Guest House Reservation Form 06sept2016
Guest House Reservation Form 06sept2016
Guest House Reservation Form 06sept2016
Thermodynamics
of Alloys
Statistical
Therlllodynalllics
of Alloys
N. A. Gokcen
Albany Research Center
Bureau of Mines
U.S. Department of the Interior
Albany, Oregon
N. A. Gokcen
Albany, Oregon
Contents
xi
xii Contents
CHAPTER 7 SEMICONDUCTORS
Introduction ........................................ 195
Distribution of Electrons ............................. 199
Motion of Electrons and Holes. . . . . . . . . . . . . . . . . . . . . . . . 200
Dopants in Semiconductors. . . . . . . . . . . . . . . . . . . . . . . . . . . 202
Energy of Electrons and Holes ........................ 203
Concentration of Electrons. . . . . . . . . . . . . . . . . . . . . . . . . . . . 204
Concentration of Holes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 205
Equilibrium Constant for Charge Carriers. . . . . . . . . . . . . . . 206
Correlation of Band-Gap Data. . . . . . . . . . . . . . . . . . . . . . . . 211
Derived Thermodynamic Properties. . . . . . . . . . . . . . . . . . . . 213
Ionization Equilibria. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 214
Fermi Energy in Doped Semiconductors. . . . . . . . . . . . . . . . 216
p-n Junctions. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 216
Effects of Applied Current on p-n Junctions. . . . . . . . . . . . 218
Solar Cells. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 222
Simple Equivalent Circuit. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 224
Tandem Solar Cells . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 226
Calculation Procedure. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 227
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 236
INDEX.......................................................... 321
Statistical
Thermodynamics
of Alloys
1
Thermodynamic Background
Introduction
The areas of thermodynamics with which this book is concerned are those
dealing with the interrelationships of energy and the changes in properties
of matter upon mixing various components to form various phases. A
component in this book is usually a pure element, but occasionally it may
be an intermetallic or a metal-metalloid compound of well-defined
stoichiometry. A single phase is a substance that has uniform physical and
compositional properties. At a phase boundary these properties change
abruptly.
The concepts of thermodynamics are based on empirical observations
of the macroscopic properties of matter and the results of these observations
are expressed in mathematical functions. Statistical thermodynamics of
solutions strives to obtain thermodynamic relationships based on the
molecular behavior of matter. The mathematical background needed for
our purposes will be developed as we proceed; a detailed background may
be found elsewhere.1.2 A basic knowledge of thermodynamics, as presented
in Refs. 1, 3, or 4, is essential in pursuing the subject matter in this book.
We shall define a term as representing a clear and concise phenomenon,
property, or concept, and designate it by a symbol. All the symbols in this
book are defined as they occur in the text, and are compiled and redefined
in Appendix F. A thermodynamic system, or briefly a system, is the substance
under investigation, separated from the surroundings by rigid or movable
boundaries that mayor may not exchange energy or matter or both with
the surroundings. The composition of a system is usually expressed in mole
fractions, Xi, but the molality, i.e., moles per kilogram solvent, is used for
solutions of electrolytes, and particles per cubic centimeter, for semiconduc-
tors. An extensive property {6 is a function of two of the three variables
2 Chapter 1
Fundamental constants:
Ice + water + vapor point
(triple point of water = 0.0100°C) 273.1600 K
Molar volume of perfect gas
we, 1 atm) 22,413.83 cm' mole-I,
Avogadro number 6.022045 x 1023 mole-I,
(molecules mole-I)
Gas constant R 8.31441 J mole-I K- ' ,
1.98719 cal mole-I K- ' ,
82.05684 cm 3 atm mole-I K -I,
Boltzmann constant k = R/ No 1.380662 X 10- 23 J K- '
Faraday constant F 96484.6 e g-equivalenC ' ,
23,060.4 cal
aFrom D. H. Whiffen, Manual of Symbols and Terminology for Physiochemical Quantities and Units, 1979
Ed., IUPAC, Pergamon Press (\979). See also Dimensions/ NBS, January 1974, a pamphlet by NBS.
Thermodynamic Background 3
G may be expressed by
(1.2)
(1.6)
dH = d (E + PV) = T dS + V dP (1.9)
Comparison of equations (1.6) and (1.9) shows that for pure condensed
substances in their stable states at P = 1 atm = 101,325 Pa, the standard
entropy SO is calculated from
SO-SO(T=O)= i
o
T Co
-P-dT
T
(LlO)
So = fT,
o
C °dT + (AH
-p-
T
-_\-
T\
0) + fT -p
T, T
__ 0) + fT -p-
C/o dT + (AH
2 CliO dT 2_
T2 T2 T
(Ll1)
where Cp 0, C~ 0, and C;o are the heat capacities in the temperature ranges
indicated by the limits of integration.
Thermodynamic Background 5
G = GO( P = 1) + RT In P (1.15)
This is the definition of fugacity wherein GO refers to the unit fugacity, and
again GO is a function of T only. For example, if the equation of state is
PV = ZRT where Z is the compressibility factor, and if for moderate
pressures Z = 1 + aP, with a as a function of T, i.e., a = a(T), then the
substitution of V = ZRTI P = (1 + aP)RTI Pin dG = VdP yields dG =
RT(d In P + a dP); hence, with dG = RTd In F, we obtain din (FI P) =
a dP, and since for P ~ 0, FI P = 1, the integration of this equation from
P ~ 0 to higher values of P yields
For example, at 300 K and toO atm, the correct value of F for hydrogen is
106.3 and equation (1.17), with Z = 1.061 at 100 atm, also yields 106.3. At
much greater pressures of hydrogen, e.g., hydrogen released from metals
and alloys into confined cracks or crevices, it is necessary to use Z as a
power series in P and carry out the foregoing procedure in detail to obtain
the values of F.
6 Chapter 1
(1.18)
(1.19)
( 1.20)
(1.21)
( 1.22)
Thermodynamic Background 7
and call OJ the partial molar Gibbs energy of i, then equations (1.20) and
(1.21) respectively become
(1.23)
(1.24)
(1.25)
(1.26)
( 1.27)
Equations (1.25) and (1.27), first derived by Gibbs (1875) and later indepen-
dently by Duhem (1886), are called the Gibbs-Duhem relations. It should
be noted that n j (i = 1,2, ... ) are the variables of state when a phase is open.
It is important to bear in mind that equations (1.18) to (1.27) are also
applicable to other thermodynamic properties such as E, H, S, and Cpo
Activity
where Ft refers to the fugacity over the pure condensed phase, and G7(g)
is a function of temperature only, i.e., G~(g) = G~( T) and G~(l, or c) =
G~(P, T). However, GW, or c) is so weakly dependent on P that for most
purposes this dependence for condensed phases will be generally ignored
at ordinary pressures. When the liquid or the solid is a solution containing
component i as one of the components, then
(1.29)
8 Chapter 1
where Fi without the superscript * is the fugacity over the solution. Solving
for G~(g) from equation (1.28), i.e., G~(g) = GW, or c) - RT In FT and
eliminating G~(g) in equation (1.29) leads to
The term G~(l, or c) refers to the standard state, which is the pure stable
condensed phase at 1 atm. Therefore, G~ for brevity without the parenthetic
notations is a function of T only. For the standard state, the activity is
always unity.
The dimensionless quantity that relates the activity to the mole fraction
is denoted by Yi and called the activity coefficient; it is defined by
(1.32)
where the approximate equality is due to the close equality of F;/ FT and
PJ PT at low pressures.
A component in a solution obeys Raoult's law (1887) when its activity
and mole fraction are equal, i.e., a i = Xi, or Yi = 1. It has been observed
experimentally that Raoult's law is obeyed in the limiting case when Xi
approaches unity, i.e., ai ~ Xi or Yi ~ 1 when Xi ~ 1. The usefulness of this
limiting law is that Yi is unity in a small but finite range of composition
approaching Xi ~ 1 because otherwise Yi = 1 is the definition of the standard
state. The activity scale based on the pure component is called the Raoultian
activity.
The activity of component i becomes also proportional to its mole
fraction when Xi approaches zero, i.e., a i = Y~Xi for Xi ~ 0 with y~ becoming
a constant, hence the superscript (0). Combination of this equation with
the definition of activity yields
where the last equality is the classical statement of Henry's law (1800), i.e.,
the pressure of i over a dilute solution of i is proportional to its mole
fraction if Fi = Pi and Ft = Pt, with k' as the proportionality constant.
Substitution of a i = Y~Xi in equation (1.31), with G~ == G~(l. or c), gives
where G; is given by
G; = G7 + R T In y7 (1.35)
( 1.36)
a
where i is called the Henrian activity. Since Oi is independent of the choice
of standard states, Oi in equations (1.31) and (1.36) are equal at any
concentration; therefore,
• a· y.
G~ - G = -RTln~=
I I a; -RTln-'
/; ( 1.37)
(1.38)
Therefore,
(1.39)
Thus, the actIvIty on one scale differs from that on another scale by a
constant factor at a given temperature when the concentrations are in mole
fractions. The activities and the activity coefficients based on Raoult's law
and Henry's law are summarized as follows:
Standard Reference
Basis Definition state state
It should be noted that the standard state for both activity bases is the state
of unit activity. The state of unit activity for the Henrian scale is physically
10 Chapter 1
A
1.0,---r---,---,-----"
/ B
/
,---r------,;-----,---" 1.0
HENRY'S LAW/
"'v/
/
0.5 /
1.0
/ t
/ a, 05
/ 10'-
/ /
. /'\ 05
0,
/ /RAOULT'S
,b / / LAW
/
/ ____ __ __ ____
O~_~ __ ~ __ ~ __ ~_ 0 o~ ~ ~~ ~ ~o
o 05 1.0 o 05 10
Xz
Figure 1.1. Activity of component 2 in hypothetical binary solutions. (A) Deviation from
Raoult's law is positive; from Henry's law, negative. (B) Deviation from Raoult's law is
negative; from Henry's law, positive. Vertical right scale is for a 2 with standard state as pure
component 2, and vertical left scale for ti2 with reference state x 2 -> O. The activity curve for
component 1 is not shown, but it is similar to that for component 2 with a, = 1 at x 2 = 0, and
a, = 0 at x 2 = 1, and having a shape as dictated by the Gibbs-Duhem relation.
nonexisting, except in the unlikely case when the solution is ideal in the
Raoultian scale. Further, there is no need for the reference state for the
Raoultain scale, but for the Henrian scale, the reference state is the infinitely
dilute solution of i in a solvent. The activities based on these scales are
illustrated in Fig. 1. The deviation from Raoult's law is positive for com-
ponent i when 'Yi is greater than 1, and negative when 'Yi is smaller than 1.
Similarly, the deviation from Henry's law is positive when}; is greater than
1, and negative when}; is smaller than 1. Generally, when 'Yi> 1, then
}; < 1, and vice versa in the vicinity where the deviation from Henry's law
starts.
Equation (1.13) may be written in the same form for either Oi or G~;
hence, at constant T and composition,
Thermodynamic Background 11
Both V; and V~ are small for condensed phases, and their difference is even
smaller; therefore, a j and 'Yj for real solutions vary insignificantly at pressure
changes of a few bars. For an ideal solution, 'Yj = 1; hence, \1= V~.
The activity a j varies with temperature according to
Ina·= OJ - O~ Fl; -
H~ Sj - S~
(1.41)
I RT RT R
( 1.42)
where the second and third terms after the first equal sign cancel out because
aH;/aT= Cpj, and as;/aT= Cp;/T. We write a similar equation for Of and
subtract it from equation (1.42) to obtain
For an ideal solution a j = X j and the left side of this equation is zero because
Xj is a variable independent of T in OJ = OJ(P, T, Xl> X2, .. .); hence, Hj =
Hf, and the substitution of this equality in equation (1.41) gives
When X2 and Y2 are very small, the left side becomes Y2 - X2 and further,
T becomes very close to Tcr; therefore,
If the solubility Y2 is nearly zero, we obtain the familiar equation for the
depression of freezing point for many ambient solutions. For alloys,
however, when Y2 is greater than X2, the melting point is elevated, and when
Y2 is smaller than X 2, the melting point is depressed. Equation (1.45) can
be rewritten for liquid-vapor equilibria and a similar procedure can be
followed to obtain an equation identical in form with (1.47) for the elevation
of the boiling point.
Molar Properties
A solution of c components may be considered as formed according to
For the solution, the molar Gibbs energy is G = L xlij, and for the pure
components, G = L XjG~ , so that the Gibbs energy of formation of solution
tlG for reaction (1.48) is
I
0
H·)
I
aln~
= -RT2 ~t... x·--·
aT '
I
( ii. _ H~
I I
= -RT2 aIn
aT
')Ij)
(1.50)
*For alloys, 1 mole of solution is identical with 1 gram atom of solution.
Thermodynamic Background 13
- 0 aIn 'Yi)
S--S·=-R Ina·+T-- (
(1.51)
I I aT I
Generally, the accuracy of data on 'Yi for alloys does not justify a rigorous
partial derivative in equations (1.50) and (1.51); instead, a In 'YJaT::::
!lIn 'YJ!l is often used for solutions of metals to obtain !lH and !lS. A
much more accurate procedure is to obtain !lH by calorimetry at various
concentrations and then to calculate !lS from the data on a i by using
!lS = R L Xi In a i - (!lH/T). Figure 1.2 illustrates !lG, 0 1 - G~, O2 - G~,
!lH, and !lG for the binary system AI (component 1) and eu (component
2) obtained by plotting the data compiled by Hultgren et al. 5
0:
"-
<f>
<I
Lo
-4
o
~
Ci
U
.:£
-8
G e = G(reaI) - G(ideaI);
From equations (1.31) and (1.32) with 'Yi =;t. 1 for real solutions, and with
'Yi = 1 for ideal solutions, we derive
6~ = RT In a i - RT In Xi = RT In 'Yi (1.53 )
where the molar enthalpy of formation tJ.H is the same as the excess enthalpy
of solution, He. Equations (1.53)-(1.56) and G~ = fi~ - TS~ lead to
Se = - R ~
L..
(
x· In 'Y'
I I
aaT
In 'Yi)
+ Tx·--
I
(1.58)
Binary Systems
The extensive excess Gibbs energy (lJe = (lJe(p, T, n h n2) for a binary
system is related to G~ by equation (1.22), i.e.,
(1.59)
(1.60)
Substitution of X2/ dXl for (n l + n2)/ dn l from this equation into equation
(1.60) gives
(1.61)
16 Chapter 1
(1.62)
-e
G2 = G e+ (1 - )(aGe)
X2 -- ( 1.62a)
aX2 P, T
-2
o
E
:; -4
u Figure 1.3. Excess molar Gibbs
-'" energy, G e , and excess partial molar
Gibbs energies, G~ and G~, for Al-Cu
system at 1373 K. Subscript 1 is for AI;
2, for Cu. The tangent line on the curve
-6 for G e at X 2 = 0.4 intersects the vertical
coordinate at X 2 = 0 to yield G~, and
at X 2 = 1 to yield G~, as shown by the
dashed lines with arrows.
Note that G~ is equal to G e plus
(1 - x 2 ) times the slope, aG e /ax 2 , as
required by equation (1.62a), and
0.2 0.4 0.6 0.8 1.0 similarly for G.. but the slope in this
case is positive because aGejax\ =
X2 -aG e /ax 2 •
Thermodynamic Background 17
G~ = I (q - 1)Aqx~ ( 1.64)
q=2
(1.68)
( 1.69)
(1.70)
(1.71)
(1.72)
[In general, AI2 is not equal to A 21 ; and the term DJ2XIX2 may be written
as x l xz(D2l x l + DJ2X2 - E\2XIX2) if additional terms are required.] It is clear
that G~ can be obtained from G~ by interchanging the subscripts 1 and 2
and agreeing that DI2 = D21 since D21 does not appear in equation (1.70).
For DJ2 = 0, we have the cubic equations, and for DJ2 = 0 and AJ2 = A2h
we have the familiar quadratic equations of the type given by equation
(1.69) since A 21 x I + AJ2x2 = AJ2. Various additional equations are summar-
ized by Hala et al. 6 ; however, no specific equation can be claimed to have
much greater advantages over other equations in accurate representations
of data.
The effect of temperature on G e and G~ is adequately accounted for
by assuming that each coefficient Au, D ij, and so on is a linear function of
temperature, e.g., AI2 = aJ2 - f312 T. The constant term aJ2 times the compo-
sitional variables contribute to He and lif, and f32 times the same variables
contribute to se and S~. This is self-evident since a(G e /T)/a(1/T) = He,
and aGe/aT = _Se, and similarly for lif and S~.
by using a galvanic cell for measurements of G~(Mg). The results have been
fitted with the following equation:
The corresponding equations for FI~ and S~ can readily be obtained from
this equation. In addition, the equation for the molar Gibbs energy G e can
be obtained by using equation (1.63) from which He and Se can be obtained.
Therefore, the measurements of G~ = RT In 'Yl at various temperatures and
compositions yield G~, FI~, S~, G~, FI~, S~, G e, He. and se.
Ternary Systems
The equations selected for G e of a ternary system must again satisfy
the boundary conditions in that G e = 0 when Xi = 1 (i = 1,2,3). When this
condition is satisfied, the power series beginning with the second-order
terms automatically satisfy the Gibbs-Duhem relation. The derivation of
G~ from G e requires ee e
= (nl + n2 + n3) G and again writing
(1.73)
20 Chapter 1
where n \ signifies that all nj other than n l are held constant during differenti-
ation. The partial differential of XI is
(1.75)
oe = x l xiA 21 x I + A I2 X2 - D 12 XI X2)
( aX2) = __ X2 . (1.77)
aX3 r X2 + X3'
Thermodynamic Background 21
(1.78)
The second of these equations can be obtained from the first by replacing
the subscript 2 with 3. The resulting equations for OJ (i = 1,2,3) by this
procedure are as follows:
( 1.82)
The exchange of subscripts 2 and 3 yields a similar equation for 0;. The
aditional terms in equation (1.76) contribute to 07 as required by equation
( 1.75).
The third-order Margules equation can be obtained from the foregoing
equations by setting Dij = 0 and C l = C 2 = C 3 = C. The results are given here
for convenience and as a check on equations (1.79)-(1.81):
Equation (1.83) contains only one ternary term, C, which must be determined
from ternary data. It is possible to set C = 0 and use only the binary
coefficients to obtain 07 for the ternary alloys with success for liquid alloys,
but marginal results for solid alloys.
The second-order Margules equations can be obtained by setting Al2 =
A2i> A13 = A3i> A 23 = A 32 , and C = 0 in the preceding equations; the results
are
(1.87)
Thermodynamic Background 23
(1.88)
(1.89)
(1.90)
Equation (1.87) contains only the constants from the binary systems; there-
fore, it contains no terms with XIX2X3' The interchange of any set of two
subscripts does not change equation (1.87) since it assumed that Aij = Aji;
therefore, only in such sets of equations, G~ can be obtained from G~ by
interchanging the subscripts 1 and 2, and likewise, G;, by interchanging
the subscripts 1 and 3. Various other types of equations and computational
methods exist, as listed elsewhere in detail. l ,6,9,1O The foregoing equations
are also valid when other molar properties such as llH, IlS, and Se replace
Oe, after which the related partial molar properties can be derived by the
same procedure.
Interaction Parameters
(1.91)
where the identity signs define eY) called the Wagner interaction parameters
or coefficients, 1,12 e~3) giving the effect of component 3 on G~, and e~2>, that
of component 2 on G~. This differentiation is easily carried out after
eliminating XI in equations (1.89) and (1.90), and then setting X2 and X3 to
zero. To generalize eY) for all values of i andj,including i = j, it is sufficient
to write eli) as (aGUax;); the results are
( 1.92)
(1.93)
(1.94 )
Thermodynamic Background 25
= RT In 'Y~ + £ e~i)x;
;=2
(1.95)
where it is assumed that the plot in this range is linear. Similar equations
for component 3 can be obtained by interchanging the subscripts 2 and 3.
A more interesting method, when experimentally possible, is to determine
G~ as a function of composition in sufficient ranges of concentration to
determine A;j in equations (1.88)-(1.90) and then use equations (1.91)-(1.93)
to determine eY).
The values of Oe (in calf g-atom) for the three binary alloys are fitted with
equation (1.70) as follows:
The results for Hg-Zn by Kozin et al.17 are in agreement with equation
(1.99). The foregoing equations could have been represented as functions
of temperature but the ternary data to be used later here are available for
one temperature only,18 i.e., 673 K; therefore, equations (1.98)-(1.100), valid
for 673 K, are adequate for our purposes. The calculations show that the
values of 'Yi for each binary system vary by a factor of approximately 2
from Xi = 0 to Xi = 1; hence, the deviation from Raoult's law is not severe.
The values of a l for Hg in the ternary system Hg-Sn-Zn were deter-
mined by Nigmetova et al. 18 by measuring the vapor pressure of Hg over
various liquid alloys at 673 K. We wish to use equations (1.76) and (1.79)
with C i = 0, i.e., to obtain ternary equations by using the binary coefficients,
in order to test the resulting values with the experimental ternary data. 18
The results for oe and G~ are as follows:
(1.101)
The result from this equation for G~ = RT In 'Yl (at Xl = 0.4, X 2 = 0.3,
X3 = 0.3) is 337.4 cal/g-atom, from which 'YI = 1.287. The experimental
result is 'Yl = 1.285 as read carefully from an appropriate figure by Nig-
metova et al. 18 The agreement is good because the deviation from ideality
is not severe.
Thermodynamic Background 27
(1.103)
where n' means that the numbers of moles other than that inside the
parentheses are regarded as constants. The total differential of G =
p + pY- TS is
dG = dp + P dY + Y dP - T dS - SdT (1.105)
-
G-(a~
- ---
,- an, p,T,n' - an,
(aE) s,v,n'
(1.107)
Likewise, starting with ij = f(S, p, n" ... ), S = fW, Y. n" ... ), and .c\ =
f( y. T, n" ... ), and following a similar procedure, the following additional
definitions of 0, can be derived:
-,- (aH)
G - --=-
an, S,P,n' -
( S)
--T ---
an,
---
E,v,n' -
(aA)
an, V,T,n'
(1.108)
28 Chapter I
References
Single-Component Equilibria
The equality of Gibbs energies in equation (2.1) requires that dGI = dGII,
and since dG = V dP - S dT, it is evident that
dP aH
or - = - - (2.4)
dT TaV
than Vi, as in melting of most metals and vaporization of all elements and
compounds, then dT / dP is positive, i.e., an increase in P causes an increase
in T. On the other hand, when VII is less than Vi, as in melting of As, Sb,
Bi, and H 20, an increase in P causes a decrease in T.
A phase diagram for a single-component system is shown in Fig. 2.1.
Solid and vapor phases coexist along the curve AB; liquid and vapor phases,
along Be; and solid and liquid phases, along BO. The areas represent the
regions of stability for the single phases, and the lines, the coexisting two
phases. At B, all three phases are in equilibrium, and B is thus called a
triple point. The relative slopes of the lines at B are significant in thermody-
namics. The slope of BO is usually very steep because fl V is very small;
in Fig. 2.1 fl V, and hence the slope, is taken to be positive. At the triple
point B, the temperature is designated as TB and the slope of the sublimation
curve AB at T B is
(2.5)
I
I
:~
~I
A '-I I
T-
Figure 2.1. Phase equilibria for one-component system.
Pbase Equilibria and Pbase Diagrams 31
usually about lOOO-fold larger than that of a condensed phase under equili-
brium conditions. Similarly, for the vaporization curve BC,
dP = (AHva p) P (2.6)
dT RT~ B
where AHvap is the enthalpy of vaporization. Equations (2.5) and (2.6) are
also valid at points other than B, i.e., at all possible values of P and T.
According to the first law of thermodynamics, the enthalpy of sublimation
at B is equal to the sum of the enthalpies of melting AHm and vaporization
AHvap, or AHsubl = AHm + AHvap; hence, AHsubl > AHvap. Therefore,
equations (2.5) and (2.6) show that the slope of AB is greater than the slope
of BC at their point of intersection. This requirement necessitates that the
phase boundaries intersect each other at angles less than 180° as shown in
Fig. 2.1. Consequently, the supercooled liquid and superheated solid must
have higher vapor pressures than the corresponding equilibrium phases.
For example, at TE , P E is greater than P F , and since Gg = GO,8 + RT In P,
then G8(over solid at E) > G8(over liquid at F) and for the coexisting
phases Gg(at E) = OS and G8(at F) = G 1; therefore, the preceding in-
equality is identical with G S > G 1 at TE • The solid at E must therefore
transform into liquid at F either directly or by vaporization and condensa-
tion, because the phase with the lower Gibbs energy is the liquid phase.
Thus, the extended portions of all the curves terminate in regions where
they represent nonequilibrium conditions. A pure substance may have
several additional triple points among its solid allotropes at various tem-
peratures and considerably higher pressures than the solid-liquid-vapor
triple point. It will be seen later that the intersecting curves at a triple point
in composition-temperature phase diagrams also obey the requirement that
they intersect one another at angles less than 180°.
Multicomponent Equilibria
(2.7)
32 Chapter 2
(2.11)
Phase Equilibria and Phase Diagrams 33
(2.12)
where (1) the change in Gibbs energy is zero, (2) pressure and (3) tem-
perature are fixed and uniform from one phase to another, and (4) the
system must be closed, and therefore the amount of each component
n), n2, ... ,nc is fixed as indicated by the subscript n. The last condition
requires that
Multiplication of the first equality in this set of equations by OJ, the second
by O~, ... , the last by O~, and subtraction of the results from equation
(2.11) after setting df(5, dP, and dT = 0, yields
(2.13)
For every possible value of each dn{, which is generally nonzero, equation
(2.13) is zero if, and only if, the coefficient of each dn{ is zero; therefore,
o~ = O~' = ... = at
[c( ¢ - 1) equations] (2.14)
G~ = O~' = ... = O~
34 Chapter 2
Phase Rule
The total number of intensive variables that can define each phase is
c - 1 composition variables, and pressure and temperature. For c/> coexisting
phases, therefore, the total number of intensive variables defining the system
is
Y=c-c/>+2 (2.17)
This equation is the formal statement of the well-known phase rule originally
derived by 1. W. Gibbs (1875). The degrees of freedom, or the variance, Y,
represents the number of unrestricted variables of state which may be a set
of variables out of P, T, and c - 1 compositions. It must be emphasized
that Y can be zero or a positive number, but never a negative number. It
should be remembered that the composition variables refer to each
individual phase and not to the bulk composition of a heterogeneous system
of two or more phases. The permissible number of composition restrictions
could be all in one phase or in several phases; for example, fixing the mole
Phase Equilibria and Phase Diagrams 35
(2.18)
Therefore, the absence of some components in some of the phases does not
alter the phase rule. It is often convenient to fix the pressure and decrease
the degrees of freedom by one in dealing with condensed phases such as
for substances with low vapor pressures. The phase rule then becomes
Phase Diagrams
T A
B
T
t c
T Figure 2.2. Hypothetical binary phase
diagrams for liquid (L), solid (S), and
S + L phase fields. Upper curve in each
panel is "liquidus," and lower curve,
"solidus."
Phase Equilibria and Phase Diagrams 37
(1 ) x2 - (2 )
point. The curve showing the start of solidification, called the liquidus, is
tangent to the curve showing the completion of freezing, called the solidus,
at the same extremum point in Fig. 2.2(8) and (C). This requirement is
necessitated by the Gibbs-Konovalow theorem,t-6 which can be proved by
using the Gibbs energy G written as G = nlO I + n202, and functionally as
G = G(P, T, nJ, n2). The total differentials of these equations are
dG = n l dOl + n2 d0 2 + 0 1 dn l + O 2 dn2
dG = YdP - $dT+ 0 1 dn l + O 2 dn2
These equations yield Y dP - $ dT = n l dOl + n2 d02, and after division
with (nl + n2), and then imposing the constant-pressure restriction, we
obtain
(2.20)
38 Chapter 2
a + L
a + ABx ABx + E
CA) XB - CB)
Figure 2.4. Hypothetical phase diagram forcatatectic (metatectic) and eutectic transformations,
and for compound AB capable of dissolving its components to limited extents.
where S is the molar entropy. This equation, rewritten for two phases, yields
where OJ is the same for both phases for a given temperature. Rearrangement
of the preceding equations and division by dx~ yields
(2.21)
PROOF. The compositions of both phases are equal, hence X!I - x! and
X11 - x~ in equation (2.21) are both zero; further, dOl! dx~ and d0 2! dx1
are both finite because 0 1 and O2 are continuous functions of T and X2,
Phase Equilibria and Phase Diagrams 39
and Sl - Sll is finite and nonzero; therefore, dT / dx~ must be zero. Equation
(2.21) can be divided by dX~1 instead of dx~, and the same argument can
be followed to show that dT / dX~1 is also zero. Therefore, both curves must
be horizontally tangent to each other and this completes the proof.
Eutectic-Type Reactions
The liquidus and solidus lines are often depressed as shown in the
upper left of Fig. 2.3 in such a way that at a particular composition and
temperature, the liquid and two solids, i.e., three phases, coexist. The
reaction upon cooling is liquid(LI) ~ solid(a) + solid(I3), which is called
the eutectic reaction. Above and below the eutectic temperature, one phase
must disappear according to the phase rule. There are other similar reactions
involving three phases with various possible combinations. Each reaction
is named according to the states of aggregation of phases and the number
of reactant phases. The suffix "tectic" is used for reactions involving one
Eutectic
" \~i ~ Liquid(L\) -+ solid(a) + solid(f3)
~
Monotectic Liquid(L3) -+ liquid(L2) + solid(e)
Syntectic Ly /f3\ 42
Liquid(L\) + liquid(L2) -+ solid(f3)
Catatectic 7
(metatectic) ~ 'tI 4 SoIid(f3) -+ soIid( a) + liquid(L)
Eutectoid
1 'fl ~ Solid(f3) -+ soIid(a) + soIid(A)
Peritectoid
P M~ Solid( a) + solid(A) -+ solid( 1/)
40 Chapter 2
or two liquid phases, and the suffix "tectoid" for three solid phases. Various
types of three-phase reactions are presented in Figs. 2.3 and 2.4 and summar-
ized in Table 2.1 with the coexisting phases marked as in these figures.
These transformations consist of (1) eutectic-type, in which the reactant
is a single phase, and (2) peritectic-type, in which the reactants consist of
two phases. Thus, there are four eutectic-type and three peritectic-type
transitions in the preceding list.
A critical point exists when one phase dissociates into two phases at
a point where the phase boundary has a horizontal inflection point, i.e.,
when aT/ aX2 and a2T / ax~ are both zero as shown in Fig. 2.3 for L ~ LI + L2,
L ~ L2 + L3 , and 8 ~ 8 1 + 82. It should be noted that the maximum point
in the 17-region does not have an inflection point and the phase boundaries
for 17 and u are tangent to each other at their common maximum points.
The maximum point in the center of Fig. 2.4 is for a congruently melting
compound, AB, capable of dissolving its component elements to limited
extents as indicated by the phase region AB x.
The limits of solubility of one solid phase in another solid phase is
sometimes called the solvus curve, or briefly, the solvus. Thus, the E-phase
boundary below the eutectic in Fig. 2.4 is the solvus, which is the limits of
solubility of ABx in E. However, the E-phase boundary above the eutectic
is the solidus.
The preceding diagrams contain all the possible phase equilibrium
types encountered in the condensed phase binary systems. Various compila-
tions of phase diagrams exist,S-ll and recently evaluated diagrams are
published in the Bulletin of Alloy Phase Diagrams (a bimonthly journal
begun in 1980, published by ASM-NBS).
Erroneous Diagrams
A number of important aspects of the phase equilibria summarized in
Figs. 2.2-2.4 must be observed to avoid errors in drawing the phase
boundaries. All such errors violate (a) the phase rule, (b) the Gibbs-
Konovalow theorem, and (c) the requirement that the extended portions
of the phase boundaries terminate in the two-phase regions. Examples of
these violations are illustrated in Fig. 2.5 and discussed as follows.
a. Along Be, a, Lh L 2 , and f3 coexist and thus violate the phase rule
since the degree of freedom Y with four coexisting phases is -1 and Y
cannot assume a negative value. This error can be corrected by joining LI
and L2 at one point on the straight line Be. On D, a, 10, A, and f3 coexist
because at E there are more than two curves and one straight line intersecting
one another; the phase boundary curve ME must therefore not terminate
at E. Accordingly, there must be no more than one straight line and two
Phase Equilibria and Phase Diagrams 41
Figure 2.5. Errors in a hypothetical phase diagram violating phase rule or other thermodynamic
principles.
Lever Rule
a
o
.!: -50
o b
u
c5
<J -100
second component. Let meA) and m(B) represent gram atoms of AI and
A2 at A and B, respectively, with the restriction that meA) + m(B) = 1.
Atomic balance requires that
meA) _ XB - Xc CB
(2.22)
m(B) Xc - XA AC
which is known as the lever rule. Note that Xc representing the bulk
composition of two phases is not a composition variable for thermodynamic
properties of each phase such as dO and dGj • Equation (2.22) may be
transformed into a useful form by adding 1 to both sides of the first equality
and simplifying by using meA) + m(B) = 1 so that
(2.23)
where the second relationship is obtained from the first by using m(B) =
1 - meA).
where the last term is the molar Gibbs energy of mixing of an ideal solution,
dO(ideaI) = RT(xl In XI + x21n X2) with RT taken to be 600 cal/mole
(301.93 K) as a convenient simple quantity. The remaining coefficients in
equation (2.24) were obtained by using one maximum point M and two
minimum points near A and near B. The tangent line ACB to the curve for
44 Chapter 2
which is in accord with the lever rule because flO of a mixture of two
phases is the sum of flO of its constituent phases. Equation (2.25) is linear
and it is represented by the straight line between A and B. Likewise, the
straight line DEF represents flO of two phases having the compositions
corresponding to D and F. The value of flO(at E) is larger than that of
flO(at C); therefore, any straight line joining two points on the curve AMB
has higher values of flO than those corresponding to the straight line ACB
at the same values of Xc; hence, a phase mixture along DEF is unstable
with respect to that along ACB for the same bulk composition.
The curve for flO in Fig. 2.6(b) has two inflection points, one at I
(X2 = 0.331) and the other at J (X2 = 0.791) as can be shown by substituting
these values of X2 in the second derivatives of equation (2.24) with respect
to X 2 • These points are called the spinodes, which are important in kinetics
of nucleation and growth of new phases from the supersaturated single
phases. The equations for the activities of components can be derived and
the results can be plotted versus X2. It can be shown that the maximum and
the minimum points in such activity versus composition diagrams for both
components coincide with the spinodes.
The shape of the curve in Fig. 2.6(b) changes with increasing tem-
perature as required by the phase diagram. Thus, A and B approach each
other and finally minimum, maximum, and the inflection points coincide
at the horizontal inflection point Q in Fig. 2.6(a) as required by the phase
diagram.
The diagram shown in Fig. 2.6(a) is for a single solid phase decomposing
into two solid phases. For all other types of transformation, it is customary
to represent flO versus X2 for each phase at each selected temperature on
the same diagram. If the phases in equilibrium are solid and liquid, the
Phase Equilibria and Phase Diagrams 4S
convention for writing /lG for each phase through the entire range of
composition at a selected temperature T is as follows:
where G(I) and G(s) are the molar Gibbs energies of the liquid and solid
solutions. Equation (2.26) at X2 = 1 becomes /lG(I) = G(I) - G 2(s) when
the stable pure phase for component 2 is solid at T, and since for X2 = 1,
G(I) is the same as G 2(1) for the pure liquid 2, then /lG°(I) is identical
with /lG2,m of melting 'for pure component 2. If, however, the stable phase
for component 2 is liquid, /lG(I) in equation (2.26) is then zero at X2(1) = 1.
Reconsider equations (2.26) and (2.27) at a temperature T greater than
the melting points of both components so that
Addition of xI(s)Gi(s) + xz(s)G 2(s) to the right side of equation (2.29) and
then subtraction of these terms from the same side, followed by a simple
rearrangement, yields
/lG(s) = G(s) - XI(S)G~(s) - X2(S)G 2(s)
(2.30)
where /l Gtm is the standard molar Gibbs energy of fusion of pure component
i. The first three terms after the equal sign can be transformed into
RTxl(s) In al(s) + RTx2(S) In a2(s) by using the definition of activity, i.e.,
Gj(s) = Gi(s) + RTln aj(s) and observing that G(s) = XI(S)G I + X2(S)G2.
Likewise, the three terms on the right side of equation (2.28) can be rewritten;
the results for equations (2.28) and (2.29) are therefore
/lG(I)= RTxI(l) In al(I) + RTxz(l) In az(1) (2.31)
/lG(s) = RTxI(s) In a1(s) + RTxz(s) In a2(s)
- x1(s)/lGi,m - X2(S)/lG 2,m (2.32)
A simple example for these equations is now presented, with the following
arbitrarily assigned values:
TI,m = 900K /lH':,m = 900R /lSi,m =R
46 Chapter 2
If, in addition, the solution is ideal so that the activities can be set equal
to the mole fractions, equations (2.31) and (2.32) become
Equations (2.34) and (2.35) are plotted in Fig. 2.7 for T = 1000 K. We note
that 6.G(s) for the solid in Fig. 2.7 is higher than 6.G(l) for the liquid;
-'G.~ 2.m
800
o
..,....... -.1G 1• m = 100 R
-800
-1600~---L----~--~----~--~
o 0.2 0.4 0.6 0.8 1.0
X2
Figure 2.7. Diagram for 40(1) and 40(s) at 1000 K from equations (2.34) and (2.35),
respectively. 40(S) > 40(1) throughout, and solid phase is thus unstable relative to liquid
phase. Phase diagram for this system is in lower portion of Fig. 2.8.
Phase Equilibria and Phase Diagrams 47
hence, the liquid phase is stable relative to the solid phase at 1000 K. The
difference between the vertical intercepts (at x\ = 1 and at X 2 = 1) for those
curves represent ~G~.frz = +100R and ~G~.frz = +700R where ~G~frz =
-~G~m with the subscripts frz and m referring to freezing and melting,
respectively. The positions of the curves below 300 K are reversed, i.e., the
curve for G(I) is above that for ~G(s) because the solid phase is stable
relative to the liquid phase below 300 K as can be shown by plotting a
different but similar figure.
Next to be considered are a set of curves at 550 K for which the liquid
and solid phases coexist at appropriate concentrations. These curves are
given by
Equations (2.40) and (2.41) are represented in the upper portion of Fig.
2.8 by the curves marked LIQUID and SOLID respectively. The straight
line tangent to both curves gives the compositions of solid and liquid at
X2 = xis) = 0.450 and X2 = X2(I) = 0.709, respectively. Above the tangent,
a pair of phases have a higher value of ~G than the phases at the points
of tangency as discussed in conjunction with Fig. 2.6. The liquid phase is
stable from X2 = 0.709 to X2 = 1, and the solid phase is stable from X2 = 0.0
to X2 = 0.450, as required by the relatively lower values of ~G represented
by the lower sections of the curves. At the point of intersection of the curves,
~G(I) and ~G(s) are equal but 0\(1) and O\(s), as well as Oil) and Ois),
are not equal; therefore, there is no equilibrium at this point. The vertical
intercepts of the curves in Fig. 2.8 correspond to ~G~.m = 350R and
-~G2.m = 250R, as indicated in the figure. The solidus and liquidus points
given by the upper portion of Fig. 2.8 are shown in the lower portion. The
repetition of the foregoing procedure for OJ at various temperatures gener-
ates the entire phase diagram as shown in the lower portion of Fig. 2.8.
48 Chapter 2
:--- 0
600 +I1G, m = 350R
C) ~
~ -200
~
-400
-600
'""'-
o
I
I
en l
I
on 01
'000 ~ '"':1
~ I 0 I
'" 1 "",1
XI xl L
750 ~ I I
I I
I~ I
I- I _~_
500
Figure 2.8. Upper diagram is for
s S + L "'"
~G(I) and ~G(s) at 550 K from
equations (2.36) and (2.37),
250 ~--~----~----~----~--~ respectively. Phase diagram in
o 0.2 0.4 0.6 0.8 1.0 lower portion is obtained from
X2 equations (2.43) and (2.44).
The phase boundaries in Fig. 2.8 have been drawn by equating OJ(l) =
O;(s) and then substituting various values of T to solve for Xj(s) and
Xj(l). In this region G~ is solid but G 2 is liquid, The liquid phase is
ideal, and for 0.(1) it is evident that 0.(1) = G~(1) + RT In x.(1), and like-
wise, O.(s) = G~(s) + RT In x.(s), and the equality of 0.(1) and O.(s)
yields
900 xl(l)
l--=ln-- (2.43)
T xl(s)
(2.44)
These equations represent the solidus and liquidus curves. For example, at
550 K, XI(l)/XI(s) = 0.5292, and xil)/x2 (s) = 1.5755, and then substitution
of 1 - xl(l) = x 2(l) and 1 - xl(s) = x 2(s) reduces the number of unknowns
in the first set of equations by two, leading to X2(S) = 0.450 and X2(l) = 0.709;
these points are indicated in Fig. 2.8.
Deviations from ideality modify the curves for AG represented in the
preceding figures; however, the principles involved in the representations
are basically the same. The curves similar to those in Figs. 2.6-2.8 for
nonideal solutions require substitution of al = I' IX1 and a 2 = I'2X2 in
equations (2.31) and (2.32), and expansion to generate a set of terms
expressed by G e = RT(xl In 1'1 + x2ln 1'2)' Appropriate analytical equations
for Ge(l) and Ge(s) must then be added to equations (2.34) and (2.35),
respectively, to represent them as functions of composition and temperature
for nonideal solutions. A detailed example of such a procedure will be
given in Chapter 6 in thermodynamic calculations of the iron-carbon phase
diagram.
a+p
the two-phase regions where AG is lower for the stable phases than the
unstable phases. Figure 2.10 shows the diagrams at Te and TIll indicated
on the phase diagram in Fig. 2.9. At Tn three phases coexist as shown by
the single tangent line to the three curves. As the temperature decreases,
the curve for the liquid, AGO), moves up and the remaining curves move
down so that the a-phase at e is in equilibrium with the {3-phase at h for
TIll. Assume that the liquid and a phrases are supercooled as shown by
the extended dashed portions of the phase boundaries in Fig. 2.9. The liquid
at g, and the a-phase at f must then coexist as shown in Fig. 2.9 and in
Fig. 2.10, by the tangent line at f and g. The location of the curve for AG(l)
at TIll necessitates that the tangent at f and g be in the two-phase region,
and since the values of AG for the two phases at f and g are higher than
Phase Equilibria and Phase Diagrams 51
o 1.0
c d
those at e and h, the phases at f and g are unstable. The extended portions
of the phase boundaries in Fig. 2.9, terminating in the two-phase regions,
are therefore properly constructed. This is possible when the angles between
the adjoining curves and the eutectic line at b and at d are less than 1800
as stressed earlier.
(2.45)
_all_G---,-(a_t_C..:..) = o. (2.49)
aXA '
For this purpose we label X 2 in equation (2.24) once as X A = x 2 (A) and then
as XB = x2(B) to obtain the equations for IlG(at A) and IlG(at B), respec-
tively. These equations must then be substituted in equation (2.25) and then
in equation (2.49) to obtain two simultaneous equations with two unknown
mole fractions XA and XB' The solution of these equations yields XA = 0.170,
XB = 0.919 at RT = 600 cal/mole (301.93 K). The calculational detail\ shows
that Xc cancels out in setting equation (2.49) to zero.
Phase Equilibria and Phase Diagrams S3
that of Cd, 1.52 A, and Cd is therefore about 19% larger than Cu, and only
1.7 at. % Cd is soluble in Cu. In contrast, the atomic radius of Zn is 1.37 A,
which is 7% larger than that of Cu, and Cu-Zn form extensive solid
solutions. The radii recommended by Hume-Rothery are half of the closest
distances of approach of atoms in crystals of the pure elements, but in the
examples cited here, the conclusion is not significantly affected when other
atomic radii are used.
Within the favorable range of 15%, known as the favorable size-factor,
other properties of atoms play important roles; consequently, the favorable
size-factor is not sufficient for substantial terminal solubilities. For example
Mg (1.60 A) and Sb (1.61 A) have nearly identical atomic radii, but their
mutual solubilities are less than 0.04 at. %. This rule is therefore a negative
rule; i.e., if the size difference between two metals exceeds 15%, then the
mutual solubilities are limited, but if the size difference is less than 15%,
the mutual solubilities mayor may not be extensive. Theoretical justification
for this rule has been made by Friedel, Blandin, Eshelby, and others.2
Rule 2, electronegativity factor. Formation of stable intermetallic com-
pounds, more appropriately stable intermediate phases, will restrict terminal
solid solubilities. The possibility of formation of such compounds increases
with increasing differences in the electronegativities of component metals.
In general, when the difference in Pauling electronegativities exceeds
±0.4 volt, the solubilities are restricted even when the size-factor is favorable.
This rule is also known as the rule of electronegativity effect.
Rule 3, electron concentration factor. In many alloy systems an important
factor that determines the extent of solubilities in terminal and intermediate
phases is the electron concentration, which is usually expressed in terms of
electrons per atom of alloy, e/at. The compositional range of existence of
each phase having a particular crystal structure is frequently observed to
correspond to a narrow range of electron concentration.
Computation of the e/at. ratios is made by adding the electrons con-
tributed by each pure component element and then dividing the sum by
the number of atoms in the alloy. The number of electrons contributed to
e/at. by each element is not a universally accepted quantity. Usually, but
not always, the outermost sand p orbitals are considered to contribute to
the electron concentration. 24 Indeed, in the B-subgroup elements with paired
d-electrons, it is generally agreed that only the s- and p-orbitals contribute
electrons. (d-electrons are completely paired when there are ten such elec-
trons in the same orbital.) However, the convention is by no means universal
for the transition and noble elements in which the inner orbitals are assumed
to contribute to the electron concentration. Any method of selection is likely
to be controversial, but the main objective is to develop a consistent empirical
picture correlating the phase regions with the e / at. ratio. A set of assigned
Phase Equilibria and Phase Diagrams 55
values for selected elements, commonly used for computing e/at. ratios,
are listed in Table 2.2.
For a better fit of experimental results, numerous investigators have
presented arguments for assigning variable valencies to Cr, Mn, Fe, Co,
and Ni. When phase diagrams are plotted as temperature versus e/at.,
several interesting features and similarities among various systems are
exhibited. Unfortunately, there is no general agreement on the electron
concentration ranges for various phases encountered in various binary
diagrams. In addition, frequently a phase of known crystal structure in one
binary diagram does not appear in another binary diagram in the same
range of e/at. ratio. Evidently, this as well as other rules encounter greater
degrees of success for the elements in particular groups in the periodic
chart. Greater degrees of divergences and greater numbers of exceptions
occur when greater numbers of metals are considered. It is therefore advis-
able to apply these rules with appropriate modifications to limited selected
groups of elements for greater degrees of success.
The Cu-Zn diagram is shown in Fig. 2.11, in which Cu and Zn
contribute one and two electrons per atom, respectively. Two simple struc-
tures among six encountered in the Cu-Zn system are (1) the bcc ,a-phase
that exists at 1.36 to 1.55 in e/at. (this phase is sometimes written as CuZn
to show its roughly equiatomic composition), and (2) the cph e-phase that
occurs at 1.78 to 1.87 in e/at. (sometimes designated as CuZn3). Such phases
are frequently called electron phases, and they often show no definite
stoichiometry to be called electron compounds. A few examples of electron
phases are listed in Table 2.3 where (1) the limit of maximum solid solubility
of Cu corresponds to e/at. = 1.4, (2) bcc structure appears as Zn is added
in Cu when e/at. = 1.5, (3) y-phase boundary corresponds to e/at. = 1.6
to 1.65, and (4) cph boundary for the e-phase is variable but e/at. lies
about 1.8 for most binary alloys of the noble metals. For each group of
elements a similar scheme can be devised with different values of e/at. for
Group e/at.
Table 2.3. Ranges of Electron/ Atom Ratios for Selected Electron Phases
fee
upper Minimum "y-phase a eph
Alloy boundary bee boundary boundary boundary (e)
"Complex cubic.
b Lower limit is 1.0.
'Terminal Zn-rich 1)-phase is also cph, and contains only 3 at.% maximum Cu, with e/at. = 1.97 minimum;
see Cu-Zn phase diagram (Fig. 2.\1).
Phase Equilibria and Phase Diagrams 57
•c
1500
.. , ..
\
'
......
,.'
N.-Cr
1300
J •
1100 I
r\: . I
I I F. -v
, I Fc-Cr
I I ~-)(-)l-X C.-V
I I Q •• 0 Co-C.
NI.V
Fc-V~! :
• • • • • N:-Cr
-~I
Fc-Cr _ _ 1
900 ,
t
\ I
'. I
'-
References
Statistical Thermodynamics
of ways we can select one letter from the first set is 3, and from the second
set, 2, so that there are 3 x 2 ways of making both selections in the stated
sequence, i.e., AE, AF, BE, BF, CE, and CF. An obvious extension of this
axiom is for more than two sets when there are Nh N 2, N 3, ... independent
particles; the total number of arrangements is then N\ x N2 X N3 X ••••
The number of distinct ways of arranging N different objects all taken
together is N!, and this result is called the permutation of N different
objects. The deduction is easy for three letters A, B, C, because the first
letter can be selected three different ways, after which the second letter can
be selected two ways, leaving one.letter to be selected last so that we have
ABC, ACB, BAC, BCA, CAB, and CBA, or 3! = 6. Among the N objects,
if N\ were alike, then N\ ! of all the arrangements would be indistinguishable;
therefore, the total number of distinct arrangements (or distribution) is
(3.1)
Boxes: gh g2,.··, gj
Energy levels: eh e2, ... , ej
Occupation numbers: Nh N 2, ... , N j
Each set of gj of the ith set of identical boxes differing from other boxes
represents the degeneracy of that level as given by quantum mechanics, i.e.,
the number of single-particle wavefunctions that yield the corresponding
single energy level. Note that gh g2, ... are the degeneracies, but eh e2, ...
are themselves not degenerate. The set Nh N 2, ... , N taken altogether is
j
Fermi-Dirac Statistics
in it. The number of boxes, each containing one particle, is Nt. and the
number of empty boxes is g\ - N\. The number of ways, Dt. for arranging
g\ boxes, of which N\ are alike and each contains one particle, and g\ - Nt.
which are empty and also alike, is given by equation (3.1), i.e.,
(3.2)
D= n
j
j=\
D = n
j
j
,.
g.!
j=\ Nj!(gj - N;)!
(3.3)
In D =.L j [
gj In (g.)
~ + N In (g. - N)]
-'--.-'
j (3.4)
,=\ g, N, N,
The particles obeying equation (3.3) or (3.4) are called fermions, typical
examples of which are electrons, neutrons, and protons which, according
to the Pauli exclusion principle, cannot occupy the same state with another
particle, or no two such particles can have the same quantum numbers.
Bose-Einstein Statistics
If we remove the restriction in the preceding section that each box may
contain no more than one particle, then the distribution becomes completely
different. Each box may now contain any number of particles; thus, for five
particles and four boxes, one of the possible arrangements is
A AAA A
Four boxes are generated by inserting the dashed, movable partitions, and
since there are three of these partitions, the number of movable partitions
is one fewer than the number of boxes so that we have g\ - 1 partitions.
The total number of objects to be arranged is 8, or g\ - 1 + N\. The number
64 Chapter 3
(3.5)
D= n; (g;-I+N;)! (3.6)
;=\ N;!(g;-1)!
and when the Stirling approximation is used after neglecting 1 in (g; - 1),
we obtain
In D =L; {- g; In (g)
,=\
- .-'-. + N; In
g, + N,
(g + N)}
-'--.-'
N,
(3.7)
The particles (e.g., photons) that obey equation (3.6) are called bosons.
Boltzmann Statistics
Consider equation (3.2) for fermions when g; is very much greater than
N;, i.e., g; » N j ; hence, most of the boxes are not occupied and equation
(3.2) can be rewritten as
(3.9)
In D = L [N j In(gJ N j ) + NJ (3.11)
i=1
The particles obeying equation (3.10) are called the corrected boltzons, and
for the most part we shall be concerned with these particles. In the historical
development of equation (3.10) the right-hand side contained N! as a factor
immediately after the equal sign and such particles were simply called the
uncorrected boltzons. We shall have no specific use for uncorrected boltzons.
The preceding equations for D are functions of N j only since degeneracies
are generally not variables.
Distribution Laws
and
(3.14)
This assumption states that only the maximum term of D in very close
proximity of E contributes significantly to D, or stated differently, if D
were plotted versus energy for a closed system, D would sharply peak out
at the prescribed value of E for the system. The justification for using the
maximum term for D is that the resulting statistical thermodynamic
equations accurately describe the observable thermodynamic properties.
For the maximum value of D, d In D is zero; hence,
i aln D
d In D = 0 = L - - dNi (3.15)
i~1 aNi
Li (alnD aN
--. + a - .- {3ei
)
dNi = 0 (3.16)
,=1 aN, aN,
Since each dNi is finite and nonzero, equation (3.16) is satisfied if, and
only if, the coefficient of each dNi is zero; hence,
aln D aN
- - + a - - {3e = 0 (3.17)
aNi aNi '
aln D
--+ a - {3e = O' (i=I,2, ... ) (3.18)
aNi "
Equations (3.12), (3.13), and (3.18) provide the simultaneous equations for
solving the unknowns N i , a, and (3.
For fermions, the first term on the left side of equation (3.18) is equal to
In(gi - N i ) -In N i ; because gi is independent of Nio therefore,
Ni 1
or - = ---:;:--- (fermions)
gi e -ex e{3e, + 1 '
(3.19)
Likewise,
(bosons) (3.20)
Statistical Thermodynamics 67
and
(boltzons) (3.21)
Equations (3.19)-(3.21) are called the distribution laws for systems consist-
ing of independent particles.
The values of a and f3 can now be obtained for boltzons. For this
purpose, we rewrite equation (3.12) by using equation (3.21) so that
(3.22)
(3.23)
and call q the molecular partition function. If the energy, Ej, is separable
to Ej and Ej' then q = qll; = (L gj e-/3£,) (L gj e-/3£j). For example, Ej and Ej
may refer to kinetic and potential energy, respectively. Equation (3.23) can
be substituted in equation (3.22) to write
eOq = N (3.24)
We solve for eO from this equation and substitute it into equation (3.21)
to obtain N j and then use equation (3.13) to obtain
(3.25)
~A~""'B--i-I- Y
e
x
1-1 cm-f
Figure 3.1. One-centimeter cube containing an ideal gas.
A molecule at "A" may move toward "B" to collide with the surface at
"B" and then rebound and travel to "e" and return to "A" again to repeat
its journey. In this process, each molecule travels a distance of 2 cm for
each collision on the selected surface dBe. When a molecule rebounds from
a surface, its average velocity changes from +u to -u and the change of
its momentum at the surface becomes mu - ( - mu) = 2mu. According to
Newton's laws of motion, the force exerted on a surface by a moving body
is equal to the rate of change of momentum and since this force acts on
dBe which has an area of 1 cm 2 , it also represents the pressure. The time,
t, necessary for each collision is t = (2 cm)/(u cm/sec) = 2/ u, and thus the
rate of change of momentum is 2mu/(2/u) = mu 2 which is the pressure
due to each molecule. For all the molecules moving in the direction of y,
i.e., N /3 V, the pressure is
P = (N /3 V)mu 2 (3.26)
(3.28)
Statistical Thermodynamics 69
The energies, E;, are functions of volume only. The reversible work
(dw) done on a system is given by dw = L N;( dEJ dV) dV = L N; dE;. From
E = N;E;, dE = L (E; dN; + N; dE;); however, from the first law of
thermodynamics, dE = dqth + dw where qth is the thermal energy exchange.
Consequently,
(3.30)
E·
dS = ~~dN (3.31)
'-T I
aln D
~ - - dN =
'- aN; I
~
'-
(f3E. dN - a dN)
I I I
(3.32)
The last term is zero because for a closed system a L dN; = a dN = o. The
left side of this equation is equal to L d In D where D is a function of N;
only. If the most probable value of D is the only significant contribution
to L d In D, then the summation can be replaced by the most probable
value of Dmp which we designated again as D so that
1 E·
d In D = f3 ~
'-
E dN
I I
= -k'-T
~ ~ dN
I
(3.33)
70 Chapter 3
dS = kd In D (3.34)
The integratin of this equation yields the definition of the entropy, i.e.,
S = kIn D (3.35)
(3.36)
where the last equality was obtained by using equations (3.12) and (3.13).
The Helmholtz energy is given by
q
A = E - TS = -kNTln-- kNT (3.38)
N
G = -kNTln!L (3.39)
N
Nj e" e- f3e ,
(3.41)
gj e-" e f3e , +1 1 + e" e- f3e ,
(3.43)
Nj
(3.45)
gj [e f3 (e,-G)] +1
£.=6
N =I Nj = 'I gj; (T~ 0) (3.46)
Ej=O
e f3 (G-e,)
1 + e f3 (G e,)
(3.47)
72 Chapter 3
Nj f3(G-g) N -f3g
-=e '=-e' (3.48)
gj q
1
(bosons) (3.49)
Again for ej » G and high temperatures, this equation becomes the same
as equation (3.48) for boltzons.
Gibbsian Ensembles
Equations (3.12), (3.13), and (3.23) through (3.25) show that q refers
to a closed system described by constant E, N, and V. A microcanonical
ensemble is therefore a closed system defined by E, N, and V. (Canonical =
standard.)
Consider equation (3.38), which can be rewritten as
qN
A = -kTln- (3.50)
N!
A = -kTln Q; (3.51)
where equation (3.39) is used on the right side, and 0, like G, is a function
of N, P, and T. We shall generally have very little use of E and 0 in the
topics considered in this book, and frequently use q and Q.
(3.54)
where i is the index that counts the distinguishable and degenerate states
individually, one by one. According to quantum mechanics, the energy of
a free particle is given by
(3.55)
(3.56)
74 Chapter 3
(3.57)
where ABC = V is the volume of the box. The energy E is obtained from
E = -N (alnaf3 q) l.5N
(3.59)
v f3
The energy of an ideal gas is given by E = 1.5 NkT, hence, again f3 = 1/ kT.
The free electrons obey the same treatment except a spin degeneracy
factor of 2 (electron spin up and spin down) must appear in equation (3.58);
thus,
J(J + 1)h 2
Ej = 8rr 2 I (J = 0, 1, ... ) (3.61)
Statistical Thermodynamics 75
where J is the rotational quantum number. The equation for qrot is similar
to equation (3.57) and can be integrated with J as the variable from zero
to infinity for sufficiently large values of T to obtain
(3.62)
H rot = RT (3.63)
-Lhp/kT _ 1
L
_ 00
qvib - e - 1 -hp/kT (3.64)
L=O - e
The last equality is obtained easily because the summation represents the
sum of a geometrical series. For sufficiently high temperatures and small
frequencies, hv/ kT is small and qvib is very closely approximated by
The Gibbs energy and enthalpy per mole from equation (3.65) are
H=RT (3.66)
E A B
-=--+- (3.68)
No 12 r6 r
Statistical Thermodynamics 77
E(min) 2B B B
-e = = --+-=--
Nor~2 r~2 r~2
from which B = er~2, and then A = 2er~. Therefore, equation (3.68) becomes
(3.69)
This equation, known as the Lennard-lones 6-12 potential, has been used
extensively for gases and liquids (see Refs. 5 and 6). It represents an
intermolecular pair potential in gases, but the concept has been extended
to the condensed phases.
0.6.---",-.---.----.---.---.---.---.
0.4
0.2 r-
t O.S 1.0 1.2 1.4 1.6 I.S 2.0 2.2
i. 0
No
- 0.2
-0.4
-0.6
- 0.8
- 1.0
-1.2~--~--~--~--~--~--~--~--~
Figure 3.2. Lennard·Jones 6-12 potential for a pair of molecules. Vertical scale is based on
£ = J, and horizontal scale, TO = 1 in equation (3.69). Minimum point is at T = TO = 1 and
E(min)/No = -£ = -I.
78 Chapter 3
Machlin 7 has used a function similar to equation (3.68) for the structural
stability of alloy phases, lattice parameter defects, and other related
properties:
a {3
E=--+-
r4 r8
(3.70)
-E=
No
(1
Me 2 - - +r---)
m I
r mrm
(3.71)
In general, all the potentials near the minimum energy point can be
closely approximated by an empirical quadratic function, i.e.,
(3.72)
References
Theories of Solutions
The most general and simple theory intended to fit liquid and solid solutions
is the regular solution theory developed by vanLaar and Lorenz (1925),
Heitler (1926), and Hildebrand (1928). The term regular solution was
proposed by Hildebrand (1929) for solutions described by random
and yet nonideal behavior. This proposed behavior was properly called
the zeroth approximation to the regular solutions by Guggenheim,l as
distinct' from the first approximation, to be discussed later in this
chapter. Early developments in this field are presented in a number of
publications. 1- 3
(4.2)
(4.4)
We assume that the energy E is nearly identical with the enthalpy H, because
the pressure-volume product PV in the definitional relationship H =
E + PV is negligibly small for condensed phases at ordinary pressures.
Likewise, the enthalpy for xjN molecules of pure condensed i is xjH~ =
(ZN /2)x j e jj where H~ refers to 1 mole of pure i. The corresponding enthalpy
for pure j is xjH'j = (ZN /2)xAj, and for the mixing process, i.e.,
(4.5)
Theories of Solutions 83
where equation (4.4) is substituted for H to obtain the last equality, and
W;j, called the exchange energy, is defined by
The enthalpy change, fl.H, is also called the molar enthalpy of formation
of solution. The arguments presented for the foregoing equations can be
readily extended to muIticomponent solutions to show that
(4.8)
(4.10)
Or---r---r---r---r-~
", .. 1..6
t -0.2
6G A
RT
-0.4
..........
1.0
t B
Gz 0.4
Figure 4.1. Regular solutions for various
values of a = NW / RT. Upper curves are for
0.2 e.G/ RT. Dashed horizontal line is tangent at
J and K to curve for e.G/ RT, with a = 2.6.
O~~~~~~--~~ L is critical point. Lower curves are for activity
o 0.2 0.4 0.6 0.8 1.0 G 2 of component 2; activity G, is symmetric
x
2
- to G 2 about vertical line through X2 = 0.5.
(4.11)
* e.G(ideal) is sometimes unnecessarily called the configurational Gibbs energy and e.S(ideal),
configurational entropy.
Theories of Solutions 85
(4.13)
(4.14)
are both zero for a = 2 at X 2 = 0.5; this is the critical point. At a > 2, the
curves for ll.G/ RT show two minima; e.g., for a = 2.6, the minimum points
are at J with X2 = 0.124 and at K with X2 = 0.876 and J and K are located
symmetrically with respect to each other. In the range of composition
between J and K, the solution separates into two phases, where the composi-
tion of each phase is constant but the relative proportions of these phases
vary from J to K. The tangent line to the curve at J and K is horizontal and
the intercepts with the vertical axes are given by ll.G 1/ RT = ll.G 2 / RT =
ll.G/ RT, with the values of ll.Gi calculated either at J or at K. The negative
values of a yield curves similar to that for a = 1 but with sharper dips.
Equation (4.14) for a> 2 is zero at the inflection points; e.g., for
a = 2.6, X2 = 0.252 and X2 = 0.748 are the inflection points, which are called
the spinodes.
The equation for ll.G2 can be derived from equation (4.12); the
result is
(4.15)
Preliminary Concepts
The first approximation to the regular solutions has been obtained in
its correct form by permutation of molecules. 4 The earlier approximation
86 Chapter 4
(ZN/2)x; = numberofj-jbonds
For this distribution, we denote for brevity that
(4.16)
a b
Figure 4.2. Permutation of bonds in two·dimensional crystals. Bonds of i-j type in (a) form
spliced molecules of the types in (b) after permutation.
Theories of Solutions 87
where Y is the parameter that gives the correct number of i-j bonds when
multiplied by Z. If we accept the concept of half-bonds for the moment,
then there are 2ZY half-bonds, half of which, i.e., ZY, have to be i half-bonds
emanating from the molecules of i, and the other half, ZY, have to be j
half-bonds emanating from the molecules of j. However, prior to mixing,
the pure i molecules have ZNj half-bonds of i type; therefore, after mixing,
the i half-bonds, not connected to the j half-bonds, are ZNj - ZY, and these
half-bonds must meet with each other to form i-i bonds. Likewise, the total
number of j half-bonds connected to the j half-bonds is Z~ - ZY. The
reader can verify these relationships by using unidimensional crystals, i.e.,
beads in necklaces as in Fig. 4.3. The total number of whole bonds of each
type in the mixture is therefore
i-j bonds = ZY
i-i bonds = (Z/2)(N j - Y) (4.18)
/~"
Figure 4.3. Molecules in unidimensional pure crys·
tals i andj, and their solution i-j; Z = 2. i-j bonds =
Zy = 4 in the bottom configuration; hence, Y = 2,
N j - Y = 4, and ~ - Y = 2.
\.
--
i-j
()---()- --
.J
88 Chapter 4
Distribution of Molecules
One-Dimensional Crystals
We shall first obtain the distribution of molecules for Z = 2 and then
extend it to Z = 4 to 6, where Z = 6 occurs in simple cubic and two-
dimensional hexagonal crystals. We take another set of arrangements shown
in Fig. 4.4 to derive the total number of arrangements for given values of
Theories of Solutions 89
a b
Figure 4.4. Arrangements of molecules in unidimensional crystals (Z = 2). Y = 3 for (a),
Y = 2 for (b). Open circles are i molecules; black circles, j molecules.
The same formula is also applicable to the j molecules; hence, for both
permutations we have
We remove the restriction on the system that I, II, and III have fixed
positions by rotating each configuration by one molecular position at one
time and repeat this procedure N times. When the rotations are constructed
on paper for all the configurations for a given value of Y, it is seen that
each configuration is unnecessarily repeated by a factor of Y. The number
of ways of arranging the molecules for a given value of Y is therefore N / Y
times the preceding expression, i.e.,
This equation is similar to but not identical with the Ising equation for
ferromagnetism. A stringent test for equation (4.23) is that the sum of D
for all the permissible values of Y is exactly equal to the sum of all the
possible configurations given by the random distribution, D rm :
yo
D rm = L D = N!/Nj!~! (4.24)
Y=m
Two-Dimensional Crystals
A two-dimensional crystal lattice with a coordination number of 4 is
shown in Fig. 4.5. The molecules in the crystal occupy the points of
intersection of the lattice. It is also possible to construct a close-packed
hexagonal two-dimensional crystal with Z = 6 but we shall not be concerned
with it because of geometrical difficulties in enumerating various configur-
ations. The dangling bonds in a crystal having a limited number of molecules
cause a relatively large error in our enumerations but for a very large number
of molecules, again the error becomes quite negligible. We can eliminate
this error by tying the dangling bonds i, 2, 3, and 4 with the lattice points
1 2 3 4
T T T T
I I I I
I I I
a' I --- a
A 8
b' eI --- b
I
I
c' L
---I C
f 9
d' 0 C
-~ d
l' 2' 4'
Figure 4.5. Two-dimensional crystal lattice with Z = 4.
Theories of Solutions 91
1',2', 3', and 4', respectively, away from the reader on the reverse side of
the page and in such a manner that we form three squares, e.g., 1,2, 1',2'
form a square. This process is best visualized if the lattice is stretched and
wrapped over a sphere so that 1,2, 1',2' become quadrangular in shape.
Similarly, a, b, c, and d are joined to the lattice points a', b', c', and d',
respectively, to form three more squares. The four corners A, B, C, and 0
form their square in this process so that there are altogether 16 squares for
16 lattice sites. (There are, likewise, N cubes for N lattice sites in a simple
cubic system for which Z = 6.) The corners A and C, as well as Band 0
are diagonally opposite to each other. This arrangement shows that any
configuration, such as the dashed L located on efg, returns to its original
position after four diagonal jumps along BO.
lt is convenient to formulate a procedure which can make the enumer-
ation of configurations clear and systematic. For this purpose, we take the
number of one of the component molecules either equal to, or larger than
the other without losing generality, e.g., N j ; " ~, and write a bond balance
for ~ molecules from equation (4.I6), i.e.,
This equation states that there are altogether Z~ half-bonds joined to the
j molecules, and two j-j of these are the j-j half-bonds and the remainder,
ZY, are the i-j half-bonds. Equation (4.25) is useful in computing Y from
the known numbers of j-j bonds. As an example, we consider N j = ~ = 8
in some detail. The number of configurations for all the permissible values
of Y has been counted and listed in Table 4.1. A stringent check on the
results is that the total number of configurations for all the possible values
of Y be equal to that given by equation (4.24).
When j-j is zero, Y is 8 and j molecules occupy diagonally opposite
positions, leaving the remaining positions for the i molecules. The sets of
positions for i and j molecules may be all interchanged once; therefore,
the total number of configurations is 2. The configurations for the values
of 1 and 2 for j-j are geometrically impossible as can be verified by
construction. The next configuration contains the cluster for which j-j = 3,
and there is only one such cluster as shown in Fig. 4.6{a). The four remaining
j molecules are surrounded entirely by the i molecules and they can be
arranged only one way. The entire configuration can be rotated four times
in 90° steps and placed in 16 different ways in the lattice, or stated differently,
Fig. 4.6{a) can be rotated in four ways and placed in 16 different positions.
The resulting number of configurations is 4 x 16 = 64. The configurations
containing the clusters of j-j = 4 are shown in Fig. 4.6{b)-{e). A selected
molecule in the cluster shown in Fig. 4.6{b) can be placed in 16 different
92 Chapter 4
I I
+ • • •
.L L'
• • • I I
a b c d e
........
L1
u
-1 -- ~ ]
0 ~
I
u
0
f 9 h I J
Figure 4.6. Arrangements of molecules for three, four, and five j-j bonds for N, = Nj = 8 and
Z=4.
lattice points and for each position the remaining three j molecules may
be arranged in four different positions and thus yield 4 x 16 = 64 configur-
ations. Figure 4.6{c) and (e) give 16 and 8 configurations, respectively,
because of symmetry, and Fig. 4.6{d) gives 32 configurations. The total
number of configurations for j-j = 4 is therefore 120. The configuration for
j-j = 5 is shown in the remaining parts of Fig. 4.6. Each cluster and its
mirror image in (f), (g), and (h) can be rotated four times and placed in
16 different sites to give 4 x 2 x 16 = 128 configurations for each figure and
384 configurations for three figures. The arrangement in (i) can be rotated
four times and thus yield {4 + 4 + 4)16 = 192 arrangements. The total for
j-j = 5 is then 576 configurations. The process of enumerating the configur-
ations for j-j = 6 is a very time-consuming task. There are three single
clusters of 6 j-j bonds involving squares with two attached bars that can
be rotated four ways and yield 3 x 4 x 16 = 192 configurations, and 15
one-piece clusters similar to (f) giving 1216 configurations. In addition,
there are nine two-piece clusters similar to (i) yielding 704 configurations,
all adding up to 2112 configurations. The results for j-j = 7 to 12 are listed
in Table 4.1 and it is readily seen that the total number of configurations
is equal to D rm • Similar enumerations for N j = 9 and ~ = 7 are also listed
in Table 4.1. The last column for the permutation of bonds, g, will be
discussed on page 107.
Theories of Solutions 93
Actual
arrange-
System j-j bonds Y y/y* ments D g
aD is the distribution of molecules; g, the distribution of bonds; D,m is given by equation (4.24).
Table 4.2. Arrangements of Molecules for Various Values of N j and N j and for Z = 4
Actual Djactual
System j-j bonds Y YjY* arrangements arrangements
Actual D/actual
System j-j bonds Y y/y* arrangements arrangements
where I is a fraction. For sufficiently large values off, e.g., I> 50, Stirling's
approximation may be used to obtain the numerical value of I!. The
maximum value of YI y* is given by (Nj + Nj)1 N j and decreases with
decreasing values of Njl N j for our convention of choosing N j ;;;. Nj. The
complexity of constructing and counting the configurations for high values
of both Nand Y I y* has limited our enumerations to relatively low values
of YI Y* for high values of N. The values of 10g(D/actual arrangements)
are plotted versus Y I y* as distinct points in Fig. 4.7. The results may be
96 Chapter 4
8
G N,' 84, NJ'16
0 N,' 90, Nj'IO I
I
'V N,' 92, Nj ' 8 I
0 N, '95, NJ' 5 / 6
ON,' 97, NJ' 3 I
~ N,' 884, NJ'16 I s
ON,' 890 NJ'IO I
9 N,' 892, NJ' 8 4~
SCALE ill u
en
D<N,' 895, Nj , 5
(} N,' 898, NJ' 2 /
I
/ 2
3
- =l. .I.oLF>J-=~: :" -
ZERO o
w
<i
u
2
en
summarized as follows: (1) All the points for a given value of N may be
fitted by a curve independent of the value of NJ ~. The dotted curves pass
through the actual points for N = 16 and N = 36 and the solid curves
ending in broken terminal portions on the left and on the right pass through
the points for N = 100 and 900. The dotted line for N = 25 is so close to
that for N = 36 that it has not been drawn except on the right side where
there is a small systematic difference between the two sets as shown by the
Theories of Solutions 97
broken lines. The nature of the solid lines ending in broken lines for N = 16
and 36 will be discussed later, but in this paragraph the dotted curves for
N = 16 and 36 and the solid curves for N = 100 and 900 will be considered.
It is clear that D / actual arrangements is a function of Nand Y / Y* and
independent of NJ~. (2) The range of Y / Y* wherein the points follow
a concave-up curve becomes wider with increasing N on both sides of the
vertical line passing through Y/ y* = 1, or Y = y*. (3) The minimum point
at Y = Y* approaches the abscissa and becomes tangent to it with increasing
N. (4) The initial points on the left, marked as 1 close to the vertical axis
in Fig. 4.7, scatter more than the others because of the coarseness of the
clusters. A certain degree of scattering is also due to the fact that the values
of N j and ~ are rather small, and as they approach Avogadro's number
the scattering would undoubtedly smooth out.
We next consider the solid/broken lines for N = 16 and 36 in Fig. 4.7.
If we lower the dotted curves vertically so that they become tangent to the
abscissa at Y = Y*, we obtain very nearly the set of solid and broken
curves. We observe that as N increases, the distinction between the dotted
and solid curves disappears, and the dotted curve becomes tangent to the
abscissa as shown by the curves for N = 100 and 900. The solid portions
of these curves which extend from about 0.6 to 1.4 for Y / y* have actually
been obtained in the following manner. We correct equation (4.23) to obtain
Dc so that D/ Dc represents the solid portions of the curves, or Dc is the
actual number of arrangements for sufficiently large values of N. The
correction c, and the resulting corrected distribution function Dc are as
follows:
b = (Z - 1)/(2Z - I) (4.28)
D = ________~(_N~j_-_l~)!~(N~j-__l~)!_N__________
c (Y+c)!(Y+c-l)!(Nj - Y-c)!(~- Y-c)!
_ Nj!~!
(4.29)
- [( Y + c) !f( N j - Y - c)! (~ - Y - c)!
ijoth c and b were obtained by trial and error with the requirement that
c = 0 for Z = 2 and for Y / y* = 1. The empirical forms of c and b are not
unique because other similar formulas can be obtained for reasonably close
values of Dc. The correction given by equation (4.27) is negative for
Y/ Y* < 1, positive for Y/ Y* > 1, and zero for Y = y* because equation
(4.23) represents the correct number of configurations at Y/ y* = 1. This
98 Chapter 4
Actual D/actual
Nj Nj D rm j-j y y/y* arrangements D arrangements
+ 10
-10
-20
-30
-40
-50
that c is a very small and negligible fraction of each set of terms in each
pair of parentheses in equation (4.29) when N; and ~ become large enough
to be comparable to Avogadro's number. For example, at YI Y* = 1.4, c
is about 7.3% of Y + c for N; = ~ = 50, but for N; = ~ = 450, it is only
2.2 % of Y + c. Likewise, at Y1 y* = 1.2 the corresponding percentages are
4.1 and 1.2 for N = 100 and N = 900, respectively.
The configurations for the simple cubic system having Z = 6 are much
more difficult to enumerate. The error from the dangling bonds is eliminated
by a procedure similar to that for Z = 2 or 4, but for the limited number
of clusters we shall consider, it is important to observe that the number of
cubes is equal to the number of molecules and the total number of bonds
is ZN 12. A limited number of enumerations are listed in Table 4.3. The
results in the last column show that equations (4.27)-(4.29) are also satisfac-
tory for Z = 6. A greater number of enumerations with j-j = 2 to 5 and
~ = 4 to 6 is possible but quite time-consuming.
The variation of Dc with Y1 y* for N; = ~ = 450 and Z = 4, or an
equimolecular crystal of 30 x 30, is shown by the upper curve in Fig. 4.8.
The circles represent the actual points from Fig. 4.7 for D/actual arrange-
ments and show that DI Dc and D/actual arrangements are identical and
therefore Dc is equal to the actual number of arrangements. The lower curve
for log (g1actual arrangements) versus Y1 Y* will be discussed on page 107.
Equations
(H = E) (4.30)
where e;;, ejj , and e;j are the energies for i-i, j-j, and i-j types of bonds,
respectively, and further, H = E. Substitution of equation (4.6) for W = W;j
in equation (4.30) gives
(4.31)
(4.32)
100 Chapter 4
The value of Y for which this equation is the maximum term in equation
(4.33) is obtained by differentiating In Q(max) with respect to Y and setting
the result to zero:
where the approximate equality holds for large values of Y; therefore, the
term b y b - I is not included in this equation. As N = Ni + ~ approaches
Avogadro's number, yb / Y approaches a value close to zero in the range
of 0.5 < Y / y* < 1.5. In addition, e in each of the logarithmic terms is
negligible in comparison with Y, Ni - Y, and ~ - Y; therefore, we could
have obtained an equation equivalent to (4.31) by using D of equation
(4.23) instead of Dc in equations (4.29) and (4.33). We rearrange equation
(4.34) and neglect e to derive
y2 = e- W / kT = e- NW/ RT (4.36)
(Ni - Y)(~ - Y)
where R = Nk is the gas constant. The left side of this equation is the law
of mass action or the pseudo-equilibrium constant for the following
reaction:
This reaction states that one molecule of i with Z half-bonds of i-i type
and one molecule of j with Z half-bonds of j-j type, both on the left and
shown with solid half-bonds, react to form one molecule of i with Z
half-bonds of i· .. j type and one molecule of j with Z half-bonds of j . .. i
type shown with dotted lines. The number of molecules of the reactants are
(Ni - Y) and (Nj - Y) and the products are Y and Y for the respective
species in reaction (4.37). The corresponding energy effect is clearly W =
(Z/2)(2eij - eii - ejj ). Reaction (4.37) is our short-range order-disorder
reaction.
First Approximation
n = e W / kT (4.38)
(1 - n) y2 - NY + NiNj = 0 (4.39)
(+ sign unacceptable)
(4.40)
The set with the plus sign after N is not acceptable for two reasons: (1)
The maximum value of Y for Ni = Nj is Y = Ni for a perfectly ordered
crystal in which the number of j-j bonds is zero [(see equation (4.25)].
However, for n < 1, the plus sign would give Y > N i • For example, if
n = 0.64 and Ni = Nj = N /2, then Y = 2.5N (2) The value of Y cannot
be negative but for n > 1 it becomes negative for the plus sign in equation
(4.40) and positive for the minus sign. Therefore, we accept equation (4.40)
with the minus sign after N
The square root in equation (4.40) is inconvenient; hence, for simplicity
we introduce
(4.41)
102 Chapter 4
(4.42)
(4.43)
(4.45)
The lower integration limit for the left side is zero because at very high
temperatures or for l/T = 0, Ge/T is zero. We wish to express d(T- 1 ) in
terms of f3 and for this purpose we differentiate equation (4.38) as follows:
dn = n(W/k)d(rl) (4.46)
We solve for n from equation (4.41) and then eliminate n from equation
(4.46) to obtain
(4.47)
Theories of Solutions 103
where 1 - 4XiXj is equal to (Xi - XJ2 after we use (Xi + xj ? = 1 for a binary
system. The substitution of equation (4.47) into equation (4.45) gives
G=
_
T
e
Nk f
f3~1
f3
({3
4x·xf3
+ 1)({3 + Xi
IJ
df3
- Xj)({3 + Xj - Xi)
(4.48)
Ge = NkTxi In [ f3 +
(I
X - X.]
/ + NkTxj In
[f3 +( JX - X]
)I (4.50)
Xi f3 + 1 Xj f3 + 1
- e
Gi = NkT In [f3 +( Xi - Xj]
); G-ej = NkT In [f3 +( Xj - Xi]
) (4.51)
Xi f3 + 1 Xj f3+1
j
S e = 2 NXiXj W - Nkx In [f3 + Xi - X] - N kx . In [f3 + Xj - Xi] (4.52)
T(f3 + 1) Xi(f3 + 1)
I J xj (f3 + 1)
For an ideal solution W = 0, f3 = n = 1, and equation (4.52) is zero as
expected, because se for a random solution is zero by definition. The
equations for 5~ and 5j are
- - [f3 + Xj -
s-e =NWxj Xi]
-Nkln [f3 + Xi - Xj] (4.53 )
I T f3 (f3 + 1) Xi (f3+ 1)
104 Chapter 4
(4.54)
He = 0.5NW/(~ + 1) (4.56)
As a convenient test for equations (4.55) and (4.56), we take the experimental
value of G e and solve for ~ and then compute W from equation (4.57).
Next we substitute ~ and W into equation (4.56) to calculate He. The
reason for using this procedure is the simplicity of computation. The
calculated and the experimental values of He are compared in Table 4.4
for a number of randomly selected alloys.4,7 It is evident that the first
approximation is good when G e and He do not differ greatly, i.e., when
the entropy term in G e = He - TS e is small. There is a substantial degree
aOe and He are in calories for one gram atom (or mole) of solution at XI = x, = 0.5
(I cal = 4.184J).
Theories of Solutions lOS
of success in the results and in fact, for about half of the alloy systems for
which the data are reliable/ the calculated and the experimental values of
He agree fairly well.
Approximate Equations
The remaining partial molar properties can be derived from the molar
properties as shown in Chapter 1. It is important to remember that these
equations exclude the majority of alloys for which W/kT is larger than 0.5,
i.e., INWI > 0.5RT = 1000 cal/mole at 1000 K. However, it is likely that
for the alloys for which INW / RTI > 0.5, the criteria for regular behavior
cannot generally be met; hence, equations (4.58) may be useful for systems
in which the deviations from ideality are not great.
Short-Range Order
The relationship expressing Y* for the random distribution can be
obtained by setting the exchange energy W to zero in equation (4.38) to
obtain n = 1, leading to (3 = 1, from equation (4.41) and y* = Nx;xj from
equation (4.42). The ratio Y / y* is therefore
Y/ Y* = 2/({3 + 1)
106 Chapter 4
y f3 - 1
a=l--=-- (4.59)
s y* f3+1
The equations for dilute binary regular solutions according to the first
approximation can be obtained by setting one of the mole fractions close
to zero. Since the equations derived in the previous sections are symmetric
with respect to composition, it is sufficient to consider the case for Xj -'» O.
The equations for He and H7 are zero as expected, and the equation for
H'} is
where n = e W/kT has been used to obtain the last equality. Since equations
(4.60) and (4.61) are equal, it is evident that
Consequently, in dilute solutions, the solute can have only one type of
configuration in which the j molecules are entirely surrounded by the solvent
molecules, and this arrangement corresponds to the random dispersion of
the j molecules in the solvent i molecules. The foregoing relationships can
also be derived by using equations (4.58).
The first approximation that existed prior to 1970 was based on the
permutation of bonds as developed mainly by Bethe 5 and by Guggenheim.!
It was shown earlier in this chapter that such a permutation leads to
physically impossible spliced molecules. While the method has been the
subject of numerous papers, some based on abstruse mathematical argu-
ments, the most elegant and rigorous method is due to Guggenheim.!
The numbers of i-i, j-j, and i-j bonds used by Guggenheim are also
given by equation (4.18). The corresponding bonds for the random distribu-
tion involve y* of equation (4.16). The energy E is also given by equation
(4.30). However, Dc used in equations (4.29) and (4.32) are quite different,
since the bonds are used in permutation. Guggenheim assumed that an
observer is capable of distinguishing an i-j bond from a j-i bond and
therefore there are ZY/2 of i-j and ZY/2 of j-i bonds. There are altogether
ZN/2 bonds consisting of Z(N; - Y)/2, ZY/2, ZY/2, and Z(Nj - Y)/2
bonds of i-i, i-j, j-i, and j-j bonds, respectively. The permutation of these
bonds is given by
, (ZN/2)!
(4.63)
g = [Z(N; - Y)/2]!(ZYj2)!(ZY/2)![Z(Nj - Y)/2]!
seen later. The sum of g for all values of Y must be equal to all the possible
configurations, i.e.,
The sum in equation (4.64) is then replaced by its maximum term g(max) =
hg'(max). To accomplish this, hg' is differentiated with respect to Y and
the result is set to zero to solve for the corresponding value of Y. This
solution gives Y = Y* = Nxjxj, and when Y* is substituted in g(max) =
N!I (Nj! ~!) = hg', an equation is obtained for h. The substitution of h
into g = hg' gives
(4.65)
Q = L ge- E / kT (4.66)
y
The sum in equation (4.67) can be replaced by its maximum term to obtain
the equation for Y. For this purpose, aIn Q'I aY is set to zero in the same
way that equation (4.34) was set to zero. The solution of the resulting
equation for a In Q'I aY = 0 yields
It is now evident that whether we had used g' or g = hg', we would still
obtain the same preceding equation because Guggenheim assumed that h
was independent of Y. The left side of equation (4.68) is the equilibrium
constant for the following reaction among the bonds:
where Stirling's approximation for factorials has been used. The value of
D rm is expressed by
where the expression in the center is for any value of Y obtained by using
In( 1 + a) = a and the last equality is for Y = L/4. It is seen that D and
Dc differ by a much smaller factor than g and Dc. For Z = 2, equation
(4.70) becomes comparable to equation (4.72) as log L is much smaller than
Lb. This is evidently due to the fact that L and the number of permutable
bonds become identical for Z = 2. However, the permutation of bonds
again creates spliced molecules.
The thermodynamic equations resulting from equation (4.68) require
that n of equation (4.38) must now be written as
n = e2W/ZkT (4.73)
g
Ge(bond permutation) = - 2 -
ZNkT{ Xi In [b+X.-X]
(P ~ 1)J +
-Xi Xj
[b+X-X]}
In -xj (P J+ 1) I
(4.74)
y2 = e- W / ZkT .
(Ni - Y)( ~ - Y) ,
1 1 1 1
- (i-i) + - (j-j) = - (i . .. j) + - (j ... i) (4.76)
2 2 2 2
where the coefficient 1/2 in this reaction comes from the fact that W / Z
contains 1/2 as shown in equation (4.75). The difference between equation
(4.75) and Guggenheim's equation (4.68) is that the latter contains 2 W / Z =
Theories of Solutions III
(eij+ eji - eii - ejj ) in its exponent, or a quantity twice as much as in equation
(4.75). The mathematical procedure and some of the postulates in the
surrounded atom model are interesting; however, the resulting thermody-
namic relations, starting with equation (4.75), are subject to the same
criticism as equation (4.74) based on the permutation of bonds.
(4.77)
where the last two terms are for the j-k system and identical in form with
the preceding terms. The equations for {3ij require writing Wij • .• for pairwise
interactions of net numbers of atoms. This relatively complicated equation
can be simplified by using equation (4.58), i.e.,
Liquid ternary alloys of some elements such as Mn, Fe, Co, and Ni are
expected to obey qualitatively this equation.
The liquidus for these binary systems at the melting point in the vicinity of
the compound has a temperature peak, the viscosity of liquid is a maximum,
and the conductivity is a minimum. Further, the enthalpy and entropy of
mixing show sharp minima in the vicinity of the compound. These properties
indicate the existence of a compound species such as AB in the liquid phase
of the binary system A-B; hence, the system may be considered as consisting
of three species, A, B, and AB in the liquid phase. Other compounds, ABno
may also exist, but the treatment presented here for the equiatomic com-
pound AB would require only minor modifications to extend it to other
intermetallic compounds, AB n • The existence of compound species along
with A and B forms the basis of associated solutions.
The solid phase, AB(s), the simplest compound to be considered here,
is in equilibrium with the liquid phase at the liquidus composition in which
the activities of AO) and BO) are denoted by a l and a2, respectively. The
accompanying equilibrium among A(I), B(I), and AB(s) is
where the activity of solid AB(s), denoted by a~, will be taken as unity,
because AB(s) is assumed to be a pure compound. Liquid ABO) is formed
by A and B in the liquid phase at the equiatomic composition, XI = X2 = 0.5,
according to
We emphasize that aJ. a2, and a 3 without superscripts always refer to the
liquid phase. If the compound formation is also very strong in the liquid
phase so that the activity of AB(I) can be taken as unity only in equation
(4.80) at equiatomic composition, then the sum of the two preceding
reactions and their ~Go is
The resulting ~G~ refers to the melting of 1 mole of AB(s) into ABO);
therefore,
A single point on the liquidus yields one value of XIX2, from which a can
be calculated if tlH':n and tlS':n are known from measurements, e.g., from
calorimetry. After the evaluation of a, the activity coefficients can be
calculated from In "1 = aX2 and In "2
= axi. Calculations show that a varies
slowly with temperature along the liquidus of a number of binary systems.
This equation, first derived by Wagner, II was later used by Vieland,12
Thurmond,13 and other investigators. A modified treatment has also been
presented by Jordan.14
The regular associated solution model does not comply with the require-
ments for regularity because the compound AB is so much larger in size
than A and B, and the distinction between the A-B bond within the
compound AB and A-B bond in AB-A between the compound and A are
difficult to reconcile with the equality of bond energies for the same type
of bonds. The term regular in this case originates from Oe expressed as a
function of composition corresponding to the zeroth approximation to the
regular solutions, often modified by various investigators to enhance the
success of representation of experimental data.
A more accurate and elaborate treatment is presented by Hsieh et al. 15
correlating thermodynamic data l5 - 19 with the phase diagram. The results
for tlO and tlH of aHoy formation per gram atom of Sn-Te aHoy are shown
in Fig. 4.9, and the phase diagram, in Fig. 4.10.
The regular associated solution models can be simplified by dividing
the system into two pseudobinary systems because the solid phase is assumed
to be nearly stoichiometric with or without the assumption that the com-
pound AB exists in the liquid phase. If the solid phase is indeed
stoichiometric, then it can be shown that the discontinuity in the liquidus
114 Chapter 4
o :\. . . .(
~
Gibbs energy of mixing y
~
t; Rakotamava et 01.
-5
\ 'il
Enthalpy of mixing
<:> Nakamura et 01.
'il Rakotamavo et 01.
/I ~
{iij
I
O\'i7v;z 'fJ.~
\:\ I
-10 0. ~
-15
c\t; ~ 'I
\\ f
-20 ~~ ,I,
\~~
'il~
-25
\y Figure 4.9. Gibbs energy and enthalpy
offormation ofSn-Te alloys at 1100 K.
- 27. 5 '--..L.-....I--.l---I.---JI...-.L.--"-.-L...--J..-.J (Courtesy of Y. A. Chang. 15 Data of
0.0 0.2 0.4 0.6 0.8 1.0 Nakamura et al. 16 and Rakotomavo et
Sn X Te - Te al. 17 )
1000- -
L
References
Long-Range Order
Chapter 2 dealt with the phase equilibria in alloys involving only the
first-order phase transitions. The Gibbs energy of a given multicomponent
system is a function of its variables of state P, T, n h n 2 , ••• , nc, i.e.,
121 = 6(P, T, n" n2, ... , nC>. This function is continuous for the first-order
phase transitions but its derivative, with respect to one of its variables,
becomes discontinuous upon a first-order transition. The variable of the
greatest importance is the temperature; therefore, we limit our discussion
to the derivatives of 121 with respect to T. A first-order transition is accom-
panied with a discontinuity in the first derivatives of 121; thus,
a(?i a(G/T)
-=-s·
aT '
or
a(1/ T)
= H (5.1)
or
a2 ( (?i/ T) = C (5.2)
a(1/ T)aT P
energy of atoms may not often be sufficient to move the atoms into their
ordered state below ambient temperatures. The clustering may occur in
liquid and solid alloys but the long-range order may occur only in solid
solutions; therefore, this chapter is largely concerned with solid solutions.
a b
:0
I I
I
If)
~------ ',-- ~------
/ /
/ /
/ /
d
c
Figure 5.1. (a) Disordered structure in body-centered cubic crystals; each site is occupied
randomly by either type of atom. (b) Ordered AB type of alloy; corner atoms occupied by A
atoms form cubes interpenetrating cubes formed by B atoms occupying body-centered posi-
tions. (c) ZnS-type tetrahedral ordered structure projected on two-dimensional coordinates;
A = Zn, B = S. (d) Two-dimensional AB 2-type hexagonal close-packed structure. Each A atom
is entirely surrounded by B atoms, and every third atom in any direction is an A atom.
Thus, if all A atoms are distributed randomly, r is zero and the probability
of finding A atoms on a-sites is equal to its mole fraction. The parameter
r is unity for a perfectly ordered alloy and all the A atoms are on the
a-sites, i.e., N~ = N A = L. The recognition of two or more different types
of sites is essential in formulating the properties of ordered solutions. The
sets of a- and b-sites are sometimes called the a- and b-sublattice respectively.
Each site has Z neighbors, Z being the coordination number. The more
complicated cases of unequal numbers of A and B atoms and a- and b-sites
will be discussed later.
The parameter r for the preceding simple case with equation (5.4) and
XA = 0.5 leads to
(5.6)
.
Probabtlty of AA = N"AN~ =
LL ( 1 - r 2)/ 4 (5.7)
Likewise,
The probability of AB pairs is the sum of two probabilities, i.e., the probabil-
ity of A on a-sites times Bon b-sites, i.e., 0.25(1 + r)2, plus the probability
of A on b-sites times B on a-sites, i.e., 0.25(1 - r)2; hence,
(5.12)
L!
D=--------- (5.14)
a [(1 + r)L/2]![(1 - r)L/2]!
The distribution Db for the atoms on the b-sites is identical, i.e., Da = Db;
hence, the overall distribution Dr is
D =DD = [
L! ]2 (5.15)
r a b [(1 + r)L/2]![(1 - r)L/2]!
124 Chapter S
G = H -Thin Dr (5.16)
ZL Lr 2 W
G(r) = 4 (eAA + eBB + 2eAB) + -2-
- LkT[21n 2 - (I + r) In(I + r) - (I - r) In(I - r)]
(5.19)
Likewise, from equations (5.13) and (5.18), or directly from equation (5.19),
we obtain
Lr 2 W
G(r) - G(r = 0) = -2- + LkT[(I + r) In(I + r) + (I - r) InO - r)]
(5.20)
The equilibrium state of the alloy corresponds to the value of r which makes
aG/ar zero:
aG
-ar = LrW + LkT[ln(I + r). - In(I - r)] =0 (5.21)
(iG(r)
ar2
.
--=LW+LkT - - + - - =0'
1+ r 1 - r '
(1 1) for r = 0 and T = Te (5.22)
Long-Range Order 125
Substitution of r ~ 0 yields
(W<O) (5.23)
1+
- r
-= exp (-rw)
- - = exp (2rTe)
-- (5.24)
I- r kT T
-rw)
exp ( - - -1
r= kT == tanh(-rw) == tanh (rTe) (5.25)
-rw) 2kT T
exp (kT +1
where the identity sign defines tanh. The numerical computation of Tel T
from this equation is very simple. For this purpose, it is necessary to select
an arbitrary value of a = rTel T and read the value of tanh a from an
appropriate calculator or obtain it from the ratio immediately after the first
equal sign in equation (5.25). Then, tanh a is the value of r, and this value
substituted in a = rTel T yields the value of T from the experimentally
known value of the critical temperature. The reader can verify that for the
equiatomic Cu-Zn alloy having Te = 742 K, with a = 0.7, tanh 0.7 =
0.6044 = r, and then T = rTel a = 640.66 K.
Equation (5.25) has a solution that is r = 0 for any value of Tel T or
WI k. When T > Te , r = 0 is the only solution because above the critical
temperature, disorder prevails. At T < Te , there is another solution in the
range of 0 < r < 1, and this root corresponds to the minimum in O(r) as
can be shown by using equation (5.22) with (5.23) as follows:
a2 0(r)
- - = -2LkTe + LkT - - +--
ar2
(1
1+ r 1 - r
1) (5.26)
126 Chapter 5
1.0 ----
--Gokcen
'" "
----Zeroth '
.8
,,
approximation ",
.6
,
\
\
"- \
\
\
.4 \
\,
.2 Figure 5.2. Variation of long-range order param-
eter with temperature. Tc is 742 K. Curve for
quasi-chemical approximation is not shown
0 0.2 0.4 0.6 0.8 1.0 because it follows closely the zeroth approxima-
T/Tc tion curve. (From Gokcen. 12 )
Heat Capacity
The heat capacity C (= Cp ) can be obtained by differentiation of
equation (5.13) with respect to temperature:
4 - ' - Experimental
---Gakcen
~
- - - - - Zeroth approximation
-.J
···········Quasi-chemical
~3 approximation
<l
2
\ '<:::::
04 1.1
T fTc
Figure 5.3. Variation of heat capacity with temperature for equiatomic Cu-Zn alloy. Tc is
742 K, and measured heat capacities of pure components were subtracted from that of alloy
to obtain experimental curve. t.C is the same as the excess heat capacity C e as shown on page
139. (From Gokcen. 12 )
where CCr = 0) is the heat capacity of disordered alloy. The heat capacity
C(r = 0) ofa random disordered solution is the same as the heat capacities
of its component elements. * The values of rand T are related by equation
(5.25) and shown in Fig. 5.2; therefore, for each value of r, T must be
computed and then the result must be substituted in equation (5.28) to
obtain de. The result for the previous example, for which T = 640.66 K,
Te=742K, or T/Te = 0.8634, and r=0.6044, is simply dC/R = 1.174.
Above the critical temperature, T> Te , r is zero and dC rapidly approaches
zero because of r2 in the numerator of equation (5.28). The values of dC
calculated from equation (5.28) are plotted in Fig. 5.3. The results will be
discussed in conjunction with those obtained by the refined methods to be
considered later in this chapter.
*t.H in equation (4.5) for a random disordered solution is independent of temperature; hence,
at.H/aT = C(r = 0) - C (component elements) = O.
128 Chapter 5
Here the equation number (S.Sa) is used to indicate its correspondence with
equation (S.5).
The probability of A-A bonds is now N~N~/ L2 = xi(1 - r2), that of
B-B is (x~ - xir2), and that of A-B is x A(1 + r)(x B + xAr) plus XA(1- r) x
(XB - xAr) or a total of 2X AXB+ 2xir2. Multiplication of each of these
probabilities with the total number of all bonds, ZL, yields the number of
each type of bond, and multiplication with the corresponding bond energy
yields the total energy of each type of bond; the sum of all energies is then
(S.lSa)
Long-Range Order 129
(S.21a)
for r = 0 (S.22a)
(W< 0) (S.23a)
The resulting equation for G(r) - G(r = 0) can be derived by using the
foregoing procedure; the result is
G(r) - G(r = 0) = NkTx'ir 2W + NkT[(x'i + xAxBr) In(x'i + xAxBr)
+ 2XAXB(1 - r) In(xAxB - xAxBr) + XB(XB + xAr)
(5.20b)
An ordered phase such as AuCu3 has XA = 0.25 and XB = 0.75 and the
equilibrium values of r are given by
(PG = NkTc
ar2 8
(8 +~)
kTc
= O.
'
for r = 0 (5.22b)
from which Tc = - W j(8k). We give these relations for the sake of complete-
ness without making use of them.
First Approximation
*R is used here and in writing the equations for permutations (Da , Db, and D,); avoid
confusion with the gas constant; see also pages 143 and 145.
132 Chapter 5
D = R! Q! (5.31)
a (R - X)!X! (Q - X)!X!
where Da is the distribution for the a-sites. The distribution, Db, for the
b-sites from [(iii), (iv)] and [(v), (vi)] is identical with Da; consequently,
the overall distribution Dr is
-DD - [
R!Q! ]2 (5.32)
Dr - a b- (R-X)!(X!)2(Q-X)!
D = _ _ _-'-(2_R....:.)-'!('-2~Q.;_)!_ __ (5.34)
r (2R - 2X)![(2X)!f(2Q - 2X)!
Long-Range Order 133
D = _ _---=[_L-'-(1_+---'r)-=..]-=..![L_(-'-:,-I_-_r--,-)=-]!_ __
(5.35)
r (L + Lr - 2X) ![(2X) !f(L - Lr - 2X)!
D = _ _--.:.(2_L--'.)_!_ _ (5.15a)
r (L+Lr)!)(L-Lr)!
This equation states that R = O.5(L + Lr) and Q = O.5(L - Lr) can be
permutated for any value of R, Q, and r, irrespective of the value of X.
When r = I, then R = L; hence, both X and Q must be zero since there
are no A-A bonds, and Dr = I as expected from equation (5.35). Conversely,
if X = 0 or if there are no A-A bonds, R has to be equal to L by geometrical
requirements because when R is less than L, it can be easily shown by one-
and two-dimensional crystal constructions that X cannot remain zero if the
permutation given by equation (5.15a) were carried out for other values of
R than R = L. The GBW method assumes that Rand Q may have values
unrestricted by the values of X so that equation (5.15a) is valid for any
value of X, and this assumption is analogous to the random distribution
permitted in the zeroth approximation to the long-range disordered regular
solutions.
All A atoms have ZNA/2 bonds (or ZNA half-bonds) belonging to A
atoms. If for convenience we define Y such that ZY /2 are the A-B bonds
emanating from A atoms, then the remaining bonds are the A-A bonds;
hence, (A-A bonds) + ZY /2 = ZNA/2, but A-A bonds are equal to
(2X)Z/2 from (i) and (iii); consequently,
2X+ Y= NA = L (5.36)
For the disordered alloy, r is zero but the solution is not necessarily random,
and substitution of r = 0 in equation (5.35) gives
LlL!
D=------:: (5.37)
r [(L - Y)!Y!f
which is exactly equation (4.23) for equiatomic solutions for the first
approximation to the regular solutions.
134 Chapter 5
Equation (5.35) is not yet useful for our purposes when the solution
is highly ordered or Y is considerably high as was shown in equation (4.29).
The correction suggested there may be simplified as FL == [F(r)t in the
numerator of equation (5.35):
[L(t + r)]![L(l - r)]!F L
Dr = Dr(corrected) = (L + Lr - 2X)![(2X)!f(L - Lr - 2X)!
(5.38)
The requirement for F(r) is that F(r) = 1 for both r = 0 and r = 1, or
In F = 0 in both cases. The exponent Lover F = F(r) is for convenience
in deriving the succeeding equations from (5.38). The form of In F obeying
these requirements, and with one adjustable parameter,II.12 g, is
In F = g(l + r) In(l + r) + g(l - r) In(l - r) (5.39)
It will be seen later that without g, it is impossible to make the exchange
energy, W, the same in the ordered solution and the disordered solution of
the same alloy. Further, g is sufficiently small to make F a correction
equivalent to c in equation (4.29).
The energy E of the alloy is Ze hl /2 times each term in (i)-(viii) in
equation (5.30) with the subscripts hi representing AA and AB, or BB, and
ehh the corresponding bond energy; the result is
2X t ± J n + r2(l - n)
L 1- n
2X
(5.43 )
L
(5.44)
[ G(r, rl)]
T T-'~O
=-
T r'~o
(H)
( - I=0
-SrT
'
) (5.46)
(5.47)
For the right side of equation (5.45), the integration procedure is similar
to that used for the long-range disordered regular solutions with r = 0 as
in Chapter 4. The integration of the right side of equation (5.45) requires
expressing d(T- I ) by using b of equation (5.44) as follows:
d
W)
(- 2kT =
bdb
b 2 _ r2 (5.49)
Next, equations (5.40), (5.43), and (5.49) are substituted in equation (5.45)
to obtain
f T-'
T-'=O
H dT- I = f
T-'
T-'=O
[ZLe AB dr l +
(1
2Lk(1 - r2)bdb
+ b)(b + r)(b - r)
] (5.50)
The left side of equation (5.45) is equal to [G(r, T)/ T] + S(r, r l = 0) from
equation (5.46). The first term on the right side of equation (5.50) is ZLeAB/ T,
and the last term can be integrated by parts to derive
G(~ T) + S(r, rl = 0)
=
ZeAB b+r b- r +
Lk [ - - + (1 + r) In--+ (1- r) In-- - 2In--
1 bJ (5.51)
kT 1+ r 1- r 2
Substitution of equation (5.48) for the entropy on the left side and
rearrangement of the result gives
G(r, T) ZeAB
LkT = kT - In F + (1 + r) In( b + r)
_l_aG=_alnF+ln(b+r) =0 (5.53)
LkT ar ar b- r
where all the remaining terms including those containing ab/ ar cancel out.
For r = 0, this equation requires that aF/ar be zero since F(r = 0) = 1.
Differentiation of equation (5.53) and thereafter substitution of r = 0 at the
critical temperature gives
Long-Range Order 137
For r = 0, equation (5.44) gives be = nO. s = e-W/2kTc, and the use of equation
(5.53) and the rearrangement of equation (5.54) yields
d 1n F
2
(W)
0.5 --;J;2 = exp 2kTe ; (5.55)
G(ideal) Z
LkT = (eAA + eBB) 2kT -21n 2; (5.57)
Ge ( 2e W/2kT )
-k-= 21n 2-2In(1 + b) =21n W/2kT (5.58)
L T I+e
This equation, where 2Lk is the gas constant, is identical 13 with equation
(4.55).
The correction term, In F, in equation (5.39) may now be substituted
in equation (5.55) to obtain the value of g for any system. We take, as an
example, the Cu-Zn system in the vicinity of equiatomic composition,
known as beta brass. The value of W / k is the only parameter permitted to
be determined from experimental data in any approximation to the regular
disordered solutions. The compilation of Hultgren et al. 14 extrapolated a
short distance to the equiatomic composition yields G e /2LkT = -1.202 at
773 K, which substituted in equation (5.58), gives W / k = -2678. The regu-
lar solution model assumes that W / k is independent of temperature. The
critical temperature Te is 742 K. The explicit functional form of F(r) is not
138 Chapter S
known and its determination would require extensive and laborious enumer-
ation of configurations for one- and two-dimensional crystals. I I Such a task
is expected to be very difficult and time-consuming. However, the function
F(r) suggested in equation (5.39) is a very small correction meeting the
conditions that it vanish for r = 1 and r = 0, and further, it is a much smaller
factor in Dr than any factorials that yield terms of the order of In M! =
M In M for any factorial term M!. We shall soon see that g is also a number
of considerably smaller than unity.
The successive derivatives of In Fare
aIn F 1+ r
--=gln--' (5.59)
ar 1 - r'
g = exp (~)
2kTc
= exp(-1.8046) = 0.16454 (5.60)
l+r b+r
0.164541n-- = I n - - (5.61)
l-r b-r
A numerical example for this equation is useful. Let r = 0.80; the left side
of this equation is then 0.36153, so that
b + 0.8
1.4355 = - - (5.62)
b - 0.8
LW(1-r2)
E = H = LZeAB - ----'-----'- (5.63)
l+b
The total differential of H = H(r, T) is
(5.68)
(5.71)
Substitution of equations (5.68), (5.69), and (5.71) in (5.65) gives the final
equation for C e / R:
(5.73)
This value, and others obtained similarly, are plotted in Fig. 5.3. It is evident
that the results are in fair agreement with the closely concordant and
independent experimental values of Moser,16 and Sykes and Wilkinson 17
plotted as C e / R = [C(alloy) - C(component elements)]/ R. The values
calculated from the quasi-chemical method are also plotted in Fig. 5.3,
indicating that these results are not significantly different from the zeroth
approximation of GBW.
We now justify the functional form of F(r) by the way of summary of
the foregoing procedure. Equation (5.39) for F = F(r) must satisfy (I) the
boundary conditions that In F be zero for r = 0 and r = 1, (II) equation
(5.53) for all the equilibrium values of r, and (III) equation (5.54) for r = 0
and T = Te. For example, In F = g(1- r) In(1 + r) would satisfy all the
requirements but not (II) for r = o. It may be stated that equation (5.39)
is not unique, but it is a useful and relatively small correction meeting the
foregoing requirements.
Long-Range Order 141
D = _ _ _.....::['-L....:...(1_+----'r)-=.]-=.![_L-'-(1_-----=-=-r)-=-]!_ __ (5.74)
r (Y+Lr+c)![(L- Y-c)!f(Y-Lr+c)!
1 1+ 1 1= 21 1
A--A B--B B--A
and accounting for all the possible lower hierarchy of configurations down
to single bonds. While with increasingly complex clusters the method would
minimize the error originating from the permutation of single bonds and
Long-Range Order 143
(5.5c)
(5.5d)
These atoms are shown and summarized in Fig. S.S for NA = 8, NB = 12,
XI = 0.4, and L = 10, for sufficiently large numbers of atoms to avoid
objectionable fractional values for items other than r, X A , and XB in equations
(S.Sc), (S.Sd), and (S.30a). The determination of X is given in the figure
caption, and that of X' will now be presented as an additional example.
Atoms of B on b-sites have 8 bonds with B atoms, and the latter are on the
a-sites. Eight bonds are shared by B on b-sites, and B on a-sites; hence,
8/2 = 4 = X'. The remaining parts of equations (S.30a) are determined by
the same procedure, and the results are listed in Fig. S.S.
The a-sites contain R atoms of A, and Q' atoms of B, and the distribu-
tion of these atoms on the a-sites is similar to equation (S.31), i.e.,
R!Q'
D = ------"------ (S.31a)
a (R - X)!X!(Q' - X')!X'!
0: A atoms ( i) X=2
.6,.: 8 atoms ( ii) R-X=2
Likewise, Db is given by
R'!Q!
D - ------"------- (5.31b)
b - (R' - X')!X'!(Q - X)!X!
(5.5f)
These atoms are shown and summarized in Fig. 5.6 for N A = 6 = La,
NB = 18 = L b , XA = 0.25, and r = 5/9. Again, these numbers are selected
to avoid fractional values for all the items in equations (5.5e), (5.5f), and
(5.30b). The enumeration R - X is illustrated in the caption for Fig. 5.6;
and this value also yields X because R is given by equation (5.5e).
146 Chapter 5
• : a-sites (iii) X= I
x : b-sites (iv) Q -X= I
NA = 6 = XAN =N/4 =La (v) X' =12
Figure 5.6. Net numbers of A and B atoms on a- and b-sites for NA = 6 = La' NB = 18 = L b ,
and r = 5/9. Six numbered bonds involve A atoms on a-sites with B neighbors; therefore,
R - X = 3 for part (ii). Remaining parts in equation (5.30b) are obtained similarly and
summarized above. All the numbers are selected to avoid fractional atoms on a- and b-sites
and fractional neighbors.
R!Q'
D=-----"'----- (5.31c)
a (R - X)!X!(Q' - X")!X"!
R'!Q!
D - ------"------ (5.31d)
b - (R' - X')!X'!(Q - X)!X!
are Ag2AI, AIFe 3, Au 3Cu, AUCU3, CoPt 3, CU3Pd, FePd 3, Fe 3Pt, FePt 3, Ni4Mo,
and Ni2Mo, to name a few. The preceding equations for NA < N B, La = Lb =
L, and for (NA = La) < (NB = L b ) have not yet been applied to such alloy
systems.
References
Interstitial Solutions
Introduction
n
y=-=_.
r
x=---=
n y
.
xle
y=-- (6.1 )
eN e' n+N y+l/e' I-x
.. fy
a =yx = - - (6.2)
1-y
where y and f are the activity coefficients for their respective concentra-
tions, and the activity can be measured in terms of a gas phase potential
or pressure in equilibrium with the solution, such as H2(g) for dissolved
hydrogen [H], CH 4 /H 2, or CO/C0 2 gas mixtures for dissolved carbon [C]
as will be discussed later. We emphasize here that the symbols without
subscripts refer to the interstitial element unless they are qualified by immedi-
ately succeeding words in parentheses.
Diatomic gas molecules dissociate to dissolve as monatomic solutes in
metals. Small solubilities of a diatomic gas such as H 2(g) in metals of Group
VIA (e.g., Cr, Mo) and those to the right side in the periodic chart on
page 318 can be assumed to take place in two steps: (1) by dissociation of
H2(g) into monatomic gas, H(g) (or as superficially adsorbed H), requiring
a large positive value for tlH~ of dissociation, and (2) by dissolution of
H(g) in a metal, generally requiring a relatively small and usually negative
value for tlHll of solution. The value of tlH for the overall process is
tlH~ + tlHll , and this sum is a positive quantity because tlH~ is generally
much greater than ItlHlll; consequently, the solubility for a given pressure
increases with increasing temperature in accord with Le Chatelier's prin-
ciple. The elements in Groups II A-VA, as well as Pd, La, and Ta, dissolve
up to 105 times larger amounts of hydrogen than the remaining poor metal
solvents. For most of these good solvents for hydrogen, again H2(g) dissolves
as monatomic H, but tlH = tlH~ + tlHll is negative so that the solubilities
decrease with increasing temperature. [An italic H refers to enthalpy; a
roman H refers to dissolved hydrogen. Monatomic gaseous hydrogen is
always followed by (g), i.e., H(g).] Palladium dissolves large amounts of
hydrogen, and at certain pressures and temperatures up to 292°C, two phases
coexist with gaseous hydrogen. The phase low in hydrogen is the a-phase,
and that high in hydrogen, the ,8-phase. Above 292°C, known as the critical
temperature, Pd and H form a single solid solution. Hydrogen in palladium
is released without much difficulty even below ambient temperatures when
the pressure is decreased. For this and many other interesting reasons, and
for many technical applications, the solubilities of hydrogen and its isotopes
in palladium have been thoroughly investigated. 1- 5 It is therefore appropri-
ate to devote the next section to the thermodynamics of the Pd-H system
with more details than usual for other systems considered in this book.
Interstitial Solutions 151
_l
,""-1
I
100
I I ~ p
~;1.-
~.'-II
00
+' /
....
198·[
'.~OQ-~-I ~4---'~/ , I/ t 1
I;~~J"'~'~'-'-"'-"'~-i-J~q,'-i
P-'" •~ 188 .[
- _ __- / I
'-g
10 10
~
~ --..=t-.-.-r---.-.-.~~ -/
2oQO{Il-f--o-o-i--o-
143·[ O~o- - -....I
I / I
' t
C 0 100 ·C " / I
I{.-" l ii I /
,--
I 0--0
J
--_-0=0-0.-0
160·[
0--0-0-
¥ -..I
A-w- + -
I' / ! I
if \
110·[
0.1
i /
1---
I "";(.)(-::=\: )(->(-)(-
1 - - ' - )(.~)(!.:A. r-/+
~_IC"': 70·[ \ +
I \ !
, I
+-+ +-+-1
0.01 " ;{
1-;~ -+-1 10·[
1
o 01 02 03 O~ 05 06 07
Figure 6.1. Dissolved hydrogen as r = H/Pd atomic ratio versus gaseous hydrogen (H 2 )
pressure P at various fixed temperatures. For phase boundaries, see Table 6.1. (From Frieske
and Wicke4 with permission.)
Pd-H System
Hydrogen dissolves in palladium as shown in Fig. 6.1, where dissolved
hydrogen as r = HI Pd atomic ratio is plotted 5 versus gaseous diatomic
hydrogen pressure P in atmospheres. The a-phase on the left coexists with
the ,B-phase on the right at each temperature below 292°C and at each
pressure, which is known as the plateau pressure. Neutron diffraction and
other studies indicate that hydrogen occupies the octahedral sites in both
a- and ,B-phases. The number of octahedral sites is the same as the number
of Pd atoms; therefore, N is also the number of octahedral sites. Hydrogen
must also occupy other sites when nl N is greater than 1 at very high
pressures.
The a-phase has a face-centered cubic structure, slightly expanded by
dissolved hydrogen, and the ,B-phase has a highly distorted face-centered
cubic structure. The increase in volume of Pd upon transition from a to ,B
is about 1.6 cm 3I g-atom H. (The volume of pure Pd is close to 8.87 cm 3 / g-
atom.)
The most extensive recent study of the Pd-H system has been carried
out by Frieske 4 •5 * in the range of 20 to 300°C at hydrogen pressures from
0.01 to 140 atm. The solubilities and the phase boundaries were determined
*The results quoted in References 4 and 5 are from H. Frieske, Dissertation, University of
Munster (1972).
152 Chapter 6
~ D absorption
400 ~ • desorption X lloo
.~ 0 absorption ,:
~~ p ~
300 10
200
X· 1(f ~
cm3/mol ~ atm
100 1-;)oF----+---+-------1~-_+_ -+----+----1 0.1
'liD
of-----+----+------+----+----+-l'-~-=J~[T-flIIPl!!I 0.01
o 01 02 03 04 05 06 07
Figure 6.2. Magnetic susceptibility X and diatomic hydrogen pressure P, versus r = H/Pd
atomic ratio at 120°C. (From Frieske and Wicke' with permission.)
Interstitial Solutions IS3
The activities of both Pd and PdHo.5 in reaction (6.3) decrease with increasing
temperature because both phases dissolve hydrogen; hence, P is nearly
equal to the equilibrium constant ofreaction (6.3); therefore, equation (6.4)
is simply an empirical equation without the activity corrections for PdHo.s
and Pd. Other equations for different ranges of temperature are available;
e.g., tlHo = 9325 cal/mole H 2, and tlSo = 21.8 cal/K-mole H2 have been
reported for -80 to +50°C by Wicke and Nernst. 7 (For lower temperatures,
see, e.g., Lynch and Flanagan. 8 ) The values of tlHO agree well with the
tlHo = 9440 cal/4 moles PdH o.5 based on the calorimetric data of Nace
and Aston. 9 Equilibria between a-phase and H 2 , or D2 have been investi-
gated in the range of approximately 0 to 100°C by Wicke and Nernse whose
results agree very well with those of Clew ley et al. lo :
H2 23,970-21,500 r 25.6
D2 22,820-21,500 r 25.4
T2 22,050-21,500 r 24.6
832
In Kp(T) = T - 6.25 (6.6)
aH(r) r as
InP= - - - + 2 I n - - + - (6.7)
RT 1- r R
The results are listed in Table 6.2. for aH(r) and as.
The gas-phase equilibria, e.g., H2 + D2 = 2HD, can be combined with
these equations for use in isotope separation. The critical temperature for
the Pd-D system is 276°C (292°C for H 2), and the critical pressure is 35 atm
(19.7 atm for H 2).
The Pd-H and related equilibria and other properties have been dis-
cussed elsewhere in full detail. l-8
start the alloy formation process with pure monatomic gaseous hydrogen
H(g) at 1 atm, in the same way as we start with pure metals at their standard
states in forming all metallic alloys. The basis of treatment is n atoms of
Hand N atoms of metals providing eN interstitial sites. Let ZII/2 and Z12
be the respective numbers of H-H and H-M bonds per atom of dissolved
H; ell and e12, the corresponding energies per bond; and eJ. the energy of
gaseous atomic hydrogen. The probability of finding a dissolved H atom
at a selected site is n/ eN and the probability of finding another H atom
next to the first atom is n 2/(eN)2. The total maximum number of bonds,
if all the interstices were occupied, is ZII cN /2, and multiplication with
n 2/ (eN)2 yields the actual number of existing H-H bonds, i.e., Zlln 2/(2eN).
The number of H-H bonds times ell, the H-H bond energy, yields
Zlle ll n 2/2cN, which is the energy due to the H-H bonds. The number of
H-M bonds per atom of H is Z12, and the total number of H-M bonds is
Z12n, which, multiplied by the bond energy e 12 , yields Z12e12n. The energy
of the metal is Z22Ne22/2 = Nh'2, and the enthalpy of the alloy is
D = _(,-c_N--'..)_!_
(6.9)
(eN - n)n!
+ N(h~ - Ts~)
Here, h~ - Ts~ = O~I N is the standard Gibbs energy per atom of metal.
The last set of terms represent - kT In D, expanded by using the Stirling
approximation for factorials.
Equation (6.11) can be simplified by using the following notation:
n2 e
0= -A-2-+ nO + NO~
eN
- RT[eN In eN - (eN - n) In(eN - n) - n In n] (6.13)
-
0= [a~
-
an P,T,N
=O
e
-2A
--+n R T l n -n- -
e eN eN-n
(6.14)
Substitution of r = nl N gives
- • 2A r
0=0 --2 r+ RTln-- (6.15)
e e-r
( 6.17)
° 2A r
0.5RT In P = 0 - 0.50 0 (H2) - -2 r + RT I n - - ( 6.18)
c c-r
(6.19)
° ° 2A r
O.5RTlnP = flH -TflS --2r+RTln-- (6.20)
c c-r
Recall that the symbols without subscripts refer to the solute, and Prefers
to H2(g). Substitution of Kp and cy = r in equation (6.18) and elimination
of d - 0.5GO(H2) yields
• 2A y
6 - G = RTln a = - - y + RTln-- (6.22)
c l-y
-af'f = G- 2 = + Ay + cRTln(l -
2
G~ y) (6.23)
aN
• • 2A y
6 + RTln a = G --y + RTln-- (6.24)
c l-y
Interstitial Solutions 159
On the Henrian scale for the activity, as y approaches zero, i.e., y ~ 0, then
f ~ 1, y/(1 - y) ~ y, and equation (6.24) becomes 6 = d + RT In y.
Comparison with the Henrian definition of activity 6 = GO + RT In y at
y ~ 0 and a ~ y shows that GO in equation (6.24) is the standard Gibbs
energy with the reference state asf ~ 1 at y ~ o. Equation (6.24) is therefore
identical with equation (6.22).
The activity on the Henrian scale in equation (6.24) is given by
6 - G" = RT In a for any concentration; however, the activity g on the
Raoultian scale, in which the pure solid is the standard state for the
interstitial, is obtained by subtracting GO from both sides of equation
(6.24):
- ° 2A
G - GO = RT In g(Raoult) = G - GO - - y + RTin -y- (6.25)
c 1-y
The equation for GO - GO, as well as for A, must then be obtained by fitting
the activity data with equation (6.25) at various temperatures and concentra-
tions. It is quite frequently possible to express GO - GO as a linear function
of temperature, i.e.,
(6.26)
- 2A
G e == RTlnf= - - y (6.27)
c
(6.28)
This equation can also be derived by using equation (6.27) in the Gibbs-
Duhem relation, dO~ = -[x/(l- x)] doe = 2Ay dy and integrating it from
y = 0 to y, observing that O~ is zero at y = 0 since a2 ~ 1 with y ~ o.
The molar excess Gibbs energy G e = xO e + (1 - x) O~ and the molar
Gibbs energy G = xO + (1 - x) O 2 can be easily obtained by substituting
the partial properties in these equations but we shall have no specific use
for such equations.
Alternative derivations of equations (6.24) and (6.27), e.g., that given
by Hillert and Staffansson/ 5 are also based on the zeroth approximation
to the regular interstitial solutions. In a number of applications, it is assumed
that A is a linear function of temperature, i.e., the bond energy ell is no
longer constant but varies linearly with temperature. This is a useful
empirical concept, as will be shown later in this chapter.
G- - 0.5GO(H 2 ) 2Ar I r
0.5 In P = -- + n-- (6.30)
RT RT 1- r
The miscibility gap between a and f3 phases disappears above the critical
temperature, Te , which is 565 ± 2 K in Fig. 6.1, where a horizontal inflection
appears in the P versus r curve. The first and second derivatives of equation
(6.30) with respect to r are zero at the critical point:
a In P 1 1 2A
0.5--=-+----=0; (6.31)
ar r 1 - r RT
a2 ln P 2r - 1
0.5 - - 2 - = ( = 0; (re = 0.5) (6.32)
ar 2)2
r - r
The first of these equations determines A in terms of Te and r = re, and the
second, the critical value of re , which is 0.5. Unfortunately, the experimental
value of re is 0.25, far below re = 0.5 from equation (6.32). An artifice
defended by a number of investigators is that there are two sets of sites
available for dissolved hydrogen: the first set is 0.6N of the total available
set of N, and the second set is O.4N. This artifice has been generally rejected l6
because all the N sites are equivalent. As evidence for the existence of two
Interstitial Solutions 161
such sites, a distinct change in the magnetic behavior of the Pd-H system
is often cited, i.e., Pd-H alloys change sharply from paramagnetic state to
diamagnetic state when r exceeds 0.6 or after 0.6N sites are occupied by
dissolved H. Substitution of 0.6N for N in the foregoing equations would
simply replace r by () = r/0.6. Equation (6.32) with () would then yield
(}c = 0.5 and rc = 0.3, which is not too far from the experimental value in
Table 6.1. The sites in excess ofO.6N, i.e., O.4N sites, are said to be available
for H at high pressures and low temperatures, and for this range it is
proposed that an empirical equation linear in r is obeyed:
In P = A + Br (6.33)
2A 1 6NoE'r
--+--+ =0 ( 6.35)
RT r - r2 RT
(6.36)
The values of rc = 0.25 and Tc = 565 are now required for determination
of A and 6NoE'/ RT; substitution of these values in the preceding equations
yields
where the first two terms on the right represent the solubility of H2 in the
a-phase for very small concentrations of r when In[r/O - r)) = In r, and
when the last two terms become negligible so that
d - 0.5GO(H2) r 1140
---R-T--'--''''- == In -po-.s == -T- - 6.35 (6.39)
The first approximation to the regular solutions in its correct form has
not yet been applied to the interstitial solutions. If elI in equation (6.11) is
taken to be zero, then the random distribution given by equation (6.9) would
be valid, and the zeroth and first approximations would be identical.
Hydrogen Storage
100
350 + + + + + + a' + ·c
K
a 50
300 + 6 +
250 + + + + +
a. p -50
,200 + + + + I
t
T -100
150 + + + +
, .y
-150
100 + + + + Y
.
-200
....... 7,-
50 + + + + '
"
C+L 1L:
"
-250
""
"
01 02 03 04 05 0.6 07 08
r- H/Ta
Figure 6.3. Tantalum-hydrogen system; r = H/Ta atomic ratio. (From Kobler and Welter20
with permission.)
(6.40)
(6.41)
Interstitial Solutions 165
t (ee)
250 150 60
F,Ti-F, Ti H
-
E
co
0..
0.1
0.01
0.001 ~~--~----~----~--~~--~~--~----~
1.8 2.2 2.6 3.0
IOOO/T
Figure 6.4. Dissociation pressures of metal-hydrogen systems. (From Wiswall 24 with per-
mission.)
In the first case, A and B form such a stable phase that hydrogen is displaced
as a gas, and in the second case, hydrogen and B share strong bonding with
A. In the latter case, hydrogen from ABnH, is released as follows:
(6.42)
166 Chapter 6
N
X
E
'0 10
W
a:
::>
(f)
(f)
w
a:
a..
z
0
I-
<t
U
0
(f)
(f)
0.1 '----'--:-'":::---'-"'---l---L.---l_-'------L----l._L-...L--L---l
0.2 0.4 0.6 0.8 1.0 1.2
ATOM RATIO H/(Fe + Til
Figure 6.S. Desorption pressure in FeTi-H system. (From Wiswa1l 24 with permission.)
( 6.43)
If, however, both A and B are stable hydride formers, more complex
reactions may often occur.
Interstitial Solutions 167
The alkali elements and their alloys form stable hydrides, unsuitable
for hydrogen storage at temperatures below 300°C. Alkaline earth elements
(Ca, Br, Sr) also form stable hydrides, but an alloy phase of calcium, CaNis,
is of possible interest because it stores hydrogen up to CaNisH6 and
dissociates at room temperature in the vicinity of 1 to 15 atm. Other calcium
alloys also have possible uses because they are similar to CaNis.
Hydrides of magnesium and its alloys are more promising media for
storage. Magnesium dihydride, MgH2' contains 7.65% by weight hydrogen
and dissociates at 287°C by adsorbing 17.8 kcalj mole H 2. The waste heat
from a combustion chamber or from a hot exhaust given off by a motor
can be used to supply the necessary heat for dissociation. The hydride is
re-formed at higher temperatures and pressures. The intermetallic phase,
26
24
I
• Pr"S5Y'"
[aIm/
22
20
18
16
12
t= 40'C
t = 20'C
2 3 4 s 6 7
Hydrogen Concentration
(at H/mol LaNiSI
Figure 6.6. Absorption-desorption isotherms for LaNis at various temperatures. (From
Kuijpers and van Mal 31 with permission.)
168 Chapter 6
Group IVA elements, Ti, Zr, and Hf, form binary hydrogen metal alloys
of considerable stability; therefore, they are not suitable for hydrogen
storage. Some metal binary alloys of the type TiFe have potential applica-
tions because they absorb H2 up to TiFeH2 and release it as shown in Fig.
6.5. The pressure hysteresis is high and the concentration of H is only 1.91 %
by weight. Powdered TiFe is not pyrophoric and not very expensive, and
for stationary storage it is possibly the best storage medium according to
Wiswal1. 24
Interstitial Solutions 169
Practical Applications
The processes requiring hydrogen storage and use are in their infancy.
Nevertheless, possible uses have already either been proposed or made in
a few significant cases. Hydrogen concentration cells are the simplest power
generators that take advantage of a pressure difference between anode and
cathode. It is simply a concentration cell that generates power by consuming
hydrogen at a high pressure and releasing it at a lower pressure. Such cells
are already in use in space vehicles. The use of metal hydrides for providing
steady hydrogen pressures for the anode and cathode and thus storing or
generating considerable amounts of power in large- and small-size batteries
has been proposed.
Production and storage of hydrogen and its reuse in fuel cells may
provide a method for electric utility power leveling. A small experimental
model was developed by using TiFe for hydrogen storage, and an enlarged
research and development program ensued. The results appear to be
promising.
Experimental automotive propulsion has been achieved by using
TiFeH n • About 7.7 kg of useable H 2, stored in 1016 kg of TiFeH m has been
demonstrated to be capable of providing a range of 121 km, at a sustained
speed of about 80 km/hr for a bus weighing about 6.8 metric tons. The hot
exhaust gases were used to dissociate the hydride. The spent hydride was
recharged to 80% of its capacity in 15 min and full capacity in 1 hr. 24 This
performance is far superior to that obtained by acid battery automotive
propulsion.
Magnesium hydride, MgH2' has more than three times the available
hydrogen from TiFeHl.9 on the basis of hydrogen per unit weight of alloy.
It would permit longer range for automobiles with a lighter storage medium.
Its dissociation enthalpy and temperature are rather high, demanding special
engines with high exhaust temperatures for this purpose.
Ultrapurification of hydrogen, fractionation of H, 0, and T, can readily
be achieved by using metal diaphragms from hydride-forming metals. Metal
hydrides can be used for gas compression for power generation in turbines,
for heat storage, and even for air conditioning. Other uses are discussed by
Wiswall/ 4 Reilly/5 and Hoffman et al. 26
Fe-C System
The Fe-C system has been investigated by more than 200 researchers
because of the vast industrial importance of iron alloys. We cite a few
selected recent papers, summaries, and compilations from which the reader
may find all the remaining publications.
The phase diagram as computed by Ohtani et al. 33 from their thermody-
namic equations based on selected data for the activity of carbon and phase
boundary compositions is shown in Fig. 6.7. This diagram contains small
Interstitial Solutions 171
Mol Fraction
1800
L + Graphite
1700
1400
Fe 3 C
u
1300 1 1600
::.::
0
w-
1500
1200 w
0:: 0::
::> ::>
J-- Y 1400 ~
<{ 1100 0::
0::
W w
a... Y+Graphite or Fe 3C 1300 ~
~ 1000 w
w
J-- J--
1200
900 p 0.0223
p' 0.0206
S 0.764
s' 0.671 1100
800
740 0 C
700 Ifp~~s'=_~-=-=-=-=-=-=-=-~-=-=-=-=72::-7-=O::c-=-=-=-=--=-,=--:::I 1000
a + Graphite or Fe 3 C
I
I 900
600~--~--~~~~~~--~--~~
Fe 1.0 2.0 3.0 4.0 5.0 6.0 7.0
CARBON CONTENT, wt pct
Figure 6.7. Calculated phase diagram for Fe-C(gr) and Fe-Fe 3 C systems. (Adapted from
Ohtani el al. 33 with permission.) Broken lines are for Fe-Fe 3 C system.
improvements over the previous diagrams,35-41 but represents the best, and
in some respects the most recent results obtained by thermodynamic and
metallurgical investigations. The evaluation of the data and the resulting
summary in the form of thermodynamic relationships require a great deal
of experimental expertise and judgment. The more recent evaluations by
Nishizawa and co-workers/ 3.34 Agren,35 Harvig/ 6 and Chipman 37 are
required reading for the interpretation, evaluation, and selection of ther-
modynamic data and the construction of phase boundaries. We assume that
the necessary thermodynamic equations are available as given by Ohtani
et a/. 33 and proceed to compute the phase diagram from these equations.
We shall hrst discuss the Fe-C(gr) [iron-graphite] system in detail and then
outline the Fe-Fe3C [iron-cementite] system.
172 Chapter 6
a-Phase
The a-phase of pure iron is the body-centered cubic phase that exists
from ambient temperatures up to 1184 K. Its Curie temperature is 1043 K.
It dissolves very small amounts of carbon in equilibrium with graphite.
Therefore, with A == A", g(Raoult) = 1, and yl(l - y) """ y, equation (6.25)
can be rearranged to obtain grG O - "G" = RTlnSy" = -ll.H" + Tll.S", where
the last equality is from equation (6.26), and s on Sy" is for graphite
saturation. The recent results of Hasebe et a1. 34 , as reinterpreted here, are
• sxO'
grGo_"G """ RTlnSy" """ RTln-= -99,750+ 33.6T
3
where ll.H" = 99,750, and ll.S" = 33.6, and all the energy units are in joules
in the remaining sections of this chapter. The last set of three terms in
equation (6.44) is a purely empirical correction representing the effect of
paramagnetism-ferromagnetism on the solubility of graphite in the range
of 800 to 1200 K. This correction is derived from H~(ferromag.)
H 2(paramag.) == ll.H 2(f.p.) determined by calorimetric measurements and
evaluated and summarized. 42 ,43 Nishizawa and co-workers 33 ,34 have pro-
posed that the contribution to equation (6.44) is ll.H 2(f.p.) x (-500/1043)
where -500 is the rate of change of Curie temperature in kelvins per atomic
fraction of carbon, and 1043 is the Curie temperature of pure iron. The
solubility of carbon in a-Fe is very small; therefore, -500 cannot be
determined accurately; hence, it must be regarded as a reasonable correla-
tion factor that relates ll.H 2(f.p.)/1043 to the Gibbs energy contribution in
equation (6.44). Other types of interpretation of magnetic effects on the
phase diagrams are discussed by Chuang et a1. 44 ,45 Equation (6.44) is based
on the experimental results from 823 K (0.0010 wt% C) to near the eutectoid
temperature of 1013 K (0.0206 wt% C) and the extended solubility calcula-
tions based on the activity of carbon in the 'Y-phase as will be seen later.
The a-Fe and graphite phase boundary is satisfactorily represented by
equation (6.44) from about 800 to 1013 K. At the eutectoid temperature of
1013 K, equation (6.44) yields Sy" = 0.000319 (0.0206 wt% C).
The a-phase (= ferrite) is in equilibrium with the 'Y-phase (= austenite)
from 1013 to 1184 K. The a I 'Y phase boundary in this temperature range
requires the activity of C(gr) in the 'Y-phase, to be discussed in the next
Interstitial Solutions 173
ad = ad + RTlnya ( 6.45)
where ya is so small that the Henrian activity coefficient is unity, and further,
the linear term in ya in equation (6.25) is zero, i.e., A a = O. The superscript
a may be replaced with 8 to obtain the corresponding equation for the
8-phase which is also bcc, and stable from 1665 to 1809 K. The superscript
a in ad is not essential; it is used here for emphasis.
y-Phase
The 'Y-phase is face-centered cubic in structure, constituting the most
important area of the iron-carbon diagram. The activity of carbon, 9 'Y, has
been determined 46 - 49 by equilibration with gaseous mixtures of either carbon
monoxide plus carbon dioxide or hydrogen plus methane:
P4
2H2 + [C in Fe]
II
= CH 4 , K =-- (6.47)
P g. p2
The values of the activity are obtained from equation (6.46) or (6.47) by
using the experimental values of pU P2 or P4 / p2 and K~ or K~I calculated
from equation (6.48) or (6.49). The activity data are fitted with equation
(6.25) to obtain eGo - gr GO) and A 'Y, both of which are usually expressed
174 Chapter 6
yY
= 45,360 - 18.4T + (57,400 + l1.2T)yY + R T l n - -y ; (J/mole)
1-y
(6.50)
wherein
Yo- - groo = 45,360 - 18.4T (6.51)
ot I 'Y Boundary
The partial molar Gibbs energy "'0 of carbon in a-Fe from equation
(6.45) is equal to yo in y-Fe from equation (6.50) so that "'G- - groo +
RT In y'" is the left side of equation (6.50). The use of equation (6.44) to
eliminate gr 0° - '" 0- yields
( 6.55)
where A" is zero in accord with equation (6.45) and A Y is given by equation
(6.52). The equality of the left sides in these equations yields
The left side of this equation is the standard Gibbs energy of a to 'Y
transformation for pure iron, which must include the magnetic effects. The
equations representing YG~ - "G~ are complex; therefore, we present the
values in Table 6.4 in close temperature intervals as calculated from the
lengthy equations given by Agren 35 (cf. also Orr and Chipman 43 ).
Equation (6.56) with the data in Table 6.4, and equation (6.53) can be
solved simultaneously for each temperature to obtain the equilibrium values
of yet and y Y and thus calculate the a / 'Y phase boundary. These equations
are transcendental and require a computer or a programmable calculator
for iterative computations. Equation (6.44) gives the ferrite/graphite satur-
ation, equation (6.50) with yo
= grG o gives the austenite/graphite satur-
ation, and equation (6.56), based on "0 2 = Y0 2 , gives the equilibrium
concentrations of iron in ferrite and austenite. Simultaneous solution of
these equations for T, yet, and yY yields the eutectoid temperature and the
compositions of the eutectoid phases. A simple iterative computation for
this purpose is to calculate yet and yY from equations (6.44) and (6.50),
respectively, for each temperature in 1 K intervals from 1010 to 1015 K; the
set of values for one of the temperatures satisfying equation (6.56) yields
the simultaneous solution of equations (6.44), (6.50), and (6.56). Thus, it
can be shown that yet = 0.000319 (0.0206 wt% C) and yY = 0.0314
(0.671 wt% C) at T = 1013 K also satisfy equation (6.56) whose left is
Liquid Phase
The ')'-liquid and 8-liquid phase boundaries require the activity of
carbon in the liquid phase. The liquid is considered to be close-packed in
structure for which c = 1. The equation for 10 recommended by Ohtani et
al.,33 based on selected sets of data on gas-phase and dissolved carbon
equilibria,51-53 in conformity with equation (6.25), is
I
10 = grG O + 24,000 - 16.4T + (52 + 50.6T)yl + RT In ~ (6.57)
l-y
where
(6.60)
wherein AI of equation (6.58) has been used. The result for the
austenite/liquid boundary is
The left side of this equation is the standard Gibbs energy of melting for
pure y-Fe, which can be obtained from the compiled thermodynamic data42
by the following steps:
IIH~ - 'YH~ = 837 J/mole (at 1665 K); 8H~ - 'YH'2 = -6622 + 4.48T
6S~ - 'YS~ = -32.728 + 4.48 In T
The hypothetical melting point of y-Fe is obtained by setting the left side
of this equation to zero and solving for T; the result is 1797 K (l524°C)
which will be used later.
Equation (6.65) provides the values for the left side of equation (6.61)
which can be solved simultaneously with equation (6.59) to obtain the
values of y'Y and / in the range of 1424 to 1767 K, and thus calculate the
y/( y + liquid) and (y + Iiquid)/Iiquid boundaries. For example, at 1500 K,
equation (6.65) yields 2286 J/mole, and it can be shown that y'Y = 0.0810
(1.71 wt% C) and yl = 0.1758 (3.64wt% C) satisfy both equations (6.59)
and (6.61).
A small and very likely error of ±50J/mole in IG~ - 'YG~ causes large
errors in the values of y'Y and yl calculated by using equation (6.61).
Therefore, a greater degree of accuracy is required in equation (6.65) for
pure iron to achieve a higher degree of accuracy in the results from equation
(6.61). A comparable error in (21,360 + 2.0T) of equation (6.59) causes
considerably lower errors in the calculated values of y'Y and yl.
178 Chapter 6
Peritectic Equilibrium
The liquid, austenite, and 8 phases are in equilibrium at the peritectic
temperature. The equalities of partial molar Gibbs energies 1l0( = "0) =
"YO = 10, and 1102 = "Y02 = 102 yield four equations to solve for T, yll, y"Y,
and /. The equation for 8-Fe is the same as that for a-Fe without the
magnetic correction terms; hence, 110 = "YO yields equation (6.53) without
the second, third, and fourth terms from the end, i.e.,
This is the first equation. Further, "YO = 10 provides equation (6.59), which
is the second equation. For iron, 1102 = "Y02 yields the right side of equation
(6.56) with the superscript a replaced with 8, and the left side given by
equation (6.64), so that
in the selection of such data for pure iron (see also Schiirmann and
Schmid 54 ).
The 8/8 + austenite/ austenite boundaries were obtained by the same
procedure as that for the ferrite/ (ferrite + austenite) / austenite boundaries,
and the 8/ (8 + liquid)/Iiquid boundaries, as that for the austenite/
(austenite + liquid)/liquid boundaries by using the relevant preceding
equations.
Liquid/Graphite Boundary
The liquidus of graphite saturation, i.e., liquid/liquid + graphite phase
boundary in Fig. 6.7, is obtained by setting' G - gr 0° to zero in equation
(6.57) for each selected temperature. Thus, at 1650 K, y' = 0.2326
(4.76 wt% C), and at 1900 K, y = 0.2526 (5.15 wt% C). The compositions
of phases at the eutectic temperature were calculated earlier.
Cementite
The cementite phase, Fe3C, denoted as the 8-phase for brevity, is a
metastable phase, considered to have no deviation from stoichiometry as
the activities and temperature vary. Experimental evidence indicates that
deviations from stoichiometry exist, but all the phase diagram calculations
are based on perfect stoichiometry. The melting point of Fe3C is therefore
assumed to be congruent. The Curie point is 485 K and the structure is
orthorhombic. The solubility data for Fe3C in austenite and the CO-C0 2
equilibrium data 34.55 have been used to obtain the standard Gibbs energy
of formation of Fe 3C from pure '}'- Fe and graphite. Cementite, as a separate
phase, is in equilibrium with austenitic Fe and its carbon; therefore, from
the general relationship 0 = njGj + njGj,
(6.68)
The left side of this equation is the standard Gibbs energy of formation of
Fe3C; the first and second sets of parentheses are given by equations (6.55)
and (6.50) respectively, wherein yY must refer to the Fe3C saturation of
austenite. The result derived by Ohtani et at. 33 is
(6.72)
Now, 8Go is formed by the liquid phase and the graphite so that equation
(6.68) must be rewritten as 8Go = 3'G2 + 'G. Subtraction of (3'G~ + graO)
from both sides of this last simple equation gives
(6.73)
We use the right side of equation (6.72) for the left side of equation (6.73),
and then substitute equations (6.60) and (6.57) for their respective terms
in parentheses on the right side of equation (6.73), and rearrange the result
to obtain
This equation represents the liquidus for the cementite saturation of liquid
iron, as shown by the broken curve above 1417 K in Fig. 6.7. At 1475 K,
y' = 0.2516 (5.13 wt% C) is obtained from the preceding equation. At y' =
1/3 (x' = 0.25), which is the composition of Fe3C, this equation is satisfied
at 1524.1 K, and this is the melting point of Fe 3C. The eutectic temperature,
y"Y, and y' are calculated from equations (6.71), (6.74), and (6.61); the
results are T = 1417K,y"Y = 0.0991 (2.09 wt%), andy' = 0.2124 (4.37 wt%).
Interstitial Solutions 181
(6.75)
The terms in the first set of parentheses are given by equation (6.54). The
remaining terms require writing equation (6.45) as "'G - grGo =
"'G- - grGo + RT In y"', wherein "'G- - grG O is given by equation (6.44). Then
the calculation of the ferrite/ cementite boundary follows the same procedure
as that for the austenite/cementite boundary. The ferrite/austenite boun-
daries remain unaltered since they do not involve either graphite or the
cementite above the eutectoid temperature. The calculation of the eutectoid
temperature and compositions with Fe3C is similar to that with graphite.
Despite enormous numbers of investigations on the Fe-C and Fe-Fe3C
systems, new and more precise thermodynamic and metallurgical data are
needed in various regions of these systems. Improved modern techniques
invite reinvestigation of discordant results recently discussed by various
investigators. 33 -38
(6.76)
where C(g) is the monatomic gas, Vi is the vacant quasi-lattice site with;
atoms of Band Z - ; atoms of A, and C i is the interstitial atom occupying
this vacancy. The monatomic gaseous C is in equilibrium with its pre-
dominating diatomic species according to equation (6.16) for C 2 (g):
(6.77)
The change in energy for reaction (6.76) is simply E~ - E~s, where E~as is
taken as zero; hence, the energy with respect to gaseous atomic C is simply
E~. If C i at site; moves to site i + 1, with; + 1 atoms of B, Wagner writes
(6.78)
(6.79)
where all eCB are equal to each other, and all eCA are also equal to each
other. Recall that A and B form an ideal solution; hence, eAA, eAB, and eBB
are equal or they may be taken as zero without affecting our argument.
Substitution of equation (6.81) in equations (6.79) and (6.80) gives
(6.83)
For pure A and pure B as solvents, two boundary conditions exist for this
equation, Le., when i = 0 we have pure A and when i = Z we have pure
B, and therefore,
Yi = eci+l - i
ec (i = 0, 1, ... , Z - 1) (6.85)
YHI - Yi = h (6.86)
i-I
Yi - Yo = L (YHI - yJ = ih (6.87)
i=O
184 Chapter 6
E~+\ - E~ = ih + Yo (6.88)
i-Ii~O i+l
(Ec - Ed ==
i
Bc -
i 0
Ec =
(i-\i~O ) +
•
lh
•
lYo (6.89)
i i(i-1)h. 0
EC = 2 + 'Yo + Bc (6.90)
z Z(Z -1)h
Ec = 2
+ Zyo+ EC (6.91)
(6.92)
qc = N
Z
L
i=O
D~X~-iX~· e-E~/kT; Z']
[D'z-- (Z-i)!i!
. (6.93)
qc) Nc qc
G = -kTln ( - = -RTln-· (NckT = RT) (6.94)
c Nc Nc '
Nc
Gdrandom) = + RT In IV = RT In Xc; (e~ = 0) (6.95)
(6.96)
Equation (6.94) for the molar Gibbs energy of C in pure A, G~, is now
A)NC
G~=-kTln ( ~: =RTlnx~+Ncec (6.97)
( 6.98)
(6.100)
where Xc and Ie without superscripts refer to the alloy. This equation shows
that the activity of C over the entire range of binary composition must also
be the same for a fixed pressure of gaseous C 2 ; therefore, the activity
coefficient Ic in the alloy is defined by
or Ic =X~
-, (6.103)
Xc
(6.104)
Interstitial Solutions 187
The term inside the first set of brackets is (f~)-i/Z from equation (6.104),
and that in the first set of parentheses is x~1 K' from equation (6.99).
Substitution of these terms in equation (6.105) with II fc = xci x~, yields
the following Wagner equation:
(6.106)
1
fc = x~ + 4X~XB(f~)-1/4. e3h/2kT + 6x~x~(f~)-1/2 . e 2h / kT
Z [ i XA
i~O D z (f~)I/Z
]Z-i[ (f~)I/Z ]i
XB
[e
(Z-i)ih/2kT
]
(6.108)
f~
This is the original Wagner equation for which we shall have no specific
application in this book, because it is much simpler and preferable to use
f~ = 1 as the reference state, taken to be an infinitely dilute solution of C
in pure A. An unusual and interesting derivation of equation (6.108) is also
given by Blander and Saboungi.64
0
-2
\
\0
-
-4
~
0
c
-l
-6 ~
"'-9
-8
"<0"o~
...... ~Q)
0.0 1.0
Ag XPb
Pb
Figure 6.9. Variation of activity coefficients of sparingly dissolved oxygen,/o' in liquid Ag- Pb
alloys at 1273 K. Reference state is oxygen in pure Ag for'which I-;g = I. Data of Jacob and
Jeffes 6 '; solid curve is for hi k = 278.8 in equation (6.106). (Adapted from Chiang and Chang61
with permission.)
Interstitial Solutions 189
discussed earlier, and for oxygen and sulfur, the following reactions are
investigated:
(6.109)
(6.110)
(6.111)
The value of Kp for Cu is the same; therefore, for the same value of H 2/H 2 0,
f~u is given by
(1473 K)
Equation (6.106) can now be used with Z = 6 to obtainfo at XAg= Xcu = 0.5.
The result is fo(xcu = 0.5) = 0.00652, which is located on the upper curve'
in Fig. 6.8. Similar calculations for the Ag-Pb-O system are shown in
Fig. 6.9.
and Cu-Zn, and in Laves phases such as MgCU2, MgZn2' MgNh, and
MgZn2. This effect is due to the decrease in e~ with increasing screening
of protons (H+) by electrons in the alloys:
References
1. G. Alefeld and J. Voelkl, editors, Hydrogen in Metals I, and II, Springer-Verlag, Berlin
(1978).
2. G. A. Lewis, The Palladium Hydrogen System, Academic Press, New York (1967).
3. W. M. Mueller, J. P. Blackledge, and G. G. Libowitz, Metal Hydrides, Academic Press,
New York (1968).
4. H. Frieske and E. Wicke, Ber. Bunsenges. Phys. Chern. 77, 50 (1973).
5. E. Wicke and H. Brodowsky, with H. Zuchner, in Hydrogen in Metals II, edited by G.
A1efeld and J. Voelkl, Springer-Verlag, Berlin (1978).
6. T. B. Flanagan, S. Kishimoto, and G. E. Biehl, in Chemical Metallurgy-A Tribute to Carl
Wagner, edited by N. A. Gokcen, Metall. Soc. AIME, p. 471 (1981).
7. E. Wicke and G. H. Nernst, Ber. Bunsenges. Phys. Chern. 68, 224 (1964).
8. J. F. Lynch and T. B. Flanagan, 1. Phys. Chern. 77, 2628 (1973).
9. D. M. Nace and J. G. Aston, 1. Am. Chern. Soc. 79, 3619,3623,3627 (1957); J. G. Aston,
Engelhard Ind. Tech. Bull. 7, 14 (1966).
10. J. D. Clewley, T. Curran, T. B. Flanagan, and W. A. Oates, 1. Chern. Soc. Faraday Trans.
1 69,449 (1973).
11. S. Schmidt, in Hydrogen in Metals II, edited by G. Alefeld and J. Volkl, Springer-Verlag,
Berlin (1978).
12. G. Sicking, Ber. Bunsenges. Phys. Chern. 76, 790 (1972).
13. J. R. Lacher, Proc. R. Soc. London Ser. A 161, 525 (1937).
14. M. Shamsuddin and O. J. KJeppa, 1. Chern. Phys. 71, 5154 (1979); W. A. Oates and R.
Ramanathan, in Proceedings, 2nd International Congress on Hydrogen in Metals, Paris,
1977, Paper 2All, Pergamon Press, Elmsford, New York (1978); G. Bourreau, O. J.
KJeppa, and K. C. Hong, 1. Chern. Phys. 67, 3437 (1977).
15. M. Hillert and L.-1. Staflansson, Acta Chern. Scand. 24, 3618 (1970); see also M. Hillert
and M. Jarl, Metall. Trans. AIME 6A, 553 (1975).
16. C. Wagner, Z. Phys. Chern. Abt. A 193, 386, 407 (1944).
17. See, e.g., B. Baranowski, Part II in Hydrogen in Metals, edited by G. A1efeld and J. VOlkl,
Springer-Verlag, Berlin (1978), p. 157.
18. M. J. B. Evans and D. H. Everett, 1. Less-Common Met. 49, 123 (1976).
19. T. Takeshita, W. E. Wallace, and R. S. Craig, Inorg. Chern. 13,2283 (1974).
20. U. Kobler and J. M. Welter, 1. Less-Common Met. 84, 225 (1984).
192 Chapter 6
21. U. Kobler and T. Schober, 1. Less-Common Met. 60,101 (1978); see also T. Schober and
H. Wenzl, in Hydrogen in Metals II, edited by G. Alefeld and J. Volkl, Springer-Verlag,
Berlin, p. 12 (1978).
22. T. N. Veziroglu and J. B. Taylor, editors, Hydrogen Energy Progress V: Proceedings of the
5th World Hydrogen Energy Conference, Toronto, Canada, 15-20 July 1984, Pergamon
Press, Elmsford, New York (1984).
23. J. O. Bockris, Energy: The Solar Hydrogen Alternative, Wiley, New York (1977).
24. R. Wiswall, in Hydrogen in Metals II, edited by G. Alefeld and J. Voelkl, Springer-Verlag,
Berlin, p. 201 (1978).
25. J. J. Reilly, Z. Phys. Chern. 117, 155 (1979).
26. K. C. Hoffman, J. J. Reilly, C. H. Waide, R. H. Wiswall, and W. E. Winsche, Int. J.
Hydrogen Energy 1, 133 (1976).
27. G. G. Libowitz, H. F. Hayes, and T. R. P. Gibb, J. Phys. Chern. 62,76 (1958).
28. J. H. N. van Vucht, F. A. Kuijpers, and H. C. Bruning, Philips Res. Rep. 25, 33 (1970);
H. H. van Mal, Philips Res. Rep. Suppl. 1 (1976).
29. H. H. van Mal, K. H. J. Buschow, and F. A. Kuijpers, J. Less-Common Met. 32,289 (1973).
30. J. L. Anderson, T. C. Wallace, A. L. Bowman, C. L. Radosevich, and M. L. Courtney,
Hydrogen Absorption by ABs Compounds, Los Alamos Sci. Lab., Rep. LA-5320-MS (1973).
31. F. A. Kuijpers and H. H. van Mal, J. Less-Common Met. 23, 395 (1971).
32. G. Bambakidis, editor, Metal Hydrides, Plenum Press, New York (1981).
33. H. Ohtani, M. Hasebe, and T. Nishizawa, Trans. Iron Steel Inst. Jpn. 24, 857 (1984).
34. M. Hasebe, H. Ohtani, and T. Nishizawa, Met. Trans. 16A, 913 (1985).
35. J. Agren, Metall. Trans. AIME lOA, 1847 (1979).
36. H. Harvig, Jernkontorets Ann. ISS, 157 (1971).
37. J. Chipman, Metall. Trans. AIME 3,55 (1972).
38. O. Kubaschewski, Iron Binary Phase Diagrams, Springer-Verlag, Berlin (1982).
39. R. Hultgren, P. D. Desai, D. T. Hawkins, M. Gleiser, and K. K. Kelley, Selected Values
of the Thermodynamic Properties of Binary Alloys, ASM, Metals Park, Ohio (1973).
40. M. Benz and J. F. Elliott, Trans. Metall. Soc. AIME 221,323 (1961).
41. M. Hansen and K. Anderko, Constitution of Binary Alloys, McGraw-Hili, New York
(1958); First Supplement by R. P. Elliott (1965); Second Supplement by F. A. Shunk (1969).
42. R. Hultgren, P. D. Desai, D. T. Hawkins, M. Gleiser, K. K. Kelley, and D. D. Wagman,
Selected Values of the Thermodynamic Properties of the Elements, ASM, Metals Park, Ohio
(1973).
43. R. L. Orr and J. Chipman, Trans. Metall. Soc. AIME 239,630 (1967).
44. Y.- Y. Chuang, Y. A. Chang, and R. Schmid, Acta Metall. in press.
45. Y.-Y. Chuang, R .. Schmid, and Y. A. Chang, Acta Metall. in press.
46. R. P. Smith, J. Am. Chem. Soc. 68, 1163 (1946).
47. H. Schenk, M. G. Frohberg, and E. Jaspert, Archiv. Eisenhiitenw. 36, 683 (1965).
48. K. Bungardt, H. Preisendanz, and G. Lehnert, Arch. Eisenhuettenwes. 35, 999 (1964).
49. S. Ban-ya, J. F. Elliott, and J. Chipman, Trans. Metall. Soc. AIME 245, 1199 (1969); Met.
Trans. lA, 1313 (1970).
50. L. B. Pankratz, J. M. Stuve, and N. A. Gokcen, Thermodynamic Datafor Mineral Technology,
Bureau of Mines Bulletin 677 (1984).
51. F. D. Richardson and W. E. Dennis, Trans. Faraday Soc. 49,171 (1953).
52. S. Ban-ya and Y. Matoba, Physical Chemistry of Process Metallurgy, Interscience, New
York, pp. 373-402 (1961).
53. T. Mori, K. Fujimura, H. Okajima, and A. Yamanchi, Tetsu To Hagane 54, 321 (1968).
54. E. Schiirmann and R. Schmid, Arch. Eisenhuettenwes. 50, 101 (1971).
55. E. Scheil, T. Schmidt, and J. Wiinning, Arch. Eisenhuettenwes. 32, 251 (1961).
56. C. Wagner, Acta Metall. 21, 1297 (1973).
IDterstitial SolutioDS 193
57. L. Pauling, The Nature of the Chemical Bond, Cornell University Press, Ithaca, New York
(1960).
58. U. Block and H. P. Stiiwe, Z. Metallled. 60, 709 (1969).
59. R. J. Fruehan and F. D. Richardson, Trans. Metall. Soc. A/ME 245, 1721 (1969).
60. E. S. Tankins and N. A. Gokcen, High Temp. Sci 4, 393 (1972); E. S. Tankins, Metall.
Trans. A/ME 1,2637 (1970).
61. K. T. Jacob and J. H. E. Je/les, Trans. /nst. Min. Metall. CSO, 32 (1971); see also 1. Chem.
Thermodyn. 3, 433 (1971),5,365 (1973).
62. T. Chiang and Y. A. Chang, Metall. Trans. A/ME 78,453 (1976).
63. J.-c. Mathieu, F. Durand, and E. Bonnier, 1. Chim. Phys.62, 1289, 1297 (1965); B. Brion,
J.-C. Mathieu, P. Hicter, and P. Desre, 1. Chim. Phys.66, 1238, 1745 (1970).
64. M. Blander and M.-L. Saboungi, in Chemical Metallurgy-A Tribute to Carl Wagner,
edited by N. A. Gakcen, Metall. Soc. A/ME, p. 223 (1981).
65. N. A. Gokcen, Thermodynamics, Techscience, Hawthorne, California (1975).
66. N. A. Gokcen, Trans. Metall. Soc. A/ME 206, 1558 (1956); 197, 191 (1953).
67. M. R. Baren and N. A. Gokcen, in Advances in Sulfide Smelting, V. 1 Basic Principles,
edited by Y. H. Sohn, D. B. George, and A. D. Zunkel, AIME, Warrendale, Pennsylvania
(1983) p. 41; see also G. Urbain, W. Burgmann, and M. G. Frohberg, C. R. Acad. Sci.
Ser. C 263(8), 595 (1966).
68. K. T. Jacob and C. B. Alcock, Acta Metall. 20, 221 (1972).
69. H. Brodowsky and E. Poeschel, Z Phys. Chem.44, 143 (1965).
70. H. Brodowsky and H. Husemann, Ber. Bunsenges. Phys. Chem. 70, 626 (1966).
71. F. G. Jones and R. D. Pehlke, Metall. Trans. A/ME 2, 2655 (1971).
72. S. J. Wang and H. J. Grabke, Z. Metalld. 61, 597 (1970).
73. W. Siegelin, K. H. Lieser, and H. Witte, Z Elektrochem. 61, 359 (1957).
74. H. Schnabl, Ber. Bunsenges. Phys. Chem. 68, 549 (1964).
7
Semiconductors
Introduction
Crystal Structure
a b
c d
Figure 7.1. (a) Diamond lattice structure with outline of unit cell; (b) view of unit cell in [100]
direction; (c) view in [111] direction; (d) view in [110] direction. (From Green 7 with
permission.)
Semiconductors 197
shown in Fig. 7.1(b). Two other useful directions, [111] and [110], are
shown by (c) and (d), respectively. The crystal structure for Ge and Si is
called the diamond structure as indicated in Figs. 7.1(a) and 7.2(A). For
most of the useful Group III and Group V compounds, such as GaAs, the
crystal structure is quite similar, but designated as the zinc blende structure.
Figure 7.1 (b) indicates that these semiconductors are tetrahedral, i.e., each
atom has four closest neighbors and therefore the structure can be represen-
ted as shown in Fig. 7.2. Other interesting semiconductors such as CdS and
ZnS are in the wiirtzite structure, and PbS and PbTe are in the rock-salt
structure, but we shall not be concerned with the last two compounds in
this chapter since we wish to limit ourselves to the structures shown in
Figs. 7.1 and 7.2.
Band Structure
A B
• E leclron, eo • E leclron
eGe or Si eGo
GAs
Figure 7.2. (A) Structure of germanium and silicon; (B) structure of GaAs; CdS and ZnS are
identical with GaAs. Each atom shares four electrons with its neighbors. When an electron,
eo, jumps to the conduction band, it forms a free electron e- and leaves a hole e+ in the
valence band as shown in (A). Double lines from two neighboring atoms signify one bond of
previous chapters.
198 Chapter 7
Conduction band
4 states/atom
empty at 0 K
p-Ievel
s -level
w
r~
(!)
a::
w
z Valence band
Figure 7.3. Schematic energy levels of Ge
w 4 states/atom
full at 0 K and Si as a function of distance between
atoms. Band-gap energy, E g , separates
Spacing in crystal valence and conduction bands.
by condensed atoms, is shown in Fig. 7.3. Half of the s- and p-Ievels merge
into a lower band filled with four electronic states, completely full at 0 K;
the resulting lower band is called the valence band. The remaining four
electronic states formed by the remaining four split levels merge to form
the conduction band. As the temperature increases beyond 0 K, or photons
of proper energy levels are injected into the semiconductor, electrons are
excited into the conduction band. The energy Eg required to excite an
electron into the conduction band is called the forbidden band-gap energy,
or briefly the band-gap energy, and often band-gap when a specific number
follows this term. The covalent bond energy of semiconductors such as Si
w
-----Ef------ BAND
Eg EDGES
o
1~(;;~!~f_'w*~r~m~!if~~jN1;~!~1~";
CRYSTAL DIMENSION
Figure 7.4. Simple band diagram for a semiconductor. Energy is measured from valence band
edge; hence, E g , E" and energy of e - are positive, and energy of e + is negative.
Semiconductors 199
and Ge is high; hence, their entropies of fusion are much larger than those
of metals since each atom in Fig. 7.2 contributes four electrons and shares
four electrons with four neighbors, and thus completes the octet. The
structure of most interesting semiconductors is tetrahedral, or the diamond
lattice type. Covalent compounds of Group II and Group VI (II-VI), e.g.,
CdS and ZnS, or III-V, such as GaAs and InSb, also have four electrons
per atom as in Si and Ge. For simplicity, the band structure is drawn in
block diagrams as shown in Fig. 7.4. The upper band is the conduction
band for occupancy by the electrons that carry electric current. The lower
band is for the valence electrons, and as will be seen later, the lower band
is also for the holes, which also carry electric current. Here the horizontal
axis is not important, but it is sometimes designated as the crystal dimension,
but the important point is the vertical axis which represents the energy
required to excite an electron to the conduction band.
Distribution of Electrons
N 1
g == 4>_(E_) = 1 + exp[(E _ _ E r)/ kT] "" exp[ -(K - Ee)/ kT] (7.1)
I
lLJ
~
z
Q
~ 0.5 . . . .
al
a::
I-
Figure 7.5. Distribution of fermions as CIl
a function of energy at 0 K and at is
T> 0 K. At E_ = Ee and T> 0,
<IdEe) = 0.5. At T = 0 K, <ldE_) = 1
for E_ < E e, and tP_(E_) = 0 for E_ > Ef
Ee· ENERGY, E
200 Chapter 7
where the equality holds for the general case, and the approximate equality
holds for the values of E_ - Er several times larger than kT so that "1" in
the denominator can be neglected. The electrons have a finite probability
of occupying the conduction band when their energy E is equal to or greater
than the band-gap energy Eg. The states for 0 < E_ < Eg are forbidden;
therefore, electrons with E_ < Eg have no allowed states and they must
return to the valence band. For pure and mildly alloyed (i.e., doped)
semiconductors, E_ - Er » kT is always satisfied, and the approximate
equality in equation (7.1) is valid; such semiconductors are called nondegen-
erate semiconductors. Conduction of electricity is due to the electrons in
the conduction band and the holes in the valence band. As the temperature
increases, the number of electrons and holes, hence the conductivity of
semiconductors increases, whereas the conductivity of metals decreases with
increasing temperature. When an electron leaves the valence band of a
semiconductor to occupy the conduction band, it leaves a positive charge
in the valence band that acts in every respect as a positively charged electron,
e +. Attempts to explain the properties of holes by any other scheme than
the positively charged particles have not been successful. The electrons and
holes are collectively called the carriers since both carry electricity. The
band-gap of a semiconductor such as Si is 1.12 eV, whereas the band-gap
of an insulator such as diamond is 7 eV. In a metal, such as Na or K, Er
is within one of the allowed bands for the valence electrons, and nearly all
the valence electrons move freely as an electron gas and conduct electricity.
(7.2)
\ I
\ ,/
',.... ",,'"
>-
<.!)
lr
W
Z
W
Eg
Figure 7.6. Energy-momentum diagram
MOMENTUM
for indirect band-gap semiconductors.
For semiconductors such as silicon, a
direct band-gap may also exist as
between Py = 0 and D.
where Ey is the energy of the valence band edge, E+ is the energy of the
hole, and m+, the effective mass of the hole. Here P is measured from the
top of the valence band which is taken to be zero. Note that the band-gap
energy Eg is given by
(7.4)
Solid curves in Fig. 7.6 are for an indirect band-gap semiconductor, e.g.,
silicon and germanium, for which Pc is not zero. In a direct band-gap
semiconductor such as GaAs, Pc = 0, i.e., the extrema of the two curves
occur at the same value of p, but again with Eg = Ec - Ey. The vertical
band-gap or direct band-gap energy for Si or Ge is about twice as high as
its indirect band-gap energy E g , but for direct-gap type semiconductors
such as GaAs, vertical band-gap is identical with the usual band-gap energy,
E g • We shall have no particular application for the vertical band-gap energy
for indirect gap semiconductors; hence, the band-gap will always refer to
the lowest value of Eg which may be direct (vertical) or indirect (nonvertical).
Actual momentum-energy diagrams for the possible values of p in different
crystal directions are complex and need not be discussed here for our
purposes. The electron and hole masses are the average masses that fit the
properties under consideration as will be seen later.
202 Chapter 7
Dopants in Semiconductors
I n
Figure 7.7. (I) Phosphorus as ionized n·dopant; free electron e- given up by P is in the
conduction band. (II) Boron as ionized dopant; free hole e+ in the valence band is formed
upon acquisition of an electron from top of the valence band by B, which then forms B- ion.
Semiconductors 203
the electrons, e -, are the majority carriers, and the holes, e +, are the minority
carriers.
Addition of a dopant such as boron from Group III into a substitutional
lattice site has an opposite effect on the free electrons. Boron has only three
(s + p) electrons, and in bonding with the neighboring silicon atoms, a boron
atom attempts to borrow one electron from its neighbors and generate a
positive particle free to move in silicon. Thus, the borrowed electron on B
leaves a deficiency of one electron in the neighboring Si atoms and this
deficiency is the same as generation of a hole in the silicon valence band.
A p-dopant is also called an acceptor because it accepts electrons from the
semiconductor and thus creates holes in the valence band. Again, for a
given temperature and for light doping with B, the product n_n+ is nearly
the same as that for pure Si but the total number of carriers (n_ + n+)
increases dramatically by doping with B. The silicon doped with Group III
elements is called p-doped or p-type silicon and B is called a p-dopant.
Some elements may act as either p-, or n-dopant; e.g., Si as dopant in GaAs
may act amphoterically, i.e., if Si occupies the Ga lattice site in GaAs, it is
an n-dopant, whereas when it occupies the As lattice site, it is a p-dopant.
This can be paraphrased by Si ~ Si+ + e- when Si is an n-dopant in GaAs
and by Si ~ Si- + e+ when Si is a p-dopant. In both cases, the dopant
ionizes to yield an ion and a charge carrier. A Group IV element such as
carbon as a dopant may act as a neutral atom in Si. The ionization of a
dopant is a function of temperature, but beyond some range of temperature,
ionization is very nearly complete as will be discussed later. At 0 K, each
dopant captures its carrier and becomes neutral. For example, p+ captures
an electron and B- captures a hole to become neutral substitutional atoms.
The ionized dopant atoms above 0 K cannot move rapidly in a semI-
conductor to contribute significantly to electrical conductivity.
The net kinetic energy of an electron was given by equation (7.2). The
electrons in the conduction band are free particles, and according to
quantum mechanics of free particles, they are subject to the Pauli exclusion
principle; this energy is given by
(7.5)
(7.6)
where N+ is the density of states for holes. The negative sign arises from
the fact that the energy of holes is measured downward from the top of the
valence band; hence, it is always negative.
Concentration of Electrons
The energy E_ of electrons in the conduction band has an additional
term which is the band-gap energy Eg; i.e., for the electrons in the conduction
band, Eg must be added to its net kinetic energy because the energy is
measured from the valence band, Ey; hence,
(7.7)
Differentiation of this equation after writing the left side as (E_ - Eg)I.S,
and rearrangement of the result yields
(7.8)
where p(E_) is the density of states per unit energy at E_ so that dN_ =
p(E_) dE_ is the number of states within a narrow energy dE_. The con-
centration of electrons n_ in the conduction band is then the integral of
cf>_dN_, Le.,
Note that at energies E_ < E g , the electrons do not have enough energy to
hop in the conduction band. The integration is carried out by substituting
Semiconductors 20S
and observing that (E r - Eg) is constant for a given temperature, and further,
dE_ = d(E_ - Eg). Substitution of these relationships in equation (7.9)
yields
U fooE~Eexp (EgkT-
exp ( E - E )
.
4".
n_=J;3(2m_)\.5
r - E )
-
x (E_ - Eg)05d(E_ - Eg) (7.10)
(7.11)
where the density of state N _ is the coefficient of the exponential term, i.e.,
(7.12)
Concentration of Holes
The Fermi distribution function for holes cfJ+ is symmetric to that for
electrons and symmetric with respect to Er in energy so that cfJ+ = cfJ+(E+)
is given by
(7.13)
Here E+ is measured from the top of the valence band; hence, the numerical
value of E+ is negative, and the approximate equality is valid when E r -
E+ » kT. The energy of a hole is the same as its kinetic energy because
holes need not overcome a barrier such as the band-gap energy to acquire
additional energy. The density of states p+ within a narrow energy difference
d(E+) is obtained by differentiating equation (7.6); the result is
(7.14)
206 Chapter 7
n+ = fo
-00 p+cP+ dE+
47T
= f1 (2mS· 5
fO
-00 ( - E+)0.5 exp[(E+ - Er)/ kT] dE+
2
= h3(27Tm+kT)15exp(-ErlkT) (7.15)
where the integral is from -00 to 0 because the permissible energies of holes
are measured from the top of the valence band in the negative direction.
The product term before the exponential term is called the density of states
for holes, i.e.,
(7.16)
For other semiconductors, the correction from the logarithmic term might
be an order of magnitude higher, but the preceding approximate equality
is still valid.
where eo is the electron bound in the valence band that generates e- (i.e.,
electron) in the conduction band and e+ (i.e., hole) in the valence band,
Semicooductors 207
and eo acts in the same way as a 1-1 electrolyte in water. The electron eo
will always be referred to as the bound electron to avoid confusion with e -.
The band-gap of a semiconductor can be measured by the injection of
photons whose energy is gradually increased toward the band-gap of the
semiconductor. No change in the conductivity of semiconductors is detect-
able when the photon energy is less than the band-gap energy, but when
the photon energy is equal to or greater than the band-gap energy, the
conductivity of semiconductors sharply increases due to the generation of
charge carriers. The number of quanta of photons may be kept to a minimum
during the measurements so that the concentrations of e- and e+ at thermal
equilibrium are not significantly disturbed, and further, in principle, reaction
(7.18) can generate a photon or corresponding energy upon the reversal of
reaction (7.18) so that equilibrium can be assumed. The energy thus involved
is therefore diJ:ectly related to the formation of the reaction products from
one reactant eo; hence, aG = Eg. Another convincing argument that Eg
is a chemical potential or a Gibbs energy per particle is presented by
Thurmond,IO and paraphrased by the following equation:
Eg = (-
iJhVN~)
- = hv = aG- (7.19)*
iJn_ S,V;x
*Precise measurement of band-gap energies may involve phonons, which will not be discussed
here.
208 Chapter 7
where q_ and q+ are the partition functions, and V is the molar volume of
the semiconductor in cubic centimeters. The concentrations of electrons
and holes may be expressed by n_1 N_ and n+1 N+ which are dimensionless
properties similar to mole fractions. 1O The particle Gibbs energies G_ and
G+ as functions of these concentrations are
n_
G- = GO- +kTln-
N_ (7.23)
(7.24)
where G~ and G~ are the standard Gibbs energies on the Henrian scale
for dilute solutions. 11 It should be noted that nl N for electrons and holes
are considerably smaller than unity for intrinsic and lightly doped semicon-
ductors so that Henry's law is obeyed, or the activity coefficients are unity.
This treatment is not concerned with heavily doped semiconductors for
which the activity coefficients differ from unity, and modified statistical
thermodynamic treatments become necessary. The standard state is the
hypothetically existing state for which nl N is unity and G = GO for _ and +
particles. It is important to stress here that in these hypothetical standard
states, both charge carriers are free particles, or boltzons, obeying classical
statistics because e- is above the Fermi level and e+ is below the valence
band edge.
It is necessary to assume here that the bound electrons participating
in reaction (7.18) are those right on the valence band edge, and their Gibbs
energy is Go = Gt + kT In Xo where Gt is a standard Gibbs energy and Xo
is a very small fraction of all valence electrons. If Xo is of the order of 10-4 ,
then a small fraction, such as 1O-5xo, ending up in the conduction band in
an intrinsic semiconductor cannot change the value of Xo significantly.
Therefore, the term kT In Xo can be added to Gt to write Go = Go for the
electrons participating in reaction (7.18). The equilibrium in reaction (7.18)
requires that
(7.25)
Semiconductors 209
(7.26)
Equations (7.11) and (7.15) can be used to obtain the logarithmic term in
equation (7.26), eliminate E r, and thus show that
(7.27)
where ± is for "either + or -" but not both, No is Avogadro's number, and
G~(at 0 K) is zero as T approaches zero in - T In(q±/ No). Substitution of
appropriate values in q± yields
G~ + G~ - G~(O K) - G~(O K)
= -2kT(74.753 + 1.5In(m) + 1.5 In T + In V) (7.29)
where (m) = (m_m+)o.s is the geometric mean of the carrier masses, which
usually increases very slowly with temperature. The carrier masses used
here are called the density of state carrier masses. It will be assumed here
that the temperature dependence of (m) is negligible, in reasonable agree-
ment with the summary presented by Thurmond. lO
The electrons in the valence band move in a highly restricted volume
bound by the atomic core on one side and the electron cloud of the
neighboring atoms on the other side. The partition function of this restricted
cloud of particles is difficult to formulate but it is proposed here that
(7.30)
210 Chapter 7
where Vv is the restricted volume available for bound electrons very close
to the valence band edge, which is assumed to be a very small fraction A
of the semiconductor volume V. Further, Vv may contain other corrections
such as electron spin effects, and those due to the splitting of bands in the
semiconductor, and interactions of electrons within each band. It was shown
in Chapter 3 that the contributions of rotation of a bound electron about
an atomic core and its vibration with the core are quite negligible because
the moment of inertia of an electron and the distance between the electron
and the shielded proton are both very small. Thus, A V consists of very
thin peripheral ribbonlike volumes joined to each other in such a way that
the motion of bound electrons is very much limited and may involve
occasional exchange of positions with the neighboring peripheral volumes.
The corresponding Gibbs energy for a valence electron from equation
(7.30) is
=
(m)
kT ( 18.642 - 3 In - + In -A - 1.5 In T ) (7.32)
mo V
representing the experimental values ranging from 200 to 420 K for silicon.
Precise values are not yet available above 420 K.
Equation (7.33) is based on only one parameter, A, and it is quite likely
that the errors involved in (m), mo, and their temperature dependence as
well as V, Vv , and their temperature dependence, could introduce corrections
for the terms in equations (7.32) and (7.33). Therefore, it is assumed that
the coefficients in equation (7.33) are adjustable within reason if the
accuracy of band-gap measurements are indeed as accurate as correctly
stated by Thurmond 1o; consequently,
(7.34)
This equation is represented in Fig. 7.8. Likewise for Ge, the corresponding
equation is
400
0.0
0.0
-0.02
-0.1
-0.04
-0.2
-0.06
> >
~
0- -0.3 ~
w
w
<J -0.08 <J
-0.4
-0.10
-0.5
-0.12
-0.6
-0.14
0.0 1400
T/K
Figure 7.8. Variation of band·gap energy with temperature. Points are experimental data from
various sources, evaluated and presented by Thurmond 1o ; curves are from equation (7.35) for
Si, and equation (7.38) for GaP.
Semiconductors 213
Equation (7.38) is plotted in Fig. 7.8. Equations (7.37) and (7.38) have been
obtained by substituting the experimental best-fit data at 300 and 800 K
from a smooth curve. The experimental results for GaAs were obtained
from 0 to 1000 K, and for GaP, from 0 to 1273 K. Both sets of data are
perfectly represented by equations (7.37) and (7.38). Accurate representa-
tions of data are important in deriving the thermodynamic properties
obtained by differentiation of the equations for !lEg as functions of tem-
perature. However, the form of the selected equation is also very important. 12
Further justification for equation (7.34) will be dealt with in the remain-
ing sections. It is now significant to point out that below approximately
73 K, !lEg becomes zero for Si according to equation (7.35), in agreement
with experimental data, and similarly, !lEg is zero for Ge, GaAs, and GaP
below 33, 17, and 61 K, respectively. The data for Si particularly support
this conclusion as shown in Fig. 7.8. We call these points "characteristic
temperatures" for temperature independence of band-gap energies. This is
reminiscent of the disappearance of temperature dependence of a number
of other physical properties in solids. The lower temperature limit of physical
validity of equation (7.34) is therefore when the experimental values of !lEg
become independent of temperature. Nevertheless, equation (7.34) con-
tributes very little to !lEg as 0 K is approached because T appears as a
coefficient in both terms.
a!lGO
- -- = !lSo = B- A + B In T (7.39)
aT
where !lSO is the standard entropy change for reaction (7.18). The results
for the foregoing semiconductors are
(7.43)
214 Chapter 7
It is evident that as the temperature decreases below exp[(A - B)/ B], e.g.,
below 27 K for Si and below 12 K for Ge, the numerical value of t:.so
changes from positive to negative. Further, t:.so approaches -00 as T
approaches zero in accord with thermodynamic computation for free elec-
trons \3 and the fact that the standard states must be chosen as dictated by
the partition functions for boltzons in equations (7.21) and (7.22).
The values of t:.eo for the standard heat capacities are given in equations
(7.35)-(7.38), and they indicate that t:.eo is not greatly different from 1.5k
as required for the free particles or boltzons at constant volume.
The equation for the standard enthalpy change, t:.Ho, can be readily
obtained from the definitional equation t:.Go = t:.Ho - Tt:.So. The result
from equations (7.34) and (7.39) is
(7.44)
Ionization Equilibria
It was stated earlier that a dopant atom ionizes to yield a free or moving
carrier and a stationary charged ion. The energy levels of both the donors
and acceptors lie within the band-gap of semiconductors. The energy
required to obtain an electron is measured from the donor level to the
conduction level as a positive quantity, and this energy is a small fraction
of the band-gap, usually about 0.1 eV or less, as shown in Fig. 7.9. The
energy required to obtain a hole is measured from the valence band to the
Semiconductors 21S
acceptor level as shown in Fig. 7.9 and it is usually about 0.1 eV or less.
Actually, each of these levels has closely spaced levels, but for simplicity
it is usually shown as a single level.
The ionization equilibria, as typified by the following reactions for
phosphorus and boron in silicon, are as follows:
(7.45)
(7.46)
(7.47)
(7.48)
It was shown by equation (7.17) that the Fermi level (Fermi energy)
in an intrinsic semiconductor lies in the close neighborhood of E g /2. In
n-doped semiconductors, the Fermi level varies with both doping level and
temperature according to equation (7.11) with (7.12) so that
(7.49)
Increasing the n-doping level, hence increasing n_, increases n_/ N_ toward
unity and Er increases toward E g • Further, as temperature decreases, the
right side of equation (7.49) decreases because n_/ N_ is less than unity for
reasonable degrees of doping and In( n_/ N _) varies slower than T; therefore,
decreasing temperature also increases Er toward E g •
An increase in the p-doping level has the reverse effect as shown by
using equation (7.15), i.e.,
(7.50)
p-n Junctions
+ c
DISTANCE FROM DISTANCE FROM
~J~U~N~C~T~IO~N~~O~__J~U~N~C~T~I~O~N___
Figure 7.10. (A) Fermi levels of n- and p-type semiconductors. (B) Fermi levels coincide after
junction formation; conduction and valence band are bent as shown. Near the junction,
negative ions e and positive ions EEl prevent excessive migration of carriers by establishing
the electrostatic potential shown in (C).
218 Chapter 7
Figure 7.11. Energy levels of electrons and built-in potential Vb; across an unbiased junction.
Semiconductors 219
-.+
/ ~e- I "'\
V<O
~r---------~~--------
z....t
__ 0
01
"--1 p - type j::1
UI
n-type
~
ZI
=>1
-:I .....
B
Figure 7.12. Energies in a reverse-biased p-n junction. Applied voltage, V, is considered to
be negative in this configuration; eV is the applied energy.
The net electronic current to the right side of the p-n junction is obtained
by subtracting equation (7.54) from equation (7.53); thus,
a exp
_(E; kT-E:) -_ exp(-EkT 10
g) =
- Irs (7.56)
where a exp(E: / kT) = 10 , and the right side of this equation is called the
reverse saturation current Irs; therefore,
(7.57)
This equation is for an ideal junction, i.e., ideal diode; however, two
adjustable parameters A and B are needed to make this equation more
general, i.e.,
(7.58)
where A and B are between 1 and 2, i.e., 1 :0;;; A :0;;; 2 and 1 :0;;; B :0;;; 2. This
equation is known as the Shockley diode equation. 17
It can be shown by using Fig. 7.13 that equation (7.55) is also valid
for a forward-biased p-n junction. In this case, the electrons on the right
+. -
/" I
e - , "'\
V>O
br---------~~--------~ 0
~
_I
P-type 1-1
ul
n - type .....-I
ZI
::::l...l
~I
A
f g::E;-Ef -eV
-t.---T
+ +
I :
Eg -Ef 1
+_____ ~ J-ev
Ef I _
Ef
B
Figure 7.13. Forward-biased p-n junction.
Semiconductors 221
side must overcome the energy from the Fermi level E r on the right to the
conduction band level of the left side, and this energy is equal to E; - E: -
eV because Er for the right side is now raised by ev'
The coefficient 10 in equation (7.58) is largely a property of electrons;
hence, it may be assigned l8 a value of 6.03 x 109 milliamp/cm 2, and for a
simple ideal diode, i.e., A = B = 1, equation (7.58) becomes
Irs = 6.03 X 109 exp( -1.5 X 1.60219 X 10- 19/1.38062 X 10- 23 X 300)
= 3.808 X 10- 16 mA/cm 2 (7.60)
DARK I-V-
v
o
ILLUMINATED I
J\ Voe
I-V"",/
Solar Cells
The solar cells are p-n junction devices that convert solar energy
directly into electricity by the photovoltaic effect (PVE). This occurs when
the energy of photons is equal to or higher than the band-gap of the
semiconductor. Such photons are said to be ionizing radiation. Pure single-
crystal semiconductors are transparent to photons with energies smaller
than the band-gap energies. The ionizing radiation generates holes and
electrons in excess of those corresponding to the thermal equilibrium. The
excess electrons on the p-side in Fig. 7.10 are attracted by the positive ions
on the n-side of the junction and the excess holes generated on the n-side
are attracted by the negative ions on the p-side of the function and thus
the charges are separated at the p-n junction where a dipole layer (or
space-charge region) exists.
The charge carriers thus separated are the minority carriers, and their
separation increases the overall concentration of holes on the p-side and
electrons on the n-side. Consequently, the p-type semiconductor becomes
a source of positive electricity, and the n-type semiconductor, a source of
electrons. A shorted solar cell yields a current that is called the short-circuit
current, I se , but when a cell has no external load, it generates a voltage
called the open-circuit voltage Voe' The maximum power obtainable from
Semiconductors 223
a cell is the product of maximum power current Imp and maximum power
voltage Vrnp , which are situated at the knee of the I-V curve shown in Fig.
7.14. Note that this curve is in the fourth quadrant because the external
current flows in the same way as in the reverse-biased junction in Fig. 7.12
while the measured potential is opposite in direction. The potential
difference between the two ends of the space-charge region sets an upper
limit to the magnitude of voltage produced by photons. This difference is
determined by the difference in the Fermi levels of p- and n-semiconductors
and, hence, by the doping levels. The potential difference is high when the
doping levels are high and the junction profile is sharp in concentration
gradient. The lifetimes of the photogenerated carriers should be sufficiently
long so that they can be collected by the conductors attached to the p- and
n-sides.
A diagram of a typical silicon solar cell and a photograph of a commer-
cial cell are shown in Figs. 7.15 and 7.16. The top layer, exposed to the
solar radiation, has collector bars attached to the surface by various tech-
niques for collecting the electrons. The surface area covered by the bars is
usually of the order of 3 % of the top surface area. The bottom layer can
be made to have collector bars for tandem or stacked-up cells, but for single
cells, it is usually covered or soldered with a metallic layer. A space-qualified
silicon solar cell of 2 x 2 cm can generate about 55 mW at approximately
0.55 volt, and 100 rnA at air mass zero (AMO) radiation outside the atmos-
pheric region of the earth. Various aspects of solar cells, their manufacture
and uses are discussed well in a number of monographs,s-9 and in detail
in the IEEE Photovoltaic Specialists Conference Records. 19 The remaining
sections of this chapter are largely based on a paper by Gokcen and
Loferski. 18
p-type
Bock contoct/
Figure 7.15. Diagram of a solar cell.
224 Chapter 7
where A will be used hereafter for e/ kT for brevity. The derivation of this
equation is based on the fact that the I-V curve of a nonilluminated junction
in Fig. 7.14 is displaced downward by Ise, but otherwise the curve maintains
its shape. When I, = 0, V = Voe , and Ise is then correlated with Voe from
Semiconductors 225
Isc r------__
Imp
tI
(7.64)
The equivalent circuit of such a cell is shown in Fig. 7.18. The maximum
SOLAR
-1
~R_D m_p~~mp
CELL
, -__ ___
Figure 7.1S. Simple equivalent circuit of a solar cell. Dynamic impedance of cell is RD. For
maximum power voltage Vmp • resistance at mp should match RD.
226 Chapter 7
(7.66)
(7.67)
The maximum power TImp is Imp from this equation multiplied by Vmp in
volts; the result is
The ratio TImp/ (Ise VoC> is called the fill factor. It is now appropriate to
proceed to tandem or stacked solar cells.
(7.69)
In order to simplify the notation in this chapter, the band-gap of the cell
being considered is denoted by E g , and that of the cell acting as a filter, by
E. It is evident that for a selected cell, Eg is fixed and E may assume various
values in excess of E g •
SemiCODductors 227
The reason for the efficiency increase provided by tandem solar cell
systems lies in the fact that they bring about a closer match between the
maximum energy photons hllmax absorbed in a given cell and the energy
gap Eg of the absorbing semiconductor. The excess energy hll - Eg of an
absorbed photon degenerates into heat. However, as the number of cells
increases, the photon flux available for absorption in anyone semiconductor
in the stack decreases which leads to a lowering of photovoltaic conversion
efficiency of each cell. The total photon flux, incident on the tandem cell
stack, can be increased by recourse to concentrating mirrors and lenses,
thus circumventing this last-named efficiency reduction. It becomes evident,
therefore, that the tandem cell concept is particularly compatible with solar
concentrator systems. This is potentially of great importance in view of
studies of large-scale solar energy conversion systems which show that the
combination of high levels of concentration with high-efficiency cells is one
possible way to achieve economically competitive photovoltaic solar cell
systems?O
The next section explores the theoretical limits on solar energy conver-
sion efficiencies at various temperatures, various concentration ratios, and
various numbers of cells in the tandem cell stack. Four temperatures (T =
200,300, 400, and 500 K), four concentration ratios (C = 1, 100, 500, and
1000 suns), and cell numbers up to 24 are considered. The energy gaps
of the photovoltaically active semiconductors which result in optimum
efficiencies are also presented.
Calculation Procedure
(7.70)
where nph(hll) is the photon flux (per cm 2 ) per hll. Table 7.1 also includes
the short-circuit current which would flow in a cell having an absorption
cutoff at hll = E if the collection efficiency were unity, i.e., if every photon
...
Table 7.1. Limiting Short-Circuit Current as a Function of Energy Gap Eg for Air Mass Zero (AMO) Solar Radiation a ~
Solar Solar
Isc(E) energy I,c(E) energy
E = hI! A (/Lm) 10- 14 Nph(E) (mA/cm 2) (mW/cm2) E = hI! A (/Lm) 10- 14 Nph(E) (mA/cm 2) (mW/cm2)
aThe photon flux Nph(E), defined as the number of photons per em' per second having energy greater than E, is taken to be zero at E = 7 eY. Short·circuit current
l,c(E) = eNph(E). The column labeled "Solar energy" is the solar energy density delivered by photons having energies in excess of E = hll.
.1=-n
.
ooool
Semiconductors 229
having an energy hll > E contributed one pair of minority carriers to the
short-circuit current. Thus,
(7.71)
where e is the charge on the electron in coulombs. For example, Table 7.1
shows that for a cell in which the AMO sunlight is incident on a photovoltai-
cally active semiconductor with an energy gap Eg = 3.00 eV, the limiting
value of Ise is 3.993 mAl cm 2 •
If a cell based on a photovoItaicaIly active semiconductor having an
energy gap Eg has interposed between itself and the solar source a filter
which cuts off all photons with energy greater than E, where E > Eg, then
the limiting short-circuit current which this cell can produce is obtained
from the relation
For example, again referring to Table 7.1, if E = 3.40 eV, and Eg = 3.00 eV,
then Ise(3.00 eV, 3.40 eV) = (3.993 - 2.006) mA/cm 2 = 1.987 mA/cm 2 • For
the required calculations, Ise(Eg, E) will be used in equations (7.66) and
(7.68) instead of I se , and Eg will be used in the relationship for Ise in
equation (7.59) to calculate Irs. The solar energy conversion efficiency TJ of
a single cell is
(7.73)
where 135.3 mWI cm 2 is the AMO solar spectrum intensity, which is the last
entry in Table 7.1, and C is the concentration ratio expressed by the number
of suns.
Calculation for an example with Eg = 1.5 at 300 K for a single cell and
C = 1 requires first Irs from equation (7.59), which was found earlier to be
Irs = 3.808 X 10- 16 mAl cm 2 • The value of Ise = 35.227 mAl cm 2 is listed in
Table 7.1 for E = Eg = 1.5 eV; therefore, Isel Irs = 35.227/3.808 X 10- 16 =
9.2508 X 10 16 ; also, A = el kT = 11,604.9/300 = 38.6829, and substitution
of these values in the logarithm of equation (7.66) yields
E
(eY) 0.60 0.70 0.80 0.90 1.00 1.10 1.20 1.30 1.40 1.50 1.60 1.70 1.80 1.90 2.00 2.10 2.20 2.30 2.40 2.50 2.60 2.70 2.80 2.90 3.00 3.20 3.40 3.60 3.80 4.00 6.90
0.60 0.00
0.70 0.10 0.00
0.80 0.33 0.48 0.00
0.90 0.60 1.10 0.89 0.00
1.00 0.86 1.66 1.75 1.09 0.00
1.10 1.14 2.25 2.66 2.31 1.47 0.00
1.20 1.41 2.82 3.54 3.50 2.96 1.71 0.00
1.30 1.68 3.38 4.41 4.67 4.44 3.48 1.99 0.00
1.40 1.93 3.89 5.19 5.73 5.78 5.10 3.87 2.07 0.00
1.50 2.17 4.36 5.91 8.71 7.01 6.59 5.62 4.05 2.16 0.00
1.60 2.39 4.81 6.60 7.64 8.19 8.01 7.28 5.95 4.28 2.29 0.00
1.70 2.60 5.22 7.23 8.48 9.26 9.30 8.79 7.69 6.23 4.45 2.31 0.00
1.80 2.79 5.60 7.80 9.25 10.22 10.47 10.16 9.26 8.01 6.42 4.46 2.29 0.00
1.90 2.97 5.97 8.36 10.01 11.18 11.63 11.53 10.82 9.77 8.38 6.61 4.62 2.47 0.00
2.00 3.14 6.29 8.84 10.66 12.Dl 12.63 12.69 12.16 11.28 10.06 8.46 6.64 4.65 229 0.00
2.10 3.30 6.60 9.32 11.30 12.81 13.60 13.84 13.48 12.76 11.71 10.28 8.63 6.80 4.59 2.42 0.00
2.20 3.44 6.87 9.73 11.85 13.51 14.44 14.82 14.60 14.04 13.13 11.85 10.33 8.65 6.58 4.54 2.21 0.00
2.30 3.56 7.11 10.08 12.33 14.12 15.17 15.68 15.59 15.16 14.38 13.22 11.84 10.28 8.33 6.41 4.20 2.07 0.00
2.40 3.67 7.33 10.41 12.78 14.68 15.85 16.48 16.51 16.19 15.53 14.49 13.22 11.78 9.95 8.15 6.05 4.02 2.03 0.00
2.50 3.78 7.54 10.72 13.19 15.20 16.49 17.22 17.36 17.15 16.61 15.68 14.52 13.19 11.47 9.77 7.78 5.86 3.96 2.01 0.00
2.60 3.88 7.73 11.01 13.58 15.69 17.08 17.92 18.16 18.05 17.61 16.78 15.73 14.50 12.88 11.29 9.40 7.58 5.78 3.92 1.98 0.00
2.70 3.97 7.90 11.28 13.94 16.14 17.62 18.55 18.89 18.87 18.53 17.80 16.83 15.70 14.18 12.68 10.88 9.15 7.44 5.67 3.82 1.90 0.00
2.804.05 8.06 11.51 14.25 16.53 18.10 19.11 19.53 19.60 19.34 18.69 17.81 16.76 15.32 13.90 12.19 10.54 8.92 7.23 5.44 3.61 1.76 0.00
2.904.11 8.18 11.69 14.49 16.83 18.46 19.54 20.02 20.15 19.95 19.37 18.55 17.57 16.19 14.83 13.18 11.60 10.04 8.41 6.69 4.91 3.12 1.400.00
3.004.17 8.29 11.85 14.71 17.11 18.79 19.93 20.47 20.66 20.52 19.99 19.24 18.31 16.99 15.69 14.10 12.58 11.07 9.50 7.84 6.12 4.38 2.71 1.35 0.00
3.20 4.25 8.44 12.09 15.02 17.50 19.27 20.49 21.11 21.38 21.32 20.88 20.21 19.36 18.13 16.92 15.41 13.96 12.54 11.06 9.48 7.83 6.18 4.58 3.30 2.01 0.00
3.40 4.30 8.54 12.24 15.23 17.77 19.59 20.86 21.54 21.87 21.86 21.47 20.86 20.07 18.89 17.73 16.28 14.89 13.52 12.09 10.57 8.98 7.38 5.83 4.60 3.37 1.45 0.00
3.60 4.35 8.63 12.36 15.40 17.97 19.84 21.15 21.87 22.25 22.29 21.94 21.37 20.63 19.49 18.38 16.97 15.62 14.30 12.91 11.43 9.89 8.33 6.83 5.64 4.45 2.61 1.24 0.00
3.80 4.38 8.69 12.47 15.53 18.15 20.05 21.40 22.15 22.56 22.64 22.33 21.80 21.09 19.99 18.91 17.54 16.23 14.95 13.59 12.15 10.64 9.12 7.65 6.50 5.34 3.58 2.27 1.10 0.00
4.00 4.40 8.74 12.53 15.62 18.26 20.18 21.56 22.34 22.77 22.87 22.59 22.08 21.39 20.32 19.27 17.92 16.63 15.37 14.05 12.63 11.14 9.64 8.20 7.08 5.94 4.22 2.97 1.84 0.78 0.00 ('"l
6.90 4.44 8.81 12.63 15.76 18.43 20.39 21.80 22.61 23.08 23.22 22.97 22.50 21.85 20.81 19.79 18.48 17.23 16.01 14.71 13.33 11.88 10.42 9.01 7.92 6.82 5.17 3.99 2.93 1.94 1.23 0.00 DO
..,:r;-
aT = 300 K; A = 1. B = 1; solar concentration = C = 1. .
-.I
Semiconductors 231
T/ = (31.414/135.3)100% = 23.22%
This value is listed in Table 7.2 at the bottom of the column for Eg = 1.5.
The same value is also listed in Table 7.3 for a single cell. Table 7.2 has
been obtained by the foregoing procedure with Ise of equation (7.72)
calculated from Table 7.1. Each column after the first column gives the
efficiency of each cell for the impinging energy E on the first column. For
example, the efficiency of a cell with Eg = 1.5, having a filter of E ~ 2, is
10.06 as this filter would absorb radiation greater than 2.0 eV. A trial-and-
error method for two tandem cells on a computer would yield E = 2.0, and
Eg = 1.2 for the highest efficiency, E being the band-gap of the semiconduc-
tor which acts as a filter for the cell with Eg = 1.2, the latter receiving
radiation equal to or smaller than 2.0 eV. The efficiency for the filter cell is
19.79 from Table 7.2; moving horizontally to the left from the point Eg = 2.0
at the top of the column (efficiency 0.000), and reaching Eg = 1.2, we obtain
12.69; therefore, the total efficiency of two tandem cells is 19.79 + 12.69 =
32.48, as listed in Table 7.3. Any other combination of cells would yield a
lower efficiency.
Table 7.3 shows how sensitive the overall efficiency is to variations in
the selected band-gaps and the number of cells. The contribution to the
efficiency is dramatic up to six cells, after which it levels off as shown in
Table 7.3 and represented in Fig. 7.19.
Increasing the operating temperature of the cells decreases their efficien-
cies as shown in Fig. 7.19. The contribution to the efficiency from the smaller
Eg semiconductors decreases more rapidly with increasing temperature than
for larger Eg semiconductors, because the reverse saturation current, Irs.
No. of cells 2 3 3 4 5
Eg(eV) 1.5 2.0,1.2 2.3, 1.6, 2.4,1.6, 2.5, 1.8, 2.6, 2.0, 1.6,
1.0 1.0 1.3,0.9 1.2,0.8
Efficiency 23.22 32.48 37.42 37.39 40.45 42.45
No. of cells 6 6 6 7
Eg(eV) 2.7,2.3,1.9, 2.8, 2.4, 2.0, 2.6, 2.2, 1.8, 2.9,2.5,2.1, 1.8,
1.5, 1.1, 0.8 1.6, 1.2,0.8 1.4, 1.0,0.8 1.5, 1.1,0.8
Efficiency 43.82 43.67 43.64 44.86
No. of cells 7 7 13 26
Eg(eV) 3.0,2.6,2.2, 2.8, 2.4, 2.1, 0.8 to 3.0 in 0.7 to 3.0 in
1.8, 1.4, 1.0, 1.8, 1.5, 1.2, intervals of intervals of 0.1,
0.8 0.8 0.2,3.4 3.2,3.4
Efficiency 44.70 44.67 48.13 50.09
l
60
~
50
I 0-
0
/~ J0
- h~ ! ...-----0
40
I /,,/,-0
u o
J
0.
'itl30
J/I
-<>=
0.0-
I Sun
A 200 K
20
0/
¢
o
o
¢
300 K
400 K
500 K
100
4 8 12 16 20 24
No. OF CELLS
Figure 7.19. Variation of efficiency, T/ (%), with number of cells at various temperatures. Solar
intensity is one sun.
70
6/6
60
I ~O- 0
-u
50
!IO~O_
1:/°.___
0-
r:-
f/
40 0-
6
500 Suns
iff
200 K
30 0 300 K
~
o 400 K
0 500 K
20
0 4 8 12 16 20 24
No. OF CELLS
Figure 7.20. Variation of efficiency with number of cells at various temperatures. Solar intensity
is 500 suns.
achievable for solar cells on satellites, space laboratories, and space stations.
The efficiency of a Carnot engine operating between 5750 K and 200 K is
[(5750 - 200)/5750]100% = 96.5%, indicating that tandem solar cells are
not far from ideally achievable efficiencies. At solar concentrations greater
than 1000 suns and fewer than a dozen tandem cells, the minority carrier
concentration becomes comparable to the majority carrier concentration
and equation (7.58) and the derived relations are no longer valid.
The foregoing results are for AMO solar radiation. If the AMI or AM2
spectrum had been used (AMI being the radiation when the sun is vertically
above the horizontal cells at sea level), the results would have been similar
in the sense that high solar concentration and low cell temperature would
lead to efficiencies approaching 70%. However, the range of useful band-
gaps on the high band-gap side would have been cut off at smaller values
of Eg and the combination of Eg values required for optimized efficiency
would be different from that appropriate to the AMO spectrum, but the
234 Chapter 7
/-
7°r-"--II--TI--~I-r-I~I~~I==]I==I=~==~1
60 I-
I~o-
~ 50- !IO~___--~_
-
1000 suns
30~r
6. 200 K
-
o 300 K
o 400 K
<> 500 K
20 <> I 1 I 1 I L L J I 1 I
o 4 8 12 16 20 24
No. OF CELLS
Figure 7.21. Variation of efficiency with number of cells at various temperatures. Solar intensity
is 1000 suns.
No. of cells
2 3 6 12
1 sun
200K 25.3 35.4 40.6 47.6 51.4
300K 19.2 26.7 30.6 35.3 37.9
400K 14.4 19.7 22.7 25.7 27.3
100 suns
200K 30.1 42.0 48.3 56.2 60.0
300 K 25.2 35.1 40.1 46.7 49.9
400K 21.1 29.2 33.3 38.3 40.6
500 suns
200 K 32.1 42.5 51.1 59.5 63.2
300 K 27.7 38.5 44.0 51.0 54.2
400K 24.0 33.2 37.8 43.5 46.2
Semiconductors 235
400 500
~------~~------~~------~~-'70
60
,
",
"", , 50
- ,, -
,,
,
U U
Q. Q.
~50
~
" , 1000 suns 40 ~
40
"
, ''(6
t ---t 100 suns
"
celis,
...!OO suns
30
"" " , """ ,
' ..... " ..... ,
30 20
200 300 400 500
TEMPERATURE. K
Figure 7.22. Variation of efficiency with temperature. For clarity, results for 500 suns are not
shown.
References
Engel-Brewer Theories
Promotion Promotion
Con- energy. Crystal Con- energy. Crystal
Element figuration keal/mole structure" Element figuration keal/mole strueture
aGround state has zero promotion energy; other values are the lowest energies for each configuration as given by Brewer.
·Stable erystal structure at ambient pressure; see page 318.
Engel-Brewer Theories 239
Basic Principles
The basic principle of the E-B theory lies in correlating the unpaired
s- and p-electrons with the crystal structures. An element crystallizes in
body-centered cubic (bcc) structure when it has one outer unpaired electron.
The alkali elements and Cr and Mo have one outermost s-electron each,
and they crystallize in bcc structure. In Cr and Mo, the ground state is dSs;
the next lowest level with an unpaired s-electron and one p-electron, d 4 sp,
requires 71 and 80 kcal/mole for promoting one d-electron to the p-orbital
for Cr and Mo, respectively. These are considered to be high promotion
energies, particularly because of no change in the total number of unpaired
d- plus s-electrons in d 5 s and in d 4 sp configurations; hence, Cr and Mo
cannot crystallize in any other structure than bcc, in agreement with observa-
tions.
For magnesium, the picture is different, because its ground state outside
the closed shell is 3s 2 , or briefly S2 (Table A.I); this state is not suitable
for bonding because S2 is a paired group of two electrons, unavailable for
240 Appendix A
The foregoing considerations lead to the following E-B rules. (1) The
crystal structure depends on the total number of unpaired (s + p) electrons;
the crystal structures and the numbers of unpaired (s + p) electrons are as
follows:
• bcc: 1 electron
• cph: 2 electrons
• fcc: 3 electrons
• diamond structure: 4 electrons
We shall see later that these integral numbers of electrons per atom will be
modified to cover a range of fractional numbers for each crystal structure
encountered in some metals and in all alloys. For brevity, Brewer used I,
II, and III for bcc, cph, and fcc, respectively. (2) The bonding energy
depends on the number of unpaired do, So, and p-electrons available for
bonding. (3) The electronic structure in solid phases is close to that in the
gaseous atomic element promoted as necessary to yield the appropriate
numbers of (s + p) electrons. The electrons are promoted with as small
amounts of energy as possible for a given crystal structure with as many
bonding do, So, and p-electrons as feasible. The promotion is mainly justified
(i) on the basis of appropriate numbers of (s + p) electrons and (ii) on the
basis of maximum number of unpaired electrons obtained by reasonable
promotion energies.
The basic assumption that the excited atoms in gaseous state also exist
in condensed state cannot be proven by existing experimental techniques.
The cph crystal structure in Li and Na at low temperatures cannot be
explained by this theory because no reasonable promotion energy can
provide two (s + p) electrons for cph structure; to do so would require
promoting one electron from the closed noble gas shell to the p-orbital
beyond the outer s-orbital. Likewise, for Ca and Sr, fcc structure requiring
three (s + p) electrons cannot be reconciled with E-B rules. Despite these
and other shortcomings, the E-B concepts have achieved remarkable success
in correlating the behavior of metals and alloys, and in estimating the
compositional ranges of alloy phases as will be seen later.
In its refined form, the E-B theory considers that the structures of all
the elements, and particularly the transition elements and their alloys, will
tolerate some deviation from the integral numbers of 1, 2, and 3, (s + p)
electrons for bcc, cph, and fcc structures, respectively. The postulated values
242 Appendix A
Transition Elements
energies in Table A.l for d 5 sp are not too large, considering the fact that
there are seven bonding or unpaired electrons for d 5 sp instead of five for
d 6 s; hence, cph structure is favored, in agreement with observations. For
Fe, d 7 sand d 6 sp configurations have acceptable promotion energies but
cph structure corresponding to d 6 sp does not exist at ambient pressures.
For Ru, d 7 s configuration should yield bcc as in Fe, but this form of Ru
does not exist. However, the promotion energy for the d 6 sp configuration
for Ru is rather high, in fact higher than that for Fe, but the corresponding
cph structure is the observed form of Ru at all temperatures, whereas cph
structure is not observed for Fe at ordinary pressures. The bonding of extra
d- and p-electrons offset the high promotion energy of d 5 sp structure because
4d electrons have much greater bonding capability than 3d electrons as
shown in Fig. A.l.
The configuration for the cph form of Co is evidently d 7 sp since the
ground state d 7 S2 is not suitable for bonding. The promotion energy for
d 6sp2 for fcc is high and at high temperatures, solid Co is in fcc structure.
For Rh and Ir, the promotion energy for the d 6 sp2 configuration is high
but promotion to d 6 Sp2 is necessary for the fcc structure that is observed
at all temperatures. The configuration proposed by Brewer for Cu, Ag, and
Au is d 8sp2, i.e., three (s + p) electrons, for which the promotion energy is
fairly high. These metals exist only in fcc structure.
Co V Cr Mn Fe Co Ni Cu Zn
60
0
!l0
e
c
2 40
A ....
u
~
30
Q.
Ci 20
u
.a
10
0
0 234!143 o
No. of unpaired electrons per atom
Figure A.I. Valence· state bonding enthalpy (- energy) in kilocalories per mole of electrons.
Upper curves are for E,p, bonding enthalpies of s· and p·electrons, versus the elements on
upper horizontal axis. Lower curves are bonding enthalpies versus numbers of unpaired
d·electrons on lower horizontal axis. (From Brewer4 with permission.)
Sr Y Zr Nb Mo Tc Ru Rh Pd 49 Cd
60
I I
!5 s,~
"0
E
c
.....
::!
U 40
B ~
Q.
......
"0
30
20
o 2 3 4 4 3 2 o
No. of un~arred 4d electrons ~er atom
8e Le Ht Te W Re as Ir Pt Au Hg
70
"0
.,:LV-
5' 60
c: 6s,p
2
C .
~ 50
i 40
'0
u
"'" 30 oLe
o bee
of 48, the promotion energy, yields Eatomiz = 161 kcaljmole, indicating that
the cph form is about 35 kcalless stable than the bcc form. Ifwe use d 3 .3 SpO.7
configuration, the lower limit of 1.7 (s + p) electrons for cph, the result
would be Ebonding = 52 x 1.7 + 35 x 3.3 = 204, and Eatomiz = 204 - 48 x
0.7 = 170, where 0.7 originates from the fact that 0.7 mole of d 3 sp and
0.3 mole of d 4 s make up d 3 .3 Sp O.7. Again, even with the use of the lower
limit for 1.7 (s + p) electrons, the result does not make the cph form as
246 Appendix A
stable as the bcc form. The foregoing figures indicate that as the number
of unpaired d-electrons decreases from five in each transition series, the
bonding energy per d-electron increases.
Calculated
values in Experimental values
Cr, Mo,
Solute and W in Cr in Mo in W
Tc 50 50
Re 50 48.5 42 37
Ru 25 34 30 23
Os 25 31 19 18.5
Rh 16.7 -15 -20 -6
Ir 16.7 -12 16 -10
Pd 12.5 -5 6.5 -5
Pt 12.5 -10 15 4
ment between the calculated and experimental values becomes larger as the
difference between the group numbers of the solvents and solutes increases.
The transition elements with fewer than five d-electrons act as a sink
for the d-electrons of the elements having more than five d-electrons. In
addition, the (s + p) electron concentration may be lower than 1.5 so that
the structure for bcc alloys of Hf( d 3 s )-Ru( d 7 s) is d 4 .S s, for Hf as the solvent,
corresponding to 62.5 at.% Hf as calculated from 5.5 = 4x + 8(1 - x). For
Ru as the solvent, the structure is dS'ss, and 6.5 = 4x + 8(1 - x) yields
37.5 at. % Hf. Calculation of the appropriate numbers of (s + p) electrons
is discussed in detail in Ref. 4. If for a known binary system the electron
concentration and configuration are fitted, then for a nearby system the
same concentration and configuration can be used to estimate the unknown
solubility. In essence, the known maximum solubilities may dictate the
electronic structure, and the unknown maximum solubilities may be based
on the rules as adjusted according to the rules derived from similar binary
systems if they have been established by experiment.
The Laves phases, for which the ideal compositional range is AB2 to
AB s, generally having cubic, hexagonal, and complex structures, are formed
not only on the basis of electron to atom ratio, but also on the basis of
atomic radius ratio of A to B as discussed elsewhere in detail. 20,23,2S No
definite (s + p) electron concentration has been assigned to these phases
by Brewer. Some of these, such as LaNi s, are useful and interesting media
for hydrogen storage as was discussed in Chapter 6.
248 Appendix A
Other Phases
Several other phases exist in addition to I, II, and III, within the
electron concentration range up to 3.5. They follow, with increasing electron
concentration, the structures indicated by
Cubic Cr3Si
The cubic Cr3Si structure (Strukturbericht A15)* is such that each Si
atom is surrounded by 12 Cr atoms without Si-Si bonds. In addition to Cr,
the elements of IVA to VIA may be substituted for Cr, and the elements
of VIlA, VIllA to VB may be substituted for Si. These phases are ordered
phases in which atomic radii should not differ by more than 15%. Excep-
tional A15 phases are Zn 4 Sn and MoTc. In Cr3Si-type phases, the (s + p)
electron concentration is 1.5 to 1.75 according to Brewer when the d
configuration is chosen properly. Thus, for Cr3Si, there are 7 (s + p) elec-
trons per 4 atoms or 1.75 (s + p) electrons per atom. For Cr3Pt, there are
6/4 = 1.5 electrons per atom, assuming that Pt does not contribute any d
atoms, but if Pt retains d 6 structure, or yields only one d-electron, then
1.75 (s + p) electron concentration still prevails.
*These structures are discussed in standard texts. l7 • IS See also Nevitt. 20 AI5 and 0' phases
compete with bcc and cph phases when atomic sizes are favorable.
250 Appendix A
a Phase
The a phase occurs at the atomic radius ratios ofO.93 to 1.15 according
to Nevitt. 20 It is formed by quasi-hexagonal layers of atoms, with a tetragonal
unit cell. The (s + p) concentration is generally 1.2 to 1.9, overlapping bcc
and cph structures. The range of (d + s + p) concentration is between 5.6
and 7.7 electrons per atom, with a very large number of phases occurring
in the vicinity of 6.6 electrons. The assignment of nonintegral configurations,
such as dS'sspo.s for Mn, d 6 . Sspo.s for Fe, retains the range of (s + p)
recommended by Brewer. For example, Cr-Mn, Mn-Fe, Mn-Re, Mn-Tc,
Mo-Tc, Re- W binary systems form a phases.
X or a-Mn Phase
The remaining phases, P, }.t, R, and X, are very closely related to the
a phase in their crystal structures. The P, }.t, and R phases are labeled as
the }.t phase by Brewer in his multicomponent diagrams. The electron
concentrations for these phases are determined from the phase diagrams
where they occur. The X phase for Mn is at 7 (d + s + p) electrons per
atom, and for the binary alloys, the range is 6.3 to 7.0 (d + s + p) electrons
per atom. Since all the transition elements forming X phases assume the d S
configuration, the (s + p) concentration range is from 1.3 to 2.0, overlapping
the a and cph phases.
Mn (Fe) Co Ni Cu
Ru Rh Pd
Os Ir Pt Au
This is explained on the basis that the elements to the right, e.g., Ni, can
contribute electrons from the d-orbital to the s- and p-orbitals, and thus
stabilize the fcc (d X Sp2) structure.
Engel-Brewer Theories 251
Effect of Pressure
Brewer postulated that increased pressure will allow the nuclei to move
closer to permit the d-orbitals to overlap better and stabilize the crystal
structure corresponding to the largest number of bonding d-electrons. For
the metals having fewer than five d-electrons per atom, e.g., Ti in cph form
(d 2sp), increasing pressure would favor d 3s corresponding to the bcc form
of Ti, in agreement with experimental results. Similar results have been
verified for Zr and Hf. In general, for an element with fcc and d 3 Sp2 structure,
increasing pressure would favor d 3 sp2 ~ d 4 sp ~ d 5 s, or fcc ~ cph ~ bcc.
For d 4 sp2, increasing pressure would stabilize the structures corresponding
to d 5 sp which has the largest number of unpaired d-electrons. The elements
that have bcc structure at ordinary pressures, V, Nb, Ta, Cr, Mo, and W,
remain bcc up to the highest pressures investigated, because (1) for V, Nb,
and Ta, d 4 s structure cannot be changed to d 5 , and (2) for Cr, Mo, and
W, the d 5 s structure already has the optimum number of d-electrons. On
the other hand, for elements having more than five d-electrons per atom,
e.g., d 7s, increasing pressure would favor d 7s ~ d 6sp ~ d 5sp2 ~ , or bcc ~
cph ~ fcc. In summary, increasing pressure would tend to stabilize the
structures toward the maximum number of unpaired, or bonding d-elec-
trons, i.e., toward d m where m tends as close to five as possible. 3 ,15,21
Multicomponent Diagrams
Re (7J
Os (8)
W (6)
Ir (9)
Pt (10)
100 80 60 40 20 0
AT,% IN
Figure A,2. Optimum phase boundaries in alloys of W with Re, Os, Ir, and Pt, Numbers in
parentheses are (d + s) electrons. I = bee, II = cph, and III = fcc. (Courtesy of L. Brewer;
see also Reference 3.)
series of Group IVA, VA, and VIA elements alloyed with the remaining
groups of elements to the right,
One such diagram is shown in Fig. A.2 for the alloys of W, where the
horizontal axis represents at. % W. The vertical axis is the total outer electron
concentration, which increases from 7 for Re to 10 for PC A horizontal line
from Os to the left represents the binary W-Os system for which the
maximum compositional ranges of existence for various phases occur at
the intersections of this line with the boundaries for the single phases. The
areas marked I, II, and III are for bee, cph, and fcc, respectively; u and X
represent the phases designated by these letters.
Such diagrams are not isothermal since the maximum range of composi-
tion for one phase is generally not at the same temperature as that for
another phase. In the W-Os system, the range of composition for II is from
o to 53 at.% W, The upper horizontal line is for W-Re and the lower
horizontal line is for W-Pt. The range of composition for II in W-Re is 0
to 20 at. % W but this range refers to a different temperature than the
preceding range for II in W-Os. The horizontal line that can be drawn
starting halfway between Os and Ir refers to an electron concentration of
8.5 at 0 at. % Wand this line is for the ternary alloy of W-Os-Ir in which
the atomic ratio of Os to Ir is unity, This same line is also for the ternary
W-Re-Pt, and pentanary W-Re-Os-Ir-Pt system, for all of which the
electron concentration is 8.5 at 0 at. % W. The boundaries in this diagram
are largely obtained from the binary diagrams, and their validity for the
ternary and multicomponent systems has not yet been tested. Where experi-
mental data were absent, E- B rules were used to obtain the phase boundaries
in Fig. A.2. Since each point on a phase field refers to a different temperature,
Engel-Brewer Theories 253
it is not possible to draw the tie lines joining one point on one phase
boundary to another point on another phase boundary. It is also not possible
to designate the areas between the single-phase fields as the multiphase
fields for the ternary and multi component alloys.
Concluding Remarks
The basic E-B postulate that bcc, cph, and fcc correspond respectively
to one, two, and three (s + p) electrons is not applicable to cph forms of
Li and Na, and fcc forms of Ca and Sr. These metals have relatively
simple electronic structures; hence, the inability of E- B postulates to account
for these structures is fundamental. Nevertheless, the numbers of structures
and transition phenomena successfully explained by these theories are quite
impressive.
In addition to the electronic configuration, Brewer considered addi-
tional important factors, i.e., the size factor, and attractive and repulsive
force factors [see equation (B.19)]. All these factors, considered judiciously,
yield the appropriate solubility or stability limits for various structures and
their energies as discussed in Brewer's publications.
There are many aspects of the E-B approach that disagree with the
methods of Friedel,I3,14 Nevitt/o and Miedema et al. 22 It is nevertheless
generally felt that the E-B theories and postulates represent the most
comprehensive existing treatment of solubility ranges for various phases in
binary and multi component systems. The theories and schemes to be
developed in the future will likely take advantage of the E-B theories, along
with other theories, correlations, and postulates. The immediate problem
is to test the validity of the diagrams by obtaining experimental data on a
number of binary and multi component alloys of the transition elements for
which E-B theories claim the greatest success.
Additional recent publications by Brewer et al. 23 - 27 are recommended
to the reader interested in pursuing this topic further.
References
Introduction
A A A A
A A A A
A A A A
+
A A A A
a
b c
Figure B.1. Wigner-Seitz cells for two-dimensional metals and alloys. (a) Cells for pure metals
A and B; they are mixed to form a mechanical mixture as shown in (b). The cell sizes (areas)
are identical in (a) and (b). (c) The mechanical mixture becomes an alloy phase and phase
boundaries move to smooth out the discontinuity of electron density in (b). In this figure
n~, < n~,; hence, the cell boundaries in (b) move to make the B cells larger as shown in (c).
Volume effects are related to the change transfer; hence, to the differences in the elec-
tronegativities.
n ws
l/3
<P V2 /3 n~/ <P V2 /3
aElectronegativity, cP, is in volts; electron density, nw>o is in electrons per (0.529 A)3; molar volume, V, is
in cm 3 at room temperature. For values related to compressibility, ¢, see, e.g., V. S. Fromenko and G. W.
Samsonov, Handbook a/Thermionic Properties, Plenum Press, New York (1966), and D. E. Eastman, Phys.
Rev. B 2, I (1970); for ¢, V, and nw>o see K. A. Gschneidner, Solid State Phys. 16,275 (1964), and V. L.
Moruzzi, J. F. Janak, and A. R. Williams, Calculated Electronic Properties 0/ Metals, Pergamon Press,
Elmsford, New York (1978).
258 Appendix B
closer to each other, but it must be emphasized that the adjustments in nws
are essential for the usefulness of the proposed semi empirical method. The
difference .:lnws is not known exactly, but it is assumed that .:In!!s3 for the
alloy is equal to [(n~s)1/3 - (n!s)1/3] for the pure components A and B.
The second effect is due to the electronegativity of elements, i.e., the
potential of electrons in metals, 4>. The more electronegative elements tend
to attract the electrons from the less electronegative elements upon alloying.
The electronegativity scales differ depending on the types of measurement,
such as the standard electrode potential method, ionization of monatomic
gases, and formation of halogen bonds. 7- 9 Since the electronegativity of
electrons in pure component A is different from that in pure component B,
the electrons tend to spend more time about the more electronegative atom
after alloying. Therefore, the potential of electrons is lowered upon alloying,
and consequently .:lH is lowered by -P(4)A - 4>0)2 = -p(.:l4>f where pis
an empirical constant. A term of this type has been used previously by
Pauling.7 Again, .:l4> for an alloy is not known, but it is assumed that
.:l4> = 4>A - 4>0 where 4>A and 4>0 refer to pure A and pure B, respectively.
The initial values of 4>, as used by Miedema et al., were the experimentally
measured values in a number of compilations cited in Table B.t. However,
these values have been adjusted slightly for empirical fitting of the experi-
mental values for .:lH. The adjustments are small in most cases,I-5 e.g., the
adjusted values for Ti and Zr are approximately 15% smaller than the
experimental values, but the remaining values for the transition elements
differ less than 15 % .
The contribution to .:lH resulting from the changes in electron con-
centration and electronegativity for certain classes of alloys is expressed by
.:lH = f(XA, V)g(XA' nws )[q(.:ln!!s3f - p(.:l4»2] (B.1)
where f(XA, V) is dependent on XA and volume V, and g(XA' nws) on XA
and nws as will be seen later, and p and q are empirical constants. The
next task is the determination of p and q.
Empirical Coefficients
The determination of empirical coefficients p and q requires experi-
mental data on .:lH, and in the absence of such data, binary phase diagrams
from which the algebraic sign of .:lH can be estimated. Initially required
information is whether .:lH is positive or negative. In the absence of
experimental data for .:lH, the sign of .:lH was obtained by the following
established empirical rules: (1) .:lH is negative for all the binary alloys in
which intermetallic compounds or ordered phases have been observed, and
(2) .:lH is positive for all the binary alloys when there are no intermetallic
Estimation of Enthalpy of Alloy Formation 259
n w,
'/3 4> V2 !.' n!:/ 4> V2 /3
"For units, see footnote to Table B.1. Values of V 2 !> for H, Si, Ga, Ge, C, N, As, Sb, Bi are for hypothetical,
close-packed, metallic structures. The numbers in parentheses indicate the valence in the group headed
by that element.
compounds and the mutual solubilities are less than 10 at. %. For all the
binary alloys of the elements in Tables B.1 and B.2, except the alloys of
(Class a element + Class (3 element) in Table B.3, equation (B.t) is valid.
For I1H = 0, equation (B.t) requires that a plot of 11¢ versus I1n 1/3 should
leave positive I1H values, indicated by +, on one side of the straight line
passing through the origin, while negative values of I1H, indicated by
Be B C N
lIA Transition metals, 1lIA- VIllA IB 0.4 1.9 2.1 2.3
Ca Sc Ti V Cr Mn Fe Co Ni Cu Mg AI Si
0.4 0.7 1.0 1.0 1.0 1.0 1.0 1.0 1.0 0.3 0.4 1.9 2.1 -
Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Zn Ga Ge As
0.4 0.7 1.0 1.0 1.0 1.0 1.0 1.0 1.0 0.15 1.4 1.9 2.1 2.3
Ba La Hf Ta W Re Os Ir Pt Au Cd In Sn Sb
0.4 0.7 1.0 1.0 1.0 1.0 1.0 1.0 1.0 0.3 1.4 1.9 2.1 2.3
Th U Pu Hg n Pb Bi
0.7 1.0 1.0 1.4 1.9 2.1 2.3
"To obtain r/ p for a binary solid alloy formed by one Class a and one Class f3 element, multiply the
numbers under the elements. For liquid alloys rip is reduced by a faclor of 0.73. For all other alloys.
r/ p ; O. Class f3 elements. except those in the first column. contain p-type outer electrons.
260 Appendix B
should be on the opposite side. A mapping of this type for solid alloys is
shown in Fig. 8.2, wherein the slope of the straight line yields
tlrf> 12 (B.2)
1tln!fs3 = : = 9.4
For liquid alloys of the same types of components, a similar mapping exists,4
yielding again q/p = 9.4. Not all the alloys are on the correct side of the
straight line in Fig. B.2, and in a similar figure for liquid alloys, but there
are very few such alloys.
The corresponding analysis of the results on tlH for (Class a + Class
(3) alloys shows that an additional large term, r, is required in equation (B. 1):
(V!
-- -:~
20
Zn,Cd,Hg AI,Ga,In, Tl
i 15
M
Ivl
10
05
°0 02
- 0 02 04 A 'I) 0.6
nws---+
5n,Pb AS,Sb,Bi
t
AQI
15
Ivl
10
0.5
-
02 04 06 o 02 04 06
An'3_ An "3
ws ws
Figure B.3. Mapping of f1cf> in volts and f1n!!; in (density units)!/3 for positive, +, and negative,
-, values of f1H according to equation (B.3), Solid curve separates + and - for binary solid
alloys of (Class a + Class (3), Class f3 elements are listed in each panel. (From Miedema and
de Chatel 5 with permission.)
262 Appendix B
the value of rip. Further analysis has shown that qlp is again 9.4; hence,
the straight lines with this slope that can be drawn through the origin are
asymptotes of the hyperbola at very large values of arl>. The results for rip
for solid alloys are obtained from the scheme in Table B.3, i.e., by mUltiplying
the number under the Class a element with the number under the Class 13
element; e.g., for Ag-Zn, rip = 0.15 x 1.4 = 0.21. The values of rip for
liquid alloys should be 0.73 x rip = r(l}1 p(l} where r(l} and p(l) refer to
the liquid alloys. Comparison of the four plots in Fig. B.3 shows that rip
increases with increasing number of outer p-electrons of Class 13 elements.
Calculation of ilH
(B.4)
where VA and VB are the molar volumes of the pure elements and x~ =
1 - x~. Since the cell volumes change upon alloying, VA for pure A, and
VB for pure B cannot be the same as VA(alloy} and VB(alloy} in the alloy,
Alloy type p
"Transition elements (trans.) are shown in Table B.3; noble elements (nob.)
are Cu, Ag, and Au; others are designated as remaining elements (remn.).
Listed values, with q" n:!;, and y2/3 in Tables B.I and B.2, yield tlH in
kJ/g-atom (4.184 J = 1 cal).
Estimation of Enthalpy of Alloy Formation 263
but the use of VA and VB for the pure elements in equation (B.4) is assumed
to be valid for a large number of alloys. It is further assumed thatf(xA, V)
follows the zeroth approximation to the regular solutions with x~ substituted
for Xi:
f(XA, V) == f(x~, x~) = x~x~ (B.5)
This equation justifies the use of V;:3 in equation (B.7), because ilHA(XA ~
0)/ ilHB(XB ~ 0) is roughly proportional to (VAi VB)2/3 in the majority of
cases considered by Miedema et al. although there are notable exceptions.
Comparison of equations (B.7) and (B.8) shows that
(B.9)
(B.IO)
264 Appendix B
This equation is used only for large values of (CPA - CPB) and for alkaline
and alkaline earth metals and small-atom stongly electronegative elements
such as B(boron), C, H, and N. For example, VA(pure) = 5.5 and
VA (alloy) = 4.98 for A = Li, B = AI, and X B = 0.5. In contrast, VA (pure) =
4.8 and VA (alloy) = 4.68 for A = Ti, B = Fe, XB = 0.5, and UA = UB = 0.04;
the resulting difference in VA(pure) and VA (alloy) is considered to be
negligible for Fe-Ti. A few examples of calculated and experimental values
of aH are listed in Table B.5 for solid and liquid alloy phases of equiatomic
composition. For example, aH for liquid equiatomic Au-Zn alloy can be
Table B.S. Calculated and Experimental Values of tlH of Formation for 1 g-atom
of Equiatomic Solid and Liquid Alloy Phases. Results Are for Long-Range
Disordered Alloys, Except as Explained in Footnote a
kJ/g-atom kJ/g-atom
"Ordered structures. Structure of AI-Au is uncertain, but assumed to be ordered in this calculation.
Magnitudes of !l.H cannot always predict the existence or the absence of long-range order.
bResults for liquid AI-Au, Au-Zn, Cu-Fe, Hg-Pb, and K-Pb and all those for solid phases were recomputed
for this book, and experimental data were obtained from R. Hultgren, P. D. Desai, D. T. Hawkins, M.
Gleiser, and K. K. Kelley, Selected Values of the Thermodynamic Properties of Binary Alloys, ASM, Metals
Park, Ohio (1973), and from O. Kubaschewski and C. B. Alcock, Metallurgical Thermochemistry, Fifth
Edition, Pergamon Press, Elmsford, New York (1979).
Estimation of Enthalpy of Alloy Formation 265
Ordered Phases
The value of x~ increases for ordered structures because a much larger
fraction of the surface due to B is in contact with A. A purely empirical
relation expressing this fraction, denoted by IB, is
Note that for disordered solutions IB = x~, but equation (B.ll) must replace
x~ in equation (B.7). Substitution of equations (B.8) and (B.ll) in equation
(B.7) yields another form for I1H, i.e.,
(B.12)
where XA may assume any value for I1H; however, I1HA(xA ~ 0) after the
second equality refers to X A ~ O. For the ordered Cu-Zn alloy of equiatomic
composition at 700 K, I1H(exp) = -12.1 kJ/g-atom, whereas I1H(calc) =
-10.5 kJ/g-atom from equation (B.12), in contrast with I1H(exp) =
-7.7 kJ/g-atom and I1H(calc) = -5 kJ/g-atom for the disordered phase as
listed in Table B.5.
The average number o"solid ordered phases or compounds in a binary
system correlates roughly with the magnitude of I1H; thus, there is often
one compound when I1H is between -4 and -10; three compounds when
I1H is between -20 and -40; five compounds when I1H is more negative
than -75, all I1H in kJ/g-atom of A + B. Here the model proposed by
Miedema et al. cannot estimate I1H unless experimental results indicate
that the phases are either ordered or long-range disordered.
266 Appendix B
tJ.Hph , kJ/g-atom
Table B.6. Calculated and Experimental Values of AHo for Selected Ordered Binary
Phases of H, C, N, Si, and Ge a
• AHo is in kJ/g-atom of alloy formed from the elements in their standard states (Refs. 3, 10, II).
Alloys of Hydrogen
alkaline metal, r / p is not zero for H but equal to 3.9 for the [H( cp metal) +
transition element] alloys. For example, r / p = 3.9 x 0.7 for the La-H alloys.
The calculations of V;';3(alloy) and fs for H(cp metal) are different
from those for other elements because these parameters are calculated by
a different procedure. The analytical form adopted for fs is identical with
equation (B.11) but equation (B.10) is rewritten as
(B.13)
The experimental value is -42 kJ/g-atom, in very close agreement with the
calculated value. The increase in the volume of Ti upon absorption of
hydrogen is calculated from
where V(alloy) is the volume of the alloy, and the partial molar volume V;
is assumed to be the same as Vj(alloy, II). The volume of pure 0.33Ti is
3.47 cm 3; hence, 11 V = 4.73 - 3.47 = 1.26 which is the increase in the volume
of Ti after the dissolution of hydrogen, and the calculated value, 11 VI V =
1.26/3.47 = 0.36 or 36%, agrees well with the experimental value, 31 %.
Similar calculations for LaO.36HO.64 yield x~ = 0.705; x~ = 0.295;
V;.(3(alloy) = 7.727; Vi(3(alloy) = 1.826; IB = 0.3966; I1H = -119 kJ/g-
atom; and I1HO(from H 2) = -55 kJ/g-atom alloy, in fair agreement with
the experimental value of -67 kJ1g-atom. The calculated value of 11 VI V is
8%, in poor agreement with the experimental value l l of 19%. The results
for a number of binary hydrides selected from Bouten and Miedema l l are
listed in Table B.6.
Ternary Hydrides
much less stable hydride than A. The generalized equation for estimation
of dH;(ABnHv+w) by reaction of ABn with H 2 (g), as indicated by the
subscript g on dH, is as follows ll :
+ w)H 2 + ABn = ABnHv+w
O.5(v
dH;(ABnHv+w) = dH;(AH v) + dH;(BnHw) (B.15)
- (1 - F) dH(AB n)
where v, w, and F depend on stoichiometric coefficient n and on the atomic
radius of A, i.e., whether A is one of the first or second group of metals as
shown in the first column of Table B.7. The empirical term F is a measure
of bonding between A and B atoms in the ternary hydride. Thus, when n
is greater than 5, and v is sufficiently high, H atoms completely surround
A and the bonding between A and B is greatly weakened so that F = 0; in
this case, equation (B.15) refers to the following reactions:
This equality is possible only if dHH _H for the last reaction is zero.
"The two groups of metals have considerably differing atomic sizes in aUoyed states.
Estimation of Enthalpy of Alloy Formation 271
reactant, e.g.,
Hz(g) + (l/3)LaNi s = (l/3)LaNisH6 (B.17)
For PH,=l, LlGo=RT In PH,=O, and (l/3)[LlHg (LaNi s H6)]=
LlH~(B.17)= T LlS~(B.17) for this reaction as indicated by equation (B.17).
Examination of the available data \Z for a number of hydrides indicates that
LlS~ = -125 ± 25 J/mole Hz-K for reaction (B.17) and other similar reac-
tions, each with a single metal at ambient temperatures, and this value is
very close to the negative value of the standard entropy, So, of H 2 (g),
indicating that the standard entropies of LaNis and LaNis H6 are not greatly
different from each other. This value of LlS~(B.17) corresponds to
T LlS;(B.16)=LlH;(B.16)=-37.5±7.5 kJ/mole Hz. Therefore, for the
binary and ternary hydrides PH,> 1 atm for LlH;(B.17»-37.5 kJ/mole
Hz, and PH ,<l atm for LlH;(B.17)<-37.5 kJ/mole Hz, i.e., the more
exothermic hydrides are more stable as expected. In general, the plateau
pressure or the equilibrium pressure PH" which is the pressure at which
two solid phases LaNi s and LaNisH6 coexist with Hz, does not exactly refer
to reaction (B.17), because the LaNi s phase is not pure, i.e., it dissolves a
small amount of hydrogen, and the number of H atoms in LaNisH6 decreases
with increasing temperature. Nevertheless, these effects for many hydrides
are not too great, and from a knowledge of LlH;(B.17) and LlS;(B.17)=
-125 J/mole Hz-K, it is possible to estimate the equilibrium pressure PH,
for various binary and ternary hydrides. A more exact calculation is possible
when PH, at one temperature, T\, and LlH; (B.17) for the reaction are known
from which LlS;(B.17) can be calculated from LlGo=RT In PH,=
LlH;(B.17)-T\ LlS;(B.17) and then the resulting values of LlH;(B.17) and
LlS; (B.17) can be used as constants to calculate PH, at higher and lower
temperatures in a possible range of roughly ±200 K.
Concluding Remarks
(B.19)
274 Appendix B
(B.21)
where the equal sign determines the critical point Tc and the inequality is
for the existence of immiscibility in the binary system (cf. also Fig. 4.1).
Motel has found that for an overwhelming majority of binary liquid alloys,
equation (8.21) is valid; therefore, it can be used to predict the presence
or absence of immiscibility in liquid alloys for which phase diagrams have
not yet been determined. The values of b for this purpose range from 1 to
6 as tabulated and discussed in detail by Mott.
Estimation of Enthalpy of Alloy Formation 275
References
1. A. R. Miedema, 1. Less-Common Met. 32, 117 (1973).
2. A. R. Miedema, R. Boom, and F. R. deBoer, 1. Less-Common Met. 41, 283 (1975).
3. A. R. Miedema, 1. Less-Common Met. 46, 67 (1976).
4. R. Boom, F. R. deBoer, and A. R. Miedema, 1. Less-Common Met. 45, 237 (1976),46,271
(1976).
5. A. R. Miedema and P. F. de Chatel, in Theory of Alloy Phase Formation, edited by L. H.
Bennett, Metall. Soc. AIME, p. 344 (1980).
6. E. Wigner and F. Seitz, Phys. Rev. 43, 804 (1946).
7. L. Pauling, The Nature of the Chemical Bond, Second Edition, Cornell University Press,
Ithaca, N.Y. (1960).
8. J. A. Alonso and L. A. Girifalco, in Theory of Alloy Phase Formation, edited by L. H.
Bennett, Metall. Soc. AIME, p.484 (1980); see also L. H. Bennett and R. E. Watson, in
Theory of Alloy Phase Formation, p. 390.
9. C. H. Hodges, 1. Phys. F 7, L247 (1977); see also Ref. 5, p. 503.
10. P. C. P. Bouten and A. R. Miedema, 1. Less-Common Met. 65, 217 (1979).
11. P. C. P. Bouten and A. R. Miedema, 1. Less-Common Met. 71, 147 (1980).
12. H. H. van Mal, K. H. J. Buschow, and A. R. Miedema, 1. Less-Common Met. 35, 65 (1974).
13. A. R. Miedema, K. H. J. Buschow, and H. H. van Mal, 1. Less-Common Met. 49, 463, 473
(1976). See also K. H. J. Buschow, H. H. van Mal, and A. R. Miedema, 1. Less-Common
Met. 42, 163 (1975).
14. J. H. N. van Vucht, F. A. Kuijpers, and H. C. A. M. Bruning, Philips Res. Rep. 25, 133 (1~70).
15. J. C. Phillips, in Theory of Alloy Phase Formation, edited by L. H. Bennett, Metal!. Soc.
AIME, p.330 (1980).
16. L. Kaufman, Calphad 1,300 (1977).
17. K. C. Mills, Thermodynamic Datafor Inorganic Sulphides, Selenides, and Tellurides, Butter-
worths, London (1974).
276 Appendix 8
Several empirical attempts have been made to correlate the excess partial
molar enthalpies and entropies of solution for metals, and the standard
enthalpies and entropies of reactions of gases with their dilute solutions in
metals. We shall limit this appendix to the correlation of (1) the excess
partial molar enthalpy, aft = H~, with the excess partial molar entropy,
S~, in dilute solutions of solute metals i in solvent metals and (2) aH· with
as· in equilibria of gaseous oxygen with dissolved oxygen in metals as
discussed in two recent publications. I ,2 Thermodynamic behavior of a solute
metal A in a solvent metal B dictates the processes for effective removal of
A from B in order to purify B. Further, dilute solutions of hydrogen,
carbon, nitrogen, oxygen, and sulfur as solutes generally play important
technological roles in purification of metals. Therefore, in the absence
of data, it is useful to estimate thermodynamic properties of such dilute
solutions.
Table c.l. Values of H7 and S~ for Dilute Binary Solutions of Solute i in Metals a
(continued)
280 Appendix C
·Solute and solvent in pure states have the same state of aggregation. Greatly uncertain values are in
parentheses. States of aggregation are denoted by I for liquid and s for solid.
the proposed correlation have been labeled as Au-Fe, Bi-Na, Cd-Au, and
so on. The values of ii~ and S~ calculated from Margules-type equations
for concentrated solutions and extrapolated to infinite dilution have been
disregarded. All the data are for the systems in which both the solute and
the solvent have the same state of aggregation. The arithmetic average of
ii~1 S~ = K, listed in the seventh column of Table C.l, was found to be
K = 3400 K and the line showing the empirical relation
(Jig-atom) (C.l)
- = -( 1--;;T) = -
G~ H~ S~(K - T); (Jig-atom) (C.2)
Co-Si
Au-Cd '"
CD
-30 .0
a;
ID Cd-Au
CD -40 ~
a
(J) ID
+20 Cd-Sb Au-Fe
eD
Cu-Sn
(J) +10 •
8
Pd-Fe
-SO -60 -40 eD
,
• (J)
+40 +60
• '"
Pd-Cu CD
Pt -Au 8<ll Pt-AI ll.ii;lkcal (g alom)-'
e -10
Fe-AI
8
I.
'"
_~I
Au-Pd _/1)
e.
:>:
-20
• a;
Cu-Pd
..
~"I
0
~
Cd-Au
(I)
b
Figure c.t. Correlation of jj~ and S~ for infinitely dilute binary solutions of i in solvent
metals. (~jji = jj~.) (From Kubaschewski' with permission.)
Thermodynamic Properties in Dilute Solutions 283
metals 2 ,13,14 such as Ag, Co, Cu, Fe, Ni, Pb, and Sn yielding I1H- and I1S-
of adequate accuracy to devise a method of estimation.
The usual method of equilibration is either with (H 2 + H 2 0) or (CO +
CO 2 ) mixtures (see Chapter 6) to obtain
• Xo··
O.50ig) = [0]; I1G o = -RT In rn- = I1Ho - TI1S o (C.3)
VP02
Table C.2. Standard Gibbs Energy Change for O.50 2 (g) = [0] in Selected Metals a
where -9.151 is for the change in the standard state from atomic percent
to atomic fraction, tlH~98 and tlH~ are in cal/ g-atom oxygen, and tlS~ is
in cal/g-atom-K. The equations for the transition elements in which the
d-orbitals are partially filled are
The difference between equations (C.5) and (C.7) is 6.0 cal/ g-atom-K, which
must be considered as an empirical term because theoretical justifications
for this value are tenuous. The numerical values obtained from equations
(C.4)-(C.7) are listed in Table C.2, indicating that the experimental and
estimated values of tlG~(exp) and tlG~(estm) are in good agreement, except
for the Ni-O system. The estimated equations for 18 transition metals are
listed by Fitzner/ whose entropy term must be changed as follows to bring
his results in line with the notation used in this appendix: tlSE(Fitzner)
-9.151 = tlS~(this appendix). The estimated results for tlG~ of oxygen in
most of the transition metals are probably within 15 kcal/ g-atom[ 0] of the
experimental value for tlG~. However, the available data for some of these
metals are scanty and greatly scattering. Therefore, new experimental data
are needed for oxygen in the transition metals to test the foregoing empirical
equations and, if necessary, to modify them to attain greater degrees of
success for prediction of thermodynamic properties of dissolved oxygen.
The method may also be extended to other interstitial solute elements such
as H, C, N, and S.
References
11. L. L. Seigle, C. L. Chang, and T. P. Sharma, Metall. Trans. AIME lOA, 1223 (1979).
12. D. A. Stevenson and J. Wulff, Trans. Metall. Soc. AIME 221,271 (1961).
13. See, e.g., E. Fromm and E. Gebhard, Gase und Kohlenstoff in Metallen, Springer-Verlag,
Berlin (1976).
14. T. Chiang and Y. A. Chang, Metall. Trans. AIME 7B, 453 (1976).
15. E. S. Tankins, N. A. Gokcen, and G. R. Belton, Trans. Metall. Soc. AIME 230, 820 (1964).
16. F. D. Richardson, 1. Iron Steel Inst. London 166, 144 (1950).
17. K. Fitzner, K. T. Jacob, and C. B. Alcock, Metall. Trans. A/ME 88, 669 (1977).
18. L. B. Pankratz, J. M. Stuve, and N. A. Gokcen, Thermodynamic Data/or Mineral Technology,
Bureau of Mines Bull. 677 (1984).
D
(continued) :
N
Condensed phase, ~
~
298.15 K Phase transitions Ideal gas, 298.15 K
Atomic
Element weight C·p S· Tm t.H:;' Tv t.H~ Cop S· t.Hf o t.Gf·
Ti 47.88 5.980 7.320 1943 3.30 3562 100.6 5.839 43.066 113.00 102.342
...a
Tl 204.383 6.290 15.340 577 0.990 1746 39.215 4.968 43.226 43.25 34.936
1:
"
Tm 168.9342 6.460 17.690 1818 4.025 2220 45.6 4.968 45.412 55.50 47.235
U 238.0289 6.612 12.000 1407 2.185 4407 110.92 5.663 47.724 127.00 116.349
V 50.9415 5.950 6.915 2202 5.461 3694 106.8 6.217 43.544 123.20 112.279
W 183.85 5.800 7.800 3695 8.5 5828 196.92 5.092 41.549 203.40 193.338
Xe 131.29 161.3918 0.55 165.Q3 3.02 4.968 40.530 0 0
y 88.9059 6.34 10.62 1795 2.724 6.181 42.869 100.7 91.085
Yb 173.04 6.39 14.30 1092 1.830 1467 30.800 4.968 41.352 36.35 28.284
Zn 65.38 6.070 9.950 692.73 1.750 1180 27.565 4.968 38.451 31.17 22.672
Zr 91.22 6.186 9.320 2128 5.000 4682 139.11 6.367 43.315 148.30 138.164
...~
E
The binary phase diagrams selected 1•2 in this appendix are for convenient
reference in pursuing the topics covered in this book. The objective is to
present a representative number of various types of diagrams with various
types of phase equilibria.
Selected thermodynamic data from Hultgren et ae are for the excess
gram atomic (or molar) enthalpy and Gibbs energy of formation from which
the corresponding partial molar properties and activities can be calculated.
The procedure yields values as accurate as those listed by Hultgren et al.
at the composition intervals of 0.1 if a power series of the type gIVen by
equation (1.63) with five terms is used for He and for se, and the results
are then combined in G e = He - TSe to obtain five terms dictated by the
data for G e • We illustrate the procedure by using only two parameters:
(E.!)
(E.2)
(E.4)
(E.5)
The same procedure also yields the equations for G~ and G~, as well as S~
and S~.
1200 ~
1234- L
./
V I
"\~ I"'----.. ./
~
V
(Cu)
"
1100
(Ag) ;0.141
~ ...-'" 1052· ",
0.399 0.951\
1000
/ \
900
800 I
700
I
Ag 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 Cu
XCu
1200
110 0
1; ~ p
\.-
1000 / Y2 (Cu)
~33.25°
/
V
897-
Erj
900 / 864·
80 0
\8~
0.025 .0.17.3.
I--- - e.~~T ~
8 ~~ O.BOE
I'\(AI) I~I
~8 TJi
70 0 0.671
.60.. 636-
~2 0.504
60 0
~
500
'" -------
I I I I I
9~3~~O I
II
900 ~
----
I--- L
---
800 ~
r-- I---
r--
692.6550
700
..........
655 0 ~~
(AI), 624.6 0 {All z /0.665 0.887"'" p.976\
600 ,../
0.395 -.....
(Zn)~
.....- ~/ 548"
,/ 0.594 0.986
500
400 I
(
I
I
300
200
Wf. "10 Go
10 ~O 30 40 SO 60 70 8f' 9.0
00
GoA$
1~38·
UqJd
I~ 00
~
/V ~
Vr
\
\
10 00
817 GOA
~
8 00
• 810·
6 00
I
A"" GoA.r Liq. + Ga.At
\
4 00
00
GaA$~Go ~9'5.
_1- I
o 10 ~O 30 40 SO '070 80 90/00
Ar.". Ga
Figure E.4. As-Ga system.
298 Appendix E
L 20~
2000
,,'
- - l--- I---
.... 7
~ .... ,,
1800
V"'"
/"
/::---
/'
1600
~
1400
V - 1~2S· -
0.60 r---......
r1336.15. (Au, Pt)
1200 V '\
Au 0.1
/
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
\
Pt
XPt
xPt: 0.11* 0.20 0.30 0.42* 0.79* 0.80 0.90, at 1423 K
G8 : 572 863 1121 1368 1123 1084 657, calfg-atom
Figure E.5. Au-Pt system. Asterisks indicate phase boundaries; data are for single solid phase
region with solid Au and Pt as standard states.
Selected Binary Phase Diagrams 299
OK ,
10 20
I
30 40 50 60 70 80 90
",,1364-
1300
L ,/
/
1200
V
......V ~
/ ~
'" \
1100
1000
I {l--
\
900 i\
800
722" \
700 1'0.99-
597· (Te)-
600
594.18-
f-(Cd)
Cd 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 Te
X Te
Figure E.6. Cd-Te system.
300 Appendix E
500 I
450H/'---+--+--t-----i:--_+--+--t---+--+_---1
400~-4-_+-~-~-+--+--~_;-_+-~
10 30 60 70 80 90
"'- --
~
~S·
ISOO
-
t;:,.
1400
b/' // I
/
~LiqUid-
I
I
1\
,,,
\
IJOO
« + Liquid
\
1200 1-1'
I
I
\sl
1100 I
I 083 0
1000
900
1 + e
\
800 a
~
--- --- - - - 1----
~agl;;(l ~/;;;,,~~ - - --- --- ---
700
600
aL
SOO
0 10 20 30 40 SO 60 70 80 90 100
Wt.% Cv
--
10 20 30 40 50 60 70 80 90
-- -
II I II I I II II 1726"
--
L ~ ~-
1600
L-- ~
-- -- - -~
1400
~56.55°
1200
{Cu, Nil
1000
....... (Cul
800
6310
600
./
-. /
400 -
CURIE TEMP.
/
./
200
/'/
/'
At. % Sn
·c /0 cO ,JO 40 SO 60 70 80 90 100
1100
~
\\ \
1000
900
800
1\ 798.1\
700
\ "'~ P ~
7SS'
~
Liquid
"I / ~ 640~
600 <X
6' \) -I~ f4 ~
~
~\;f~
'r-SJ £ I.iquid ~
""-
SeO ~
SOO
L 41S·
"'" r\
\
400
/ lSO'
/
! \
,JOO
1J ee7'
cOO 189' 186°-
:
ISO 1'-17' 1
o 10 cO ,JO 40 SO 60 70 80 90 100
Wt % Sn
Wt%Si
·c S 10 IS ..0 "S JO 40 SO 60 10 80 90
1600
ISOO "- I I 1 I
~ ~ ~
~~ + ~
1410°
\'\ /
1400
V
\
"'" \'' ii'O./ '.
I\. /'
1275°
IJOO
/ ~ IS-,L
..... I 02 d 1212° 08'
"'. t
, .. 00
r-£'9f
1/00
Y
1/ '\ I-Fo.. S,
1'090·
,-FoS, f-- Fo 2IH}
1000
II J '07rl ~FoSSiJ
r2!L
I 99~
I at 1/; 0&,
960"
900
700
600
I--
,,
S40· \
,
SOO
\
\
400 \
o 10 ..0 JO 40 SO 60 70 80 90 100
At%Si
,.....904
L
800· /
~
800 773
".--
700 V 0.683
600
'/
500 I
t,..42S.76°
'r-I'
428-
400
:.... (In) (Sb)'"
'"
900 t...
i\ ""V 834.4°
800 \11 o.
K~·0345
(Mg) .~
-
700
{J
I""
"
600
900 ."...,.
L
~
~ !
800
./
/ I
700 I
V (Sbl-f
SOO.s- ./
or
600
~V 524.4- 1
500
~o.o58
(P1bl
\
\
400
Pb 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 Sb
XSb
600.6"
r-- r--.
600
"'"
'"
L
.............
~
i'-..... ~ 5)5.06
500
\ 456 0 ~ ~ .- ~ u(.~
(Pb)
V 029
0.74 0.985
If
400
900
L
V
6"9. J/,
800
700 II 690.81·
692.6S~~
III~ISI 591.J.
~9977
I
600 I
0.022.
(Z7)-
500 f--CPbl
Pb 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 Zn
XZn
References
List of Symbols
1. Latin Symbols
(Symbols not defined here are empirical constants usually used on only one
or two pages.)
2. Greek Symbols
a phase designation
a defined by equations (3.24) and (3.40), i.e., a = 01 kT
a NWI RT in equation (4.12)
as short-range order parameter defined by equation (4.59)
f3 phase designation
f3 II kT in Chapter 3 only
f3 defined by equation (4.41) and used only in Chapters 4-6
'Yi activity coefficient, equation (1.32); 1'7 when Xi ~ 0
.y activity coefficient in equation (6.2)
.:l difference; final state minus initial state
a partial differential
8A solubility parameter 8~ = .:lEAl VA where .:lE is enthalpy of
vaporization or sublimation
ei energy level, only in Chapter 3; energy of an interstitial atom
in Chapter 6
e\j) Wagner interaction parameter, equation (1.91)
71 efficiency, percent, equation (7.71)
K empirical parameter equal to 3400 K, equation (C.1)
A defined as el kT in Chapter 7
A phase designation
316 Appendix F
Periodic Table
(Mendeleev Table of the Elements)
IA
I H
0.115
0.46 CPH
6Hv, (I)",
2.20 FCC
TIA
26 Fe
1. ' Alomic 99.3 BCC }cryotal
radius (2)- 1.28 FCC
3 Li 4
38.41
Be , ~.83 BCC
struclure, (5)
II No 12 Mg
25.75 35.0
1.92 BCC 1.60 CPH
0.93 CPH 1.31
3,' 3,' IDA riA Y.A lQA ilIA / Y.III A",
19 K 20 Co 21 Sc 22 Ti 23 V 24 Cr 25 Mn 26 Fe 27 Co
21.31 42.5 BCC 113.0 123.2
90.3 BCC BCC 95.0 67.1 BCC 99.3 BCC 101.5 FCC
2.3B BCC 1.97 FCC 1.64 1.47 1.36 BCC 1.28 BCC 1.12 FCC 1.28 FCC 1.25 CPH
0.82 1.00 1.36 CPH 1.54 CPH 1.63 1.66 1.55 CUBIC 1.83 BCC 1.88
4.' 40' 3d ' 4. 2 3d 24. 2 3d l 4.' 3d 54.' 3d54.' BCC 3d'4,2 3d!4"
37 Rb 38 Sr 39 Y 40 Zr 41 Nb 42 Mo 43 Tc 44 Ru 45 Rh
19.33 39.2 BCC 100.7 14B.3 175.2 157.5 158 153.6 133.1
2.51 BCC 2.15 1.81 BCC 1.60 BCC 1.47 BCC 1.40 BCC - CPH 1.34 CPH 1.34 FCC
FCC CPH CPH
0.82 0.95 1.22 1.33 1.60 2.16 1.90 2.20 2.28
5.' 5,2 4d'5.' 4d 25.' 4d45. ' 4d 55. ' 4d 55s' 4d!5s' 4d'5,'
55 Cs 56 Bo [\ 72 Hf 73 To 74 W 75 Re 76 Os 77 Ir
18.32 43.0 148.0 BCC 186.9 203.4 184.0 111.61 159.0
2.70 BCC 2.24 BCC \ 1.59 CPH 1.47 BCC 1.41 BCC 1.38 CPH 1.35 CPH 1.35 FCC
0.79 0.89 1.30 1.50 2.36 1.90 2.20 2.20
("9..z. \
t \\
6.' 6,2 4t'45d 26s' 4f '4 5d l Ss' 4f'45d 4SS2 4f '4 5d 5S.' 4f '45d 66,2 4f'45d!S,'
87 Fr 88 Rol\
-
-
-
- \%
0.70
78'
0.90
78'
"90 \ ~ 57 Lo 58 Ce 59 Pr 60 Nd 61 Pm 62 Sm
- 8CC
..A \ 103.0 BCC 101.0 BCC 85.0 78.3 49.4
FCC 1.82 FCC 1.83 BCC 1.82 BCC - CPH - BCC
~ 1.87 CPH CPH
'" 1.10 CPH 1.12 CPH 1.13 1.14 1.13 1.17 RHOM
\~
\01 89 Ac 90 Th 91 Po 92 U 93 NP 94 Pu
FCC 142.73 8ce 145.0 127.0 BCC 111.1 BCC? 82.5 BCCa,_
\ 1.10
1.80
1.30
FCC BCC
1.50 BC-TElR
1.38 TETR
1.38 ORTR
- TETR
1.36 ORTR
- FCC y ,.
1.28 NON.,11
4
5f 6d'7. 2 5f 8 782
46 Pd 47 Ag 48 Cd 49 In 50 5n 51 5b 52 Te 53 I 54 Xe
90.4 68.09
58.15 11.99 62.1 50.S
26.13 10.02 3.02 fC- BC- FCC
1.31 fCC 1.44 fCC 1.52 CPH 1.51 TETR 1.58 TETR 1.61 RHOM 1.43 TRI US ORTR 2.18
2.20 1.93 1.69 1.18 1.96 DIAN 2.05 2.10 2.6S -
4d lO 4d10s.1 4dI05.' 4d 105,'pl 4d 105,'p' 4d 10s,'p' 4~10s,'p' 4d I05,'pl 4d I05,'p'
78 Pt 79 Au 80 Hg 81 TI 82 Pb 83 Bi 84 Po 85 At 86 Rn
135.1 81.5 14.S1 BC-TETR 43.25 BCC 4S.15 49.5 - - 3.92
1.38 fCC 1.44 FCC 1.55 RHOM 1.11 1.15 fCC 1.82 RHOI 1.40 CUBIC -
CPH
- fCC
2.28 2.54 2.00 2.04 2.33 2.02 2.00 2.20 -
4f 14 5d,&,1 4f 145dIO S,1 4f I4 5dI06,' 4f145d"61'pl 4f l'5dloS,'p' 4f145dl0&s'P' 4t1 45dIOS,",' 4fl'5dI06s',s 4f 145d I06,'p'
63 Eu 64 Gd 65 Tb 66 Dy 67 Ho 68 Er 69 T~170 Yb 71 Lu
55.5 BCC, 36.35 BCC 102.2 BCC
44.9 95.0 92.9 69.4 11.9 BCC 15.8
BCC BCC BCC
2.04 BCC 1.80 1.11 1.11 1.16 CPH 1.15 CPH 1.14 CPH 1.93 fCC 1.13 CPH
1.20 1.20 CPH 1.20 CPH 1.22 CPH 1.23 1.24 1.25 1.10 CPH 1.21
4f'6.' 4f 15d l6s' 4f'6,' 4f 106,' 4f"6s' 4f"6,' 4f ll6s' 4f 6,'
14 4f"sd IS,'
95 Am 96Cm 97 Bk 98 Cf 99 Es 100Fm 101 Md 102 No 103 Lr
61.9 92.6 - -
fCCr --
- - - -
- fCC, - fCCI - fCC' - - - - -
1.30 CPH 1.30 CPH 1.30 CPH! 1.30 CPH? 1.30 1.30 1.30 1.30 -
5fll" 5f 76d l 1s' 5f l 1l' 5f l0 1l' 5f"18' 5f l'1I' 5f l'1I' 5f141s' 5f146b,'
321
322 Index