Mg5 Al2 Si3O10 OH + 2 K+ + 8 H+ + 3 SiO2 ⇔ 2 KAlSi3O8 + 5 Mg2 + + 8 H2O .
8 (8.259)
Mg- chlorite K- feldspar
a5
Mg 2 +
K 259 = . (8.260)
a2+ ⋅ a8 +
K H
Again, multiplying and dividing by a2H⫹ and transforming into logarithmic nota-
tion, we obtain
a + aMg 2 +
2 log K + log K 259 = 5 log 2 . (8.261)
aH + a
H+
We can now build up the activity diagram, adopting as Cartesian axes log
aK⫹ ⫹ pH (abscissa) and log aMg2⫹ ⫹ 2pH (ordinate; see figure 8.31). The kaolin-
ite–muscovite equilibrium in such a diagram is defined by a straight line parallel
to the ordinate, with an intercept at 12 log K249 on the abscissa. Analogously, the
equilibrium between muscovite and K-feldspar is along a straight line, parallel to
the preceding one and with an intercept at 12 log K252 ; the kaolinite–Mg-chlorite
equilibrium lies along a straight line parallel to the abscissa, with intercept at
1
5 log K254 .
1
on the ordinate axis is 15 log K256 (cf. eq. 8.258). Analogously, for the Mg-
chlorite–K-feldspar equilibrium, the univariant curve has slope 25 and intercept
at 15 log K259 (cf. eq. 8.261). As shown in figure 8.31, the intersections of univariant
equilibria contour the stability fields of the various solids in reaction. A fluid of
composition A in figure 8.31 is in equilibrium with kaolinite; fluid B, falling on
the kaolinite–muscovite boundary at T ⫽ 350 °C, is clearly buffered by the alter-
ation of mica; and a fluid of composition C is in invariant equilibrium with the
original paragenesis. Translating the above observations into log Q/K ⫽ f (T )
notation (section 8.19), for a fluid of composition C we observe convergence of
the saturation curves of muscovite, K-feldspar, and Mg-chlorite toward log Q/
K ⫽ 0 at T ⫽ 350 °C (note that, because amorphous silica has fixed unitary
activity, the components are reduced to three in the computation of the degree of
Geochemistry of Aqueous Phases / 587
Table 8.25 Equilibrium constants used for construction of figure 8.31 (activity diagram).
freedom and, because P and T are fixed, three phases determine invariance in a
pseudoternary system).
To construct the diagram, we must know the various equilibrium constants
at the P and T of interest. It is thus necessary to compute the standard molal
Gibbs free energies of water, solid phases and aqueous ions in reaction. For solids,
the methods described in section 3.7 must be adopted. For aqueous ions, the
HKF model (section 8.11) or similar types of calculations are appropriate. To
avoid error progression and inconsistencies, the database used in calculating the
Gibbs free energies of both aqueous species and solid compounds must be inter-
nally consistent. In the case of figure 8.31, the constants were calculated at differ-
ent T and P conditions along the liquid-vapor univariant curve using the database
of Helgeson et al. (1978) (computer package SUPCRT ). The resulting values are
listed in table 8.25. It must be emphasized that extensive compilations of activity
diagrams in chemically complex systems are already available in the literature.
For instance, the compilation of Helgeson et al. (1976) for the system MgO-CaO-
FeO-Na2O-K2O-Al2O3-SiO2-CO2-H2O-H2S-H2SO4-Cu2S-PbS-ZnS-Ag2S-HCl
presents activity diagrams drawn for various T conditions (0, 25, 60, 100, 150,
200, 250, 300°C; P ⫽ 1 bar) and is an excellent tool for interpreting water-rock
equilibria at shallow depths.
The arguments treated in the two preceding sections were developed in terms of
simple equilibrium thermodynamics. The “weathering” of rocks at the earth’s
surface by the chemical action of aqueous solutions, and the complex water-rock
interaction phenomena taking place in the upper crust, are irreversible processes
that must be investigated from a kinetic viewpoint. As already outlined in section
2.12, the kinetic and equilibrium approaches are mutually compatible, both being
based on firm chemical-physical principles, and have a common boundary repre-
sented by the “steady state” condition (cf. eq. 2.111).
of one of the main ions involved in the process. Denoting as mi the molality of
solute ion i and as ni the number of moles of i in solution, molality is related to
the mass of solvent by
ni
mi = Ω, (8.262)
nW
where nw are the moles of water solvent and ⍀ are the moles of H2O in 1 kg of
water (⍀ ⫽ 55.51). Differentiating equation 8.262 with respect to time t gives
(Delany et al., 1986), where dnw /dt is the consumption (or production) rate of
water molecules in the system. This term is often negligible with respect to the
other elemental balances.
The rate of a generic reaction j is represented by its absolute velocity v j, which
may be obtained from experiments, provided that the mass of solvent is held
constant throughout the experiment and no additional homogeneous or heteroge-
neous reactions concur to modify the molality of the ion in solution (Delany et
al., 1986):
nW dmi
vj ≅ × , (8.264)
Ωvi , j dt
Vs dc
vj ≅ × i. (8.265)
vi , j dt
When more than one reaction affects the mass of the ion i in solution (e.g., disso-
lution and/or precipitation of heterogeneous solid assemblages; homogeneous re-
actions in solution), the overall change in mass of component i is given by the
summation of individual absolute reaction velocities v j, multiplied by their respec-
tive stoichiometric reaction coefficients:
T
dni
dt
= ∑ vi , j vj . (8.266)
j =1
Geochemistry of Aqueous Phases / 589
dni
= vi , j vj . (8.267)
dt j
dmi Ω T vW , j
= ∑ vi , j − vj (8.268)
dt nW j =1 nW
(Delany et al., 1986), where vW, j is the number of water molecules involved in the
jth reaction.
dξ j
vj = . (8.269)
dt
In section 2.12 we recalled the significance of the “overall reaction progress variable”
which represents the extent of advancement of simultaneous irreversible processes in
heterogeneous systems (Temkin, 1963). The overall progress variable is related to progress
variables of individual reactions by
T
ξ = ∑ ξj (8.270)
j =1
(Delaney et al., 1986). It follows from equations 8.269 and 8.270 that we can also define
an “overall reaction rate” v, which is related to individual velocities by
T
v= ∑ vj (8.271)
j =1
(Delany et al., 1986). The “relative rate of the jth reaction” j with respect to the overall
reaction progress variable is given by
dξ j
σj = (8.272)
dξ
vj
σj = (8.273)
v
and
dξ
v= . (8.274)
dt
We can generally group the rate laws adopted to describe dissolution and
precipitation kinetics into three main categories (Delany et al., 1986):
dmiT s
( ),
n
= k + mieq − mi (8.275)
dt Ω
where dmTi is the differential increment of total molality of ion i in solution (free
ion plus complexes), k⫹ is the rate constant of forward reaction (i.e., dissolution;
cf. eq. 2.110), s is the active surface area, and n is an integer corresponding to the
degree of reaction. The corresponding equation for precipitation is
dmiS s
( ),
n
− = k − mi − mieq (8.276)
dt Ω
where mSi is the total molality of i in the precipitating solid and k is the rate
constant of the backward reaction.
The simplest case of compositional dependence is the zero-order reaction, in
which the concentration gradient is not affected by concentration. Denoting the
molar concentration of the ith element (or component) as ci (and neglecting sur-
face area and volume of solution effects), we have
∂ ci
= −k+ . (8.277)
∂ t T,P
Geochemistry of Aqueous Phases / 591
ci ,t = ci0 − tk + , (8.278)
∂ ci
= − kci ,t . (8.279)
∂ t T,P
( )
ci ,t = ci0 exp − kt . ( 2.280)
In this case, the rate constant has units of t⫺1 and is usually termed “half-life of
the reaction” (t1/2 ), the amount of time required to reduce the initial concentra-
tion c0i , to half:
ln 2
t1 2 = . (8.281)
k
As we will see in detail in chapter 11, the production of radiogenic isotopes fol-
lows perfect first-order reaction kinetics.
In molar notation, and referencing to the equilibrium concentration cieq as-
suming n ⫽ 1 (i.e., first-order reaction), equation 8.275 can be translated into
s
dci
( ).
n
= k + cieq − ci (8.282)
dt Vs
Assuming that s and Vs remain constant throughout the process, the solution of
a first-order equation in the form of equation 8.282 (n ⫽ 1) is
c eq − ci s
ln ieq = − k+ t ,
0 Vs
(8.283)
ci − ci
Figure 8.32 Rates of dissolution of biogenic opal in seawater at T ⫽ 25 °C, for various
amounts of suspended opal (percent mass). Reprinted from Hurd (1972), with kind per-
mission from Elsevier Science Publishers B.V, Amsterdam, The Netherlands.
relationship. T and Vs being equal, the negative slope depends on the percent
amount of suspended opal (hence on the active surface area), but remains con-
stant throughout the process.
By “active surface area” we meant the kinetically active part of the total surface area.
According to Helgeson et al. (1984), this area is restricted to etch pits. Alternative esti-
mates of surface areas may be obtained from measurements of specific surface area s#
whenever solid particles have a narrow size range. The specific surface area for spherical
particles is given by
6
s# = , (8.284)
ρD
where is grain density and D is diameter. For cubic particles, the same equation applies
after substitution of edge length a for diameter:
Geochemistry of Aqueous Phases / 593
6
s# = . (8.285)
ρa
Experimental measurements of surface area are usually carried out by BET gas absorp-
tion techniques (see Delany et al., 1986, for an appropriate treatment).
In some case the rate laws in meq or ceq are of high order—i.e., n ⬎ 1. If ci is
lower than the equilibrium concentration cieq, the solution of equation 8.282 is
( n − 1)s k t.
(c ) ( )
1− n 1− n
i
eq
− ci − cieq − ci0 = + (8.286)
Vs
To deduce the degree of reaction, one may use equation 8.286 with different inte-
ger values of n. The value of n that yields a straight line in a (cieq ⫺ ci )1⫺n versus
t plot is the order of reaction.
The equilibrium concentration cieq to be introduced for this purpose in equa-
tion 8.286 is the one asymptotically achieved in a ci versus t plot, at high values
of t (see Lasaga, 1981a, for a more appropriate treatment of high-order reaction
kinetics).
(Lasaga, 1984), the rates of dissolution can be expected to increase with aH⫹ in
acidic solutions. The experiments, in fact, denote the pH dependency of rate con-
stants as simple proportionalities of type:
( )
v+
k + ∝ a H+ , (8.288)
v j = s ∑ ki ∏ ak
v k,i
. (8.289)
i k
Plummer et al. (1978) utilized, for instance, an equation in the form of equation
8.289 to describe the dissolution of calcite. According to the authors, the simulta-
neous reactions concurring to define the absolute velocity of dissolution are
Figure 8.33 shows in detail the effect of the single rate constants on the for-
ward velocity at the various pH-PCO2 conditions (T ⫽ 25 °C). Although the indi-
vidual reactions 8.290.1–8.290.3 take place simultaneously over the entire compo-
sitional field, the bulk forward rate is dominated by reactions with single species
in the field shown: away from steady state, reaction 8.290.1 is dominant, within
the stippled area the effects of all three individual reactions concur to define the
overall kinetic behavior, and along the lines labeled 1, 2, and 3 the forward rate
corresponding to one species balances the other two.
Geochemistry of Aqueous Phases / 595
Figure 8.33 Reaction mechanism contribution to the total rate of calcite dissolution re-
action as a function of pH and PCO2 at 25 °C. From L. N. Plummer, T. M. L. Wigley, and
D. L. Pankhurst (1978), American Journal of Science, 278, 179–216. Reprinted with per-
mission of American Journal of Science.
HCO −3 ⇔ H + + CO 23 −
CaCO 3 ⇔ Ca 2 + + CO 23 − ( 8.292.3 )
we have
k1 K 2 k2 K 2 k3 K 2
k4 = + a HCO − + a OH− , (8.293)
KC KC K1 3 K C KW
K2
k4 = k1 +
K C
1
a H+ 2 3
(
k2 a H CO 0 + k 3 .
) (8.294)
The forward rate constants (cm/sec) are semilogaritmic functions of reciprocal tem-
perature (K) according to
444
log k1 = 0.198 − (8.295.1)
T
2177
log k2 = 2.84 − (8.295.2)
T
log k 3 = −5.86 −
317
T
(T < 298 K ) (8.295.3)
log k 3 = −1.10 −
1737
T
(T > 298 K ) . (8.295.4)
The corresponding activation energies deduced from the Arrhenius slope (see section
8.21.2) are listed in table 8.27.
SiO 2 → SiO 02 .
solid aqueous
(8.296)
[
SiO2 + v H2O → SiO2 ⋅ v H2O * → SiO20 + v H2O, ] (8.297)
solid aqueous
Because the energy of the activated complex is higher than the energies of
reactants and of the final reaction products, it is crucial to establish how many
activated complexes may form at given P-T-X conditions. This can be done by
application of quantum chemistry concepts.
Limiting the system to all accessible states of energy for the molecules of interest, the
probability of finding an individual molecule B in the state of energy j is
Pj =
(
exp − ε j kT ),
(8.298)
Q
− Ei
Q= ∑ exp
kT
(8.299)
i
(cf. section 3.1). The partition function Q is related to the partition function qB relative to
the individual molecule B by
−ε j
qB = ∑ exp
kT
(8.300)
j
N
Q = q B B = QB N B ! (8.301)
(Lasaga, 1981b), where NB is the total number of B molecules in the system. Because
the chemical potential is related to the partial derivative of the Helmholtz free energy at
constant volume:
∂F ∂ ln(Q N B !)
µB = = − kT (8.302)
∂N B T ,V ∂N B T ,V
recalling the Stirling approximation (eq. 3.116) and applying equation 8.301, we obtain:
qB
µ B = − kT ln (8.303)
NB
v A A + vB B → vC C + vD D. ( 8.304 )
vA vB vC vD
q q q q
− kT ln A − kT ln B = − kT ln C − kT ln D . (8.305)
NA NB NC ND
N Cv C ⋅ N Dv D q Cv C ⋅ q Dv D
= . (8.306)
v v
N A A ⋅ N Bv B q A A ⋅ q Bv B
Finally, dividing both sides of equation 8.306 by the volume of the system, we obtain
(q ) ⋅ (q )
vC vD
c Cv C ⋅ c Dv D C V D V
K = = . (8.307)
c v A ⋅ c Bv B (q V ) ⋅ (q V)
vA vB
A
A B
Equation 8.307 is the basis of all applications founded on the TST approach.
A
v j = sk + q + 1 − exp + (8.308)
ηRT
A
− v j = sk − q − 1 − exp + . (8.309)
ηRT
Equation 8.308, representing the forward reaction rate, is written in terms of for-
ward direction parameters: A⫹ is thermodynamic affinity (cf. section 2.12), q⫹ is
the kinetic activity product for the left side of the microscopic reaction (i.e., in
the case of reaction 8.297, q⫹ ⫽ aSiO2 ⭈ avH2O ), and is a stoichiometric factor
relating the stoichiometric coefficient of the solid reactants in the macroscopic
and microscopic reaction notations (in the case of reactions 8.296 and 8.297, this
term is 1, because both reactions have the same stoichiometric coefficient.
Figure 8.34 Arrhenius plot of reaction rates for equilibrium of analcite ⫹ quartz ⇔ al-
bite in NaCl, and NaDS-bearing solutions, after Matthews (1980). Preexponential factors
and activation energies can be deduced from the fitting expression. NaDS ⬅ Na2Si2O5.
−Ea
k = k 0 exp . (8.310)
RT
∂ ln k E
=− a.
( )
∂1T R (8.311)
Plotting the natural logarithm of the reaction rate against the reciprocal of abso-
lute temperature (Arrhenius plot), the activation energy of the reaction is repre-
sented by the slope of equation 8.311 at the T of interest multiplied by ⫺R. If Ea
is constant over T, the semilogarithmic plot gives rise to a straight line, as illus-
trated in figure 8.34.
Activation energies of some important geochemical reactions are listed in
table 8.27.
600 / Geochemistry of Fluids
Table 8.27 Activation energies of geochemical reactions (kJ/mole) (after Lasaga, 1981a, 1984)
Because reaction rates are also pressure-dependent (although they are much
less P-dependent than T-dependent), the volume of activation of reaction Va can
be deduced in an analogous fashion from the partial derivative
∂ ln k V
=− a . (8.312)
∂P RT
Ca2⫹, and Na⫹. Their concentration ratios are remarkably constant, as is the bulk
Geochemistry of Aqueous Phases / 601
amount of salts. This parameter is usually expressed as salinity (S) and represents
the percent by weight of dissolved inorganic matter per 103 g of seawater. In the
salinity calculation, the amount of Br⫺ and I⫺ is replaced by an equivalent
amount of Cl ⫺, and carbonate ions HCO⫺ 2⫺
3 and CO3 are converted into oxides.
The chlorinity of seawater (Cl ) is the amount of Cl⫺ and other halogens content
(expressed in g per kg of seawater), equivalent to the total amount of alkalis. This
parameter is practically determined by titration with AgNO3 , and the current
definition of chlorinity is “weight in g of Ag necessary to precipitate the halogens
(Cl ⫺, Br⫺, I⫺ ) in 328.5233 g of sea water” (Stumm and Morgan, 1981). Table
8.28 lists the mean composition of seawater at salinity 35. The first data column
lists the amount of salts in g per kg of solution. In column (2), weight is related
to chlorinity; column (3) shows the molarity/chlorinity ratio.
The salinity of seawater ranges from 33 to 37. Its value can be derived directly
from chlorinity through the relationship
However, salinity values are easily obtained with a salinometer (which measures
electrical conductivity and is appropriately calibrated with standard solutions and
adjusted to account for T effects). The salinity of seawater increases if the loss of
H2O (evaporation, formation of ice) exceeds the atmospheric input (rain plus
rivers), and diminishes near deltas and lagoons. Salinity and temperature concur
antithetically to define the density of seawater. The surface temperature of the
sea reflects primarily the latitude and season of sampling. The vertical thermal
profile defines three zones: surface (10–100 m), where T is practically constant;
thermoclinal (100–1000 m), where T diminishes regularly with depth; and abyssal
602 / Geochemistry of Fluids
(⬎ 1000 m), with constant low T (2–4 °C). Density at atmospheric level is usually
denoted by parameter :
kg
σ = Density 3
− 1 × 1000 (8.314)
dm
The HCO⫺ 3 amount listed in table 8.28 actually represent carbonate alkalinity, calcu-
lated neglecting CO32⫺. The alkalinity [Alk] of a solution represents its capacity to neutral-
ize strong acids in solution. Alkalinity can be defined as “the sum of equivalents of all
species whose concentrations depend on the concentration of H⫹ ions in solution, minus
the concentration of H⫹ ions in solution.” For a solution containing borates and carbon-
ates, alkalinity can be expressed as
where the terms in square brackets denote molality. If we consider only carbonate equilib-
ria, equation 8.315 reduces to
and, in the absence of CO32⫺ (at the normal pH of seawater, the carbonate ion is virtually
absent; see figure 8.21),
Figure 8.35 shows the redox state and acidity of the main types of seawaters.
The redox state of normal oceanic waters is almost neutral, but they are slightly
alkaline in terms of pH. The redox state increases in aerated surface waters. Sea-
waters of euxinic basins and those rich in nutrients (eutrophic) often exhibit Eh-
pH values below the sulfide-sulfate transition and below carbonate stability limits
(zone of organic carbon and methane; cf. figure 8.21). We have already seen (sec-
tion 8.10.1) that the pH of normal oceanic waters is buffered by carbonate equi-
libria. At the normal pH of seawater (pH ⬵ 8.2), carbonate alkalinity is 2.47
mEq per kg of solution.
Geochemistry of Aqueous Phases / 603
Figure 8.35 Mean Eh-pH values for various types of seawaters. A ⫽ normal oceanic
waters; B ⫽ oxidized surface waters; C ⫽ euxinic basins; D ⫽ eutrophic waters. Dashed
line: field of natural waters according to Baas Becking et al. (1960).
Table 8.30 Mean composition of seawater for a salinity of 35 (Turekian, 1969). Values
expressed in ppb.
Figure 8.36 Changes in quantities of phosphorus, nitrogen, and silicon in seawater with
depth.
Figure 8.37 REE concentration profiles with depth in Atlantic and Pacific Oceans.
Table 8.30 shows the chemistry of seawater compiled by Turekian (1969) for
major, minor, and trace constituents, expressed in parts per billion (ppb) at a
mean salinity of 35. The listed values are estimates of mean amounts in solution,
whereas elemental concentrations actually vary with depth. The most conspicu-
ous variations are observed in the first 200 m from the surface, where photosyn-
thetic processes are dominant and phosphorus and nitrogen are fixed by plankton
and benthos, as well as silica and calcium, which constitute, respectively, the skel-
etons of planktonic algae (diatom) and the shells of foraminifera and mollusks.
Figure 8.36 shows how the elemental amounts of P (generally present as com-
⫺
plexes of the phosphate ion PO3⫺ 4 ), N (present as NO3 complexes), and Si vary
with depth in the various oceanic basins. Concentrations are expressed in micro-
moles per liter of solution. Note the relative constancy of concentrations below
1000 m.
Trace elements also exhibit systematic changes in concentration with depth.
Figure 8.37 shows, for instance, concentration profiles of rare earth elements
(REE) determined by De Baar et al. (1985) in Pacific and Atlantic waters. Note
that the concentration profiles differ for the various elements in the series; in
particular, the amount of Ce is quite high in surface waters of the Atlantic Ocean.
Geochemistry of Aqueous Phases / 607
steady state, the inputs of matter exactly balance the outputs. In the “seawater
system,” inputs are constituted primarily by:
dC
J d = −D × . (8.319)
dX
Ja = ρ × V . ( 8.320 )
Geochemistry of Aqueous Phases / 609
Figure 8.39 Box model for steady state chemistry of seawater. Numbers in boxes: 14C/
12
C ratio normalized to atmospheric value. L: specific transfer rate between atmosphere
and ocean; R: transfer rates between oceanic reservoirs. Reprinted from W. S. Broecker,
R. D. Gerard, M. Ewing, and B. C. Heezen, Geochemistry and physics of ocean circula-
tion. In Oceanography, M. Sears, ed., copyright 䉷 1961 American Association for the
Advancement of Science. Reprinted with permission of the American Association for the
Advancement of Science, Washington.
Steady state models are based on the concept of the mean time of residence
of the elements in the system. Mean time of residence for a steady state is
defined as
M
τ = ,
∑ Ji (8.322)
i
where mass M and fluxes have the same units (e.g., M ⫽ moles, Ji ⫽ moles/year).
The “fractional time of residence” is relative to a given input:
M
τi = , (8.323)
Ji
610 / Geochemistry of Fluids
1 1 1 1
= + +L . (8.324)
τ τ1 τ 2 τn
Indexing input and output fluxes as Jin and Jout, respectively, the mass varia-
tion in time dt is given by
dM
= J in − J out . (8.325)
dt
Assuming the removal rate of a given element from the system to be proportional
(at first approximation) to the elemental abundance in the system, through pro-
portionality constant Kt :
J out = K t × M , (8.326)
dM
= 0; J in = K t × M , (8.327)
dt
1 M
= = τ. (8.328)
Kt J in
Geochemistry of Aqueous Phases / 611
The time of residence thus represents the reciprocal of the constant of re-
moval from the system. The times of residence of elements in the “seawater sys-
tem” are generally evaluated on the basis of equation 8.327. The main input flux is
fluvial: the fractional time of residence based on fluvial input is thus the primary
constituent of the mean time of residence (cf. eq. 8.322). An alternative way of
evaluating the time of residence is based on output flux due to sedimentation,
applying equation 8.326 (the two methods give concordant results). Table 8.31
lists times of residence of elements in seawater, expressed in years and arranged
in decreasing series. Long times of residence (e.g., those of Cl, Br, Na, B, and
Mg) are indicative of low reactivity in the system; short times of residence are
connected with oversaturation and high reactivity (e.g., La, Pb, Al, Fe, and Th).
CHAPTER NINE
PV = nRT , (9.1)
n= ∑ ni (
= n X1 + X 2 + X 3 + L + X i , ) (9.2)
i
where X1 , X2 , . . . ,Xi are the molar fractions of the various gaseous species.
The activity of a perfect gas, as for any substance, is unitary, by definition, at
standard state. Moreover, for a perfect gas, activity is (numerically) equivalent to
pressure, at all pressures. Let us consider the relationship existing, with T held
constant, between the chemical potential of component i in gaseous phase g at 1
bar ( i,1,T,g ) and at pressure P ( i,P,T,g ):
P
µ i , P,T , g = µ i ,1,T , g + ∫ Vi , g dP. (9.3)
1
Because for a pure phase partial molar volume V i,g is equivalent to molar volume
V g , and since we have defined as standard state the condition of the pure compo-
nent at P ⫽ 1 bar and T of interest:
µ i ,1 , T , g ≡ µ i0, g , ( 9.4 )
612 /
Geochemistry of Gaseous Phases / 613
we have
P
µ i , P,T , g = µ i0, g + ∫ Vg dP. (9.5)
1
Because
RT ln a i , g = µ i , P , T , g − µ io, g , ( 9.6 )
P
dP
RT ln a i , g = ∫ RT
P
, (9.7)
1
RT ln a i , g = RT ln P. ( 9.8)
Hence
a i , g = P. ( 9.9)
P
RT ln a i , g = RT ln (9.10)
Pr
and
P
ai,g = . (9.11)
Pr
A real gas does not obey the equation of state of perfect gases (eq. 9.1) be-
cause of the interactions among gaseous molecules. The higher the pressure and
the lower the temperature, the greater the deviation from ideality. At very low P,
intermolecular distances are great and interactions between gaseous molecules
are virtually absent. For almost all gaseous species of geochemical interest, the
approximation to a perfect gas is acceptable only over a very small P range.
Interaction forces among gaseous molecules are generally subdivided into
614 / Geochemistry of Fluids
Figure 9.1 Molecular interaction potentials in Stockmayer’s (1941) model for H2O vapor.
(a): antiparallel dipolar moments; (b): parallel dipolar moments. Reprinted from D.
Eisemberg and W. Kauzmann, The Structures and Properties of Water, 1969, by permission
of Oxford University Press.
µ 2 f 4 εσ 6 4 εσ 12
U AB = − − + . (9.12)
R3 R6 R 12
The first term on the right in the sum of potentials in equation 9.12 is the electro-
static term (represented here as a simple dipole moment). Factor f is the angular
function of dipole orientation:
( )
f = sin θ A sin θ B cos φ A − φ B − 2 cos φ A cos φ B . (9.13)
Angles and in equation 9.13 define the dipole orientation with respect to a
coordinate system centered on the oxygen nuclei. The second term on the right
Geochemistry of Gaseous Phases / 615
P
RT ln fi = RT ln P + ∫ (Vi )
− Vperfect dP , (9.14)
Pr
where V i is the molar volume of the real gas and V perfect is the volume occupied
by one mole of perfect gas under the same P and T conditions. The relationship
between fugacity f i and pressure Pi of a real gas i is given by
fi = Pi Γi . (9.15)
( )
P Vi − Vperfect dP
Γi = exp ∫ . (9.16)
P RT
r
fi
ai = , (9.17)
fi 0
where fi0 is fugacity at standard state. Fugacity fi0 ⫽ 1 only when Pi ⫽ 1 bar and
T and P are such that ⌫i ⬅ 1 (cf. eq. 9.15). If the adopted standard state is, for
instance, pure gas at the P and T of interest and P is high (or T is low), fi0 ⬆ 1,
and hence ai ⬆ f i .
Figure 9.2 shows how fugacity coefficient ⌫i of H2 , O2 , and CO2 behaves with
increasing P at various T (0 °C, 100 °C, 200 °C). Tables 9.1 and 9.2 list discrete
values attained by ⌫i for H2O and CO2 gases under various P and T conditions.
Values of ⌫i for gases of geochemical interest for which experimental data are
Figure 9.2 Fugacity coefficients of H2, O2, and CO2 plotted as functions of P at various
T. Reprinted from Garrels and Christ (1965) with kind permission from Jones and Bartlett
Publishers Inc. Copyright 䉷 1990.
Geochemistry of Gaseous Phases / 617
Table 9.1 Fugacity coefficients of CO2 under various P and T conditions (Mel’nik, 1972).
T (K)
P (atm) 400 500 600 700 800 900 1000 1100 1200 1300 1400 1500
500 0.57 0.85 1.02 1.11 1.15 1.17 1.19 1.20 1.20 1.20 1.20 1.20
1,000 0.63 0.95 1.16 1.28 1.34 1.37 1.38 1.38 1.37 1.37 1.37 1.37
1,500 0.82 1.20 1.44 1.56 1.62 1.62 1.62 1.59 1.59 1.57 1.55 1.53
2,000 1.16 1.60 1.85 1.95 2.00 1.95 1.91 1.86 1.82 1.78 1.76 1.71
3,000 2.46 3.00 3.18 3.15 2.95 2.08 2.46 2.34 2.29 2.19 2.09 2.04
4,000 5.49 5.60 5.60 5.18 4.67 3.89 3.55 3.31 3.16 2.95 2.75 2.63
5,000 12.5 11.3 9.88 8.52 6.92 5.75 5.01 4.57 4.27 3.89 3.63 3.39
6,000 28.4 22.1 17.4 13.9 10.5 8.32 7.24 6.31 5.75 5.13 4.68 4.27
7,000 64.2 42.7 30.2 22.4 16.2 12.6 10.5 8.91 7.94 6.92 6.17 5.50
8,000 143 81.4 51.8 35.7 24.6 18.6 15.1 12.6 10.7 9.12 7.94 7.08
9,000 319 154 88.1 56.3 38.0 28.8 22.4 17.8 14.8 12.3 10.5 9.33
10,000 703 289 148 80.9 56.2 41.7 31.6 24.6 20.0 16.2 13.5 11.8
Table 9.2 Fugacity coefficients of H2O ( ⌫i ⫻ 102; Helgeson and Kirkham, 1974).
P (kbar)
T (°C) 0.5 1 2 3 4 5 6 7 8 9
200 3.66 2.36 0.19 0.20 0.24 0.30 0.38 0.50 0.66 0.89
250 8.59 5.46 0.43 0.45 0.51 0.62 0.77 0.97 1.25 1.62
300 16.59 10.50 0.82 0.83 0.92 10.9 13.2 16.3 20.5 25.9
350 27.63 17.52 13.5 13.5 14.8 17.1 20.3 24.6 30.3 37.5
400 40.83 26.22 20.2 19.9 21.6 24.5 28.7 34.3 41.4 50.4
450 54.16 35.94 27.8 27.3 29.3 32.9 38.0 44.7 53.3 64.0
500 64.81 45.89 36.1 35.3 37.6 41.8 47.8 55.6 65.4 77.6
550 72.59 55.28 44.5 43.5 46.1 51.0 57.7 66.5 77.5 90.9
600 78.48 63.58 52.7 51.7 54.6 60.0 67.5 77.1 89.0 103.4
650 82.97 70.57 60.3 59.5 62.8 68.6 76.7 87.0 99.0 115.0
700 86.46 76.33 67.3 66.8 70.4 76.7 85.3 96 109.5 125.4
800 91.37 84.87 78.9 79.5 83.9 91.0 100.4 112 126.3 142.9
not available in the literature can be obtained through the “universal chart of
gases” (figure 9.3). Construction of this chart is based on the fact that the thermo-
dynamic properties of all real gases are similar functions of P and T, when the
parameters of the equation of state are reduced with respect to the values of
critical temperature Tc, critical pressure Pc, and critical volume Vc of the com-
pound of interest. We can thus introduce the concepts of reduced pressure Pr ,
reduced temperature Tr , and reduced volume Vr:
Pi
Pr,i = (9.18)
Pc,i
Figure 9.3 Universal chart of gases. From Garrels and Christ (1965); reprinted with kind permission from Jones and Bartlett Publishers
Inc. Copyright 䉷 1990.
Geochemistry of Gaseous Phases / 619
Tc
Ti
Tr,i = (9.19)
Tc,i
Vi
Vr,i = . (9.20)
Vc,i
Values of critical temperature and critical pressure for gaseous species of geo-
chemical interest are listed in table 9.3.
The ⌫i values allow the calculation of fugacity and, thence, the chemical po-
tential of the gaseous component at the P and T of interest, applying
f
µ i , g = µ i0, g + RT ln i0 . (9.21)
fi
620 / Geochemistry of Fluids
As we have already seen, the universal chart of gases assumes that all gaseous
species exhibit the same sort of deviation from ideal behavior at the same values
of Tr , Pr , and Vr . This fact, known as the principle of corresponding states, is ana-
lytically expressed by the “deviation parameter” (or “compressibility factor”) Zg .
For n ⫽ 1,
Zg =
PV
RT
(
= f Pr , Tr . ) (9.22)
Z g ( P,T ) = 1 + BP + CP 2 + DP 3 + L (9.23)
Z g (V ,T ) = 1 + B ′V + C ′V −2 + D ′V −3 + L. (9.24)
Figure 9.4 (A) Reduced isotherms of N2, CO2, and H2O vapor according to reduced
pressure. (B) Comparison between experimental isotherms of H2O and Tr predicted by
principle of corresponding states.
Geochemistry of Gaseous Phases / 621
Virial expansion coefficients B, C, D and B⬘, C⬘, D⬘ in equations 9.23 and 9.24 are
simple functions of T and are specific to each gaseous species. Virial expansion
equations such as equations 9.23 and 9.24 do not reproduce the behavior of real
gases in the vicinity of the critical point and under high P and T conditions with
satisfactory precision.
Deviations from ideality of real gases can also be expressed through the van
der Waals equation:
(P + aV )(V − b) = RT ,
−2
(9.25)
4
b= πd 3 (9.26)
3
32 3
b= π rg , (9.27)
3
P +
a
( )
V − b = RT .
( )
V V + b T 1 / 2
(9.28)
Parameters a and b in equation 9.28 have the same significance as in equation 9.25. How-
ever, whereas in the van der Waals equation these parameters are constant for a given
gaseous species, in the Redlich-Kwong formulation they are often assumed to be variable.
For instance, Holloway (1977) adopted a coefficient a varying with T:
a = f ( T ). ( 9.29 )
Touret and Bottinga (1979) and Bottinga and Richet (1981) proposed
a 3 a 6
a = a 1 3 − 3 + a 2 (9.30)
V V
622 / Geochemistry of Fluids
and
V −1
b = ln + b1 b2 , (9.31)
a3
a = c + dV −1 + eV −2 ( 9.32 )
c = c1 + c2 T + c 3 T 2 ( 9.33 )
d = d 1 + d 2 T + d 3T 2 (9.34)
e = e1 + e2 T + e 3 T 2 (9.35)
a = a 1 + a 2 T + a 3T −1 (9.36)
b=
(1 + b P + b P + b P ) .
1 2
2
3
3
(b + b P + b P )
4 5 6
2 (9.37)
More recently, Saxena and Fei (1987) proposed quantifying deviations from ideality
of real gases through an expression they defined as “nonvirial”:
Z g = A + BP + CP 2 + L. (9.38)
The difference between equation 9.38 and the corresponding virial expansion of equation
9.23 lies in term A, which substitutes 1 and which is a function of T, as are the remaining
coefficients. Equation 9.38 also finds application in a “corresponding states” notation:
Table 9.4 lists the parameters of the nonvirial expansion of equation 9.38 for H2O
and the parameters of the corresponding states notation for the remaining gaseous species.
Table 9.4 Parameters of nonvirial expansion of Saxena-Fei (from Ganguly and Saxena, 1987).
P
RT ln fi = ∫ (Vi )
− Vperfect dP + RT ln P , (9.40)
1
The calculation of the Gibbs free energy of a gaseous mixture can be performed,
starting from the chemical potentials of the various pure gaseous components at
the P and T of interest, assuming that the mixture is ideal:
R 12 =
1
2
(
R 11 + R 22 ) (9.43)
σ 12 =
1
2
(
σ 11 + σ 22 ) (9.44)
Geochemistry of Gaseous Phases / 625
( )
12
ε 12 = ε 11ε 22 (9.45)
( )
12
µ12 = µ11 µ 22 . (9.46)
Once potential U12 , with substitution of the various parameters in equation 9.12,
has been derived, an interaction parameter W12 can be obtained by applying
1
W12 = z U 12 − U 11 + U 22
2
( ) (9.47)
Saxena and Fei (1988) have shown that the observed deviations from ideality in
H2O-CO2 mixtures can be obtained from interaction parameter W12 by applying
RTW12 X 1X 2
Gexcess mixing = ,
(X q 1 1 + X2 q 2 ) (9.48)
where q1 and q2 are the “efficacious volumes” of the gaseous molecules H2O and CO2 .
Because these parameters cannot be derived easily from first principles, the authors
adopted the simplifications
q1 V1
= (9.49)
q2 V2
and
q 1 + q 2 = 1, ( 9.50 )
2
q 1X 1
ln γ 1 = q 1W12 1 + (9.51)
q2 X 2
and
2
q X
ln γ 2 = q 2W12 1 + 2 2 . (9.52)
q 1X 1
Note that these coefficients are real “activity coefficients” and not “fugacity coefficients”
(see section 9.1).
626 / Geochemistry of Fluids
Table 9.5 Thermodynamic parameters for main gaseous species of geochemical interest.
H 0f, P r , Tr in kJ/mole: S 0P r , T r in J/(mole ⫻ K). CP ⫽ K1 ⫹ K2T ⫹ K3T 2 ⫹ K4T ⫺1/2 ⫹ K5 T ⫺2
(data from Robie et al., 1978). Tr ⫽ 298.15 K, Pr ⫽ 1 bar
Once the excess Gibbs free energy of mixing is known, the Gibbs free energy
of the gaseous mixture is calculated as follows:
The first term implies knowledge of the chemical potentials of the gaseous com-
ponents at standard state. Thermodynamic parameters for the main gaseous spe-
cies of geochemical interest at T ⫽ 298.15 K and P ⫽ 1 bar are listed in table 9.5.
The molar volume in these T-P conditions for all species is that of an ideal gas—
i.e., 24,789.2 cm3 (or 2478.92 J/bar). The limiting T for the CP function is 1800
K. Data are from Robie et al. (1978).
Figure 9.6 Ternary atomic composition diagram for volcanic gases. Crosses: samples
from Mount Etna. From Gerlach and Nordlie (1975a), American Journal of Science, 275,
353–76. Reprinted with permission of American Journal of Science.
H : O = 2 :1 (9.54)
with stoichiometry
(C + S) : O = 2 : 1. ( 9.55)
628 / Geochemistry of Fluids
Figure 9.7 O-H-S-C compositional tetrahedron. From Gerlach and Nordlie (1975a),
American Journal of Science, 275, 353–76. Reprinted with permission of American Journal
of Science.
This indicates that the major molecular species present in volcanic gases are H2O,
CO2 , and SO2 . Some samples (mainly from Mount Etna’s emissions in 1970) also
contain appreciable amounts of H2 molecules.
Figure 9.7 shows the O-H-S-C compositional tetrahedron. In this representa-
tion, the compositional limits of volcanic gases are composed of vertical planes
with atomic ratios
C :S = 1: 2 (9.56)
and
C : H = 1 : 1, (9.57)
respectively, the H-O-C plane, and the surfaces established by the hematite-
magnetite (HM) and quartz-fayalite-magnetite (QFM) oxygen buffers:
1
3 Fe2O3 ⇔ 2 Fe3O4 + O2 (9.58)
2
Geochemistry of Gaseous Phases / 629
1
Ni+ O2 ⇔ NiO (9.60)
2
1
2 FeO+ O2 ⇔ Fe2O3 . (9.61)
2
Ar ⇔ Ar .
gas melt (9.63)
This can be conceived as “mechanical solubility,” which does not imply either
reactivity or modifications of the molecular structure of the compound of inter-
est. Several gaseous species do not exhibit solubility of the type exemplified by
equation 9.63, but more or less complex reactions with melt components. For
instance, in the case of CO2 ,
CO 2 + O 2 − ⇔ CO 23 − .
gas melt melt
(9.64)
CO 2 ⇔ CO 2
gas melt
(9.65)
632 / Geochemistry of Fluids
and
CO 2 + O 2 − ⇔ CO 23 − .
melt melt melt
(9.66)
There are several ways of expressing the solubilities of gaseous species in melts or
liquids. The Bunsen coefficient ␣g defines “the volume of reduced gas at T ⫽ 273.15 K,
P ⫽ 1.013 bar adsorbed by a unit volume of solvent at T of interest and P ⫽ 1.013 bar.”
The equation used to calculate the Bunsen coefficient is
273.15 Pg 1 1.013
α g = Vg × ,
1.013 Vmelt Pg
(9.67)
T
where Vg is the volume of reduced gas adsorbed at the T of interest, Vmelt is the volume of
solvent (melt, in our case), and Pg is the partial pressure of the gaseous species in equilib-
rium with the solvent. Note that equation 9.67 reduces to
Vg 273.15
αg = × . (9.68)
Vmelt T
The Ostwald coefficient Lg is “the ratio between volume of adsorbed gas and volume
of solvent, measured at the same temperature”:
Vg T
Lg = = αg . (9.69)
Vmelt 273.15
Henry’s constant Kh of the gas represents the ratio between the partial pressure of the
species in the gaseous phase and the molar fraction of the species in solution:
Pg
Kh = . (9.70)
X g , melt
Note that Henry’s constant differs from equilibrium constant K, which represents the ac-
tivity ratio between the two phases at equilibrium:
a g , gas Γg Pg 1 Γg
K = = × = Kh × , (9.71)
a g , melt f g0, gas X g , melt γ g , melt fg 0, gas γ g , melt
Geochemistry of Gaseous Phases / 633
where ⌫g is the fugacity coefficient in the gas phase and f 0g is the standard state fugacity.
In the case of homogeneous equilibria such as equilibrium 9.66, the relationship between
thermodynamic constant and Henry’s constant is obviously more complex.
In geochemistry, the solubilities of gaseous species in melts are usually expressed
through Henry’s constant or as fractional solubilities in moles (molar fraction of gaseous
species in the gas phase over molar fraction of gaseous species in the melt).
H 2 O + O 0 ⇔ 2OH.
gas melt melt
(9.72)
Because H2O is a basic oxide (i.e., it tendentially dissociates free oxygen; cf. sec-
tion 6.1.3), reaction 9.72 may appear contradictory. However, reaction 9.72
clearly represents the summation of partial reactions:
H 2 O ⇔ H 2 O.
gas melt
(9.73)
H 2 O ⇔ 2H + + O 2 −
melt melt melt
(9.74)
O 2 − + O 0 ⇔ 2O −
melt melt melt
(9.75)
2O − + 2H+ ⇔ 2OH.
melt melt melt
(9.76)
O− O− H O− O−
/
O − − Si − O − Si − O − + O ⇔ O − − Si − OH + O − − Si − OH. (9.77)
\
O− O− H O− O−
Figure 9.9 Amounts of molecular H2O and OH hydroxyls determined in hydrated silicate
glasses by IR spectroscopy. Reprinted from E. Stolper, Geochimica et Cosmochimica Acta,
46, 2609–2620, copyright 䉷 1982, with kind permission from Elsevier Science Ltd., The
Boulevard, Langford Lane, Kidlington 0X5 1GB, UK.
H 2 O + O 0 ⇔ 2OH
melt melt melt
(9.78)
is not completely displaced rightward. Stolper (1982) has shown that the weight
proportions of H2O and OH in melts vary nonlinearly with increasing bulk
amount of H2O. When this exceeds 5%, molecular water is dominant. Figure 9.9
shows the proportions of OH and molecular H2O in melts with increasing
amounts of dissolved H2O, experimentally determined by infrared (IR) spectros-
copy. The observed distributions are consistent with an equilibrium constant K78
between 0.1 and 0.3.
The value of constant K78 is determined in the following manner (Stolper, 1982). The
constant of equilibrium 9.78 is given by the activity ratio
2
a OH, melt
K 78 = . (9.79)
a H 2O, melt a O 0 , melt
It is assumed, at first approximation, that the activities of the three species in homoge-
neous equilibrium correspond to the relative molar fractions
nH 2O, melt
a H 2O, melt ≡ X H 2O, melt = (9.80)
nH 2O, melt + nOH, melt + nO 0 , melt
nOH, melt
a OH, melt ≡ X OH, melt = (9.81)
nH 2O, melt + nOH, melt + nO 0 , melt
Geochemistry of Gaseous Phases / 635
nO 0 , melt
a O 0 , melt ≡ X O 0 , melt = . (9.82)
nH 2O, melt + nOH, melt + nO 0 , melt
If N1 moles of component H2O mix with N2 moles of an anhydrous melt whose stoichio-
metric formula contains r moles of oxygen (i.e., r ⫽ 8 for NaAlSi3O8 , r ⫽ 2
for SiO2 , and so on), because one mole of O 0 is necessary to form two moles of OH
hydroxyls, we obtain
2 2
X OH, melt nOH, melt
K 78 = = ,
X H 2O, melt ⋅ X O 0 , melt (N 1 )(
− 0.5nOH, melt rN 2 − 0.5nOH, melt ) (9.85)
Figure 9.11 Relationship between fugacity of gaseous component H2O and amount of
same component dissolved in an albitic melt at equilibrium at T ⫽ 1100 °C (K78 ⫽ 0.2).
Interpolants in various P conditions determine value of K72, based on equation 9.87. Re-
printed from E. Stolper, Geochimica et Cosmochimica Acta, 2609–2620, copyright 䉷 1982,
with kind permission from Elsevier Science Ltd., The Boulevard, Langford Lane, Kidling-
ton 0X5 1GB, UK.
Figure 9.11 shows the effect of pressure on the solubility of H2O in an albitic
melt at T ⫽ 1100 °C, based on the experiment of Burnham and Davis (1974). At
low H2O amounts, a marked linear correlation between fugacity of H2O in the
gaseous phase and the squared molar fraction of H2O component in the melt is
observed. Note that the abscissa axis in figure 9.11 is X 2H2Ototal, melt and thus differs
from the same axis in figure 9.9. Let us now again consider equations 9.72, 9.73,
and 9.78. Clearly, the constant of heterogeneous equilibrium 9.72 is given by the
product of partial equilibria 9.73 and 9.78—i.e.,
K 72 = K 73 × K 78 . (9.86)
Geochemistry of Gaseous Phases / 637
2
K 72 rN 22 − N 12 X H 2O,melt
× = (9.87)
4 f H0 O,gas N + N 2
2 (
1 2 )
f H 2O,gas
(Stolper, 1982). If the total amount of H2O in the melt (XH2Ototal, melt ) is suffi-
ciently low, it is also true that N1 ⬍⬍ N2 and the relationship between
X 2H2Ototal, melt and f H2O,gas is simply linear (cf. eq. 9.87 and figure 9.11).
The interpolation of experimental points in figure 9.11 allows constant K72 to
be determined at each pressure. This exercise, when repeated at the various T of
interest, leads to the construction of the Arrhenius plot of figure 9.12 (Fraser,
1975b). The same figure also shows the Arrhenius trend relative to equation 9.73,
calculated for K78 ⫽ 0.2 at T ⫽ 1100 °C and K78 ⫽ 0.1 at T ⫽ 700 °C (Stolper,
1982)—i.e.,
ln K 73 = ln K 72 − ln K 78 . (9.88)
Note that constant K73 is the reciprocal of Henry’s constant when fugacity co-
efficient ⌫g ⫽ 1.
Lastly, figure 9.13 shows that the linear relationship between the square root
of activity in the gaseous phase and the amount of H2O component in the melt
638 / Geochemistry of Fluids
Figure 9.13 Relationship between activity of H2O in gaseous phase and molar amount
of H2O in melt at equilibrium. Experimental data from Burnham and Davis (1974) (䡵);
Fraser (1975b) (䊱); Kurkjian and Russel (1958) (䢇). Reprinted from B. J. Wood and D. G.
Fraser, Elementary Thermodynamics for Geologists, 1976, by permission of Oxford Univer-
sity Press.
is generally observed in all types of silicate (and also phosphate) melts (Kurkjian
and Russel, 1958; Burnham and Davis, 1974; Burnham, 1975; Fraser, 1975b,
1977).
Neglecting N1 in equation 9.87, we obtain
X H 2O,melt r 12
= K 72 . (9.89)
a 1H2O,gas 2
2
CO2 + O2 − ⇔ CO23 − .
melt melt melt
(9.90)
However, not all the CO2 is present in the melt in the form of carbonate ion:
molecular CO2 is also present, based on the heterogeneous equilibrium
Geochemistry of Gaseous Phases / 639
Figure 9.14 Weight concentration of molecular CO2, carbonate ion CO32⫺, and total CO2
dissolved in albitic melts as a function of P, under various T conditions. Reproduced with
modifications from E. Stolper, Geochimica et Cosmochimica Acta, 46, 2609–2620, copy-
right 䉷 1982, with kind permission from Elsevier Science Ltd., The Boulevard, Langford
Lane, Kidlington 0X5 1GB, UK.
CO2 ⇔ CO2 .
gas melt
(9.91)
( ) ( )
CO2 + X O 2 − ⇔ X CO 23 −+ 1 − X CO2 , ( ) (9.92)
gas melt melt melt
where X and (1 ⫺ X ) are molar fractions of molecular CO2 and carbonate ion in
the melt, respectively, with respect to the total amount of CO2 . IR spectroscopy
measurements on albitic glasses (Stolper et al., 1987) show that, at constant P,
the solubility of molecular CO2 in the melt decreases with increasing T, whereas
the solubility of carbonate ion increases: the net result is that bulk solubility is
practically unaffected by temperature. This fact, based on equilibrium 9.92, can
be interpreted as increased reactivity of free oxygen in the melt, with increasing
T. As P increases, with T held constant, we observe increased solubility of carbon-
ate ion and, more appreciably, of molecular CO2 (figure 9.14).
The solubility of sulfur dioxide in melts may be ascribed to two concomi-
tant processes:
3
SO2 + O2 − ⇔ S2 − + O2
2 (9.93)
gas melt melt gas
640 / Geochemistry of Fluids
Figure 9.15 Solubility of noble gases in silicate melts as a function of kinetic radius of
gaseous molecule (A) and temperature (B). Ordinate axis in part A: natural logarithm of
Henry’s constant; in part B, natural logarithm of equilibrium constant. Reprinted from
G. Lux, Geochimica et Cosmochimica Acta, 51, 1549–1560, copyright 䉷 1987, with kind
permission from Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington 0X5
1GB, UK.
and
1
SO2 + O2 + O2 − ⇔ SO24 − .
2 (9.94)
gas gas melt melt
Equilibria 9.93 and 9.94 are similar to those proposed by Fincham and Richard-
son (1954) for native sulfur. According to these equilibria, the solute states of
sulfur compounds depend on oxygen fugacity; the total concentration of sulfur
dissolved in the melt initially decreases with increasing fO2, based on equation
9.93, and then increases again when equilibrium 9.94 becomes dominant.
Table 9.8 Standard molar enthalpies (⌬H 0i ) and standard entropies (⌬S 0i ) of solution of noble
gases in silicate melts, after Lux (1987). (1) leucite-basanite; (2) tholeiite; (3) alkali-olivine basalt.
( ⌬H i0 ) (kcal/mole) ( ⌬S i0 ) (cal/mole)
amount of SiO2 and decreases with the amounts of MgO and CaO in the melt,
increases with melt viscosity (in agreement with the chemical effect; White et al.,
1989) and decreases with increasing density. The latter fact is in agreement with
essentially “mechanical” solubility (cf. section 9.6)—i.e., solubility is a function
of the free volume available within the liquid which in turn is (approximately)
inversely proportional to density. Lux (1987) furnished the parameters of a linear
equation which reproduces the Henry’s constant of the various gases, based on
density of melt, within the field 2.1 to 2.7 g/cm3 (Table 9.7). The resulting Henry’s
constant is in cm3 of reduced volume per g of melt/atm ⫻ 10⫺5. With P, T, and
composition held constant, solubility decreases with increasing atomic number
and can be correlated directly to the kinetic radius of noble gas ri through an
empirical equation of the type
(Kirsten, 1968), as shown in figure 9.15A. With P and composition held constant,
solubility increases slightly with T. Figure 9.15B shows an Arrhenius plot of Hen-
ry’s constant for an alkaline basalt melt in the T range 1250 ⬍ T (°C) ⬍ 1500
(Lux, 1987). From equality of potentials at equilibrium:
we can derive
∆H i0 ∆Si 0
ln K i = − + . (9.97)
RT R
By applying equation 9.97 to plots of the type shown in figure 9.15B, Lux (1987)
obtained standard molar enthalpies ⌬H 0i and standard molar entropies ⌬S i0 of
solution of noble gases in melts, as listed in table 9.8. Enthalpies of solution are
positive but close to zero, within error approximation.
Lastly, figure 9.16 shows the effects of pressure on the solubility of argon in
642 / Geochemistry of Fluids
Figure 9.16 Effects of pressure on solubility of argon in silicate melts. From White et al.
(1989). Reprinted with permission of The Mineralogical Society of America.
∆H 0 1 1 V0
ln K P ,T − ln K Pr ,Tr = −
R
− −
T Tr RT
P − Pr ( ) (9.98)
for a standard state of pure gas Ar at P ⫽ 15 kbar and T ⫽ 1600 °C, and assuming
Ar to mix ideally in the liquid over a matrix composed of Ar and O2⫺. The inter-
polation parameters of equation 9.98 are listed in Table 9.9.
Geochemistry of Gaseous Phases / 643
Gas A0 A1 A2 A3 A4 A5
He 6830.52 ⫺187,444.82 1.974201 ⫺2738.127 ⫺687.751 889.593
Ne 1421.175 ⫺55,338.362 0.185867 ⫺516.808 ⫺20.0201 27.2899
Ar 3145.791 ⫺87,149.49 ⫺516.808 ⫺1270.795 ⫺468.632 615.388
Kr 1098.849 ⫺32,333.73 ⫺20.0201 ⫺442.515 ⫺230.154 310.802
Xe 1074.854 ⫺34,604.31 27.2899 ⫺420.535 ⫺113.399 147.401
9.17, but also O2 , H2 , CH4 , NH3 , H2S, and CO2 ), the two solubility branches of
the trend, ascending and descending, are virtually straight. Because early experi-
ments on the solubility of gaseous species in water were limited to low T (typically
below 50 °C), solubility was described by means of empirical relationships accu-
rately reproducing low-T behavior—for instance, the Benson-Krause equation:
( ) Si
( )
2
ln K h,i = A T1 /T − 1 − T1 /T − 1 , (9.99)
R
1
2
A1 1
K h,i = A0 + + A2T + A3 log T + A 4 log + A5 log
T ρ ρ (9.100)
× 10 4 × Γi .
(
ln Γi = − ∫ 1 − Z g ) dP
P
. (9.101)
0
Still more recently, Giggenbach (1980) proposed simple linear equations re-
lating the molar distribution constants of the main reactive gases to T expressed
in Celsius:
Geochemistry of Gaseous Phases / 645
Species A B
NH3 1.4113 ⫺0.00292
H2S 4.0547 ⫺0.00981
CO2 4.7593 ⫺0.01092
CH4 6.0783 ⫺0.01383
H2 6.2283 ⫺0.01403
N2 6.4426 ⫺0.01416
X i,g
log K i′ ≡ log = A + BT . (9.102)
X i ,l
These constants, listed in Table 9.11, were derived by Giggenbach (1980) from
the respective Henry’s constant by applying:
Z g , H 2O
K i′ = K h,i × , (9.103)
PH 2O
where Zg,H2O is the compressibility factor of gaseous H2O. Because the resulting
values of K⬘ vary in a simple exponential fashion with T (Figure 9.18), the maxi-
mum in Henry’s constant observed for the various noble gases in Figure 9.17 can
be simply ascribed, in light of equation 9.103, to the compressibility factor of
H2O vapor.
To understand fully the relationship between Henry’s constant and the molar distribu-
tion constant of a given species between a gas and a liquid phase, we must recall the
concept of the “compressibility factor” of a gas (see section 9.3):
PV
Z g ,i = i i
. (9.104)
ni RT
ni
X i, g = . (9.105)
n i + n H 2O ( g )
ni ni
X i, g = ≈ . (9.106)
n i + n H 2O ( g ) n H 2O ( g )
646 / Geochemistry of Fluids
Pi Z g ,H 2O
X i, g = × . (9.107)
PH 2O Z g ,i
X i, g Pi Z g ,H 2O 1 Z g ,H 2O 1
K i′ = = × × = K h, i × × . (9.108)
X i, l X i, l Z g ,i PH 2O Z g ,i PH 2O
In the transition K⬘ → Kh,i, the compressibility factor of the trace component in the gas-
eous phase is usually neglected because, with the concentration so low, it is virtually 1 (see
eq. 9.23).
At shallower depths, the vapor phase separates from the liquid phase and
migrates toward the surface, sometimes heating superficial aquifers and finally
giving rise to fumarole activity. The gaseous phase is dominated by H2O mixed
with subordinate amounts of CO2 , H2S, CH4 , H2 , N2 , and NH3 . Table 9.13 lists
representative analyses of geothermal gases from the main geothermal fields of
the world. It also lists the pressure of separation of the gaseous phase from the
liquid, the steam fraction in the fluid (FH2O, f ), and the “fraction of gas in steam”
(Xg ). The latter definition is purely operative, because H2O vapor is a gas to all
intents and purposes and the gaseous phase is not unmixed at the various P and
T of interest. The need to distinguish the total amount of non-H2O gaseous com-
648 / Geochemistry of Fluids
Table 9.13 Gas composition in geothermal fluids (from Henley, 1984). Psep ⫽ pressure of
separation of gaseous phase. nd ⫽ not determined.
Concentration in Gas
(millimoles/mole total gases)
(1) Wairakei, New Zealand; (2) Tauhara, New Zealand; (3) Broadlands, New Zealand; (4) Ngawha, New
Zealand; (5) Cerro Prieto, Mexico; (6) Mahio-Tongonan, Philippines; (7) Reykjanes, Iceland; (8) Salton
Sea, California; (9) Geysers, Wyoming; (10) Larderello, Italy.
ponents from gaseous H2O (i.e., “steam”) is dictated by the fact that, as the gas-
eous phase flows through superficial aquifers, further enrichment in H2O may
take place by simple evaporation.
Cl 1
=
C 0 1 + Fg K ′ − 1 ( ) (9.109)
Cg K′
=
C0 (
1 + Fg K ′ − 1 ) (9.110)
[ ( )]
Cl −n
= 1 + ∆ Fg K ′ − 1 (9.111)
C0
Cg K′
=
[1 + ∆F ( K ′ − 1)] (9.112)
n
C0
g
Geochemistry of Gaseous Phases / 649
Cl
C0
≈ exp − K ′F g( ) (9.113)
Cg
C0
(
≈ K ′ exp − K ′F g . ) (9.114)
Equations 9.109 and 9.110 are valid for a boiling process in a closed system: the
gaseous phase develops from a liquid with initial concentration C 0 , Fg is the mass
fraction of developed gas, and K ⬘ is the mass distribution constant of the compo-
nent of interest between gas and liquid. Equations 9.111 and 9.112 describe a
multistage separation process in which n is the number of separation stages, ⌬Fg
is the mass fraction of gas separated in each stage, and K ⬘ is the mean mass distri-
bution constant of the process. Equations 9.113 and 9.114 refer to a boiling pro-
cess in an open system: the gas is continuously removed from the system as the
process advances.
The following assumptions are implicitly accepted when adopting equations
9.109 to 9.114.
Assumption 2 may lead to substantial errors, because the salinity of the liquid
normally increases during boiling. The consequent increase in ionic strength
modifies the activity coefficients of solutes. In particular, if the gaseous species
are polar and nonpolar, the activity coefficients of solutes diverge as the process
advances (see section 8.7).
Let us now consider speciation in the gaseous phases. According to the litera-
ture, there are three main reactions controlling speciation in the gaseous phase:
2NH 3 ⇔ N 2 + 3H 2
gas gas gas
(9.116)
H2 S ⇔ S + H2 .
gas solid gas
(9.117)
9323
log K 115 = 10.76 − (9.118)
T
5400
log K 116 = 11.81 − . (9.119)
T
Assuming the gaseous phase to be ideal, we can substitute partial pressures for
the thermodynamic activities of the components:
PCO2 ⋅ PH4
K 115 = 2
(9.120)
PCH 4 ⋅ PH2
2O
PN2 ⋅ PH3
K 116 = 2
2
. (9.121)
PNH
3
Pi = X i , g Pt , (9.122)
X CO2 ⋅ X H4
K 115 = 2
⋅ Pt2 (9.123)
X CH 4 ⋅ X H2
2O
X N2 ⋅ X H3
K 116 = 2
2
⋅ Pt2 . (9.124)
X NH
3
Pt = PH 2O + ∑ Pi , (9.125)
i
where the Pi terms are the partial pressures of gaseous components other than
H2O, assuming
∑ Pi = Pt X g F g K CO
′ 2,
(9.126)
i
where Xg is the fraction of gas (non-H2O) in the system and Fg is the fraction of
vapor plus non-H2O gas in the system, one has:
Geochemistry of Gaseous Phases / 651
PH 2O
Pt = .
(1 − X ′ 2
g F g K CO ) (9.127)
The vapor pressure along the univariant curve of pure water can be expressed as
a semilogarithmic function of the reciprocal of absolute temperature (Giggen-
bach, 1980):
2048
log PH2 O = 5.51 − . (9.128)
T
For aqueous solutions with mean saline amounts, a similar expression holds:
1820
log PH2 O = 4.90 − (9.129)
T
( )
T K =
5227
(
log X CH 4, g − log X CO 2, g − 4 log X H 2, g − 0.26 + 4 log 1 − X g F g K CO
′ 2 )
(9.130)
( )
T K =
1304
2 log X NH
3, g
− log X N
2, g
(
− 3 log X H2 , g + 0.79 + 2 log 1 − X g F g K CO
′ 2 )
( 9.131)
CO2 + H2 ⇔ CO + H2O
gas gas gas liquid
(9.133)
652 / Geochemistry of Fluids
(substituting H2O vapor for liquid H2O in the two equilibria and subtracting from equilib-
rium 9.132 equilibrium 9.133 multiplied by 4, we return to equilibrium 9.115).
The substitution of H2O vapor for liquid H2O simplifies the equilibrium. Because the
activity of liquid H2O is almost 1, we obtain
4
X CO, g
K 132 = .
X CH 4 , g ⋅ X CO
3 (9.134)
2,g
Based on thermodynamic data (Barin and Knacke, 1973), Chiodini and Cioni (1989) ob-
tained
13606
log K132 = 8.065 − , (9.135)
T
( )
T K =
log X CH
13, 606
+ 3 log X CO 2 , g − 4 log X CO, g + 8.065
.
(9.136)
4 ,g
As we see, Pt does not appear in equation 9.136 but can be calculated, based on equilib-
rium 9.133 and assuming
Pt ≈ PH 2O + PCO 2 (9.137)
2094
log Pt = 9.083 − + log XCO, g − log XH2, g (9.138)
T
Let us now consider equilibrium 9.117. The formation of H2S obviously im-
plies the presence of sulfur in the system (but not necessarily native sulfur). The
main crystalline sulfur compound in geothermal fields is always pyrite FeS2 . Equi-
librium 9.117 can thus be rewritten as follows:
1 7
CaSO4 + FeS2 + 3H2O + CO2 ⇔ CaCO3 + Fe3O 4 + 3H2S + O2 .
crystal crystal liquid gas crystal
3 crystal gas
3 gas (9.141)
The thermometric equation of D’Amore and Panichi (1980) has the form
( )
T K =
XCH , g XH , g
24, 775
X
.
2 log 4
− 6 log 2
− 3 log H2 S, g − 7 log PCO 2 + 36.05
XCO , g XCO , g XCO , g
2 2 2
(9.142)
In table 9.14, for the sake of completion, we list the thermodynamic parameters
of the HKF model concerning neutral molecules in solution (Shock et al., 1989).
Calculation of partial molal properties of solutes (see section 8.11), combined
with calculation of thermodynamic properties in gaseous phases (Table 9.5),
allows rigorous estimates of the various equilibrium constants at all P and T
of interest.
PART FOUR
Methods
CHAPTER TEN
The presence of trace elements in minerals can be ascribed to at least four main
phenomena (McIntire, 1963):
1. Surface adsorption. Foreign ions are kept in a diffuse sheet at the surface
of the crystal, as a result of electrostatic interaction with surface atoms
whose bonds are not completely saturated.
2. Occlusion. Adsorbed impurities on the surface of the crystal are trapped
by subsequent strata during crystal accretion.
3. Substitutional solid solution with a major component. The trace element
substitutes for a major element in a regular position of the crystal lattice.
4. interstitial solid solution. Similar to the preceding phenomenon, but here
the trace element occupies an interstitial position in the crystal lattice.
The first two processes may be operative at ultratrace concentration levels. The
first one is relevant whenever the mineral has a high surface-to-mass ratio, as
in the case of colloids. The last two processes are by far the most important in
geochemistry and can be appropriately described in thermodynamic terms.
As we saw in section 3.8.1, Raoult’s law describes the properties of an ideal solu-
tion, in which the thermodynamic activity of the component is numerically equiv-
alent to its molar concentration:
a i = Xi . (10.1)
/ 657
658 / Methods
Figure 10.1 (A) Activity–molar concentration plot. Trace element concentration range
is shown as a zone of constant slope where Henry’s law is obeyed. Dashed lines and ques-
tion marks at high dilution: in some circumstances Henry’s law has a limit also toward
infinite dilution. The intercept of Henry’s law slope with ordinate axis defines Henry’s law
standard state chemical potential. (B) Deviations from Nernst’s law behavior in a logarith-
mic plot of normalized trace/carrier distribution between solid phase s and ideal aqueous
solution aq. Reproduced with modifications from Iiyama (1974), Bullettin de la Societée
Francaise de Mineralogie et Cristallographie, 97, 143–151, by permission from Masson
S.A., Paris, France. ⌬ in part A and log ⌬ in part B have the same significance, because
both represent the result of deviations from Henry’s law behavior in solid.
µ i = µ i0 + RT ln γ i X i , (10.2)
µ i = µ i∗ + RT ln Xi . (10.3)
µ∗ − µ0 ∆µ ∗
γ h,i = exp i i
≡ exp
i
. (10.4)
RT RT
Trace Element Geochemistry / 659
The excess Gibbs free energy term ⌬i* corresponds to the amount of energy
required to transfer one mole of trace component i from an ideal solid solution
to the solid solution of interest. As shown in figure 10.1, the more i* differs from
0i , the more angular coefficient ␥h,i differs from 1, which represents the condition
of Raoult’s law. Trace element distribution between phases of geochemical interest
is governed by Henry’s law only within restricted solid solution ranges (here we
use the term “solution” instead of “mixture,” because the Henry’s law standard
state of the trace element differs from the standard state adopted for the carrier,
which is normally that of pure solvent at the P and T of interest: cf. section 2.1).
XA
Xs = (10.5)
XB
mA
Xaq = (10.6)
mB
as the molality ratio in aqueous solution, the equilibrium distribution for two
ideal solutions is given by
X aq = K ( P,T ) Xs , (10.7)
(
log ∆ = log X aq − log K + log Xs . ) (10.8)
660 / Methods
(
ln X aq = ln K + ln X s + 1 − 2X A,s ) RT
W
(10.9)
2W
log Xaq = log K′ + log Xs − XA, s log e . (10.10)
RT
X aq = KX s (10.11)
is replaced by
X I,s = K I X aq (10.12)
X II,s = K II X aq , (10.13)
Figure 10.2 Deviations from Nernst’s law in crystal–aqueous solution equilibria, as ob-
tained from application of various thermodynamic models. (A and B) Regular solution
(Iiyama, 1974). (C) Two ideal sites model (Roux, 1971a). (D) Model of local lattice distor-
tion (Iiyama, 1974). Reprinted from Ottonello (1983), with kind permission of Theophras-
tus Publishing and Proprietary Co.
types I and II in the crystal (m ⫹ n ⫽ 1), the relationship between partial distribu-
tion constants KI, KII, and K is
mK I + nK II + X aq K I K II
K = .
(
1 + X aq nK I + mK II ) (10.14)
Figure 10.3 Two ideal sites model. Normalized Rb/Na distribution between nepheline
(Xs) and hydrothermal solution (Xaq) (data from Roux, 1974).
12
1
(
K 0 1 − aXs + 1 − aXs ) + bXs
2
K = , (10.15)
2
where
K 0 = mK I + nK II (10.16 )
mK II + nK I
a= (10.17 )
K0
K I K II
b= . (10.18)
K 02
Values of equation 10.15 are plotted in figure 10.2C for various KII /KI ratios
between the limiting cases:
K II
K = K 0 when =1 (10.19 )
KI
and
n K
K = K 0 1 − X s when II = 0 . (10.20)
m KI
Trace Element Geochemistry / 663
As we see, the more KI differs from KII and the more Xs approaches m/n, the more
marked deviation from ideality becomes.
It was shown by Roux (1974) that the two ideal sites model applies perfectly
to partitioning of Rb distribution between nepheline and hydrothermal solutions.
Based on the experimental work of Roux (1974), deviation from ideality in the
normalized Rb/Na distribution between nepheline (Xs ) and hydrothermal solu-
tion (Xaq ) is detectable for Xs values higher than 10⫺2 (figure 10.3A) Distribution
constant K in the Nernst’s law range is K 0 ⫽ 0.82 (Roux, 1974), and the modifica-
tion of K with increasing Xs is well described by equation 10.20, within experi-
mental approximation (figure 10.3B).
It may be noted in figure 10.3B that prolongation of the interpolant encoun-
ters the abscissa axis at Xs ⫽ 13, which is exactly the proportion of the two struc-
tural sites where Rb/Na substitutions take place in nepheline. According to figure
10.3B and equation 10.20, it is plausible to admit that Rb is fixed almost exclu-
sively in the larger of the two structural sites.
Q=
[ (
N N − r +1 )] [ N − 2(r + 1)] L [ N − ( N A )(
−1 r +1)]
(10.21)
NA !
S mixing = k ln Q, ( 10.22 )
N B = N − NA ( 10.23 )
(
ln N ! = N ln N − 1 , ) ( 10.24 )
NA N B − rNA N B − rNA
S mixing = − k NA ln + ln . (10.25)
N r +1 N
n A nB − rn A nB − rn A
Gmixing = RT n A ln + ln . (10.26)
n A + nB r + 1 n A + nB
nB − rn A nB − rn A nB
Gexcess mixing = RT ln − nB ln . (10.28)
r + 1 n A + nB n A + nB
Differentiating equation 10.28 with respect to nA,we obtain the activity coefficient
of component AM in the solid solution (BX ,A1⫺X )M:
∂ G excess mixing −r
ln γ A =
∂ nA = (
ln X B − X A , ) (10.29)
r +1
where
NA nA
XA = = (10.30)
N n A + nB
and
NB nB
XB = = . (10.31)
N nA + nB
distributions calculated with the local lattice distortion model reproduce experi-
mental evidence fairly well. As described by Iiyama (1974), this occurs because
any enthalpic contribution to the Gibbs free energy of mixing in the solid phase
(Hmixing ), arising from solute-solvent interactions, varies linearly with XA in the
Henry’s law concentration field.
Figure 10.4 shows normalized Ba/K and Sr/K distributions between sanidine
and a hydrothermal solution (Iiyama, 1972), Li/K between muscovite and a hy-
drothermal solution (Voltinger, 1970), and Rb/Na between nepheline and a hy-
drothermal solution (Roux, 1971b), interpreted through the local lattice distor-
tion model, by an appropriate choice of the Nernst’s law mass distribution
constant K and the lattice distortion propagation factor r.
Table 10.1 lists r values calculated for normalized distributions of alkalis and
alkaline earths between silicates and hydrothermal solutions and for normalized
distributions of rare earth elements (REE) between pyroxenes and silicate melts.
666 / Methods
Table 10.2 lists upper concentration limits for Henry’s law behavior of various
trace elements in silicate crystals for specific P and T conditions. In some cases,th-
ese limits are only approximate or conjectural. The table is not exhaustive, be-
cause it reflects the state of the art in the early 1980s and is subject to substantial
modifications in response to the advance of experimental knowledge since that
time.
Henry’s law in crystals has low concentration limits (figure 10.1A), ascribable to
modifications in the solid solution process, which, at high dilution, is greatly
affected by the intrinsic and extrinsic defects in the crystal (Wriedt and Darken,
1965; Morlotti and Ottonello, 1982). Figure 10.5 shows some experimental evi-
dence concerning solid solutions of REE in silicates. Deviations from Nernst’s
law observed in figure 10.5 may be ascribed to defect equilibria between REE and
cationic vacancies present in the solid (De Vore, 1955; Wood, 1976c; Harrison
and Wood, 1980; Morlotti and Ottonello, 1982; see also chapter 4).
In logarithmic plots of simple weight concentrations, as in figure 10.5, the
Nernst’s law region appears as a straight line with an angular coefficient of 1. At
the low concentration limits for Henry’s law behavior in the solid, we observe a
sudden change in slope, and, at high dilution, the new distribution has an angular
coefficient lower than 1 (i.e., ␣ ⬍ 45°). Both the slope rupture point and the
angular coefficient valid at high dilution are dictated by the extrinsic defects in
the solid (Morlotti and Ottonello, 1982).
Correct evaluation of solution mechanisms of a given trace element in silicate
Trace Element Geochemistry / 667
Table 10.2 Upper concentration limits for Henry’s law behavior in silicates and oxides.
Abbreviations: Ab ⫽ albite; Sa ⫽ sanidine; Pl ⫽ plagioclase; Fel ⫽ alkaline feldspar; Rb-fel ⫽
Rb-feldspar; Ne ⫽ nepheline; Mu ⫽ muscovite; Ol ⫽ olivine; Di ⫽ diopside; Cpx ⫽ clinopyroxene;
Oxp ⫽ orthyporoxene; Amph ⫽ amphibole; Par ⫽ pargasite; Gr ⫽ garnet; Ilm ⫽ ilmenite;
Arm ⫽ armalcolite (from Ottonello, 1983).
crystals thus requires knowledge of the defects in the phases. As we saw in chapter
4, defect equilibria in crystals are conveniently described by referring to the f O2
condition. For instance, oxygen deficiency may generate in the crystal doubly
ionized Schottky vacancies VO·· or interstitials Mi·· wheras oxygen excess may result
||
in cationic vacancies V M. Intrinsic disorder is also always present under intrinsic
oxygen partial pressure conditions P* O2. Table 10.3 shows the energetically plausi-
ble defect equilibria that may affect the distribution of trivalent REE between
pyrope Mg3Al2Si3O12 and silicate melt (see Kröger, 1964, for exhaustive treatment
of the defect state of crystals; see also chapter 4 for the significance of the adopted
symbols). The proposed equilibria, which refer to the various f O2 conditions and
trace element concentrations of interest, can be adapted to other silicate crystals
with opportune modifications in the “carrier” notations (Mg, in this case). All
the equilibria ruling the REE distribution within the various concentration ranges
are then translated into a bulk distribution equation whose single terms on the
right each apply to a specific concentration range. For instance, constant K4 in
Trace Element Geochemistry / 669
Figure 10.5 Experimentally observed deviations from Nernst’s law behavior in solid/melt
REE distribution. (A) Weight distribution of Sm and Tm between garnet and melt, experi-
mental sources specified in figure. (B) Weight distribution of Sm between clinopyroxene
and silicate melt. (C) Weight distribution of Sm and Tm between plagioclase and silicate
melt. Systems are An55Di45 and Ab60An30Di10 at T ⫽ 1300 °C and P ⫽ 1 bar. From R.
Morlotti and G. Ottonello, Solution of rare earth elements in silicate solid phases: Henry’s
law revisited in light of defect chemistry, Physics and Chemistry of Minerals, 8, 87–97,
figures 2–3, 1982, copyright 䉷 1982 by Springer Verlag. Reprinted with the permission of
Springer-Verlag GmbH & Co. KG.
table 10.5, relative to equilibrium (4) in the same table, describes defect process
involving preexisting intrinsic defects, whereas constant K7 describes the creation
of extrinsic defects that take place when intrinsic defects are saturated (i.e., at
higher concentrations).
Deviations from Nernst’s law in log-distribution diagrams such as figure 10.5
are represented by the simple equation:
X∗ 1 X∗
ln γ A = ln A −
A
, (10.32)
X A tan α XA
where XA is the trace element concentration in the solid, X A * is the lower concen-
tration limit for Nernst’s distribution, and tan ␣ is the slope of non-Nernstian
distribution at high dilution (see table 10.4). Equation 10.32 can be interpreted
in light of the defect equilibria listed in table 10.3, assuming that slope rupture
670 / Methods
Table 10.3 Defect equilibria affecting REE solubility in pyrope and their effects in bulk garnet/melt
REE distribution (Morlotti and Ottonello, 1982)
Defect equilibrium: ⫻
MgMg ⇔ Mg··
i ⫹ VMg
||
(1)
[REE·Mg] ⱕ [VMg]*
||
KD ⫽ K4KF1/2 ⫺ K4 [REE·Mg] (6)
3 ⫻ 3 1
[REE·Mg] ⱖ [V||Mg]* REE(l3⫹) ⫹ MgMg ⇔ Mg(l2⫹) ⫹ REE·Mg ⫹ V||Mg (7)
2 2 2
KD ⫽ K7[REE·Mg]⫺1/2 (8)
B: Oxygen fugacity higher than the intrinsic oxygen partial pressure: ƒO2 ⬎ P*O2
1 ⫻
Defect equilibrium: O2 ⇔ OO ⫹ V||Mg ⫹ 2h· (14)
2 (gas)
⫺1/2
Equilibrium constant: K14 ⫹ 4[V||Mg]3 ƒO 2
(15)
⫺1/3
Defect concentration: [VMg] ⫽ [VMg]* ⫹ 4
|| || 1/3 1/6
K14 ƒO2 (16)
Bulk distribution:
(17)
C: Oxygen fugacity lower than the intrinsic oxygen partial pressure: ƒO2 ⬍ P*O2
1 ⫺
Defect equilibrium: OO⫻ ⇔ O2(gas) ⫹ V ··
O ⫹ ⫹2e (18)
2
Equilibrium constant: K18 ⫽ 4[V ·· 3 1/2
O] ƒO2 (19)
⫺1/3 1/3 ⫺1/6
Defect concentration: [VMg] ⫽ [V Mg]* ⫺ 4
|| ||
K18 ƒO2 (20)
Bulk distribution:
(21)
Table 10.4 Least-squares fits of type ln Xsolid ⫽ a ln Xliquid ⫹ b relating REE concentrations in silicate crystals and melts, as derived from existing
experimental data. Corresponding values of R2 correlation parameter are also listed (from Morlotti and Ottonello, 1982).
points X A
* represent saturation of intrinsic and extrinsic defects. In the case of
samarium, for instance,
2 ∗
∗ =
X Sm
3
[ ]
V Mg (10.33)
and
2 ∗
X Sm = V Mg
3
[ ] [ ]
− V Mg
.
(10.34)
1 − tan α 3X Sm
ln γ Sm = ln . (10.35)
tan α ∗
[ ]
2 V Mg
2w − wAA − w BB
W = N 0 AB . (10.36)
2
Figure 10.6 Onuma diagrams for crystal/melt trace element distributions. Ionic radii of
Whittaker and Muntus (1970). (A) Augite/matrix distribution, data of Onuma et al.
(1968). (B) Plagioclase/matrix distribution, data of Dudas et al. (1971), Ewart and Taylor
(1969), and Philpotts and Schnetzler (1970). Reprinted from B. Jensen, Geochimica et Cos-
mochimica Acta, 37, 2227–2242, copyright 䉷 1973, with kind permission from Elsevier
Science Ltd., The Boulevard, Langford Lane, Kidlington 0X5 1GB, UK.
carrier substitutions in the crystal of interest (Onuma et al., 1968; Jensen, 1973).
Figure 10.6A, for instance, shows the mass distribution of various trace elements
between augitic pyroxene and the silicate matrix (assumed to represent the silicate
melt), based on the data of Onuma et al. (1968). There are two maxima, at 0.79
and 1.01 Å, interpreted respectively as “optimum dimensions” of sites M1 and
M2 in the crystal (Jensen, 1973). The peaks are more clearly defined for divalent
ions, and the partition coefficients decrease more rapidly for elements with higher
formal charges, departing from the optimum dimension. Figure 10.3B shows an
Onuma diagram focused on the mass distribution of alkaline earths between pla-
gioclase and the matrix. The distribution curves indicate the marked composi-
tional control exerted by the solid: because the carrier is Ca, the mass distribution
coefficients progressively increase with the amount of anorthite component in
solid solution, whereas fractionation is not particularly affected. The main advan-
tage of Onuma diagrams is that they can elucidate the marked fractionation in-
duced by the crystal structure in crystal/melt equilibria. However, their interpreta-
674 / Methods
tion is merely qualitative and does not currently allow translation of the observed
phenomena into solution models for the solid phases of interest.
As we will see in section 10.9, trace element geochemistry has important applica-
tions in the investigation of magmatic processes. However, in partial melting and
partial crystallization models based on trace elements, the role ascribed to the
liquid phase is often subordinate to that of the solids undergoing melting or crys-
tallization. In such models, in fact, the fractionation properties of solids are
clearly defined, but the silicate melt is often regarded as a sort of reservoir that
does not exercise any structural control on trace partitioning, where trace ele-
ments are merely discharged or exhausted, depending on their relative affinities
with the solids. This sort of reasoning is fundamentally wrong and may result in
prejudice and vicious circles.
mass distribution of manganese between olivine and silicate liquid: with decreas-
ing melt acidity (expressed here as the Si/O atomic ratio), conventional mass
distribution coefficient K⬘ decreases for increased solubility of the element in the
liquid phase, as a result of the increased relative extension of the anion matrix
(see section 6.1.3).
We may now argue whether or not a silicate melt can fractionate elements of the
same group. The question is rather important for REE, whose relative fraction-
676 / Methods
and
whose Gibbs free energy change is related to partial equilibria 10.37 and 10.38 by
∆G 39 =
1
2
(
∆G 38 − ∆G 37 . ) (10.40)
a
REEO−
∆G39 = − RT ln
2( l ) .
a 3+ a 2 2 − (10.41)
REE( l ) O( l )
Writing the bulk molar amount of trivalent REE in melt n in terms of partial
amounts of polyanions (n1 ) and cations (n2 ):
2 aO2 − ⋅ K 39
γ REE =
(l )
. (10.43)
1 + aO2 − ⋅ K 39
(l )
Trace Element Geochemistry / 677
Z
∆G 39 = a + b 2 , (10.44)
AREE
where a and b are constants. This amount of energy (hence, the various dispropor-
tionation constants K39 ) can be evaluated from experimental observations on liq-
uid/liquid REE partitioning (see Ottonello, 1983, for details of calculations).
Table 10.5 shows a comparison between measured (m: Watson, 1976; Ryerson
and Hess, 1978, 1980) and calculated (c: Ottonello, 1983) mass distributions on
the basis of the amphoteric solution model. The K39 values obtained from the
model are as follows: La ⫽ 7.5, Ce ⫽ 9, Pr ⫽ 11, Nd ⫽ 14, Sm ⫽ 21, Eu ⫽ 27,
Gd ⫽ 33, Tb ⫽ 40, Dy ⫽ 50, Ho ⫽ 62, Er ⫽ 76, Tm ⫽ 95, Yb ⫽ 120, and Lu ⫽
150. Applications of these K39 values to equation 10.43 leads to the mean activity
coefficient trends observed in figure 10.8.
The plot in figure 10.8 clearly shows that silicate melts may exert marked
fractionation on REE: acidic melts with relatively greater free oxygen activity
display the highest mean activity coefficients in light REE (LREE). LREE are
thus less soluble in these melts than heavy REE (HREE). The opposite is ob-
served in basic melts, which have the highest mean activity coefficients at the
HREE level. In other words, REE relative concentrations in silicate melts simply
reflect their relative solubilities, as we saw for crystal structures.
The Eu2⫹ /Eu3⫹ ratio in silicate melts is controlled by the redox equilibrium
Morris and Haskin (1974) observed an increase in the Eu2⫹ /Eu3⫹ ratio with in-
creasing (Al⫹Si)/O and high field-strength ions in the melt. This effect may be
ascribed to the different acid-base properties of Eu2O3 and EuO (Fraser,
1975a,b). Equilibrium 10.45 may in fact be rewritten in the form
1
Eu2O3( l ) ⇔ 2 EuO ( l ) + O 2 (gas) . (10.46)
2
The activities of the two oxides in the liquid are determined by the type of
dissociation. For EuO, basic dissociation is plausible:
Table 10.5 Measured (m) and calculated (c) basic liquid/acidic liquid REE mass distribution constants (Ottonello, 1983).
La Sm Dy Yb Lu
Reference(s) m c m c m c m c m c
Watson (1976) 3.70–4.31 3.68 4.19–5.15 4.15 – – – – 4.18–7.13 5.69
Ryerson and Hess (1978) 3.08–6.90 6.86 – – 2.20–4.64 10.9 3.03–4.18 18.2 – –
(system a)
Ryerson and Hess (1978, 1980)
(system b) 13.8 14.5 – – 16.57 27.37 – – – –
Trace Element Geochemistry / 679
Figure 10.8 Mean activity coefficients of various REE in silicate melts with different
values of activity of dissolved free oxygen aO2⫺, plotted as a function of increasing atomic
number. Reprinted from Ottonello (1983), with kind permission of Theophrastus Publish-
ing and Proprietary Co.
EuO ⇔ Eu 2 + + O 2 − . (10.47 )
Eu 3+
=
a O1 4
2(gas)
K a1 2 ⋅ a 3 22− ⋅
( )
1 − X MO
( )
(10.48)
Eu 2 + c O(l )
X MO + 0.25 1 − X MO
(Wood and Fraser, 1976), where c is a constant and XMO is related to (Al⫹Si)/O
in the systems investigated by Morris and Haskin (1974) through (Fraser, 1975b):
O O
X MO = − 2 − 2 + 1 . (10.49)
Al + Si Al + Si
It must be noted that the increased stabilization of the oxidized form with the
basicity of the melt, experimentally observed by Morris and Haskin (1974) for
680 / Methods
There are few data on this point. Most information comes from the experimental
work of Watson (1976) on trace partitioning between immiscible liquids in the
K2O-Al2O3-FeO-SiO2 system at 1180 °C and 1 bar. Table 10.6 shows some con-
centration limits based on this experimental work. At present, it is impossible to
establish whether or not deviations from Henry’s law take place in the acidic or
basic liquids at equilibrium.
We have already seen that, within the range of Nernst’s law, the solid/liquid parti-
tion coefficient differs from the thermodynamic constant by the ratio of the Hen-
ry’s law activity coefficients in the two phases—i.e.,
γ h,i ,l
K i′ = Ki . (10.50)
γ h,i , s
Trace Element Geochemistry / 681
D=
( Tr Cr) s
.
( Tr Cr) l
(10.51)
The formulation of this coefficient derives from the consideration that solid/liquid
trace element distribution can be ascribed to the existence of simple exchange
equilibria of the type
where M is the common ligand. For instance, an exchange reaction of this type
may be responsible for the isomorphic substitution of Sr in plagioclase:
CaAl 2 Si 2 O 8 + Sr 2 + ⇔ SrAl 2 Si 2 O 8 + Ca 2 + .
solid melt solid melt
(10.53)
In this case, trace element and carrier occupy the same structural position both
in the solid phase and in the melt and are subject to the same compositional
effects in both phases (i.e., extension of the cation matrix in the melt and amount
of anorthite component in the solid). Figure 10.9A shows the effect of normaliza-
tion: the conventional partition coefficient of Sr between plagioclase and liquid
varies by about one order of magnitude under equal P-T conditions, with increas-
ing anorthite component in solid solution, whereas normalized distribution co-
efficient D is virtually unaffected. Figure 10.9B shows the same effect for the
Ba-Ca couple.
It is important to note that, in the example given above, trace element (Sr or
Ba) and carrier (Ca) have the same formal charge. In the case of substitution of
682 / Methods
Figure 10.9 Effect of normalization on trace element distribution of Sr (A) and Ba (B)
between plagioclase and silicate liquid. Distribution curves are based on various experi-
mental evidence (Ewart et al., 1968; Ewart and Taylor, 1969; Berlin and Henderson, 1968;
Carmichael and McDonald, 1961; Philpotts and Schnetzler, 1970; Drake and Weill, 1975).
Ab: albite; Or: orthoclase; An: anorthite.
elements with different valences, the exchange equilibrium is more complex and
the adoption of normalized partition coefficients does not eliminate the composi-
tional effects. For instance, in the case of stabilization of a rare earth in diopside,
we can write
Figure 10.10 Arrhenius plot of conventional partition coefficient (A) and thermody-
namic constant (B) for clinopyroxene/liquid distribution of Ni.
tional weight concentration ratio, according to the data of Lindstrom and Weill
(1978) and Häkli and Wright (1967): the Arrhenius plot defines two distinct
arrays, with different slopes and intercepts.
If we now consider solid/liquid Ni distribution from a thermodynamic point
of view, as suggested by Lindstrom and Weill (1978):
K 55 =
[CaNiSi O ] 2 6
, (10.56)
[NiO ][CaO ][SiO ]
2
2
[CaNiSi O ] = X
2 6 Ni,M 1 (10.57)
[SiO ] = X
2 SiO 2 , l (10.58)
(Lindstrom and Weill, 1978), where XNi,M1 is the atomic fraction of Ni in the M1
site of pyroxene, and the other X terms represent the molar fractions of oxide in
the melt. As shown in figure 10.10B, the plot of thermodynamic constant K55
obtained in this way defines a single Arrhenius trend. Evidently, in this case, the
double conversion implicit in the adoption of conventional partition coefficients
introduces an apparent complexity that is indeed ascribable to the different com-
positions of the investigated systems. These effects must be carefully evaluated
when studying trace element distributions in petrogenetic systems.
Recent appraisals of the state of art in experimental trace partitioning studies
relevant to igneous petrogenesis may be found in the review articles of Green
(1994) and Jones (1995).
µ TrM,s = µ TrM,
0
s + RT ln a TrM,s (10.61)
µ CrM, s = µ CrM,
0
s + RT ln a CrM, s (10.62)
µ TrM,l = µ TrM,
0
l + RT ln a TrM,l (10.63)
µ CrM,l = µ CrM,
0
l + RT ln a CrM,l . (10.64)
and
a TrM,l µ TrM,
0
s − µ TrM,l
0
= exp (10.67)
a TrM, s RT
and
a CrM,l µ CrM,
0
s − µ CrM,l
0
= exp . (10.68)
a CrM, s RT
Substituting molar fractions X and molalities m for activities in the solid and
in the liquid phase, respectively, we obtain
and
As we have already seen, within Henry’s law range the activity coefficient of the
trace component in the solid can be expressed as
∆µ ∗
γ TrM, s = exp
RT
, (10.71)
where ⌬* is the molar Gibbs free energy amount involved in solute-solvent inter-
actions. The (molar) partition coefficient thus has the form
∆µ ∗ ∆µ TrM
0 −1
K ′TrM = exp
RT ⋅ exp ⋅ γ TrM,l . (10.72)
RT
∆µ ∗ ∆µ TrM
0 ∆µ CrM
0 γ CrM,l
D s l = exp ⋅ exp ⋅ exp − ⋅ .
RT γ TrM,l
RT (10.73)
RT
Because
686 / Methods
∆µ TrM
0
= ∆ hTrM
0
− T∆ s TrM
0
+ P∆v TrM
0
(10.75)
∆µ CrM
0
= ∆ h CrM
0
− T∆s CrM
0
+ P∆v CrM
0
, (10.76 )
where terms in h, s, and v are, respectively, partial molar enthalpies, partial molar
entropies, and partial molar volumes (cf. section 2.5), substituting equations
10.74 to 10.76 in equations 10.72 and 10.73 and deriving the logarithmic forms
in dT, we obtain
∂ ln K ′TrM
∂T
P, X
=
1 0
RT 2
( ) ∗
hTrM, s − hTrM,l − ∆ hTrM
(10.77)
∂ ln D s l
∂T
P, X
= (
1 0
RT 2
) (
hTrM, s − hTrM,l − hCrM,
0
) ∗
s − hCrM,l − ∆ h TrM .
(10.78)
The terms in parentheses in equations 10.77 and 10.78 represent the enthalpies
of solution of TrM and CrM, respectively, in the melt at the T of interest (cf.
section 6.4). If the enthalpy of solution does not vary appreciably with T, because
the interaction term ⌬h* TrM cannot also be expected to vary appreciably with T,
integration of equation 10.77 leads to a linear form of the type
ln K ′TrM = A + BT −1 , (10.79)
∂ ln K ′TrM
∂P
T,X
=
1 0
RT
( ) ∗
v TrM, s − v TrM,l − ∆v TrM
(10.80)
Trace Element Geochemistry / 687
and
∂ ln D s l
∂P
T,X
=
RT
(
1 0
) (
v TrM, s − v TrM,l − vCrM,
0
) ∗
s − v CrM,l − ∆v TrM .(10.81)
The trace component in solid solution usually has partial molar volumes not
far from those of the carrier, for crystal-chemical reasons, and the terms in paren-
theses in equation 10.81 compensate each other. The term ⌬v* TrM, expressing the
differences in partial molar volume arising from solute-solvent interactions in the
solid, is also usually negligible. As a result, the effect of P on normalized partition
coefficients is usually very restricted. The same cannot be said for conventional
partition coefficient K⬘TrM, because the partial molar volumes of the trace compo-
nent in the solid and liquid phases (and hence their differences) may vary appreci-
ably with P (McIntire, 1963).
Figure 10.12 Effects of oxygen fugacity on oxidation state of Ti and V in Na2Si2O5 melt
at T ⫽ 1085 °C (experimental data from Johnston, 1964, 1965, and Johnston and
Chelko, 1966).
experiments of Johnston (1964, 1965) and Johnston and Chelko (1966). These
effects, which must not be confused with the previously described compositional
effect, markedly influence solid/liquid trace element distribution, because crystal
structures dramatically discriminate elements in various oxidation states. Figure
10.13 shows, for example, the effect of f O2 on the conventional partition coeffi-
cient of Cr between olivine and melt (figure 10.13A) and between subcalcic py-
roxene and melt (figure 10.13B). In both cases, the effect of f O2 is nonlinear and
varies with T in a rather complex way.
Figure 10.13 Effect of oxygen fugacity on conventional partition coefficient of Cr. (A)
Olivine/liquid partitioning; experimental data of Bird (1971), Weill and McKay (1975),
Huebner et al. (1976), Lindstrom (1976), and McKay and Weill (1976). (B) Subcalcic py-
roxene/liquid partitioning; experimental data of Schreiber (1976). Reprinted from A.J. Ir-
ving, Geochimica et Cosmochimica Acta, 42, 743–770, copyright 䉷 1978, with kind permis-
sion from Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington 0X5 1GB,
UK.
In the first case, the diffusion rates of the trace component in all phases of the
system must be sufficiently high with respect to rates of crystal accretion or melt-
ing. Figure 10.14 shows experimental data on elemental diffusivities of alkalis in
a rhyolitic melt. T conditions being equal, the diffusivities of Na and K in the
melt are four to six orders of magnitude greater in the liquid than in feldspars
(compare figures 10.14 and 4.8). We can thus expect the formation of chemically
zoned crystals, whereas possible chemical zoning in the liquid may more reason-
ably be ascribed to advection.
The partition coefficient of trace element i between a given solid phase s and
690 / Methods
liquid l in a closed system is valid on each portion of the two phases at equi-
librium:
X i,s
K′ = . (10.82)
X i ,l
In the case of interface equilibrium (open system conditions), the partition co-
efficient is valid only at the interface between solid and liquid (or at zero distance
from the interface) and at time of crystallization (or melting) t:
X i , s, 0,t
K′ = . (10.83)
X i ,l , 0,t
t
K′
t ∫ X i ,l , 0,t dt
KA = L
0
(10.84)
1
L ∫ X i ,l , x,t dx
0
Trace Element Geochemistry / 691
(Albarede and Bottinga, 1972), where t is time since the beginning of the process
(at constant accretion or dissolution rates), Xi,l,x,t is the trace element concentra-
tion measured at distance x from the interface and at time t, and L (in the case of
fractional crystallization) is half the mean distance between centers of accreting
crystals less half the mean thickness of the crystals. We will see later the effects
of erroneous evaluation of KA and K⬘ when modeling fractional crystallization
processes.
[ ]
K ′ −1
Cl = C 0 1 − F , (10.85)
dm M − m
= K′ , (10.86)
ds Q−S
where Q is the initial mass of the system, M is the mass of trace elements in the system,
S is the mass of the solid, and m is the mass of trace element in the solid. According to
Neumann et al. (1954), the integrated form of equation 10.86 is
S
K′
m = M 1 − 1 − (10.87)
Q
(when S ⫽ 0, m ⫽ 0).
Differentiating equation 10.87 in ds yields
K ′ −1
dm M S
= K′ 1 − . (10.88)
ds Q Q
[ ]
K ′ −1
Cs = C0K′ 1 − F ( 10.89 )
[ ]
K ′ −1
Cl = C0 1 − F . (10.90)
It must be noted that equation 10.89 is valid only at the solid/liquid interface, whereas
equation 19.90 is valid over the entire crystallization range.
[ ]
K′
1− 1− F
Cs = C0 . (10.91)
F
Clearly, equation 10.91 can be derived directly from equation 10.87 by substitut-
ing concentrations for masses.
Equation 10.91 also can be expressed as
[ ]
K ′ −1
C s = K AC 0 1 − F (10.92)
(the identity between eq. 10.92 and eq. 10.89 at the interface where KA ⫽ K⬘
is obvious).
The ratio between equations 10.91 and 10.90 gives the relationship between
apparent partition coefficient KA and conventional partition coefficient at inter-
face K⬘ (i.e., the integrated form of eq. 10.84):
[ ]
K′
C 1− 1− F
KA = s = . (10.93)
[ ]
K ′ −1
Cl F 1− F
Figure 10.15 Fractional differences (KA ⫺ K⬘)/K⬘ plotted as functions of degree of frac-
tional crystallization F for various values of K⬘. Reprinted from F. Albarede and Y. Bot-
tinga, Geochimica et Cosmochimica Acta, 36, 141–156, copyright 䉷 1972, with kind per-
mission from Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington 0X5
1GB, UK.
KA
K ′C 0
Cs =
[1 − (1 − K ′ )F ] (10.94.1)
and
C0
Cl = ,
[1 − (1 − K ′ )F ] (10.94.2)
where K⬘ is valid in all portions of the system (and not only at the interface, as
in the preceding case).
dm M − m + C PS
= K′ Tr
,
ds (
Q − 1 − P S ) (10.95)
where P is the mass ratio between resorbed and precipitated phases and C Tr is the
mean concentration of the trace element in the resorbed phase (see eq. 10.86 for
the remaining symbols). Neumann et al. (1954) gave the following distribution
equations, which are valid at the interface:
Cs
Cl = . (10.97)
K′
Trace Element Geochemistry / 695
Xα Xβ Xγ
Ds l = + + + L. (10.98)
K l′ α K l′ β K l′ γ
In equation 10.98, X␣, X, and X␥ are the relative mass fractions of crystallizing
phases (X␣ ⫹ X ⫹X␥ ⫹ . . . ⫽ 1), and K⬘l/ ␣, K⬘l/ , and K⬘l/ ␥ are their respective
conventional partition coefficients.
Ds/l can be substituted for K⬘ in equations 10.89 and 10.90 and equations
10.94.1 and 10.94.2.
K ′C 0
C acc = × X acc +
C0
(
× 1 − X acc )
[1 − (1 − K ′)F ] [1 − (1 − K ′ )F ] (10.99)
(
C acc = C 0 K A 1 − F ) ( ) ( )
KA −1
× X acc + C 0 1 − F
KA −1
× 1 − X acc . (10.100)
process. Note also, however, that the field of possible trace element concentrations
is quite wide (dashed area in figure 10.16), depending on the efficiency of filter-
pressing.
1. Modal melting. Solid phases melt in the same modal proportion in which
they occur in the initial solid, so that the modal composition of the resid-
ual solid remains unchanged throughout the process.
2. Nonmodal melting. The melting proportions of solid phases differ from
the modal proportions in the initial solid, so that the modal composition
of the residuum is progressively modified throughout the process.
C0
Cl =
[D 0 (
+ F 1 − D0 )] (10.101)
and
C 0D 0
Cs = ,
[D 0 (
+ F 1 − D0 )] (10.102)
where F is the mass fraction of melt in the system (not the fraction of solid, as in
the preceding cases) and D 0 is the initial solid/liquid distribution at the beginning
of melting:
X 0,α X 0, β X 0,γ
D0 = + + + L. (10.103)
K l′ α K l′, β K l′ γ
D 0 − PF
Ds l = , (10.104)
1− F
where
Pα Pβ Pγ
P= + + + L. (10.105)
K l′ α K l′ β K l′ γ
C0
Cl =
[D 0 + F 1− P ( )] (10.106)
and
C 0D s l
Cs = .
[D 0 + F 1 − D0( )] (10.107)
Because P␣,,␥ and K⬘l / ␣,,␥ represent final state conditions, equations 10.106
and 10.107 can be applied only to the entire melting range whenever melting
proportions and conventional mass distribution coefficients remain constant
698 / Methods
throughout the process. In all other cases, the equations above must be associated
with functions describing the evolution of P␣,,␥ and K⬘l/ ␣,␥ with the degree of
partial melting F.
n
C s,i = C 0 ∏
(
1 − Di + i − 1 ∆ F)
(10.108)
i =1 1− F
Di −1 − Pi −1 ∆ F
Di = (10.109)
1 − ∆F
C l , i = D i C s , i −1 (10.110)
n
1
C l ,i = ∑ C l ,i
i i =1
(10.111)
P = f (F ) (10.112)
K l′ α , β ,γ = f ( F ), (10.113)
where Cs,i is trace element concentration in the residual solid after removal of the
ith increment of liquid, Cl,i is trace element concentration in the ith increment of
liquid, C l,i is mean trace element concentration in i accumulated increments, and
⌬F is the fractional mass increment of melting.
Many natural processes are closer to fractional melting than to equilibrium
melting conditions. For instance, the leucosome bands in gneisses may be assimi-
lated to fractional melting liquids, being progressively removed from the system
as soon as they are produced. In such cases, the melting proportions of the solid
phases may change dramatically as the process advances (cf. Presnall, 1969) and
must be carefully evaluated for rigorous modeling of trace element distributions.
For this purpose, it is opportune to compare the natural system to a synthetic
analog, the melting behavior of which is known with sufficient precision. As we
Trace Element Geochemistry / 699
Figure 10.17 Evolution of liquid and solid during fractional melting of initial com-
position C in Di-Ab-An system. Curves ILC(1) and ILC(2): fractional melting liquids;
TSC: residual solids; dotted line: fractionally accumulated liquids at quoted degrees of
partial melting (from Ottonello and Ranieri, 1977).
Pα = PAb =
[13 − (9 + 40 F )
12
]
20
P = f ( F ) ≡ Pβ = PAn =
(1 − Pα ) (10.114)
2.17
Pγ = PDi = 1.17Pβ .
700 / Methods
The equations above were determined graphically using a few points on the ILC
curves of figure 10.17 and allow us to follow trace element distribution during fractional
melting of initial system C (Di15An51Ab34 ). The heavy line (TSC) in figure 10.17 describes
the evolution of the residual liquid during melting; the dotted curve represents the modal
evolution of fractionally accumulated liquids calculated by Morse (1976) with application
of the lever rule (see chapter 7). This last curve can be obtained directly by summing the
compositions of instantaneous liquids obtained through the two systems of equations
10.114 and 10.115.
1 1 X
Cl = C0 −
Ds l Ds l
− 1 ⋅ exp − Ds
L
( l ) . (10.116)
In equation 10.116, L is the length of the molten zone and X is the length of
relative displacement. The application of zone melting to geological systems must
be carried out with caution, because the modal proportions in the original solid,
in the melt, and in the precipitating solid all generally differ markedly. Moreover,
initial concentration C 0 cannot be expected to remain constant over long X tra-
jectories. Nevertheless, zone melting probably operates in the earth’s upper man-
tle: it has been suggested that the existence of the low-velocity layer is a result of
such a process. Radiogenic elements such as U, Th, and K may be efficaciously
extracted by zone melting from convecting lithospheric portions and contribute
feedback to the process by heat release associated with radioactive decay.
Trace Element Geochemistry / 701
During magma genesis and differentiation, fluid phases may develop. The fluids
may exolve from the melt for two main reasons: progressive increase of concentra-
tion of gaseous species that reach the solubility product of the fluid in question,
or decreased solubility in the melt during ascent to the surface (adiabatic decom-
pression). If we call the mass fractions of solid, melt, and fluid in the system Fs ,
Fl , and Ff , respectively, the mass balance equation relating initial trace element
concentration C 0 to the concentrations in solid, melt, and fluid is
C 0 = C l Fl + C s F s + C f F f . (10.117)
C0
Cl =
[Fl ( )
+ D s l 1 − Fl − F f + D f l F f ] (10.118)
C 0D s l
Cs =
[Fl ( )
+ D s l 1 − Fl − F f + D f l F f ] (10.119)
Cf = D f l Cl . (10.120 )
(1 Ps l −1 )
C0 Fl Ps f
Cf = 1 − (10.121)
D0, s f + v D0, s f + v
Cs =
(
C f D0, s f − Ps f F f ) (10.122)
1 − Fl
Cf
Cl = (10.123)
Df l
1 Ps l
C0 Fl Ps f
,
Cl = 1− 1−
Fl D0, s f + v (10.124)
702 / Methods
where is the constant fractional amount of fluid with respect to the initial mass
of the system.
If the fluid is present during fractional crystallization, the trace element con-
centration in the liquid is dictated by
C l = C 0 Fl
( Ds l + D f l G −1) (10.125)
Some trace elements have a marked tendency to partition into the melt during
magma genesis and differentiation. Elements showing the highest affinity for the
melt (i.e., Th, Ta, Zr, Hf, La) were defined by Treuil (Treuil, 1973; Treuil and
Varet, 1973; Treuil and Joron, 1975; Treuil et al., 1979) as “hygromagmatophile,”
and those preferentially partitioned into the melt (although not so markedly as
the hygromagmatophile ones) were defined as “incompatible.”
M x + + nL y − ⇔ MLzn − , ( 10.126 )
whose constant is . If we admit that solid/melt trace partitioning only involves cationic
species Mx⫹ (in agreement with the concept of “noncrystallizable groupings” of Ubbel-
hode, 1965, and Boon, 1971):
K =
[M ] x+
s
,
[M ] x+
l
(10.127)
* The sentence “the gas phase is not considered as an additional phase” adopted by Allegre et
al. (1977) for this purpose is rather cryptic. They probably meant that the fractionation laws valid
for solid phases are also valid for the gaseous phase, regardless of the different state of aggregation.
In our view, the gas phase is a real phase to all intents and purposes (i.e., a region of the system with
peculiar and distinguishable chemical and physical properties; cf. section 2.1). Hence, the term G is
operationally analogous (although not identical) to v in equations 10.121 and 10.124.
Trace Element Geochemistry / 703
K′s l =
[M ] γ
x+ −1
x+
s M ,s
.
[M x+
] γ + [ML ] γ
−1
x+
l M ,l
z− −1
z−
l ML , l
(10.128)
K
K ′s l = .
γ M x+ , s γ M x+ , l γ n
L y− ,l
( ) (10.129)
n
+ β Lyl −
γ M x +, l γ MLz −, l
Assuming ideal solution behavior by Mx⫹, Ly⫺, and MLz⫺ in the melt, equation 10.129
reduces to
K
K ′s = .
( )
l n (10.130)
1 + β Lyl −
Based on equation 10.130 (figure 10.18), Treuil et al. (1979) suggested that hygromagmato-
phile behavior is restricted to the field (L ly⫺ )n ⬎⬎ 1, where K⬘s /l ⬍⬍ K, whereas for values
of (L ly⫺ )n near 1 the highest variability of K⬘s /l occurs, corresponding to the field of incom-
patible elements.
Figure 10.18 shows that coordination number n of the complex (which depends on
704 / Methods
the relative charges of cation and ligand) has a strong influence on K⬘s/l. The higher the
value of n (i.e., the higher the charge of the cation for similar ligands: for example, Zr4⫹,
Hf 4⫹, Th4⫹ ), the wider the field of hygromagmatophile behavior. Moreover, if the anionic
ligand is a complex of type CO32⫺, F⫺, Cl ⫺, or OH⫺, its concentration in the melt in
saturation conditions depends on the solubility product of the gaseous phase. Modifica-
tions in fluid partial pressure may thus result in sudden changes in K⬘s/l during magma
differentiation (Treuil et al., 1979).
C l = C 0F ( ),
Ds l −1
(10.131)
Cl
≈ F −1 (10.132)
C0
C i .l C
= i .0 = constant. (10.133)
C j ,l C j,0
1
∆ = − F ( s l −1 ) F
D
(10.134)
F
[
C l = C 0 1 − F 1 Ds l F −1 ,] (10.135)
where F is the fraction of residual solid, in fact reduces to equation 10.132 when-
ever Ds/l is negligible. In the case of fractional melting, errors associated with
Trace Element Geochemistry / 705
Figure 10.19 Fractional error associated with equation 10.132. (A) Rayleigh’s crystalliza-
tion and fractional melting. (B) Equilibrium melting.
equation 10.132 are symmetric with respect to errors associated with fractional
crystallization:
1
∆ = − F ( s l ) F
l D
(10.136)
F
1 1
∆= − F (10.137)
F D sl + F + FD s l
are not negligible for any reasonable value of F (figure 10.19B) and for two hygro-
magmatophile elements i and j the relative abundances will not generally obey
equation 10.133. This is especially true for low melting conditions in which Ds/l
is comparable in magnitude to F.
The differentiation index defined by equation 10.132 is a formidable tool in
the petrologic investigation of volcanic suites. Because types and modal propor-
tions of crystallizing phases change discontinuously during fractional crystalliza-
tion (cf. Presnall, 1969), plotting the various chemical parameters as a function of
F (defined through eq. 10.132) generally gives smooth trends with discontinuities
located at particular F values corresponding to the appearance or exhaustion of
a given solid. This was exemplified quite convincingly by Barberi et al. (1975) in
the Boina magmatic series (Afar, Ethiopia). Figure 10.20 shows major element
706 / Methods
Figure 10.20 Major element concentrations in effusive products of Boina series plotted
as functions of degree of fractional crystallization, based on equation 10.132. I ⫽ first
discontinuity (transition from olivine-dominated to plagioclase-dominated fractionation);
II ⫽ second discontinuity (appearance of Fe-Ti oxides); III ⫽ third discontinuity (field
of SiO2 oversaturated trachytes: apatite and Mn-oxides precipitate); IV ⫽ beginning of
peralkalinity field. From Barberi et al. (1975), Journal of Petrology, 16, 22–56. Reproduced
with modifications by permission of Oxford University Press.
Isotope Geochemistry
Atomic nuclei are composed mainly of protons, with a mass of 1.67262 ⫻ 10⫺24
g, or 1.007276 atomic mass units (amu, relative to 12C ⫽ 12.000000), and neutrons,
with a mass of 1.008665 amu (see appendix 1 for a complete energy list of nu-
clear properties).
The number of protons Z in the nucleus establishes the identity of a chemical
element. The mass number A of a given nucleus is established by the number of
protons Z and neutrons N, added together. Neutrons are “neutral” in charge, and
their relative numbers with respect to protons are not generally fixed. Generally,
stable elements with even numbers of protons tolerate greater variability in terms
of numbers of neutrons per nucleus than elements with odd numbers of protons.
Ag atoms, for instance, are stable only when they have 47 protons and 60 or 62
neutrons. Each possible combination of nuclear charge Z and neutron number N
is referred to as a nuclide. Nuclides of the same nuclear charge (same element)
with the same number of protons Z and a different numbers of neutrons N are
defined as isotopes of that element (iso ⫽ “same,” topos ⫽ “place”—i.e., they
occupy the same position on the periodic chart of the elements). Nuclides of the
same mass number A but with different atomic numbers Z are called isobars, and
nuclides with the same number of neutrons N but different masses A are isotones.
The difference N ⫺ Z (corresponding to A ⫺ 2Z) is referred to as an isotopic
number. Finally, isomers are two or more nuclides having the same mass number
A and atomic number Z but existing for measurable times in different quantum
states, with different energies and radioactive properties.
The term “isotope” was coined by Soddy (1914) to define two or more substances of
different masses occupying the same position in the periodic chart of the elements. Soddy’s
hypothesis was adopted to explain apparent anomalies in the relative positions of three
couples of elements (Ar-K, Co-Ni, and Te-I) in the periodic chart. For instance, pot-
assium is present in nature with three isotopes with masses of 39, 40, and 41, respectively,
in the following proportions: 39K ⫽ 93.26, 40 K ⫽ 0.01, and 41K ⫽ 6.73. Because the pro-
portion of 39K is dominant, the element K occurs in nature with mass 39.102, near the
/ 707
708 / Methods
mass of isotope 39K. Ar also has three isotopes: 36Ar (0.34%), 38Ar (0.06%), and 40 Ar
(99.60%) and, obviously, the atomic weight of element Ar (39.948) is near that of isotope
40
Ar, which is the most abundant. Argon, which has an atomic number lower than that of
potassium, has a higher mass, and this explains the (apparent) anomaly of their respective
positions on Mendeleev’s periodic chart.
Figure 11.1 is a chart of nuclides with N as the ordinate and Z as the abscissa.
In this representation, isotones appear along horizontal lines and isotopes along
the same vertical line. The opposite sort of representation is known as a “Segré
chart.”
All stable nuclides fall above the N ⫽ Z line, with the exception of 1H and
3
He. Up to 157 N, the number of neutrons does not exceed the number of protons
by more than 1. Above 15 7 N, the stable nuclides diverge from the N/Z ⫽ 1 line
Hydrogen atoms and part of 4He are believed to have been created during the “Big
Bang” by proton-electron combinations. Most nuclides lighter than iron were created by
nuclear fusion reactions in stellar interiors (cf. table 11.1). Nuclides heavier than the Fe-
group elements (V, Cr, Mn, Fe, Co, Ni) were formed by neutron capture on Fe-group seed
nuclei. Two types of neutron capture are possible: slow (s-process) and rapid (r-process).
Figure 11.2 shows two enlarged portions of the Segré chart where s-process and
r-process are effective. Every box in the figure represents a possible nuclide, but only the
white boxes are occupied by stable nuclides. The diagonally ruled boxes are unstable (half-
lives of hours, minutes, or seconds) and the crosshatched ones are highly unstable. In the
s-process (upper chain in figure), neutrons generated during He burning in stellar interiors
are added to Fe-group seed nuclei. Neutron capture (n) displaces the nuclear composition
rightward until it reaches an unstable box. ⫺ decay (electron emission, see later) then
takes place and a neutron is transformed into a proton. This results in a diagonal move-
ment in the chart of nuclides (upward and leftward).
The lower chain in figure exemplifies the r-process: several neutrons are added in a
rapid sequence until a highly unstable Z-N combination is attained (crosshatched boxes),
⫺ decay then takes place.
The p-process takes place in supernovas by addition of protons followed by ⫹ decay
(positron emission). s-process nuclides are also transformed into nuclides of higher Z by
photodisintegration reactions (gamma rays eject neutrons). Photodisintegration (by cos-
mic ray particles) is also responsible for the formation of the light nuclides 6Li, 9Be, 10 B,
and 11B by fragmentation of the nuclei of carbon, nitrogen, and oxygen. Table 8.1 lists in
an extremely concise fashion the main nuclear reactions that lead to the formation of the
light nuclides.
All elements with Z ⬎ 83 (Bi) are unstable and belong to chains of radioactive
decay, or decay series. Three decay series include all radioactive elements in the
Z ⬎ 83 part of the chart of nuclides—namely, 4n, 4n ⫹ 2, and 4n ⫹ 3 (because
the decay takes place by ␣ emission with mass decrease of four units, or by 
emission with a negligible mass decrease, all nuclides within a series differ by
Figure 11.1 Chart of nuclides. •: stable, °: unstable. Reproduced with modifications from
Rankama (1954).
710 / Methods
Figure 11.2 Enlarged portions of Segré chart of nuclides, showing s-process (upper
chain) and r-process (lower chain). White boxes: stable nuclides; diagonally ruled boxes:
unstable nuclides; crosshatched boxes: highly unstable nuclides.
multiples of four, and hence they are in a 4n-type series, with n an integer). The
major branches of the three decay series are shown in figure 11.3. The parent
nuclide of 4n series is 232Th, which has 208Pb as its stable end-product. The 4n ⫹
2 series has 238U as parent nuclide and 206Pb as stable end-product, and the 4n ⫹
3 series has 235U as parent nuclide and 207Pb as stable end-product.
All naturally occurring radioactive substances (besides those specified in the
Isotope Geochemistry / 711
Table 11.1 Nucleosynthesis and the main nuclear reactions taking place in stars. * ⫽ unstable;
n ⫽ neutron; ⫹ ⫽ neutrino;  ⫹ ⫽ positron; ␥ ⫽ gamma radiation.
NUCLEOSYNTHESIS (3 minutes after the Big Bang → 1 hour after the Big Bang; log T ⬵ 9;
1
H ⫹ n → 2D ⫹ ␥ 2
D ⫹ 1H → 3He ⫹ ␥ 3
He ⫹ 3 He → 4He ⫹ 1H ⫹ 1H
HYDROGEN BURNING (first-generation stars; log T ⬵ 7)
1
H ⫹ 1H → 2D ⫹  ⫹ ⫹ ⫹ 2
D ⫹ 1H → 3He ⫹ ␥ 3
He ⫹ 3He → 4He ⫹ 1H ⫹ 1H
3
He ⫹ 4He → 7Be ⫹ ␥
7
Be ⫹ e⫺ → 7Li ⫹ ⫹ 7
Li ⫹ 1H → 8Be* → 4He ⫹ 4He
7
Be ⫹ 1H → 8B ⫹ ␥ 8
B → 8Be* ⫹  ⫹ ⫹ ⫹ 8
Be* → 4He ⫹ 4He
CARBON-NITROGEN CYCLE (second- and later-generation stars)
12
C ⫹ 1H → 13N ⫹ ␥ 13
N → 13C ⫹  ⫹ ⫹ ⫹
13
C ⫹ 1H → 14N ⫹ ␥ 14
N ⫹ 1H → 15O ⫹ ␥ 15
O → 15N ⫹  ⫹ ⫹ v ⫹
15
N ⫹ 1H → 12C ⫹ 4He
HELIUM-BURNING (red giants; log T ⬵ 8)
4
He ⫹ 4He → 8Be* 8
Be* ⫹ 4He → 12C ⫹ ␥ 12
C ⫹ 4He → 16O ⫹ ␥
16
O ⫹ 4He → 20Ne ⫹ ␥
CARBON-BURNING reactions (log T ⬵ 8.8)
12
C ⫹ 12C → 24Mg ⫹ ␥
12
C ⫹ 12C → 23Mg ⫹ n
12
C ⫹ 12C → 23Na ⫹ 1H
12
C ⫹ 12C → 20Ne ⫹ ␣
12
C ⫹ 12C → 16O ⫹ 2␣
OXYGEN-BURNING reactions (log T ⬵ 9.4)
16
O ⫹ 16O → 31S ⫹ n
16
O ⫹ 16O → 31P ⫹ 1H 31
P ⫹ ␥ → 30Si ⫹1H 30
Si ⫹ ␥ → 29Si ⫹ n
29
Si ⫹ ␥ → 28Si ⫹ n
16
O ⫹ 16O → 28Si ⫹ ␣
decay series of figure 11.3) are shown in table 11.2, together with their half-lives,
isotopic abundances, and decay products. All the listed nuclides with A ⱖ 40 are
the results of early cosmic events. Lighter radionuclides are continuously pro-
duced in the lower atmosphere by interaction of neutrons (produced by cosmic
rays and slowed to thermal energies) with atmospheric nitrogen, oxygen, and
argon nuclei. Note that ⫺ emission and ␣ decay are by far the most common
decay processes for natural radionuclides. Positron emission ( ⫹ ) is observed only
in the decay of 26Al, 36C1, 59Ni and in a small branch of 40 K (0.001%; see section
11.7.3). Electron capture (EC) is also a quite limited process.
Figure 11.3 Major branches of decay chains of heavy radionuclides. heavy arrows: ␣ decay: light arrows:  decay. a ⫽ years, d ⫽ days,
h ⫽ hours, s ⫽ seconds.
Isotope Geochemistry / 713
Nuclear mass is always lower than the sum of proton and neutron masses because
of mass-energy conversion, which transforms part of the mass into binding en-
ergy. For example, 4He has mass 4.002604, whereas the total mass of two protons
and two neutrons is 4.032981. Based on Einstein’s equation
E = mc 2 (11.1)
(where c is the speed of light: 2.99792458 ⫻ 108 m ⭈ s⫺1 ), the mass differential
714 / Methods
Figure 11.4 Mean binding energy per nucleon as a function of A. From G. Friedlander,
J. W. Kennedy, E. S. Macias, and J. M. Miller, Nuclear and Radiochemistry, copyright
1981 by John Wiley and Sons. Reprinted by permission of John Wiley & Sons.
corresponds to a binding energy of 28.3 MeV (i.e., about 7.1 MeV per nuclear
constituent). The average binding energy per nucleon is roughly constant (7.4 to
8.8 MeV), except in a few of the lightest nuclei. It increases slightly from light
elements to A ⫽ 60 (iron and nickel nuclei) and then decreases more or less
linearly with increasing A (figure 11.4).
Both protons and neutrons in the nucleus have an intrinsic angular momen-
tum resulting from spinning of nucleons,
h
P=I × , (11.3)
2π
L = m×v (11.4 )
that is an integral multiple of h/2, being associated with motion along nuclear
orbits with angular frequency ⫽ 2v. Because the intrinsic spin of protons and
neutrons is I ⫽ 12, the nuclear spin of nuclides with odd A is 12; if A is even, it is
zero or an integer. Because rotation of a charged particle produces a magnetic
moment, nuclides with nuclear spins different from zero also have a magnetic
moment of magnitude ZeP/2mc. The magnetic moment of a single proton is
Isotope Geochemistry / 715
1
taken as the unit of nuclear magnetic moment, corresponding to 658 of the Bohr
magneton (magnetic moment of electron).
Because experimental evidence indicates that the volume and total binding
energy of nuclei are proportional to the number of nucleons, it can be deduced
that nuclear matter is incompressible and that each nucleon is able to interact
only with a limited number of surrounding nucleons. This evidence leads to for-
mulation of the binding energy and mass of nuclei in terms of a “charged drop”
model—i.e., the nuclei are assimilated to charged liquid drops with surface ten-
sion. Total binding energy EB is the result of various energy contributions, each
with simple functional dependence on mass A and charge Z of the nucleus (Myers
and Swiatecki, 1966):
N − Z N − Z
2 2
E B = c1 A1 − k − c 2 A 1 − k
2 3
A A
(11.5)
− c 3Z 2 A −1 3 + c 4Z 2 A −1 + δ .
Although equation 11.5 has only six adjustable parameters, it is able to reproduce
the energy of the approximately 1200 nuclides of known mass. When binding
energy is expressed in MeV, the empirically adjusted coefficients have the follow-
ing values: c1 ⫽ 15.677 MeV, c2 ⫽ 18.56 MeV, c3 ⫽ 0.717 MeV, c4 ⫽ 1.211 MeV,
and k ⫽ 1.79. ␦ is associated with the number of protons (if A is odd, ␦ ⫻ A ⫽
0; if it is even, ␦ ⫻ A ⫽ ⫾132).
The first term on the right in equation 11.5 is the volume energy, proportional
to the number of protons and neutrons in the nucleus. It contains a symmetry
correction term, proportional to (N ⫺ Z)2 /A2, accounting for the fact that nuclear
forces are at maximum when N ⫽ Z and decrease symmetrically on both sides of
N ⫽ Z. The second term is the surface energy, which accounts for the fact that
nucleons at the surface of the nucleus have unsaturated forces that reduce binding
energy proportionally to A2/3 (and hence to the surface, A being proportional to
volume). The third term is the electrostatic Coulomb interaction, accounting for
Coulomb repulsion between protons, which have homologous charge. This term
also reduces the binding energy of the nucleus. Because the Coulomb interaction
is proportional to Z2, it becomes increasingly important with the increase of Z
and is responsible for the observed progressive relative increase of N/Z (see figure
11.1). The fourth term in equation 11.5 is a correction for Coulomb energy, re-
sulting from the nonuniform charge distribution (or “diffuse boundary”).
All nuclides are generally subdivided into four types, depending on whether they
contain even or odd numbers of protons and neutrons, as shown in table 11.3.
716 / Methods
n → p + β−. (11.6)
p → n + β+. (11.7 )
(
M = ZMP + A − Z Mn − E B , ) (11.8)
( )( ) ( )( ) ( )
2
M = f1 A Z − ZA + f 2 A Z − ZA − δ A . (11.9)
Because, for a constant value of A, f 1(A), f 2(A), and ␦(A) are also constant,
equation 11.9 is a parabola (for a given value of ␦ ) with maximum at Z ⫽ ZA :
Isotope Geochemistry / 717
∂M
=0 → ZA = −
( ).
f2 A
2 f ( A)
(11.10)
∂Z 1
Figure 11.5 shows the nuclear energy parabolas for odd (A ⫽ 125) and even
(A ⫽ 128) nuclides. These parabolas represent sections of the nuclear energy sur-
face on the Z-EB plane. Because ␦ is zero at any odd value of A, there is a single
parabola, whereas for an even A there are two parabolas, separated along the
energy axis by 2␦ /A.
The usefulness of plots such as those in figure 11.5 stems from the fact that
they give approximate values of the energy available for  decay between neigh-
boring isobars. For odd A (figure 11.5A) there is only one stable nuclide—i.e.,
the one nearest to the minimum of the curve. Note here that, as is customary,
energy parabolas are drawn in a reverse fashion, with the minima actually corre-
sponding to maxima in the Z-EB plane. Note also that in some instances the
energy minimum may be achieved at a noninteger value of Z, because equation
11.5 is continuous in Z. In these cases, the actual minimum is attained at the
integer value of Z nearest to the calculated one.
A nuclide of type 1 emitting ⫹ or ⫺ particles is transformed into a type 4
nuclide, lying on a parabola translated by 2␦/A (figure 11.5B). After a new ⫹ or
⫺ emission, the type 4 nuclide is reconverted into a type 1 nuclide, and so on.
In a  emission chain, nuclides of different nuclear charges alternate between the
two energy parabolas in the energy–nuclear charge field. The  decay process
operating on even A parent nuclides may result in two or more stable isotopes of
the even-even type. For example, in figure 11.5B, both 128 128
52 Te and 54 Xe can be
considered stable, having a binding energy lower than that of the odd parent 128 53 I.
The energy associated with  emission during nuclear decay is not constant,
but varies from a minimum value near zero to a maximum corresponding to the
energy gap between the parent nuclide and the radiogenic nuclide. This apparently
contrasts with the principle of energy conservation. However, as shown by Pauli
(1933), a third particle participates in the process. This particle is called a “neu-
trino” ( because of its very limited mass and charge. The existence of the neu-
trino (and of its corresponding “antineutrino,” ), postulated on purely theoreti-
cal grounds by Pauli, has only recently received experimental confirmation.
Beta particles do not transfer all the binding energy freed by the decay pro-
cess; part of it resides in neutrinos. The mean  emission frequency is one-third
of the energy gap involved in nuclear decay; in other words, most emitted  parti-
cles have kinetic energy corresponding to one-third of the energy gap. Only a few
particles are emitted with a frequency corresponding to maximum energy. Emit-
ted  particles may penetrate surrounding matter before losing most of their ki-
netic energy. Secondary electrons hit by emitted  particles also have high energy
and may be expelled by their localized orbitals, causing successive ionizations.
Figure 11.5 Energy versus nuclear charge plots for odd parity (A) and even parity (B) nuclides. Experimentally observed values of binding
energy differences are reported in parentheses for comparative purposes. From Nuclear and Radiochemistry, G. Friedlander and J. W. Kennedy,
Copyright 1956 by John Wiley and Sons. Reprinted by permission of John Wiley & Sons, Ltd.
Isotope Geochemistry / 719
Let us consider, for instance, the decay process of the unstable oxygen isotope
14 14 14
8 O, which is transformed into stable 7 N (figure 11.6). The 8 O decays, emitting
⫹
two series of  particles of different energies. In decay process (1) in figure 11.6,
a 14
7 N isomer forms in an unstable excited state, and reaches stability by emission
of ␥ rays of different energies. The first ␥ ray has an energy of 2.313 MeV and
leads to an excited state 1.02 MeV above the ground level. The second decay step
involves positron-electron interaction and leads to emission of two ␥ photons
with energies of 1.02 MeV, which reduce the energy of the daughter nuclide at the
⫹
7 N. The energy of the 
ground state of stable 14 particles involved in decay pro-
cess (2) is higher (4.1 MeV) and leads directly to the excited state preceding
positron-electron interaction.
We can summarize the  decay process with two examples:
40
19 K → 40
20 Ca + β − + v + 1.312 MeV (11.12 )
and
8 O + β + v + 1.655 MeV.
→ 18
18 +
9F (11.13)
p → n + β+ (11.14 )
(i.e., 1.02 MeV; cf. eq. 11.11), an electron of the inner K level enters the nucleus,
leaving behind a vacancy that is occupied by a more external electron. Vacancy
migration proceeds toward the external orbitals and is associated with X-ray
emission. An example of electron capture (EC) decay is
57 La + 56 Ba + hv ( γ ) − hv ( X ).
e − → 138
138
(11.15)
Photoelectric Effect
Emitted ␥ radiation interacts with electrons of the surrounding matter. If the fre-
quency of the emitted radiation exceeds the energy level corresponding to the
ionization potential of the element, the electron may be expelled from its localized
Isotope Geochemistry / 721
orbital (K-level electrons are normally involved in the process). Because hv is the
energy of the ␥ radiation and W is the energy necessary to expel the electron from
its orbital (or “work function of the metal,” an important parameter in mass
spectrometry), expulsion will take place only when hv is greater than W. In this
case, the expelled electron achieves kinetic energy corresponding to the surplus:
1
mv 2 = hv − W . (11.16)
2
Compton Effect
When ␥ radiation hits an electron, it is deviated from its original trajectory and,
losing energy, changes its frequency. The increase in frequency consequent upon
anelastic scattering is only a function of the angle between incident and deflected
rays and does not depend on the energy of the incident radiation. The importance
of this fact, known as the Compton effect, increases as the atomic number of the
element decreases.
Internal Conversion
Expelled electrons belong to the K-level of the emitting atom. Also in this case,
as in K-capture, the created vacancy is progressively occupied by more external
electrons and is associated with X-ray emission.
722 / Methods
P = λ ∆t, (11.17)
1 − P = 1 − λ ∆t. (11.18)
This probability remains unchanged even if the radionuclide has survived the first
elapsed time increment. Combining the single probabilities for n elapsed time
intervals, we obtain
n
N t
= 1 − λ , (11.19)
N0 n
t = n × ∆t. (11.20)
n
lim 1 + x = exp( x ) ,
n → ∞ (11.21)
n
we obtain
N
= exp( − λ t) (11.22)
N0
N 1
ln = ln = − λ t1 2 (11.23)
N0 2
ln 2 0.69315
t1 2 = = . (11.24)
λ λ
The average life of a given radionuclide corresponds to the inverse of the decay
constant—hence, to 1/ln 2 times the half-life t1/2, as obtained by integration of
the time of existence of all radionuclides divided by the initial number N 0 :
t=∞
1 t1 2 1
τ =−
N0
∫ tdN =
ln 2
=
λ
. (11.25)
t=0
Because in nature the ground state of a stable nuclide is often attained by decay
chains involving intermediate species decaying at different rates, it is worth evalu-
ating the implications of the relative magnitudes of the various decay constants
on the isotopic composition of the element.
If we call N1 the number of nuclides of the first species, decaying at rate 1 ,
the number of disintegrated nuclides of type 1 per increment of time dt is
dN1
= − λ 1N1 . (11.26)
dt
dN 2
= λ 1N1 − λ 2 N2 . (11.27)
dt
Several substitutions yield the following equation, relating the population of the
second decaying species to elapsed time t:
λ1 N0,1
N2 =
λ 2 − λ1
[ exp (− λ t ) − exp (−λ t )] + N
1 2 0,2 ( )
exp − λ 2 t , (11.28)
where N0,1 and N0,2 are the populations of species 1 and 2 at time t ⫽ 0.
724 / Methods
We can now distinguish three general cases, depending on whether the first
decaying species has a longer, a much longer, or a shorter half-life than that of
the daughter nuclide. These three cases are transient equilibrium, secular equilib-
rium, and nonequilibrium.
R T = R 1 + R 2 = c1 λ1 N1 + c2 λ 2 N2 . (11.29 )
λ1 N 0,1
N2 =
λ 2 − λ1
(
exp − λ1t ) (11.30)
and, because
( )
N 0,1 exp − λ1t = N1 , (11.31)
N1 λ − λ1
= 2 . (11.32)
N2 λ1
As shown in figure 11.8B, because the ratio N1 /N2 becomes constant, the
slopes of the combined decay curves of the two radionuclides attain a constant
value corresponding to the half-life of the longer-lived term (curves a and b in
figure 11.8B). Moreover, assuming identical detection coefficients for the two spe-
cies, their radioactivity ratio also attains a constant value of
R 1 λ 2 − λ1
= . (11.33)
R2 λ2
Isotope Geochemistry / 725
Figure 11.8 Composite decay curves for (A) mixtures of independently decaying species,
(B) transient equilibrium, (C) secular equilibrium, and (D) nonequilibrium. a: composite
decay curve; b: decay curve of longer-lived component (A) and parent radio nuclide (B,
C, D); c: decay curve of short-lived radionuclide (A) and daughter radionuclide (B, C, D);
d: daughter radioativity in a pure parent fraction (B, C, D); e: total daughter radioactivity
in a parent-plus-daughter fraction (B). In all cases, the detection coefficients of the various
species are assumed to be identical. From Nuclear and Radiochemistry, G. Friedlander and
J. W. Kennedy, Copyright 1956 by John Wiley and Sons. Reprinted by permission of
John Wiley and Sons Ltd.
N1 λ
= 2. (11.34)
N2 λ1
Assuming identical detection coefficients for the two species, the radioactivity
ratio obviously reduces to 1. This condition, known as “secular equilibrium,” is
illustrated in figure 11.8C for t1/2,1 ⫽ ∞ and t1/2,2 ⫽ 0.8 hr. Secular equilibrium
can be conceived of as a limiting case of transient equilibrium with the angular
coefficient of decay curves progressively approaching the zero slope condition
attained in figure 11.8C.
11.5.3 Nonequilibrium
The case of “nonequilibrium” occurs whenever the decay constant of the parent
nuclide is higher than the decay constant of the daughter nuclide (i.e., 1 ⬎ 2 )
and is illustrated in figure 11.8D for 1 / 2 ⫽ 10. As the parent decays, the number
of daughter nuclides progressively rises, reaching a maximum at time tm. It then
decreases in a constant slope, characteristic of its own half-life. Time tm is found
graphically by intersecting the decay curve of the parent (curve b in figure 11.8D)
and the extrapolation of the final decay curve to time zero (curve c). The same
parameter can be obtained analytically by applying the equation
ln λ 2 − ln λ1
tm = . (11.35)
λ 2 − λ1
The translational partition function Qtrans is equal to the classical one at all
temperatures
Isotope Geochemistry / 727
ε k ,i , x ε k ,i , y ε k ,i , z
Q trans = ∑ exp − − −
kT
(11.37 )
i kT kT
where k,i,x is the kinetic energy of molecule i along direction x. Also the rota-
tional partition function Qrot is equal to the classical one, with the exception of
hydrogen (see section 11.8.2).
In the harmonic approximation, Qvib is given by
(
exp − Xi 2 )
Q vib = ∏ 1 − exp( − X )
(11.38)
i i
hvi បωi
Xi = = (11.39)
kT kT
s • Q • m°
32
s•
o f = o o •
s s Q m
(11.40)
exp − Xi • 2 1 − exp − Xi •
Xi •
= ∏ .
i Xi ° ( )[
exp − Xi ° 2 1 − exp − Xi ° ( )]
•
The ratio of symmetry numbers s /s° in equation 11.40 merely represents the
relative probabilities of forming symmetrical and unsymmetrical molecules, and
•
m and m° are the masses of exchanging molecules (the translational contribution
to the partition function ratio is at all T equal to the 32 power ratio of the inverse
molecular weight). Denoting as ⌬Xi the vibrational frequency shift from isotopi-
cally heavy to light molecules (i.e., ⌬Xi ⫽ X °i ⫺ X •i ) and assuming ⌬Xi to be
intrinsically positive, equation 11.40 can be transated into
s• 1 − exp − Xi • + ∆Xi
Xi • ∆Xi
f = ∏ exp . (11.41)
s° i Xi • + ∆Xi 2 1 − exp − Xi •
728 / Methods
F = − kT ln Q (11.42 )
Based on equation 11.41, the difference between the Helmholtz free energies of
formation of two isotopic molecules with respect to their gaseous atoms depends
on the shift of vibrational frequencies between heavy and light isotope-bearing
compounds—i.e., according to Bigeleisen and Mayer (1947),
1 1 − exp − Xi •
∆F • − ∆F ° ∆ Xi
= ∑ − ∆X i + ln 1 + + ln
kT i 2 Xi •
(
1 − exp − Xi °
)
(11.43)
s•
+ ln ,
s°
•
where ⌬F and ⌬F° are the Helmholtz free energies of formation of isotopically
heavy and light molecules. If ⌬Xi is small, which is the case for all isotopes except
hydrogen, equation 11.43 can be reduced to
∆F • − ∆ F ° 1 1 1 s•
= −∑ − + ∆ Xi + ln ° , (11.44)
kT i 2 Xi • exp Xi • − 1 s
•
and the separative effect (s /s°) f becomes
s• 1 1 1
° f = 1+ ∑ − + ∆Xi . (11.45)
s i 2 Xi • exp Xi • − 1
•
If X •i is also small, then the terms in brackets in equation 11.45 approach X i /12
and the separative effect reduces to
•
s• Xi ∆Xi
° f = 1+
s
∑ 12
. (11.46)
i
•
Because ⌬Xi is positive, (s /s°) f is always greater than 1, and the heavy isotope
preferentially stabilizes in the condensed phase whereas the light isotope favors
the gaseous phase.
Isotope Geochemistry / 729
Figure 11.9 Separative effect per unit shift, plotted against nondimensionalized frequen-
cies X ⫽ h /kt ⫽ ប /kT. Reprinted from Bigeleisen and Mayer (1947), with permission
from the American Institute of Physics.
At low temperatures and light frequencies, the separative effect per unit shift
(the terms in brackets in eq. 11.44 and 11.45) approaches 12 (figure 11.9) and the
Helmholtz free energy difference approaches the differences in zero-point ener-
gies. At high T (low frequencies), the separative effect per unit shift approaches
•
zero and the total separative effect (s /s°) f approaches 1, so that no isotopic frac-
tionation is observed.
A useful approximation for the isotopes of heavy elements (Bigeleisen and
Mayer, 1947) is
s• M ∆m 2
° f = 1+ X s n, (11.47)
s 24 m 2
where m is the mass of the central atom (isotope) surrounded by n ligands of mass
M, ⌬m is the difference in isotopic masses, and nondimensionalized frequency Xs
is related to the totally symmetric frequency arising from the stretching
of bonds with central atom ns by
hνs h 2πC
Xs = = (11.48)
kT kT M
Molecule s f K
SiF4 800 1.111
1.002
SiF62⫺ 600 1.109
SnCl4 367 1.00256
1.00025
SnCl62⫺ 314 1.00281
28
SiF4 + 30
SiF62 − ⇔ 30
SiF4 + 28
SiF62 − (11.49 )
120
SnCl 4 + 118 SnCl 26 − ⇔ 118 SnCl 4 + 120 SnCl 26 − . (11.50 )
Note that the separative effect is quite significant, even for the relatively “heavy”
isotopes of tin. Note also the large cancellation effect on the isotopic fractionation
constant, arising from superimposed partial reactions such as
120
SnCl 4 + 118 Sn → 118 SnCl 4 + 120 Sn (11.51)
and
118
SnCl 26 − + 120 Sn → 120 SnCl 26 − + 118 Sn . (11.52 )
a A ° + b B • ⇔ a A• + b B ° , (11.53)
• •
where A°, A and B°, B are molecules of the same type that differ only in the
isotopic composition of a given constituent element (open and full symbols: light
and heavy), and a and b are stoichiometric coefficients.
In light of equation 11.42, the equilibrium constant for the isotopic exchange
reaction of equation 11.53 reduces to
Isotope Geochemistry / 731
(Q )
a
•
A Q A°
K 53 = . (11.54)
(Q )
b
•
B Q B°
3N − 6
(
exp − Xi 2 ),
Q= ∏ 1 − exp( − X )
(11.55)
i =1 i
where N is the number of atoms in the crystal, and the ratio of partition functions
of the isotopically heavy and light compounds (Kieffer, 1982) is
Q• 3N − 6
(
exp − Xi • 2 ) 1 − exp( −X ) . °
∏
i
=
Q° i =1 1 − exp − Xi •
( ) exp( −X 2) i
° (11.56)
•
Introducing the masses of exchanging isotopes (m and m°, respectively), the
ratio of partition functions for crystalline components can be related to that of
the primitive unit cell (Kieffer, 1982), thus defining “reduced” partition function
ratio f (whose formulation is equivalent to that obtained by Bigeleisen and
Mayer, 1947, and Urey, 1947, for gaseous molecules):
( 3 2 )r
Q • ′ Q • m°
f = = , (11.57)
Q °′ Q ° m•
ωl
− ប ω
F = 3 nNβ + kT ∫ ln 1 −
exp
g (ω )dω ,
kT
(11.58)
0
where
3 nN
1
3nNβ = −U0 −
2
∑ ប ω. (11.59)
r =1
frequency. In equation 11.59, U0 is the potential energy at the bottom of the po-
tential well, and the summation term is zero-point energy.
As shown schematically in figure 11.10, the frequency distribution for an iso-
topically light compound g°(v) [and the corresponding g°( )] can be expected to
•
differ from the frequency distribution of the isotopically heavy counterpart g (v)
•
[or g ( ].
Kieffer (1982) suggests a set of rules describing expected frequency shifts for different
types of vibrational modes upon substitution of a heavy isotope into the mineral structure.
Kieffer’s (1982) rules for 18O-16O substitution in silicates are as follows.
( 3 2 )r
3s
Xi ° m•
∏ Xi •
=
m°
, (11.60)
i =1
Because
F° − F• 3 m°
ln f = + r ln • , (11.61)
kT 2 m
•
calculations of F° and of the corresponding F for isotopically heavy compounds
(through application of the above general rules for frequency shifts induced by
isotopic substitution) allows determination of partition function ratio f and of
the corresponding “reduced partition function” (103 /r) ln f (see section 11.8.1).
120
SnCl 4 + 118SnCl 26 − ⇔ 118SnCl 4 + 120SnCl 26 − . (11.62 )
120
SnCl 4 + 2 Cl − ⇔ 120SnCl 26 − (11.63)
and
118
SnCl 4 + 2 Cl − ⇔ 118SnCl 26 − . (11.64 )
734 / Methods
To each of these two reactions will apply a reaction constant (K63 and K64 , respec-
tively) that is related to the forward and backward reaction rates by
k + ,63
K 63 = (11.65.1)
k − ,63
and
k + ,64
K 64 = . (11.65.2)
k − ,64
Because at equilibrium,
k k
∆G 62
0
= − RT ln K 62 = − RT ln + ,63 × − ,64 , (11.66)
k + ,64 k − ,63
and because we have seen that the isotopic fractionation constant at low T differs
from 1 (i.e., K62 ⫽ 1.0025 at T ⫽ 300 K; cf. table 11.4), it follows that isotopically
heavy and light molecules cannot be expected to have identical reaction rates at
low T (even in the case of relatively “heavy” isotopes such as 118Sn and 120 Sn).
The effects discussed above may be particularly important when the fractional
mass difference between isotopically light and heavy molecules is high (see the
calculations of Sutin et al., 1961). A quantitative approach to this problem was
afforded by Bigeleisen (1949) on the basis of transition state theory.
Adopting the same notation developed in the previous section, the reaction
rates of isotopically heavy and light molecules are related to the differences in
•
their respective activation energies of reaction E , E°, through
(
∂ ln k k °
•
)=E •
− E°
, (11.67)
∂T kT 2
( •
∂ ln k k ° )= ∂ ln τ ( •
τ° )+X •2
s − X s°
2
∂T ∂T 12T
3 N −6 °
(
)
1 + Xi − 1 exp Xi
°
1 ∆ Xi*
(11.68)
+ ∑ − ,
[ ( ) ]
2
i =1 exp X i° − 1 2 T
Isotope Geochemistry / 735
•
where and ° are transmission coefficients (which, for asymmetric reactions
such as reactions 11.63 and 11.64, must be evaluated experimentally), and ⌬X i*
is the vibrational shift induced by formation of the activated complex.
11.6.3 Diffusion
Some aspects of thermally activated diffusion were discussed in section 5.9.1 in
regard to intracrystalline exchange geothermometry. Concerning more precisely
isotope diffusion, there are two main aspects that may be relevant in geochem-
istry:
C x − C1 x
= erf , (11.69)
C 0 − C1
( )
12
2 Dt
736 / Methods
Figure 11.11 18O/ 16O relative concentration profile as a function of depth in adularia
hydrothermally equilibrated with an 18O-enriched fluid (H218O ⫽ 40%). Reprinted from
B. J. Giletti, M. P. Sennet, and R. A. Yund, copyright 1978, with kind permission from
Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington 0X5 1GB, UK.
where Cx is the 18O concentration at distance x from the surface of the crystal,
C 0 is the concentration at infinite distance, C1 is the concentration at the surface,
t is time, and D the self-diffusion coefficient.
Equation 11.69 was reexpressed by Giletti et al. (1978) in terms of 18O/ 16O
ratios x , 0 , and 1 as
A × erf(Y ) + ρ1
ρx = , (11.70)
1 − A × erf(Y )
where
ρ 0 − ρ1
A= (11.71)
1 + ρ0
and
x
Y = . (11.72)
2(Dt )1 2
Isotope Geochemistry / 737
Results were then interpreted in terms of the usual form for thermally activated
diffusion:
−Ea
D = D 0 exp , (11.73)
RT
Unfortunately, the interpretation of Giletti et al. (1978) does not solve the problem
of differential diffusivities of 18O and 16O. To do this, their experimental results should be
interpreted in terms of interdiffusion of 18O and 16O. Application of Fick’s first law to
interdiffusion of the two species would in fact lead to the definition of an interdiffusion
coefficient D , so that
D18O D16O
D = (11.74)
X 18O D18O + X 16O D16O
(cf. section 4.11). Because, when X18O → 0,D → D18O, interpretation of the relative concen-
tration profile in terms of self-diffusion is appropriate for the inner parts of the crystals,
where 16O is dominant, but it loses its significance near the borders of grains, where inter-
diffusion effects cannot be neglected (see figure 11.11).
The experiment of Giletti et al. (1978) has important implications for the
isotopic reequilibration of feldspars in hydrous environments. Assuming the
diffusion of 18O to obey diffusion equation 11.73, with D0 ⫽ 2.0 ⫻ 10⫺8 cm2 /s
and Ea ⫽ ⫺24 kcal/mole, and defining F as the fractional net 18O exchange oc-
curring with respect to total net exchange at equilibrium, Giletti et al. (1978)
calculated the length of time required to achieve F ⫽ 0.1 and F ⫽ 0.9 according
to temperature and particle size. The results of their calculations are shown in
figure 11.12.
Any feldspar particle of radius less than 10 m is able to exchange with a
hydrothermal fluid, essentially to completion (i.e., Fⱖ 0.9), in less than 500 years,
if the temperature exceeds 300 °C. However, exchanges for larger crystals may
span periods of time from one million to several hundred million years, depending
on size and T, before isotopic equilibrium is attained. Note also that, because of
the rather low activation energy, exchanges have no single “closure temperature.”
This fact introduces us to the next aspect of the problem.
738 / Methods
Figure 11.12 Time required for diffusional exchange of 18O between feldspar and hydro-
thermal fluid according to particle size and temperature. F ⫽ 0.1 can be taken as condition
of absence of exchanges and F ⫽ 0.9 as one of completion of exchanges. Reprinted from
B. J. Giletti, M. P. Sennet, and R. A. Yund, copyright 䉷 1978, with kind permission from
Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington 0X5 1GB, UK.
∂ C Dt − t
= 2 exp ∇ 2 C
0
(11.75)
∂t a τ
and
Isotope Geochemistry / 739
Figure 11.13 Relationship between “geochronological closure” (A) and “frozen equilib-
rium” (B). Diffusion parameters are identical, and T ⫺1 is assumed to increase linearly
with time. From M. H. Dodson, Closure temperature in cooling geochronological and
petrological systems, Contributions to Mineralogy and Petrology, 40, 259–274, figure 2,
1973, copyright 䉷 1973 by Springer Verlag. Reprinted with the permission of Springer-
Verlag GmbH & Co. KG.
∂ C* Dt* − t
= 2 exp ∇ 2 C * + L*
0
(11.76)
∂t a τ *
(Dodson, 1973). Both of these equations are developed assuming a linear de-
crease of temperature T with time t (see the analogy of treatment in section 5.9.1).
is a time constant corresponding to the time taken for D to diminish by a factor
of exp(⫺1), a is the characteristic dimension of the phase, and the asterisk de-
notes the radiogenic isotope.
Equations 11.75 and 11.76 both obey boundary conditions Cx ⫽ C0 every-
where at t0 and C ⫽ C1 at the surface for t ⱖ 0. In equation 11.76, L* is the rate
of production of the radiogenic isotope.
As we see in figure 11.13, for a stable isotope such as 18O, the mean concentra-
tion in a phase attains a constant value after the “closure” of exchanges (“closure
temperature” Tc in this case is graphically determined by intersecting the prolon-
gations of surface concentration and mean concentration; see upper dashed lines
in figure 11.13). However, the radiogenic daughter isotope progressively increases
its concentration, as a result of the virtual cessation of exchanges, beyond the
value of Tc (in this case, Tc and the corresponding tc are graphically determined
740 / Methods
R
Tc = .
Aτ Dt 0 (11.77)
E a ln 2
a
87
Rb α + 87
Srβ ⇔ 87
Srα + 87
Rb β , (11.78 )
the Rb/Sr fractional amounts in the two phases can be related to the equilibrium
constant through
− ∆G780 ( 87
Rb 87
Sr ) β
⋅K .
K 78 = exp =
RT
( 87
Rb 87
Sr )
α
γ , 78 (11.79)
The term within the square brackets in equation 11.79 is the normalized (equilib-
rium) distribution coefficient between the minerals ␣ and  (cf. section 10.8) at
the closure condition of the mineral isochron. Ganguly and Ruitz (1986) have
shown it to be essentially equal to the observed (disequilibrium) distribution co-
efficient between the two minerals as measured at the present time. K␥,78 can be
assumed to be 1, within reasonable approximation. Equation 11.79 can be cali-
brated by opportunely expanding ⌬G 078 over P and T:
Based on equations 11.79 and 11.80, the measured DRb/Sr value for each mineral
of a given isochron must furnish concordant indications in terms of the T and P
of equilibrium. Moreover, the deduced parameters can be assumed to correspond
to the condition of closure of exchanges.
Isotope Geochemistry / 741
In this section we will briefly examine the various methods adopted in geochro-
nology, all based essentially on the constancy of the decay rates of radiogenic
nuclides, which, within reasonable limits, are unaffected by the physicochemical
properties of the system. For more detailed treatment, specialized textbooks such
as Rankama (1954), Dalrymple and Lanphere (1969), Faure and Powell (1972),
Jäger and Hunziker (1979), Faure (1986), and McDougall and Harrison (1988)
are recommended.
11.7.1 Rb-Sr
The element rubidium has 17 isotopes, two of which are found in nature in the
following isotopic proportions: 85Rb ⫽ 0.72165 and 87Rb ⫽ 0.27835. Strontium
has 18 isotopes, four of which occur in nature in the following average weight
742 / Methods
proportions 84Sr ⫽ 0.0056, 86Sr ⫽ 0.0986, 87Sr ⫽ 0.0700, and 88Sr ⫽ 0.8258. 87Rb
decays to 87Sr by ⫺ emission, according to
38 Sr + β
→ 87 + v,
87 −
37 Rb (11.81)
with a decay constant of ⫽ 1.42 ⫻ 10⫺11 a (a ⫽ anna, the Latin word for
“years”), corresponding to a half-life t1/2 of about 4.9 ⫻ 1010 a.
Because the quantity of unstable parent radionuclides P remaining at time t
is related to the decay constant (cf. eq. 11.22) through
P = N 0, P exp ( − λt ) , (11.82)
[
D = P exp (λt ) − 1 , ] (11.83)
87
Srrad = 87
[
Rb exp ( λt ) − 1 . ] (11.84 )
87
Srt = 87
Sr0 + 87
[
Rb exp ( λt ) − 1 , ] (11.85)
Sr Sr
[ ]
87 87 87
Rb
= + exp ( λt ) − 1 . (11.86)
86
Sr t 86
Sr 0 86
Sr
Whenever it is possible to analyze in a given rock at least two minerals that crys-
tallized at the same initial time t ⫽ 0, equation 11.86 can be solved in t. On a
Cartesian diagram with coordinates 87Sr/ 86Sr and 87Rb/ 86Sr, equation 11.86 ap-
pears as a straight line (“isochron”) with slope exp( t) ⫺ 1 and intercept ( 87Sr/
86
Sr) 0 . As shown in figure 11.14A, all minerals crystallized at the same t from the
same initial system of composition ( 87Sr/ 86Sr) 0 rest on the same isochron, whose
slope exp( t) ⫺ 1 increases progressively with t.
Isotope Geochemistry / 743
Figure 11.14 (A) Internal Rb-Sr isochron for a system composed of three crystalline
phases of initial compositions A 0, B0, and C0 formed at time t ⫽ 0 and thereafter closed
to isotopic exchanges up to time of measurement t, when they acquired compositions
At , Bt , and Ct . (B) Effects of geochronological resetting resulting from metamorphism
or interaction with fluids. X1, X2, and X3: bulk isotopic compositions of the three rock
assemblages. In cases of short-range isotopic reequilibration, the three assemblages define
crystallization age and original (87Sr/ 86Sr)0 of the system; the three internal isochrons (con-
cordant in this example) define resetting age.
60 Nd + α
→ 143 t1 2 = 1.06 × 1011 a
147
62 Sm (11.87 )
60 Nd + α
→ 144 t1 2 = 7.0 × 1015 a
148
62 Sm (11.88 )
60 Nd + α
→ 145 t1 2 = 1.0 × 1016 a
149
62 Sm (11.89 )
Neodymium also has seven isotopes, occurring in nature in the following pro-
portions: 142Nd ⫽ 0.2711, 143Nd ⫽ 0.1217, 144Nd ⫽ 0.2385, 145Nd ⫽ 0.0830,
146
Nd ⫽ 0.1719, 148Nd ⫽ 0.0573, and 150 Nd ⫽ 0.0562. Sm-Nd dating is based on
the decay of 147Sm, whose half-life is the only one sufficiently short for geochrono-
logical purposes. The isochron equation is normalized to the 144Nd abundance:
143 Nd 143 Nd
[ ]
147
Sm
144 = 144 + exp (λt ) − 1 . (11.90)
Nd t Nd 0 144
Nd
72 Hf + β
→ 176 + v.
176 −
71 Lu (11.91)
The first applications of the Lu-Hf method were attempted about 20 years ago.
Lutetium has 22 isotopes, two of them occurring in nature in the following pro-
portions: 175Lu ⫽ 0.9741 and 176Lu ⫽ 0.0259. Hafnium has 20 isotopes, six of
which occur in nature in the following proportions: 174Hf ⫽ 0.0016, 176Hf ⫽
0.0520, 177Hf ⫽ 0.1860, 178Hf ⫽ 0.2710, 179Hf ⫽ 0.1370, and 180 Hf ⫽ 0.3524. The
isochron equation is normalized to the 177Hf abundance:
Hf Hf
[ ]
176 176 176
Lu
= + exp ( λt ) − 1 . (11.92)
177
Hf t 177
Hf 0 177
Hf
The decay constant is not known experimentally but has been deduced by com-
parisons with radiometric ages obtained by other methods. The obtained value
is (1.94 ⫾ 0.07) ⫻ 10⫺11 a⫺1, corresponding to a half-life of (3.57 ⫾ 0.14) ⫻
1010 a.
The first application of La-Ce dating to earth sciences was proposed by Ta-
Isotope Geochemistry / 745
58 Ce + β
→ 138 +v
138 −
57 La (11.93)
57 La + 56 Ba + hv ( γ ) + hv ( X ).
e − → 138
138
(11.94 )
Lanthanum has 19 isotopes with A between 126 and 144. Only two occur in
nature, in the following proportions: 138La ⫽ 0.0009 and 139La ⫽ 0.9991. The low
relative amount of 138La and its low decay rate ( ⫽ EC ⫹ ⫺ ⫽ 6.65 ⫻ 10⫺12
a⫺1 ) prevented earlier application of this dating method. Cerium has 19 isotopes,
with masses between 132 and 148, and two isomers at A ⫽ 137 and 139. Four
isotopes are found in nature, in the following proportions: 136Ce ⫽ 0.0019,
138
Ce ⫽ 0.0025, 140 Ce ⫽ 0.8848, and 142Ce ⫽ 0.1108. The isochron equation is
normalized to the 142Ce abundance:
138Ce 138Ce β −
[ ]
138
La
142 = 142 + exp( t ) − 1 . (11.95)
Ce t Ce 0 142
Ce
11.7.3 K-Ar
Potassium has 10 isotopes, three of which occur in nature in the following propor-
tions: 39K ⫽ 0.9326, 40 K ⫽ 0.0001, and 41K ⫽ 0.0673. Argon (from the Greek
verb argéos—i.e., to be inactive) is present in the earth’s atmosphere (0.934% in
mass). It is composed of eight isotopes, three of which are present in nature in
the following proportions: 36Ar ⫽ 0.0034, 38Ar ⫽ 0.0006, and 40 Ar ⫽ 0.9960.
K-Ar geochronology is based on the natural decay of 40 K into 40 Ar (EC,
 ⫽ 10.5%). However, as shown in figure 11.16, 40 K also decays into 40 Ca ( ⫺ ⫽
⫹
λ EC
40
18 Art = 18
40
Ar0 +
λ
EC + λ β
40
{ [(
× 19 K exp λ EC + λ β t − 1 , )] } (11.96)
and the corresponding age equation (obtained under the assumption that no
40
18Ar 0 is present in the mineral at the time of its formation) is
λ EC + λ β 40
1 18 Ar
t= ln 1 +
λ EC + λ β λ EC 40
19 K
(11.97)
40
(
= 1.804 × 10 9 ) ln 1 + 9.54
18 Ar
40 .
19 K
40
18 Artotal = 40
18 Arrad + 40
18 Aratm . (11.98 )
Because the 40 Ar/ 36Ar ratio of the earth’s atmosphere is 295.5 (constant), the
correction for atmospheric argon is made by measuring the amount of 36Ar:
40
18 Arrad = 18 Artotal −
40
295.5 × 36
18 Ar . (11.99 )
In practice, mass spectrometric measurements are carried out on the 40 Ar/ 38Ar
and 38Ar/ 36Ar ratios, for the reasons previously outlined (see Dalrymple and
Lanphere, 1969 for the form of the relevant chronological equation).
748 / Methods
Figure 11.16 Decay scheme for 40K, showing double decay to 40Ca and 40Ar.
40
Ar Ar
40
=
36
Ar t Ar 0
36
(11.100)
λ EC K
{ [( )] }
40
+ exp λ EC + λ β t − 1 .
λ EC + λ β Ar
36
The intercept term ( 40 Ar/ 36Ar) 0 , which accounts for igneous, metamorphic, or
atmospheric sources, is regarded as the excess contribution present at time t ⫽ 0,
whereas the second term is the radiogenic component accumulating in the various
minerals of the isochron by decay of 40 K. If all the minerals used to construct the
isochron underwent the same geologic history and the same sort of contamination
by excess 40 Ar, the slope of equation 11.100 would have a precise chronological
Isotope Geochemistry / 749
Figure 11.17 Whole rock K-Ar isochron of Tuff IB strata from Olduvai Gorge, Tanzania.
From G. Faure (1986), Principles of Isotope Geology, 2nd edition, copyright 䉷 1986 by
John Wiley and Sons. Reprinted by permission of John Wiley & Sons.
40
11.7.4 Ar/ 39Ar
Figure 11.18 Apparent K-Ar ages of minerals from Idaho Springs Formation (Front
Range, Colorado, 1350–1400 Ma) in zone subjected to contact metamorphism by intru-
sion of a quartz monzonite (Eldora stock, 55 Ma). Reprinted from S. R. Hart, Journal of
Geology, (1964), 72, 493–525, copyright 䉷 1964 by The University of Chicago, with per-
mission of The University of Chicago Press.
19 K + n→ + p.
39 39
18 Ar (11.101)
The number of 39Ar atoms formed depends on irradiation time ⌬t, density of neutron
flux ( ⑀ ) as a function of energy , and cross-section of 39K atoms ( ⑀ ):
39
18 Ar = 39
19 [
K ∆ t ∫ ϕ( ε ) σ( ε ) dε .] (11.102)
If we denote the flux production factor in square brackets in equation 11.102 as ⌽, we can
define a parameter J related to 39Ar production in a monitor sample of known age tm:
−1 −1
40
K λ EC
J = 19
m Φ
19 K
39
λ EC + λ β
(11.103)
Ar
{exp [(λ ) ] }
39
= 18
+ λβ tm − 1 .
Ar m
40 EC
18
Produced
Isotope Calcium Potassium Argon Chlorine
36
Ar 40
Ca(n, n␣) – – –
37
Ar 40
Ca(n, ␣) 39
K(n, n d ) 36
Ar(n, ␥ ) –
38
Ar 42
Ca(n, n␣) 39
K(n, d) 40
Ar(n, n d,  ⫺) 37
Cl(n, ␥,  ⫺)
41
K(n, ␣,  ⫺)
39
Ar 42
Ca(n, ␣) 39
K(n, p) 38
Ar(n, ␥ ) –
43
Ca(n, n␣) 40
K(n, d) 40
Ar(n, d,  ⫺)
40
Ar 43
Ca(n, ␣) 40
K(n, p) – –
44
Ca(n, n␣) 41
K(n, d)
40
Ar
=
[(
exp λ EC + λ β t − 1 )] , (11.104)
39
Ar J
1 40
Ar
t= ln J + 1 . (11.105)
λ EC + λ β 39
Ar
stable with respect to the time normally involved in analyses. Ages obtained
through application of equation 11.105 are normally referred to as “total argon
release dates” (cf. Faure, 1986) and are subject to the same uncertainties as con-
ventional K-Ar ages. Moreover, corrections for atmospheric Ar through equation
11.99 are complicated by the production of 36Ar by 40 Ca(n,n␣ ) reaction during
irradiation (see table 11.6). More generally, neutron irradiation is responsible for
a complex alteration of the pristine isotopic composition of Ar by nuclear reac-
tions involving the neighboring isobars, as summarized in table 11.6.
According to Dalrymple and Lanphere (1971), the appropriate 40 Ar/ 39Ar
ratio to be introduced in age equation 11.105 can be derived from the measured
ratio ( 40 Ar/ 39Ar)meas by application of
40
Ar 40
Ar 36
Ar
= − C1 + C1C 2 D − C 3
Ar meas Ar meas
39 39 39
Ar
(11.106)
( )
−1
× 1 − C4 D ,
752 / Methods
Figure 11.19 The first 40Ar/ 39Ar age spectrum. Reprinted from Turner et al. (1966), with
kind permission from Elsevier Science Publishers B.V, Amsterdam, The Netherlands.
where C1 ⫽ 295.5 is the 40 Ar/ 36Ar ratio in the atmosphere, C2 ⫽ 2.72 ⫾ 0.014 ⫻
10⫺4 is the 36Ar/ 37Ar ratio induced by interfering nuclear reactions involving Ca,
C3 ⫽ 5.9 ⫾ 0.42 ⫻ 10⫺3 is the 40 Ar/ 39Ar ratio induced by interfering nuclear
reactions involving K, C4 ⫽ 6.33 ⫾ 0.04 ⫻ 10⫺4 is the 39Ar/ 37Ar ratio induced by
interfering nuclear reactions involving Ca, and D is the 37Ar/ 39Ar ratio in the
sample, after correction for decay of 37Ar (t1/2 ⫽ 35.1 days). We refer readers to
Brereton (1970) and to Dalrymple and Lanphere (1971) for detailed accounts of
the necessary corrections.
The main virtue of the 40 Ar/ 39Ar dating method is that the isotopic com-
position of the various argon fractions released during stepwise heating of the
sample (or “age spectrum”) furnishes precious information about the nature of
excess argon and the thermal history of the sample. The first interpretation in this
sense was that of Turner et al. (1966), who analyzed a sample of the disturbed
Bruderheim chondrite and arranged the mass spectrometric 40 Ar/ 39Ar data in a
series corresponding to the cumulative percentage of released 39Ar (figure 11.19).
The data reported in figure 11.19 refer to two separate irradiations, but their
distribution is consistent with the model age spectrum of a sample composed of
spheres with a lognormal distribution of radii that underwent a brief episode of
radiogenic 40 Ar* outgassing (90%) 0.5 Ga before present (solid line in figure
11.19). According to the single-site diffusion model of Turner (1968), if we as-
sume homogeneous distribution of 39Ar, a single (thermally activated) transport
mechanism, identical diffusion rates for all Ar isotopes, and a radiogenic 40 Ar*
boundary concentration of zero, a set of model age curves can be parametrically
generated, corresponding to the various degrees of outgassing. As shown in figure
Isotope Geochemistry / 753
Figure 11.20 Theoretical age spectra based on Turner’s (1968) single-site diffusion model.
True age of sample is 4.5 Ga, and outgassing episode took place 0.5 Ga ago. Reprinted
from Turner (1968).
11.20, the theoretical age spectra thus obtained differ greatly according to the
size distribution of the outgassing particles (uniform radius or lognormal distri-
bution) if the 40 Ar* loss exceeds 40%. The first infinitesimal increment of gas
extracted represents the age of a brief thermal disturbance that affected the
sample, causing a partial loss of radiogenic argon, whereas the true age is attained
at maximum 39Ar release.
Note that Turner’s theoretical spectra for uniform spheres indicate that, if
40
Ar* loss exceeds 10%, the maximum age achieved at complete 39Ar release does
not correspond to the true age of the sample—i.e., the asymptotic rise of the
curves does not reach 40 Ar/ 39Ar ⫽ 1 for both lognormal and uniform distribution
of radii. In these cases, assessment of 40 Ar* loss must be established from the
form of the experimental age spectrum to correct appropriately for the lowering
of maximum age. For example, in the case exemplified in figure 11.20, a 40 Ar*
loss of 60% would lead to an apparent maximum age of 3.9 Ga for an aggregate
of uniform spheres, whereas the age perturbation would be greatly reduced for
an aggregate of spheres with a lognormal distribution of radii (apparent age 艐
4.4 Ga). Partial modifications of Turner’s (1968) single-site diffusion model were
presented by Harrison (1983) to account for multiple 40 Ar* losses in samples with
particularly complex thermal histories.
Further complications in 40 Ar/ 39Ar age spectra arise whenever samples con-
tain excess 40 Ar not produced by in situ decay of 40 K. In these cases, the presence
of 40 Ar violates the zero concentration boundary condition of Turner’s model,
and the first infinitesimal increments of extracted argon often display extremely
754 / Methods
Figure 11.21 Age spectra of hornblendes of a Paleozoic gabbro (367 Ma) intruded by a
granitic body during the Cretaceous (114 Ma). Samples underwent permeation of 40Ar*
from lower crustal portions and differential losses of radiogenic argon (32, 57, and 78%
respectively), proportional to distance from contact (0.3, 1, and 2.5 km). Reprinted from
T. M. Harrison and I. McDougall, Geochimica et Cosmochimica Acta, 44, 2005–2020,
copyright 䉷 1980, with kind permission from Elsevier Science Ltd., The Boulevard, Lang-
ford Lane, Kidlington 0X5 1GB, UK.
old apparent ages, which decrease rapidly in the subsequent stepward releases.
An emblematic representation of such complex age spectra is given in the study
of Harrison and McDougall (1980) on hornblendes from Paleozoic gabbro in-
truded by a granitic body during the Cretaceous. The observed 40 Ar/ 39Ar spectra
reveal loss profiles compatible with Turner’s model for an aggregate of spheres
with lognormal distribution of radii (figure 11.21) with partial losses of 31, 57,
and 78% of radiogenic argon and a true age of 367 Ma (coinciding with an isoch-
ron age of 366 ⫾ 4 Ma; Harrison and McDougall, 1980). However, the presence
of intergranular excess argon is evident from the initial steeply descending relative
concentration profiles (which in some samples of the same outcrop generate initial
apparent ages as high as 3.5 Ga; cf. Harrison and McDougall, 1980).
Additional complexities arise whenever the permeating excess argon diffuses
from intergranular boundaries to lattice sites or whenever the solid phase under-
Isotope Geochemistry / 755
Figure 11.22 (A) Saddle-shaped age spectra of calcic plagioclases from amphibolites,
Broken Hill, Australia. (B) Arrhenius plots of reactor-produced isotopes for two of the
five samples, defining existence of three diffusion domains corresponding to albite-rich
lamellae (domain 1) and anorthite-rich lamellae of different widths (domains 2 and 3).
Reproduced with modifications from T. M. Harrison and I. McDougall (1981), with kind
permission from Elsevier Science Publishers B.V, Amsterdam, The Netherlands.
11.7.5 K-Ca
The K-Ca dating method uses the ⫺ decay branch of 40 K into 40 Ca (cf. figure
11.16). Calcium has six stable isotopes occurring in nature in the following abun-
dances: 40 Ca ⫽ 0.969823, 42Ca ⫽ 0.006421, 43Ca ⫽ 0.001334, 44Ca ⫽ 0.020567,
46
Ca ⫽ 0.000031, and 48Ca ⫽ 0.001824. The high relative abundance of 40 Ca (a
result of its mass number A, which is a multiple of 4 and thus exceptionally stable;
cf. section 11.3) is one of the two main problems encountered in this sort of dating
(in Ca-rich samples, the relative enrichment in 40 Ca resulting from 40 K decay is
low with respect to bulk abundance). The other problem is isotopic fractionation
of calcium during petrogenesis (and also during analysis; see for this purpose
Russell et al., 1978). These two problems prevent extensive application of the K-
Ca method, which requires extreme analytical precision. The isochron equation
involves normalization to the 42Ca abundance
40
Ca 40
Ca
=
42
Ca t 42
Ca 0
(11.107)
{ [( }
λβ K
)]
40
+ exp λ EC + λ β t − 1 .
λ EC + λ β
42
Ca
The ( 40 Ca/ 42Ca)t ratio to be introduced into equation 11.107 (hereafter, Rcorr ) was
derived from the measured ratio (Rmeas ) by application of the “exponential law”:
P
m
Rmeas,ij = Rcorr,ij i (11.108)
mj
with exponent
Rmeas, jk
ln
Rnorm, jk
P= , (11.109)
mj
ln
mk
where pedices i, j, and k identify nuclides of mass numbers 40, 42, and 44, respectively.
Rnorm,jk is the normalizing ratio ( 42Ca/ 44Ca) taken to be 0.31221 (Russell et al., 1978), and
mi , mj, mk are, respectively, the masses of the corresponding nuclides ( 40 Ca ⫽ 39.962591,
42
Ca ⫽ 41.958618, and 44Ca ⫽ 43.955480).
Isotope Geochemistry / 757
Figure 11.23 K-Ca isochron for granite batholith of Pikes Peak (Colorado). Aliquot-
spiked samples not included in age calculation. Reprinted from B. D. Marshall and D. J.
De Paolo, Geochimica et Cosmochimica Acta, 46, 2537–2545, copyright 䉷 1982, with kind
permission from Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington 0X5
1GB, UK.
Figure 11.23 shows the isochron obtained by Marshall and De Paolo (1982)
for the granite batholith of Pikes Peak (Colorado). The effectiveness of the
double-spike technique is evident, especially when we see that aliquot-spiked
samples do not fall on the best-fit interpolant (York’s algorithm; York, 1969).
The obtained age (1041 ⫾ 32 Ma) is consistent with that previously obtained with
Rb-Sr whole rock analyses (1008 ⫾ 13 Ma; see Marshall and De Paolo, 1982, for
references). The initial ratio ( 40 Ca/ 42Ca) 0 of 151.0 is identical, within the range
of uncertainty, to upper mantle values, indicating negligible contamination by old
crust components: the relative K/Ca abundance in the earth’s mantle is about
0.01, a value too low to alter the “primordial” ( 40 Ca/ 42Ca) 0 composition.
11.7.6 U-Th-Pb
Uranium has 15 isotopes, all unstable, with A-values from 227 to 240, and two
isomers at A ⫽ 235. Natural uranium is composed of 238U (99.2745%), 235U
(0.720%), and 234U (0.0055%). Thorium has 12 isotopes, all unstable, with A-
values from 223 to 234. Lead has 29 isotopes, four of them present in nature in
the following atomic proportions: 204Pb ⫽ 1.4%, 206Pb ⫽ 24.1%, 207Pb ⫽ 22.1%,
and 208Pb ⫽ 52.4%. The 4n, 4n ⫹ 2, and 4n ⫹ 3 decay series can be schematically
represented as follows:
758 / Methods
90 Th → 82 Pb +
232 208
6α + 4 β − λ 232 = 4.9475 × 10 −11 a −1 (11, 110.1)
238
92 U → 82 Pb +
206
8α + 6 β − λ 238 = 1.55125 × 10 −10 a −1 (11.110.2 )
235
92 U → 82 Pb +
207
7α + 4 β − λ 235 = 9.8485 × 10 −10 a −1 , (11.110.3)
where 232, 238, and 235 are the combined decay constants. Equation 11.83 ap-
plied to the three parent nuclides gives
208
Pb = 232
[ (
Th exp λ 232 t − 1 ) ] (11.111.1)
206
Pb = 238
[ (
U exp λ 238 t − 1 ) ] (11.111.2 )
207
[ (
Pb = 235 U exp λ 235 t − 1 . ) ] (11.111.3)
204
Normalization to the stable isotope Pb leads to the chronological equa-
tions
Pb Pb Th
[ ( ) ]
208 208 232
= + exp λ 232 t − 1 (11.112.1)
204
Pb t 204
Pb 0 204
Pb
Pb Pb U
[ ( ) ]
206 206 238
= + exp λ 238 t − 1 (11.112.2 )
204
Pb t 204
Pb 0 204
Pb
Pb Pb U
[ ( ) ]
207 207 235
= + exp λ 235 t − 1 . (11.112.3)
204
Pb t 204
Pb 0 204
Pb
Because of the different decay rates of 235U and 238U and their constant relative
proportion in nature ( 235U/ 238U ⫽ 0.00725), a fourth chronological equation can
be derived, based on the combination of equations 11.112.3 and 11.112.2:
207
Pb 207
Pb
∗ −
207
Pb 204
Pb t 204
Pb 0
=
206
Pb 206
Pb 206
Pb
−
204
Pb t 204
Pb 0 (11.113)
= 0.00725
[ exp (λ t ) − 1] .
235
[ exp (λ t ) − 1]
238
Isotope Geochemistry / 759
Note that, because equation 11.113 is indeterminate at the limit of t ⫽ 0, its solution
under boundary conditions involves application of De l’Hopital’s rule:
f (t ) f ′ (t )
lim = lim (11.114)
t→0
g( t ) t → 0
g ′(t )
lim 0.00725
[ exp(λ t ) − 1] = lim 0.00725 [λ
235 235 ( )]
exp λ 235t
t→0
[ exp(λ t ) − 1]
238
t→0
[λ 238 exp( λ t )]
238
(11.115)
λ 235
= 0.00725 .
λ 238
760 / Methods
Figure 11.24 (207Pb/ 306Pb)* radiogenic ratio of a U-bearing system as a function of age
t (Ga).
206
Pb 206
Pb
∗ −
Pb Pb t Pb 0
[ ( ) ]
206 204 204
Figure 11.25 Th-Pb and U-Pb isotopic compositions of whole rocks from Granite Moun-
tains Formation (Wyoming) based on data of Rosholt et al. (1973). 2.79 Ga isochrons
based on Pb-Pb age are superimposed, to highlight effect of differential elemental remobi-
lization.
Figure 11.26 Concordia diagram. (A) Interpretation of straight discordia path as a result
of a single episode of Pb loss. (B) Continuous diffusion model of Tilton (1960) applied to
world minerals of a t ⫽ 2800 Ma “common” age. Reproduced with modifications from
G. Faure (1986), Principles of Isotope Geology, 2nd edition, copyright 䉷 1986 by John
Wiley and Sons, by permission of John Wiley & Sons and from Tilton (1960) by permis-
sion of the American Geophysical Union.
762 / Methods
207
Pb 207
Pb
∗ −
Pb Pb t Pb 0
[ ( ) ]
207 204 204
Each age t therefore corresponds to a certain 206Pb/ 238U, 207Pb/ 235U couple, and
their time evolution gives rise to a curved path or “concordia curve” in the binary
field (figure 11.26A and B). The age interpretation of samples falling outside the
concordia line is controversial. In the simplest acceptance, straight-line arrange-
ments (or “discordia lines”) below the concordia curve are interpreted in terms
of Pb losses or U gains in the system (an eventual U loss from the system would
extend the straight discordia path above the concordia curve). Figure 11.26A
shows, for instance, the discordant compositions of three zircons from a gneiss
of Morton and Granite Falls (Southern Minnesota; Catanzaro, 1963): because
the partial loss of Pb does not significantly affect the 206Pb/ 207Pb ratio, the dis-
cordia line is straight and intersects the concordia curve at two points, t and t⬘,
which are interpreted, respectively, as the true age of the rock (t) and the age of
remobilization (t⬘). However, figure 11.26B shows that U-bearing minerals from
five continents with 206Pb/ 207Pb ages greater than 2300 Ma plot on a single dis-
cordia line, thus indicating an apparent common crystallization age of t ⫽ 2800
Ma, followed by a common Pb loss episode at t⬘ ⫽ 600 Ma (Tilton, 1960). How-
ever, as shown by Tilton (1960), the common straight-path arrangement of the
various minerals is a result of continuous diffusion of Pb from crystals. Assuming
that diffusion obeys Fick’s first law (cf. section 4.11) and affects only Pb, the
daughter-to-parent ratios of the various minerals (assimilated to spheres of radius
r with identical and constant diffusion coefficient D and uniform U distribution)
evolve along a curved trajectory that may be approximated by a straight line for
D/r2 ⬍ 50 ⫻ 10⫺12 a⫺1. Obviously, in this case, the age of rejuvenation t⬘ is only
apparent, although initial age t maintains the same significance as in the previous
model (see Faure, 1986 for alternative interpretations of the concordia diagram).
11.7.7 Re-Os
Rhenium has 12 isotopes, from A ⫽ 185 to A ⫽ 196, and two isomers at A ⫽ 186
and A ⫽ 190. Natural rhenium is composed of 185Re (0.37398) and 187Re
(0.62602). Osmium has 28 isotopes from A ⫽ 169 to A ⫽ 196, with isomers at
A ⫽ 181, 183, 190, 191, and 192. Natural osmium has seven isotopes, all stable:
184
Os (0.00024), 186Os (0.01600), 187Os (0.01510), 188Os (0.13286), 189Os
(0.16252), 190 Os (0.26369), and 192Os (0.40958) (fractional atomic abundances
after Luck and Allègre, 1983).
The 187Re isotope is transformed into 187Os by ⫺ decay:
Isotope Geochemistry / 763
76 Os + β
→ 187 + v + 0.0025 MeV .
187 −
75 Re (11.117 )
Os Os
[ ( ) ]
187 187 187
Re
= + exp λt − 1 . (11.118)
186
Os t 186
Os 0 186
Os
Note that the decay rate derived by Luck and Allègre (1983) from the isoch-
ron of figure 11.27, assuming a common age of 4550 Ma for all investigated
samples ( Re ⫽ 1.52 ⫾ 0.04 ⫻ 10⫺11 a⫺1 ), corresponds to a half-life of 4.56 ⫾
0.12 ⫻ 1010 a, which is somewhat higher than the direct determination of Lindner
et al. (1989) but consistent with the direct determination of Payne and Drever
(1965) [(4.7 ⫾ 0.5) ⫻ 1010 a].
Figure 11.28 shows the rhenium and osmium isotopic compositions of black
shales and sulfide ores from the Yukon Territory (Horan et al., 1994). The black
shale and sulfide layers are approximately isochronous. The superimposed refer-
ence isochrons bracket the depositional age of the enclosing shales. One reference
line represents the minimum age (367 Ma) with an initial ( 187Os/ 186Os) 0 ratio
of one, consistent with the mantle isotopic composition at that age (see later).
The other reference isochron is drawn for a maximum age of 380 Ma, with
( 187Os/ 186Os) 0 ⫽ 12 (the maximum value measured in terrigenous sediments).
Further examples of application of Re-Os dating of sediments can be found in
Ravizza and Turekian (1989).
764 / Methods
Figure 11.27 Re-Os isochron for iron meteorites and metallic phase of chondrites and
earth’s mantle. Reprinted with permission from J. M. Luck and C. J. Allègre, Nature, 302,
130–132, copyright 䉷 1983 Macmillan Magazines Limited.
Isotope Geochemistry / 765
Figure 11.28 Re-Os isochron for black shales and sulfide deposits of Yukon Territory
(Canada). Reprinted from M. F. Horan et al., Geochimica et Cosmochimica Acta, 58, 257–
265, copyright 䉷 1994, with kind permission from Elsevier Science Ltd., The Boulevard,
Langford Lane, Kidlington 0X5 1GB, UK.
14
11.7.8 C
Carbon has eight isotopes, from A ⫽ 9 to A ⫽ 16. Natural carbon is composed
of the two stable isotopes 12C (0.9889) and 13C (0.0111) and of radiogenic 14C,
continuously produced in the earth’s atmosphere mainly by (n,p) reaction of slow
(or “thermal”) neutrons with 14N:
7 N+ n → 14
6 C + p.
14
(11.119)
The 14C isotope reacts directly with atmospheric oxygen to produce 14CO2 and
14
CO, and undergoes isotopic exchanges with the preexisting 12CO2-12CO and
13
CO2-13CO molecules. Mixing of the various molecules is rapidly achieved in the
atmosphere (less than 2 years). Atmospheric carbon (1.4% of exchangeable car-
bon, mainly as CO2 ) equilibrates readily with oceanic carbon (H2CO3, HCO⫺ 3,
the various soluble carbonates, and very subordinate CO2⫺ 3 , amounting in total
766 / Methods
to 95% of exchangeable carbon) and with carbon in the terrestrial biosphere and
humus (3.2% of exchangeable carbon).
The decay of 14C to 14N takes place by ⫺ emission:
7 N+ β
→ 14 + v + 0.156 MeV .
14 −
6 C (11.120 )
The relatively rapid production rate of 14C (approximately 2s⫺1 cm⫺2 ) coupled
with high exchange kinetics (the mean time of residence in the atmosphere is
relatively short—i.e. 5 to 10 years) and a comparatively long half-life of the decay
process of equation 11.120 (5730 ⫾ 40 a; Godwin, 1962) ensure that a steady
state in the relative concentrations of the three carbon isotopes is quite rapidly
achieved. The steady state isotopic composition of carbon results in specific ra-
dioactivity of 14.9 dpm/g in the atmosphere and 13.6 dpm/g in the biosphere (the
lowering of specific radioactivity is a result of exchange kinetics among atmo-
spheric, oceanic, and biogenic reservoirs and of minor isotope fractionation
effects).
Living animals and plants acquire and maintain a constant level of 14C radio-
activity during their lifetimes as a result of continuous reequilibration with atmo-
sphere and biosphere. However, when they die, these continuous exchanges cease
and 14C radioactivity progressively decreases according to the exponential law
R t = R 0 exp( − λt ) , (11.121)
where R t is 14C radioactivity measured at time t after the death of the living organ-
ism, and R 0 is 14C radioactivity in the biosphere during its lifetime. Through some
substitutions, equation 11.121 can be rewritten as the chronological equation
R
t = 8.2666 × 10 3 ln 0 . (11.122)
Rt
Although the scope of this book does not allow an appropriate treatment of stable
isotope compositions of earth’s materials (excellent monographs on this subject
can be found in the literature—e.g., Hoefs, 1980; Faure, 1986), we must neverthe-
less introduce the significance of the various compositional parameters adopted
in the literature before presenting the principles behind stable isotope geother-
mometry.
Let us consider again a generic isotopic exchange as represented by a reaction
of the type
a A ° + bB • ⇔ aA • + bB ° (11.123)
• •
(Urey, 1947), where A°, A and B°, B are, respectively, molecules of the same type
differing only in the isotopic composition of a given constituting element (open
and full symbols: light and heavy), and a and b are stoichiometric coefficients.
The “fractionation factor” ␣ represents the relative distribution of heavy and
light isotopes in the two phases at equilibrium (somewhat similar to the normal-
ized distribution coefficient adopted in trace element geochemistry; cf. section
10.8):
Figure 11.29 Specific radioactivity of 14C expressed as per-mil deviation (⌬‰) from present-day radioactivity over the last 7500 years,
derived from comparisons with dendrochronological studies. Reprinted from Damon et al. (1978a), with permission, from The Annual
Review of Earth and Planetary Sciences, Volume 6, copyright 1978 by Annual Reviews Inc. and from Suess (1965), Journal of Geophysi-
cal Research, 70, 5937–5952, copyright 1965 by the American Geophysical Union.
Isotope Geochemistry / 769
X•
°
X A
α = , (11.124)
X•
°
X B
•
where (X /X °)A is the ratio of isotopic abundance in compound A. Assuming
random distribution of isotopes in the exchanging compounds, the fractionation
factor is related to equilibrium constant K (Epstein, 1959) by
α = K1 r , (11.125)
where r is the number of exchanging isotopes per formula unit. In the simplest
case, exchange reactions are written in such a way that only one isotope is ex-
changed. In this case, the fractionation factor obviously identifies with the equi-
librium constant (i.e., ␣ ⫽ K ).
1 1 1 1
MgSi 16O3+ Mg2Si 18O4 ⇔ MgSi 18O3+ Mg2Si 16O4 .
3 4 3 4 (11.126)
opx ol opx ol
[ O] [ O]
1 3 1 4
18 16
opx ol
K 126 = , (11.127)
[ O] [ O]
1 3 1 4
16 18
opx ol
[ O]
18
opx
= X 183
O, opx
≡ X•
n
(11.128.1)
[ O]
18
ol
= X 418
O, o l
≡ X•
m
(11.128.2)
[ O]
n
16
= X 163 ≡ X° (11.128.3)
opx O, opx
[ O]
m
16
= X 164 ≡ X° . (11.128.4)
ol O, o l
770 / Methods
r = a × n = b × m = 1, ( 11.129 )
and hence
α = K 126 . ( 11.130 )
(
X• X° )
∆ AB = • A
(
− 1 × 10 3 = α − 1 × 10 3. )
(
X X° ) B
(11.131)
δA
(
X• X°
= •
) A
− 1 × 10 3 (11.132.1)
(
X X° ) st
and
δB = •
(
X• X° ) B
− 1 × 10 3. (11.132.2)
(
X X° ) st
δA 10 3 + 1 δA + 10 3
α AB = = (11.133)
δ B 10 3 + 1 δ B + 10 3
and
δ A + 10 3
∆ AB = − 1 × 10 3 . (11.134)
δ B + 10 3
Because
X
10 3 ln 1 + ≅ X, (11.135)
1000
δ A − δ B ≅ ∆ AB ≅ 10 3 ln α AB . (11.136 )
The closeness of these approximations depend on the modulus of the first term
|␦A ⫺ ␦B|, as shown in table 11.8.
11.8.1 Oxygen
Oxygen is present in nature with three stable isotopes: 16O (0.997630), 17O
(0.000375), and 18O (0.001995). The first application of oxygen isotopic fraction-
ation studies to geological problems was suggested by Urey (1947), who related
the temperature of precipitation of calcium carbonate in seawater to the (T-
dependent) isotopic fractionation induced by the exchange reaction between H2O
molecules and carbonate ion CO2⫺ 3 :
C 16 O 23 − + 3H18
2 O ⇔ C O 3 + 3H 2 O.
18 2 − 16
(11.137)
772 / Methods
Figure 11.30 Reduced partition function for various minerals calulated by Kieffer (1982)
through equation 11.61 plotted against T ⫺2. Heavy curve labeled H2O(l) is reduced parti-
tion function of water according to Becker (1971). Dashed curve is a T ⫺2 extrapolation of
high-T reduced partition curve for quartz. Mineral abbreviations: Qtz (quartz), Calc
(calcite), Albt (albite), Musc (muscovite), Enst (clinoenstatite), Anor (anorthite), Diop
(diopside), Pyrp (pyrope), Gros (grossular), Zron (zircon), Fors (forsterite), Andr (andra-
dite), Rutl (rutile). Reprinted with permission from Kieffer (1982), Review of Geophysics
and Space Physics, 20, 827–849, copyright 1982 by the American Geophysical Union.
Phase a b c
Calcite 11.781 ⫺0.420 0.0158
Quartz 12.116 ⫺0.370 0.0123
Albite 11.134 ⫺0.326 0.0104
Anorthite 9.993 ⫺0.271 0.0082
Diopside 9.237 ⫺0.199 0.0053
Forsterite 8.326 ⫺0.142 0.0032
Magnetite 5.674 ⫺0.038 0.0003
peratures (Urey, 1947; Bigeleisen and Mayer, 1947; see also section 11.6.1). In-
deed, a cubic function in T ⫺2 fits the calculated reduced partition functions within
⫾0.02 per mil, for all temperatures above 400 K.
The Kieffer model correctly predicts the systematic change of the reduced
partition functions of various minerals with structure, as indicated by Taylor and
Epstein (1962). For anhydrous silicates, the decrease in the sequence framework-
chain-orthosilicate reflects the decreasing frequency of antisymmetric Si-O
stretching modes. The internal frequencies of the carbonate ion give a high re-
duced partition function at all T. The value for rutile is low because of the low
frequencies of the Ti-O modes (Kieffer, 1982).
Combining the results of Kieffer’s model and of laboratory experiments,
Clayton and Kieffer (1991) obtained a set of third-order equations relating the
reduced partition functions of various minerals to the inverse of the squared abso-
lute temperature (table 11.9) according to
1000 Q •
fA = ln = ax + bx + cx
2 3
(11.138)
r Q° A
x = 10 6 × T −2 . (11.139 )
∆ AB = fA − f B . (11.140)
Table 11.10 Comparison of polynomial and linear fits for mineral pairs
at 1000 K. Values in italics refer to equation 11.141 (from Clayton and
Kieffer, 1991).
Qz Ab An Di Fo Mt
Ca ⫺0.38 0.56 1.65 2.33 3.19 5.74
⫺0.38 0.56 1.61 2.37 3.29 5.91
Qz 0.94 2.03 2.72 3.57 6.12
0.94 1.99 2.75 3.67 6.29
Ab 1.09 1.78 2.63 5.18
1.05 1.81 2.73 5.35
An 0.69 1.54 4.09
0.76 1.68 4.30
Di 0.86 3.41
0.92 3.54
Fo 2.55
2.62
10 6
∆ AB = a AB × . (11.141)
T2
A comparison of the two methods for T ⫽ 1000 K is given in table 11.10. For
each entry, a positive value indicates 18O enrichment in the phase on the left. The
upper entry is relative to equations 11.138 to 11.140 and the lower entry to equa-
tion 11.141. Note that the table implicitly gives the parameters of the equation of
Clayton et al. (1989), because at 1000 K equation 11.141 reduces to ⌬AB ⫽ aAB.
Figure 11.31 shows the results of equations 11.138 to 11.141 applied to the
calcite-diopside couple compared with experimental evidence and with the indi-
cations of Kieffer’s (1982) model.
Reduced partition functions for oxides were calculated by Zheng (1991) on
the basis of the “modified increment method.”
This method derives from the observation that the degree of 18O enrichment in a set
of cogenetic silicate minerals (expressed as oxygen isotope index I-18O) can be related
to bond strengths in minerals (Taylor, 1968). The oxygen isotope index for phase A is
defined by
(M M ) *
3 2
n ct
( ) ∑Z
16 18
I− 18
O = × i ct′ − O , (11.142)
2 n (M M )
A 3 2 ct
1
O 16 18 A
where the asterisk denotes a reference phase. M16 and M18 are the formula weights of
isotopically light and heavy compounds, nO and nct are the stoichiometric numbers of
oxygens and cations per formula unit. Zct is the cationic charge, and the summation ex-
tends over all cations in the formula unit, i⬘ct-O is the “normalized 18O-increment” of a
cation-oxygen bond (Zheng, 1991):
Isotope Geochemistry / 775
Figure 11.31 Oxygen isotopic fractionation between calcite and diopside as obtained
from equations 11.131 to 11.133. Experimental data points are from Chiba et al. (1989).
Dotted line: indications from vibrational calculations following Kieffer’s (1982) procedure.
From Clayton and Kieffer (1991), reprinted with kind permission of The Geochemical
Society, Pennsylvania State University, University Park, Pennsylvania.
( )
q
i ct′ − O = i ct − O i Si − O . ( 11.143 )
Z ct
Sct − O = . (11.146)
v ct Rct − O
Although equation 11.146 agrees with the principle of electrostatic valence of Pauling
(1929), normalization to the cation-to-oxygen distance Rct-O to account for bond variation
776 / Methods
with the degree of ionicity is not in line with commonly adopted generalizations (see sec-
tion 1.10.3).
The term Wct-O is related to the masses of cations and oxygen isotopes:
Wct − O =
(m ct )
+ m 18O m 18O
.
(m )m
(11.147)
ct + m 18O 16O
The reduced partition function ratio of the mineral is derived from the equations above
through
1000 Q •
fA = (
ln = I −18O ) × exp
( (
∆E 1 − I−18 O ) )
A
r Q° A A RT
(11.148)
1000 Q •
× ln .
r Q° *
∆ AB = a × 10 6 T −2 + b × 10 3 T −1 + c. (11.149)
The T-dependency selected by Zheng (1991, 1993a,b) differs from the equation of
Clayton and Kieffer (1991), and the two methods do not give concordant results,
especially at low T, as shown in figure 11.32.
Zheng (1993b) also extended the modified increment method to the evalua-
tion of reduced partition function ratios in hydroxyl-bearing silicate minerals
through
2nO − nOH 18
( I− O )
18
OH− A
=
2 nO
I− O ,
A
( ) (11.150)
where nO and nOH represent the numbers of oxygens and hydroxyl groups in the
hydrous mineral and (I⫺18O)A is the isotope index in the anhydrous counterpart.
Although largely empirical, the method of Zheng (1991, 1993a,b) has the
Isotope Geochemistry / 777
Table 11.11 Calculation of 18O increments for silicates (from Zheng, 1993a).
11.8.2 Hydrogen
Hydrogen has two stable isotopes, 11H (0.99985) and 21H (0.00015), and a short-
lived radioactive isotope 31H (tritium) produced in the atmosphere by interaction
of 14N with cosmic ray neutrons:
14
7 N + n → 13 H + 12
6 C. (11.151)
Isotope 21H (deuterium), discovered by Urey et al. (1932), is usually denoted by sym-
bol D. The large relative mass difference between H and D induces significant fraction-
ation ascribable to equilibrium, kinetic, and diffusional effects. The main difference in the
calculation of equilibrium isotopic fractionation effects in hydrogen molecules with re-
spect to oxygen arises from the fact that the rotational partition function of hydrogen is
nonclassical. Rotational contributions to the isotopic fractionation do not cancel out at
high T, as in the classical approximation, and must be accounted for in the estimates of
the partition function ratio f .
The classical expression for Qrot is
* = hcB ប2
Qrot = , (11.152)
kT 4 I kT
where B is the rotational constant in cm⫺1, I is the moment of inertia, and the remaining
symbols have their usual meanings. Bigeleisen and Mayer (1947), to account for the non-
classical behavior of hydrogen molecules, suggested expansion of the deviation from the
classical rotational partition function in the polynomial form
Qrot σ σ2 σ3
= 1 + + + + L ,
∗
Qrot 3 15 315 (11.153)
where ⫽ 1/Q*rot.
Because the partition function ratio f is defined in such a way that the classical rota-
tional and translational contributions are canceled, equations 11.40, 11.41 and 11.43 must
be modified by introducing the ratio of the deviations from classical rotational behavior
of heavy and light hydrogen molecules. For small values of , this contribution reduces to
780 / Methods
Mineral a b c a b c a b c
Quartz 4.48 ⫺4.77 1.71 ⫺0.47 0.10 0
Stishovite 0.22 1.88 ⫺0.78 4.26 ⫺6.64 2.08 ⫺0.25 1.98 ⫺0.78
K-feldspar 0.16 1.50 ⫺0.62 4.32 ⫺6.27 2.00 ⫺0.30 1.60 ⫺0.62
Albite 0.15 1.39 ⫺0.57 4.33 ⫺6.15 1.98 ⫺0.32 1.49 ⫺0.57
Anorthite 0.36 2.73 ⫺1.14 4.12 ⫺7.50 2.24 ⫺0.11 2.83 ⫺1.14
Leucite 0.22 1.91 ⫺0.79 4.26 ⫺6.67 2.08 ⫺0.24 2.01 ⫺0.79
Nepheline 0.37 2.77 ⫺1.15 4.11 ⫺7.53 2.24 ⫺0.10 2.89 ⫺1.15
Diopside 0.56 3.67 ⫺1.53 3.92 ⫺8.43 2.40 0.10 3.78 ⫺1.53
Enstatite 0.51 3.45 ⫺1.44 3.97 ⫺8.22 2.37 0.05 3.55 ⫺1.44
Hedenbergite 0.55 3.61 ⫺1.51 3.93 ⫺8.37 2.40 0.08 3.71 ⫺1.51
Ferrosilite 0.48 3.33 ⫺1.39 3.99 ⫺8.09 2.35 0.02 3.43 ⫺1.39
Jadeite 0.31 2.42 ⫺1.01 4.17 ⫺7.19 2.18 ⫺0.16 2.52 ⫺1.01
Acmite 0.30 2.38 ⫺0.99 4.18 ⫺7.14 2.17 ⫺0.17 2.45 ⫺0.99
Wollastonite 0.67 4.11 ⫺1.72 3.81 ⫺8.87 2.49 0.21 4.21 ⫺1.72
Rhodonite 0.63 3.94 ⫺1.65 3.85 ⫺8.71 2.46 0.16 4.04 ⫺1.65
Beryl 0.50 3.40 ⫺1.42 3.98 ⫺8.16 2.34 ⫺0.03 3.50 ⫺1.42
Cordierite 0.38 2.85 ⫺1.19 4.10 ⫺7.62 2.26 ⫺0.08 2.95 ⫺1.19
Almandine 0.72 4.26 ⫺1.79 3.76 ⫺9.02 2.52 0.25 4.36 ⫺1.79
Pyrope 0.73 4.30 ⫺1.80 3.75 ⫺9.07 2.52 0.26 4.40 ⫺1.80
Spessartine 0.73 4.31 ⫺1.81 3.75 ⫺9.07 2.52 0.26 4.41 ⫺1.81
Grossular 0.74 4.35 ⫺1.82 3.74 ⫺9.11 2.52 0.27 4.45 ⫺1.82
Andradite 0.72 4.29 ⫺1.80 3.76 ⫺9.05 2.52 0.26 4.38 ⫺1.80
Uvarovite 0.72 4.28 ⫺1.80 3.76 ⫺9.05 2.52 0.26 4.39 ⫺1.80
Titanite 0.67 4.11 ⫺1.72 3.81 ⫺8.87 2.49 0.21 4.21 ⫺1.72
Malayaite 0.60 3.83 ⫺1.60 3.88 ⫺8.60 2.43 0.14 3.94 ⫺1.60
Zircon 0.72 4.26 ⫺1.79 3.76 ⫺9.03 2.52 0.25 4.37 ⫺1.79
Thorite 0.79 4.50 ⫺1.89 3.69 ⫺9.27 2.55 0.32 4.61 ⫺1.89
Fayalite 0.84 4.69 ⫺1.97 3.64 ⫺9.46 2.59 0.38 4.79 ⫺1.97
Forsterite 0.93 4.95 ⫺2.09 3.55 ⫺9.72 2.64 0.46 5.05 ⫺2.09
Tephroite 0.86 4.75 ⫺2.00 3.62 ⫺9.51 2.60 0.39 4.85 ⫺2.00
Willemite 0.69 4.18 ⫺1.75 3.79 ⫺8.94 2.50 0.23 4.28 ⫺1.75
Phenacite 1.06 5.34 ⫺2.26 3.42 ⫺10.11 2.70 0.60 5.45 ⫺2.26
Sillimanite 0.22 1.89 ⫺0.78 4.26 ⫺6.65 2.07 ⫺0.25 1.99 ⫺0.78
Andalusite 0.28 2.26 ⫺0.94 4.20 ⫺7.02 2.15 ⫺0.19 2.36 ⫺0.94
Kyanite 0.32 2.52 ⫺1.05 4.16 ⫺7.29 2.20 ⫺0.14 2.63 ⫺1.05
Muscovite 0.38 2.84 ⫺1.18 4.10 ⫺7.61 2.25 ⫺0.08 2.97 ⫺1.18
Paragonite 0.37 2.77 ⫺1.15 4.11 ⫺7.54 2.24 ⫺0.10 2.87 ⫺1.15
Margarite 0.38 2.84 ⫺1.18 4.10 ⫺7.61 2.25 ⫺0.08 2.94 ⫺1.18
Illite 0.34 2.60 ⫺1.08 4.14 ⫺7.36 2.21 ⫺0.13 2.70 ⫺1.08
Phengite 0.35 2.64 ⫺1.10 4.13 ⫺7.41 2.22 ⫺0.12 2.74 ⫺1.10
Isotope Geochemistry / 781
Mineral a b c a b c a b c
Glauconite 0.49 3.34 ⫺1.39 3.99 ⫺8.11 2.34 0.02 3.44 ⫺1.39
Phlogopite 0.62 3.92 ⫺1.64 3.86 ⫺8.68 2.45 0.16 4.02 ⫺1.64
Annite 0.59 3.79 ⫺1.59 3.89 ⫺8.56 2.43 0.13 3.89 ⫺1.59
Biotite 0.64 3.99 ⫺1.67 3.84 ⫺8.76 2.46 0.18 4.10 ⫺1.67
Lepidolite 0.38 2.81 ⫺1.17 4.10 ⫺7.58 2.25 ⫺0.09 2.91 ⫺1.17
Hornblende 0.59 3.80 ⫺1.59 3.89 ⫺8.56 2.43 0.13 3.90 ⫺1.59
Tremolite 0.53 3.52 ⫺1.47 3.95 ⫺8.28 2.38 0.06 3.62 ⫺1.47
Actinolite 0.52 3.48 ⫺1.45 3.96 ⫺8.25 2.37 0.05 3.58 ⫺1.45
Cummingtonite 0.52 3.48 ⫺1.45 3.96 ⫺8.25 2.37 0.05 3.58 ⫺1.45
Grunerite 0.50 3.40 ⫺1.42 3.98 ⫺8.17 2.36 0.03 3.50 ⫺1.42
Glaucophane 0.45 3.18 ⫺1.32 4.03 ⫺7.94 2.31 ⫺0.02 3.28 ⫺1.32
Riebeckite 0.43 3.08 ⫺1.28 4.05 ⫺7.85 2.30 ⫺0.04 3.18 ⫺1.28
Pargasite 0.71 4.22 ⫺1.77 3.77 ⫺8.99 2.51 0.24 4.32 ⫺1.77
Anthophyllite 0.52 3.49 ⫺1.46 3.96 ⫺8.26 2.38 0.06 3.60 ⫺1.46
Gedrite 0.63 3.96 ⫺1.66 3.85 ⫺8.72 2.46 0.17 4.06 ⫺1.66
Kaolinite 0.19 1.68 ⫺0.69 4.29 ⫺6.44 2.03 ⫺0.28 1.78 ⫺0.69
Lizardite 0.51 3.42 ⫺1.43 3.97 ⫺8.19 2.36 0.04 3.52 ⫺1.43
Amesite 0.53 3.53 ⫺1.47 3.95 ⫺8.30 2.38 0.06 3.63 ⫺1.47
Pyrophyllite 0.08 0.86 ⫺0.35 4.40 ⫺5.62 1.87 ⫺0.38 0.96 ⫺0.35
Talc 0.28 2.27 ⫺0.94 4.20 ⫺7.04 2.14 ⫺0.19 2.38 ⫺0.94
Serpentine 0.49 3.35 ⫺1.40 3.99 ⫺8.12 2.35 0.02 3.45 ⫺1.40
Pennine 0.51 3.43 ⫺1.43 3.97 ⫺8.19 2.36 0.04 3.53 ⫺1.43
Clinochlore 0.51 3.43 ⫺1.43 3.97 ⫺8.19 2.36 0.04 3.53 ⫺1.43
Chamosite 0.42 3.04 ⫺1.27 4.06 ⫺7.81 2.30 ⫺0.04 3.14 ⫺1.27
Thuringite 0.58 3.73 ⫺1.56 3.90 ⫺8.50 2.42 0.11 3.84 ⫺1.56
Zoisite 0.43 3.07 ⫺1.28 4.05 ⫺7.84 2.30 ⫺0.04 3.18 ⫺1.28
Epidote 0.42 3.05 ⫺1.27 4.05 ⫺7.81 2.29 ⫺0.04 3.15 ⫺1.27
Vesuvianite 0.93 4.97 ⫺2.09 3.55 ⫺9.74 2.63 0.47 5.07 ⫺2.09
Ilvaite 1.05 5.32 ⫺2.24 3.43 ⫺10.08 2.69 0.58 5.42 ⫺2.24
Norbergite 0.86 4.75 ⫺2.00 3.62 ⫺9.51 2.60 0.39 4.85 ⫺2.00
Chondrodite 0.74 4.34 ⫺1.82 3.73 ⫺9.11 2.52 0.27 4.45 ⫺1.82
Humite 0.71 4.24 ⫺1.78 3.77 ⫺9.01 2.51 0.24 4.34 ⫺1.78
Clinohumite 0.70 4.19 ⫺1.75 3.78 ⫺8.95 2.50 0.23 4.29 ⫺1.75
Staurolite 0.39 2.90 ⫺1.21 4.09 ⫺7.66 2.27 ⫺0.07 3.00 ⫺1.21
Topaz 0.30 2.41 ⫺1.00 4.18 ⫺7.18 2.17 ⫺0.16 2.52 ⫺1.00
Datolite 0.15 1.44 ⫺0.60 4.33 ⫺6.21 1.99 ⫺0.31 1.54 ⫺0.60
Chloritoid 0.49 3.35 ⫺1.40 3.99 ⫺8.11 2.35 0.02 3.45 ⫺1.40
Tourmaline 0.27 2.22 ⫺0.92 4.21 ⫺6.99 2.14 ⫺0.20 2.32 ⫺0.92
Axinite 0.28 2.26 ⫺0.94 4.20 ⫺7.02 2.15 ⫺0.19 2.36 ⫺0.94
Pectolite 0.41 2.95 ⫺1.23 4.08 ⫺7.71 2.28 ⫺0.06 3.05 ⫺1.23
Prehnite 0.30 2.40 ⫺1.00 4.18 ⫺7.17 2.18 0.17 2.50 ⫺1.00
Rutile 0.79 6.82 ⫺3.59 3.45 ⫺10.60 2.55
Pyrolusite 0.64 6.16 ⫺3.23 3.61 ⫺9.94 2.19
Cassiterite 0.56 5.80 ⫺3.04 3.68 ⫺9.58 2.00
782 / Methods
Mineral a b c a b c a b c
Plattnerite 0.56 5.80 ⫺3.04 3.68 ⫺9.58 2.00
Cerianite 1.67 9.32 ⫺4.97 2.57 ⫺13.09 3.93
Thorianite 1.66 9.30 ⫺4.96 2.58 ⫺13.07 3.92
Uraninite 1.61 9.19 ⫺4.90 2.63 ⫺12.97 3.86
Ilmaite 1.36 8.61 ⫺4.57 2.88 ⫺12.38 3.53
Geikelite 1.46 8.85 ⫺4.70 2.78 ⫺12.63 3.66
Hematite 1.55 9.05 ⫺4.82 2.69 ⫺12.82 3.78
Corundum 2.00 9.94 ⫺5.32 2.24 ⫺13.71 4.28
Magnetite 1.22 8.22 ⫺4.35 3.02 ⫺12.00 3.31
Jacobsite 1.23 8.24 ⫺4.37 3.02 ⫺12.02 3.33
Magnesioferrite 1.29 8.43 ⫺4.47 2.95 ⫺12.20 3.43
Ulvöspinel 1.43 8.78 ⫺4.67 2.81 ⫺12.56 3.63
Chromite 1.63 9.22 ⫺4.91 2.62 ⫺12.99 3.87
Magnesiochromite 1.76 9.49 ⫺5.07 2.48 ⫺13.27 4.03
Hercynite 1.92 9.80 ⫺5.24 2.32 ⫺13.58 4.20
Spinel 2.06 10.04 ⫺5.38 2.18 ⫺13.81 4.34
( )
• 2
•
σ −σ° σ −σ° σ
•2
−σ°
2
1+ + + . (11.154)
3 18 90
Qrot =
π
12
1 +
[
2s x + 2s y + 2s z − (s x s y s z ) − (s y s z s x ) − (s z s x s y )
+ L,
]
sxsy sz 12
( 11.155)
∗ = ប2
s x = Qrot, x , (11.156)
4 I x kT
1 1
H 2O+ D ⇔ D 2O+ H (11.157)
2 2
Isotope Geochemistry / 783
1 1
CH 4 + D ⇔ CD 4 + H (11.158)
4 4
1 1 1 1
H 2 O + CD 4 ⇔ D 2 O + CH 4 . (11.159)
2 4 2 4
Results of his calculations are listed in table 11.13 in terms of fractionation factors
␣ and ⌬A* B. The ⌬A*B factors for methane–water vapor and hydrogen–water vapor
couples are plotted in figure 11.33 as a function of T (°C).
Note that Bottinga (1969a) defines ⌬AB * as
(see eq. 11.133 for the significance of ␣ ). Taking the water vapor as phase B, the
parameter ⌬AB * is equal to the deuterium enrichment in phase A with respect to
water, and hence is equivalent to the ␦ notation adopted in the geochemical litera-
ture, under the assumption of equilibrium.
The hydrogen isotope fractionation between OH-bearing minerals and water
is a sensitive function of T. Figure 11.34 shows the experimental results of Suzu-
oki and Epstein (1976). The interpolant expressions for muscovite-water, biotite-
water, and hornblende-water in the T range 450 to 850 °C are
22.1 × 10 6
10 3 ln α muscovite− water = − + 19.1 (11.161.1)
T2
21.3 × 10 6
10 3 ln α biotite− water = − − 2.8 (11.161.2)
T2
23.9 × 10 6
10 3 ln α hornblende− water = − + 7.9 (11.161.3)
T2
Because the slope coefficient in the equations above is practically constant, within
experimental uncertainty, the T-dependency of the hydrogen fractionation factor
for mica and amphibole can be generalized as
22.4 × 10 6
10 3 ln α mineral − water = −
T2 (11.162)
+ 28.2 + 2 X Al − 4 X Mg − 68 X Fe
(Suzuoki and Epstein, 1976), where XAl, XMg, and XFe are molar fractions of
sixfold coordinated cations. Equation 1.162 implies that hydrogen isotope frac-
tionation among coexisting mica and amphibole is temperature independent and
784 / Methods
ln ␣ 103 ln ␣ ⌬AB
*
Figure 11.33 Calculated oxygen isotope fractionation factors ⌬ *AB for methane–water va-
por and hydrogen–water vapor couples plotted against T (°C). Reprinted from Y. Bottinga,
Geochimica et Cosmochimica Acta, 33, 49–64, copyright 䉷 1969, with kind permission
from Elsevier Science Ltd., The Boulevard, Langford Lane, Kidlington 0X5 1GB, UK.
kbar water pressure in the T range 100 to 500 °C lead to the following polyno-
mial expansion:
27.5 × 10 6 7.69 × 10 4
10 3 ln α serpentine− water = − + 40.8. (11.163)
T2 T
Note that the T-dependency in equation 11.163 is inverse with respect to equation
11.162 (see, however, Wenner and Taylor, 1971).
Concerning clay minerals (illite and smectite), Capuano (1992) proposed a
simple linear dependency of the mineral-water fractionation factor over recipro-
cal temperature:
45.3 × 10 3
10 3 ln α clay − water = − + 94.7. (11.164)
T
Equation 11.164 is valid in the T range 0 to 150 °C. The same sort of T-
dependency (although less marked) was envisaged by Yeh (1980) (slope coeffi-
cient ⫽ ⫺19.6 ⫻ 103, intercept ⫽ 25).
One of the main applications of hydrogen and oxygen isotope thermometry in
geochemistry is the estimation of the reservoir temperatures of active geothermal
systems or the evaluation of the ruling T conditions during deposition or alter-
Figure 11.34 Fractionation factors versus T ⫺2 for mica-water and amphibole-water
couples. Reprinted from T. Suzuoki and S. Epstein, Geochimica et Cosmochimica Acta,
40, 1229–1240, copyright 䉷 1976, with kind permission from Elsevier Science Ltd., The
Boulevard, Langford Lane, Kidlington 0X5 1GB, UK.
Isotope Geochemistry / 787
γ HDO
ΓD = (11.165.1)
γ H2 O
γ 18
H2 O
ΓO = . (11.165.2)
γ H 2O
R activity 1 + 10 −3 δ activity
Γ = = (11.166)
R composition 1 + 10 −3 δ composition
(Horita et al., 1993a,b), where ␦ is the conventional delta value (‰). Combining
equations 11.165 and 11.166, through some substitutions, we can derive
Both the oxygen and hydrogen isotope salt effects were determined by Horita
et al. (1993a) in single-salt (NaCl, KCl, MgCl2 , CaCl2 , Na2SO4 , and MgSO4 )
aqueous solutions of various molalities, and their results can be represented in
the simple linear form
b
10 3 ln Γ = m a + , (11.168)
T
where m is molality, and a and b are fitting parameters whose values are listed in
table 11.14.
788 / Methods
11.8.3 Carbon
Fractionation factors for the two stable carbon isotopes 12C and 13C have been
calculated by Bottinga (1969a,b) for exchanges among graphite, calcite, carbon
dioxide, and methane and for graphite-CO2 and diamond-CO2 equilibria. Results
of vibrational calculations of Bottinga (1969a,b) in the T range 0 to 1000 °C are
listed in table 11.15 and graphically displayed in figure 11.36.
The main points arising from the study of Bottinga (1969a) are as follows:
Isotope Geochemistry / 789
1. The carbon isotope fractionation between CO2 and methane and between
graphite and methane is a sensitive function of temperature.
2. The isotope composition of graphite does not characterize its origin be-
cause it is not unique to any simple method of condensation.
3. The carbon fractionation factors between gaseous CO2 and CH4 are
rather large, especially at low T. Temperatures deduced by application of
13
C fractionation geothermometers are, however, reasonable for either
CO2 or CH4 produced by bacterial decomposition of organic matter and
present in gaseous emissions associated with geothermal fields (cf. tables
9 and 10 in Bottinga, 1969a).
790 / Methods
Table 11.15 Calculated fractionation factors (1000 ln ␣) for 13C exchange among carbon dioxide,
calcite, graphite, diamond, and methane (Bottinga 1969a,b). Cc ⫽ calcite; Gr ⫽ graphite;
Dm ⫽ diamond.
1000 ln ␣AB
Moreover, as we can see in figure 11.36, the fractionation factor between CO2
and graphite is practically constant below 200 °C (1000 ln ␣ 艐 14.5) and falls
progressively with increasing T. Also, the carbon fractionation factor between
CO2 and calcite has a maximum near 460 °C (1000 ln ␣ 艐 2.85) and a crossover
at 192 °C, confirmed by direct observations of natural occurrences in geothermal
areas by Craig (1963).
According to the calculations of Bottinga (1969b) for the C-O system, the
high-density form diamond should be preferentially enriched in 13C with respect
to the low-density form graphite (although the enrichment is quite limited; cf.
figure 11.36). Moreover, the CO2-diamond carbon fractionation curve exhibits a
broad maximum near 300 °C (1000 ln ␣ 艐 11).
ω max
∫ g( ω ) dω = 3 Nr (11.169)
0
kT gn
X n+1
exp( − X 2 )
ln Qvib = ∑ × ∫ ln dX ,
1 − exp( − X )
(11.170)
hc n ω n +1 − ω n Xn
where the various gn (i.e., the eigenvalues of the dynamic matrix) are the numbers of modes
in the vibrational frequency interval n⫹1 ⫺ n (see sections 3.1, 3.3, and 11.6.1 for the
significance of the remaining symbols).
Vibrational spectra generated by Bottinga (1969b) for graphite and diamond with this
procedure are shown in figure 11.37. The spectrum for graphite is based on the interionic
potential model of Yoshimori and Kitano (1956) and adopts the force constants of Young
and Koppel (1965) (assumed to be identical for 12C and 13C). The histogram of figure
11.37A represents 1.79 ⫻ 106 frequencies in the irreducible segment ( 24 1
) of the Brillouin
zone, with a frequency interval n⫹1 ⫺ n of 1 cm⫺1. Figure 11.37B shows the normalized
phonon spectra for diamond, based on the interionic potential model of Aggarwal (1967)
and the dynamic matrix equations of Herman (1959) ( 48 1
of the Brillouin zone; 3.46 ⫻ 106
frequencies). The corresponding Debye spectrum for D ⫽ 1860 K is also shown for com-
parative purposes. The jagged appearance of the phonon spectra is partly attributable to
incomplete root sampling of phonon frequencies and to the narrow bandwidth adopted
in the two histograms (Bottinga, 1969b).
Isotope Geochemistry / 793
Figure 11.37 Normalized phonon spectra for 12C and 13C in graphite (A) and diamond
(B). Reprinted from Bottinga (1969b), with kind permission from Elsevier Science Pub-
lishers B.V, Amsterdam, The Netherlands.
6.3 × 10 3
1000 ln α H * = − 0.91 (11.171.1)
2 CO 3 − CO2 ( g ) T2
1.099 × 10 6
1000 ln α HCO− − CO = − 4.54 (11.171.2 )
3 ( )
2 g T2
8.7 × 10 5
1000 ln α CO2 − − CO = − 3.4 . (11.171.3)
3 ( )
2 g T2
11.8.4 Sulfur
Sulfur has four stable isotopes: 32S (0.9502), 33S (0.0075), 34S (0.0421), and 36S
(0.0002). The 34S/ 32S sulfur isotope fractionation in sulfide minerals and solute
sulfur species is a sensitive function of temperature (Sakai, 1968; Ohmoto, 1972).
Detailed vibrational calculations of the partition function ratio and correspond-
ing isotopic fractionation factors for solute sulfur species were first attempted by
Sakai (1968). The results of his calculations are shown in figure 11.39 in terms of
the ⌬34S enrichment factor:
∆ AB ≅ 1000 ln a AB ≅ δ 34 S A − δ 34 S B (11.172)
Figure 11.38 Relationship between average 13C enrichment factor and pH in solute car-
bonates–gaseous CO2 equilibria (⌬DIC⫺CO2 ⫽ 13␦DIC⫺CO2 ). Reprinted from C. S. Romanek,
E. L. Grossman, and J. W. Morse, Geochimica et Cosmochimica Acta, 56, 419–430, copy-
right 䉷 1992, with kind permission from Elsevier Science Ltd., The Boulevard, Langford
Lane, Kidlington 0X5 1GB, UK.
the complex (i.e., strong positive relative enrichment factors for sulfate ion
SO2⫺ 2⫺
4 with respect to H2S and negative enrichment factors for sulfide ion S ; see
later in this section). Moreover, it is also evident that a simple inverse relation
with the squared temperature holds.
Because the fractionation effect between condensed sulfates and SO2⫺ 4 ions is
negligible, it can be assumed that all sulfate complexes in solution (e.g.,
HSO⫺ ⫺ ⫺
4 , KSO4 , NaSO4 , . . .) have similar ⌬ S enrichment factors, T being equal
34
(Ohmoto, 1972).
Sulfur isotope relative fractionation factors between various minerals and py-
rite are shown in figure 11.40. As we can see, the isotopic compositions of sulfates
reflect the relative fractionation effects induced by the SO2⫺ 4 groups (compare
figures 11.40 and 11.39).
Ohmoto and Rye (1979), through a critical examination of all the existing
data, proposed a set of fractionation equations relating the fractionation factors
between important ore-forming sulfides and H2S(aq) to T of formation. The pro-
posed equations obey a simple linear dependence over the inverse squared tem-
perature (K), passing through the origin:
A
10 3 ln α i − H2 S(aq) = . (11.173)
T2
Figure 11.39 ⌬34S enrichment factors for solute sulfur species based on the data of Sakai
(1968). Reproduced with modifications from Ohmoto, Economic Geology, 1972, Vol. 67,
p. 533.
A
T (K ) = . (11.174)
10 ln α i − H2 S aq
3
( )
The slope coefficients A for the various sulfides are listed in table 11.16.
Coupling thermometric equations such as equation 11.174 for different min-
erals, one obtains solid-solid fractionation expressions of the type
TAB ( K ) =
(AA − AB ).
(11.175)
10 ln α AB
3
0.40 + 0.63
Tpyrite − galena ( K ) = . (11.176)
10 3 ln α AB
Figure 11.40 ⌬34S relative enrichment factors for various sulfur minerals with respect to
pyrite. Reproduced with modifications from Rye and Ohmoto, Economic Geology, 1974,
Vol. 69, p. 828.
Figure 11.41 Variation of ␦ 34S of sulfate, H2 S(aq), and sulfide minerals in response to
H2S/SO42⫺ variations in the hydrothermal solution at T ⫽ 200 °C and ␦ 34S∑S ⫽ 0‰. Repro-
duced with modifications from Rye and Ohmoto, Economic Geology, 1974, Vol. 69, p. 828.
H2 S (aq) ⇔ H + + HS − (11.177 )
Isotope Geochemistry / 799
Figure 11.42 ␦ 34S contours superimposed on the stability fields of Fe-S-O minerals and
barite at T ⫽ 250 °C and ␦ 34S∑S ⫽ 0‰ Solid lines and their dashed extensions: ␦ 34S con-
tours for pyrite (values in brackets) and barite (values in parentheses). Dashed lines: Fe-
S-O mineral boundaries for a 10⫺1 molal concentration of sulfur in the system. Small
dashed lines: Fe-S-O mineral boundaries for a 10⫺3 molal concentration of sulfur in the
system. Dash-and-dot line: solubility product of barite. Reproduced from Ohmoto, Eco-
nomic Geology, 1972, Vol. 72, p. 555.
HS − ⇔ H + + S 2 − , (11.178)
the increase in pH favors the formation of the relatively 34S-depleted ions HS⫺
and S2⫺, and hence the precipitating sulfides attain higher ␦34S values, T and Eh
being equal.
The combined effects of f O2 and pH variation on the sulfur isotopic com-
positions of ore-forming minerals are shown in figure 11.42 for a temperature of
250 °C and ␦ 34S⌺S ⫽ 0‰. The dashed lines delineate the stability limits of con-
densed phases in the Fe-S-O system for two different total sulfur concentrations
(10⫺1 and 10⫺2 molal). On these lines are superimposed ␦ 34S contours (solid lines
800 / Methods
and their dashed extensions) for coexisting pyrite (values in brackets) and barite
(values in parentheses). ␦ 34S for pyrite can range from ⫹5 to ⫺27 per mil and
for barite from 0 to ⫹32 per mil within geologically plausible Eh-pH ranges. In
acidic-reducing conditions at the limit of the pyrite stability field, the ␦ 34S value
of sulfide is similar to ␦ 34S⌺S and rather insensitive to Eh-pH changes. Neverthe-
less, at high Eh-pH conditions, near the pyrite-hematite and pyrite-magnetite sta-
bility limits, the ␦ 34S value of sulfide is largely variable in response to limited Eh-
pH changes and much dissimilar from ␦ 34S⌺S as a result of the presence of sulfate
ion in solution, which fractionates 34S.
Figure 11.42 elucidates in a rather clear fashion the marked control exerted
by the chemistry of the fluid on the sulfur isotopic compositions of ore-forming
minerals. At the T and ␦ 34S⌺S conditions delineated in the figure, an increase in
f O2 or in pH of one log unit can cause decreases in ␦ 34S values of sulfide minerals
by as much as 20‰.
Although these chemical effects are important in deciphering the genesis of
ore minerals, it must be emphasized that the differences in the ␦ 34S values among
coexisting condensed phases (hence the fractionation factor) at equilibrium are
constant in each case because they depend only on T.
APPENDIX ONE
The International Union for Pure and Applied Chemistry (IUPAC) recommends
the adoption in scientific publications of the SI system notation. Tables A1.1 and
A1.2 present the base units and some derived units.
In geochemical literature, one often encounters non-SI units (e.g., bar, atmo-
sphere, calorie, angstrom, electron-volt, etc.), which can be converted to SI units
by the use of simple multiplicative factors, as shown in table A1.3.
Length meter m
Mass kilogram kg
Time second s
Electric current ampere A
Thermodynamic temperature kelvin K
Amount of substance mole mol, mole
Intensity of light candle cd
Energy joule J m ⭈ kg ⭈s ⫺ 2
2
Force newton N m ⭈k g⭈ s ⫺ 2
Power watt W J⭈s⫺1
Pressure pascal Pa m ⫺1 ⭈ kg ⭈s ⫺ 2
Electric charge coulomb C A⭈s
Electric potential volt V m 2 ⭈ kg ⭈s ⫺ 3 ⭈A ⫺ 1
Capacitance farad F m ⫺2 ⭈ kg ⫺ 1 ⭈s 4 ⭈A 2
Resistance ohm ⍀ m 2 ⭈ kg ⭈s ⫺ 3 ⭈A ⫺ 2
Magnetic flux weber Wb m 2 ⭈ kg ⭈s ⫺ 2 ⭈A ⫺ 1
Magnetic flux density tesla T k g⭈ s ⫺ 2 ⭈A ⫺ 1
Frequency hertz Hz s ⫺1
/ 801
802 / Appendix 1
Table A1.3 Non-SI units and their factors for conversion to the SI system
Review of Mathematics
F = F ( x1 , x 2 , K , x n ) (A 2.1)
f1 dx1 + f2 dx 2 + K + fn dx n = ∑ fi dx i . (A2.2)
i
Whenever the differential equation A2.2 can be equated with the differential of
function F:
∂F ∂F ∂F
df = dx1 + dx 2 + L dx n , (A2.3)
∂ x1 x ∂ x2 x ,Kx ∂xn
2 ,K x n 1 n x1 , x2 , K
∂F ∂F ∂F
f1 = ; f2 = ; fn = . (A2.4)
∂ x1 x ∂x 2 x ∂x n x
2, K x n x
1 ,K n 1, x 2, K
/ 805
806 / Appendix 2
The state functions presented in section 2.4 are examples of exact differential
functions of the type exemplified by equation A2.3. Generally, reversible transfor-
mations in physical systems are described by equations of this type.
For two generic independent variables xg and xh , such that
∂F
fg = (A2.5)
∂ xg x
1,K
and
∂F
fh = (A2.6)
∂ xh x
1,K
we have
∂ fg ∂ 2F
∂x = (A2.7)
hx ∂ x g∂ x h
1,K
and
∂ fh ∂ 2F
∂x = . (A2.8)
g x ∂ xh ∂ xg
1, K
∂ 2F ∂ ∂F ∂ ∂F ∂ 2F
= = = , (A2.9)
∂ x∂ y ∂ x ∂ y ∂ y ∂ x ∂ y∂ x
we obtain
∂ fg ∂f
∂x = h . (A2.10)
h x ,K ∂ xg x ,K
1 1
Equation A2.10 represents a necessary and sufficient condition for equation A2.2
to be an exact differential and is valid for any couple of conjugate variables
( f g ,xg ) and ( f h ,xh ).
Review of Mathematics / 807
( )
f x, y, z = 0. (A 2.11)
The function
(
z = z x, y ) (A 2.12 )
∂f ∂f ∂f
dx + dy + dz = 0 (A2.13)
∂x ∂y ∂z
and, from equation A2.13, keeping constant one of the three variables and assum-
ing the partial derivatives of f to be non-null, we obtain
∂ x ∂ y ∂f ∂f
= 1 = − (A2.14.1)
∂ y z ∂ x z ∂ y ∂ x
∂ z ∂ y ∂f ∂f
= 1 = − (A2.14.2)
∂ y x ∂ z x ∂ y ∂ z
∂z ∂ x ∂f ∂f
= 1 = − . (A2.14.3)
∂ x y ∂z y ∂ x ∂ z
∂ x ∂ y ∂ z
⋅ ⋅ = −1 (A2.15)
∂ y z ∂ z x ∂x y
or
∂ x ∂ z ∂ z
− = . (A2.16)
∂ y z ∂ y x ∂ x y
808 / Appendix 2
∂ x ∂ x
dx = dy + dz (A2.17)
∂ y z ∂ z y
and a further function of the two variables x and y, (x,y), such that
∂ϕ ∂ϕ
dϕ = dx + dy. (A2.18)
∂ x y ∂ y x
If we let x and y vary while holding z constant, we get, from the partial derivative
in dx of equation A2.18,
∂ϕ ∂ϕ ∂ϕ ∂ y
= + . (A2.19)
∂ x z ∂ x y ∂ y x ∂ x z
Equation A2.19 is one of the possible ways to switch from a couple of variables
to another conjugate couple. Most transforms adopted in the thermodynamics
of systems with two independent variables are based on equation A2.19 and its
modifications (Legendre transforms, cf. section 2.4).
For a function F:
(
F = F x1 , x 2 , L , x n ) (A 2.20)
dF = f1 dx1 + f2 dx 2 + L + fn dx n , (A 2.21)
where
∂F
f1 = ,
∂ x1 x
2 ,K x n
∫ ( f1 dx1 + f2 dx 2 + L + fn dx n = 0 ) (A 2.22.2)
Review of Mathematics / 809
—i.e., the cyclic integral of such a function is zero. Moreover, the value of a
generic contour integral between points A and B in a field of existence described
by variables x1 , x2 , . . . ,xn does not depend on the path followed in integration.
That is, if
(
F = F x1 , x 2 ) (A 2.23)
∂F ∂F
dF = dx1 + dx 2 , (A2.24)
∂ x1 x ∂ x2 x
2 1
we have
B B
∂F ∂F
∫ dF = ∫ dx1 +
∂ x1 x
dx 2
∂ x2 x
A A 2 1
(A2.25)
B
= ∫ f1 dx1 + f 2 dx 2 = F( B ) − F( A)
A
V = V. (A 2.26)
With V being the modulus of vector V, its unit vector U is such that
V = V U. (A 2.27)
a ⋅ b = ab cos α ab , (A 2.28)
where a and b are the moduli of the respective vectors and ␣ab is the minimum
angle between the representatives of the unit vectors.
810 / Appendix 2
c = ab sin α ab . (A 2.29)
Ux Uy Uz
a × b = ax ay az (A2.30)
bx by bz
and the mixed product of three vectors a, b, and c is given by the determinant
ax ay az
a × b ⋅ c = bx by bz . (A2.31)
cx cy cz
The vectorial operator “gradient” (with symbol nabla ∇) allows the passage
from scalar to vectorial fields. For scalar ⌽ the vector ∇⌽ (gradient of ⌽) is
given by
∂Φ ∂Φ ∂Φ
∇Φ = , , . (A2.32)
∂x ∂y ∂z
In section 1.19, for example, ∇i⌽e, j was the vectorial component of the electrical
potential of the f th dipole ⌽e, j directed toward i.
The differential operator “divergence” allows the passage from vectorial to
scalar fields. For a vector ,
∂ϕ ∂ϕ ∂ϕ
div ϕ = ∇ϕ = , , . (A2.33)
∂x ∂y ∂z
The “curl” of a vector function (or “rotation” with symbol “rot”) is a vector
that is formally the cross product of the operator and the vector. For a vector V,
Review of Mathematics / 811
Ux Uy Uz
∂ ∂ ∂
rot V = ∇ × V = . (A2.34)
∂x ∂y ∂z
vx vy vz
∂2 ∂2 ∂2
∇2 = + + (A2.35)
∂ x 2 ∂ y 2 ∂ z2
∂Xl
Xl = ∑ ∂qj
dq j . (A2.36)
j
∂p
dp = ∑ ∂ qj
dq j . (A2.37)
j
dp ∂p ∂qj ∂p
V=
dt
= ∑ ⋅
∂qj ∂t
+
∂t
. (A2.38)
j
∂V ∂p ∂q j
= , where ∂ q˙ j = . (A2.39)
∂ q˙ j ∂q j ∂t
1
T= mV 2 , (A2.40)
2
we obtain
d ∂T ∂T
ma dp = ∑ dt ∂ q˙ − ∂ q dq j . (A2.41)
j j j
F dp = − dΦ (A2.42)
we also have
∂Φ
F dp = − dq j . (A2.43)
∂qj
∑ ( Fj )
− ma j dp j = 0
(A2.44)
j
* If the work exerted by a force field on a particle moving from point A to point B is indepen-
dent of the path followed by the particle between the two points, the force field is said to be “conser-
vative.”
Review of Mathematics / 813
∂Φ d ∂T ∂T
∑ − ∂ q − dt ∂ q + ∂ q dq j = 0, (A2.45)
j j j j
d ∂T ∂T ∂Φ
− =− . (A2.46)
dt ∂ q˙ j ∂ q j ∂qj
Posing
L = T − Φ, (A 2.47)
d ∂L ∂L
− = 0. (A2.48)
dt ∂ q j ∂ q j
˙
dx d 2x
q = x; x˙ = = V; x˙˙ = =a (A2.49.1)
dt dt 2
1 KF 2
T= mx˙ 2 ; Φ= x (A2.49.2)
2 2
1 K
L= mx˙ 2 − F x 2 . (A2.49.3)
2 2
d ∂L ∂L
− =0 (A2.50.1)
dt ∂ x ∂ x
d
mx˙ + KF x = 0 (A2.50.2)
dt
814 / Appendix 2
mx˙˙ + KF x = m a + KF x = 0. (A 2.50.3)
∂L
H= ∑ ∂ q˙ j
dq˙ j − L. (A2.51)
j
L = T − Φ, (A 2.52)
∂T
H= ∑ ∂ q˙ j
dq˙ j − T + Φ, (A2.53)
j
but, because
∂T
∑ ∂ q˙ j
dq˙ j = 2 T , (A2.54)
j
we obtain
H = T + Φ. (A 2.55)
As we have seen in sections 1.17 and 3.1, the quantum mechanics Hamilto-
nian in the Schrödinger equation has the form
h2
H=− ∇ 2 + Φ. (A2.56)
8π 2 m
The error function erf x and its complementary function erfc x appear in the
solution of differential equations describing diffusive processes (see, for instance,
section 11.6.3). They are defined by the integrals
x
erf x =
2
∫ ( )
exp − y 2 dy (A2.57)
π 0
Review of Mathematics / 815
and
∞
erfc x =
2
∫ ( )
exp − y 2 dy. (A2.58)
π x
( )
erf −∞ = −1; erf 0 = 0 ; erf ∞ = 1 (A 2.59)
Because
( )
erfc −∞ = 2; erfc 0 = 1; erfc ∞ = 0. (A 2.61)
Values of functions erf x and erfc x are tabulated (cf. Gautschi, 1964) but can
also be rapidly obtained by the polynomials
( )
−8
erfc x = 1 + a 1 x + a 2 x 2 + a 3 x 3 + a 4 x 4 + a 5 x 5 (A 2.62)
erfc x =
( )
exp − x 2 1 1 3 15 105 945 10, 365
− + 2 5 − 3 7 + 4 9 − 5 11 + 6 13
π x 2x 3
2 x 2 x 2 x 2 x 2 x
(A 2.63)
The least-squares treatment of the fitting of an isochron must account for the fact
that the error in the y-coordinate of each point is correlated with the error in the
x-coordinate.
With each point in the isochron being subjected to the requirement
yi = a + bx i i = 1, K , n, (A 2.64)
816 / Appendix 2
and denoting the observations as Xi and Yi, the adjusted values of these observa-
tions as xi and yi , the weights of the various observations as (Xi ) and (Yi ), and
the correlations between the x and y errors as ri , the best slope (i.e., “age”) can
be found by means of the equation
U bVi rV
∑ Zi2Vi ω Yi + − i i
i ( )
i ω Xi ( ) αi
b= , (A2.65)
U bVi brU
∑ Z i2U i i + − i i
i ω Yi( )ω ( X i ) αi
where
Zi =
( ) ( )
ω Xi ω Yi
b ω (Y ) + ω (X ) − 2 br α
2 (A2.66)
i i i i
U i = Xi − X (A 2.67)
Vi = Yi − Y (A 2.68)
∑ Zi X i
X = i
(A2.69)
∑ Zi
i
∑ ZiYi
Y = i
(A2.70)
∑ Zi
i
( ) ( )
α i = ω Xi ω Yi . (A2.71)
Because Ui , Vi, and Zi contain b, the solution of the system A2.65 →A2.71 is
obtained by an iterative procedure, inserting an approximate value of b into these
terms and calculating a new b from equation A2.65. The best intercept is found
by use of the equation
a = Y − bX . (A 2.72)
REFERENCES
Ackermann T. and Schreiner F. (1958). Molwärmen und Entropien einiger Fettsäuren und
ihrer Anionen in wässriger Lösung. Zeits. Elektroch., 62:1143–1151.
Acree W. E. Jr. (1984). Thermodynamic Properties of Nonelectrolyte Solutions. New York: Aca-
demic Press.
Aggarwal K. G. (1967). Lattice dynamics of diamond. Proc. Phys. Soc., 91:381–389.
Ahrens L. H. (1952). The use of ionization potentials, 1: Ionic radii of the elements. Geochim.
Cosmochim. Acta, 2:155–169.
Ahrland S. (1973). Thermodynamics of the stepwise formation of metal-ion complexes in aque-
ous solution. Struc. Bonding, 15:167–188.
Ahrland S. (1975). Metal complexes present in sea water. In The Nature of Sea Water, E. D.
Goldberg, ed. Dahlem Konferenzen.
Aikawa N., Kumazawa M., and Tokonami N. (1985). Temperature dependence of intersite
distribution of Mg and Fe in olivine and the associate change of lattice parameters.
Phys. Chem. Minerals, 12:1–8.
Akamatsu T., Kumazawa M., Aikawa N., and Takei H. (1993). Pressure effect on the divalent
cation distribution in nonideal solid solution of forsterite and fayalite. Phys. Chem. Min-
erals, 19:431–444.
Akimoto S., Komada E., and Kushiro I. (1967). Effect of pressure on the melting of olivine
and spinel polymorph of Fe2SiO4. J. Geophys. Res., 72:679–689.
Akimoto S., Matsui S., and Syono S. (1976). High-pressure crystal chemistry of orthosilicates
and the formation of the mantle transition zone. In The Physics and Chemistry of Miner-
als and Rocks, R. J. G. Strens, ed. New York: John Wiley.
Albarede F. (1976). Some trace element relationships among liquid and solid phases in the
course of fractional crystallization of magmas. Geochim. Cosmochim. Acta, 40:667–673.
Albarede F. and Bottinga Y. (1972). Kinetic disequilibrium in trace element partitioning be-
tween phenocrysts and host lava. Geochim. Cosmochim. Acta, 36:141–156.
Allègre C. J., Treuil M., Minster J. F., Minster B., and Albarede F. (1977). Systematic use of
trace element in igneous process, Part I: Fractional crystallization processes in volcanic
suites. Contrib. Mineral. Petrol., 60:57–75.
Allred G. C. and Woolley E. M. (1981). Heat capacities of aqueous acetic acid, sodium acetate,
ammonia, and ammonium chloride at 283.15 K, 298.15 K, and 313.15 K: ⌬C 0P for ion-
ization of acetic acid and for dissociation of ammonium ion. J. Soln. Chem., 14:549–560.
Althaus E. (1967). The triple point andalusite-sillimanite-kyanite. Contrib. Mineral. Petrol.,
13:31–50.
Alvarez L. W. (1938). The capture of orbital electrons by nuclei. Phys. Rev., 54:486–497.
Anderson A. T. and Gottfried D. (1971). Contrasting behaviour of P, Ti and Nb in a differenti-
ated high-alumina olivine-tholeite and a calc-alkaline suite. Geol. Soc. Amer. Bull.,
82:1929–1942.
Anderson A. T. and Greenland L. P. (1969). Phosphorous fractionation diagram as a quantita-
tive indicator of crystallization differentiation of basaltic liquid. Geochim. Cosmochim.
Acta, 33:493–505.
/ 817
818 / References
Barin I., Knacke O., and Kubaschewsky O. (1977). Thermochemical Properties of Inorganic
Substances: Supplement. Berlin-Heidelberg-New York: Springer-Verlag.
Barnes V. E. (1930). Changes in hornblend at about 800 °C. Amer. Mineral., 5:393–417.
Barr L. W. and Liliard A. B. (1971). Defects in ionic crystals. In Physical Chemistry: An Ad-
vanced Treatise, vol. 10, W. Jost (series ed.).
Barron L. M. (1976). A comparison of two models of ternary excess free energy. Contrib. Min-
eral. Petrol., 57:71–81.
Bass, J. D. (1986). Elasticity of uvarovite and andradite garnets. J. Geophys. Res., 91:7505–7516.
Bass J. D. (1989). Elasticity of grossular and spessartite garnets by Brillouin spectroscopy.
J. Geophys. Res., 94:7621–7628.
Basso R. (1985). Crystal chemical and crystallographic properties of compounds with garnet
or hydrogarnet structure. Neues Jahrb. Miner. Monat., 3:108–114.
Battino R. and Clever H. L. (1966). The solubilities of gases in liquids. Chem. Rev., 66:395–463.
Baur W. H. (1970). Bond length variation and distorted coordination polyedra in inorganic
crystals. Trans. Amer. Cryst. Assoc., 6:129–155.
Baur W. H. (1978). Variation of mean Si-O bond lengths in silicon-oxygen tetrahedra. Acta
Cryst., B34:1754–1756.
Baur W. H. and Kahn A. A. (1971). Rutile-type compounds, IV: SiO2, GeO2, and a comparison
with other rutile-type compounds. Acta Cryst., B27:2133–2139.
Becker R. H. (1971). Carbon and oxygen isotope ratios in iron-formation and associated rocks
from the Hamersley Range of Western Australia and their implications. Ph.D. diss.,
University of Chicago.
Bell R. G., Jackson R. A., and Catlow C. R. A. (1992). Loewenstein’s rule in zeolite A: A com-
putational study. Zeolites, 12:870–871.
Belton G. R., Suito H., and Gaskell D. R. (1973). Free energies of mixing in the liquid iron-
cobalt orthosilicates at 1450 °C. Met. Trans., 4:2541–2547.
Berger G. W. and York D. (1981). Geothermometry from 40Ar/39Ar dating experiments. Geo-
chim. Cosmochim. Acta, 45:795–811.
Berlin R. and Henderson C. M. B. (1968). A re-interpretation of Sr and Ca fractionation trends
in plagioclases from basic rock. Earth Planet. Sci. Lett., 4:79–83.
Berman R. G. (1988). Internally consistent thermodynamic data for minerals in the system
Na2O-K2O-CaO-MgO-FeO-Fe2O3-Al2O3-SiO2-TiO2-H2O-CO2. J. Petrol., 29:445–522.
Berman R. G. (1990). Mixing properties of Ca-Mg-Fe-Mn garnets. Amer. Mineral., 75:
328–344.
Berman R. G. and Brown T. H. (1984). A thermodynamic model for multicomponent melts
with applications to the system CaO-Al2O3-SiO2. Geochim. Cosmochim. Acta, 48:
661–678.
Berman R. G. and Brown T. H. (1985). Heat capacity of minerals in the system Na2O-K2O-
CaO-MgO-FeO-Fe2O3-Al2O3-SiO2-TiO2-H2O-CO2: Representation, estimation, and
high temperature extrapolation. Contrib. Mineral. Petrol., 89:168–183.
Berman R. G. and Brown T. H. (1987). Development of models for multicomponent melts:
Analysis of synthetic systems. In Reviews in Mineralogy. vol. 17, P. H. Ribbe (series ed.),
Mineralogical Society of America.
Bernal J. D. (1964). The structure of liquids. Proc. Roy. Soc. London, A280:299–322.
Berner R. A. (1971). Principles of Chemical Sedimentology. New York: McGraw-Hill.
Berner R. A. (1980). Early Diagenesis: A Theoretical Approach. Princeton: Princeton Univer-
sity Press.
Berner R. A., Sjoberg E. L., and Schott J. (1980). Mechanisms of pyroxene and amphibole
weathering, I: Experimental studies. In Third International Symposium on Water-Rock
Interaction: Proceedings, Edmonton: Alberta Research Council.
Berthelot M. (1872). On the law which governs the distribution of a substance between two
solvents. Ann. Chim. Phys., 4th. ser., 26:408–417.
Bertorino G., Caboi R., Caredda A., Cidu R., Fanfani L., and Zuddas P. (1981). Caratteri
820 / References
idrogeochimici delle acque naturali della Sardegna meridionale, 3: Le acque della Mar-
milla e del Sarcidano. Rend. S.I.M.P, 37:951–966.
Bertrand G. L., Acree W. E. Jr., and Burchfield T. (1983). Thermodynamical excess properties
of multicomponent systems: Representation and estimation from binary mixing data.
J. Solution. Chem., 12:327–340.
Bethe H. (1929). Splitting of terms in crystals. Amer. Phys., 3:133–206.
Bieri R., Koide M., and Goldberg E. D. (1966). Noble gases contents of Pacific seawater.
J. Geophys. Res., 71:5243–5265.
Bigeleisen J. (1949). The relative reaction velocities of isotopic molecules. J. Chem. Phys.,
17:675–678.
Bigeleisen J. and Mayer M. G. (1947). Calculation of equilibrium constants for isotopic ex-
change reactions. J. Chem. Phys., 13:261–267.
Biggerstaff D. R. (1986). The thermodynamic properties of aqueous solutions of argon, ethyl-
ene, and xenon up to 720 K and 34 MPa. Ph.D. diss., University of Delaware.
Bina C. R. and Wood B. J. (1987). The olivine-spinel transitions: Experimental and thermody-
namic constraints and implications for the nature of the 400 km seismic discontinuity.
J. Geophys. Res., 92:4853–4859.
Birch F. (1966). Compressibility. In Handbook of Physical Constants, S. P. Clark Jr., ed. Memoir
97, The Geological Society of America.
Bird D. K. and Helgeson H. C. (1980). Chemical interaction of aqueous solution with
epidolite-feldspar mineral assemblage in geologic systems, I: Thermodynamic analysis
of phase relations in the system CaO-FeO-Fe2O3-Al2O3-SiO2-H2O-CO2. Amer. Jour.
Sci., 280:907–941.
Bird M. L. (1971). Distribution of trace elements in olivines and pyroxenes: An experimental
study. Ph.D. diss., University of Missouri, Rolla.
Bish D. L. (1981). Cation ordering in synthetic and natural Ni-Mg olivine. Amer. Mineral.,
66:770–776.
Bishop F. C. (1980). The distribution of Fe⫹⫹ and Mg between coexisting ilmenite and pyrox-
ene with application to geochemistry. Amer. Jour. Sci., 280:46–47.
Bishop F. C., Smith J. V., and Dawson J. B. (1976). Na, P, Ti, and coordination of Si in garnet
from peridotite and eclogite xenoliths. Nature, 260:696–697.
Blasi A., Brajkovic A., and De Pol Blasi (1983). Dry-heating conversions of low microcline to
high sanidine via a one-step disordering process. Abstr. NATO Advanced Study Inst. on
Feldspars and Feldspatoids, Rennes, France.
Blencoe J. G. (1977). Molal volumes of synthetic paragonite-muscovite micas. Amer. Mineral.,
62:1200–1215.
Blencoe J. G. and Luth W. C. (1973). Muscovite-paragonite solvi at 2, 4, and 8 kbar pressure.
Geol. Soc. Amer. Abstr. with Programs, 5:553–554.
Bocchio R., Bondi M., and Domeneghetti M. C. (1979). Reexamination of shefferite, urbanite,
and lindesite from central Sweden, Part I: Chemistry and crystal structure. Per. Min-
eral., 48:181–193.
Bockris J. O’M. and Mellors G. W. (1956). Electric conductance in liquid lead silicates and
borates. J. Phys. Chem., 60:1321–1328.
Bockris J. O’M. and Reddy A. K. N. (1970). Modern Electrochemistry. New York: Plenum
Press.
Bockris J. O’M., Kitchener J. A., and Davies A. E. (1952a). Electric transport in liquid silicates.
Trans. Faraday Soc., 48:536–548.
Bockris J. O’M., Kitchener J. A., Ignatowicz S., and Tomlinson J. W. (1952b). The electrical
conductivity of silicate melts: Systems containing Ca, Mn, Al. Discuss. Faraday Soc.,
4:281–286.
Boeglin J. L. (1981). Mineralogie et geochimie des gisements de manganese de Conselheiro
Lafaiete au Bresil et de Moanda au Gabon. These 3eme Cycle, Toulouse, 1608.
Bohlen S. R. and Boettcher A. L. (1981). Experimental investigations and geological applica-
tions of orthopyroxene geobarometry. Amer. Mineral., 66: 951–964.
References / 821
Bohlen S. R. and Essene E. J. (1977). Feldspar and oxide thermometry of granulites in the
Adirondack Highlands. Contrib. Mineral. Petrol., 62:153–169.
Bohlen S. R., Wall V. J., and Boettcher A. L. (1982). The system albite-water-carbon dioxide: A
model for melting and activities of water at high pressures. Amer. Mineral., 67:451–462.
Bohlen S. R., Wall V. J., and Boettcher A. L. (1983). Experimental investigation and geological
application of equilibria in the system FeO-TiO2-Al2O3-SiO2-H2O. Amer. Mineral.,
68:1049–1058.
Bokreta M. (1992). Energetics of garnets: Computational model of the thermochemical and
thermophysical properties. Ph.D. diss., University of Philadelphia.
Bokreta M. and Ottonello G. (1987). Enthalpy of formation of end-member garnets. EOS,
68:448.
Bonatti E. (1965). Palagonite, hyaloclastites, and alteration of volcanic glass in the ocean. Bull.
Volcanol., 28:251–269.
Bonatti E., Ottonello G., and Hamlyn P. R. (1986). Peridotites from the Island of Zabargad
(St. John)., Red Sea: Petrology and geochemistry. J. Geophys. Res., 91:599–631.
Boon J. A. (1971). Mössbauer investigations in the system Na2O-FeO-SiO2. Chem. Geol.,
7:153–169.
Boon J. A. and Fyfe W. S. (1972). The coordination number of ferrous ions in silicate systems.
Chem. Geol., 10:287–298.
Borg I. Y. (1967). Optical properties and cell parameter in the glaucophane-riebeckite series.
Contrib. Mineral. Petrol., 15:67–92.
Born M. (1920). Volumen und Hydratationswarme der ionen. Zeitschr. Physic, 1:45–48.
Born M. and Von Karman T. (1913). Hüber die Verteilung der Eigenschwingungen von Punkt-
gittern. Physik. Zeitschr., 14:65–71.
Boström D. (1987). Single-crystal X-ray diffraction studies of synthetic Ni-Mg olivine solid
solutions. Amer. Mineral., 66:770–776.
Boström D. (1988). Experimental studies of (Ni,Mg)—and (Co,Mg)—olivine solid solutions
by single crystal X-ray diffraction and solid state emf method. Ph.D. thesis, Department
of Inorganic Chemistry, University of Umeå, Umeå, Sweden.
Boström D. and Rosen E. (1988). Determination of activity-composition in (Ni,Mg)2SiO4 solid
solution at 1200–1600 K by solid state emf measurements. Acta Chim. Scand.,
A42:149–155.
Boswara I. M. and Franklyn A. D. (1968). Theory of the energetics of simple defects in oxides.
In Mass Transport in Oxides, J. B. Wachtman and A. D. Franklyn, eds. Natl. Bur. Stand.
(U.S.). Spec. Publ., 296.
Bottinga Y. (1968). Calculation of fractionation factors for carbon and oxygen isotopic ex-
changes in the system calcite-carbon dioxide-water. J. Phys. Chem., 72:800–807.
Bottinga Y. (1969a). Calculated fractionation factors for carbon and hydrogen isotope ex-
change in the system calcite-carbon dioxide-graphite-methane-hydrogen-water vapor.
Geochim. Cosmochim. Acta, 33:49–64.
Bottinga Y. (1969b). Carbon isotope fractionation between graphite, diamond and carbon di-
oxide. Earth Planet. Sci. Letters, 5:301–307.
Bottinga Y. (1985). On the isothermal compressibility of silicate liquids at high pressure. Earth
Planet. Sci. Letters., 74:350–360.
Bottinga Y. E. and Richet P. (1981). High pressure and temperature equation of state and calcu-
lation of the therodynamic properties of gaseous carbon dioxide. Amer. Jour. Sci.,
281:615–660.
Bottinga Y. and Weill D. F. (1970). Densities of liquid silicates systems calculated from partial
molar volumes of oxide components. Amer. Jour. Sci., 269:169–182.
Bowen N. L. (1913). Melting phenomena in the plagioclase feldspar. Amer. Jour. Sci.,
35:577–599.
Bowen N. L. (1915). The crystallization of haplobasaltic, haplorhyolitic and related magmas.
Amer. Jour. Sci., 4th ser., 33:551–573.
Bowen N. L. (1928). The Evolution of Igneous Rocks. Princeton University Press, Princeton, N.J.
822 / References
Bowen N. L. and Schairer J. F. (1935). The system MgO-FeO-SiO2. Amer. Jour. Sci.,
29:151–217.
Bowen N. L. and Schairer J. F. (1938). Crystallization equilibrium in nepheline-albite-silica
mixtures with fayalite. J. Geol., 46:397–411.
Bowen N. L., Shairer J. F., and Posniak E. (1933). The system Ca2SiO4-Fe2SiO4. Amer. Jour.
Sci., 26:273–297.
Boyd F. R. (1973). The pyroxene geothermometry. Geochim. Cosmochim. Acta, 37:2533–2546.
Boyd F. R. and Schairer J. F. (1964). The system MgSiO3-CaMgSi2O6. J. Petrol., 5:275–309.
Bragg W. L. (1920). The arrangement of atoms in crystals. Phil. Mag., 40:169–172.
Brereton N. R. (1970). Corrections for interfering isotopes in the 40Ar/ 39Ar dating method.
Earth Planet. Sci. Letters, 8:427–433.
Brewer P. G. (1975). Minor elements in seawater. In Chemical Oceanography. vol. 2, J. P. Riley
and G. Skirrow, eds. New York: Academic Press.
Bricker O. (1965). Some stability relations in the system Mn-O2-H2O at 25° and one atmosphere
total pressure. Amer. Mineral., 57:284–287.
Brillouin L. (1953). Wave Propagation in Periodic Structures. New York: Dover.
Brinkmann U. and Laqua W. (1985). Decomposition of fayalite (Fe2SiO4). in an oxygen poten-
tial gradient at 1418 K. Phys. Chem. Minerals, 12:283–290.
Broecker W. S., Gerard R. D., Ewing M., and Heezen B. C. (1961). Geochemistry and physics
of ocean circulation. In Oceanography, M. Sears, ed. Amer. Assn. Adv. Sci., Wash-
ington.
Brookins D. G. (1988). Eh-pH Diagrams for Geochemistry. Berlin-Heidelberg-New York:
Springer-Verlag.
Brown B. E. and Bailey S. W. (1964). The structure of maximum microcline and the sanidine-
microcline series. Norsk Geol. Tidsskr., 42/2:25–36.
Brown G. E. (1970). The crystal chemistry of the olivines. Ph.D. Dissertation, Virginia Polytech-
nic Inst. and State University, Blacksburg, Virginia.
Brown G. E. (1982). Olivines and silicate spinels. In Reviews in Mineralogy, vol. 5, P. H. Ribbe
(series ed.), Mineralogical Society of America.
Brown G. E. and Prewitt C. T. (1973). High temperature crystal chemistry of hortonolite. Amer.
Mineral., 58:577–587.
Brown G. E., Prewitt C. T., Papike J. J., and Sueno S. (1972). A composition of the structures
of low and high pigeonite. J. Geophys. Res., 77:5778–5789.
Brown G. M. (1967). Mineralogy of basaltic rocks. In Basalts, H. H. Hess and A. Poldervaart,
eds., New York: John Wiley.
Brown I. D. and Shannon R. D. (1973). Empirical bond-strength—bond-length curves for
oxides. Acta Cryst., A29:266–282.
Brown W. L. and Parson I. (1981). Towards a more practical two-feldspar geothermometer.
Contrib. Mineral. Petrol., 76:369–377.
Bruno E., Chiari G., and Facchinelli A. (1976). Anorthite quenched from 1530 °C. I. Structure
refinement. Acta Cryst., B32:3270–3280.
Buddington A. F. and Lindsley D. H. (1964). Iron-titanium oxide minerals and synthetic equiv-
alents. J. Petrol., 5:310–357.
Buening D. K. and Buseck P. K. (1973). Fe-Mg lattice diffusion in olivine. J. Geophys. Res.,
78:6852–6862.
Buiskool Toxopeus J. M. A. and Boland J. M. (1976). Several types of natural deformation in
olivine: An electron microscope study. Tectonophysics, 32:209–233.
Burnham C. W. (1967). Ferrosilite. Carnegie Inst. Wash. Yb., 65:285–290.
Burnham C. W. (1975). Water and magmas: A mixing model. Geochim. Cosmochim. Acta,
39:1077–1084.
Burnham C. W. and Davis N. F. (1974). The role of H2O in silicate melts, II: Thermodynamic
and phase relations in the system NaAlSi3O8-H2O to 10 kbar, 700 °C to 1000 °C. Amer.
Jour. Sci., 274:902–940.
References / 823
ing in minerals, III: Order parameter coupling in potassium feldspar. Amer. Mineral.,
79:1084–1093.
Carpenter M. A., Powell R., and Salje E. K. H. (1994). Thermodynamics of nonconvergent
cation ordering in minerals, I: An alternative approach. Amer. Mineral., 79:1053–1067.
Catanzaro E. J. (1963). Zircon ages in southwestern Minnesota. J. Geophys. Res., 68:2045–2048.
Catlow C. R. A. and Stoneham A. M. (1983). Ionicity in solids. J. Phys., C16:4321–4338.
Catti M. (1981). A generalized Born-Mayer parametrization of the lattice energy in orthorhom-
bic ionic crystals. Acta Cryst., A37:72–76.
Chacko T., Mayeda T. K., Clayton R. N., and Goldsmith J. R. (1991). Oxygen and carbon
isotope fractionations between CO2 and calcite. Geochim. Cosmochim. Acta, 55:2867–
2882.
Chakraborty S. and Ganguly J. (1991). Compositional zoning and cation diffusion in garnets.
In Advances in Physical Geochemistry, vol. 8, J. Ganguly, ed. New York-Heidelberg-
Berlin: Springer-Verlag.
Chakraborty S. and Ganguly J. (1992). Cation diffusion in aluminosilicate garnets: Experimen-
tal determination in spessartine-almandine diffusion couples, evaluation of effective bi-
nary diffusion coefficients, and applications. Contrib. Mineral. Petrol., 111:74–86.
Charles R. W. (1974). The physical properties of the Mg-Fe richerites. Amer. Mineral.,
59:518–528.
Charles R. W. (1980). Amphiboles on the join pargasite-ferropargasite. Amer. Mineral., 65:996–
1001.
Charlu T. V., Newton R. C., and Kleppa O. J. (1975). Enthalpies of formation at 970 K of
compounds in the system MgO-Al2O3-SiO2 from high temperature solution calorimetry.
Geochim. Cosmochim. Acta, 39:1487–1497.
Charlu T. V., Newton R. C., and Kleppa O. J. (1978). Enthalpy of formation of lime silicates
by high temperature calorimetry, with discussion of high pressure phase equilibrium.
Geochim. Cosmochim. Acta, 42:367–375.
Chatillon-Colinet C., Newton R. C., Perkins D., and Kleppa O. J. (1983). Thermochemistry of
(Fe⫹⫹,Mg)SiO3 orthopyroxene. Geochim. Cosmochim. Acta, 47:1597–1603.
Chatterjee N. (1987). Evaluation of thermochemical data on Fe-Mg olivine, orthopyroxene,
spinel and Ca-Fe-Mg-Al garnet. Geochim. Cosmochim. Acta, 51:2515–2525.
Chatterjee N. (1989). An internally consistent thermodynamic data base on minerals: Applica-
tions to the earth’s crust and upper mantle. Ph.D. diss., City University of New York.
Chatterjee N. D. (1970). Synthesis and upper stability of pargasite. Contrib. Mineral. Petrol.,
27:244–257.
Chatterjee N. D. (1972). The upper stability limit of the assemblage paragonite ⫹ quartz and
its natural occurrences. Contrib. Mineral. Petrol., 34:288–303.
Chatterjee N. D. (1974). Synthesis and upper thermal stability limit of 2M-margarite,
CaAl2[Al2Si2O10(OH)2]. Schweiz. Mineral. Petrogr. Mitt., 54:753–767.
Chatterjee N. D. (1976). Margarite stability and compatibility relations in the system CaO-
Al2O3-SiO2-H2O as a pressure-temperature indicator. Amer. Mineral., 61:699–709.
Chatterjee N. D. and Froese E. (1975). A thermodynamic study of the pseudobinary join
muscovite-paragonite in the system KAlSi3O8-NaAlSi3O8-Al2O3-SiO2-H2O. Amer. Min-
eral., 61:699–709.
Chatterjee N. D. and Johannes W. (1974). Thermal stability and standard thermodynamic
properties of synthetic 2M, muscovite KA1[AlSi3O10(OH)2]. Contrib. Mineral. Petrol.,
48:89–114.
Cheng W. and Ganguly J. (1994). Some aspects of multicomponent excess free energy models
with subregular binaries. Geochim. Cosmochim. Acta, 58:3763–3767.
Chiari G., Facchinelli A. and Bruno E. (1978). Anorthite quenched from 1630 °C. Discussion.
Acta Cryst., B34:1757–1764.
Chiba H., Chacko T., Clayton R. N., and Goldsmith J. R. (1989). Oxygen isotope fraction-
ations involving diopside, forsterite, magnetite, and calcite: Application to geother-
mometry. Geochim. Cosmochim. Acta, 53:2985–2995.
References / 825
Chiodini G. and Cioni R. (1989). Gas geobarometry for hydrothermal systems and its applica-
tions to some Italian geothermal areas. Appl. Geochem., 4:465–472.
Clark A. M. and Long J. V. P. (1971). Anisotropic diffusion of nickel in olivine. In Graham
Memorial Symposium on Diffusion Processes. New York: Gordon & Breach.
Clark J. R., Appleman D. E., and Papike J. J. (1969). Crystal-chemical characterization of cli-
nopyroxenes based on eight new structure refinements. Min. Soc. Amer. Spec. Paper,
2:31–50.
Clayton R. N. and Kieffer S. W. (1991). Oxygen isotopic thermometer calibrations. In Stable
Isotope Geochemistry: A Tribute to Samuel Epstein, H. P. Taylor Jr., J. R. O’Neil, and
I. R. Kaplan, eds., The Geochemical Society, Special Publication n. 3.
Clayton R. N., Goldsmith J. R., and Mayeda T. K. (1989). Oxygen isotope fractionations in
quartz, albite, anorthite and calcite. Geochim. Cosmochim. Acta, 53:725–733.
Clugston M. J. (1978). The calculation of intermolecular forces: A critical examination of the
Gordon-Kim model. Adv. Phys., 27:893–912.
Cohen R. E. (1986). Thermodynamic solution properties of aluminous clinopyroxenes: Nonlin-
ear least squares refinements. Geochim. Cosmochim. Acta, 50:563–575.
Coltorti M., Girardi V., and Schoerster J. (1987). Liquid immiscibility in the Archean Green-
stone Belt of Piumhi (Minas Gerais, Brazil). Lithos, 20:77–91.
Colville A. A. and Ribbe P. H. (1968). The crystal structure of an adularia and a refinement of
a structure of orthoclase. Amer. Mineral., 53:25–37.
Connolly J. A. D. (1992). Phase diagram principles and computations: A review. In Proceedings
of the V Summer School of Earth and Planetary Sciences, University of Siena Press,
Siena.
Connolly J. A. D. (1994). Computer calculation of multidimensional phase diagrams: Why and
how? In Proceedings of the VII Summer School of Earth and Planetary Sciences, Univer-
sity of Siena Press, Siena.
Connolly J. A. D. and Kerrick D. M. (1987). An algorithm and computer program for calculat-
ing composition phase diagrams. CALPHAD, 11:1–55.
Coulson C. A. (1961). Valence. (2d ed.) Oxford University Press, London.
Craig H. (1957). Isotopic standards for carbon and oxygen and correction factors for mass-
spectrometric analysis of carbon dioxide. Geochim. Cosmochim. Acta, 12:133–149.
Craig H. (1961). Standards for reporting concentrations of deuterium and oxygen-18 in natural
waters. Science, 133:1833–1934.
Craig H. (1963). The isotopic geochemistry of water and carbon in geothermal areas. In Nu-
clear Geology of Geothermal Areas, E. Tongiorgi, ed. Spoleto.
Crank J. (1975). The Mathematics of Diffusion Oxford University Press, London.
Cressey G. (1978). Exsolution in almandine-pyrope-grossular garnet. Nature, 271:533–534.
Cressey G., Schmid R., and Wood B. J. (1978). Thermodynamic properties of almandine-
grossular garnet solid solutions. Contrib. Mineral. Petrol., 67:397–404.
Cripps-Clark C. J., Sridhar R., Jeffes J. H. E., and Richardson F. D. (1974). Chain distribution
and transition temperatures for phosphate glasses. In Physical Chemistry of Process
Metallurgy. J. H. E. Jeffes and R. J. Tait, eds. Inst. Mining Met.
Currie K. L. (1971). The reaction 3 cordierite ⫽ 2 garnet ⫹ 4 sillimanite ⫹ 5 quartz as a
geological thermometer in the Ipinicon Lake Region, Ontario. Contrib. Mineral. Pet-
rol., 33:215–226.
Currie K. L. and Curtis L. W. (1976). An application of multicomponent solution theory to
jadeitic pyroxenes. J. Geol., 84:179–194.
Curtis G. H. and Hay R. L. (1972). Further geological studies and potassium-argon dating at
Olduvai Gorge and Ngorongoro Crater. In Calibration of Hominoid Evolution, W. W.
Bishop and J. A. Miller, eds. Edinburgh: Scottish Academic Press.
Czaya R., (1971). Refinement of the structure of ␣⫺Ca2SiO4. Acta Cryst., B27:848–849.
Dahl P. S. (1980). The thermal compositional dependence of Fe⫹⫹–Mg distributions between
co-existing garnet and pyroxene: Applications to geothermometry. Amer. Mineral.,
65:852–866.
826 / References
Dal Negro A., Carbonin S., Molin G. M., Cundari A., and Piccirillo E. M. (1982). Intracrystal-
line cation distribution in natural clinopyroxenes of tholeiitic, transitional and alkaline
basaltic rocks. In Advances in Physical Geochemistry. vol. 1, S. K. Saxena (series ed.),
New York: Springer-Verlag.
Dal Negro A., Carbonin S., Domeneghetti C., Molin G. M., Cundari A., and Piccirillo E. M.
(1984). Crystal chemistry and evolution of the clinopyroxene in a suite of high pressure
ultramafic nodules from the newer volcanism of Victoria, Australia. Contrib. Mineral.
Petrol., 86:221–229.
Dalrymple G. B. and Lanphere M. A. (1969). Potassium-Argon Dating W. H. Freeman, San
Francisco.
Dalrymple G. B. and Lanphere M. A. (1971). 40Ar/ 39Ar dating technique of K-Ar dating: A
comparison with the conventional technique. Earth Planet. Sci. Letters, 12:300–308.
Damon P. E., Lerman J. C., and Long A. (1978). Temporal fluctuations of 14C: Casual factors
and implications. Ann. Rev. Earth Planet. Sci., 6:457–464.
D’Amore F. and Panichi C. (1980). Evaluation of deep temperatures of hydrothermal systems
by a new gas geothermometer. Geochim. Cosmochim. Acta, 44:549–556.
Darken L. S. (1948). Diffusion mobility and their interrelations through free energy in binary
metallic systems. Trans. Met. Soc. AIME, 175:184–201.
Darken L. S. (1967). Thermodynamics of binary metallic solutions. Trans. Met. Soc. AIME,
239:80–89.
Das C. D., Keer H. V., and Rao R. V. G. (1963). Lattice energy and other properties of some
ionic crystals. Z. Physic. Chem., 224:377–383.
Davidson P. M. and Mukhopadhyay D. K. (1984). Ca-Fe-Mg olivines: Phase relations and a
solution model. Contrib. Mineral. Petrol., 86:256–263.
Davis L. L. and Smith D. (1993). Ni-rich olivine in minettes from Two Buttes, Colorado: A
connection between potassic melts from the mantle and low Ni partition coefficients.
Geochim. Cosmochim. Acta, 57:123–129.
Davoli P. (1987). A crystal chemical study of aegirin-augites and some evaluations on the oxida-
tion state of Mn. Neues Jahrb. Miner. Abh., 158:67–87.
De A. (1974). Silicate liquid immiscibility in the Deccan Traps and its petrogenetic significance.
Geol. Soc. Amer. Bull., 85:471–474.
De Baar H. J. W., Bacon M. P., Brewer P. G., and Bruland K. W. (1985). Rare earth elements
in the Pacific and Atlantic Oceans. Geochim. Cosmochim. Acta, 49:1943–1960.
Debye P. (1912). Zur Theorie der spezifischen warmer. Ann. Physik, 39(4).: 789–839.
Deer W. A., Howie R. A., and Zussman J. (1978). Rock Forming Minerals vol. 2A, (2d ed.),
New York: John Wiley.
Deer W. A., Howie R. A., and Zussman J. (1983). An Introduction to the Rock-Forming Miner-
als. Longman, Harlow Essex, England.
Deganello S. (1978). Thermal expansion from 25°C to 500°C of a few ionic radii. Zeit. Krist.,
147:217–227.
Deines P., Langmuir D., and Harmon R. S. (1974). Stable carbon isotope ratios and the exis-
tence of a gas phase in the evolution of carbonate ground waters. Geochim. Cosmochim.
Acta, 38:1147–1164.
Delany J. M., Puigdomenech I., and Wolery T. J. (1986). Precipitation Kinetics Option for the
EQ6 Geochemical Reaction Path Code Lawrence Livermore National Laboratory, Liv-
ermore, Cal., UCRL-53642.
Della Giusta A. and Ottonello G. (1993). Energy and long-range disorder in simple spinels.
Phys. Chem. Minerals, 20:228–241.
Della Giusta A., Ottonello G., and Secco L. (1990). Precision estimates of interatomic distances
using site occupancies, ionization potentials and polarizability in Pbnm silicate olivines.
Acta Cryst., B46:160–165.
Dempsey M. J. (1980). Evidence for structural changes in garnet caused by calcium substitu-
tion. Contrib. Mineral. Petrol., 7:281–282.
References / 827
nite and its bearing on the origin of some bedded cherts. Amer. Jour. Sci., 267A:114–
133.
Ernst W. G. and Wai C. M. (1970). Infrared, X-ray and optical study of cation ordering and
dehydrogenation in natural and heat-treated sodic amphiboles. Amer. Mineral.,
55:1226–1258.
Essene E. J. (1982). Geologic thermometry and barometry. In Reviews in Mineralogy, vol. 10,
P. H. Ribbe (series ed.), Mineralogical Society of America.
Eugster H. P. (1954). Distribution of cesium between sanidine and a hydrous fluid. Carnegie
Inst. Wash. Yb., 53:102–104.
Eugster H. P. and Baumgartner L. (1987). Mineral solubilities and speciation in supercritical
metamorphic fluids. In Reviews in Mineralogy. vol. 17, P. H. Ribbe (series ed.), Mineral-
ogical Society of America.
Eugster H. P. and Wones D. R. (1962). Stability relations of the ferruginos biotite, annite.
J. Petrol., 3:82–125.
Eugster H. P., Albee A. L., Bence A. E., Thompson J. B. Jr., and Walbaum D. R. (1972). The
two phase region and excess mixing properties of paragonite-muscovite crystalline solu-
tions. J. Petrol., 13:147–179.
Ewart A. and Taylor S. R. (1969). Trace element geochemistry of rhyolitic volcanic rocks, Cen-
tral North Island, New Zealand. Phenocryst data. Contrib. Mineral. Petrol., 22:127–138.
Ewart A., Bryan W. B., and Gill J. B. (1973). Mineralogy and geochemistry of younger volcanic
islands of Tonga S. W. Pacific. J. Petrol., 14:429–465.
Ewart A., Taylor S. R., and Capp A. C. (1968). Trace and minor element geochemistry of the
rhyolitic volcanic rocks, Central North Island, New Zealand. Contrib. Mineral. Petrol.,
18:76–85.
Faure G. (1986). Principles of Isotope Geology. 2d ed. New York: John Wiley.
Faure G. and Powell J. L. (1972). Strontium Isotope Geology. New York: Springer-Verlag.
Fei Y. and Saxena S. K. (1986). A thermochemical data base for phase equilibria in the system
Fe-Mg-Si-O at high pressure and temperature. Phys. Chem. Minerals., 13:311–324.
Fei Y. and Saxena S. K. (1987). An equation for the heat capacity of solids. Geochim. Cos-
mochim. Acta, 51:251–254.
Fei Y., Saxena S. K., and Eriksson G. (1986). Some binary and ternary silicate solution models.
Contrib. Mineral. Petrol., 94:221–229.
Ferguson J. and Currie K. L. (1971). Evidence of liquid immiscibility in alkaline ultrabasic
dikes at Callender Bay, Ontario. J. Petrol., 12:561–585.
Ferry J. M. and Spear F. S. (1978). Experimental calibration of the partitioning of Fe and Mg
between biotite and garnet. Contrib. Mineral. Petrol., 66:113–117.
Fincham C. J. B. and Richardson R. F. (1954). The behaviour of sulfur in silicate and aluminate
melts. Proc. Roy. Soc. London, A223:40–61.
Finger L. W. (1970). Refinement of the crystal structure of an anthophyllite. Carnegie Inst.
Wash. Yb., 68:283–288.
Finger L. W. and Ohashi Y. (1976). The thermal expansion of diopside to 800 °C and a refine-
ment of the crystal structure at 700 °C. Amer. Mineral., 61:303–310.
Finnerty T. A. (1977). Exchange of Mn, Ca, Mg and Al between synthetic garnet, orthopyrox-
ene, clinopyroxene and olivine. Carnegie Inst. Wash. Yb., 68:290–292.
Finnerty T. A. and Boyd F. R. (1978). Pressure dependent solubility of calcium in forsterite
coexisting with diopside and enstatite. Carnegie Inst. Wash. Yb., 77:713–717.
Fisher K. (1966). A further refinement of the crystal structure of cummingtonite, (Mg, Fe)7
(Si4O11)2(OH)2. Amer. Mineral., 51:814–818.
Fitch F. J., Miller J. A., and Hooker P. J. (1976). Single whole rock K-Ar isochron. Geol.
Mag., 113:1–10.
Fleer V. N. (1982). The dissolution kinetics of anorthite (CaAl2Si2O8) and synthetic strontium
feldspar (SrAl2Si2O8) in aqueous solutions at temperatures below 100°C: With applica-
tions to the geological disposal of radioactive nuclear wastes. Ph.D. diss., Pennsylvania
State University, University Park.
References / 829
Flood H. and Förland T. (1947). The acidic and basic properties of oxides. Acta Chem.
Scand., 1:952–1005.
Flood H., Förland T., and Gzotheim K. (1954). Über den Zusammenhang zwischen Konzen-
trazion und Aktivitäten in geschmolzenen Salzmischungen. Zeit. Anorg. Allg. Chem.,
276:290–315.
Foland K. A. (1974). Alkali diffusion in orthoclase. In Geochemical Transport and Kinetics,
Hoffman, Giletti, Yoder, and Yund, eds. Carnegie Institution of Washington.
Francis C. A. and Ribbe P. H. (1980). The forsterite-tephroite series, I: Crystal structure re-
finement. Amer. Mineral., 65:1263–1269.
Franz G., Hinrichsen T., and Wannermacher E. (1977). Determination of the miscibility gap
on the mean of the infrared spetroscopy. Contrib. Mineral. Petrol., 59:207–236.
Fraser D. G. (1975a). Activities of trace elements in silicate melts. Geochim. Cosmochim.
Acta, 39:1525–1530.
Fraser D. G. (1975b). An investigation of some long-chain oxi-acid systems. D.Phil. diss., Uni-
versity of Oxford.
Fraser D. G. (1977). Thermodynamic properties of silicate melts. In Thermodynamics in Geol-
ogy. D. G. Fraser, ed. Reidel, Dortrecht-Holland.
Fraser D. G., Rammensee W., and Hardwick A. (1985). Determination of the mixing properties
of molten silicates by Knudsen cell mass spectrometry, II: The system (Na-K)AlSi4O10
and (Na-K)AlSi5O12. Geochim. Cosmochim. Acta, 49:349–359.
Fraser D. G., Rammensee W., and Jones R. H. (1983). The mixing properties of melts in the
system NaAlSi3O6-KAlSi2O6 determined by Knudsen-Cell Mass Spectrometry. Bull.
Mineral., 106:111–117.
Frey F. A. (1982). Rare earth element abundances in upper mantle rocks. In Rare Earth Element
Geochemistry, P. Henderson, ed. Elesevier, Amsterdam.
Friedel J. (1964). Dislocations. Addison-Wesley, Reading, Massachusetts.
Friedlander G. and Kennedy J. W. (1956). Nuclear and Radiochemistry. New York: John Wiley.
Friedlander G., Kennedy J. W., Macias E. S., and Miller J. M. (1981). Nuclear and Radiochemis-
try. New York: John Wiley.
Frondel C. F. (1962). The System of Mineralogy of James Dwight Dana and Edward Salysbury
Dana. 7th ed., vol. 8, Silica Minerals, New York: John Wiley.
Fujii T. (1977). Fe-Mg partitioning between olivine and spinel. Carnegie Inst. Wash. Yb.,
76:563–569.
Fujino K., Sasaki S., Takeuchi Y., and Sadanaga R. (1981). X-ray determination of electron
distribution in forsterite, faialite and tephroite. Acta Cryst., B37:513–518.
Fumi F. G. and Tosi M. P. (1957). Naor relations between Madelung constants for cubic ionic
lattices. Phil. Mag., 2:284–285.
Fumi F. G. and Tosi M. P. (1964). Ionic sizes and Born repulsive parameters in the NaCl-type
alkali halides, I: The Huggins-Mayer and Pauling forms. J. Phys. Chem. Solids, 25:31–43.
Fung P. C. and Shaw D. M. (1978). Na, Rb and Tl distributions between phlogopite and sani-
dine by direct synthesis in a common vapour phase. Geochim. Cosmochim. Acta,
42:703–708.
Fursenko B. A. (1981). High pressure synthesis of the chromium bearing garnets
Mn3Cr2Si3O12. Doklady Acad. Sci. USSR, 250:176–179.
Fyfe W. S., Turner F. J., and Verhoogen J. (1958). Metamorphic reactions and metamorphic
facies. Geol. Soc. Amer. Mem., vol. 75, 253 pp.
Gabis V., Abelard P., Ildefonse J. P., Jambon A., and Touray J. C. (1979). La diffusion a haute
temperature dans les systems d’interet geologique. In Haute Temperatures et Science
de la Terre. Centre Regional de Publications de Toulouse, Editions du C.N.R.S., 163–
174.
Ganguly J. (1973). Activity-composition relation of jadeite in omphacite pyroxenes: Theoretical
deductions. Earth Planet. Sci. Letters, 19:145–153.
Ganguly J. (1976). The energetics of natural garnet solid solution, II: Mixing of the calcium
silicate end members. Contrib. Mineral. Petrol., 55:81–90.
830 / References
Ganguly J. (1977). Crystal chemical aspects of olivine structures. N. Jahrb Miner. Abh.,
130:303–318.
Ganguly J. (1979). Garnet and clinopyroxene solid solutions and geothermometry based on
Fe-Mg distribution coefficient. Geochim. Cosmochim. Acta, 43:1021–1029.
Ganguly J. and Cheng W. (1994). Thermodynamics of (Ca,Mg,Fe,Mn)-garnet solid solution:
New experiments, optimized data set, and applications to thermo-barometry. I.M.A.,
16th General Meeting (Abstracts)., Pisa, Italy.
Ganguly J. and Kennedy G. C. (1974). The energetics of natural garnet solid solutions, I: Mix-
ing of the aluminosilicate end members. Contrib. Mineral. Petrol., 48:137–148.
Ganguly J. and Ruitz J. (1986). Time-temperature relation of mineral isochrons: A thermody-
namic model, and illustrative examples for the Rb-Sr system. Earth Planet. Sci. Let-
ters, 81:338–348.
Ganguly J. and Saxena S. K. (1984). Mixing properties of aluminosilicate garnets: Constraints
from natural and experimental data and applications to geothermo-barometry. Amer.
Mineral., 69:79–87.
Ganguly J. and Saxena S. K. (1987). Mixtures and Mineral Reactions. Berlin-Heidelberg-New
York: Springer-Verlag.
Ganguly J., Cheng W., and O’Neill H. St. C. (1993). Syntheses, volume, and structural changes
of garnets in the pyrope-grossular join: Implications for stability and mixing properties.
Amer. Mineral., 78:583–593.
Garrels R. M. and Christ C. L. (1965). Solutions, Minerals, and Equilibria. New York: Harper
and Row.
Garrels R. M. and Thompson M. E. (1962). A chemical model for seawater at 25°C and one
atmosphere total pressure. Amer. Jour. Sci., 260:57–66.
Gartner L. (1979). Relations entre enthalpies en enthalpies libres de formation des ions, des oxydes
de formule MnNmO2. Utilization des frequences de vibration dans l’infra-rouge Doct. Ing.
Université de Strasbourg, 193 pp.
Gasparik T. (1985). Experimentally determined compositions of diopside-jadeite pyroxene in
equilibrium with albite and quartz at 1200–1350 °C and 15–34 kbar. Geochim. Cos-
mochim. Acta, 89:865–870.
Gasparik T. and Lindsley D. H. (1980). Phase equilibria at high pressure of pyroxenes con-
taining monovalent and trivalent ions. In Reviews in Mineralogy, vol. 7, P. H. Ribbe
(series ed.), Mineralogical Society of America.
Gasperin M. (1971). Crystal structure of rubidium feldspar, RbAlSi3O8. Acta Cryst.,
B27:854–855.
Gasperin J. and McConnel J. D. (1984). Experimental delineations of the C1̄–I1̄ trasformation
in intermediate plagioclase feldspar. Amer. Mineral., 69:112–121.
Gast P. W. (1968). Trace element fractionation and the origin of tholeiitic and alkaline magma
types. Geochim. Cosmochim. Acta, 32:1057–1086.
Gautschi W. (1964). Error function and Fresnel integrals. In Handbook of Mathematical Func-
tions, M. Abramowitz and I. A. Stegun, eds. National Bureau of Standards, Wash-
ington.
Geiger C. A., Newton R. C. and Kleppa O. J. (1987). Enthalpy of mixing of synthetic
almandine-grossular and almandine-pyrope garnets from high-temperature solution
calorimetry. Geochim. Cosmochim. Acta, 51:1755–1763.
Gelinas L., Brooks C., and Trzcienski W. E. Jr. (1976). Archean variolites-quenched immiscible
liquids reexamined: A reply to criticisms. Can. Jour. Earth Sci., 14:2945–2958.
Gerlach T. M. and Nordlie B. E. (1975a). The C-O-H-S- gaseous systems, Part I: Composition
limits and trends in basaltic cases. Amer. Jour. Sci., 275:353–376.
Gerlach T. M. and Nordlie B. E. (1975b). The C-O-H-S-gaseous systems, Part II: Temperature,
atomic composition, and molecular equilibria in volcanic gases. Amer. Jour. Sci.,
275:377–394.
Gerlach T. M. and Nordlie B. E. (1975c). The C-O-H-S-gaseous systems, Part III: Magmatic
References / 831
gases compatible with oxides and sulfides in basaltic magmas. Amer. Jour. Sci.,
275:395–410.
Ghent E. D. (1976). Plagioclase-garnet-Al2SiO5-quartz: A potential geobarometer-
geothermometer. Amer. Mineral., 61:710–714.
Ghent E. D., Stout M. Z., Black P. M., and Brothers R. N. (1987). Chloritoid bearing rocks
associated with blueschists and eclogites, northern New California. J. Met. Geol.,
5:239–254.
Ghiorso M. S. (1984). Activity/composition relation in the ternary feldspars. Contrib. Mineral.
Petrol., 87:282–296.
Ghiorso M. S. and Carmichael I. S. E. (1980). A regular solution model for meta-aluminous
silicate liquids: Applications to geothermometry, immiscibility, and the source regions
of basic magma. Contrib. Mineral. Petrol., 71:323–342.
Ghiorso M. S., Carmichael I. S. E., Rivers M. L., and Sack R. O. (1983). The Gibbs free energy
of mixing of natural silicate liquids; an expanded regular solution approximation for the
calculation of magmatic intensive variables. Contrib. Mineral. Petrol., 84:107–145.
Ghose S. (1982). Subsolidus reactions and microstructures in amphiboles. In Reviews in Miner-
alogy, vol. 9A, P. H. Ribbe (series ed.), Mineralogical Society of America.
Ghose S. and Wan C. (1974). Strong site preference of Co2⫹ in olivine Co1.10Mg0.90SiO4. Con-
trib. Mineral. Petrol., 47:131–140.
Ghose S. and Weidner J. R. (1971). Mg2⫹-Fe2⫹ isotherms in cummingtonites at 600 and 700
°C (abstract). Trans. Amer. Geophys. Union, 52:381.
Ghose S. and Weidner J. R. (1972). Mg2⫹-Fe2⫹ order-disorder in cummingtonite (Mg,Fe)7Si8-
O22(OH)2. A new geothermometer. Earth Planet. Sci. Letters, 16:346–354.
Ghose S., Choudhury N., Chaplot S. L., and Rao K. R. (1992). Phonon density of states and
thermodynamic properties of minerals. In Advances in Physical geochemistry, vol. 10,
S. K. Saxena (series ed.). Berlin-Heidelberg-New York: Springer-Verlag.
Ghose S., Kersten M., Langer K., Rossi G. and Ungaretti L. (1986). Crystal field spectra and
Jahn-Teller effect of Mn3⫹ in clinopyroxene and clinoamphiboles from India. Phys.
Chem. Minerals, 13:291–305.
Gibbs J. W. (1906). On the equilibrium of heterogeneous substances. In The Scientific Papers of
J. Willard Gibbs. Vol. 1: Thermodynamics. Longmans Green (reprinted by Dover, 1961).
Giggenbach W. F. (1980). Geothermal gas equilibria. Geochim. Cosmochim. Acta, 44:2021–
2032.
Giggenbach W. F. (1987). Redox processes governing the chemistry of fumarolic gas discharges
from White Islands, New Zealand. Appl. Geochem., 2:143–161.
Gilbert M. C., Helz R. T., Popp R. K., and Spear F. S. (1982). Experimental studies of amphi-
bole stability. In Reviews in Mineralogy, vol. 9B, P. H. Ribbe (series ed.), Mineralogical
Society of America.
Giletti B. J., Sennet M. P., and Yund R. A. (1978). Studies in diffusion, III: Oxygen in feld-
spars—an ion microprobe determination. Geochim. Cosmochim. Acta, 42:45–57.
Glasser F. P. (1960). Einige Ergebnisse von Phasengleichgewwichtsuntersuchungen in der Sys-
temen MgO-MnO-SiO2 und CaO-MnO-SiO2. Silikattechnik, 11:362–363.
Godwin H. (1962). Half-life of radiocarbon. Nature, 195:984.
Goldman D. S. and Albee A. L. (1977). Correlation of Mg/Fe partitioning between garnet and
biotite with 18O/16O partitioning between quartz and magnetite. Amer. Jour. Sci.,
277:750–767.
Goldschmidt V. M. (1923). Geochemical laws of the distribution of the elements. Videnskaps.
Skrift. Mat.-Natl. Kl., 3:1–17.
Goldschmidt V. M., Barth T., Lunde G., and Zachariasen W. (1926). Geochemische Verteilung-
sgesetze der Elemente, VII: Die Gesetze die Kristallchemie. Det. Norske. Vid. Akad.
Oslo I, Mat. Natl. Kl., 2:1–117.
Goldsmith J. R. (1980). The melting and breakdown reactions of anorthite at high pressures
and temperatures. Amer. Mineral., 66:1183–1188.
832 / References
Goldsmith J. R. and Heard H. C. (1961). Subsolidus phase relations in the system CaCO3-
MgCO3. J. Geol., 69:45–74.
Goldsmith J. R. and Newton R. C. (1969). P-T-X relations in the system CaCO3-MgCO3 at
high temperatures and pressures. Amer. Jour. Sci., 267A:160–190.
Gordon R. G. and Kim Y. S. (1971). Theory of the forces between closed-shell atoms and mole-
cules. J. Chem. Phys., 56:3122–3133.
Gordy W. (1946). A new method of determining electronegativity from other atomic properties.
Phys. Rev., 69:604–607.
Gordy W. (1950). Interpretation of nuclear quadrupole couplings in molecules. J. Chem.
Phys., 19:792–793.
Gottardi G. (1972). I Minerali. Serie Geologia, Boringhieri, Bologna.
Gramaccioli C. M. and Filippini G. (1983). Lattice dynamical calculations for orthorhombic
sulfur: A non-rigid molecular model. Chem. Phys. Letters, 108:585–588.
Grandstaff D. E. (1977). Some kinetics of bronzite orthopyroxene dissolution. Geochim. Cos-
mochim. Acta, 41:1097–1103.
Grandstaff D. E. (1980). The dissolution rate of forsterite olivine from Hawaiian beach sand.
In Third International Symposium on Water-Rock Interaction: Proceedings, Alberta Re-
search Council, Edmonton.
Graves J. (1977). Chemical mixing in multicomponent solutions. In Thermodynamics in Geol-
ogy, D. E. Fraser, ed. Reidel, Dordrecht-Holland.
Grebenshchikov R. G., Romanov D. P., Sipovskii D. P., and Kosulina G. I. (1974). Study of a
germanate hydroxyamphibole. Zhur. Prikl. Khimii, 47:1905–1910.
Greegor R. B., Lytle F. W., Sanstrom D. R., Wang J., and Schultz P. (1983). Investigation of
TiO2-SiO2 glasses by X-ray absorption spectroscopy. J. Non. Cryst. Solids, 55:27–43.
Green D. H. and Hibberson W. (1970). The instability of plagioclase on peridotite at high pres-
sure. Lithos, 3:209–221.
Green E. J. (1970). Predictive thermodynamic models for mineral systems, I: Quasi-chemical
analysis of the halite-sylvite subsolidus. Amer. Mineral., 55:1692–1713.
Green S. H. and Gordon R. G. (1974). POTLSURF: A program to compute the interaction
potential energy surface between a closed-shell molecule and an atom. Quantum Chemis-
try Program Exchange No. 251, Indiana University.
Green T. H. (1994). Experimental studies of trace-element partitioning applicable to igneous
petrogenesis—Sedona 16 years later. Chem. Geol., 117:1–36.
Greenwood H. J. (1967). Wollastonite: Stability in H2O-CO2 mixtures and occurrence in a con-
tact metamorphic aureole near Salmo, British Columbia, Canada. Amer. Mineral.,
52:1669–1680.
Greenwood N. N. (1970). Ionic Crystals, Lattice Defects and Non-stoichiometry. New York:
Chemical Pub. Co.
Greig J. W. (1927). Immiscibility in silicate melts. Amer. Jour. Sci., 5th Ser., 13(73).: 1–44 and
13(74).: 133–154.
Greig J. W. and Barth T. F. W. (1938). The system Na2O-Al2O3-2SiO2 (nepheline, carnegieite).-
Na2O-Al2O3-6SiO2. (albite). Amer. Jour. Sci., 5th. ser., 35A:93–112.
Griffin W. L. and Mottana A. (1982). Crystal chemistry of clinopyroxenes form St. Marcel
manganese deposit, Val d’Aosta, Italy. Amer. Mineral., 67:568–586.
Grover J. (1977). Chemical mixing in multicomponent solutions: An introduction to the use of
Margules and other thermodynamic excess functions to represent non-ideal behaviour.
In Thermodynamics in Geology, D. G. Fraser, ed. D. Reidel, Dordrecht-Holland.
Grover J. E. and Orville T. M. (1969). The partitioning of cations between coexisting single
and multisite phases with application to the assemblage: Orthopyroxene-clinopyroxene
and orthopyroxene-olivine. Geochim. Cosmochim. Acta, 33:205–226.
Grundy H. D. and Ito J. (1974). The refinement of the crystal structure of a synthetic non-
stoichiometric Sr-feldspar. Amer. Mineral., 68:1319–1326.
Guggenheim E. A. (1937). Theoretical basis of Raoult’s Law. Trans. Faraday Soc., 33:205–226.
References / 833
Harrison T. M. and McDougall I. (1982). The thermal significance of potassium feldspar K-Ar
ages inferred from 40Ar/39Ar age spectrum results. Geochim. Cosmochim. Acta, 46:1811–
1820.
Harrison W. J. (1977). An experimental study of the partitioning of samarium between garnet
and liquid at high pressures. In Papers Presented to the International Conference on Ex-
perimental Trace Elements Geochemistry, Sedona, Arizona.
Harrison W. J. (1978). Rare earth element partitionings between garnets, pyroxenes and melts
at low trace element concentration. Carnegie Inst. Wash. Yb., 77:682–689.
Harrison W. J. (1981). Partition coefficients for REE between garnets and liquids: Implications
of non-Henry’s law behavior for models of basalt origin and evolution. Geochim. Cos-
mochim. Acta, 45:1529–1544.
Harrison W. J. and Wood B. J. (1980). An experimental investigation of the partitioning of REE
between garnet and liquid with reference to the role of defect equilibria. Contrib. Min-
eral. Petrol., 72:145–155.
Hart R. (1970). Chemical exchange between seawater and deep ocean basalts. Earth. Planet.
Sci. Letters., 9:269–279.
Hart S. R. (1964). The petrology and isotopic-mineral age relations of a contact zone in the
Front Range, Colorado. J. Geol., 72:493–525.
Harvey K. B. and Porter G. B. (1976). Introduzione alla Chimica Fisica Inorganica. Piccin Edi-
tore, Padova.
Haselton H. T. and Newton R. C. (1980). Thermodynamics of pyrope-grossular garnets and
their stabilities at high temperatures and pressures. J. Geophys. Res., 85:6973–6982.
Haselton H. T. and Westrum E. F. Jr. (1980). Low-temperature heat capacities of synthetic
pyrope, grossular, and pyrope60grossular40. Geochim. Cosmochim. Acta, 44:701–709.
Haselton H. T., Hovis G. L., Hemingway B. S., and Robie R. A. (1983). Calorimetric investiga-
tion of the excess entropy of mixing in analbite-sanidine solid solutions: Lack of evi-
dence for Na, K short range order and implications for two feldspar thermometry. Amer.
Mineral., 68:398–413.
Hawthorne F. C. (1976). The crystal chemistry of the amphiboles, V: The structure and chemis-
try of arfvedsonite. Canadian Mineral., 346–356.
Hawthorne F. C. (1981a). Crystal chemistry of amphiboles. In Reviews in Mineralogy, vol. 9A,
P. H. Ribbe (series ed.), Mineralogical Society of America.
Hawthorne F. C. (1981b). Some systematics of the garnet structure. J. Solid State Chem.,
37:157–164.
Hawthorne F. C. and Grundy H. D. (1976). The crystal chemistry of the amphiboles, IV: X-ray
and neutron refinement of the crystal structure of tremolite. Canadian Mineral.,
14:334–345.
Hazen R. M. (1977). Effects of temperature and pressure on the crystal structure of ferromag-
nesian olivine. Amer. Mineral., 62:286–295.
Hazen R. M. and Burnham C. W. (1973). The crystal structure of 1 layer phlogopite and annite.
Amer. Mineral., 58:889–900.
Hazen R. M. and Finger L. W. (1978). Crystal structure and compressibilities of pyrope and
grossular to 60 kbar. Amer. Mineral., 63:297–303.
Hazen R. M. and Finger L. W. (1979). Bulk-modulus volume relationship for cation-anion
polyhedra. J. Geophys. Res., 84:6723–6728.
Hazen R. M. and Finger L. W. (1982). Comparative Crystal Chemistry. New York: John Wiley.
Hazen R. M. and Prewitt C. T. (1977). Effects of temperature and pressure on interatomic
distances in oxygen-based minerals. Amer. Mineral., 62:309–315.
Hazen R. M. and Wones D. R. (1972). The effect of cation substitutions on the physical proper-
ties of trioctahedral micas. Amer. Mineral., 57:103–129.
Helgeson H. C. (1967). Solution chemistry and metamorphism. In Research in Geochemistry,
vol. 2, P. H. Abelson (series ed.). New York: John Wiley.
Helgeson H. C. (1969). Thermodynamics of hydrothermal system at elevated temperatures and
pressures. Amer. Jour. Sci., 267:729–804.
References / 835
oxygen and hydrogen isotopes in aqueous salt solutions, I: Vapor-liquid water equilibra-
tion of single salt solutions from 50 to 100°C. Geochim. Cosmochim. Acta, 57:2797–2817.
Horita J., Cole D. R., and Wesolowski D. J. (1993b). The activity-composition relationship of
oxygen and hydrogen isotopes in aqueous salt solutions, II: Vapor-liquid water equili-
bration of mixed salt solutions from 50 to 100°C and geochemical implications. Geo-
chim. Cosmochim. Acta, 57:4703–4711.
Hovis G. L. (1974). A solution calorimetric and X-ray investigation of Al-Si distribution in
monoclinic potassium feldspars. In The feldspars, W. S. MacKenzie and J. Zussman, eds.
Manchester: Manchester University Press.
Hovis G. L. and Waldbaum D. R. (1977). A solution calorimetric investigation of K-Na mixing
in a sanidine-analbite join-exchange series. Amer. Mineral., 53:1965–1979.
Hsu L. C. (1968). Selected phase relationships in the system Al-Mn-Fe-Si-O-H: A model for
garnet equilibria. J. Petrol., 9:40–83.
Huckenholtz H. G. and Knittel D. (1975). Uvarovite: Stability of uvarovite-grossularite solid
solutions at low pressure. Contrib. Mineral. Petrol., 49:211–232.
Huckenholtz H. G., Schairer J. F., and Yoder H. S. Jr. (1969). Synthesis and stability of ferri-
diopside. Mineral. Soc. Amer. Spec. Paper, 2:163–177.
Huebner J. S. (1982). Phyroxene phase equilibria at low pressure. In Reviews in Mineralogy, vol.
7 (2d ed.), P. H. Ribbe (series ed.), Mineralogical Society of America.
Huebner J. S. and Papike J. J. (1970). Synthesis and crystal chemistry of sodium-potassium
richterite, (Na, K)NaCaMg5Si8O22(OH,F)2. Amer. Mineral., 55:1973–1993.
Huebner J. S. and Sato M. (1970). The oxygen fugacity-temperature relationships of manganese
oxide and nickel oxide buffers. Amer. Mineral., 55:934–952.
Huebner J. S., Lipin B. R., and Wiggins L. B. (1976). Partitioning of chromium between silicate
crystal and melts. Proc. Seventh Lunar Sci. Conf., 1195–1220.
Huheey J. E. (1975). Inorganic Chemistry: Principles of Structure and Reactivity. New York:
Harper and Row.
Hurd D. C. (1972). Factors affecting solution rate of biogenic opal in sea water. Earth Planet.
Sci. Letters, 15:411–417.
Hutner R. A., Rittner E. S., and Du Pré F. K. (1949). Concerning the work of polarization in
ionic crystals of the NaCl type, II: Polarization around two adjacent charges in the rigid
lattice. J. Chem. Phys., 17:204–208.
Iczkowski R. P. and Margrave J. L. (1961). Electronegativity. J. Amer. Chem. Soc., 83: 3547–
3551.
Iiyama J. T. (1972). Fixation des elements alcalinoterreux, Ba, Sr et Ca dans les feldspaths;
etude experimentale. Proc. 24 Congres Geol. Internat., Section 10, 122–130.
Iiyama J. T. (1973). Behavior of trace elements in feldspars under hydrothermal conditions. In
The Feldspars, W. S. MacKenzie and J. Zussmann, eds. Manchester: Manchester Univer-
sity Press.
Iiyama J. T. (1974). Substitution, deformation locale de la maille et equilibre de distribution
des elements en tracs entre silicates et solution hydrothermale. Bul. Soc. Fr. Mineral.
Cristallog., 97:143–151.
Irefune T., Ohtani E., and Kumazawa I. (1982). Stability field of knorringite at high pressure
and its application to the occurrence of Cr-rich pyrope in the upper mantle. Phys. Earth
Planet. Interiors, 27:263–272.
Irving A. J. (1974). Geochemical and high-pressure experimental studies of garnet pyroxenite
and granulite xenoliths from the Delegate basaltic pipes, Australia. J. Petrol., 15:1–40.
Irving A. J. (1978). A review of experimental studies of crystal/liquid trace element parti-
tioning. Geochim. Cosmochim. Acta. 42:743–770.
Isaak D. G. and Graham E. K. (1976). The elastic properties of an almandine-spessartite gar-
net and elasticity in the garnet solid solutions series. J. Geophys. Res., 81:2483–2489.
Isaak D. G., Anderson L., and Oda H. (1992). High-temperature thermal expansion and elas-
ticity of calcium-rich garnets. Phys. Chem. Minerals, 19:106–120.
838 / References
Ivanov I. P., Potekhin V. Y., Dmitriyenko L. T., and Beloborodov S. M. (1973). An experimen-
tal study of T and P conditions of equilibrium of reaction: Muscovite-Kfeldspar ⫹ co-
rundum ⫹ H2O at P(H2O)⬍P(total). Geokhimiya, 9:1300–1310.
Jäger E. and Hunziker J. C. (1979). Lectures in Isotope Geology Berlin-Heidelberg-New York:
Springer-Verlag.
Jäger E., Niggli E., and Wenk E. (1967). Alterbestimmungen an Glimmern der Zentralalpen.
Beitr. Geol. Karte Schweitz, NF134, Bern.
Jahn H. A. and Teller E. (1937). Stability of polyatomic molecules in degenerate electronic
states. Proc. Roy. Soc. London, A161:220–235.
JANAF (1974–1975). Thermochemical Tables U.S. Department of Commerce, National Bureau
of Standards, Institute for Applied Technology, Supplements.
Jaoul O., Froidevaux C., Durham W. B., and Michaud M. (1980). Oxygen self-diffusion in
forsterite: Implications for the high temperature creep mechanism. Earth. Planet. Sci.
Letters, 47:391–397.
Jaoul O., Poumellec M., Froidevaux C., and Havette A. (1981). Silicon diffusion in forsterite:
A new constraint for understanding mantle deformation. In Anelasticity in the Earth,
F. D. Stacey et al., eds. Godyn Ser., vol. 4, AGU, Washington, D.C.
Jenne E. A. (1981). Geochemical modeling: A review. In Waste/Rock Interactions Technology
Program, Pacific Northwest Laboratory, Publication 3574.
Jenne E. A., Girovin D. C., Ball J. W., and Burchard J. M. (1978). Inorganic speciation of silver
in natural waters-fresh to marine. In Environmental Impacts of Nucleating Agents Used
in Weather Modification Programs, D. A. Klein, ed. Dowden-Hutchison and Ross, Stras-
burg, Pennsylvania.
Jensen B. (1973). Patterns of trace elements partitioning. Geochim. Cosmochim. Acta, 37:2227–
2242.
Johnson J. W. and Norton D. (1991). Critical phenomena in hydrothermal systems: State, ther-
modynamic, electrostatic, and transport properties of H2O in the critical region. Amer.
Jour. Sci., 291:541–648.
Johnson J. W., Oelskers E. H., and Helgeson H. C. (1991). SUPCRT92: A Software Package for
Calculating the Standard Molal Thermodynamic Properties of Minerals, Gases, Aqueous
Species, and Reactions from 1 to 5000 bars and 0° to 1000°C Earth Sciences Department,
L-219, Lawrence Livermore National Laboratory, Livermore, California.
Johnston W. D. (1964). Oxidation-reduction equilibria in iron-containing glass. J. Amer. Ceram.
Soc., 47:198–201.
Johnston W. D. (1965). Oxidation-reduction equilibria in molten Na2O-2SiO2 glass. J. Amer.
Ceram Soc., 48:184–190.
Johnston W. D., and Chelko A. (1966). Oxidation-reduction equilibria in molten Na2O-2SiO2
glass in contact with metallic copper and silver. J. Amer. Ceram. Soc., 49:562–564.
Jones J. H. (1995). Experimental trace element partitioning. In Rock Physics and Phase Rela-
tions. A Handbook of Physical Constants, T. J. Ahrens, ed., AGU, Washington, D.C.
Jost W. (1960). Diffusion in Solid State Physics. New York: Academic Press.
Kamensky I. L., Tokarev I. V., and Tolstikhin I. N. (1991). 3H-3He dating: A case for mixing
of young and old groundwaters. Geochim. Cosmochim. Acta, 55:2895–2899.
Karpinskaya T. B., Ostrovsky I. A., and Yevstigneyeva O. (1983). Synthetic pure iron garnet
skiagite. Internat. Geol. Rev., 25:1129–1130.
Kato K. and Nukui A. (1976). Die Kristallstructur des monoclinen Tief-trydymits. Acta Cryst.,
B32:2486–2491.
Katz A. and Matthews A. (1977). The dolomitization of CaCO3: An experimental study at
252–295°C. Geochim. Cosmochim. Acta, 41:297–308.
Kawasaki T. and Matsui Y. (1977). Partitioning of Fe2⫹ and Mg2⫹ between olivine and garnet.
Earth. Planet. Sci. Letters, 37:159–166.
Kelley K. K. (1960). Contributions to the data on theoretical metallurgy, XIII: High tempera-
ture heat content, heat capacity and entropy data for the elements and inorganic com-
pounds. U.S. Bur. Mines Bull., 584, 232 pp.
References / 839
Lindsley D. H. and Anderson D. J. (1982). A two pyroxene thermometer. Proc. 13th Lunar
Planet. Sci. Conf., 887–906.
Lindsley D. H. and Dixon S. A. (1976). Diopside-enstatite equilibria at 850 to 1400°C, 5 to 35
kbar. Amer. Jour. Sci., 276:1285–1301.
Lindsley D. H., Grover J. E., and Davidson P. M. (1981). The thermodynamics of the
Mg2Si2O6-CaMgSi2O6 join: A review and a new model. In Advances in Physical Geo-
chemistry, vol. 1, S. K. Saxena (series ed.), New York: Springer-Verlag.
Lindstrom D. J. (1976). Experimental study of the partitioning of the transition metals between
clinopyroxene and coexisting silicate liquids. Ph.D. diss., University of Oregon.
Lindstrom D. J. and Weill D. F. (1978). Partitioning of transition metals between diopside and
coexisting silicate liquids, I: Nickel, cobalt, and manganese. Geochim. Cosmochim.
Acta, 42:801–816.
Liu K. K. and Epstein S. (1984). The hydrogen isotope fractionation between kaolinite and
water. Chem. Geol. (Isot. Geosci. Sec.)., 2:335–350.
Liu L. and Bassett A. (1986). Elements oxides and silicates. In Oxford Monographs on Geology
and Geophysics, vol. 4, P. Allen, H. Charnock, E. R. Oxburg, and B. J. Skinner, eds. New
York: Oxford University Press.
Loewenstein W. (1954). The distribution of aluminum in the tetrahedra of silicates and alumi-
nates. Amer. Mineral., 39:92–96.
London F. (1930). Properties and application of molecular forces. Z. Phys. Chem., B11:
222–251.
Long P. E. (1978). Experimental determination of partition coefficients for Rb, Sr and Ba be-
tween alkali feldspar and silicate liquid. Geochim. Cosmochim. Acta, 42:833–846.
Luck J. M. and Allègre C. J. (1983). 187Re-187Os systematics in meteorites and cosmological
consequences. Nature, 302:130–132.
Lumpkin G. R. and Ribbe P. H. (1983). Composition, order-disorder and lattice parameters of
olivines: Relationships in silicate, germanate, beryllate, phosphate and borate olivine.
Amer. Mineral., 68:164–176.
Lux G. (1987). The behavior of noble gases in silicate liquids: Solution, diffusion, bubbles and
surface effects, with application to natural samples. Geochim. Cosmochim. Acta,
51:1549–1560.
Lux H. Z. (1939). Acid and bases in fused salt bath: The determination of oxygen-ion concen-
tration. Z. Elektrochem., 45:303–309.
Mackrodt W. C. and Stewart R. F. (1979). Defect properties of ionic solids, II: Point defect
energies based on modified electron gas potentials. J. Phys., 12C:431–449.
Mah A. D. and Pancratz L. B. (1976). Thermodynamic properties of nickel and its inorganic
compounds. U.S. Bur. Mines Bull., 668:1–125.
Maier C. G. and Kelley K. K. (1932). An equation for the representation of high temperature
heat content data. J. Amer. Chem. Soc., 54:3243–3246.
Manacorda T. (1985). Elementi di Meccanica Ondulatoria. Istituto di Matematiche Applicate
Ulisse Dini, Pisa.
Mantoura R. C. F., Dickson A., and Riley J. P. (1978). The complexation of metals with humic
materials in natural waters. Estuarine Coastal Mar. Sci., 6:387–408.
Maresch M. V., Mirwald P. W., and Abraham K. (1978). Nachweis einer Mischlüke in der Olivi-
nreihe Forsterit (Mg2SiO4)-Tephroit (Mn2SiO4). Fortschr. Miner., 56:89–90.
Marshall B. D. and De Paolo D. J. (1982). Precise age determination and petrogenetic studies
using the K-Ca method. Geochim. Cosmochim. Acta, 46:2537–2545.
Marshall W. L. and Franck E. U. (1981). Ion product of water substance, 0–1000°C, 1–10,000
bars: New international formulation and its background. J. Phys. Chem. Ref. Data,
10:295–303.
Martin R. F. (1970). Cell parameters and infra-red absorption of synthetic high to low albites.
Contrib. Mineral. Petrol., 26:62–74.
Martin R. F. (1974). Controls of ordering and subsolidus phase relations in the alkali feldspars.
References / 843
Navrotsky A. and Akaogi M. (1984). a-b-c phase relations in Fe2SiO4-Mg2SiO4 and Co2SiO4-
Mg2SiO4: Calculation from thermochemical data and geophysical applications. J. Geo-
phys. Res., 89:10135–10140.
Navrotsky A. and Loucks D. (1977). Calculations of subsolidus phase relations in carbonates
and pyroxenes. Phys. Chem. Minerals, 1:109–127.
Navrotsky A., Geisinger K. L., McMillan P., and Gibbs G. V. (1985). The tetrahedral frame-
work in glasses and melts: Inferences from molecular orbital calculations and implica-
tions for structure, thermodynamics and physical properties. Phys. Chem. Minerals,
11:284–298.
Navrotsky A., Peraudeau G., McMillan P., and Coutures J. P. (1982). A thermochemical study
of glasses and crystals along the joins silica-calcium aluminate and silica-sodium alumi-
nate. Geochim. Cosmochim. Acta, 46:2039–2047.
Nernst W. (1891). Distribution of a substance between two solvents and between solvent and
vapor. Z. Phis. Chem., 8:110–139.
Neumann H., Mead J., and Vitaliano C. J. (1954). Trace element variation during fractional
crystallization as calculated from the distribution law. Geochim. Cosmochim. Acta,
6:90–99.
Newnham R. E. and Megaw H. D. (1960). The crystal structure of celsian (barium feldspar).
Acta Cryst., 13:303–312.
Newton R. C. (1987). Thermodynamic analysis of phase equilibria in simple mineral system.
In Reviews in Mineralogy. vol. 17, P. H. Ribbe (series ed.), Mineralogical Society of
America.
Newton R. C. and Haselton H. T. (1981). Thermodynamics of the garnet-plagioclase-Al2SiO5-
quartz geobarometer. In Thermodynamics of Minerals and Melts, R. C. Newton, A. Na-
vrotsky, and B. J. Wood, eds. New York: Springer-Verlag.
Newton R. C. and Perkins D. (1982). Thermodynamic calibration of geobarometers based on
the assemblages garnet-plagioclase-orthopyroxene (clinopyroxene)-quartz. Amer. Min-
eral., 67:203–222.
Newton R. C. and Wood B. J. (1980). Volume behavior of silicate solid solutions. Amer. Min-
eral., 65:733–745.
Newton R. C., Charlu T. V., and Kleppa O. J. (1977). Thermochemistry of high pressure gar-
nets and clinopyroxenes in the system CaO-MgO-Al2O3-SiO2. Geochim. Cosmochim.
Acta, 41:369–377.
Newton R. C., Charlu T. V., and Kleppa O. J. (1980). Thermochemistry of high structural state
plagioclases. Geochim. Cosmochim. Acta, 44:933–941.
Newton R. C., Geiger C. A., Kleppa O. J., and Brousse C. (1986). Thermochemistry of binary
and ternary garnet solid solutions. I.M.A., Abstracts with Program, 186.
Newton W. and McCready N. (1948). Thermodynamic properties of sodium silicates. C. Phys.
Coll. Chem., 52:1277–1283.
Nishizawa O. and Akimoto S. (1973). Partitioning of magnesium and iron between olivine and
spinel, and between pyroxene and spinel. Contrib. Mineral. Petrol., 41:217–240.
Nordlie B. E. (1971). The composition of the magmatic gas of Kilauea and its behavior in the
near surface environment. Amer. Jour. Sci., 271:417–473.
Novak G. A. and Colville A. (1975). A linear regression analysis of garnet chemistries versus
cell parameters (abstract). Geol. Soc. Amer. S. W. Section Meeting, Los Angeles, 7:359.
Novak G. A. and Gibbs G. V. (1971). The crystal chemistry of silicate garnets. Amer. Min-
eral., 56:791–825.
Nylen P. and Wigren N. (1971). Stechiometria CEDAM, Padova.
Oba T. (1980). Phase relations in the tremolite-pargasite join. Contrib. Mineral. Petrol.,
71:247–256.
Ohmoto H. (1972). Systematics of sulfur and carbon isotopes in hydrothermal ore deposits.
Econ. Geol., 67:551–578.
Ohmoto H. and Lasaga A. C. (1982). Kinetics of reactions between aqueous sulfates and sul-
fides in hydrothermal systems. Geochim. Cosmochim. Acta, 46:1727–1745.
References / 847
Ohmoto H. and Rye R. O. (1979). Isotopes of sulfur and carbon. In Geochemistry of Hydrother-
mal Ore Deposits, 2d ed., H. L. Barnes, ed. New York: John Wiley.
Oka Y. and Matsumoto T. (1974). Study on the compositional dependence of the apparent
partitioning coefficient of iron and magnesium between coexisting garnet and clinopyro-
xene solid solutions. Contrib. Mineral. Petrol., 48:115–121.
Okamura F. P. S., Ghose S., and Ohashi H. (1974). Structure and crystal chemistry of calcium
Tschermak’s pyroxene CaAlAlSiO6. Amer. Mineral., 59:549–557.
O’Neill H.St. C. and Wood B. J. (1979). An experimental study of Fe-Mg partitioning between
garnet and olivine and its calibration as a geothermometer. Contrib. Mineral. Petrol.,
70:59–70.
O’Neill M. J. and Fyans R. L. (1971). Design of differential scanning calorimeters and the
performance of a new system. Norwalk, Connecticut, Perkin-Elmer Corp., 38 pp.
O’Nions R. K. and Smith D. G. W. (1973). Bonding in silicates: An assessment of bonding in
orthopyroxene. Geochim. Cosmochim. Acta, 37:249–257.
Onken H. (1965). Verfeinerung der Kristallstruktur von Monticellite. Tsch. Min. Petr. Mitt.,
10:34–44.
Onuma N., Higuchi H., Wakita H., and Nagasawa H. (1968). Trace element partition between
two pyroxenes and host volcanic rocks. Earth Planet. Sci. Letters, 5:47–51.
Orville P. M. (1972). Plagioclase cation exchange equilibria with aqueous chloride solution:
Results at 70°C and 2000 bars in the presence of quartz. Amer. Jour. Sci., 272:234–272.
Osborn E. F. (1942). The system CaSiO3-diopside-anorthite. Amer. Jour. Sci., 240:751–788.
Osborn E. F. and Tait D. B. (1952). The system diopside-forsterite-anorthite. Amer. Jour. Sci.,
Bowen Volume, 413–434.
Ottonello G. (1980). Rare earth abundance and distribution in some spinel peridotite xenoliths
from Assab (Ethiopia). Geochim. Cosmochim. Acta, 44:1885–1901.
Ottonello G. (1983). Trace elements as monitors of magmatic processes (1) limits imposed by
Henry’s law problem and (2) Compositional effect of silicate liquid. In The Significance
of Trace Elements in Solving Petrogenetic Problems and Controversies, S. Augusthitis, ed.
Theophrastus Publications, Athens.
Ottonello G. (1986). Energetics of multiple oxides with spinel structure. Phys. Chem. Miner-
als., 13:79–90.
Ottonello G. (1987). Energies and interactions in binary (Pbnm) orthosilicates: A Born parame-
trization. Geochim. Cosmochim. Acta, 51:3119–3135.
Ottonello G. (1992). Interactions and mixing properties in the (C2/c) clinopyroxene quadrilat-
eral. Contrib. Mineral. Petrol., 111:53–60.
Ottonello G. and Morlotti R. (1987). Thermodynamics of nickel-magnesium olivine solid solu-
tion. J. Chem. Thermodyn., 19:809, 818.
Ottonello G. and Ranieri G. (1977). Effetti del controllo petrogenetico sulla distribuzione del
Ba nel processo di anatessi crustale. Rend. S.I.M.P., 33:741–753.
Ottonello G., Bokreta M. and Sciuto P. F. (1996). Parameterization of energy and interactions
in garnets: End-member properties. Amer. Mineral., 81:429–447.
Ottonello G., Bokreta M., and Sciuto P. F. (in preparation). Parameterization of energy and
interactions in garnets: Mixing properties.
Ottonello G., Della Giusta A., and Molin G. M. (1989). Cation ordering in Ni-Mg olivines.
Amer. Mineral., 74:411–421.
Ottonello G., Piccardo G. B. and Ernst W. G. (1979). Petrogenesis of some Ligurian peridotites,
II: Rare earth element chemistry. Geochim. Cosmochim. Acta, 43:1273–1284.
Ottonello G., Princivalle F., and Della Giusta A. (1990). Temperature, composition and f O2
effects on intersite distribution of Mg and Fe2⫹ in olivines. Phys. Chem. Minerals,
17:301–312.
Ottonello G., Della Giusta A., Dal Negro A., and Baccarin F. (1992). A structure-energy model
for C2/c pyroxenes in the system Na-Mg-Ca-Mn-Fe-Al-Cr-Ti-Si-O. In Advances in
Physical Geochemistry, S. K. Saxena (series ed.), vol. 10. Berlin-Heidelberg-New York:
Springer-Verlag.
848 / References
Sr and Ba, with application to anorthosite and basalt genesis. Geochim. Cosmochim.
Acta, 34:307–322.
Piccardo G. B. and Ottonello G. (1978). Partial melting effects on coexisting mineral com-
positions in upper mantle xenoliths from Assab (Ethiopia). Rend. S.I.M.P., 34:499–526.
Pitzer K. S. (1973). Thermodynamics of electrolytes. I Theoretical basis and general equations.
J. Phys. Chem., 77:268–277.
Pitzer K. S. (1983). Dieletric constants of water at very high temperature and pressure. Proc.
Natl. Acad. Sci. U.S.A., 80:4575–4576.
Plummer L. N., Wigley T. M. L. and Parkhurst D. L. (1978). The kinetics of calcite dissolution
in CO2-water system at 5° to 60°C and 0.0 to 1.0 atm CO2. Amer. Jour. Sci., 278:179–216.
Pluschkell W. and Engell H. J. (1968). Ionen und Elektronenleitung in Magnesiumorthosilikat.
Ber. Dtsch. Keram. Ges., 45:388–394.
Poldervaart A. and Hess H. H. (1951). Pyroxenes in the crystallization of basaltic magmas.
J. Geol., 59:472–489.
Pople J. A. (1951). The structure of water and similar molecules. Proc. Roy. Soc., A202:323–336.
Popp R. K., Gilbert M. C. and Craig J. R. (1976). Synthesis and X-ray properties of Fe-Mg
orthoamphiboles. Amer. Mineral., 61:1267–1279.
Potter R. W. and Clynne M. A. (1978). The solubility of noble gases He, Ne, Ar, Kr and Xe in
water up to the critical point. J. Soln. Chem., 7:837–844.
Pourbaix M. (1966). Atlas of Electrochemical Equilibria in Aqueous Solutions Pergamon Press,
Oxford.
Powell M. and Powell R. (1977a). Plagioclase-alkali feldspar thermometry revisited. Min.
Mag., 41:253–256.
Powell R. and Powell M. (1977b). Geothermometry and oxygen barometry using iron-titanium
oxides: A reappraisal. Min. Mag., 41:257–263.
Powenceby M. I., Wall V. J. and O’Neill H. St. C. (1987). Fe-Mn partitioning between garnet
and ilmenite: Experimental calibration and applications. Contrib. Mineral. Petrol.,
97:116–126.
Prausnitz J. M., Lichtenhaler R. N., and De Azvedo E. G. (1986). Molecular Thermodynamics
of Fluid Phase Equilibria. New York: Prentice-Hall.
Presnall D. C. (1969). The geometrical analysis of partial fusion. Amer. Jour. Sci., 267:1178–
1194 Prewitt C. T., Papike J. J. and Ross M. (1970). Cummingtonite. A reversible, non-
quenchable transition from P21/m to C2/m symmetry. Earth. Planet. Sci. Letters,
8:448–450.
Prigogine I. (1955). Introduction to Thermodynamics of Irreversible Processes. 2d ed. New York:
Interscience Publishers.
Provost A. and Bottinga Y. (1974). Rates of solidification of Apollo 11 basalts and hawaiian
tholeiite. Earth. Planet. Sci. Letters, 15:325–337.
Putnis A. and McConnell J. D. C. (1980). Principles of Mineral Behavior. Blackwell Scientific
Publications, Oxford.
Rae A. I. M. (1973). A theory for the interactions between closed shell systems. Chem. Phys.
Lett., 18:574–577.
Raheim A. (1975). Mineral zoning as a record of P, T history of Precambrian metamorphic
rocks in W. Tasmania. Lithos, 8:221–236.
Raheim A. (1976). Petrology of eclogites and surrounding schists from the Lyell Highway–
Collingwood River area, W. Tansmania. Geol. Soc. Austr. Jour., 23:313–327.
Raheim A. and Green D. H. (1974). Experimental determination of the temperature and pres-
sure dependence of the Fe-Mg partition coefficient for coexisting garnet and clinopyro-
xene. Contrib. Mineral. Petrol., 48:179–203.
Raheim A. and Green D. H. (1975). P, T paths of natural eclogites during metamorphism, a
record of subduction. Lithos, 8:317–328.
Rakai R. J. (1975). Crystal structure of spessartite and andradite at elevated temperatures.
M.Sc. diss., University of British Columbia.
850 / References
Ramberg H. (1952). Chemical bonds and the distribution of cations in silicates. J. Geol.,
60:331–355.
Ramberg H. and Devore G. (1951). The distribution of Fe2⫹ and Mg in coexisting olivines and
pyroxenes. J. Geol., 59:193–210.
Rammensee W. and Fraser D. G. (1982). Determination of activities in silicate melts by Knud-
sen cell mass spectrometry, I: The system NaAlSi3O8-KAlSi3O8. Geochim. Cosmochim.
Acta, 46:2269–2278.
Rankama K. (1954). Isotope Geology Pergamon Press, London.
Ravizza G. and Turekian K. K. (1989). Application of the 187Re-187Os system to black shale
geochronometry. Geochim. Cosmochim. Acta, 53:3257–3262.
Rayleigh J. W. S. (1896). Theoretical considerations respecting the separation of gases by diffu-
sion and similar processes. Philos. Mag., 5th ser., 42:493–498.
Reddy K. P. R., Oh S. M., Major L. D. and Cooper A. R. (1980). Oxygen diffusion in forsterite.
J. Geophs. Res., 85:322–326.
Redlich O. and Kister A. T. (1948). Thermodynamics of non-electrolyte solutions: x-y-t rela-
tions in a binary system. Ind. Eng. Chem., 40:341–345.
Reed M. and Spycher N. (1984). Calculation of pH and mineral equilibria in hydrothermal
waters with application to geothermometry and studies of boiling and dilution. Geo-
chim. Cosmochim. Acta, 48:1479–1492.
Reuter J. H. and Perdue E. M. (1977). Importance of heavy metal-organic matter interactions
in natural waters. Geochim. Cosmochim. Acta, 41:325–334.
Ribbe P. H. (1983a). The chemistry, structure and nomenclature of feldspars. In Reviews in
Mineralogy, vol. 2 (2d ed.), P. H. Ribbe (series ed.), Mineralogical Society of America.
Ribbe P. H. (1983b). Aluminium-silicon order in feldspars: Domain textures and diffraction
patterns. In Reviews in Mineralogy, vol. 2 (2d ed.), P. H. Ribbe (series ed.), Mineralogical
Society of America.
Richardson F. D. (1956). Activities in ternary silicate melts. Trans. Faraday Soc., 52:1312–1324.
Richardson S. W. (1968). Staurolite stability in a part of the system Fe-Al-Si-O-H. J. Petrol.,
9:467–488.
Richardson S. W., Gilbert M. C. and Bell P. M. (1969). Experimental determination of kyanite-
andalusite and andalusite-sillimanite equilibria: The aluminum silicate triple point.
Amer. Jour. Sci., 267:259–272.
Richet P. and Bottinga Y. (1983). Verres, liquides, et transition vitreuse. Bull. Mineral.,
106:147–168.
Richet P. and Bottinga Y. (1985). Heat capacity of aluminium-free liquid silicates. Geochim.
Cosmochim. Acta, 49:471–486.
Richet P. and Bottinga Y. (1986). Thermochemical properties of silicate glasses and liquids: A
review. Rev. Geophys. Space Phys., 24:1–25.
Rickard D. T. (1975). Kinetics and mechanism of pyrite formation at low temperatures. Amer.
Jour. Sci., 275:636–652.
Rimstidt J. D. and Barnes H. L. (1980). The kinetics of silica-water reactions. Geochim. Cos-
mochim. Acta, 44:1683–1699.
Ringwood A. E. (1969). Phase transformations in the mantle. Earth Planet. Sci. Letters,
5:401–412.
Ringwood A. E. and Major A. (1970). The system Mg2SiO4-Fe2SiO4 at high pressures and
temperatures. Phys. Earth Planet. Int., 3:89–108.
Rittner E. S., Hutner R. A. and Du Pré F. K. (1949). Concerning the work of polarization in
ionic crystals of the NaCl type. II. Polarization around two adjacent charges in the rigid
lattice. J. Chem. Phys., 17:198–203.
Robie R. A., Hemingway B. S. and Fisher J. R. (1978). Thermodynamic properties of minerals
and related substances at 298.15 K and 1 bar (105 pascal) pressure and at higher temper-
atures. U.S.G.S. Bull., 1452, 456 pp.
Robie R. A., Bethke P. M., Toulmin M. S. and Edwards J. L. (1966). X-ray crystallographic
References / 851
Roux J. (1974). Etude des solutions solides des néphélines (Na,K)AlSiO4 et (Na,Rb)AlSiO4.
Geochim. Cosmochim. Acta, 38:1213–1224.
Rubey W. W. (1951). Geologic history of seawater. Geol. Soc. Amer. Bull., 62:1111–1147.
Russell W. A., Papanastassiou D. A. and Tombrello T. A. (1978). Ca isotope fractionation on
the earth and other solar system materials. Geochim. Cosmochim. Acta, 42:1075–1090.
Rye R. O. and Ohmoto H. (1974). Sulfur and carbon isotopes and ore genesis: A review. Econ.
Geol., 69:826–842.
Ryerson F. J. and Hess P. C. (1978). Implications of liquid distribution coefficients to mineral-
liquid partitioning. Geochim. Cosmochim. Acta, 42:921–932.
Ryerson F. J. and Hess P. C. (1980). The role of P2O5 in silicate melts. Geochim. Cosmochim.
Acta, 44:611–624.
Sack R. O. (1980). Some constraints on the thermodynamic mixing properties of Fe-Mg ortho-
pyroxenes and olivines. Contrib. Mineral. Petrol., 71:257–269.
Sahama Th. G. and Torgeson D. R. (1949). Some examples of applications of thermochemistry
to petrology. J. Geol., 57:255–262.
Sakai H. (1968). Isotopic properties of sulfur compounds in hydrothermal processes. Geochem.
J., 2:29–49.
Sakai H. and Tsutsumi M. (1978). D/H fractionation factors between serpentine and water at
1000° to 500°C and 2000 bar water pressure and the D/H ratios of natural serpentines.
Earth Planet. Sci. Letters, 40:231–242.
Salje E. (1985). Thermodynamics of sodium feldspar I: Order parameter treatment and strain
induced coupling effects. Phys. Chem. Minerals, 12:93–98.
Salje E. (1988). Structural phase transitions and specific heat anomalies. In Physical Properties
and Thermodynamic Behaviour of Minerals, E. K. H. Salje, ed. Reidel Publishing
Company.
Salje E., Kuscholke B., Wruck B., and Kroll H. (1985). Thermodynamics of sodium feldspar,
II: Experimental results and numerical calculations. Phys. Chem. Minerals, 12:99–107.
Samsonov V. (1968). Handbook of the Physicochemical Properties of the Elements. New York-
Washington: IFI/Plenum.
Sanderson R. T. (1960). Chemical Periodicity. New York: Reinhold.
Sanderson R. T. (1966). Bond energies. J. Inorg. Nucl. Chem., 28:1553–1565.
Sanderson R. T. (1967). Inorganic Chemistry. New York: Reinhold.
Sartori F. (1976). The crystal structure of 1M lepidolite. Tsch. Min. Petr. Mitt., 23:65–75.
Sarver J. V. and Hummel F. A. (1962). Solid solubility and eutectic temperature in the system
Zn2SiO4-Mg2SiO4. J. Amer. Ceram. Soc., 45:304–314.
Sasaki S., Fujino K., Takeuchi Y., and Sadanaga R. (1980). On the estimation of atomic charges
by the X-ray method for some oxides and silicates. Acta Cryst., A36:904–915.
Sasaki S. Y., Takeuchi Y., Fujino K., and Akimoto S. (1982). Electron density distribution of
three orthopyroxenes, Mg2Si2O6, Co2Si2O6 and Fe2Si2O6. Zeits. Krist., 158:279–297.
Sato Y., Akaogi M., and Akimoto S. (1978). Hydrostatic compression of the synthetic garnets
pyrope and almandine. J. Geophys. Res., 83:335–338.
Saxena S. K. (1968). Distribution of iron and magnesium between coexisting garnet and clino-
pyroxene in rocks of varying metamorphic grade. Amer. Mineral., 53:2018–2021.
Saxena S. K. (1969). Silicate solid solution and geothermometry. 3. Distribution of Fe and Mg
between coexisting garnet and biotite. Contrib. Mineral. Petrol., 22:259–267.
Saxena S. K. (1972). Retrival of thermodynamic data from a study of intercrystalline and intra-
crystalline ion-exchange equilibrium. Amer. Mineral., 57:1782–1800.
Saxena S. K. (1973). Thermodynamics of Rock-Forming Crystalline Solutions. Berlin-
Heidelberg-New York: Springer-Verlag.
Saxena S. K. (1977). A new electronegativity scale for geochemists. In Energetics of Geological
Processes, S. K. Saxena and S. Bhattacharji, eds. New York-Heidelberg-Berlin:
Springer-Verlag.
Saxena S. K. (1979). Garnet-clinopyroxene geothermometer. Contrib. Mineral. Petrol.,
70:229–235.
References / 853
Seck H. A. (1972). The influence of pressure on the alkali-feldspar solvus from peraluminous
and persilicic materials. Fortschr. Mineral., 49:31–49.
Seifert S. and O’Neill H.St. C. (1987). Experimental determination of activity-composition re-
lations in Ni2SiO4-Mg2SiO4 and Co2SiO4-Mg2SiO4 olivine solid solutions at 1200 K and
0.1 MPa and 1573 K and 0.5 GPa. Geochim. Cosmochim. Acta, 51:97–104.
Seward T. M. (1981). Metal complex formation in aqueous solutions at elevated temperatures
and pressures. Phys. Chem. Earth, 13:113–132.
Shannon R. D. (1976). Revised effective ionic radii and systematic studies of interatomic dis-
tances in halides and chalcogenides. Acta Cryst., A32:751–767.
Shannon R. D. and Prewitt C. T. (1969). Effective ionic radii in oxides and fluorides. Acta
Cryst., B25:925–946.
Shaw D. M. (1970). Trace element fractionation during anatexis. Geochim. Cosmochim. Acta,
34:237–243.
Shaw D. M. (1978). Trace element behaviour during anatexis in the presence of a fluid phase.
Geochim. Cosmochim. Acta, 42:933–943.
Shmulovich K. I., Shmonov V. M., and Zharikov V. A. (1982). The thermodynamics of super-
critical fluid systems. In Advances in Physical Geochemistry. vol. 2, S. K. Saxena (series
ed.). New York: Springer-Verlag.
Shock E. L. and Helgeson H. C. (1988). Calculations of the thermodynamic and transport
properties of aqueous species at hight pressures and temperatures: Correlation algo-
rithms for ionic species and equations of state predictions to 5Kb and 1000°C. Geochim.
Cosmochim. Acta, 52:2009–2036.
Shock E. L. and Helgeson H. C. (1990). Calculations of the thermodynamic and transport
properties of aqueous species at high pressures and temperatures: Standard partial
molal properties of organic species. Geochim. Cosmochim. Acta, 54:915–946.
Shock E. L. and Koretsky C. M. (1995). Metal-organic complexes in geochemical processes:
Estimation of standard partial molal thermodynamic properties of aqueous complexes
between metal cations and monovalent organic acid ligands at high pressures and tem-
peratures. Geochim. Cosmochim. Acta, 59:1497–1532.
Shock E. L., Helgeson H. C. and Sverjensky B. A. (1989). Calculation of thermodynamic and
transport properties of aqueous species at high pressures and temperatures: Standard
partial molal properties of inorganic neutral species. Geochim. Cosmochim. Acta,
53:2157–2183.
Simkin T. and Smith J. V. (1970). Minor element distribution in olivine. J. Geol., 78:304–325.
Simons B. (1986). Temperatur und Druckabhangigkeit der Fehlstellenkonzentration der Olivine
und Magnesiowustite Habilitationsschrift, Christian-Albrechts Universitat, Kiel.
Sipling P. J. and Yund R. A. (1976). Experimental determination of the coherent solvus for
sanidine-high albite. Amer. Mineral., 61:897–906.
Sjoberg E. L. (1976). A fundamental equation for calcite dissolution kinetics. Geochim. Cos-
mochim. Acta, 40:441–447.
Skinner B. J. (1956). Physical properties of end-members of the garnet group. Amer. Mineral.,
41:428–436.
Skinner B. J. (1966). Thermal expansion. In Handbook of Physical Constants, S. P. Clark Jr.,
ed. Memoir 97, The Geological Society of America.
Slavinskiy V. V. (1976). The clinopyroxene-garnet geothermometer. Dokl. Akad. Nauk USSR,
231:181–184.
Smith D. M. and Stocker R. L. (1975). Point defects and non-stoichiometry in forsterite. Phys.
Earth. Planet. Int., 10:183–192.
Smith J. V. (1983). Some chemical properties of feldspars. In Reviews in Mineralogy, vol. 2 (2d
ed.), P. H. Ribbe (series ed.), Mineralogical Society of America.
Smith J. V. and Yoder H. S. Jr. (1956). Experimental and theoretical studies of the mica poly-
morphs. Min. Mag., 31:209–235.
Smyth J. R. (1973). An orthopyroxene structure up to 850 °C. Amer. Mineral., 58:636–648.
References / 855
Smyth J. R. (1974). The high temperature crystal chemistry of clinohyperstene. Amer. Min-
eral., 59:1069–1082.
Smyth J. R. and Bish D. L. (1988). Crystal Structures and Cation Sites of the Rock-Forming
Minerals Allen and Unwin, Boston.
Smyth J. R. and Hazen R. M. (1973). The crystal structure of forsterite and hortonolite at
several temperatures up to 900°C. Amer. Mineral., 58:588–593.
Smyth J. R., Smith J. V., Artioli G., and Kvick A. (1987). Crystal structures of coesite at 15 K
and 298 K from neutron and X-ray single crystal data: Test of bonding models. J. Phys.
Chem., 91:988–992.
Sockel H. G. and Hallwig D. (1977). Ermittlung kleiner Diffusionkoeffizienten mittels SIMS in
oxydischen Verbindungen. Mikrokim. Acta, 7:95–107.
Soddy F. (1914). The Chemistry of the Radio-Elements. Pt. II Longmans & Green, London.
Speer J. A. (1984). Micas in igneous rocks. In Reviews in Mineralogy, vol. 13, P. H. Ribbe (series
ed.), Mineralogical Society of America.
Speidel D. H. and Osborn E. F. (1967). Element distribution among coexisting phases in the
system MgO-FeO-Fe2O3-SiO2 as a function of temperature and oxygen fugacity. Amer.
Mineral., 52:1139–1152.
Spencer K. J. and Lindsley D. H. (1981). A solution model for coexisting iron-titanium oxides.
Amer. Mineral., 66:1189–1201.
Stebbins J. F., Carmichael I. S. E. and Moret L. K. (1984). Heat capacities and entropies of
silicate liquids and glasses. Contrib. Mineral. Petrol., 86:132–148.
Steele I. M. and Smith J. V. (1982). Ion probe analysis of plagioclase in three howardites and
three eucrites. Geochim. Cosmochim. Acta, 42:959–971.
Steele I. M., Hutcheon I. D. and Smith J. V. (1980a). Ion microprobe analysis and petrogenetic
interpretations of Li, Mg, Ti, K, Sr, Ba in lunar plagioclase. Geochim. Cosmochim. Acta,
Suppl. 14, 571–590.
Steele I. M., Hutcheon I. D. and Smith J. V. (1980b). Ion microprobe analysis of plagioclase
feldspar (Ca1⫺xNaxAl2⫺xSi2⫹xO8). for major, minor and trace elements. VIII Int. Congr.
X-ray Optics Microanalysis, Pendell Pub. Co, Midland, Michigan.
Steele I. M., Smith J. V., Raedeke L. D. and McCallum I. S. (1981). Ion probe analysis of Still-
water plagioclase and comparison with lunar analyses. Lunar Planet. Sci., 12:1034–1036.
Stevenson F. J. (1976). Stability constants of Cu2⫹, Pb2⫹ and Cd 2⫹ complexes with humic acids.
Soil. Sci. Soc. Amer. Jour., 40:665–672.
Stevenson F. J. (1982). Humus Chemistry: Genesis, Composition, Reactions. New York: Wiley
Interscience.
Stevenson F. J. (1983). Trace metal-organic matter interaction in geologic environment. In The
Significance of Trace Elements in Solving Petrogenetic Problems and Controversies, S. Au-
gustitis, ed. Theophrastus Publications, Athens.
Stewart D. B. and Ribbe P. H. (1969). Structural explanation for variations in cell parameters
of alkali feldspars with Al/Si ordering. Amer. Jour. Sci., 267A:444–462.
Stewart R. F., Whitehead M. A. and Donnay D. (1980). The ionicity of the Si-O bond in
quartz. Amer. Mineral., 65:324–326.
Stockmayer W. H. (1941). Second virial coefficients of polar gases. J. Chem. Phys., 9:398–402.
Stolper E. (1982). The speciation of water in silicate melts. Geochim. Cosmochim. Acta,
46:2609–2620.
Stolper E., Fine G., Johnson T., and Newmann S. (1987). Solubility of carbon dioxide in albitic
melt. Amer. Mineral., 72:1071–1085.
Stolper E., Walker D., Hager B. H. and Hays J. F. (1981). Melt segregation from partially mol-
ten source regions: The importance of melt density and source region size. J. Geophys.
Res., 86:6261–6271.
Stormer J. C. Jr. (1975). A practical two feldspar thermometer. Amer. Mineral., 60:667–674.
Strip K. F. and Kirkwood J. G. (1951). Asymptotic expansion of the rotational partition func-
tion of the asymmetric top. J. Chem. Phys., 19:1131–1133.
856 / References
Stull D. R. and Prophet H. (1971). JANAF Thermochemical Tables Data series, Washington
D.C., 37:1–1141.
Stumm W. and Morgan J. J. (1981). Acquatic Chemistry. New York: John Wiley.
Sueno S., Papike J. J. and Prewitt C. T. (1973). The high temperature crystal chemistry of trem-
olite. Amer. Mineral., 58:649–664.
Sueno S., Papike J. J., Prewitt C. T. and Brown G. E. (1972). Crystal structure of high cumming-
tonite. J. Geophys. Res., 77:5767–5777.
Suess H. (1965). Secular variations of the cosmic-ray produced carbon-14 in the atmosphere
and their interpretation. J. Geophys. Res., 70:5937–5952.
Sumino Y. (1979). The elastic constants of Mn2SiO4, Fe2SiO4 and Co2SiO4, and the elastic
properties of olivine group minerals at high temperature. J. Phys. Earth, 27:209–238.
Sumino Y., Nishizawa O., Goto T., Ohno I., and Ozima I. (1977). Temperature variation of
elastic constants of single crystal forsterite between 190 and 400°C. J. Phys. Earth,
28:273–280.
Sutin N., Rowley J. K. and Dodson W. (1961). Chloride complexes of iron (III) ions and the
kinetics of chloride-catalyzed exchange reactions between iron (II) and iron (III) in light
and heavy water. J. Phys. Chem., 65:1248–1252.
Sutton S. R., Jones K. W., Gordon B., Rivers M. L., Bajt S., and Smith J. V. (1993). Reduced
chromium in olivine grains from lunar basalt 15555: X-ray Absorption Near Edge Struc-
ture (XANES). Geochim. Cosmochim. Acta, 57:461–468.
Suzuki I. and Anderson O. L. (1983). Elasticity and thermal expansion of a natural garnet up
to 1000 K. J. Phys. Earth, 31:125–138.
Suzuki I., Seya K., Takei H., and Sumino Y. (1981). Thermal expansion of fayalite, Fe2SiO4.
Phys. Chem. Minerals, 7:60–63.
Suzuoki T. and Epstein S. (1976). Hydrogen isotope fractionation between OH-bearing miner-
als and water. Geochim. Cosmochim. Acta, 40:1229–1240.
Szilagyi M. (1971). Reduction of Fe3⫹ ion by humic acid preparations. Soil Sci., 3:233–235.
Tanaka T. and Masuda A. (1982). La-Ce geochronometer: A new dating method. Nature,
300:515–518.
Tanger J. C. and Helgeson H. C. (1988). Calculation of the thermodynamic and transport prop-
erties of aqueous species at high pressures and temperatures: Revised equation of state
for the standard partial molal properties of ions and electrolytes. Amer. Jour. Sci.,
288:19–98.
Tardy Y. (1979). Relationship among Gibbs energies of formation of compounds. Amer. Jour.
Sci., 279:217–224.
Tardy Y. and Garrels R. M. (1974). A method of estimating the Gibbs energies of formation
of layer silicates. Geochim. Cosmochim. Acta, 38:1101–1116.
Tardy Y. and Garrels R. M. (1976). Prediction of Gibbs energies of formation: I-Relationships
among Gibbs energies of formation of hydroxides, oxides and aqueous ions. Geochim.
Cosmochim. Acta, 40:1015–1056.
Tardy Y. and Garrels R. M. (1977). Prediction of Gibbs energies of formation of compounds
from the elements, II: Monovalent and divalent metal silicates. Geochim. Cosmochim.
Acta, 41:87–92.
Tardy Y. and Gartner L. (1977). Relationships among Gibbs energies of formation of sulfates,
nitrates, carbonates, oxides and aqueous ions. Contrib. Mineral. Petrol., 63:89–102.
Tardy Y. and Viellard Ph. (1977). Relationships among Gibbs energies and enthalpies of forma-
tion of phosphates, oxides and aqueous ions. Contrib. Mineral. Petrol., 63:75–88.
Taylor H. P. Jr. (1968). The oxygen isotope geochemistry of igneous rocks. Contrib. Mineral.
Petrol., 19:1–71.
Taylor H. P. Jr. and Epstein S. (1962). Relation between 18O/16O ratio in coexisting minerals of
igneous and metamorphic rocks. Part I. Principles and experimental results. Geol. Soc.
Amer. Bull., 73:461–480.
References / 857
Temkin M. (1945). Mixtures of fused salts as ionic solutions. Acta Phys. Chim. URSS, 20:411–
420.
Temkin M. (1963). The kinetics of stationary reactions. Akad. Nauk. SSSR Doklady,
152:782–785.
Tequı̀ C., Robie R. A., Hemingway B. S., Neuville D. R. and Richet P. (1991). Melting and
thermodynamic properties of pyrope (Mg3Al2Si3O12). Geochim. Cosmochim. Acta,
55:1005–1010.
Tetley N. W. (1978). Geochronology by the 40Ar/39Ar technique using HIFAR reactor Ph.D. Dis-
sertation, Australia National University, Canberra.
Thompson A. B. (1976). Mineral reactions in pelitic rocks, II: Calculations of some P-T-X (Fe-
Mg). phase relations. Amer. Jour. Sci., 276:425–444.
Thompson J. B. Jr. (1969). Chemical reactions in crystals. Amer. Mineral., 54:341–375.
Thompson J. B. Jr. (1970). Chemical reactions in crystals: Corrections and clarification. Amer.
Mineral., 55:528–532.
Thompson J. B. Jr. and Hovis G. L. (1978). Triclinic feldspars: Angular relations and the repre-
sentation of feldspar series. Amer. Mineral., 63:981–990.
Thompson J. B. Jr. and Waldbaum D. R. (1969). Mixing properties of sanidine crystalline solu-
tions. III Calculations base on two phase data. Amer. Mineral., 54:811–838.
Thompson J. B. Jr., Waldbaum D. R. and Hovis G. L. (1974). Thermodynamic properties re-
lated to ordering in end member alkali feldspars. In The Feldspars, W. S. MacKenzie
and J. Zussman, eds. Manchester: Manchester University Press.
Thompson P. and Grimes N. W. (1977). Madelung calculation for the spinel structure. Phil.
Mag., 36:501–505.
Thorstenson D. C. (1970). Equilibrium distribution of small organic molecules in natural wa-
ters. Geochim. Cosmochim. Acta, 34:745–770.
Tilton G. R. (1960). Volume diffusion as a mechanism for discordant lead ages. J. Geophys.
Res., 65:2933–2945.
Tokonami M., Horiuchi H., Nakano A., Akimoto S., and Morimoto N. (1979). The crystal
structure of pyroxene type MnSiO3. Min. Jour., 9:424–426.
Toop G. W. and Samis C. S. (1962a). Some new ionic concepts of silicate slags. Can. Met.
Quart., 1:129–152.
Toop G. W. and Samis C. S. (1962b). Activities of ions in silicate melts. Trans. Met. Soc.
AIME, 224:878–887.
Tosi M. P. (1964). Cohesion of ionic solids in the Born model. Solid State Phys., 16:1–120.
Tossell J. A. (1980). Theoretical study of structures, stabilities and phase transition in some
metal dihalide and dioxide polymorphs. J. Geophys. Res., 85:6456–6460.
Tossell J. A. (1981). Structures and cohesive properties of hydroxides and fluorides calculated
using the modified electron gas ionic model. Phys. Chem. Minerals, 7:15–19.
Tossell J. A. (1985). Ab initio SCF MO and modified electron gas studies of electron deficient
anions and ion pairs in mineral structures. Physica, 131B:283–289.
Tossell J. A. (1993). A theoretical study of the molecular basis of the Al avoidance rule and of
the spectral characteristics of Al-O-Al linkages. Amer. Mineral., 78:911–920.
Tossell J. A. and Vaughan D. J. (1992). Theoretical Geochemistry: Application of Quantum Me-
chanics in the Earth and Mineral Sciences. New York-Oxford: Oxford University Press.
Tossell J. A., Vaughan D. J. and Johnson K. H. (1973). X-ray photoelectron, x-ray emission
and UV spectra of SiO2 calculated by the SCF-X␣. scattered wave method. Chem. Phys.
Lett., 20:329–334.
Tossell J. A., Vaughan D. J. and Johnson K. H. (1974). The electronic structure of rutile, wustite
and hematite from molecular orbital calculations. Amer. Mineral., 59:319–334.
Touret J. and Bottinga Y. (1979). Equation d’etat pour le CO2; application aux inclusion car-
boniques. Bull. Mineral., 102:577–583.
Treuil M. (1973). Critères petrologiques geochimiques et structuraux de la genèse et de la
858 / References
crocline and plagioclase and its effect on geothermometric calculations. Amer. Min-
eral., 62:687–691.
Whittaker E. J. W. and Muntus R. (1970). Ionic radii for use in geochemistry. Geochim. Cos-
mochim. Acta, 34:945–956.
Wilhem E., Battino R., and Wilcock R. J. (1977). Low pressure solubility of gases in liquid
water. Chem. Rev., 77:219–262.
Will T. M. and Powell R. (1992). Activity-composition relationships in multicomponent amphi-
boles: An application of Darken’s quadratic formalism. Amer. Mineral., 77:954–966.
Willaime C. and Gaudais M. (1977). Electron microscope study of plastic defects in experimen-
tally deformed alkali feldspar. Bull. Soc. Fr. Mineral. Cristallogr., 100:263–271.
Williams R. J. (1971). Reaction constants in the system Fe-MgO-SiO2-O2 at 1 atm between 900
and 1300°C: Experimental results. Amer. Jour. Sci., 270:334–360.
Winchell A. N. (1933). Elements of Optical Mineralogy. New York: John Wiley.
Winter J. K., Okamura F. P. and Ghose S. (1979). A high temperature structural study of high
albite, monalbite, and the analbite-monalbite phase transition. Amer. Mineral.,
64:409–423.
Witte H. H. and Wölfel E. (1955). Electron distribution in rock salt. Zeit. Phys. Chem.,
3:296–329.
Wohl K. (1946). Thermodynamic evaluation of binary and ternary liquid system. Trans. Amer.
Inst. Chem. Eng., 42:215–249.
Wohl K. (1953). Thermodynamic evaluation of binary and ternary liquid systems. Chem. Eng.
Progr., 49:218–219.
Wolery T. J. (1983). EQ3NR. A Computer Program for Geochemical Aqueous Speciation-
Solubility Calculations: User’s Guide and Documentation Lawrence Livermore Labora-
tory, Livermore, Cal., UCRL-53414.
Wolery T. J. (1986). Some Forms of Transition State Theory, Including Non-Equilibrium Steady
State Forms Lawrence Livermore Laboratory, Livermore, Cal., UCRL-94221.
Wones D. R. (1967). A low pressure investigation of the stability of phlogopite. Geochim. Cos-
mochim. Acta, 31:2248–2253.
Wones D. R. (1972). Stability of biotite: A reply. Amer. Mineral., 57:316–317.
Wones D. R. and Dodge F. C. W. (1977). The stability of phlogopite in the presence of quartz
and diopside. In Thermodynamics in Geology, D. G. Fraser, ed. Dordrecht-Holland,
Reidel.
Wones D. R. and Eugster H. P. (1965). Stability of biotite: Experiment theory and application.
Amer. Mineral., 50:1228–1272.
Wood B. J. (1974). Crystal field spectrum of Ni2⫹ in olivine. Amer. Mineral., 59:244–248.
Wood B. J. (1976a). The reaction phlogopite ⫹ quartz ⫽ enstatite ⫹ sanidine ⫹ H2O. In Prog-
ress in Experimental Petrology, G. M. Biggar, ed. Natural Environment. Res. Coun-
cil:Publ. series D.
Wood B. J. (1976b). The partitioning of iron and magnesium between garnet and clinopyro-
xene. Carnegie Inst. Wash. Yb., 75:571–574.
Wood B. J. (1976c). Samarium distribution between garnet and liquid at high pressure. Carne-
gie Inst. Wash. Yb., 75:659–662.
Wood B. J. (1980). Crystal field electronic effects on the thermodynamic properties of Fe2⫹
minerals. In Advances in Physical Geochemistry, vol. 1, S. K. Saxena (series ed.), New
York-Heidelberg-Berlin: Springer-Verlag.
Wood B. J. (1987). Thermodynamics of multicomponent systems containing several solid solu-
tions. In Reviews in Mineralogy. vol. 17, P. H. Ribbe (series ed.), Mineralogical Society
of America.
Wood B. J. (1988). Activity measurements and excess entropy-volume relationships for pyrope-
grossular garnets. J. Geol., 96:721–729.
Wood B. J. and Banno S. (1973). Garnet-orthopyroxene and orthopyroxene-clinopyroxene rela-
tionships in simple and complex systems. Contrib. Mineral. Petrol., 42:109–124.
References / 861
Wood B. J. and Fraser D. G. (1976). Elementary Thermodynamics for Geologists Oxford Univer-
sity Press.
Wood B. J. and Kleppa O. J. (1981). Thermochemistry of forsterite-fayalite olivine solutions.
Geochim. Cosmochim. Acta, 45:529–534.
Wood B. J. and Kleppa O. J. (1984). Chromium-aluminum mixing in garnet: A thermochemical
study. Geochim. Cosmochim. Acta, 48:1373–1375.
Wood B. J. and Strens R. G. J. (1972). Calculation of crystal field splittings in distorted coordi-
nation polyhedra: Spectra and thermodynamic properties of minerals. Min. Mag.,
38:909–917.
Wood B. J., Hackler R. T. and Dobson D. P. (1994). Experimental determination of Mn-Mg
mixing properties in garnet, olivine and oxide. Contrib. Mineral. Petrol., 115:438–448.
Wood J. A. (1979). The Solar System Prentice-Hall, Englewood Cliffs, N.J.
Woodland A. B. and O’Neill H. St. C. (1993). Synthesis and stability of garnet and phase rela-
tions with solutions. Amer. Mineral., 78:1002–1015.
Wriedt H. A. and Darken L. S. (1965). Lattice defects and solubility of nitrogen in deformed
ferritic steel. Trans. Met. Soc. AIME, 233:111–130.
Yeh H. W. (1980). D/H ratios and late-stage dehydration of shales during burial. Geochim. Cos-
mochim. Acta, 341–352.
Yoder H. S. and Eugster H. P. (1954). Phlogopite synthesis and stability range. Geochim. Cos-
mochim. Acta, 6:157–185.
York D. (1969). Least-squares fitting of a straight line with correlated errors. Earth Planet. Sci.
Letters, 5:320–324.
Yoshimori A. and Kitano Y. (1956). Theory of the lattice vibration of graphite. J. Phys. Soc.
Japan, 11:352–361.
Young J. A. and Koppel J. U. (1965). Phonon spectrum of graphite. J. Chem. Phys., 42:357–364.
Yund R. A. (1974). Coherent exsolution in the alkali feldspars. In Geochemical Transport and
Kinetics, Hofmann, Giletti, Yoder, and Yund, eds. New York: Carnegie Inst. Washing-
ton and Academic Press.
Yund R. A. (1983). Diffusion in feldspars. In Reviews in Mineralogy, vol. 2 (2d ed.), P. H. Ribbe
(series ed.), Mineralogical Society of America.
Yund R. A. and Anderson T. F. (1974). Oxygen isotope exchange between potassium feldspar
and KCl solution. In Geochemical Transport and Kinetics, Hofmann, Giletti, Yoder, and
Yund, eds. New York: Carnegie Inst. Washington and Academic Press.
Yund R. A. and Tullis J. (1983). Subsolidus phase relations in the alkali feldspars with emphasis
on coherent phases. In Reviews in Mineralogy, vol. 2 (2d ed.), P. H. Ribbe (series ed.),
Mineralogical Society of America.
Zhang Z. and Saxena S. K. (1991). Thermodynamic properties of andradite and application to
skarn with coexisting andradite and hedenbergite. Contrib. Mineral. Petrol.,
107:255–263.
Zheng Y. F. (1991). Calculation of oxygen isotope fractionation in metal oxides. Geochim. Cos-
mochim. Acta, 55:2299–2307.
Zheng Y. F. (1993a). Calculation of oxygen isotope fractionation in anhydrous silicate minerals.
Geochim. Cosmochim. Acta, 57:1079–1091.
Zheng Y. F. (1993b). Calculation of oxygen isotope fractionation in hydroxyl-bearing silicates.
Earth Planet. Sci. Letters, 120:247–263.
INDEX
/ 863
864 / Index