Chapter-04-Applications-Thermodynamics-Earth 4744 0 PDF
Chapter-04-Applications-Thermodynamics-Earth 4744 0 PDF
Chapter-04-Applications-Thermodynamics-Earth 4744 0 PDF
White Geochemistry
Chapter 4: Applications of Thermodynamics
I
n the previous 2 chapters, we developed the fundamental thermodynamic relationships and
saw how they are applied to geochemical problems. The tools now in our thermodynamic
toolbox are sufficient to deal with most of the phenomena we will encounter in the second half
of this book. They are not sufficient, however, to deal with all geochemical problems. In this
chapter, will add a final few thermodynamic tools. These allow us to deal with non-ideal behavior
and exsolution phenomena in solids and silicate liquids. With that, we can use thermodynamics to
determine the pressure and temperature at which rock assemblages formed, certainly one of the
most useful applications of thermodynamics to geology. Along the way, we will see how thermo-
dynamics is related to one of the most useful tools in petrology: phase diagrams. Finally, we return
to the question of non-ideal behavior in electrolyte solutions and examine in more depth the prob-
lems of ion association and solvation and how this affects ion activities. Deviations from ideal be-
havior tend to be greater in solutions of high ionic strength, which includes such geologically im-
portant solutions as hydrothermal and ore-forming fluids, saline lake waters, metamorphic fluids,
and formation and oil field brines. We briefly examine methods of computing activity coefficients
at ionic strengths relevant to such fluids.
4.2 Activities in Non-Ideal Solid Solutions
4.2.1 Mathematical Models of Real Solutions: Margules Equations
Ideal solution models often fail to describe the behavior of real solutions; a good example is wa-
ter and alcohol, as we saw in Chapter 3. Ideal solutions fail spectacularly when exsolution occurs,
such as between oil and vinegar, or between orthoclase and albite, a phenomenon we will discuss
in more detail shortly. In non-ideal solutions, even when exsolution does not occur, more complex
models are necessary.
Power, or Maclaurin, series are often a convenient means of expressing complex mathematical
functions, particularly if the true form of the function is not known. This approach is the basis of
Margules† equations, a common method of calculating excess free energy. For example, we could
express the excess volume as a power series:
Vex = A + BX 2 + CX 22 + DX 23 + K 4.1
where X2 is the mole fraction of component 2.
Following the work of Thompson (1967), Margules equations are used extensively in geochemis-
try and mineralogy as models for the behavior of non-ideal solid solutions. We will consider two
variants of them: the symmetric and asymmetric solution models.
4.2.1.1 The Symmetric Solution Model
In some solutions, a sufficient approximation of thermodynamic functions can often be obtained
by using only a second order power series, i.e., in equ. 4.1, D = E = ... = 0. Now in a binary solu-
tion, the excess of any thermodynamic function should be entirely a function of mole fraction X 2 (or
X 1, however we wish to express it). Put another way, where X 2 = 0, we expect Vex = 0. From this
we can see that the first term in Equ. 4.1, A, must also be 0. Thus equation 4.1 simplifies to:
‡ The term regular solution is often used to refer to symmetric solutions. In that case, what we termed a regular
solution is called a strictly regular solution.
DG (kJ/mol)
or WG = RT ln h
then 4.15 is Henry’s Law. Thus the 0
interaction parameter can be related DG real
to the parameters of Henry’s Law,
and activity coefficient. In the Mar- -2
gules representation, a solution that is
ideal throughout is simply the special
case where A= B = C = D = ... = 0. -4
4.2.1.2 The Asymmetric Solution
DGi ideal
Model -6
Many real solutions, for example 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1
mineral solutions with asymmetric XAb
solvi, are not symmetric. This corre-
sponds to the case where D in equa- Figure 4.1. Alkali feldspar solid solution computed at 600°
tion 4.01 is nonzero; i.e., we must C and 200 MPa (2 kb) using the data of Thompson and
carry the expansion to the third order. Waldbaum (1969). DGreal = DGideal + DGexcess .
It can be shown that in this case the 0
excess free energy in binary solutions
is given by:
G ex = (WG1 X2 + WG2 X1 ) X1 X2 4.16
500∞
C
(You can satisfy yourself that this may
550∞C
-1
be written as a power-series to the
600∞C
third order of either X 1 or X 2.) The
two coefficients are related to the
DG (kJ/mol)
n
G excess can be positive or negative. G excess depends on
the lambda terms, or interaction parameters, which in a
turn depend on composition (for non-Henry's Law T b
solutions). If for some combination of T and X (and two coexisting
P), the second derivative of G excess becomes negative solutions
and its absolute value is greater than the RT/X1X2
term, inflection points appear in the G-X curve. Exso-
lution is thermodynamically favored (Figure 4.5b) in
this situation. 0 X2 1
Suppose now that phases a and b are not pure sub- Figure 4.5. (a) Schematic isothermal, iso-
stances, but are each limited solutions of both compo- baric G-X plot for a real solution showing
nents 1 and 2, as is illustrated in Figure 4.5b. Two in- DG of mechanical mixing, ideal mixing and
flection points appear, one at X a2 and one at X 2b . We excess mixing. (b) Sum of ideal and excess
can draw a straight line that is tangent to the free en- mixing free energies shown in (a). Tangents
ergy curve at both inflection points. This line is the to the minima give the chemical potentials
free energy of a mechanical mixture of the two lim- in immiscible phases a and b. (c). T-X plot
ited solutions a and b. Phase a is mostly component for same system as in (b). Solid line is the
solvus, dashed line is the spinodal. Ex-
1, but contains X a2 of component 2. Similarly, phase
solution may not occur between the
b is mostly component 2 but contains 1 – X 2b of com- spinodal and solvus because the free energy
ponent 1. The mechanical mixture of a and b has less can locally increase during exsolution.
free energy than a single solution phase everywhere From Nordstrom and Munoz (1986).
between X a2 and X 2b . It is therefore thermodynami-
se
points such as B but at inflection points
DGunmix 1
pha
(where ∂2G/∂X2 = 0) such as S. The locus A"
of such points is plotted in Figure 4.5c as B" A' + B'
the gray line and is known as the spinodal.
A' G
B
4.4 Thermodynamics and Phase B'
Diagrams
B
A phase diagram is a representation of A+
the regions of stability of one or more
phases as a function of two or more ther-
modynamic variables such as temperature,
pressure, or composition. In other words, X2
if we plot 2 thermodynamic variables such Figure 4.6. A small portion of a G-X plot illustrating
as temperature and pressure or tempera- the origin of the spinodal. The process of exsolution of
ture and composition, we can define an two phases from a single solid solution must overcome
area on this plot where a phase of interest an energy barrier. As exsolution from a solution of
is thermodynamically stable. Figure 4.7 is composition C begins, the two exsolving phases have
an example of a T-P phase diagram for a 1 compositions that move away from C, e.g., A’ and B’.
component system: SiO2. The diagram But the free energy of a mechanical mixture of A’ and
shows the SiO2 phase stable for a given B´ has greater free energy, by DGunmix than the original
combination of pressure and temperature. single solution phase. Exsolution will therefore be in-
Figure 4.8 is an example of a simple T-X hibited in this region. This problem does not occur if
the original solution has composition C’.
diagram for the two-component system
= 1- 4.25
T DSi , m
Since enthalpies of fusion, rather than
1400 An + L
entropies, are the quantities measured,
equation 4.25 may be more conveniently
expressed as:
Di + L
Ti , m Ti , m R ln ail, m 1300
= 1- 4.26
T DHi , m An + Di
Example 4.2 shows how the approximate
phase diagram for the diopside-anorthite 1200
system (Figure 4.9) may be constructed 0 0.2 0.4 0.6 0.8 1
using this equation. X Di Di
It must be emphasized that in deriving
An
equation 3.66, and hence the equations 4.25 Figure 4.9. Computed phase diagram for the system
and 4.26, we made the assumption that the Anorthite-Diopside (CaAl2 Si2 O8 –CaMgSi2 O6 ). The eu-
solid was a pure phase. This assumption is tectic occurs at XDi = 0.7 and 1334°C. The dashed lines
a reasonably good one for ice, and for beyond the eutectic give the apparent melting points of
anorthite-diopside binary system, but it is the components in the mixture.
* The heat of fusion is often designated by DH . I have chosen to use the subscript m to avoid confusion with heat
f
of formation, for which we have already been using the subscript f.
We then calculate T for every value of X An and X Di . This producestwo curves on a T-X plot, as
shown in Figure 4.09. The curvesintersect at the eutectic , or lowest point at which melt may exist
in the system.
Comparing our result with the actual phase relationships determined experimentally
(Figure 4.9), we see that while the computed phase diagram is similar to the actual one, our
computedeutectic occursat X Di = 0.70 and 1335° C and the actual eutectic occursat X Di ª 0.56 and
1274° C. The difference reflects the failure of the several assumptions we made. First, and most
importantly, silicate liquids are not ideal solutions. Second, the entropies and enthalpies of
fusion tend to decrease somewhat with decreasing temperature, violating the assumption we
made in deriving equation 4.25. Third, the diopside crystallizing from anorthite-diopside
mixtures is not pure, but contains some Al and an excess of Mg.
not generally valid. Should the solid or solids involved exhibit significant solid solution, this
assumption breaks down and these equations are invalid. In that case, melting phase diagrams can
still be constructed from thermodynamic equations, but we need to model solid solution as well as
the liquid one. Section 4.4.2.1 below illustrates an example (Anorthite-Albite) we the two solutions
can be modeled as ideal.
4.4.2 Thermodynamics of Phase Diagrams for Binary Systems
In a one component system, a phase boundary, such as the melting point, is univariant since at
that point two phases coexist and ƒ = c- p + 2 = 1 - 2 + 2 = 1. Thus specifying either temperature or
pressure fixes the other. A three-phase point, e.g., the triple point of water, is invariant. Hence
simply from knowing that three phases of water coexist (i.e., knowing we are at the triple point),
we know the temperature and pressure.
In binary systems, the following phase assemblages are possible according to the Gibbs Phase
Rule (ignoring for the moment gas phases):
Phases Free compositional variables
Univariant 2 solids + liquid, 2 liquids + solid, 3 solids or liquids 0
Divariant 1 solid + 1 liquid, 2 solids, 2 liquids 0
Trivariant 1 solid or 1 liquid 1
– –
When a G -X diagram is drawn, it is drawn for a specific temperature and pressure, i.e., G -X are
isobaric and isothermal. Thus we have already fixed two variables, and the compositions of all
phases in univariant and divariant systems are fixed by virtue of our having
– fixed T and P. Only in
trivariant systems are we free to choose the composition of a phase on a G -X diagram.
DG mAb
X sAb = X Ab
l
e RT
4.33
DG mAn RT
X s
An =X e l
An 4.34
Thus the fraction of each component in the melt can be predicted from the composition of the solid
l l s s
and thermodynamic properties of the end members. Since X An = 1 – X Ab and X An = 1 – X Ab , we can
combine equations 4.33 and 4.34 to obtain:
(1 - X )e DG mAn RT
e DG m
Ab
l
Ab = 1 - X Ab
l RT
4.35
DG mAn
1- e RT
and rearranging yields: X l
Ab = 4.36
e DG m - e DG m
Ab An
RT RT
The point is that the mole fraction of any component of any phase in this system can be predicted from the
thermodynamic properties of the end-members. We must bear in mind that we have treated this as an
ideal system; i.e., we have ignored any G excess term. Nevertheless, the ideal treatment is relatively
successful for the plagioclase system.
4.5 Geothermometry and Geobarometry
An important task in geochemistry is estimating the temperature and pressure at which mineral
assemblages equilibrate. The importance extends beyond petrology to tectonics and all of geology.
Here we take a brief look at the thermodynamics underlying geothermometry and geobarometry.
Geothermometry and geobarometry involve two nearly contradictory assumptions. The first is
that the mineral assemblage of interest is an equilibrium one, the second is that the system did not
re-equilibrate during the passage through lower P and T conditions that brought the rock to the sur-
face where it could be collected. As we will see in the next chapter, reaction rates depend exponen-
tially on temperature, hence these assumptions are not quite as contradictory as they might seem. In
this section, we will focus only on “chemical” thermobarometers. In Chapter 9, we will see that
temperatures can also be deduced from the distribution of isotopes between phases.
4.5.1 Theoretical considerations
In general, geobarometers and geothermometers make use of the pressure and temperature de-
pendence of the equilibrium constant, K. In Section 3.9 we found that DG° = -RT ln K. Assuming
that DCp and DV of the reaction are independent of temperature and pressure, we can write:
DG o = DH To ,Pref - TDS To ,Pref + DVTo,Pref ( P - Pref ) = - RT ln K 4.37
where the standard state of all components is taken as the pure phase at the temperature and pres-
sure of interest, and the enthalpy, entropy and volume changes are for the temperature of interest
and a reference pressure (generally 0.1 MPa).
Solving 4.37 for ln K and differentiating the resulting equation with respect to temperature and
pressure leads to the following relations:
Ê ∂ ln K ˆ DH To ,Pref + DVTo,Pref ( P - Pref )
Á ˜ = 4.38
Ë ∂T ¯ P RT 2
Ê ∂ ln K ˆ DVTo,Pref
and Á ˜ = 4.39
Ë ∂P ¯ T RT
These equations provide us with the criteria for reactions that will make good geothermometers
and geobarometers. For a good geothermometer, we want the equilibrium constant to depend
heavily on T, but be approximately independent of P. Looking at equation 4.38, we see this means
the DH term should be as large as possible and the DV term as small as possible. A fair amount of
P, GPa
investigations for the past 25 years. The
general principal is illustrated in Figure 1.5
4.15, which shows the concentration of
Al in orthopyroxene (opx) coexisting 1.0 1 2 4 6 8 10 12 14 16 Crystals
with olivine (forsterite) and an Fo + Sp +
aluminous phase, anorthite, spinel, or 0.5 Liquid
garnet. The Al content of opx depends Fo + An
almost exclusively on pressure in the 0
presence of anorthite, is essentially in- 400 600 800 1000 1200 1400 1600
dependent of pressure in the presence T, ∞C
of spinel, and depends on both Figure 4.15. Isopleths of Al in orthopyroxene (thin red
temperature and pressure in the lines; weight percent) coexisting with forsterite plus
presence of garnet. Orthopyroxene- and aluminous phase in the CMAS (Ca-Mg-Al-Si) sys-
garnet equilibria has proved to be a tem. After Gasparik (1984).
particularly useful geobarometer.
Garnet is an extremely dense phase. So we might guess that the DV of reactions that form it will
be comparatively large, and therefore that it is potentially a good geobarometer. The concentration
of Al in opx in equilibrium with garnet may be used as a geobarometer if temperature can be
independently determined. Although there has been a good deal of subsequent work and re-
finement of this geobarometer, the underlying thermodynamic principles are perhaps best illus-
trated by considering the original work of Wood and Banno (1973).
Wood and Banno (1973) considered the following reaction:
Mg2 Si2 O6 + MgAl2 SiO6 ® Mg3 Al2 Si3 O12 4.40
Opx Solid Solution ® Pyrope Garnet
In developing a geobarometer based on this reaction, they had to overcome a number of problems.
First, the substitution of Al in orthopyroxene is a coupled substitution. For each atom of Al substi-
tuting in the M1 octahedral site, there must be another Al atom substituting for SiO2 in the tetrahe-
dral site. Second, there was a total lack of thermodynamic data on the MgAl2SiO6 phase compo-
nent. Data was lacking for a good reason: the phase does not exist and cannot be synthesized as a
pure phase. Another problem was the apparent non-ideal behavior of the system, which was indi-
cated by orthopyroxenes in Fe- and Ca-bearing systems containing less alumina than in pure MgO
systems at the same pressure.
The equilibrium constant for reaction 4.40 is:
aMg 3 Al 2 Si 3O12
K= 4.41
aMg 2 Si 2O 6 aMgAl 2 SiO 6
where the activities in the denominator represent the activities of the enstatite and the hypothetical
aluminous enstatite phase components in the enstatite solid solution. In the pure MgO system (i.e.,
no CaO, FeO, MnO, etc.), the numerator, the activity of pyrope, is 1, of course, and we may write:
DG o = RT ln( aMg 2 Si 2O 6 aMgAl 2 SiO 6 ) = DH o - TDS o + ( P - Pref ) DV o 4.42
(compare equation 4.37). For an ideal case, this may be rewritten as:
RT ln( X Mg 2 Si 2O 6 X MgAl 2 SiO 6 ) = DH o - TDS o + ( P - Pref ) DV o 4.43
Di Hd
1000 900 70 0
800
1300 120 11
0 00
En Fs
Figure 4.17. Comparison of calculated (solid lines) and experimentally observed (red dashed lines)
phase relationships between clino- and orthopyroxene shown in the ‘pyroxene quadrilateral’, a part
of the CaSiO3–MgSiO3–FeSiO3 system. Di: diopside, En: enstatite, Hd: hedenbergite, Fs: ferrosilite.
Lines show the limit of solid solution at the corresponding temperatures (° C). Experiments show
solid solution to be complete in the iron end-members at 700 ° C.
Despite its complexities, the system has been modeled with some success using a symmetric solu-
tion model developed by Wood (1987). There are two octahedral sites in both ortho- and clinopy-
roxenes, generally called M1 and M2. Ca2+ occurs only in the M2 site, while Fe and Mg can oc-
cupy either site. Ignoring pigeonite and components other than Ca, Mg and Fe, we can treat mix-
ing in the M2 and M1 sites separately. Mixing in the M2 site can be treated as a ternary Mg, Fe,
and Ca solution. In a symmetric ternary solution consisting of components A, B, and C, the activi-
ties of the components may be calculated from:
(
RT ln g A = X B2WGAB + XC2WGAC + X B XC WGAB + WGAC - WGBC ) 4.45
where W GAB is the A-B binary interaction parameter, etc. Mixing of Fe and Mg between the M1 and
M2 sites was treated as a simple exchange reaction:
FeM2 + MgM1 ® FeM1 + MgM2
with DH of 29.27 kJ/mol and DS of 12.61 j/mol. Using this approach, Wood calculated the tempera-
ture dependence of the solvus in shown in Figure 4.17. The model fits experimental observation
reasonably well for the Mg-rich pyroxenes, but significant deviations occur for the Fe-rich pyrox-
enes.
4.5.2.3 Exchange Reactions
The third type of thermobarometer we will consider is the exchange reaction. These thermo-
barometers depend on the exchange of two species between phases. We will consider two exam-
ples of these.
The Roeder and Emslie olivine-liquid geothermometer is a rather simple one based on the equi-
librium between magma and olivine crystallizing from it. Consider the exchange reaction:
MgOOl + FeOliq ® MgOliq + FeOOl
where Ol denotes olivine and liq denotes liquid. We can write the equilibrium constant for this re-
action as:
liq
X Ol
FeO X MgO
KD = 4.46
X liq Ol
FeO X MgO
90Fo-10Fa 80Fo-20Fa
20
1350
18
16 1300
14
Mole % MgO in Liquid
1250 70Fo-30Fa
12
1200
10
1150 60Fo-40Fa
8
1100
6 1050 50Fo-50Fa
4 1000
40Fo-60Fa
2 30Fo-70Fa
20Fo-80Fa
0 4 8 12 16 20
Mole % FeO in Liquid
Fig. 4.18. Olivine saturation surface constructed by Roeder and Emslie (1970).
Recalling our criteria for a good geothermometer, we can guess that this reaction will meet at least
several of these criteria. First, olivine exhibits complete solid solution, so we might guess we can
treat it as an ideal solution, which turns out to be a reasonably good assumption. We might also
guess that the molar volumes of forsterite and fayalite and of their melts will be similar, meaning
the DV term, and hence pressure dependence, will be small. As it turns out, the DH term, which is
related to the difference in heats of fusion of forsterite and fayalite, is also relatively small, so the ex-
change reaction itself is a poor geothermometer. However, we can consider two separate reactions
here:
MgOliq Æ MgOOl and FeOliq Æ FeOOl
and we can write two expressions for KD. This was the approach of Roeder and Emslie (1970), who
deduced the following relations from empirical (i.e., experimental) results:
Ol
X MgO 3740
log liq
= - 1.87 4.47
X MgO
T
X Ol 3911
log FeO
liq
= - 2.50 4.48
X FeO T
These KD's are much more temperature dependent than for the combined exchange reaction. Sub-
tracting equation 4.47 from 4.48 yields:
3911 3740
TFeO = and that based on MgO is: TMgO =
Ê X Ol ˆ Ê X MgO
Ol
ˆ
log Á FeO
liq ˜
+ 2.50 log Á liq ˜ + 1.87
Ë X FeO ¯ Ë X MgO ¯
We find that the temperatures of the two methods agree within 6°, which is fairly good. This in-
dicates the analyzed olivine probably was in equilibrium with the liquid.
u lv g n e t
1
yFe 2 TiO 4 + (1 - y )Fe 3O 4 + O 2 s.s.
ma
osp ite
ine s.s.
4
4.50
l–
Ê3 ˆ
® yFeTiO 3 + Á - y˜ Fe 2 O 3
Ë2 ¯ FeO Fe 3O 4 Fe 2O 3
which describes equilibrium between the ul- Figure 4.19. The TiO2–FeO–Fe2O3 ternary
vospinel–magnetite (titanomagnetite) and ilmen- system. Phases are: FeO: wüstite; Fe2O 3:
ite–hematite solid solution series. The equilibrium hematite; TiO2: rutile; Fe2TiO4: ulvospinel;
constant expression may be written as: Fe2O 4: magnetite; FeTiO3: ilmenite. The sys-
tem also includes the FeTi2O5—Fe2TiO5 solu-
tion, which is not shown.
–5 y 3/ 2 - y
a FeTiO a Fe
K= 3 2O 3
4.51
y
a Fe 2 TiO 4
a1Fe- y3O 4 fO1/2 4
p 40
Us Usp 60 The original Buddington and
sp 20 p 80
Ups 85
p0 Lindsley geothermometer was
–10 Us U based on empirical (experimental)
Us
1 00 -
observations of compositional
m
1
He He
( )
Gex = WG1 X 2 + WG 2 X1 X1 X 2
for each solution series. When pressure is neglected, the free energy interaction parameter expres-
sion (equation 4.08) simplifies to:
WG = WH – TWS 4.56
Values for WH and WS were obtained from least-squares fits to experimental data. The parameters
obtained are listed in Table 4.1.
Substituting equations 4.56 and 4.16 into the free energy of solution expression (DG excess =
DGideal – DGreal), the following equation can be obtained:
(
Kexch = XUsp X Hem
2
) 2
X Mt X Ilm , DHo = 27.799 kJ/mol, DSo = 4.1920 J/K-mol
* A magma consists not only of a liquid, but any suspended gas or crystalline phases as well.
(a) (b)
Figure 4.22. (a) structure of pure silica glass and (b) a silica-rich glass
with additional component ions.
† For clarity, we have simplified Ghioso's equation by neglecting H O, which they treated separately.
2
Table 4.3. Interaction Parameters for the Ghiorso Regular Solution Model
SiO2 TiO 2 Al2 O 3 Fe 2 O 3 MgCr2 O 4 Fe 2 SiO4 Mg2 SiO4 CaSiO3 Na 2SiO3 KAlSiO4 Ca3 (PO4 )
TiO 2 26267
Al2 O 3 --39120 -29450
Fe 2 O 3 8110 -84757 -17089
MgCr2 O 4 27886 -72303 -31770 21606
Fe 2 SiO4 23661 5209 -30509 -179065 -82972
Mg2 SiO4 3421 -4178 -32880 -71519 46049 -37257
CaSiO3 -864 -35373 -57918 12077 30705 -12971 -31732
Na 2SiO3 -99039 -15416 -130785 -149662 113646 -90534 -41877 -13247
KAlSiO4 -33922 -48095 -25859 57556 75709 23649 22323 -17111 6523
Ca3 (PO4 )2 613892 25939 52221 -4214 5342 87410 -23209 37070 15572
H2 O 30967 81879 -16098 31406 28874 35634 20375 --96938 10374 43451
Values are in kJ/mol. From Ghiorso and Sack (1995).
1
-DGjop + RT ln aj p - RT  n p ,i ln Xil =  n p .i  X jWGi, j -   X X W k, j 4.66
i i j 2 j k,k π j k j G
The quantities on the left-hand side of the equation are terms that can be calculated from the compo-
sitions of coexisting solids and liquids and solution models of the solids. The right hand side con-
tains the unknowns. One statement of equation 4.66 can be written for each component in each
solid phase at a given temperature and pressure. With enough experiments, values for the interac-
tion parameters can be extracted from the phase relations. Ghiorso et al. (1983) and Ghiorso and
Sack (1995) used a statistical technique called least squares† to determine the interaction parameters
from a large number of published experimental data. Ghiorso and Sack (1995) also noted that the
absence of a phase in an experiment provides thermodynamic information about that phase, i.e.,
that its free energy must be higher than that of the phases that are present. Their approach made
use of this information as well, though discussion of that aspect of their method would take us too
far afield. The interaction parameters they determined are listed in Table 4.3.
With values for the interaction parameters, the model can then be used to predict the assemblage
of solids, their compositions, and the liquid composition that will be present in the system as a func-
tion of temperature and pressure. The equilibrium condition for a magma, as for any other system,
is the condition where Gibbs Free Energy is at a minimum. Thus the problem becomes finding a
composition for the liquid and coexisting solids that minimizes G at a particular temperature and
pressure. In other words, we want to find values of G l and G f1, G f2, ... G fn such that G sys is minimal
where:
Gsys = Gl + Â Gj 4.67
j
Inherent in the problem is finding which solids will be in equilibrium with the liquid for a given
bulk system composition at specified temperature and pressure. In Ghiorso’s approach, an initial
guess is made of the state of the system. This is done by taking the liquid composition as equal to
the system composition and estimating what phases are likely to be in equilibrium with this liquid.
† ‘Least squares’ is a numerical technique that attempts to minimize the square of the difference between
calculated a nd observed value of some parameter. The square is taken to give greater weight to large deviations.
Thus least squares techniques yield results where there are relatively few large deviations between the calculated
and observed value of the parameter of interest. We discuss this technique further in Chapter 8.
90 Tridymite
Plagioclase
Percent Crystallized
70 Na2O MgO
Liquid + K2O
50 Augite
Orthopyroxene Figure 4.23. Calculated compositional evolution of
30 magmas from Thingmuli, Iceland using the regular
10
Pigeonite solution model of Ghiorso compared with actual
Spinel analyses plotted on a FeO–MgO–Na2O+K2O (“AFM”)
ternary diagram. Solid line is for 0.1 MPa pressure,
SiO2 ¥ 13 Shergotty dashed line is for 200 MPa pressure. Both assume ƒO 2
WSeight Percent
TiO2 B O O
B B MgO repetitive matrix calculations and would not be
Na2O O
O
OO
F F FF
Fe
FF 3
2 O practical without a computer, but they can easily be
FOH
OF
HH H HH HH performed on the current generation of computers.
1000 1100 1200 Figure 4.23 and 4.24 illustrate application of the
Temperature, ∞C Ghiorso model. Figure 4.23 compares the calculated
compositional evolution during fractional crystalliza-
Figure 4.24. (a) Cumulative mineral pro- tion of magmas from Thingmuli, eastern Iceland,
portions and liquid composition predicted with actual analyses. During fractional crystalliza-
by the “MELTS” model of Ghiorso and
tion, solid phases are isolated from the liquid and
Sack (1975) compared with experimental
therefore do not continue to equilibrate with the
determined values of Stolper and
magma after they have formed. Figure 4.24a shows
McSween (1979) for the Shergotty
the calculated cumulative proportions of spinel,
meteorite. Shergotty is one of the SNC
olivine, pigeonite, plagioclase, and clinopyroxene
class of meteorites believed to have come
that crystallized from the meteorite Shergotty as well
from Mars (see Chapter 10). From
as the composition of the remaining liquid as a
Ghiorso et al. (1994).
‡ A Taylor series expansion of a function ƒ(z) in the vicinity of some point z = a has the form:
(z - a) (z - a)2
f ( z) = f ( a ) + '(
f a ) + f ''( a ) + K
1! 2!
where ƒ’ and ƒ” are the first and second derivatives of ƒ with respect to z.
or: ( ) (
m NaCl = m oNa + + m oCl - + RT ln m Na + + ln mCl - + RT ln g Na + + ln g Cl - )
Though we can certainly determine the concentrations of Na and Cl in solution, how do we
independently determine their activity coefficients? Since we cannot create a pure Na + solution or
a pure Cl– one, we cannot say what part of the non-ideality of NaCl solution is due to Na+ and what
part is due to Cl–. The practical solution then is to assign all non-ideality equally to both ions. This
leads to the concept of the mean ion activity coefficient:
g ± = (g Na + g Cl - )1/ 2 4.72
Thus the mean activity coefficient of a salt is the multiplicative mean of the activity coefficients of its
component ions. Equation 4.71 then becomes:
(
m NaCl = m oNa + + m oCl - + RT ln m Na + + ln mCl - + ln g ±2 )
Equation 4.72 is valid for 1:1 salts (i.e., 1 cation for each anion). A general expression for the mean
activity coefficient of a salt of composition An+ Bn– is:
+ -
g ± = (g n+ g n- )1/n 4.73
where n is the sum of the component positive and negative ions:
n = n+ + n– 4.74
Mean activity coefficients have the advantage that they are readily measurable (through electro-
chemical means or solubility, for example). Given a well-behaved salt, such as KCl, where the rela-
m ± = m o± +
(
RT ln an+ + ln an-
+ -
) 4.76
n
Rearranging once more, we obtain:
( )
+ - 1/n
m ± = m o± + RT ln an+ an- 4.77
Comparing this relationship with equation 4.71, we define a mean ionic activity such that:
( )
+ - 1/n
a± = an+ an- 4.78
We can also define mean ionic molalities such that a± = g± m±. Substituting a + = g + m + , and a– =
g–m –, we find the mean ionic molality is then:
( )
+ - 1/n
m ± = mn+ mn- 4.79
Mass balance requires that:
m+ = n+ m and m – = n –m 4.80
Substituting this into equation 4.79, we see that:
( )
+ - 1/n
m ± = m nn+ nn- 4.81
Let’s returning to our NaCl example. Dissociation is essentially complete and n+ and n– are
unity, so that:
mNa+ = mNaCl and mCl – = mNaCl
Since n = 2: m ± NaCl = m NaCl
2
= m NaCl
* The use of KCl as a reference for determining mean ion activity coefficients is based on the observation that K +
and Cl– have about the same effective radius and ion mobility and is known as the MacInnes Convention. Like
that of Debye-Hückel, h owever, this approach breaks down at high ionic strength.
aNaCl
activity coefficient as it’s sometimes referred to, of NaCl
would be the square root of the product of the
0.2
component activity coefficients according to equation
4.78, as would the mean ionic activity. The individual 0.2 0.4 0.6 0.8 1.0
ion activities can be measured in a number of ways. 2
Therefore, the above relationships allow calculation of mNaCl
0.6
the mean ionic activity coefficient from measurable (a)
quantities.
0.4
aNaCl
For strong electrolytes, i.e., salts that completely
dissociate, it can also be shown that mean activity
coefficient and mean activity of the salt are related to its
0.2
activity coefficient and activity by:
g = g n± 4.82 0.1 0.2 0.3 0.4 0.5 0.6
and a = an± mNaCl
4.83
We can modify the Debye-Hückel equations to obtain Figure 4.25. (a) Relationship between
activity and molality of NaCl in
mean ion activity coefficients as follows:
aqueous solution. The activity is very
Debye-Hückel Extended Law: low and the “Henry’s Law Slope” is
almost 0 at low concentrations. (b)
Relationship between activity and the
square of molality of NaCl in aqueous
Example 4.6: Mean Ionic Parameters for solution.
a fully dissociated electrolyte
If the molality of a CaCl2 solution is 0.3 M - Az+ z- I
and the activity coefficients of Ca2+ and Cl– are log10 g ± = 4.84
0.5 and 0.7 respectively, calculate the activity 1 + Bå I
and mean ionic molality of CaCl2 in the solu- Limiting Law:
tion. Assume that CaCl2 fully dissociates. log10 g ± = - Az+ z- I 4.85
Answer: For CaCl2, n + = 1, n– = 2, and n = 3. where å is taken as the sum of the radii of the
So we can use equation 3.58 to calculate mean anion and cation, i.e., å = å+ + å– .
ionic molality:
1/3 4.7.1.1 Relationship between Activity and
m ± CaCl2 = mCaCl2 11 22 = 41/4mCaCl2 Molality of a Salt
Substituting 0.3 for m, we find that m± = 0.4762 Let’s consider the relationship between ac-
M. tivity and molality of a salt in an electrolyte
We then use equation 4.81 to calculate the solution such as a NaCl solution. Figure 4.27a
mean ionic activity coefficient: illustrates this relationship. What we immedi-
+ – 1/n 1/3
n n ately notice is that the slope in the Henry’s
g±= g+ g–
( ) = (0.510.72) = 0.625 Law region is essentially zero, which is not at
The mean ionic activity is then: all what we expect for Henry’s Law behavior.
a ± = g±m ± = 0.625 ¥ 0.4762 = 0.2980 It can easily be shown that the
relationship in 4.25a is a simple consequence
and the activity of CaCl2 is: of the dissociation of the NaCl into Na + and
n 3 Cl - ions. From 3.75 we have:
acaCl = a± = g±m± = 0.2980 = 0.0263 M
2
When we plot activity versus the square of molality, we obtain a linear relationship (Fig. 4.25b).
Generalizing this result for dissociation of a substance into a positive ion A and negative ion B,
such as:
+ -
An+ Bn- ® n+ A z+ + n– Bz– A + B - ® n + An + n + Bn
n n
a AB µ mnAB 4.90
For example, n is 3 for CaCl2, 4 for FeCl3, etc. 1.0
Now let’s see what happens if we substi-
tute the mean ion activity for activity. Since: 0.9
n
a = a AB
± 0.8
We have: a = mnAB or
n
a± µ m AB 25∞C
±
This is the relationship that we observed in 0.7
Figure 4.25, so we see that the mean ionic 100∞C
activity accounts for the effects of dissocia- 0.6
tion. g±
0.5
4.7.2 Activities in High Ionic Strength
Solutions 0.4 200∞C
As we noted in Chapter 3, there are two
phenomena that are not considered in the 0.3
Debye-Hückel approach that result in inaccu-
racies in the Debye-Hückel equation at ionic 0.2 300∞C
strengths above about 0.1 m. This is illus-
trated in Figure 4.26, which shows the exper- 0.1
imentally determined mean ion activity coef-
ficient for NaCl as a function of ionic strength 0
0.1 0.10 1.0 10.0
and temperature. At low temperatures, the Ionic Strength, mol/kg
activity begins to increase about ionic
Figure 4.26. Observed mean ion activity coeffi-
strengths of 1 m, whereas Debye-Hückel
cient, g± , of NaCl as a function of ionic strength and
predicts continual decrease. The activities of
many electrolytes eventually exceed 1 at high temperature (solid lines; data from Helgeson, 1981)
compared with value predicted by the Debye-
concentrations. The difference between the
Hückel Law (computed as (gNa + gCl– )1/2 ).
aw = 1 – 0.04m
Table 4.3. Ion Solvation Numbers
Figure 4.27 illustrates the effect of solvation on the
Species h Species h activity coefficient. As may be seen, solvation sub-
Li+ 2.3 OH – 7.6 stantially affects the activity coefficient at ionic
Na + 3.3 F– 6.7 strengths above about 0.5 m.
K+ 2.3 Cl– 2.7
Rb + 2.3 Br_ 1.7 4.7.2.2 Effects of Ion Association
Mg 2+ 8.9 CO 32 - 14.4 An ion pair can be considered to have formed
Ca2+ 8.9 SO 24 - 10.4 when ions approach closer than some critical dis-
Cd2+ 6.3 tance, rc, where the electrostatic energy, which tends
Ba2+ 9.2 to bind them, exceeds twice the thermal energy,
+ – – +–
+
Solvation shells intact Partial disruption of solvation shells Disruption of solvation shells
Figure 4.28. In formation of ion pairs, the solvation shells may remain intact or be partially or
totally disrupted. The former results in an outer sphere ion pair, the latter results in an inner
sphere ion pair.
ions that associate to form ion pairs or complexes. The associate of this fraction of ions as ion pairs
will be thermodynamically equivalent to that fraction of the substance not dissociating to begin
with. The fraction of free ions is then 1 – a. Equation 4.80 becomes:
m+ = (1 – a)n+ m and m– = (1 – a)n–m 4.96
where m is the molality of the solute. We can rewrite equation 4.78 as:
[ ]
1/n
a± = [(g+m+ )n (g–m– )n ]1/n a± = (g +m + )n (g -m - )n
+ – + -
4.97
Substituting 4.95 into 4.96 and rearranging, we obtain:
1/n
( ) [ Ï
] [(1 - a )n m ] ¸
1/n
Ì (1 - a )n m
n+ n-
n+ n- + -
a± = g g + - ý
Ó þ
A little more rearranging and we have:
1/ n
( ) [ Ï
] (n ) (n ) ¸
1/n
Ì (1 - a )m
(n + +n - ) n+ n-
n+ n- + -
a± = g g + - ý
Ó þ
Finally, since n = n+ + n– , we obtain:
1/n
( ) (1 - a )m ÏÌÓ n + ( ) (n ) ¸
1/n n+ n-
n+ n- -
a± = g g + - ý 4.98
þ
We can recognize the last term as m ± . Since a ± = g±m ±, we see that the mean ionic activity
coefficient will be
( )
1/n
g ± = (1 - a ) g n+ g n-
+ -
4.99
for an incompletely dissociated electrolyte. Thus the mean ion activity coefficients are reduced by a
factor of 1 – a. Provided we have appropriate stability constants for the ion pairs or complexes, a
can be calculated and an appropriate correction applied.
Now consider a CaSO4 solution of which some fraction of the Ca2+ and SO 24 - ions, a, associate to
form CaSO o4 . The ionic strength of the this solution would be
I=
(1 - a )
2
(
4mCa 2+ + 4m SO 2-
4
)
153 October 17, 2001
W. M. White Geochemistry
Chapter 4: Applications of Thermodynamics
Thus the ionic strength is reduced by a factor
of 1 - a as well.
0.8
Ion pairs and complexes need not be 0.7
neutral species (AlCl2+, for example). When
they are not, they will contribute to ionic 0.6
strength. A general expression for ionic
strength taking account of ion associations 0.5
must include charged ion pairs and
complexes: g ± 0.4
a max = 0
1È ù
I = ÍÂ (1 - a )mi zi2 + Â c n zn2 ú 4.100 0.3
2Î i n û
where ai is the fraction of each ion involved
0.2
in ion associations, and cn is the concentra- a max = 0.4
tion of each ion pair or complex and zn is its
0.1 a max= 0.2
charge. We could use this result directly in
0
the Debye-Hückel equation to make an im- 0 1 2 3 4 5
proved estimate of ionic strength, and hence
of the single ion activity coefficient.
I (apparent)
Figure 4.29. Effects of ion association on the activity
Figure 4.29 illustrates the effect of ion pair
coefficient. Mean ion activity coefficient of CaCl2 for
formation for a hypothetical CaCl2 solution in 2+
which some fraction of the ions combine to varying extents of– ion association. Fraction of Ca
ions forming CaCl was assumed to increase linearly
form ion pairs. The fraction of Ca2+ ions
with ionic strength up to a maximum value (amax) at
forming CaCl– was assumed to increase
linearly with ionic strength up to the maxi- = 5 m. Solid line shows electrostatic term (Debye-
I
Hückel) after correction for ion association, dashed
mum value shown.
line shows the combined electrostatic and solvation
If the formation of ion pairs depends on term.
the ratio of thermal to electrostatic energy,
we might expect that ion pair formation will
decrease with temperature. However, the relative permittivity of water decreases with
temperature, allowing increased electrostatic interaction between ions, and this effect dominates
over the increased thermal energy of ions. As a result, the extent of ion association increases with
temperature. Increasing pressure, on the other hand, favors dissociation of ions.
4.7.2.2 Alternative Expressions for Activity Coefficients
There have been a number of attempts to develop working equations that account for all the
effects on activity coefficients at high ionic strength. Many of these are ultimately based on the
specific ion interaction theory of Brønsted (1922). Brønsted proposed an equation of the form:
log g i = am1/ 2 + bi m 4.101
where a is a constant that is independent of the solute ions and b is the “specific ion interaction pa-
rameter” and is different for each ionic species. Guggenheim (1935) replaced the first term on the
right with a simplified form of the Debye-Hückel equation and the second term with the
summation of ion-ion interaction parameters:
- zi2 A I
log g i = + 2 Â bi , k m k 4.102
1+ I k
where bik is parameter describing the interactions between ions i and k. For natural waters with
many species, the Guggenheim equation becomes complex. Also starting from Debye-Hückel,
Truesdell and Jones (1974) proposed the following simpler equation:
Species Conc
Example 4.7. Activity Coefficients in a Brine
g/kg
The following concentrations were measured in a shield brine from Sudbury,
Canada at 22° C. Calculate the activity coefficients of these species using the Na 18.9
Truesdell-Jones equation. K 0.43
Answer: Our first task is to convert g/kg to molal concentrations. We do this by Ca 63.8
dividing by molecular weight. Next, we need to calculate ionic strength Mg 0.078
(equation 3.73) which we find to be 5.9 m. Calculation of activity coefficients is SO 4 0.223
then straightforward using the parameters in Tables 3.2 and 4.6. Finally, we HCO3 0.042
apply a correctionfor the decreased concentrationof water (equation 4.91). Our Cl 162.7
final spreadsheet is shown below.
- zi2 A I
1.25 log g i = + bi I 4.103
1.05 1 + Båi I
MgCl2 The first term on the right is identical in
0.85
ies
2. Show that: Gexcess = WG1X2 + WG2X1 X1X2 may be written as a 4 term power expansion, i.e.:
G ex = A + BX 2 + CX 22 + DX 32
3. Construct G-bar–X diagrams for a a regular solution with W= 12 kJ (W is the interaction
perameter in a non-ideal solution) at 100° temperature intervals from 200 to 700° C. Sketch the
corresponding phase diagram.
4. Interaction parameters for the enstatite-diopside solid solution have been determined as follows:
WH-En = 34.0 kJ/mol, WH-Di = 24.74 kJ/mol (assume WV and WS are 0).
a.) Use the asymmetric solution model to calculate DGreal as a function of X2 (let diopside be com-
ponent 2) curves for this system at 100 K temperature from 1000 K to 1500 K. Label your curves.
b.) What is the maximum mole fraction of diopside that can dissolve in enstatite in this tempera-
ture range:?
c.) Sketch the corresponding T-X phase diagram.
5. Sketch G-bar–X diagrams for 1600° C, 1500° C, 1300° C, and 1250° C for the system Diopside-
Anorthite (Figure 4.8). Draw tangents connecting the equilibrium liquids and solids.
6. Suppose you conduct a 1 atm melting experiment on a plagioclase crystal. Predict the mole
fractions of anorthite in the liquid and solid phases at a temperature of 1425° C. Assume both the
liquid and solid behave as ideal solutions. Albite melts at 1118° C, anorthite at 1553°C. DH m for
albite is 54.84 kJ/mol; DHm for anorthite is 123.1 kJ/mol.
8. Determine the temperature and oxygen fugacity of equilibration for the following set of
coexisting iron-titanium oxides in lavas from the Azores:
titanomagnetite s.s. phase ilmenite s.s. phase
mole percent magnetite mole percent hematite
7. Starting from equations 4.23 and 4.62, use the fundamental relationships between free energy, en-
tropy, enthalpy, and the equilibrium constant to derive the temperature
dependence of the titan–omagnetite–ilmentite exchange (equation 4.63). Species Conc
g/kg
10. In a melt having a composition, in wt %, of: Na + 63.00
SiO2 58.12% TiO2 0.92% K+ 6.15
Al2O3 16.47% Fe2O3 1.82% Mg 2+ 2.77
MgO 5.62% FeO 9.94% CaO 7.11% Ca2+ 44.6
Use the Ghiorso regular solution model and the interaction parameters in Cl– 200.4
Table 4.3 to: 2-
– – SO4 0.13
a.) calculate the G ex and DG mixing for this composition at 1300°C. HCO 3
–
0.03
11. An analysis of an oil field brine from Mississippi with a temperature of 150° C is shown to the
right. Calculate the activities of these species at that temperature using the Truesdell-Jones equation.
12. Show that for a strong electrolyte, i.e., one in which dissociation is complete and:
m– = n–m and m + = n+m
where m is the molality of the solute component An+Bn-, that:
1/n
m ± = m nn+ n–n
+ –
13. Mean ionic activity coefficients were measured for the following solutions at an ionic strength
of 3: gKCl = 0.569, gNaCl = 0.734, gNa2 CO 3 = 0.229. Assuming gCl – = gK – = g±KCl,
what is the activity coefficient of CO 32 - ? g±
I, m observed
14. Calculate the electrostatic and solvation contributions to the mean ionic 0.001
activity coefficient of MgCl2 at concentrations of 0.0033, 0.01, 0.033, 0.05, 0.1, 0.003 0.887
0.33, 0.5, and 1 using the Debye-Hückel equation and Robinson and Stokes 0.006 0.847
(equ. 4.92) equ. respectively. Plot results, as well as gelect+solv = gelect *gsolv as 0.01
a function of ionic strength. 0.015 0.78
0.03 0.716
15. Calculate the mean ionic activity coefficient for NaCO3 using the Debye-
0.06 0.644
Hückel and Truesdell-Jones equations and compare your results with the ob-
0.1
served values to the right. Overall, which fits the data better?
0.15 0.541
0.3 0.462
0.6 0.385
1
1.5 0.292
3 0.229
6 0.182