Sie sind auf Seite 1von 112

SOLID STATE PHYSICS. VOL.

56

Transition-Metals and Their Alloys

R . E . WATSON AND M . WEINERT

Department of Physics, Brookhaven National Laboratory


Upton, NY 11973-5000 USA

I. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2
1. Hybridization and Bonding Energies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
II. Mostly the Elemental Metals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2. The Crystal Structures . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3. Cohesion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7
4. Some Other Applications of the Friedel Model . . . . . . . . . . . . . . . . . . . . . . . . . . . 13
5. Lattice Instabilities, High-Temperature, and Other
Excited Phases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15
III. Alloying Trends: Electronegativity, Size. and Valence Electron Count . . . . . . . . . . . . . 21
6. Electronegativities . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 22
7. Atomic Size . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
8. Valence Electron Counts and Valence Electron Character . . . . . . . . . . . . . . . . . . 30
IV. Wigner-Seitz Cells, the Frank-Kasper Phases. and the Bernal Environments . . . . . . . . 34
9. Wigner-Seitz Cells (Voronoi or Dirichlet Polyhedra) . . . . . . . . . . . . . . . . . . . . . . 34
10. The Frank-Kasper Phases and Closely Related Structures . . . . . . . . . . . . . . . . . . 36
11. The Bernal Environments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
I2. Atomic Environments and Polymorphism in
Transition-metal-rich-Metalloid Glasses . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 48
V. Charge "Transfer" and Alloying . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
13. Charge Transfer as Sampled by Optical and Photoelectron Experiments . . . . . . 50
14. Charge Tailing and the Local Sampling of Orbital Populations . . . . . . . . . . . . . . 59
VI. Electronic Structure and Phase Predictions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
15. Approaches for Alloy Calculations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
16. Density Functional Theory . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
17. Phase Stability of Binary Alloys: Zr-AI . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75
18. Ternary Alloys: Transition-metal Aluminides . . . . . . . . . . . . . . . . . . . . . . . . . . . . 79
19. Entropy Contributions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 85
VII. Magnetic Moments and Site Volumes . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 93
20. Fe and Mn . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 94
21. The Iron Nitrides . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
22. The Rare-earth-Transition-metal Hard Magnets . . . . . . . . . . . . . . . . . . . . . . . . . . 98
VIII. Band Gaps in Metallically Bonded Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
23. FeSi and Other Materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 100
24. Transition-metal Aluminides and the BiF~ Structure . . . . . . . . . . . . . . . . . . . . . . . 102
IX. Epilogue . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Acknowledgments . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 109
Appendix: Population Analyses:
Standard and Site Specific . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 110

ISBN 0-12-607756-8 Copyright 9 2001 by Academic Press


ISSN 0081 - 1947/01 $35.00 All rights of reproduction in any form reserved.
2 R.E. WATSON AND M. WEINERT

I. Introduction

In this article, we will discuss the transition-metals, emphasizing the phases in


which they and their alloys form, the competition among these phases, and their
magnetism. At the onset, it is appropriate to quote from the first words of Wigner
and Seitz in the first volume of this series: ~

If one had a great calculating machine, one might apply it to the p r o b l e m of


solving the Schr/Sdinger equation for each metal and obtain thereby the interesting
physical quantities .... Presumably the results would agree with the experimentally
determined quantities and nothing vastly new would be learned .... It would be
preferable instead to have a vivid picture.., a simple description of the essence
of the factors which determine cohesion and an understanding of the origins of
variation in properties from metal to metal.

During the time since Wigner and Seitz wrote, we have yet to acquire the "great
calculating machine," though for many purposes we do very well with what has
become available. What is more, the results from our calculating machines often
can offer insight. Simple pictures, when correct, were and remain invaluable. Many
of these rely on ideas which predate the observation of Wigner and Seitz, having
been employed by workers such as Debye, Friedel, Hume-Rothery, Pauling, and
Slater.
The study of transition-metals and their ordered alloys has a long history 2~ and
cover a wide range of topics. We will consider elements having open d shells as
well as the noble metals. Although the latter have filled d-bands, their bonding
shares important features in common with the elements to their left. In addition
to transition-metal-transition-metal alloys, alloys with such main group elements
as Si and A1 also will be encountered. The rare earths and the actinides with their
open f, as well as open d, shells will only be of passing concern.
This article is not an exhaustive survey of the field. Instead, we will discuss a
number of different topics that we consider important for understanding the
properties of transition metals and their alloys, and will illustrate them using
particular examples. Although we are sure that others would have chosen different

E. E Wigner and E Seitz, Solid State Physics. 1, 97 r 1955).


2 In an earlier volume in this series, D. G. Pettifor, Solid State Physics, 40. 43 (1987). has dealt with
many important issues concerning the electronic structure of both the simple metals and the transition-
metals.
3 Other articles appearing in this series pertain to the subjects in this article. For example, see the articles
on disorder in alloys by H. Ehrenreich and L. M. Schwartz, 31, 150/1976), and by D. De Fontaine,
47, 33 (1994): on pair potentials (and beyond) by A. E. Carlsson. 43, 1 11990): on electronic structure
using the tight-binding-recursion approach by V. Heine. D. W. Bullett. R. Haydock, and M. J. Kelly
in 35 (1980): and on the physical properties of the transition-metals and of the other elements by
K. A. Gschneidner, Jr., 16, 275 (1964).
TRANSITION-METALS A N D THEIR ALLOYS 3

/ \

iI \

/ ,,, i \
i " i \
I ,,, i

I i i \ ~
I i i /
i I I ~ 1 i
I i

9,, / i
9,, / \
,, . \ i
,, i
,, ~ 1

FIG. 1. The effect of hybridization between a pair of electron levels and betw'een a 2-fold and a
4-fold nearly degenerate set of levels: In each case the center of gravity of the energy of the sets of
levels is preserved upon mixing.

topics, we hope that the article provides an entree into this vast area of transition-
metal alloys.

1. HYBRIDIZATION AND BONDING ENERGIES

Throughout this article the hybridization between electron levels and the conse-
quent bonding energy will be encountered. The ideas are simple, but not always
correctly understood. Consider the two-level situation plotted in the lefthand part
of Fig. 1 that often is found in elementary texts. The two levels may be associated
with orbitals centered on the adjacent atoms of a diatomic molecule or they might
reside at the same atomic site in a crystal where they are allowed to mix by
symmetry. If the orbitals mix, there will be high-lying orbital character mixed into
the lower orbital and vice versa. The center of gravity of the energies of the two
remains fixed and so bonding energy is gained only. if the lower bonding level is
occupied and the higher lying antibonding level empty. 4 The fight panel shows a
slightly more complicated case, i.e., different numbers of nearly degenerate levels,
with approximately equal mixing matrix elements for the nearly degenerate states.
Again the center of gravity holds fixed and the maximum bonding energy is gained
if all the lower "bonding" levels are filled. This bonding energy is reduced to zero
as the higher lying "antibonding" states are successively filled. This case is germane
to a large class of transition-metal alloys where the d bands of one constituent lie
well below those of the other. For compounds away from 1:1 composition, the two
sets of levels are different in number and, as will be seen, strong alloy formation

4 That the center of gravity stays fixed is a general result that follows from the fact that the trace of
a matrix is invariant. Although arbitrary mixing of these orbitals to form the hybridized states will
shift the eigenvalues of the states, their sum, and hence center of gravity, is not shifted, Of course.
including hybridization with additional states will cause shifts unless they too are included in the
accounting of the center of gravity.
4 R.E. WATSON AND M. WEINERT

will occur at the composition where the low-lying hybridized levels are filled and
the higher lying are empty.

II. Mostly the Elemental Metals

2. THE CRYSTAL STRUCTURES

The crystal structures in which the elemental transition-metals are found form a
well-ordered pattern:

Sc Ti V Cr Mn Fe Co Ni Cu
hcp(bcc) hcp(bcc) bcc bcc a Mn bcc hcp fcc fcc

Y Zr Nb Mo Tc Ru Rh Pd Ag
hcp(bcc) hcp(bcc) bcc bcc hcp hcp fcc fcc fcc

La/Lu Hf Ta W Re Os Ir Os Au
hcp(bcc) hcp(bcc) bcc bcc hcp hcp fcc fcc fcc

where the structures in parenthesis are high-temperature phases. Mn, Fe, and Co
have structures different from the other members of their columns, which is
generally attributed to magnetism: i.e., based on the corresponding 4d and 5d
elements, nonmagnetic Mn and Fe would be expected to be hcp and Co would
be expected to be fcc. The a Mn structure is closely related to the topologically,
or tetrahedrally, close packed phases (tcp), to be discussed later. These structures
involve atomic sites of differing volumes and hence are appropriate to an alloy.
Mn forms in such a structure with the Mn atoms at different sites having different
magnetic moments (with the larger moment in the larger site). Thus a Mn may
be deemed to be a sort of "alloy" but with atoms differing in moments rather than
in Z. Except for the V and Cr columns, the transition-metals form in the close-
packed fcc and hcp structures. The reason that the V and Cr columns are bcc may
be understood visually by inspecting the densities of states for the three structures
(see Fig. 2), i.e., the Fermi level falls in the characteristic deep hollow in the bcc
density of states for the V and Cr column elements. Having the occupied one-
electron states pushed down and away from the Fermi level at the expense of the
unoccupied states helps stabilize the bcc structure relative to its close-packed
competition. Similarly, there is a bonding-antibonding hollow in the hcp density
of states at band fillings appropriate to the Mn and Fe columns that is more clearly
established than that for the fcc structure, thus stabilizing the hcp structure relative
to the fcc for the elements in these columns.
TRANSITION-METALS AND THEIR ALLOYS 5

, i 9 9 , , - 9 9 , . . . . . . T . . . . .

hcp Ru

~0

f c c

%...,
0

"~0

bcc

\
-8 -6 --4 -2 0 2
Energy (eV)

FIG. 2. The calculated densities of states for Ru in the hcp, fcc, and bcc structures, all at the same
volume. The observed phase is hcp and the lattice volume and c/a ratio were determined variationally.
The energy zero is the Fermi level.

9 5
Pettifor pioneered the quantitative estimates of the structural trend employing a
"hybrid nearly-free electron tight-binding" scheme where the two essential param-
eters were e d, the d band center of gravity, and the band width W,t. Using canonical
bands for the fcc, hcp, and bcc structures, the total energies of the competing
structures were calculated as sums over occupied one-electron energies, i.e., con-
duction electron-conduction electron interactions were neglected. Except for pre-
dicting that the noble metals were bcc by a small energy difference, this scheme
correctly yielded the structural trend. To correct the prediction for the noble metal
structures, Pettifor introduced a repulsive hard core term which penalizes the bcc
d band center of gravity. (For a given volume, the nearest neighbor distances are
smaller in bcc than fcc, thus such a term can be chosen to affect the bcc d bands
more than the fcc ones.) Subsequent calculations, employing muffin-tin potentials
(spherical potentials within atomic sites and constant potentials in the interstitial
region outside), have also tended to predict the noble metals to be bcc; present-day

5 D. G. Pettifor, J. Phys. C, 3, 367 (1970).


6 R.E. WATSON AND M. WEINERT

\\\
240O
' . . . - -- / Y f

\ i I I
\ I I I
\ / I

,r
( i (fcc)
E 2000 hex

1600
Mo 20 40 60 ~0 Ir
Atomic % Ir

FIG. 3. The Mo-lr phase diagram.

treatments that account for the aspherical character of the valence electron charge
and potential (so-called "full potential" methods) predict the correct structure for
the noble metals as well as the correct trend in the remainder of the transition-metal
row.
Noting the existence of a Mn, the structural trend as a function of electron filling

hcp --> bcc ~ tcp --~ hcp ---> fcc

seen in the elemental structures above also holds for transition-metal-transition-


metal alloys providing that they are not too strongly bound. As an example, the
phase diagram for Mo-Ir is shown in Fig. 3. Two tcp phases, the A15 (Cr3Si) and
the cr (Cr-Fe), appear between the bcc Mo-rich phase and the two intermediate
hexagonal phases. Except for the Laves phases, of which more will be said later,
these two tcp phases along with a Mn are the most commonly encountered. When
more than one occurs in an alloy system, the sequence is generally

A15 --> o"---> a Mn

as a function of increasing alloy d-band filling. For the more strongly bound
systems, this trend loses out to the occurrence of stoichiometric compounds.

6 T. Massalski, Binary Alloy Phase Diagrams, Second Ed. (American Society for Metals, Metals Park,
OH, 1990); B. Predel, Landolt-BOrnstein New Series. Vol. IV/5a (Springer-Verlag, Berlin, 1991):
W. G. Moffatt, The Handbook of Binar 3" Phase Diagrams (Genium Publishing, Schenectady).
TRANSITION-METALS AND THEIR ALLOYS 7

3. COHESION

The cohesive energy of an elemental solid is the difference between the total energy
per atom in the solid and the energy of the free atom in its ground state. A "'good"
estimate of the cohesive energy, then, requires treating the solid and the atom with
equal accuracy, which has not proved easy.
Wigner and Seitz estimated 7 the cohesion of the alkali metal sodium. They
introduced Wigner-Seitz polyhedra (or the spheres of equal volume) and defined
one-electron wave functions inside from which Bloch functions were constructed.
The potentials and then the cohesive energies were calculated within the Fock
scheme with an exchange hole centered at the position of the sampling electron.
This procedure yielded a cohesive energy one-third that of experiment and a
lattice constant 15% too large. They then generated a correlation hole between
electrons of anti-parallel spin, which yielded a cohesive energy in much better
agreement with experiment.
Assuming the 3d electrons to be corelike, Fuchs applied ~ the Wigner and Seitz
scheme to Cu. Copper was chosen because of the availability of calculated free
atom wave functions. Fuchs introduced a repulsive d - d core term. The resulting
calculation gave a cohesive energy half that of experiment, a lattice constant 20%
too large, and a compressibility in excellent accord with experiment. Fuchs
concluded that at the stable lattice volume, the free-electron band encourages
contraction, while d electrons favor expansion. This conclusion is opposite to the
9
modern view; for example, using Andersen's atomic sphere approximation
(ASA) that allows one to define partial pressures associated with different bands.
Pettifor observed l~ that the d bands of a transition-metal encourage contraction,
but the increase in kinetic energy of the diffuse s-p bands at smaller bondlengths
favors expansion.
The early endeavors did not include d bonding directly. Friedel, invoking argu-
ments by Wigner and Seitz, described ~3.~2the trend in cohesion across the transition-
metal rows with a simple model. Friedel considered only the d band densities of
states and replaced them by a rectangular density of states. Although this approx-
imation is not correct in detail (cf. Fig. 2), especially for the bcc structure, it is
reasonable in that the density of states is dominated by "'d'" bands of reasonbly
well-defined extent and magnitude. With this approximation, the cohesive energy

7
E. Wigner and F. Seitz, Phys. Rev., 43, 804 (1933) and ibid.. 46. 509 (1934).
8 K. Fuchs, Proc. Roy. Soc. A, 151,585 (1935) and ibid., 153. 622 11936).
9 0 . K. Andersen, Solid State Commun., 13, 133 (1973).
~0D. G. Pettifor, J. Phys. F: 8, 219 (1978).
~ J. Friedel, in The Physics of Metals I, J. Ziman, ed. (Cambridge Unix. Press, Cambridge. 1969), p. 340.
J2 Also see E Cyrot-Lackmann, J. Phys. Chem. Solids, 29, 1235 t1968) and F. Ducastelle and F.
Cyrot-Lackman, J. Phys. Chem. Solids, 32, 285 (1971).
8 R.E. WATSON AND M. WEINERT

may be written (with N - 10 for d bands)"

w
H~ ~-- ~--~n(N- n) + nEd + C, (1)

where n is the number of electrons in the band and W is the band width. The first
term gives a parabolic trend in the cohesive energy, with the maximum occurring
at half filling. The constant C term includes such contributions as correcting the
free atom reference energy for the difference between the ground multiplet state
and the state most appropriate for comparison with the solid. The second, Ea,
term appears since the center of gravity of the band in the solid may differ from
that of some reference level. These last two terms have the effect of shifting the
peak in the cohesive energy from n / N = 1/2 to (1/2 + Ed/W), with a maximum
cohesive energy of (WN/8)[1 + (2Ed)/W] + C. Thus, the shift caused by Ed also
provides a means to take into account the fact that the transition-metals do not
have integral numbers of d electrons. Physically, one expects that the parameters
W, Ea, and C will vary across a transition-metal row. However, assigning them
constant values, Eq. (1) provides semiquantitative agreement with the experimen-
tal cohesive energies, as shown in Fig. 4 for the 4d row. The value of W = 6 eV
used to fit the 4d values is comparable to the Rh band width. That the trend in
cohesive energy is given correctly across the whole row (with the possible excep-
tion of Ag) by such a simple model is rather remarkable.
There are consequences arising from this trend. The heat of formation of an
AxBl_x alloy or compound is

AHu - E [ A x B l _ x ] - x E [ A ] - (1 - x ) E [ B ] , (2)

I I I I I I I 1 I
.& - - - ~

~
.," \
/ \
\
6
!

> /
! \
\
/ \
r
\\

~4
\
\ 9
.,-, 9 Experiment \
\
0 Friedel Model \\\
-0= 2

0 1 1 I I 1 I i 1 1

Y Zr Nb Mo Tc Ru Rh Pd Ag

FIG. 4. Comparison of the Friedel model, Eq. (1), with the experimental cohesive energies for the
4d elements. A band width W = 6 eV was assumed.
TRANSITION-METALS AND THEIR ALLOYS 9

0.5 I I " [

0.0
E
O

O CsC1
-0.5 [] CuAuI

-1.0

1 1 1 1 1 1 1

PtLu PtHf PtTa PtW PtRe PrOs PtIr

FI6.5. Calculated heats of formation, employing muffin-tin potentials, for the 50/50 Pt-5d-element
binary-alloy phases taken in the CsC1 and CuAu-I structures.

where E[A] and E[B] are the total energies (per atom) of the elemental solids of A
and B and E[AxB]_x ] the total energy of the compound. Thus, elements in the middle
of the transition element row, such as W, have the most to lose, and hence the least
to gain upon alloy formation. Also, if two transition-metals were to create a common
band, one might expect the largest heats of formation when the two elements are
from opposite ends of the row so that the band filling of the combined bands is
near the maximum, while the elemental energies are small. These effects can be
seen in Fig. 5, where calculated heats of formation 1314 for ordered 50/50 compounds
of Pt with its 5d neighbors are shown. Non-binding is seen to extend from W to
the right. (PtIr binds weakly as a disordered fcc phase). It is to be emphasized that
the cohesive energies of those compounds which fail to form may be substantial,
but they lose out to the combined cohesion of the elemental metals.
The Wigner-Seitz sphere approach has been extended to include d and non-d
electrons in several ways. One is the renormalized atom approach ~5~6 which treats
the estimate of cohesion in steps. The first step is to prepare an atom by taking
it from its ground multiplet state to the center of gravity of the d"s" configuration
considered most relevant to the solid, and is obtained from the experimental

13 R. E. Watson, J. W. Davenport. and M. Weinert. Phys. Rex'. B, 36, 6396 (1987).


l,~ The same trends (plotted with a scale factor) have been shown by A. R. Williams. C. D. Gelatt,
and V. L. Moruzzi, Phys. Rex~ Lett., 44, 429 (1980). These calculations employed the augmented-
spherical-wave method, yielding spherical solutions for the CsCI and CuAu-I structures within volume
filling overlapping Wigner-Seitz spheres.
15 L. Hodges, R. E. Watson, and H. Ehrenreich. Phys. Rex: B, 5, 3953 (1972).
~6C. D. Gelatt, H. Ehrenreich, and R. E. Watson, Phys. Rex'. B, 15, 1613 (1977).
10 R.E. WATSON AND M. WEINERT

multiplet spectra. ~7Then for this atomic configuration, the total energy difference
between Hartree-Fock calculations for a free atom and a "renormalized" atom
(an atom with its electron orbitals cut off and "renormalized" within a Wigner-
Seitz sphere) is obtained. The potential derived from the renormalized charge is
used to obtain band levels necessary to construct a tight-binding band structure
and, in turn, to estimate the difference between the energy of the renormalized
atom and that of the band structure. The potential is {-dependent, with a full d
Coulomb hole centered on and completely within the Wigner-Seitz sphere when
calculating d band effects, and an s hole otherwise, s-d hybridization is included
and its effect on the cohesive energy is taken to be the change in energy associated
with turning its s-d matrix elements off in the tight-binding scheme.
Cohesive energies (assuming all the metals to be fcc) for Cu and the 4d elements
are shown in Fig. 6. The agreement between calculation and experiment is very
good and, as expected from Friedel's observations, d-d bonding prevails. In the
middle of the row where multiplet effects are largest, the atomic preparation makes
a substantial contribution. The renormalization contribution is, in large part, equiv-
alent to the second term of Eq. (1), although it peaks in the middle of the row,
rather than increasing monotonically across it. In contrast, the bonding energy
associated with the non-d electrons alone, is near zero at the middle of the row.
As noted, these estimates were done for the fcc structure: in the cases of bcc Nb
and Mo, the energy gained upon going to the bcc structure will reduce the
extraordinary agreement seen between the calculated and experimental heats.
Finally, the s - d hybridization contribution is reasonably constant, except for Ag
where it is near zero. This small value results from the position of the d bands
and not the atomic configuration: For Ag, the top of the filled d bands is 4 eV
below the Fermi level, while for Cu the distance is half as great, resulting in
greater s-d hybridization. This hybridization energy is estimated to be respon-
sible for half of the cohesion of Cu, and is comparable in magnitude to that in
the other transition-metals.
Using an ASA-based scheme that treats the d bands self-consistently, but neglects
s-d hybridization and the atomic preparation energy, Pettifor obtained z~4d cohesive
energies in essential agreement (cf., Fig. 7) with those of Fig. 6. As Pettifor noted,
the good agreement between the two sets of calculations is, in part, due to the rough
cancellation between these two terms, as can be seen in Fig. 6. At the same time
Moruzzi, Janak, and Williams ~s~'~2~ did self-consistent Korringa-Kohn-Rostoker

iv Atomic Energy Levels, C. E. Moore, ed. (U.S. National Bureau of Standards, Circular 467, 1971 ).
~s j. F. Janak, V. L. Moruzzi, and A. R. Williams, Phys. Rev. B, 12, 1257 t1975).
~'~V. L. Moruzzi, A. R. Williams and J. F. Janak, Phys. Rev. B, 15, 2854 (1977).
-~~ L. Moruzzi, J. F. Janak, and A. R. Williams, Calculated Atomic Properties of Metals (Pergamon,
New York, 1978).
TRANSITION-METALS AND THEIR ALLOYS 11

m I ~
L.--.i
i ' I ' r' , i
- Q o lml "
>
_ 13 _.
6
9 H
&m,w

4
- H -
> 9 Q
- o 9 -

= 2
9 Calculation 9
m

13 Experiment
, ,, , , I I , I m 1_

>
Z r. Z ~
4
= = | E i i i conduction
i..
-= i= =
E_
E
E
=-.
~:~
-=
-.
-- E
:
b/

I-T.l
0

renormalization
atomic preparation

Y Zr Nb Mo Tc Ru Rh Pd Ag Cu

FIG. 6. Components of the cohesive energies calculated 1r for Cu and for the 4d elements. The
experimental energies are denoted by the open boxes and the total calculated energies by the filled
boxes.

(KKR) 21"22muffin-tin local density calculations for a large number of metals. Their
cohesive energies are also in agreement with the ones above. These authors included
a form of atomic preparation energies by determining the spin-polarized local
density ground state atomic configurations. Granted that the three calculations differ
substantially in at least some details, the extent of agreement between them is
surprising as is the agreement with experiment. Results of more rigorous all-electron

21 j. Korringa, Physica, 13, 392 (1946).


--W. Kohn and N. Rostoker, Phys. Rev., 94, 1 l ll (1954).
12 R . E . WATSON A N D M. W E I N E R T

I I 1 I I I ! 1 I

,.,8 " Z~-._._ .... o ....A : : ~ - -

/,~-".f " ~ \ ,

(D
Experiment '
,~. . . . 3 Gelatt, et al. \
o2
r,.) o ..... o Pettifor \>
~-----~ Moruzzi, et al. ~ -

0 I 1 1 1 1 1 I I 1
Y Zr Nb Mo Tc Ru Rh Pd Ag

FIG. 7. Comparison of the calculated 4d cohesive energies by Gelatt et al.,~6 Pettifor,~0 and Moruzzi
et al. ~8"19"2~to the experimental values,

LDA calculations, 23 including aspherical terms in the charge density and potential,
are presented in Fig. 8. Here the atomic preparation involved using experimental
spectral data to estimate the centers of gravity of various atomic configurations
and then using LDA calculations for those free-atom configurations and for the
solid. While results for the d" and d " - ~ s reference configurations agree, those for
the d " - 2 s 2 do not, and the spread across much of the row is measurable. Overall,
the disagreement between calculation and experiment is much worse than that
encountered for the more rudimentary calculations and its sign indicates that LDA
has underestimated the total energies of the free atoms relative to those of the
elemental solids.
As will be seen later, the situation is better when dealing with the heat of formation
of a compound, which involves differences in energies of the compound and the
elemental reference solids, Eq. (2), and not the atomic energies directly. The conclu-
sion is that there apparently is better cancellation of errors among different solids
than between the solid and the free atom. Such an observation can be rationalized by
noting that, in metallic systems at least, the local environments and bonding in the
elemental system and the compound are reasonably similar, whereas there are large
differences between the free atom and solid ones. A simple measure supporting this
argument is to note that cohesive energies, i.e., the difference between the solid and
free atoms, are on the order of 4-8 eV, while heats of formations for metals are
significantly smaller, with a large heat being on the order of 1 eV.

23 R. E. Watson, G. W. Fernando, M. Weinert, Y. J. Wang, and J. W. Davenport, Phys. Re~: B, 43, 1455 (1991).
TRANSITION-METALS AND THEIR ALLOYS 13

I I 1 I l 1 1 1 I
A dn
0 d n-1
11 n dn-2 s 2
+ Exper.
10

13,,

"6 /'
e
"~ 5

4-

I 1 I 1 1 1 I I
Y Zr Nb Mo Tc Ru Rh Pd Ag
n-1
FI6.8. The calculated cohesive energies of the 4d transition metals using the d"-:s:, d s, and d n
atomic configurations as reference states. The experimentally based multiplet promotion energy
corrections were used to calculate the difference between the atomic ground state and the center of
gravity for each configuration.

4. SOME OTHER APPLICATIONS OF THE FRIEDEL MODEL

The Friedel model has also been utilized to illuminate transition-metal alloying
trends. VarlTla 24 and Pettifor 25 employed it to describe the heats of formation of the
alloys. It was also utilized in a semi-empirical scheme 262v to predict heats. These
estimates involved inserting Eq. (1), defined for both the elemental solids and

24 C. M. Varma, Solid State Comm., 31,295 (1979) and in Theom" of Alloy Phase Formation, L. H.
Bennett, ed. (AIME, Warrendale, PA, 1980).
25 D. G. Pettifor, Solid State Comm., 28, 621 (1978) and Phys. Rev. Lett., 42, 846 (1979).
26 R. E. Watson and L. H. Bennett, Phys. Rev. Lett.. 43, 1130 (1979).
27 R. E. Watson and L. H. Bennett, CALPHAD, 5, 25 (1981): ibid., 8, 307 (1984).
14 R.E. WATSON AND M. WEINERT

the compound, into Eq. (2). The bandwidth of the compound Was involves both
the d-bands positions and bandwidths attributable to each component separately
in the compound, 28 i.e.,

WAB _ [ 1 2 x ( 1 - x ) ( E A - E B )2 + ( x W a + ( 1 - X ) W B ) -'] i~2 . (3)

Factors, such as the positions, E~ and Es, of the (shifted) d centers of gravity in
the reference elemental solids and their shifts on going to the compounds, are
important to the predictions. The results have semiquantitative success in tracing
the trends in the heats.
Another application of the model is to the relative solubilities of transition-metal
alloys, i.e., A in B versus B in A. In discussing noble metal alloys, Hume-Rothery and
coworkers pioneered the study of the role of valency in alloy phase behavior. They
observed 29 a "relative valency" rule, for example, if lower valence Ag is alloyed with
a higher valence element, say Sb, the higher valence element has greater solubility
in the lower valent metal than vice versa. The rule works well for the select group
of alloys considered by Hume-Rothery, but as he later noted, 3~it must be abandoned
as a general rule. Inspecting the terminal solubilities for a much broader array of
systems, Gschneidner observed 3~ that, more often than not, the rule does not hold.
Goodman assessed 3e the relative solubilities between the transition-metals and, as
seen in Fig. 9, a pattern emerges, namely, that the metal having closer to half-filled
d-bands tends to be the more soluble. It is to be emphasized that not all the attributions
are rigorous since phase diagrams were inspected, usually at elevated temperatures,
32
and assumptions sometimes had to be made when reading them.
To a first approximation, the Friedel model predicts 33 that at dilute concentrations
whichever alloy constituent is closest to having half-filled d-bands has the more
favorable heat of solution, i.e., will be the more soluble. The detailed boundary
between the two regions depends on the treatment of the shifts in the d-band centers
of gravity; plausible treatments of the shifts place the boundary at where it is seen
in Fig. 9. It is important to remember that a number of factors were omitted from
these Friedel model estimates, including atomic size. crystal structure, the question
of how the terminal solubility limits are affected by competition from adjacent
phases of intermediate composition, and the relative elasticities of the two hosts.

28 M. Cyrot and E Cyrot-Lackman, J. Phys. F, 6, 2257 (1976).


29 W. Hume-Rothery, G. W. Mabbott, and K. Channel Evans, Phil. Trans. Royal Sot., l_zmdon Ser. A,
233, 1 (1937).
3o W. Hume-Rothery, R. E. Smallman. and C. W. Haworth, Structure of Metals and Alloys, 5th ed.
(Institute of Metals, London, 1969).
3~ K. A. Gschneidner, Jr., in Theor 3" of Alloy Phase Formation, L. H. Bennett, ed. (AIME, Warrendale,
PA, 1980).
32 D. A. Goodman, L. H. Bennett, and R. E. Watson, Scripta Metail., 17, 91 (1983).
,~3R. E. Watson, L. H. Bennett, and D. A. Goodman, Acta Metail., 31, 1285 (1983).
TRANSITION-METALS AND THEIR ALLOYS 15

ELEMENT B
de d8 d7 d6 d5 d4
Ni Pd Pl Co Rh Ir Fe Ru Os MnTc Re Cr MoW
,,,
V NbTo
T, ,
d3 Zr 0 O O . 0 0 0 0 0 0 0
Hf 9 9 oko,,\\'~..o...o.-!:: o o o o o 9 o 9
v 9 9 9 9 9 OL~'Ox,'~_!:;.0_.];0.0:.0
,._,, 9 0 O!
o 0 i.
To 9 9 9 O O O~t,X~Xq~i.i.:Q._.,Oit.O 0 0
E Or,tO O O 0 O O 0 O O ~ ~ ' X ~ "'
L ~ ~olo 9 9 9 o 9 9 9 o ~ ~ q
E ~LO 9 9 9 9 9 9 9 9
ENd61MrO O O O O 0 0 0
T Tc0 0 9 0 0 9
[
k'x\\~
9 A MORE SOLUBLE INB
A Re 0 O Q 9 9 0 O0 0 B MORE SOLUBLEIN
FeO O I O OI~
d7]RuO O O O 9
Os0 0 0 0 0 0
c~[oo F - - -
~8 ~hlo 9 o /

FIG. 9. Experimentally observed relative solubilities of transition-metal alloys, as indicated by


phase diagrams and other data. The dotted region indicates systems where the alloy constituents are
equally far from having half-filled d bands, while the hatched region indicates a boundary where no
clear bias appears in the solubilities. Some systems are completely miscible at high temperatures,
but display a bias at low temperatures.

Despite these omissions, the model has picked up the main trend. The success of
the Friedel-based estimates suggests that relative size is of little relevance as to
whether A is more soluble in B or vice versa.

5. LATTICE INSTABILITIES, HIGH-TEMPERATURE, AND OTHER


EXCITED PHASES

As noted earlier, the transition-metal elements of the Sc and Ti columns have the
bcc structure at high temperatures. Zener 3z and Friedel 3s noted that the bcc structure
is more openly packed than the fcc or hcp lattices, and thus should have lower

34 C. Zener, Phys. Rev., 71,846 (1947).


35 j. Friedel, J. Phys. (Paris) Lett., 35, L59 (1974).
16 R.E. WATSON AND M. WEINERT

/ ./
" ,,o'"
0 ~
i
9 i
i
0 : "-., ',ly
9 .,.--" ~,,I

9 @
,!- , 9 .,' ic 0
i i

/" ./
fcc bcc

FIG. 10. The Bain path is an (001) tetragonal distortion connecting the fcc (cla = ,4~) and bcc
(c/a = 1) lattices. The solid (dashed) lines denote the conventional fcc (bcc) unit cells. The two
endpoint structures are drawn assuming a volume-conserving distortion. (For clarity, the sites at the
centers of the front faces of the cells have been omitted.)

energy phonons. This implies greater entropy due to phonon excitations and, hence,
the stabilization of the bcc structure at elevated temperatures. This conclusion
follows providing, of course, that the stabilization occurs before the system melts.
There is an alternative class of descriptions which, instead of invoking soft modes,
attributes 36'37"38'39"4~such Martensitic transitions to fluctuations related to anharmo-
nicities. There are also bcc-stabilizing entropy contributions coming from electronic
excitations 41 for the elements in the Sc and Ti columns. A complication in describing
the phonon entropy or developing an anharmonic construct based on phonons is
that the excited bcc phases of these elements are structurally unstable to local
distortions at zero temperature. Also, in the CALculation of PHAse Diagram,
CALPHAD, community 42 the modeling of phase diagrams assumes that excited
phases, such as hcp Mo which enters the construct of the intermediate hcp phases
in Fig. 3, are structurally stable. However, hcp Mo, among others, is unstable.
To investigate such instabilities, consider distortions that take one crystal struc-
ture into another. One can exploit general symmetry arguments 43 to define con-
ditions under which the energy displays an extremum along such a distortion
path. For example, the Bain distortion shown in Fig. 10 is a tetragonal distortion

36p. j. Clapp, Phys. Status Solidi b, 57, 561 (1973).


3v j. A. Krumhansl and R. J. Gooding, Phys. Re~; B, 39. 3047 (1989).
38j. A. Krumhansl, Solid State Commun., 84, 251 (1992).
39 E Lindg~d and O. G. Mouritsen, Phys. Re~; Lett., 57, 2458 (1986).
4~ C. Kerr and M. J. Rave, Phys. Rev. B, 48, 16234 (1993).
41 R. E. Watson and M. Weinert, Phys. Rev. B, 30. 1641 (1984).
4,~L. Kaufman and H. Bernstein, Computer Calculations of Phase Diagrams (Academic Press, New
York, 1970); also see the journal CALPHAD.
43E J. Craievich, M. Weinert, J. M. Sanchez, and R. E. Watson, Phys. Re~; Lett., 72, 3076 (1994).
TRANSITION-METALS AND THEIR ALLOYS 17

that takes the bcc lattice, with c/a - 1, into the fcc structure at c/a = 4r2. The
structural energy of systems constrained to have tetragonal symmetry can be
written in terms of the lengths a and c of the two independent axes. To first order
in the displacements 6a and 6c, a general distortion which maintains the tetrag-
onal symmetry will cause a change in energy of

6U 6U (4)
6 U ( a, c ) = 2-~a 6 a + --~c 6 C.

For a volume-conserving deformation, 6a and 6c are related to first order by

a~c = - 2 c 6 a . (5)

Combining these results yields

6U(a,c) = 26a 6I-~a-Ca 6~]. (6)

For general tetragonal symmetry, there is no simple relationship between 6U/6a


and 6U/6c since there are no symmetry operations connecting the a and c axes.
It then follows that the c/a values for which U is an extremum depend on details
of the system. If, however, there is an increase in the symmetry of the system,
say to cubic, additional relationships between 6 U/6a and 6 U/6c may exist. For
an elemental solid at c/a = 1, the structure is bcc and the symmetry group is Oh.
Then

6U =
6U (7)

and 6 U is zero by symmetry. Similarly, when c/a = 4 ~ , the structure is fcc and

6U 1 6U
6-~ = ,,~ 6a" (8)

Again, 6 U is zero by symmetry. (A different derivation for these cubic cases has
been given by Kraft et al. 44). For an ordered alloy along the same distortion path,

44T. Kraft, R M. Marcus, M. Methfessel, and M. Scheffler, Phys. Rev. B, 48. 5586 (1993).
18 R.E. WATSON AND M. WEINERT

0.5
Mo

E
.-,-a
>.
O

0.0
fcc bcc bcc hcp
L i
1.4 1.2 1.o 0 1
c/a 8
FIG. 11. The calculated relative energies of Mo along the Bain and the bcc ~ hcp distortion paths.

the symmetry at these c/a values depends on the atomic packing. For example,
consider an A B alloy with the A atoms at the corners of the cell and the B at the
centers. For c/a = 1, this is the CsC1 structure which has full cubic symmetry and
hence the energy will again have an extremum. However, for c/a - ,[2, we have
the CuAu-I structure, which is tetragonal due to the layered packing of the atoms
and no such extremum condition holds for the energy.
Similar symmetry observations hold for other distortion paths connecting high
symmetry structures. Not all paths around the high symmetry structures, however,
will show extrema since many of the excited phases are actually saddle points.
The bcc and hcp structures can be connected with a path ~5 involving a base-
centered orthorhombic two-atom unit cell where the distortion is commonly
described as a combination of a bcc N-point T~ phonon and a long-wavelength
shear. A particular choice of path between the bcc and hcp structures can be
defined that maintains certain near-neighbor bondlengths. The energies along both
the Bain and bcc-hcp paths are shown for Mo in Fig. 11. The cell volumes along
the paths are equal to that determined for bcc Mo because the volume conserving
assumption (cf. Eq. (5)) is needed to relate the different structural parameters.
However, if the volume is relaxed, the basic structure is the same since the volume
conserving requirement is only necessary localh' around each high symmetry
structure. The bcc, fcc, and hcp points all show energy extrema along these special
paths, as they must, with the bcc stable and the others unstable to local distortions
(at zero temperature).
This behavior presents difficulties for phase diagram constructs that assume
the excited phases to be locally stable. The situation is even worse here since hcp

45 p. j. Craievich, M. Weinen, J. M. Sanchez, and R. E. Watson, Phys. Rev. B, 55, 787 (1997).
T R A N S I T I O N - M E T A L S AND THEIR ALLOYS 19

1 ' ' 9 I '

0.1
E
o
>.
~D

0.0 bcc--+hcp "~ bcc--> fcc ( Bain )

bcc hcp bcc fcc

0 1 06 I0 14
8 c/a
FIG. 12. The calculated relative energies of Ti along the Bain and the bcc <--+ hcp distortion paths.

Mo is seen to be slightly less bound than fcc" In the CALPHAD modeling as


normally done, the excited elemental phase energies are employed; with the hcp
less bound than the fcc, the intermediate hcp alloy phases of Mo-Ir in Fig. 3 are
then predicted not to occur.
Much smaller bcc-hcp and bcc-fcc energy differences are seen in Fig. 12 for Ti
taken along the same distortion paths. Here it is the low-temperature close-packed
phase which is stable while the high-temperature bcc phase is not. The systems
shown in Figs. 11 and 12 will inevitably show second minima since the number of
minima minus maxima must equal one. In the case of Ti, the second minimum
along the Bain distortion path corresponds to a compressed body-centered tetrag-
onal phase that is encountered in Pa and in Sn under pressure.
The local instability of the bcc phase along the two paths implies that it has
phonons with ~ < 0 along at least these two swaths in k-space. This makes straight-
forward estimates of the phonon entropy difficult. Vibrational entropy, however, is
clearly responsible for the stabilization of the bcc phase.
The Gibbs energy, including the electronic, but not vibrational, entropy, as a
function of temperature for Ti is shown in Fig. 13. At an electronic temperature
of---2525 K, which is above the actual hcp --+ bcc transition temperature, the hcp
phase is no longer a local minimum and b o t h the bcc and hcp phases are mechan-
ically unstable. At 3350 K, the bcc phase is lower in energy but is still locally
unstable. At--4000 K, the bcc phase is finally mechanically stable as well. Thus,
for Ti at least, electronic excitations are sufficient to cause the transition, though
at a temperature well above melting and the observed transition temperature.
The distortion energy of RuNb along the Bain path, Fig. 14, provides an example
of the effect of the symmetry on the shape of the energy curve. RuNb has its
calculated minimum 43 at a c/a value close to the observed one. There is an energy
extremum at the CsC1 (c/a - 1) structure, as required by symmetry: at a c / a
20 R.E. WATSON AND M. WEINERT

9 315K
Ti A 2525K
[] 3 3 5 0 K
0.1 moo 9 4000K

E
o
>
v

0.0 [3s [] [] []
8mmm
8
(ff)oo o o

bcc hcp

o i

FIG. 13. The Gibbs energy of Ti, as a function of temperature, along the bcc ~-> hcp distortion
paths. Electronic, but not vibrational, entropy contributions are included.

0.10
1'
E
o
0.05
>

>,

(1,)
t-
I,,U
0.00

-0.05 ' ' ' ' '


0.8 1.o 1'.2 1'.4
c/a

FIe;. 14. The C alculated 4~" energy of RuNb along the Bain distortion path. The c/a at which the
compound is observed to occur is marked, as well as the CsCl (the maximum, c/a = 1) and the cubic
CuAu-I (no extremum, c/a = ~/2) structures.
TRANSITION-METALS AND THEIR ALLOYS 21

appropriate to the cubic CuAu-I structure (c/a - ,,/5), there is no hint of an extre-
mum. Although often described as CuAu-I, Wigner-Seitz cell constructs (which
will be discussed in Section IV) show the atoms, for the observed c/a, to have the
14 nearest neighbors characteristic of a bcc environment rather than the 12 appro-
priate to the fcc. The majority of the compounds designated "'CuAu-I "~ have c/a
ratios in the vicinity of~/2 and 12-fold coordination. However, RuNb is one of a
number of phases better understood as a tetragonal distortion of the bcc, rather
than the fcc, lattice. If it were not for the presence of the energy maximum, one
would expect an E versus c/a dependence with less curvature and with the mini-
mum shifted to the left: the proximity of the maximum affects the elastic properties
of the compound.
Other compounds that undergo Martensitic transitions can be described simi-
larly. 45 For example, both PtV and PtTi are in the AuCd structure at low tempera-
tures, i.e., have energy minina at the AuCd structure. For PtV, the high-temperature
CuAu-I phase is also a local minimum (locally stable). In contrast, the high-
temperature CsC1 structure for PtTi is locally unstable at zero temperature.
Based on internal energy considerations, the high-temperature bcc phases of
the elemental transition-metals, as well as the high-temperature phases of a
number of compounds undergoing Martensitic transitions, are locally unstable.
Vibrational (and electronic) entropy must be responsible for the stabilization
of these high-temperature phases. However, describing the vibrational entropy
in terms of phonons is problematic since the mechanical instabilities imply
unstable phonons. To date, a completely satisfactory resolution of this problem
is still lacking.

III. Alloying Trends: Electronegativity, Size,


and Valence Electron Count

In the days before the advent of modern-day quantum mechanical calculations,


there were widespread efforts to understand the phase stabilities and properties of
alloys in terms of physical factors such as the relative sizes of the atomic constit-
uents. To this day, the results of both experiment and theory are often discussed in
terms of these factors. "Structural maps," with coordinates such as the relative sizes
of the constituents and their electronegativity differences, are drawn to ascertain
where some type of phase occurs and to predict what yet unobserved phases might
be expected. Sometimes such correlations work well, but often matters are more
complicated than what is assumed in the constructs and in the definitions of the
parameters upon which they are based. Three of the most popular parameters and
their relation to transition-metal alloying are inspected in this section: electroneg-
ativity, size, and valence electron count.
22 R.E. WATSON AND M. WEINERT

6. ELECTRONEGATIVITIES

It has been long recognized that there is polar character in the bonding of solids
and molecules, and a diversity of electronegativity scales have been devised in
an attempt to quantify this effect. In first approximation, the difference in two
atoms' electronegativities is taken to be a measure of the direction and magnitude
of the charge flow, were it to be allowed, between a pair of otherwise noninter-
acting atoms. In the normal scheme of things, the square of an electronegativity
difference is proportional to the energy gain associated with such flow.
Mulliken based his scale 46 on the average of an atom's ionization energy (the
cost to give up an electron) and its electron affinity (the energy associated with
gaining an electron or, equivalently, the ionization energy of the -1 charged ion):

~p.~t = e(0) + e(-1) (9)


2

It so happens that electron affinities are small compared with the ionization
energies and it is common to define this scale in terms of the latter alone. (Note
that the Mulliken scale is already in energy units.) Coulson has argued 47 that this
scale has, in its simplicity, a precise theoretical and experimental foundation.
However, utilization of electronegativities is normally done for atoms involved
in bonding in solids or molecules, and these environments affect charge transfer
behavior. In addition, a Mulliken scale depends on the charge of the atom, on
what sort of valence electrons are involved in the transfer, and arguably also on
the amount of charge transfer. A Mulliken scale, based on the neutral atom
ionization energies, is shown for the 5d transition-metal row in Fig. 15. This row
was chosen because there has been a long-standing issue as to whether gold is,
or is not, the most electronegative of the metals. Since the upper end of the 5d
row is more electronegative than the other transition- and non-transition-metals,
the Mulliken scale would suggest that Au is, indeed, the most electronegative,
i.e., that it attracts valence charge from any other metal.
Perhaps the most famous of the electronegativity scales is that of Pauling, 4s who
employed experimental thermochemical data to estimate the extra ionic contributions
to the bonding between unlike atoms. Inevitably, this scale tended to shift depending
on the data sampled, but it has the virtue of taking its measure from interacting atoms.
Pauling's scale is also displayed in Fig. 15. (One of the complications of comparing

46 R. S. Mulliken, J. Chem. Phys., 46. 497 (1949): H. A. Skinner and H. O. Prichard, Trans. Faraday
Soc., 49, 1254 (1953).
-~7
C. A. Coulson, Valence (Oxford Unix'. Press, London. 1952),
4s L. Pauling, The Nature of the Chemical Bond, 2nd ed./Cornell Univ. Press, Ithaca, 1948): and 3rd
ed. (1960).
TRANSITION-METALS AND THEIR ALLOYS 23

| I 1 I 1 I I' I 1 ~

/ ~7\ /

/+I 7 " ~ 9 Pauling


/ ~ O Miedema
I --d-/.,//~ + Gordy 0
.,-4
>
...
o-f v"
1 I 1 1 I I 1, 1 1,. 1...

0 I I I I 1 1 I 1 1 1
/

If~" 7" 9 Mull/ken


V.//~" / " O d-band hybridization

La Lu Hf Ta W Re Os Ir Pt Au

F~G. 15. Six electronegativity scales, as discussed in the text, plotted fl)r the 5d roy< The various
scales have been rescaled to bring them into registry for purposes of comparison. The dot-dash line
in the bottom plot is a repeat of Pauling's scale, given lk)r comparison.

different sets is that they involve different zeroes and different units depending
on how a given electronegativity scale was derived. For the purpose of compar-
ison, some of the electronegativities in Fig. 15 have been scaled so that their values
for La and/or Pt are set to Pauling's.)
Gordy and Thomas took another path, au namely they expressed the electrone-
gativity as the potential resulting from the unscreened nuclear charge of a bonded
atom as sampled by a valence electron situated at the covalent radius of the atom.
This approach, then, involves establishing rules for the screening of the nuclear
charge by the core and the other valence electrons. They also defined a second
scale simply based on the known work functions, 0, of the metals. Both of these
scales appear in Fig. 15. The work function makes sense in that it is a measure of
the position of the chemical potential within the crystal with respect to the vacuum
zero. Arguably one would prefer the reference to be some "zero" within the
crystal, but there are serious questions as to how to define this zero. An additional
complication is which 0 to use since the work function of metals show a crystal

-~9W. Gordy and W. J. O. Thomas. J. of Chem. Phys.. 24. 439 (1956).


24 R.E. WATSON AND M. WEINERT

face-dependence of-10%. More recently, Miedema devised ~~ an effective Hamil-


tonian for the heats of formation of compounds where one term involved the
difference in electronegativities of the constituents. He started with the experimental
work functions and adjusted them so as to better reproduce the known heats of
formation. This set is also in the figure. The differences relative to Gordy and
Thomas are mostly due to having more modem work function data and, only
secondarily, because of the adjustments.
The final electronegativity scale represented in the figure is based on a band
theory estimate. 52 Consider a test atom inserted into a host metal. The atom has
two electron levels, one at ~E above the host's Fermi level and one ~E below.
Assuming a constant mixing matrix element, the one above hybridizes into the
host's occupied levels and reduces the host's occupied electron count. The hybrid-
ization of the host's unoccupied levels into the one below increases that count.
Available calculated transition-metal densities of states were employed in the
estimates and the results for the 5d row are given in Fig. 15. There is one catch
in the plot: An almost empty d-band metal can only gain charge by such hybrid-
ization and a filled d-band can only lose. This is opposite in sign to what is
plotted. The sign shown in the plot assumes that the non-d conduction electrons
contribute a screening charge flow which is larger than and opposite to that of
the d electrons. This issue will be discussed in a later section.
Of the five classes of electronegativities represented in the figure, three have
gold as the most electronegative of the metals. Otherwise, they are more remark-
able in their similarities than their differences, granted their diverse origins.
One issue confronting treatments of electronegativity scales has already been
encountered in the d-band estimate: What electrons are involved? This issue has
been of particular concern when dealing with main group elements. Moffitt 53
employed a Mulliken scheme in some approximate molecular orbital calculations.
A main concern of his was differentiating between Jr and cr bonding effects. Sub-
54.55.56
sequently, there has been a quantum defect (pseudopotential-type) treatment,
which is a natural descendent of Gordy and Thomas's effective Z approach. In this
newer scheme, separate s and p electronegativity terms are derived and their
average is used as an atom's electronegativity. In addition, their difference is taken

50 E R. deBoer, R. Boom, W. C. M. Mattens, A. R. Miedema, and A. K. Niessen, Cohesion In Metals,


Transition Metal Alloys (North-Holland, Amsterdam, 1988).
51 For a critical review of the Miedema scheme by a metals theorist, see Pettifor. z Also see Williams
et al., ~4 J. R. Chelikowsky, Phys. Rev. B, 25, 6506 (1982). and A. R. Williams, C. D. Gelatt, and
V. L. Moruzzi, Phys. Rev. B, 25, 6509 (1982).
52 R. E. Watson and L. H. Bennett, Phys. Rm: B, 18. 6439 (1978).
53W. Moffit, Proc. Roy. Soc. A, 196, 570 (1949).
54 G. Simons and A. N. Bloch, Phys. Rm: B, 7, 2754 (1973).
55 j. St. John and A. N. Bloch, Phys. Re~: Lett., 33, 1095 (1974).
56 E. S. Machlin, T. R Chow, and J. C. Phillips. Phys. Rev. Lett., 38, 1095 (1977).
TRANSITION-METALS AND THEIR ALLOYS 25

as a measure of relative p versus s bonding tendencies. This difference has been


found useful when rationalizing structural trends in s - p bonded compounds.
Schemes such as these involve single one-electron energies (or ionization energies)
that can introduce 57 a charge dependence to the electronegativity scale.
If charge transfer goes as the difference in electronegativities, then the energy
associated with this transfer goes as the difference squared. Such a term appears
9 "~0 51
in Miedema's remarkably successful empirical Hamiltonlan, " which has been
applied to the heats of formation of the 1:1 compounds, as well as to the energetics
of surfaces, interfaces, and vacancies. (A second term in this Hamiltonian, that
opposes bonding, arises from a Wigner-Seitz-type argument concerning the
energy cost of bringing the charge densities at the surfaces of the elemental solids'
cells to a common density.) Considering the quality of the results relative to the
cost of getting them, Miedema's scheme is an extraordinary "best buy." However,
the calculated heats for the 1:1 compounds are at best semiquantitative, bear little
relation to the electronic structure(s) of the metals, 5~ and upon going to other
composition ratios the scheme falters.

7. ATOMIC SIZE

The relative sizes of the atomic constituents are a factor in determining the com-
pound phases that occur. The measure of these sizes may be the volumes (per atom)
of the elemental solids, the radii of touching spheres in the same solids, or some
other measures such as covalent bond lengths. The appropriate choice depends on
the nature of the compound. If, for example, Si or P is involved in metallic bonding
with transition-metals, then the volume should not be that of covalent elemental
Si or P, but that appropriate to them in metallic bonding. There have been consid-
erable efforts over the years to define such sizes for covalent versus metallic versus
ionic binding.
One application has been Darken and Gurry's consideration 5s of the roles of
differences in size and in electronegativity to the substitutional solubility of
one element in another. The idea was that too large an electronegativity difference
would encourage ordered compound formation and that too large a size difference
would make it energetically difficult for a host to accommodate the impurity.
Using the Goody-Thomas electronegativity scale and 12-fold coordinated metallic
radii as coordinates, Waber e t al. 59 constructed maps where substantial alloy
solubility, defined as greater than 5%, was differentiated from little or no solubility.
In their survey, Waber and coworkers found that size, as an apparent controlling

5vR. E. Watson, L. H. Bennett. and J. W. Davenport. Phys. Rev. B. 27. 6428 (1983t.
58L. Darken and R. W. Gurry, Physical Chemistry o[ Metals t McGraw-Hill, New York. 1953).
59j. T. Waber, K. Gschneidner. A. C. Larson. and M. Y. Price. Trans. Metall. Soc.. 227, 717 i 1963).
26 R . E . WATSON A N D M. W E I N E R T

....
[ ' I ' I l ' t ' I '

4-
E + ~'~ 9 Fe host + 4g + + AI host
5 +
::+
"O

%oo+
+ ~9
"~ 4 ++ 4+
+(3- +
+ o

3
._o + + O +
+ +
++
2 , I0 , i , I , , I , I , I ,
0.5 1
9
1.5 2.0 2.5 0.5 1.() 1.5 2.0 "~"
-.~

Metallic radius (A) Metallic radius (,~)

FIG. 16. Gurry-Darken plots of the solubilities of the elements in Fe and AI hosts, plotted as a
function of electronegativity and size (see text). Circles indicate substantial solubilities of 5% or
more, crosses indicate low or no solubility. The filled circles indicate the host.

factor, was more important than was electronegativity to the occurrence of sub-
stitutional alloys. Two of their cases have been redrawn and appear in Fig. 16,
where the coordinates are now the Miedema electronegativity scale and a set of
metallic radii. 6~ Shown are those cases for which solubility data, including lack
of solubility, existed when Waber et al. did their plots. (The lone exception is Ga
in A1 which Waber et al., indicated to have weak solubility, but which current
phase diagrams show to meet the 5c~ criterion.) In the case of Fe as the solvent,
most of the soluble impurities lie close to Fe: those that are poorly soluble lie
further away. There are only a few truly misplaced cases. The tendency to do a
good job of predicting those impurities that are soluble appears to be typical for
transition-metal hosts. As seen in the other half of Fig. 16, the situation is quite
different for aluminum as the host. The plot is notable in how poorly it does.
The problem is that relying on electronegativity differences alone underestimates
the strength of bonding between main group elements such as AI, Ga, Si, and Ge,
and the transition-metals. Instead of having a substitutional alloy, there is a ten-
dency to have a two-phase mixture of the pure main group element plus a well-
ordered adjacent compound. The tendency to have such extra binding has been
recognized for some time: An extra term to account for this appears in Miedema's
Hamiltonian for those cases when transition and main group elements are involved
in compound formation.
When correlating some effective physical parameter with some experimental
trend, one has to be wary of apparent success: When plotted across the periodic

~) Table of Periodic Properties of ttle Elements (Sargent-Welch. Skokie. IL. 1968).


TRANSITION-METALS AND THEIR ALLOYS 27

, , , i , , , , , , , , , , , , , , 9 , , T' 9 " , ~ 9 , 9 '"'r , 9 , T , , "t


9 Like neighbors
O U n h k e neighbors .]

10
ct} CuAu- I o CrB (Zr Site)

o
..Q
c-
o') 5 <> 9

Z 0
O15
i_

ElO
CsCI CrB (AI site)
Z 0

9 <> 9 09 o~
i , , 1 j ~ ~ t , , , i , , , 1 , i , 1 0 i 1
2 4 6 8 0 2 4 6 8 0
Distance (a.u.) Distance (a.u.)

FIG. 17. The calculated near-neighbor distributions f o r Z r A 1 in t h e CrB, CsC1, and CuAu-I struc-
tures. For the CrB observed phase, t h e Z r a n d A1 s i t e s h a v e d i f f e r e n t environments: for the higher
symmetry CsC1 and CuAu-I structures, both sites have the same distributions of like and unlike
neighbors.

table, two quite different parameters may mimic each other. For example, across
the transition-metals, the inverse of the volume and the electronegativities show
similar behavior. In the case of the Gurry-Darken plots, the choice of volume
makes sense and it is gratifying to see how well it works, although reassuring
that the correlations are less than perfect.
The relative sizes of the constituent atoms is also important to the formation
of ordered alloys. Many workers have been concerned with such issues, notable
among these Pearson. ~ He has considered how constituent size controls unit
cell dimensions in ill-packed structures such as the Laves phases 62 and, more
extensively, how phases are stabilized or destabilized due to the relative sizes of
the constituent atoms.
Consider the competition between structures encountered in the 50/50 com-
pounds of the transition-metals with each other and with the main group elements.
The CsC1 (B2) structure is the most commonly encountered. It is a bcc lattice
with one type of atom at the comers and the other at the body center. Its near-
neighbor radial distribution is plotted in Fig. 17. This structure has eight unlike

6~ W. B. Pearson, Acta Crvst B, 24. 7, 1415 (1968): Phil. Trans. Roy. Sot., 298. 415 (1980): J. Less
Common Metals, 77, 227 (1981): ibid., 81, 161 (1981): Acta Crvst. B, 37, 1174 (1981): J. Less
Common Metals, 80, 59 (1981): J. Phys. Chem. Solids. 43. 547 (1982).
62 W. B. Pearson, Phil. Mag. A, 46, 379 (1982).
63W. B. Pearson, The Crystal Chemistta" attd Physics of Metals (J. W i l e y a n d S o n s . N e w Y o r k . 1972):
a l s o s e e W. B. Pearson, Computer Modelling of Phase Diagrams, L. H. B e n n e t t . e d . ( T h e M e t a l l u r g i c a l
Soc., Warrendale, PA, 1986), p. 163.
28 R.E. WATSON AND M. WEINERT

I
I

0' I

Zr I Al I AI, Zr, CrB


- structure

FI6. 18. ZrA1 in the CrB structure. The lattice consists of AI~ and Zr~ atoms lying in one plane
(filled circles) that alternate with planes of AI, and Zr, atoms I open circles). The solid lines connect
representative Zrl and AI~ sites to their closest neighbors. The two nearest AI neighbors to A1 lie in
adjacent planes, while the 10th and l lth nearest neighbors are also in-plane AI~ atoms.

nearest neighbors and, slightly farther away, six like neighbors. The next shell of
neighbors is much farther away. The plots for this and the other structures are
based on calculations for ZrAI, a system where there is a substantial difference
in size between the constituents. Another common structure is CuAu-I, already
encountered in the discussion of the Bain distortion. This fcc-based lattice distorts
tetragonally because the packing results in the twelve nearest neighbors being
broken up into a set of eight unlike neighbors out of plane and four like neighbors
in-plane. Generally, the unlike out-of-plane neighbors are slightly closer than the
in-plane like neighbors, resulting in a decrease of the c/a ratio from its ideal
value. Both of these structures have the same neighbor distribution functions for
the two species of atoms, and thus cannot accommodate the fact that the smaller
A1 atoms might prefer shorter AI-A1 distances than those appropriate to Zr-Zr.
Neither of the structures occurs for ZrA1. although the CuAu-I is calculated ~ to
be within 0.02 eV/atom of the observed CrB structure and the CsCI structure lies
0.17 eV/atom above.
The CrB is a centered orthorhombic structure with two internal coordinates not
determined by symmetry that define the atomic positions in the unit cell. The
structure for ZrA1 is plotted in Fig. 18. There is the indication of incipient hex-
agonal layer packing, suggesting that the structure may be thought of as "approx-
imately close-packed." The Zr atom has 13 nearest neighbors, with both like and
unlike neighbors at distances essentially those of the CsCI structure. In contrast,
the smaller A1 has 2 very close A1 neighbors and 7 Zr atoms at the CsC1 distance.
In addition, there are 2 more distant A1 atoms that perhaps count as nearest
neighbors. The A1 site is thus either 9- or 11-fold coordinated. (Wigner-Seitz cell
constructs for the 50/50 and metal-rich borides, such as CrB, to be discussed later,

64 M. Alatalo, M. Weinert. and R. E. Watson. Pitvs. Re~: B. 57. R2009 (1998).


TRANSITION-METALS AND THEIR ALLOYS 29

indicate that the small atom site normally has the 9-fold Bernal environment.)
The ability of the CrB structure, and others such as the FeB, to allow close small
atom-small atom approaches, while at the same time maintaining reasonable
unlike neighbor distances, appears to be essential to their occurrence. Differences
in size, as measured by differences in atomic coordination are a feature common
to many ordered alloys.
Changes in size often accompany compound formation. According to Vegard's, or
Haag's, law, atoms maintain their elemental volumes, i.e., the volume of an A~B~_,
compound is

V[AxB~_,] - x V [ A I + (1 - x ) V[B]. (10)

Considering a variety of compounds forming in the CsC1 and NaC1 structures,


Kubaschewski and Alcock observed ~'~ deviations from Vegard's law of as much
as -80% loss in volume (in the alkaline earth oxides). These contractions corre-
lated quite well with the observed heats of formation of the compounds. Much
66.
less severe effects are seen In the 50/50 compounds between transition-metals.
The bulk of these compounds have volumes which are within +1 c~ of the law.
For magnetic alloys such as MnPd, however, there are lattice expansions of as
much as 4%. These arise from the Invar effect, i.e., lattice expansion due to the
presence of magnetic moments. Contractions of 2 to 8% are associated with
strong compound formation when the most electropositive and polarizable of the
transition-metals (e.g., Sc, Y, La, or Lu) are one of the alloy components. Overall,
the observed volume contractions in the nonmagnetic alloys are small and display
some scatter, but fits yield a plausible-looking electronegativity scale 66 across the
transition-metal rows.
Thus, deviations from Vegard's law are signatures of changes in the strength of
bonding and/or of the presence of magnetic moments. It is to be emphasized that
deviations from Vegard's law due to bonding are not simple measures of bonding
strengths because they depend on how the size of the constituent atoms respond
to that bonding; for example, the relatively polarizable Sc, Y, La, and Lu display
relatively large changes in size for given changes in valence electron count com-
pared to atoms at the middle and upper ends of the transition-metal rows.
Although size is an intuitive concept, making measures of size quantitative is
difficult. The sizes of both the elemental and compound unit cells are well defined.
However, separating the volume of the compound into volumes appropriate to
each element can be done in different ways. One approach based on Wigner-Seitz
constructs will be discussed later. In electronic structure calculations, one can use

65O. Kubaschewski and C. B. Alcock, Metallugrical Thermoclwmistrv/Pergamon Press. Oxford, 1979).


66 R. E. Watson and L. H. Bennett, Acta Metall.. 30, 1941 (1982).
30 R.E. WATSON AND M. WEINERT

the virial thereom 67~~ to define the pressure in terms of surface integrals of the
self-consistent wave functions and potential over an atomic Wigner-Seitz cell
boundary. By breaking the unit cell into atomic cells, it is then possible to define
partial atomic pressures. The relative sizes of each atomic cell can be varied until the
atomic partial pressure is zero. Another related approach 6s is to define "virial frag-
ments" around each atomic site by those closed surfaces such that f i - V p ( r ) - 0,
where p(r) is the valence charge density. These surfaces define saddle points in
the density and represent a natural partitioning of a system. Both of these approaches
applied to elemental systems will give the standard values. How well such proce-
dures work for bulk alloys and would agree with each other and with other measures
of size has not been extensively investigated. From what little is known, the differ-
ences can be expected to be measurable and would qualitatively impact conclusions
concerning such matters as charge transfer. ~
In Section V, changes in the charge at atomic sites upon compound formation
will be sampled over Wigner-Seitz spheres. Some of these compounds suffer devi-
ations from Vegard's law. In those cases, where the deviations are not too severe,
the elemental sphere volumes of the two alloy components can be scaled by the
same factor.

8. VALENCE ELECTRON COUNTS AND VALENCE ELECTRON CHARACTER

The phase diagram for the Mo-Ir alloy system shown in Fig. 3 had the same phases
that occurred for the elemental transition-metals at the same band fillings. We also
encountered Friedel's treatment ~ of cohesion in terms of the d-bands alone. Such
observations might seem to suggest that transition-metal trends can be understood
in terms of the d-bands alone. In contrast, Pauling 4s in the 2nd edition of "The
Nature of the Chemical Bond" considered the trend in transition-metal volumes
(shown for the 4d row in Fig. 19) and observed:

It is interesting to note that the minimum in the curve for each period falls at about
the ninth element in the period, and the following explanation of this can be given.

~,7D. A. Liberman. Phys. Re~: B. 3. 2081 (1971t: R. M. Nieminen and C. H. Hodges. J. Phys. F, 6,
573 (1976).
6s S. Srebrenik and R. F. W. Bader, J. Chem. Phys.. 63. 3945 (1974).
69An early experimental example of the problem is provided by J.-L. Staudenmann, R Coppens. and
J. Muller, Solid State Comm.. 19. 29 (1979). The valence electron distribution for V~Si {an A15
superconductor) deduced from x-ray diffraction v~as used to estimate the site electron counts for
different apportioning depending on the relative size of the V and Si atoms. Using conventional Si
and V metallic radii, essentially zero charge transfer was to be inferred, whereas for their preferred
choice of Rv/Rs, ~ 1.5, each Si lost -2 electrons. For a recent discussion of determining valence
charge distributions from diffraction, see P. Coppens. X-ray Charge Densities and Chemical Bonding
(Oxford University Press, New York, 1997).
TRANSITION-METALS AND THEIR ALLOYS 31

I I I I I I I I I I I-

80

e,-,
6O

r
E
= 40

20

0 I I 1 1 1 1 1 1 I 1 I 1

Rb Sr Y Zr Nb Mo Tc Ru Rh Pd Ag Cd

FEG. 19. Crystalline volumes of the 4d transition-metals Rb to Ag.

The orbitals which are available for bond formation by the transition elements are
the outermost d, s, and p orbitals, which are nine in number: and with nine orbitals
available per atom the maximum number of bonds can be formed, since with fewer
electrons there is a deficiency of bond electrons and with more a deficiency of bond
orbitals, some nine orbitals being necessarily occupied by unshared pairs.

Although modified somewhat in later editions, the argument continues to


invoke a hybridized dSs~p ~ composite with the more extended s and p important
to the bonding. This view has had credence in the metallurgical community.
While the non-d wave function character is of significance in transition-metal
formation, the non-d counts yielded by present-day electronic structure calcu-
lations are markedly less than those expected from Pauling's arguments, as can
be seen from the partial densities of states (calculated within non-overlapping
atomic spheres) for hcp Ru shown in Fig. 20. The total number of s + p + d
electrons seen here is less than that in Fig. 2 because the interstitial region is
omitted from the count and the high ( components of the charge within the
spheres are not shown here. (These high ~ components will later be seen to be
significant to any argument concerning the nature and amount of charge transfer
upon alloy formation.) Although exact values of the local partial density of states
depend somewhat on the details of the calculations, for example the choice of
sphere radius, the general conclusions about the relative weights of different ~'
components are unaffected. The problem with Pauling's observation is that lattice
volumes are not determined by the strength of bonding alone.
There are several features of the Ru partial densities of states which are typical
of the transition-metals in general. First, the density of states is dominated by
the d component. The onset of the conduction bands has s-like character below
32 R.E. WATSONAND M. WEINERT

I
......... p

~ d

N1

--|---I"'i

-8 -6 -4 -2 0 2
E n e r g y (eV)

Fie. 20. The s, p, and d partial densities of states within the atomic (muffin-tin) sphere for hcp
Ru. These are taken from the same calculation that gave the total density of states appearing in Fig. 2.

the d-bands, followed by significant d-non-d hybridization at the bottom of the


d-bands. Hybridization has largely swept out the non-d character from the upper
reaches of the high d density of states, and there is some attendant d character
in the states above. It is easy to place the top of the "d" bands; in some cases it
is more useful to account for the d-band occupancy by counting hole states down
from the top of the d-bands. Such a count is, of course, a count of hybridized
band states and n o t a pure d electron count. Finally, there is no hint of Pauling's
s-p3-d 5 hybridized orbitals. Instead, there is of the order of one non-d electron in
the states below the top of the d-bands. Even at something like one in number,
the non-d wave function character plays an observable role in alloying.
The Mo-Ir system represented in Fig. 3 involves moderate binding; stronger
binding would be manifested by line compound formation. Binding favoring com-
pound formation is expected to be maximized by having the maximum number of
unlike nearest neighbors. For a disordered system, and to some extent for ordered
ones, the number of unlike nearest atom pairs goes as x(1 - x), where x is the
fraction of one component, with the maximum at a 50/50 composition. The Pd-Zr
system, whose phase diagram is shown in Fig. 21, is measurably more strongly
bound than Mo-Ir. Yet its maximum bonding, as manifested by the melting tem-
peratures, is skewed well away from 50/50. This behavior follows from bonding
arguments v~like those encountered in Section 1: Metals from the Sc and Ti columns
have their d-band centers of gravity lying some 4-5 eV higher than those of elements

v0R. E. Watson, M. Weinert. J. W. Davenport, and G. W. Femando. Phys. Rex'. B. 39. 10761 (1989).
TRANSITION-METALS AND THEIR ALLOYS 33

20001
/ ,~PdaZr Pdzzrzr-PdSystem' I
1
1800~
1600~,~
/'/ ! \ / Pd,,Zr~
~Pd'~il ~Zr //I J!

1400 PdZr2
oc
1200

1000

800 (z

6000 10 20 30 40 50 60 70 80 90 100
Pd at % Zr Zr

Fxo. 21. The Pd-Zr phase diagram."

Pd3Zr Total

>
if)
0

.,..., 4
if)
v
tar)
.,..,
2
0

if)
r- 0
a
4
Pd

0
--'0 -5 0 5
Energy (eV)

FIG. 22. The total (per atom) and local densities of states for Pd3Zr.
34 R.E. WATSONAND M. WEINERT

from the righthand end of the row. The lower lying "'Pd'" bands do not have enough
holes, about a half an electron worth in elemental Pd, to accommodate the Zr
valence electrons. As can be seen in Fig. 22, with three Pd atoms to each Zr, the
Zr charge can be accommodated in the bonding levels, primarily Pd d in character,
and the higher lying antibonding levels are empty. The filling of the Pd holes arises
primarily from the hybridization of Zr character into the low lying bands, and not
much from any net transfer of charge. The extent to which bonding states are filled
and antibonding ones are empty, is often critical to the competition among alloy
phases of differing composition or among different structures at the same compo-
sition. For example, if Pd is replaced by Ru, each Ru can accommodate the electrons
from one Zr, with the result that (except for the terminal Ru phase) the Ru-Zr alloy
system has its maximum melting temperature at 50/50 composition.

IV. Wigner-Seitz Cells, the Frank-Kasper Phases,


and the Bernal Environments

9. WIGNER-SEITZ CELLS (VORONOI OR DIRICHLET POLYHEDRA)

The Wigner-Seitz cell about some lattice point is that region which is closer to
that point than to any other. When Wigner and Seitz considered 7 bcc sodium, with
only one atom in the primitive cell, the Wigner-Seitz cell appropriate to the lattice
was identical to the cell constructed from perpendicular bisecting planes associated
with lines connecting an atom to all others. When there is more than one atom in
the primitive cell, the situation is. of course, different. The Wigner-Seitz cell drawn
for the full lattice cell will reflect the lattice symmetry. Those drawn for individual
atomic sites will be smaller and, as a rule, have lower symmetry than that of the
lattice. Going to reciprocal space, the first Brillouin zone involves the same construct
and thus is the Wigner-Seitz cell of the associated reciprocal lattice. 7.
Of concern in this section will be the topologies of the cells of individual
atomic sites, but these cells also are intimately connected to questions of size
as discussed above. The topology of such a cell involves the number of facets,
the number of edges on a given facet, and the arrangement of the facets with
respect to one another. Sites having different site symmetries, in the same or
different lattices, may have topologically identical Wigner-Seitz cells, albeit some-
what distorted, and sites of the same symmetry may have topologically different
cells. As an example of the latter, consider RuNb in the CuAu-I structure, which
was discussed in Sec. 5. In its ideal cubic form, the system consists of alternating

7. For further discussion of this issue, see, e.g.. N. W. Ashcrofl and N. D. Mermin. Solid State Physics
(Saunders College. Philadelphia, 1976).
TRANSITION-METALS AND THEIR ALLOYS 35

planes of A and B atoms along the z direction on an fcc lattice. Because of the
atomic packing, the symmetry is tetragonal and distorts along the Bain distortion
line. If the c/a ratio of the lattice is sufficiently close to the fcc one, the Wigner-
Seitz cell will have the 12 facets (neighbors) characteristic of the fcc lattice. If,
however, as for RuNb, the c/a is in the vicinity of that appropriate to a bcc lattice,
there will be 14 facets, characteristic of bcc packing. Thus, alloys of the same
nominal structure--and with identical site symmetries--have differing local
environments because of differing values of c/a. The difference between the 12-
fold and 14-fold environments for the CuAu-I structure imply different band
structure character. Conversely, metalloids involved in transition-metal-rich alloy
formation can occupy sites of different symmetries, but show a strong preference
for Wigner-Seitz cells of some particular topology.
What information is indicated by the cell? Of course, the cell has the symmetry
of the site. Its facets indicate which atoms are nearest neighbors and the number
of facets, the number of nearest neighbors. The respective areas of the facets
provide some suggestion of the relative importance to bonding of the neighbors.
The number of edges on a facet, q, usually 4, 5, or 6, indicate the number of
nearest neighbors shared by a near-neighbor pair. All things being equal, the larger
the number of shared neighbors, the larger the circumference of the ring they will
form and the closer the two atoms sharing them may lie. Both the closeness of
approach and the number of shared common neighbors can have consequences
for the bonding and for any magnetic coupling. The number of facet edges,
averaged over all the facets of all the Wigner-Seitz cells appropriate to some given
structure, has been used by Nelson and Spaepen 7273 as a measure of ordering in
the Frank-Kasper phases and related glass systems, and will be discussed in the
next section.
The traditional Wigner-Seitz cell construct that uses perpendicular bisecting
planes assumes constituent atoms of equal size. As a rule this is not the case in a
compound and a "radical construct" has been devised 747~76 to deal with this situ-
ation. The scheme involves 7v assigning spheres, with radii reflecting assigned sizes,
to the different atoms. Then points are determined where three or more neighboring

7: D. R. Nelson, Phys. Rm" B, 28, 5515 (1983).


73 D. R. Nelson and F. Spaepen, Solid State Physics, 42, 1 (1989).
74 W. Fischer, E. Koch, and E. Hellner, Neues. Jahrb. Mineral. Monatsh., 227 (1971).
75 j. L. Finney, J. Comput. Phys., 32, 137 (1979).
76 B. J. Gelattly and J. L. Finney. J. Non-Crvst. Solids, 50, 313 (1982).
77To be more explicit, given spheres of chosen radius about each site. a "'radical plane" is defined
between the center atom and a neighbor by where tangent lines of common length from the two
spheres touch. If the sphere radii happen to be chosen so that the spheres touch, then the plane will
pass through that point. Comers of the Wigner-Seitz cell are where three (or more) planes intersect
and these comers are easy to locate. The scheme picks up facets even when they do not intersect the
line between the two neighbors. The resulting cells are volume filling.
36 R.E. WATSON AND M. WEINERT

sites share a common comer of the center atom's cell. The construct naturally picks
up facets which do not intersect the line between a pair of neighbors: If these are
neglected, the set of constructed cells will not be volume filling. If three neighbors
are responsible for some comer, the comer will be said to be "'triangulated" and
will have three edges going into it. Most Wigner-Seitz cells are made up of
triangulated comers. On the other hand, four (or more) edges arise from four (or
more) coplanar nearest neighbors, vs Such is the case for the fcc cell, with the
consequence that an infinitesimal shear will break up this coplanarity and the cell
will go from having 12 to 14 facets (nearest neighbors). A number of cells are
illustrated in Fig. 23. The notation ( a , b , c . . . . ) indicates a cell having a three-
edged, b four-edged, c five-edged .... facets. Only occasionally is there more than
one possible arrangement encountered for a given set of facets. Wigner-Seitz cells
for structures based on the fcc, bcc, and hcp lattices tend to be constructed of
4- and 6-edged facets. Those of the topologically close-packed Frank-Rasper
phases are predominantly made up of 5-edged facets, with some cells having a
few 6-edged ones.

10. THE FRANK-RASPER PHASES AND CLOSELY RELATED STRUCTURES

Obviously, the individual cell volumes and the number of facets may change with
changes in the assigned atomic radii. Some cells are stable in the number and
arrangement of facets over chemically reasonable ranges of choices of radii. Such
"topological stability" is characteristic of the Wigner-Seitz cell assignments for the
Frank-Rasper phases. These phases are an important class of systems in which
transition-metal alloys form. Unlike the fcc, bcc, and hcp structures, where atoms lie
in lattice planes, those in the Frank-Rasper structures do not, with the consequence
is that they tend to be brittle. These structures have environments with differing atomic
coordination and readily accommodate, in fact need, atomic constituents of differing
size. Included in this class of systems are an important group of hard magnets and
the A 15 materials, such as Nb)Sn, which was a high-temperature superconductor prior
to the advent of the cuprates.
Noting the existence of some of these phases, Frank and Rasper devised 79 a
construct involving the (0,0,12,0), (0,0,12,2), (0,0,12,3), and (0,0,12,4) Wigner-
Seitz cells. (A (0,0,12,1) is topologically impossible.) Using these as building
blocks they constructed this important class of structures--both observed and
hypothetical.

78 Actually, J. D. Bernal (private communication) has pointed out that there are occasions when four
or more facets will intersect at a common vertex even though there are not four (or more) coplanar
neighbors. Coplanarity is expected to be the normal source of such non-triangulation.
79 F. C. Frank and J. S. Rasper, Acta Crvst., 11, 184 (1958): ibid., 12, 483 (1959).
TRANSITION-METALS AND THEIR ALLOYS 37

Frank-Kasper Bemal

(0,0,12,0)

(0,4,4)

fee
(0,12,0) (0,0,12,2)

(0,3,6)

bcc (0,0,12,3)
(0,6,0,8)
(0,2,8)

(0,0,12,4)
FIG. 23. The Wigner-Seitz cells for the fcc and bcc lattices, and the Frank-Kasper and Bernal
environments. The latter are somewhat distorted because the,, v,ere derived for actual sites in crystals,
e.g., the (0,2,8,0) is for P in V3P and the (0,0,12,0) and the (0,0,12,21 are for Fe sites in Nd-Fe systems.
See the text for the definition of the labels.

While considering order in glasses, Nelson 7273 made several observations con-
cerning tetrahedral, or topologically, close-packed ordering, which are relevant to
the Frank-Kasper phases. Consider a pair of atoms, with a bond length L between
them. Let these share other near neighbors with the same bond lengths L. If two
such neighbors are added, one has a tetrahedron. Adding a third adds another
tetrahedron with the initial bond line shared as a common edge. The number of
tetrahedra which may share a common edge is

q..p = 2 n ' / c o s - l ( 1 / 3 ) - 5.10429 .... (11)

which is a non-integral number. Thus, such packing is frustrated in real (flat) space
though, as Nelson observes, it can be accommodated in suitably curved space.
38 R . E . W A T S O N A N D M. W E I N E R T

, !

o A15
5.110
+
+
v (Mnfii) O
(lD +
+
5.105
-+t-
"~ O (D
qtcp
Mg~AI,,ZnI,O
5.100 O(133
Laves O Observed
+ Hypothetical

5.095 , I , 1 , L ,
0 50 100 150 200
A t o m s / u n i t cell

FIG. 24. A " N e l s o n " plot of (q), the average number of common nearest neighbors on a bond line,
as a function of unit cell size for both the observed and the hypothetical Frank-Kasper phases, q,,.p
is the value for tetrahedral close-packing.

He goes on and notes that this frustration can be partially accommodated in a


structure in real space by starting with 5-fold bond lines and introducing an
occasional 6-fold,-72 ~ disclination, a disclination being a rotational defect. Fur-
ther, the average numbers of shared neighbors, (q), in the Frank-Kasper phases are

5.1 < (q) < 5 . 1 1 1 1 1 1 .... (12)

9 8o
which bracket q~,,cvery nicely. Thus, gi noting their crystallinity, the Frank-Kasper
8~
phases may be deemed close relatives of a tetrahedrally ordered "ideal glass" in
the sense of having disclinations. A Nelson map, indicating (q) as a function of
unit cell size for the observed Frank-Kasper structures, and for the hypothetical,
but as yet not observed, structures constructed by them, appears in Fig. 24. There
is some suggestion that the values of (q) approach q,,r, as the unit cell becomes
larger, implying that the larger unit cell structures might be better approximations
to the ideal glass.
As already noted, the Frank-Kasper building blocks involve zero, two, three,
or four 6-fold bond lines or disclinations. The (0,0,12,2), has two bond lines lying
in an approximately straight line; the (0,0,12,3) has the three bond lines 120 ~
away from each other in a plane; and the (0,0,12,4) has them arranged tetrahe-
drally. Since there are no sites with a single disclination, the disclinations do not
halt at a site, but instead thread through the crystal. For example, the Cr3Si (A15)
structure in which Nb3Sn forms is constructed from (0,0,12,0) and (0,0,12,2)

80 For comparison, (q) is 4 and 5.14 for the fcc and bcc structures, respectively.
sl Nelson attributes this observation to an unpublished report of J. F. Sadoc.
TRANSITION-METALS AND THEIR ALLOYS 39

1 " T ;

24 Mn
4
u')
n-
l 11/Ht~
o 4- 24 Mn TIT the 14th neighbor
rn
"I-
{..9
I.d _
z 8 MnTT !
u. 4 -
o
n- ., Aq

m12-

Z 8- 2MnI

4-

l I I I t

0.25 0.30 0.35 0.40 0.45


NEIGHBOR D I S T A N C E , nm

FIG. 25. The near-neighbor distribution functions for the various sites in a' Mn.

cells, and the disclination lines associated with the (0.0.12.2) extend in the x. v,
or z directions across the crystal, s-~
The c~ Mn phase occurs in the phase diagram displayed in Fig. 3. This phase
is often listed as a Frank-Kasper system, though for years there has been a question
of whether it is such a phase. The o~Mn structure involves four different sites. The
problem is associated with the Mn III site and its near-neighbor distribution
function, as illustrated in Fig. 25. It apparently has 13 nearest neighbors, and hence
the structure would not be a Frank-Kasper phase. A Wigner-Seitz cell construct

82 The chains involve compressed lines of the larger, more distortable, atoms do,xn to a very close
approach given their metallic radii. These nearest-neighbor chains are considered to be essential to
the superconductivity in these systems. A variety of other crystal structures have /0.0.12.2) chains.
although they do not otherwise meet the criteria for being Frank-Kasper phases. Among these are
the Mn3Si.~ and W,~Si3 structures, with compounds such as Pt~Zr~ and Ir~Zr~ tk)rming in the Mnr
structure. The chains lie in a single direction in these structures, as they do for Rh4Zr~ (the Pd4Pu~
structure) where it is the smaller atom that now lies in the compressed (0.0.12.2) lines. This system
[L. H. Bennett, R. M, Waterstrat. L. J. Sv~artzendruber. L. A. Bendersky. H. J. Brown. and R. E.
Watson, J. Appl. Phys. 87. 6016 (2000)] shows an incommensurate density wave normal to the chains
and is a weak paramagnet. Replacing 1 or 2% of the Rh by Pd causes the density wave to disappear.
but the addition of another 2% causes it to reappear and the system becomes a weak room temperature
ferromagnet. Increasing the Pd concentration to 6c~ sees the disappearance of ferromagnetism and
the appearance of weak superconductivity. Yet more Pd leads to the disappearance of both the density
wave and the superconductivity. Such rapid changes in cooperative phenolnena with concentration
are unusual in metallic systems. Structures with chain.,, involving both alloy constituents are also
possible and might lead to interesting properties.
40 R.E. WATSON AND M. WEINERT

indicates 83 the site to be 14-fold coordinated, but with an (0,1,10,3) environment


where each Mn III is connected to another by a 4-fold, +72 ~ disclination line.
As is indicated in Table 1, Mn I and Mn II have (0,0,12,4) environments and the
Mn IV has (0,0,12,0). The sites thus range from 12- to 16-fold. The atoms adjust
to the differences in site volumes by assuming different magnetic moments. The
(0,0,12,0) has an essentially zero-valued moment while those of the 16-fold sites
are substantial. It would appear that a Mn is in some senses an "alloy," but one
involving varying magnetic moments rather than differing atomic charge. Despite
the presence of a few 4-fold disclinations, (q) = 5.124, a value close to, but larger
than qtcp.
There are other systems that bear a resemblance to the Frank-Kasper phases,
including such hard magnets s4 as SmCo5 (the CaCu5 structure), Nd2Fel7, and
NdzFe14B. 85 A tabulation of the Wigner-Seitz cells for a number of the 3d-rare
earth and 3d-actinide systems also appears in Table 1. A Nelson plot (Fig. 26) of
(q) for some of these structures shows that they are close to the tcp value. In these
systems, the large rare-earth atoms lie in 20- or higher-fold environments, mostly
(0,0,12,8) or (0,0,16,4). The 3d atoms are usually in Frank-Kasper Wigner-Seitz
cells, but a 13-fold (0,1,10,2) environment does crop up.
The (0,1,10,2) is related to its Frank-Kasper neighbors: If one takes an (0,0,12,0)
environment and brings an extra atom in between several of the existing neighbors
such that a pair of comers connected by a common edge of the Wigner-Seitz cell
are cut out and replaced by a new facet, the resulting environment is the (0,1,10,2). 86
The 12 atoms already present, of course, relax their positions in the presence of
the new neighbor while retaining the topology of the new cell. If this cell is similarly
cut by the introduction of yet another neighbor, the (0,0,12,2) arises and, with
additional similar cuts, one has the sequence:

(0,0,12,0) ~ (0,1,10,2) --~ (0,0,12,2) ---) (0,0,12,3) --) (0,0,12,4).

The (0,1,10,2) site, with its short-ranged 4-fold +disclination lines that begin and
end at these types of sites, would appear to be the missing, topologically possible,

83 R. E. Watson and L. H. Bennett, Scripta Metall. 19, 535 (1985).


84 For a discussion of the principal disclination nets associated with some of these systems and their
relation to easy axes of magnetization, see R. E. Watson, L. H. Bennett, and M. Melamud, J. Appl.
Phys., 63, 3136 (1988).
85 For a review of the important R2Fe~4B class of materials, see J. F. Herbst, Rev. Mod. Phys., 63,
819 (1991).
86 Alternatively, the (0,1,10,2) cell can be constructed by cutting out a pair of edges which share
a common corner in the (0,0,12,0) cell. In the (0,1,10.2) cell, the single 4-fold facet has the two
6-fold facets as neighbors on opposite edges. Similarly, in the Bernal-connected (0,2,8,1) cell, the
two 4-fold facets adjoin opposite edges of the single 6-fold facet. A cut, involving two edges (each
ending at one of the 4-fold facets) and sharing a common corner, turns the (0,2,8,1) into the Frank-
Kasper (0,0,12,0).
,-t

0 2: Z ~ 0
d~
,_..
,.-. ..
..
uu ~ q
t~

:Z
c"" a,,
r ~ 9
9 9

rrl
..-I
~ N =~
~ fJ

,.....
- ~

~_. >
"4. "/.

",I..

r~ 0

.7~

~.~ ~ 4-- ~,..~ I,~ ,'--' ',.~,~ ~ '--"

rrl

A
"4.
S.o o o o o o o o o o o o o o o o o
;,., - . ~ .o .~ -. o i.-, ~
-. o F- ~. o.o b .- o
OC~

"H
42 R.E. WATSON AND M. WEINERT

TABLE 1. THE WIGNER-SEITZCELLS OBrAINEI)*a FOR SOME FRANK-KASPERAND RELATED


PHASES. THE ATOMS AND THEIR SITE DESIGNATIONSARE INI)ICATED(CONTINUED)

Phases with high-coordination large-atom sites


ThMn~2 Th 2a (0.0.16.4)
MnI 8f (0.0.12.0)
MnlI 8i (0.0.12.2)
MnIII 8j (0.0.12.0)
BaCdll (e.g.. NdCo,Si:) Ba 4a (0.0.20.2)
CdI 4h (0.0.12.2)
CdII 32i (0.0.12.0)
CdllI 8d (0.0.12.0)
Others Ce:Mn.Nis. CezNi~-Si~

|
~0 '
Mn

5.110
0 cMg:,AI,,,

5.105
qtcp
0 Nd.Fe._
O
Be,_Ru,
5.100 EX)aX3D O
\
CaCu, 0 Nd:Fe: B

5.095 , l , l , l l

0 50 100 150 200


Atoms/unit cell

FIG. 26. Plot of the average number of common nearest neighbors on a bond line. (q). as a function
of unit cell size for observed phases closely related to the Frank-Kasper structures, q,,r, is the count
for tetrahedral close-packing.

13-fold member of the Frank-Kasper sequence. This environment does occur in


crystal structures bearing close resemblance to the Frank-Kasper phases, as can
be seen in Table 1.
Granted the substantial variation in the coordination numbers with the different
sites in these structures, one expects that the relative sizes of the atomic constituents
plays a significant role in compound formation, as discussed in Section 7. This
conjecture is tested s7 for the Laves structures, with their 12- and 16-fold environ-
ments, in the "structural map" in Fig. 27. The coordinates are the ratios of the
volume of the minority elemental metal divided by that of the majority and an

s7 R. E. Watson and L. H. Bennett. Acta Metall.. 32. 477 ( 1984): ibid.. 32. 491 (1984).
TRANSITION-METALS AND THEIR ALLOYS 43

_
1 I 1 1 I

I
I o I
I I
I
I ~
Mn2La Ix
Io
3.0 I
I ~
I
I
o i
I o o I x
I o oo o o
x
c
I
I o I
2.5 x+ I 0
,j . . . . O~ I X
Pt2Lo I co
0
X IO O 9 i
9 I
xx ~ o I
X I 0 0
~" 2.0 x o o." oi I
Rh2Sc /D :|
I *o,; 'I ~ x ~*
~o I ,, 9 ; I V2Zr
; J o xx
t i ~x o
# #~ V2Hf _
1.5
)x
x
x

15
[I V2To xx
x
x :~ XX
X

1.0 N x X ~r -

X X

- [ I 1 I 1

2 3 4 5 6
Nh

FIG. 27. The occurrence and non-occurrence of AB 2 transition-metal Laves phases as a function
of the volume ratio V(A)/V(B) and the effective d-band hole count. denote the cases for which Laves
phases have not been reported to occur (the two triangles indicate Frank-Kasper phases which occur
at AB2 composition but in non-Laves structures). Occurrences are indicated by open circles for the
MgCu2 structure, closed circles for the MgZn2 and half-filled circles when both occur. A atoms were
taken from the Sc, Ti, V, and Cr columns and the B from the V. Cr. Mn. Fe. Co. Ni, and Cu columns
of the transition-metal elements.

average effective d-band hole count. A circle denotes a reported 2"1 Laves phase,
and a x either that a 2"1 phase has not been reported or it occurs in a different
structure. It appears that the Laves phases occur only if the v o l u m e ratio is in excess
of 1.34 and only within a well-defined range of d-band hole counts. The errors are
few, most notably those of V with Ta, Hf, and Zr. Rh_~Sc is another apparent error.
A l t h o u g h Rh2Sc has not been reported experimentally and might be losing out to
the stronger c o m b i n e d binding of adjacent RhSc and Rh3Sc c o m p o u n d s , it w o u l d
be surprising if it does not exist as a stable Laves phase. Inside the region w h e r e
the Laves phases do exist, there is a well-defined pattern for the occurrence o f the
44 R.E. WATSON AND M. WEINERT

competing Laves structures. Johannes and coworkers ss have very satisfactorily


reproduced this pattern using a model d-band Hamiltonian.
The Laves structures are the one group of topologically close-packed phases
where the small 12-fold sites are in a substantial majority. Having only 12- and
16-fold sites of quite different size, the Laves structures appear at 2:1 stoichiom-
etry with little or no range in composition. The situation is quite different for the
other Frank-Kasper phases, a Mn included. They have cells of intermediate, i.e.,
14- and/or 15-fold, coordination and can accommodate constituents that differ
only moderately in size. These systems thus can have a range in stoichiometry, as
seen for Mo-Ir in Fig. 3. In addition, more than one such phase may occur within
the stoichiometry range. The ranges observed in the binary transition-metal alloys,
excluding those containing Mn, appear in Fig. 28. The plot has the same abscissa
as that used in Fig. 27. To the fight are shown the volume ratios V(A)/V(B), where
the A atom is the one to the left in the periodic table. When 1.2 < V(A)/V(B) < 1.4,
the formation of these non-Laves phases is particularly favorable. Larger compo-
sition ranges occur for such volume ratios. In the cases of Fe-Tc and Fe-Re at the
bottom, the favorable volume ratios appear to allow the occurrence of such phases
at what are apparently less than optimum average d-band fillings.
It appears that an interplay of volume and d-band fillings is essential to the
appearance of the transition-metal Frank-Kasper phases. Not surprisingly, granted
the Wigner-Seitz cells involved, the Laves phases require larger volume ratios. This
difference in site volumes, in turn, implies that the systems cannot readily accom-
modate anti-site defects and that the phases occur at or close to ideal stoichiometry.
On going from one compound to another, they may--and do--appear over a wider
range of average d-band fillings than do the other Frank-Kasper phases.
This class of structures is not limited to the transition, rare-earth, and actinide
metals; a number of binary and ternary alloys involving the transition-metals with
main group elements, in particular with A1 and Si, also occur. These also may be
placed on plots similar to Fig. 28, provided that fictional d electron counts roughly
equal to that of Co are arbitrarily assigned s to the main group elements.

11. THE BERNAL ENVIRONMENTS

/ 11 9o,91
Bernal in his classic modeling of glassy systems with clay oans, observed inter-
stices having the (0,4,4,0), (0,3,6,0), and (0,2,8,0) Wigner-Seitz cells shown in Fig. 23.
These had been known to crystallographers for some time as metalloid environments

ss R. L. Johannes, R. Haydock, and V. Heine, Phys. Rm: Lett.. 36, 372 (1976).
89 L. H. Bennett and R. E. Watson. Hi~h-Temperature Alloys." Theory and Design, J. O. Steiger, ed.
(The Metallurgical Soc., Warrendale. PA. 19841, p. 103, in particular see Fig. 3.
90 j. D. Bernal, Nature. 185, 68 (1960).
,l j. D. Bernal, Proc. Roy. Soc. London. A 280, 299 t1964).
-.z oO z ~ o ~ # _ ~ ~ 9~ ~ ~> ~_ ~ ~ ~
< < -.
o ~ o ~ -.~ ~ . . . .
> _ ~ . ~ ~, I_._..J I ~ ~ L - - - JL
_ Jl IL
. . . .
~ i,.., ~ g ,
~_~ A m @ ~'~ ~~"
~~~ I

A ~" ~
I
I
I , II
~--2~ 9
~.~
,.~ ~.
z,~.
:=, I' I i
I,I I 9 I iI 9
~'~'~.~._ I

~ -~ E.~

,-""

'-~ ~ 0

~ ~ ~.~ ~ O0 9 9 9 9 9 9 9 9 9 9 9

"" ~ 0 oO 0 0 O0~ 0 0 O0 O0
O0
9 O~ r , o 9 9 9 9 9 9
~ O 9 9 oO 9
OO . . . . . 9
- l o g - "
O 9
DO ol 9
,.... ,.~-% F I~

r '1

< ~ ~..q" # ~ ~o o ~ ~ o ~ :~ o E :~ ~ z ,:,- ~ z ~ < z o~ N - ~ ~"'


46 R.E. WATSONAND M. WEINERT

in transition-metal-metalloid crystalline compounds, and also appear to be relevant


to the modeling of metal-metalloid glasses. Warren in his early x-ray diffraction
studies of glasses invoked 92 a "quasicrystalline" model for glasses, long before such
a term was applied to icosahedral phases. More recently, for example, Gaskel193 has
constructed two models of the amorphous Pd-Si system where the basic building
block is the (0,3,6,0) Si environment found in the Pd-rich crystalline compounds.
The Bernal environments are, in some senses, the "mirror images" of the
Frank-Kasper ones. The (0,4,4,0) has four 4-fold bond lines, or in Nelson's
observation, 7273 +72 ~ disclinations. These, like the f o u r - d i s c l i n a t i o n s of the
(0,0,12,4) cell are arranged roughly tetrahedrally. The three +disclinations of the
(0,3,6,0), like the three of the ( 0 , 0 , 1 2 , 3 ) , are coplanar and roughly 120 ~ apart,
and the two of the (0,2,8,0) are linear like the two of the (0,0,12,2). The I 1-fold
environment in the Bernal sequence might be expected to be an (0,1,10,0), but
like the (0,0,12,1) is topologically impossible. Cutting the (0,2,8,0), as to add
another neighbor, can yield s6 an (0,2,8,1). There is a series of cuts that takes the
(0,4,4,0) to the (0,0,12,4). This sequence of Wigner-Seitz cells is: (0,4,4,0) ---)
((0,3,6,0) --) (0,2,8,0) -~ (0,2,8,1) .-) (0.0,12.0) ---) (0.1,10,2) --) (0,0.12,2) ---)
(0,0,12,3) ~ (0,0,12,4).
In the transition-metal-rich transition-metal-metalloid compounds, the metal-
loid can be expected to have transition-metal nearest neighbors, with the number
of neighbors reflecting the relative sizes of the constituents. Consider N, C. B, P.
and Si 94 in transition-metal rich systems:

N a n d C: N is always to be found in the octahedral (0,6,0,0) environment, as is


carbon when the 4d and 5d elements are involved. C tends to be in the Bernal
(0,3,6,0) environment when involved with the smaller 3d elements in the middle
of the row and in the (0,6,0,0) at the beginning and end of the row. However,
when alloyed with Sc, C is in the (0,4,4,0) environment.

B: Boron is normally in a Bernal environment, particularly the (0,3,6.0), and only


once appears in an octahedral environment. The strong preference for the (0,3,6,0)
has some consequences: In the rare-earth-3d hard magnet systems, the metals are
normally arranged in planar arrays. In Nd2Fej4B (see Table 1), however, the B and
its (0,3,6,0) environment jams the structure, resulting 9~~6in corrugated atomic layers.

P: P is generally found in the Bernal environments, most commonly the (0,3,6,0),


when involved with the 3d and about half the time in the same environments with

~-~B. E. Warren, J. Appl. Phys.. 8. 645 (1937).


~~R H. Gaskell. J. Non-Crvst. Solids. 32. 207 (1979).
~'~R. E. Watson and L. H. Bennett. Phys. Rec. B 43. 11642 (1991).
9sC. B. Shoemaker. D. P. Shoemaker. and R. Fruchart. Acta Costallo~. Sec. C. 40, 1665 (1984).
,~6j. F. Herbst, J. J. Croat, and W. B. Yelon. J. Appl. Phys.. 57. 4086 (1985).
TRANSITION-METALS AND THEIR ALLOYS 47

the 4d and 5d elements. In the remainder of cases, both triangulated and non-
triangulated environments are found, usually low coordination 6-, 7-, and 8-fold
environments.

Si: This is the largest of the atoms being considered here. It is most often to be
found in the Bernal, plus the (0,2,8,1) and (0.0,12,0) environments. There are
also a variety of other triangulated and non-triangulated Si environments which
are, most often, 10-fold in coordination.

On going to the 50/50 compounds, metalloids having greater near-neighbor


coordination than 6 are likely to find other metalloids among these nearest neigh-
bors. Nevertheless, the 50/50 3d borides keep ~4 their (0,3,6.0) environments, except
for TiB, which is in the NaC1 structure and has the (0,6,0,0) environment. Si and
R on the other hand, are found in non-triangulated environments which are mostly
10-fold--(0,6,4,0)---for the phosphides and 13-fold---(0,6,6,1 )--among the silicides.
As expected, the size of a metalloid appears to affect the coordination num-
ber(s) of its environment(s). When the coordination number is right, a Bemal
environment tends to prevail. Further, a good ru le')7.~s is that the densities of an
ordered compound and a glass, at the same composition, are essentially the same.
There have been a number of other considerations of the role ~:~s~') ~J ~~ of atomic
size in glass formation. Central to this is the choice of reference atomic volumes
for the elemental constituents which are appropriate to metallic bonding. Teatum
and coworkers estimated such 12-fold coordinated sizes ~~ for metalloids includ-
ing those above. Turnbull 9z~)s observed that for metalloids, such as P, Si, and Ge,
alloyed with late transition-metals from the Fe, Co, Ni, and Cu columns, the
average atomic volume in the glass is that of the transition-metal. This conclusion,
in large part, may be due to these metalloids having metallic volumes (based on
estimates by Teatum et al.) that are essentially the same as those of the 3d elements
Fe through Cu. ~~ Whatever the detailed choice of reference volumes, there
101
appear to be observable deviations from Vegard's law when the metalloids are
involved in compound formation with elements at the beginning of the transition-
metal rows.
Johnson and Williams considered ~"~the variation in the atomic volume of a glass
as a function of metalloid concentration for several transition-metal-metalloid systems.

~7 D. Turnbull, Scripta Metall., 11, 1131 I 1977).


9~ D. Tumbull, Scripta Metall., 15, 1039/198] ).
99 R H. Gaskell, Acta Metall., 29, 1203 {1981).
~~176 L. Johnson and A. R. Williams, Phys. Re~: B. 20, 1640 i1979).
~0~ L. H. Bennett and R. E. Watson, Scripta Metall., 16, 1379 (1982).
10e E. Teatum, K, Gschneidner, and J. T. Waber, Los Alamos Report LA-2345 /unpublished. 1960),
as reported in W. B. Pearson, The Crystal Chemistry and Phv.~i~'s ~[Metals (J. Wiley and Sons. New
York, 1972).
48 R.E. WATSON AND M. WEINERT

They observed the volumes to vary linearly with concentration. The glass volumes
are thus describable in terms of a pair of fixed elemental volumes; however, these
transition-metals were some 5% larger than the elemental volumes. This change
in volume is small compared with the 40-50% change associated with taking a
system from 12-fold metallic packing to the covalent structures. If one insists
that the transition-metals must be assigned their elemental volumes, then the
metalloid volumes must vary with concentration. However, the metalloid con-
centration ranges of concern were up to 30%. and over this range experience with
the crystalline compounds suggests that the metalloids have nearest neighbors
consisting only of transition-metals. Meanwhile the transition-metals sustain
increases in metalloid neighbor counts of up to something around 3 per metal
atom. This observation suggests that if any site should see a change in volume,
it should be the transition-metal site. Johnson and Williams also monitored the
nearest-neighbor distance as obtained from x-ray radial distribution functions.
The distances showed a maxima at concentrations (-20-30%) at which the
electrical resistivities showed breaks in slope. They deemed these concentrations
to correspond to an "ideal amorphous structure," with the glasses to either side
involving defects in that structure. Whatever the nature of this packing, it involves
metalloids, and of necessity is different from the ideal topologically close-packing
9 72 73
envisaged ' by Nelson and Spaepen. The volumes and bonding of the constit-
uents thus affect the packing which may occur.

12. ATOMIC ENVIRONMENTS AND POLYMORPHISM IN


TRANSITION-METAL-RICH--METALLOID GLASSES

Polymorphism has been known for some time in at least a few transition-metal-
metalloid systems such as Fe-B ~~ and Ni-P. ~~176 In the latter case, different
methods of fabrication yielded two different glassy phases over a 14 to 25%
phosphorus composition range, as manifested by different P Knight shifts and
different densities. The Knight shifts were distinct from one another, with only
one varying with metalloid composition9 One phase proved to be much easier to
fabricate. Annealing it led to an irreversible change of structure in the bulk to the
other phase. This transition was accompanied by a discontinuous change in the
Knight shift. The observed densities and the P Knight shifts of the glasses and of
the compound Ni3P indicated a similarity between the compound and the stabler,
94
less easy to fabricate, glass. Subsequently, Wigner-Seitz cells were constructed ~
for Ni2E Ni3P, and Ni~2P5 that showed the P in the first two compounds to be in

103j. L. Walter, S. E Bartram, and I. Mella, Matet: Sci. Eng.. 36. 193 (1978).
~~ H. Bennett, H. E. Schone, and R Gustafson, Phys. Rev., B 18, 2027 (1978).
~05 D. S. Lashmore, L. H. Bennett, H. E. Schone, R Gustafson. and R. E. Watson, Phys. Rev. Lett.,
48, 1760 (1982).
TRANSITION-METALSAND THEIR ALLOYS 49

(0,3,6,0) Bernal sites, while in the latter P is in the non-triangulated (8,0,0,0) and
(0,6,4,0) environments. The question naturally arises whether the propensity for
a system to form compounds with metalloids in distinctly different classes of
environments is indicative of a propensity for polymorphism in the glasses. Fe-B,
which is also reported ~~ to be polymorphic, has Fe-rich compounds with boron
in Bernal environments, except for perhaps metastable Fe~3B6, which has B in a
non-triangulated (0,8,0,0) environment. Two other systems, Cr-C and Mn-C, form
in the same structure, but otherwise the carbons are in Bernal environments in
other phases, again perhaps indicating a propensity for these to form polymorphic
glasses. These are the only two other systems among the carbides or borides that
appear to be likely candidates for polymorphism. Among the transition-metal-
rich phosphides and silicides, a number of possibilities exist. Here, however,
matters are complicated by the interplay of size and coordination number in the
environments of these metalloids.

V. Charge "Transfer" and Alloying

In Section III, the relationship of electron count and electronegativity scales to


bonding trends were discussed. Underlying both of these parameters--even in the
choice of language--is the concept of charge transfer. In ionic systems, the
connection between charge transfer and bonding is rather direct and well-known.
For metallically bonded alloys, the connection is less obvious. As discussed earlier,
electronegativity differences traditionally are viewed as measures of both charge
transfer and the strength of bonding in compounds. Thus, the heats of formation
of Fig. 5 mirror the electronegativities of Fig. 15.
Charge transfer plays an important role not just in models of bonding trends, but
is the starting point for many other more complicated models of the electronic
properties of materials. For example, the formal valence of atoms, such as 02- or Cu 2+,
are often taken as an accurate discription of the electronic structure of an atom in a
solid. Understanding when and where such simplifications are acceptable is essential.
Although charge transfer is an intuitive concept, quantifying it is difficult, and different
ways of looking at the same data can lead to vastly different conclusions.
The goal of this section is not to relate charge transfer to the energetics of alloy
bonding, since this was done indirectly already in the discussions of electron count
and electronegativity. Rather, the goal is to investigate the concept of charge transfer
itself in alloys, and to consider the questions:

9 How to define charge transfer?

9 Where is the charge transferred to or from?

9 How is charge transfer related to physical measurements?


50 R.E. WATSON AND M. WEINERT

9 What are the physical consequences of charge transfer?

9 Is charge transfer a useful concept?

Although we cannot fully answer these questions, we will show that some of the
issues are quite general and that they need to be kept in mind whenever "charge
transfer" is invoked.

13. CHARGE TRANSFER AS SAMPLED BY OPTICAL


AND PHOTOELECTRON EXPERIMENTS

For several thousand years goldsmiths have alloyed small amounts of silver into
gold to produce white gold. The Au 5d bands (and core levels) lie lower relative
to the Fermi level in the alloy than they do in pure Au. The lowering of the upper
edge of the 5d band increases the energy of the absorption cutoff, and hence is
responsible for the whitening. In the simplest view, an observed increase
(decrease) in the binding of a level indicates an increasingly attractive (repulsive)
potential associated with charge transfer off (onto) the sampling site. This would
imply that Au has lost charge to Ag, contrary to the notion that Au is the more
electronegative of the two elements. Observed Au core level shifts ~~ ~0v.10s when
it is alloyed with other metals almost invariably also show an increase in binding.
(For a discussion of the interpretation of core level shifts, see e.g., Spanjaard
et a/. 1~ o r Egelhoff. ~~
These complications are not limited to gold alloys. In a classic paper, ~j~ Steiner
and Htifner obtained the core level shifts of both sites as a function of concen-
tration for a number of binary alloys grown by coevaporation. The alloys involved
constituents from the Ti, Ni, and Cu columns. Their results for the Pd-Zr system
(already encountered in Figs. 21 and 22) appear in Fig. 29. The shifts at both
sites are substantial for a metallic system, are indicative of increased binding,
and are of common sign. Half of the systems considered by Steiner and Htifner
had shifts of common sign. Since charge cannot very well flow on or off both
constituents simultaneously, the simplest version of level shifts does not hold for
such transition-metal alloying.
As has been already discussed in this article, bonding in such alloys is rather
more complex. The sign of a core level shift need not be indicative of the direction

1r R. E. Watson, J. Hudis, and M. L. Perlman, Phys. Re~: B, 4. 4139 (1971).


~o7T. S. Chou, M. L. Perlman, and R. E. Watson. Phys. Ret: B. 14. 3248 (1976).
~os T. K. Sham, M. L. Perlman, and R. E. Watson. Phys. Re~: B. 19, 539 { 1979).
10~D. Spanjaard, C. Guillot, M. Desjonqueres. G. Treglia, and J. Lecante, Surf Sci. Rep., 5, 1 (1985).
110W, F. Egelhoff, Sulf Sci. Rep., 6, 253 (1987). (1978).
l l~ E Steiner and S. Htifner, Acta Metall., 29, 1885 (1981).
TRANSITION-METALS AND THEIR ALLOYS 51

_
t I 9 . . . .

>.
r

,,,,_,

o
r

0.0 0.5 1.0


Pd concentration

FIG. 29. Steiner and H~ifner's curves summarizing their observed 3d51: core level shifts for the Pd-Zr
system ~ as a function of alloy composition. A positive sign for the shifts indicates increased binding
of the core levels.

of any net charge transfer. Other factors besides interatomic charge transfer are
obviously important. A short, nonexhaustive, list follows. ~-~

1. Changes in the screening of the final state hole: The photoemission process
leaves the system in an excited state. The energy associated with the
relaxation of the electrons around the core hole--the difference between
the unrelaxed Koopman energy and the measured binding energy--can be
quite large in magnitude. ~ ~ 4 ~ These final state relaxations will contrib-
ute to the core-level shifts only through the difference in relaxation between
different environments. These differences are reasonably small for metallic
systems, although not always negligible. ~ ~ 7

~2 M. Weinert and R. E. Watson, Phys. Re~: B, 51. 17168 11995).


I~ A. R. Williams and N. D. Lang, Phys. Re~: Lett.. 40, 954 (1978).
ll4A. R. Williams and N. D. Lang, Phys. Re~: B. 16, 2408 (1977).
~ M. Weinert, J. W. Davenport, and R. E. Watson. Phys. Rev. B, 34. 2971 (1986).
n~6j. R. Smith, E J. Arlinghaus, and J. G. Gay. Phys. Re~: B, 26, 1071 (1982).
~7 M. Alden, H. L. Shriver, and B. Johansson, Phys. Re~: Lett.. 71, 2449 (1993).
52 R.E. WATSONAND M. WEINERT

2. Changes in reference level, i.e., the Fermi energy: While changes in the
reference level may seem odd at first, modification of bandwidths and band
shapes due to changes in volumes and/or chemical bonding may change
the position of the Fermi level relative to the center of gravity of the bands,
etc. Such changes in reference levels will contribute to measured shifts.

3. Intra-atomic, e.g., sp ~ d, charge interchange: Because of the different


spatial extent of the two groups of valence electrons, they make different
contributions to the Coulomb potential.

4. Redistribution of charge due to bonding and hybridization: Significant


changes in the spatial distribution of charge, and hence the potential it
produces, can occur even without inter- or intra-atomic charge transfer. This
effect seems underappreciated in the literature. ~s For example, consider a
monatomic solid. The bonding orbitals have increased density in the inter-
stitial bond region between the atoms which, from charge neutrality consid-
erations, implies less charge close to the nucleus. Antibonding states, on the
other hand, have increased density closer to the nucleus and less in the
interstitial region. These charge rearrangements give rise to increased
(decreased) core-level binding energies for the bonding (antibonding) case.
Therefore, relative changes in bonding between systems can cause core-level
shifts in a completely covalent, i.e., no charge transfer or interchange, regime.

While separating out these various (initial state) contributions unambiguously


is impossible since they are usually interrelated, it is still worthwhile to attempt
such divisions in order to have a picture of the bonding and of the associated
core-level shifts themselves. Without such an understanding, a meaningful inter-
pretation of experimental level shifts borders on the impossible.
For the gold alloys M6ssbauer isomer shift data ~~ exists. The isomer shift
measures the change in s (and relativistic P~/2) charge density character at the
nucleus that arises from changes in valence electron count and any adjustments
of the core electrons to the changes in bonding. The data indicated that the alloys
had increased Au s valence electron counts. Granted that the 5d electrons are more
compact and have substantially larger Coulomb interactions with the core than
do the valence s and p electrons, a simple d-non-d charge compensation picture
emerges. The Au 5d band, though filled, is chemically active. Hybridization of
wave function character of the other constituent into the Au d-band lowers the

~8 M. Strongin, M. W. Ruckman, M. Weinert, R. E. Watson, and J. W. Davenport. Metallic Alloys:


Experimental and Theoretical Perspectives, J. S. Faulkner and R. G. Jordan. eds. (KluwerAcademic,
Dordrecht, 1994), p. 37.
~IgE H. Barrett, R. W. Grant. M. Kaplan. D. A. Keller. and D. A. Shirley. J. Chem. Phys.. 35. 1035
(1963): L. D. Roberts, R. L. Becker, F. E. Obenshain. andJ. O. Thomson, Phys. Rex:. 137.A895 (1965).
TRANSITION-METALS AND THEIR ALLOYS 53

1 1 1 I I 1 1 I 1
0.2 l~--m _

Other site ,, ,, "


. . . . . . ,,~ , ~ ~
0.0
"<1

-0.2
I , , ,

1 I I I | 1 1 I 1 ._]
0.2
Au site []

0.0

/ /

-0.2
1 1 I 1 1 1 I 1 1

Lu Hf Ta W Re Os Ir Pt bcc Au

FIG. 30. Individual d charge transfer components of Au - 5d transition-metal alloys in the CsCl
structure. Mulliken populations (dashed lines) and WS sphere integrations (solid lines) are shown.
The d charge count changes associated with taking Au from the fcc to the bcc structure are shov~n
to the right.

net d count. Of course, there also may be intra-atomic sp ~ d interchange accom-


panying the bonding. The isomer shift, then, indicates a compensating increase in
non-d count. Renormalized atom estimates ~(~ ~07.~0s of the Coulomb interactions of
the d and non-d electrons with the core (including relaxation effects associated with
the changing counts) and of the valence s contact density, combined with the
experimental data, yielded small net charge transfers onto the Au site. In most cases
these were inconsequential in magnitude, but did indicate Au to be the more
electronegative element, although barely. The full tale of the bonding in these
systems, however, is more complicated.
Calculated values 12~ for the d-count change, Aria, for Au alloyed with the 5d
elements in the CsC1 structure appear in Fig. 30, One advantage of resorting to
calculation is that one can look at trends and scan a sequence of alloys even if
some do not exist: W, Re, Os, and Ir do not form alloys with Au and the remaining
four form in quite different phases. Two separate estimates of &zj are plotted. One
arises from a simple integration of the d density over a Wigner-Seitz sphere while
the other is the result of a Mulliken population analysis ~e~ (see the Appendix).
Both samplings indicate that both alloy constituents lose d count across most of

12~ E. Watson, J. W. Davenport, and M. Weinert. Phys. Re~: B, 35, 508 (1987).
121R. S. Mulliken, J. Chem. Phys., 23, 1833 (1955).
54 R.E. WATSON AND M. WEINERT

1 I I i i 1 I

0
O -

0
I

O 0 9
0 9

-1
O
I 1 I l I 1 1 I

Lu Hf Ta W Re Os Ir Pt

FIG. 31. The calculated initial-state chemical shifts for the 4f,, core levels associated with going
from the elemental metals to the compounds in the CsCI structure. Filled circles denote the Au
shifts and the open indicate the shifts for the other alloy component. Increased core level binding
is associated with negative shifts.

the row. The calculated initial state core level shifts (i.e., final state screening is
neglected) shown in Fig. 31 display an increase in binding except for those few
cases (Ir and Pt) where Anj is already positive. There is some scatter in the shifts,
and there is the question of which set of ~zj to employ, but dividing the calculated
shift by zSa~dyields an effective Coulomb d term of the order of 5 eV. This is a
reasonable number for the interaction of the valence d with the core, but does
not imply that other contributions to the core level shifts are inconsequential.
Calculations 13 for Pt alloyed with its 5d neighbors show similar results for the
other transition-metal site, though with a somewhat smaller effective Coulomb
interaction.
At the Pt sites, however, the situation is quite different: The core level binding
increases on the order of 1/2 eV for PtHf through PtRe, but And is positive and
essentially zero-valued. These ~z,~ would be expected to contribute small decreases
in binding. Changes in s and p counts also anticorrelate with the core level shifts.
The explanation 13 is that a situation similar to that encountered in Sect. 8 for Pd3Zr
arises: For alloys across much of the 5d row, the low-lying Pt-like 5d bands are
filled; the combination of this filling with the hybridization of the other alloy
constituent into these bands leads to very small changes in the Pt d counts. These
bands are then no longer pinned to the Fermi level. As the relatively low density
of states bands immediately above are filled, the Pt 5d center of gravity shifts
down relative to the Fermi level to accommodate the remaining valence charge.
The core levels will tend to follow, though not precisely. The result is an increase
in the distance between the Pt core levels and the Fermi energy. A measurable part
of the Pt level shifts appears to arise from this shift in the reference Fermi level.
TRANSITION-METALS AND THEIR ALLOYS 55

The Pt-5d systems were also investigated in the CuAu-I structure. Here the An~
remain small, but reversed in sign, consistent with the calculated level shifts.
However, the effective Coulomb interactions, again obtained by dividing the shifts
by the Anj, are -20-25eV, which is of the order of an unscreened Coulomb
interaction between an interior core level and a d electron. This value is unreason-
ably large when screening of the d transfer is taken into account. Thus, while now
of the fight sign, the Pt 5d count changes again do not appear to play a primary
role in the core level shifts.
Inspection of the bonding of adlayers on surfaces, and the associated core
level shifts, offers an interesting juxtaposition with the behavior in the bulk and,
more importantly, allows consideration of the bonding appropriate to compounds
which are not stable in the bulk (see Fig. 5). Metastable alloys of such unstable
systems are being fabricated and are finding increasing utilization in industry.
Rodriguez and Goodman studied ~22123 the core-level shifts of Cu and Pd adlayers
on a number of 4d and 5d substrates. With the exception of Pd-Ta, the adlayer
and substrate metals do not form bulk compounds. However, because of their
smaller surface energies, Cu and Pd do wet the substrates, and form stable
adlayers. Only the core-level shifts of the adlayers were obtained, though the
shift of the Ta substrate under the Pd adlayer had been reported elsewhere for
P d - T a . 124"125 If the core-level shifts of the adlayers are interpreted ~22123 in terms
of the standard charge transfer picture, then one would require a new electrone-
gativity scale for the adlayers opposite to that found for bulk systems. Some of
these results appear in Fig. 32, along with the experimental value ~2~ for Pd in
crystalline PdTa. Also shown are calculated level shifts ~2 for the 50/50 com-
pounds in the CuAu-I structure ~:7 and for Cu and Pd as adlayers on Ta(1 10).
The shifts for the adlayer systems agree in sign, although smaller in magnitude,
than those of the compounds, as would be expected from the fact the adlayers
have fewer unlike neighbors than the bulk compounds. Theory and experiment
are in accord for those cases where both are available. The shifts for Cu and Pd,

122 j. A. Rodriguez and D. W. J. Goodman, J. Phys. Chem., 95, 4196 (1991).


123 j. A. Rodriguez and D. W. J. Goodman, Science, 257, 897 (1992).
~24M. W. Ruckman, V. Margai, and M. Strongin, Phys. Reu B, 34, 6759 (1986).
~25Also see Fig. 4 in M. W. Ruckman and M. Strongin, Acc. Chem. Res., 27, 250 (1994).
126F. V. Hillebrecht, J. C. Fuggle, P. A. Bennett, Z. Zolnierek. and C. Freiburg, Phys. Reu, B 27, 2179
(1983).
J27The calculations for the ordered alloys were done in the CuAu-I structure (undistorted fcc lattice).
This simple structure appeared to be the most reasonable choice for the systems of concern for several
reasons: i) The calculated heats for the CsCI structures were less bound, ii) PdTa is reported as being
in the CuTi structure and, at high temperatures, in a disordered fcc phase, iii) The calculated heat of
formation of PdTa in the CuTi structure is only 0.01eV/atom more bound than that for CuAu-I. it')
PdTa is the only transition-metal-transition-metal system reported to occur (along with the noble
metals alloyed with Ti, Zr, and Hf) in the CuTi structure. For these reasons, the CuAu-I structure
appeared to be the better representative choice.
56 R.E. WATSON AND M. WEINERT

0.5

0.0 w

~" -O.5
O

1.o Pd

0.5
\ 9
0.0
-0.5
Ta W Re Ru I'r
Pd or Cu site, bulk
S - - C Other site, bulk

FIG. 32. Calculated initial state shifts for both sites for compounds in the CuAu-I structure, and
for Pd/Ta(110) and Cu/Ta(110) adlayer systems. Experimental core-level shifts for Cu and Pd adlayers
on various substrates (filled diamonds), for the Ta substrate under Pd iopen diamond), and tbr Pd in
bulk PdTa (+) are also indicated. Positive values indicate increased binding energy. The open squares
indicate the calculated W shifts in the bulk alloys when fcc W, rather than bcc W, is used as a
reference. (After Ref. [112].)

as by now might be expected, are to increased binding. The dips seen for the
W shifts are associated with W being bcc rather than close-packed: As shown
in the figure, W falls into line with its close-packed neighbors to the right if
fcc W, rather than bcc W, is employed as the reference metal.
Two samplings of the calculated changes in d-electron counts for the compounds
are shown in Fig. 33. One sampling involves the change in the count within the
atomic spheres, while the other involves the "locally-based" charge, as will be
discussed in the next section, inside the Wigner-Seitz spheres. The differences
between the two choices have little effect on the conclusions here. Except for the
failure of the Cu and Pd &~,~values to bend down at the lefthand sides of the plots,
the d counts and the level shifts display quite similar trends. If one takes the ratio
of the shifts to the z~nd, one again gets values of 20-25 eV for Pd, 10-12 eV for
Cu, and--7 eV for the others. There is some suggestion that the variation in the
level shifts is being driven by the variation in &7~ for the others but that other
factors play a substantial role for Pd and Cu. Hence, the core-level shifts often
indicate the sign, but not necessarily the magnitude, of An~.
In viewing the shifts, it is desirable ~~2to separate the contributions to the change
in the potential into intra- and extra-atomic terms. The "standard" models of core-
level shifts consider basically the intra-atomic contributions alone, i.e.. effect of
charge transfer on the on-site binding. The extra-atomic terms can be thought of
TRANSITION-METALS AND THEIR ALLOYS 57

-0.1o - _a_

- _ Cu
-0.05 f
o.oo

t- 0.05 t
Ta W Re (SS Ir

-0.10 9 Pd
-0.05

0.00
r~ W Re Ru Ir

FIG. 33. Calculated changes in the on-site d-electron count ~ nj for the bulk alloys represented in
Fig. 32. Positive values indicate increased d count at a site. Note the inverted scale to allow direct
comparison with the previous figure. The solid lines indicate the results for Pd and Cu sites, and the
dashed, the other sites in the compounds. The 9 and + correspond to the direct integration of the d
density over the atomic sphere: gq and x correspond to the on-site-only sampling.

as Madelung-type (long-ranged Coulomb) contributions arising from changes in


the charge distribution throughout the crystal.
The intra-atomic terms, as sampled by an inner core electron, may be written:

(13)

or its corresponding ~ decomposition. The n and (1/r) involve integrations out


to some sphere radius, such as the atomic or the Wigner-Seitz. It is changes in
such local samplings of the amount (n) and radial shape ( ( l / r ) ) of the valence
charge which are traditionally thought to be responsible for the level shifts. The
situation is more complex when extra-atomic contributions are considered.
For adlayers, ~2 the extra-atomic contributions to the potential are seen to be
larger in magnitude than the intra-atomic for any reasonable choice sphere radii.
In the case of Cu on Ta(110), the two terms are of the same sign, while for
Cu/Rh(001) and Pd/Ta(110), they are found to be of opposite signs. In marked
contrast with bulk alloys, the largest contributions to the intra-atomic terms come
from the s electron charge. As for the substrate atoms at the interface, the intra-
atomic terms are relatively more important and, in fact, are found to dominate in
Pd/Ta and Cu/Rh. These results suggest that substrate core-level shifts provide
a better measure of the local chemical effects than do the adlayer shifts. This
conclusion is not altogether surprising since the adlayer is more heavily affected
by the presence of the vacuum and by interference between these effects and the
interaction with the substrate.
58 R.E. WATSON AND M. WEINERT

The breakup into intra- and extra-atomic contributions has relevance to bulk
compounds as well. Consider Pd3Zr, which was encountered before. Using the
atomic radius as the cut-off to separate intra- and extra-atomic contributions, extra-
atomic terms are found to dominate at the Pd sites, and extra- and intra-atomic
terms are roughly equal and of the same sign for the Zr sites. The Pd intra-atomic
terms are small because An is small and the d contributions to the first two terms
of Eq. (13) largely cancel each other. Among the intra-atomic terms at the Zr site,
the And(1/r)d is substantial and is of the same sign and magnitude as the total
shift and the extra-atomic contribution, but is largely canceled by the other (
contributions.
For PdZr in the CsC1 and CuAu-I structures, the contributions to the core-level
shifts for the two structures differ in some significant details: The extra-atomic
(intra-atomic) contribution dominates at the Pd (Zr) site in the CsC1 structure, while
extra- and intra-atomic terms are about equal in the CuAu-I structure. Again,
And and its contribution to the shifts are small at the Pd sites, while the njA ( 1/r) d
contributions are significant and smaller than--but of the same sign as--the total
shifts. Although differing by a factor of four in magnitude, the ~zd terms are
again important at the Zr sites in the two structures with signs consistent with the
level shifts.
When Pd is replaced by Rh in these systems, the most striking feature is that the
Rh zSa~dterm is significant and opposite in sign to that on the Zr site. Except for Zr
in Rh3Zr, all And terms have signs in accord with the core-level shifts, and the extra-
atomic contributions dominate. The Rh sites share the distinction of having small
ndA(1/r)d terms: for the Pd and Zr sites these terms are typically 1/2 to 1 eV, so,
while not dominant, they are of experimental significance.
The above inspection of only six compounds suggests the complexity of the
bonding effects and their effect on the potential which contributes to a core-level
shift. First, there is the competition between intra- and extra-atomic contributions,
with the latter often prevailing. Second, there is the competition between the first
and second terms of Eq. (13), i.e., changes in occupations vs. changes in shape.
When it comes to the d electrons, the second term is of experimental significance,
though usually it will only dominate when the first term is small. Third, there is
the competition between charge effects arising from charge of differing ~. In three-
quarters of the sites encountered in the six compounds, the zSaTjterm is of a sign
consistent with the level shift. It is thus tempting to attribute the sign, if not the
magnitude, of a shift to changes in d count. However, in eight of these nine cases,
there are non-d intra-atomic contributions of opposite sign, which when summed,
are larger than the d term taken alone. This puts the attribution of core-level shifts
to changes in charge transfer on a shaky physical basis, even if the correlation
between the sign of And and the shifts holds much of the time.
Although core-level shifts as a measure of charge transfer are rather imper-
fect, the shifts can provide useful information about bonding and energetics
TRANSITION-METALS AND THEIR ALLOYS 59

in metals. ~28'~29Core-level binding energies are the difference between the ground
state and a state with a core hole in it. For metals, it is expected that the screening
of the core hole is quite efficient, leading to the Z+I model, 13~ which assumes
that the final state of a core hole at an atom of charge Z is equivalent to an impurity
atom of charge Z+I with an extra occupied valence electron in its vicinity. Then,
under these assumptions, the measured core-level shifts are related to differences
in heats of formation. This model appears to work well, even if the core electron is
not from deep in the core. It has been applied successfully to estimating jjLj3~ the
heats of formation of alloys where the Z of the constituents differs by one; it appears
to be particularly useful when considering level shifts between atoms at surfaces and
in the b u l k 13'~'135"136"~37and relating the observations to trends, for example, in surface
segregation.

14. CHARGE TAILING AND THE LOCAL SAMPLING OF ORBITAL POPULATIONS

Modern electronic structure calculations can give rather accurate wave function
information. Given the wave functions, one can expand them in terms of different
angular momentum components about a site. If the atoms are well separated, this
decomposition yields the standard s, p, d... atomic states. For a solid, however,
one finds that such an expansion typically requires ~ > 8 for reasonable conver-
gence. The high-f components are associated with the tailing of orbitals associ-
ated with neighboring atomic sites. They are not coming from bands of intrinsic
- 4-, 5-, or even 8-1ike character. The sums of the charge associated with f > 3
13.120.138
components for the 5d elemental metals are shown in Fig. 34. It is seen
that there is substantial charge in these sums and it varies significantly across the
5d row.

128C. E Flynn, J. Phys. F, 10, 2315 (1980).


129C. E Flynn, Phys. Rev. B, 14, 5294 (1976).
~30N. M~u'tensson and B. Johansson, Solid State Commun., 32, 791 (1979).
~3~B. Johansson and N. M~,rtensson, Phys. Rev. B, 21, 4427 (1980).
L~zE Steiner, S. Htifner, N. M~.rtensson, and B. Johansson, Solid State Commun., 37, 73 (1981).
~33E Steiner and S. H~ifner, Solid State Commun., 37, 79 ( 1981 ). This paper provides a lucid example
(for the Pd-Ag system) of the process of extracting an alloy heat of formation as a function of
concentration, from core-level shifts.
134V. Kumar, D. Tom~inek, and K. H. Bennemann, Solid State Commun., 39, 987 (1981).
135W. E Egelhoff, Jr., Phys. Rev. Len., 50, 587 (1983).
136N. M~rtensson, A. Stenborg, O. Bjtirneholm, A. Nilsson, and J. N. Andersen, Phys. Rev. Lett.. 60,
1731 (1988).
~37 M. Weinert, R. E. Watson, J. W. Davenport, and G. W. Fernando, Ordering at Surfaces and
Interfaces, A. Yoshimori, T. Shinto, and H. Watanabe, eds. (Springer-Verlag, Berlin, 1992), p. 321.
~38R. E. Watson, M. Weinert, and G. W. Fernando, Phys. Re~; B, 43, 1446 (1991 ).
60 R . E . W A T S O N A N D M. W E I N E R T

0.3 l l l , i i , , ,

9= 0.2
-<
9

~D
m0.1
AI

0.0 L L 1 I I I I I I
Lu Hf Ta W Re Os Ir Pt Au

FIG. 34. The calculated sums of ,t > 3 contributions to the electron counts residing in the Wigner-
Seitz spheres of the elemental 5d metals.

What happens in a compound when atom A has atoms of type B as nearest


neighbors? One expects that the "tailing" charge density sampled at A will, in
first approximation, be characteristic of B, and that the amount of charge will
depend on the volume over which A samples the tail of B atom density, i.e.,

tail VA tail
nAi,,AB = 77-,n B (1> 3), (14)
vB

where V8 is the elemental B metal volume and VA is the volume characteristic of A


in the compound. As shown in Fig. 35, this accurately reproduces the tailing
calculated for the Au compounds in the CsC1 structure. Granted the assumptions,
the agreement is remarkable. More importantly, the changes in tailing charge counts
on going from the elemental metals in Fig. 34 to the compounds ("Other site"
curve) in Fig. 35 are as much as a quarter an electron, which is quite substantial.
The question that naturally arises is What part of the charge transfer and of the
redistribution of charge, of all ~, is intrinsic to the site at which it resides ? It is
the intrinsic charge transfer which is being implicitly considered in simple models
of bonding: Electronegativity scales and models of d-band bonding are related to
the intrinsic transfer. In contrast, experiments and calculations normally give infor-
mation about the actual (spatial) distribution of charge in a solid and what might
be called "apparent" charge transfer. Such apparent charge transfers are often used
in simple models that assume "intrinsic" charge transfer effects: the results are
often contrary to intuition. A simple example, already encountered, is the case
where core-level shifts are of the same sign at both sites in a binary alloy, implying
charge flow onto (or off of) both sites simultaneously. Another example is the use
TRANSITION-METALS AND THEIR ALLOYS 61

I l I I I I I I I

[3
0.3

,ha

O
= 0.2

AI
~O. 1

0.0 l l 1 I 1 1 1 L.. I

Lu Hf Ta W Re Os Ir Pt Au
FIG. 35. The sums of f > 3 contributions to the electron counts residing in the Wigner-Seitz spheres
of Au and the other element, as obtained from LASTO calculations ~:' ~3s for compounds in the CsC1
structure (the solid lines). The dashed curves are estimates employing Eq. (14), using the charges in
Fig. 34 and assuming that V~ has its elemental value.

of calculated atomic populations to describe the bonding in transition-metals. Often


much of the apparent charge transfer in an alloy, i.e., the sum of on-site and tailing
terms, is of p electron character. Taken at face value, this result would imply that
changes in p bonding are at least as important a s - - i f not more important than--
changes in d bonding. However, as will be noted shortly, this p-like charge is
primarily tailing in character.
Decomposing the charge in a solid into atomic-like pieces is difficult and it is
clearly not unique. Since we are attempting to obtain atomic-like charges, it would
seem that the simplest way of proceeding in a calculation is to use a linear
combination of atomic orbitals (LCAO) picture and employ one of the various
population analyses 121139~4~used by quantum chemists. Unfortunately, population
analyses such as those by Mulliken or L6wdin are really measures of apparent
charge transfer: There is no information as to whether the actual charge at a given
site comes from on-site or from the neighbors since the projection is onto a basis
set of Bloch sums of localized orbitals which already contain the overlap of tails
from the neighboring sites. To make this distinction more concrete, consider the
simplest case of a monoatomic system with one orbital per site. It is clear that
any of the standard population analyses will give a weight of one to this orbital;

~39E. W. Stout and E Politzer, Theor. Chim. Acta, 12, 379 I1968).
1~ P. O. L6wdin, J. Chem. Phys., 18, 365 (1950).
62 R.E. WATSON AND M. WEINERT

they (by construction) do not address the issue of separation into on-site and
tailing contributions. While it is true that such an analysis will give a measure of
intrinsic charge transfer (and was used in Fig. 30), the results can be significantly
skewed and provide an incomplete picture of what is going on. A related approach
to discussing charge transfer, commonly used in the various augmented schemes
(e.g., augmented plane wave, Korringa-Kohn-Rostoker, linear muffin-tin orbital,
augmented spherical wave, and linearized augmented Slater-type orbital), is to
use the f-dependent contributions to the charge density inside atomic spheres for
the analysis. While this is a unique and correct decomposition of the f character
of the density within the site, there is no information as to w h e r e the wave function
character, responsible for the charge, has come from, in any simple chemical picture.
While these different approaches are useful as an indication of charge transfer,
they all measure the apparent transfer and are not strictly appropriate for discus-
sions concerning electronegativity scales and the like. We wish to go beyond such
analyses and determine how much of the apparent charge transfer, obtained with
either the LCAO or the atomic sphere population analyses, is intrinsic to the wave
function character located on the site and what is merely due to the overlap of
tails from neighboring sites. We have already seen significant tail charge residing
in the high ~ components of the charge. In free transition-metals atoms, there
are typically 15 1 to 1.5 electrons outside the Wigner-Seitz sphere appropriate to the
elemental solid. In a simple picture, this would overlap onto the other sites in
the solid. This estimate of tailing contributions agrees with modem estimates ~38
of the tailing charge obtained with the scheme described in the Appendix and below.
A traditional quantum chemical population analysis takes a set of basis functions
and estimates their contributions to the charge throughout space. Here the concern
is different; the goal is to analyze the local charge contributions at a single site
and to determine the changes in charge and bonding attributable to electrons
intrinsic both to the site itself and to its neighboring atoms. One approach is to
separate the wave function into on-site and tailing contributions, ~38from which the
various contributions to the charge density are estimated. The separation of terms,
as discussed in the Appendix, involves the use of a localized basis set of atomic-
like functions, either analytical, numerical, or a combination of both. Having
divided the wave function character into on-site and tailing terms, one then has
the problem familiar to quantum chemists, namely that squaring the wave function
to obtain the charge density yields not only on-site and tailing terms, but also a
cross (overlap) term. In general, the charge in the overlap terms is significant. The
question is how to apportion these overlap contributions. A number of choices
have been made for population analyses taken over all space. For example,
Mulliken suggested 12~ apportioning the overlap equally among the other terms,
while in a modified Mulliken scheme ~39it is apportioned according to the relative
weights of the on-site and tailing terms. Similar choices can be made for the
overlap term in this local site sampling. In the case of transition-metal alloys,
TRANSITION-METALS AND THEIR ALLOYS 63

i ~
"L)

Au, I r o - /
, ,\ Other site
o
0.4
i
/
/
/
/

0.2 , Q
/

" O O
,Q
_ __@---o~
A
0.0
lid
0~ t o
"o- - o - o
-0.2

-0.4
O'
/

I I I . I I I 1 j 1 j
-~--? 1 .l

Hf Ta W Re Os Ir Pt Hf Ta W Re Os Ir Pt

FIG. 36. Calculated 13s charge transfer for Au and Ir alloyed with the 5d elements in the CsCI
structure. The open symbols, circles for the Au and diamonds for the lr alloys, are the total changes
in charge within the Wigner-Seitz sphere (positive values indicate increases in charge). The filled symbols
are the intrinsic charge transfer after tailing and the modified Mulliken apportioning of the overlap
charge have been subtracted off within the same spheres.

the on-site d count can be an order of magnitude greater than the d tailing so that
the modified Mulliken apportioning appears to be the more appropriate.
The consequences of the large amount of tailing charge for Au and Ir alloyed
with other 5d elements can be seen in Fig. 36. The charge transfer ~~8 is " accounted
for in two different ways. The open symbols indicate the total change in charge
within the Wigner-Seitz spheres, with Au always showing a gain in charge. There
is an indication that the rare earths, if they had been included to the left, might
show a net gain from Au. The similar curve for Ir lies parallel to and below the
Au one. The curve for Pt (not shown) lies between the ones of Au and Ir, but
slightly closer to Au.
Inspected casually, the total charge transfer suggests that Au is more electro-
negative that Pt which, in turn, is more electronegative than Ir. However, it also
appears that Ta and Hf and the rare earths to their left are more electronegative
than the metals in the middle of the transition-metal row. A different picture arises
if the change in charge intrinsic to the site is estimated by subtracting off the tailing
and overlap contributions (estimated in the modified Mulliken approximation).
The results are the filled symbols which are smaller in magnitude and, as often as
not, opposite in sign to the total transfer. Now Au loses charge to Pt, It, Os, and
perhaps Re. The Miedema and, to a lesser extent, the d-band electronegativities
of Fig. 15 are consistent with this trend. Also, Hf and Ta now show the greatest
propensity to lose valence charge. Changes in p-like charge counts are important
64 R.E. WATSON AND M. WEINERT

to the accounting: Typically the elemental metals have -0.8 electrons of p-like
charge, more than three-quarters of which comes from tailing and overlap. The
changes in the p-like tailing and overlap with alloying are significant on the scale
of the changes in intrinsic charge count estimated in Fig. 36.
One of the extreme cases of transition-metal alloying is that involving the elec-
tropositive alkali metals. Of the noble and transition-metals, only Au forms ordered
compounds with all of the alkalis. TM AuLi, AuRb, and AuCs form in the CsC1
structure. The latter two are semiconductors due to the large separation between Au
sites that causes gaps to form in the intrinsic Au bands. The gaps are reduced, but
not filled by hybridization with the alkali. The bonding in these systems has been the
object of a number of theoretical investigations. ~z-'~~B'l'H14~'~z6'147~zsOne feature of
the compounds is their extreme volume contractions. Deviations from Vegard's law,
Eq. (10), of 40-50% percent are typical. In the case of the observed Cs, Rb, and
K compounds, as well as the Li- and Na-rich ones, this contraction results in the
volume of the compound being less than that of the alkali metal itself. In esti-
mating charge transfer one must first apportion the volume in the compound. One
choice, biased toward Au, is to assign the elemental Au volume to Au in the
compound. This choice, then, implies substantial contractions of the alkali vol-
umes on going from the elemental metals to the compounds. Despite the bias,
the Au sites in the 50/50 compounds in the CsC1 structure lose 148 between ~and
1
_23electrons in the compound, except for AuLi where the loss is only about ~e.
This observation will not be changed by any reasonable reapportioning of the
crystal volume(s) between sites.
A clue to the apparent charge transfer may be seen in the charge tailing
contributions 148 plotted for AuCs and AuNa in Fig. 37. Apart from details, the
tailing charge at the alkali sites in the compounds resembles that of elemental
Au with a slow decay in the higher f contributions. On the other hand, the rapid
decrease in high ~ contributions at the Au site in the compound resembles those
of the elemental alkali metals. (The disparities involve more pronounced d terms
at the alkali sites in the compounds and more pronounced p in the elemental alkalis).
Calculations for Au in the compounds with the alkalis removed, i.e., for simple
cubic Au at the compound lattice constants, show that the tailing contributions go

14~ In addition to Au, Ag and Pt form ordered compounds with Li and Na: Ir, Pd, and Rh, however,
form compounds with Li alone.
142A. Hasegawa and M. Watabe, J. Phys. F, 7, 75 (19771.
143 H. Overhof, J. Knecht, R. Fisher, and F. Hensel, J. Phys. F, 8, 1607 (1978).
144E Meloni and A. Balderschi, Physics of Narrow Gap Semiconductors. E. Gornik, H. Heinrich,
and L. Palmetshoffer, eds. (Springer, Berlin, 1982).
145 N. E. Christensen and J. Kollar, Solid State Commun., 46, 727 (1983).
1,~6j. Robertson, Phys. Rex'. B, 27, 6322 (1983).
147C. Koenig, N. E. Christensen, and J. Kollar, Phys. Rev. B, 29, 6481 (1984).
~48R. E. Watson and M. Weinert, Phys. Rex: B, 49, 7148 (1994).
TRANSITION-METALS AND THEIR ALLOYS 65

9 iL
0.4 ,' " , AuNa: Au site iI [ ,' 9 ', AuNa: Na site

0.2
o', !i 9 " " i

= 0.0 '
=-

0.4 " ', AuCs: Au site i. AuCs: Cs site i


i
/

0.2 , ', i L ' "~ "t

"'Q.. .~ . . "U_O_O. "


0.0 9 9 --.- "9-o-o i o ,o__o
0 i 2 3 4 .~ 6 7 ~ 0 i ~ 3 -i ~ L 7 s

FIG. 37. Calculated f-decomposed charge tailing contributions to the site charges for AuNa and
AuCs. The open circles ( 9 are the contributions for the elemental metals (at their elemental volumes)
and the filled (O) for the sites in the compounds.

to zero, indicating that the alkalis are responsible for the tailing at the Au sites
in the compounds. Going from alkali to Au tailing contributions at the alkali sites
results in huge increases in tailing charge, despite a halving of the sampling
volumes in the compounds. Reductions of similar magnitude occur for Au. Allo-
cation of some of the overlap terms to the off-site charge would increase the
magnitude of these changes. Even without such additions, subtraction of the
tailing terms from the total site charges leads to a sign reversal, where the alkalis
now have intrinsic charges which are positive and those of Au are negative. Thus,
with this analysis in terms of tailing charge (in contrast to the more standard
population analyses), preconceived notions of charge transfer from the electrop-
ositive alkalis to gold are sustained.
There are hybridization effects which are also of significance to any accounting
of charge in the Au-alkali compounds. In elemental Au, for example, a small
fraction of the p-like charge (-0.15 electron) is neither of tailing nor overlap in
origin, but intrinsic, and is involved in hybridization with the d-bands. When Au
is alloyed, particularly with the heavier alkalis, this p charge is reduced to near
zero. Also, Cs and to a less extent Rb and to a still lesser extent K, display increases
in d count which come from the hybridization of alkali d character, residing several
eV above the Fermi level, into the occupied valence bands. The increase is of
the order of 0.1 of a d electron. This d increase, plus the Au p depletion, has a
measurable effect on the charge balance in the compounds. Such an appearance
of d character has been reported 149 in calculations for Cs adsorbed on W(001),
66 R.E. WATSON AND M. WEINERT

where its presence affects the shape of the charge at the surface and, in turn, the
work function.
There are indications of real changes in the shape of the charge in the Au-alkali
compounds as well. As an example, heats of formation were calculated for AuCs
using both a muffin-tin potential (where the potential is spherical inside the atomic
sphere and constant outside) and a full potential (where aspherical terms in the charge
density and potential are accounted for throughout space). In view of the high
symmetry and good packing of the CsCI structure, rather similar heats were to be
expected. However, full-potential results yield a calculated h e a t - 0 . 5 eV/atom
more bound. This is a large change and indicative of significant modification in
charge shape and character on going from the elements to the compound.
Several lessons can be drawn from these different examples and analyses. In the
literature, there has been a certain tendency to presume that the charge at an atomic
site is associated with orbitals at that site. The discussion above suggests that such
an assumption is not always justified. Hybridization is important to any charge
accounting. For example, in Cu, Au, and Pt (when its d-bands are filled upon
compound formation) hybridization causes the d counts in their filled d-bands to
be measurably less than 10. These changes due to the hybridization are important
to the chemistry of their alloying. It is also clear that any assessment of the charge
behavior is to some extent hostage to the scheme employed, e.g., population anal-
yses require basis sets and assumptions concerning the apportioning of overlap
contributions.

Vl. Electronic Structure and Phase Predictions

Up to now the discussion has been about various parameters such as size and
charge transfer that affect the occurrence of different phases. Already in this article
several examples of binary phase diagrams (Figs. 3 and 21) have been shown.
There are a number of general features: As a function of composition, a number
of ordered compounds exist, often in intrinsically different crystal structures.
Some structures occur over an extended composition range, while others only
over a very narrow range. Some phases exist from low temperatures all the way
up to the melting temperature, while others exist either only at high or only at
low temperatures. Finally, the melting temperature is not uniform across the full
composition range, nor is it a linear interpolation between the elemental melting
temperatures.
Not all phase diagrams show common behavior, but these types of features are
genetic to many phase diagrams of interest. Understanding is an important goal.

149 E. Wimmer, A. J. Freeman, M. Weinert, H. Krakauer, J. R. Hiskes, and A. M. Karo, Phys. Re~:
Lett., 48, 1128 (1982).
TRANSITION-METALS AND THEIR ALLOYS 67

0.0

,,, ~ Stable /
' ~ ' " ~(~,, @Unstable /
9

> -0.5

-1.0
A B
A l-xB,

FIG. 38. Schematic of the heats of formation for various potential phases (a. 13. 7- 5) for a
hypothetical A~_,B, alloy. The tie lines connecting the stable phases are also shown.

An essential aspect of phase stability is the competition among competing phases:


It is not enough to be strongly bound, but for an alloy to be stable, it must be
more stable than other phases in a given concentration range. This last qualifi-
cation is necessary since the values of AH by themselves are not enough infor-
mation to determine stability.
The heats of formation for a hypothetical A-B alloy are shown in Fig. 38. The
various ordered phases, a,/3, 7, 5, all have AH < 0. (The notation is simply as a
label for the discussion.) Naively, this would suggest that these phases are all
stable. However, AH < 0 only means that the compound is stable relative to a
two-phase mixture of A and B. As drawn, there are only two stable phases, and
these two are not necessarily the ones with the largest heats of formation. Just
as AH is a measure of the stability of a two-phase mixture of the pure elements,
each stable phase must lie below the tie line connecting any two other phases.
In the figure, the phase stability is dominated by the heat of the 7 phase. Although
the heat for 5 is the second largest, it is still less than a two-phase mixture of 7
and pure B. On the other hand,/3 is stable relative to a two-phase mixture of 7 and
A, and thus would be stable if only these alloys were considered. However, the
presence of the a phase suppresses the /3 phase even though the heat for 13 is
more binding than for o~. Thus, from this simple set of heats, we would expect
two stable phases (a, 7) and two-phase regions in between. Because of the sig-
nificantly larger binding of 7, we would expect a peak in the melting temperature,
15O
as seen in the Pd-Zr case (Fig. 21).

~s0The heat of formation AH is an enthalpy, not a free energy. Limiting the consideration to All alone
neglects entropic contributions, so this discussion of phase stability is restricted to low temperatures.
68 R.E. WATSON AND M. WEINERT

Although this schematic phase diagram may seem artificial, this type of com-
petition among phases is quite common. Depending on how close a metastable
phase is to the tie line that suppresses it at zero temperature and what the entropy
contributions are, the metastable state can affect the properties of the alloy in,
for example, affecting the propensity toward ordering. The fact that M-/by itself
is not sufficient to determine stability, or even relative stability, means that it may
be necessary to consider the whole concentration range in order to understand
the phase diagram even over a small range.

15. APPROACHES FOR ALLOY CALCULATIONS

Obtaining the thermodynamical data such as A H needed to understand the phase


diagrams is experimentally difficult. In addition, this area of research has been
decreasing in activity. Although theory provides a means for obtaining some of
these results, it is not a complete substitute for experiment. An advantage that theory
does have, however, is that it is not limited to materials that will form; instead it
is possible to consider (metastable) phases that are inaccessible to experiment. The
ability to look at trends (using Wigner and Seitz's "'great computing machine")
across a whole class of related systems hopefully provides more insight into the
underlying physics and chemistry than is possible from looking at a single material.
Over the years many different theoretical approaches for dealing with alloys have
been used, often depending on the system under consideration. One class of systems
studied extensively has the same underlying structure throughout the full compo-
sition range" for example, Au-Pd on a fcc lattice. In these types of systems, ran-
domness and the ordering of the constituents are the important aspects. For such
systems approaches that start from the assumption of a fixed underlying lattice are
reasonable. Such approaches, however, will not be appropriate when there is strong
compound formation that involves different lattice structures.
In the case where the atoms in an A~_,B, alloy are randomly distributed on a
fixed lattice, the conceptually simplest approach is the "'virtual crystal" approxima-
tion (VCA). The physical picture is that looking in any direction, there is a random
distribution of A and B atoms. Hence, on average, the potential that an electron will
see is the concentration average of the potentials for A and B atoms. This observation
leads to the approximation of replacing the lattice containing a random distribution
of potentials by a lattice of averaged potentials on each site, and then solving the
electronic structure problem for this "virtual crystal." One implementation, appro-
priate for use in self-consistent calculations, is to consider each atom on the
lattice to have an effective atomic number of Z - (1 -x)Z~ + xZe. This approx-
imation then treats each site equivalently, thereby ignoring differences in the local
electronic structure and local environment.
TRANSITION-METALS AND THEIR ALLOYS 69

A more sophisticated approach based on the same underlying physics is the


coherent potential approximation (CPA). J~J Here the actual electronic structure of
each site is considered, but the effect of the other atoms is treated in an effective
manner. This means that the distinction between an A and B site is maintained,
and the electronic structure is solved for the A and B atoms embedded in an
averaged host. It can be shown that the CPA is the best single-site representation
for the random alloy ~5~ and that deviations from randomness can be treated
analytically 152 by perturbation theory. The CPA can be extended to include more
of the actual local environment by considering small clusters of A and B atoms.
The CPA uses the random alloy as the basis from which to consider ordering.
Thus, for systems that are dominated by disorder effects and have only a weak
tendency for ordering, the CPA is a reasonable approximation and provides an
excellent way of describing the effects of disorder.
A different, but conceptually straightforward, approach for dealing with dis-
order is to consider a lattice with a large number of sites and explicitly calculate
the properties for different configurations and then average. For a large enough
lattice, this should be a reasonable representation. As a practical matter, however,
treating enough large systems to ensure a reasonable average is computationally
difficult, although not impossible. A practical solution to this problem is to use
the results from sets of smaller calculations to build up the results for the larger
systems. The justification for such a procedure is given by the cluster varia-
tion 153'~54 (CVM) method. In the cluster expansion methods, ~5s~56 a spin-like
occupation number 0", is_.)assigned to each lattice point i in the crystal. The set of
N occupation numbers o" = { o'~ ..... o-x }, where N is the number of sites in the
crystal, fully specify the configuration of the system. In the cluster expansion
method, the energy (or any function of a configuration -> o" ) is written as a weighted
sum of the multisite cluster functions r ~)"

E(3-) - y__, v,,r (3-).


tl

The cluster functions form a complete basis in configurational space and the
expansion coefficients V,, are called effective cluster interactions (ECI). In general,

~s~j. S. Faulkner, Progress in Materials Science, Vol. 27, J. W. Christian. E Haasen, and T. B. Massalski,
eds. (Pergamon Press, New York. 1982). p. 1: and references therein.
152 E Ducastelle, Alloy Phase Stabilitx, Vol. 163, NATO Advanced Study Institute Series E: Applied
Sciences, ed. G. M. Stocks and A. Gonis t Kluwer, Boston. 1989), p. 293.
153 R. Kikuchi, Phys. Rev., 81,988 (1951).
~.s4D. de Fontaine, Solid State Physics, 34. 73 (1979).
155 j. M. Sanchez, E Ducastelle, and D. Gratias, Phvsica, 128A, 334/1984).
156j. M. Sanchez, Phys. Rev. B, 46, 14013 t1993).
70 R.E. WATSONAND M. WEINERT

FIG. 39. Pair, three-body, and four-body clusters of lattice points used in the cluster expansion for
the bcc lattice.

the ECIs are of short range, which allows the expansion to be truncated at a
reasonably small number of clusters. The two-, three-, and four-body clusters for
the bcc lattice are shown in Fig. 39. It may be shown that the ECIs have the
symmetry of the lattice. This requirement allows one to group ~,,(~') in sets of
equivalent clusters. The average over the set of all cluster functions related by
the symmetry of the lattice is called a correlation function, and if multiplied by
the degeneracy of the ECI, can be used for the expansion.
The ECIs can be determined by matrix inversion or by a least-square fit using
the calculated energies of selected (ordered) structures at various volumes, and their
correlation functions. ~-~vSpecial structures ~~s that have correlation functions that
match the ensemble averages of the random alloy are particularly efficient for
determining these ECIs. If total energies are calculated for different volumes and,
for example, for different c/a ratios, 4-~ then these ECIs will have a volume and
c / a ratio dependence also. Given these ECIs, the energy for any ordered or
disordered structure can be calculated. Thus, this approach allows for the efficient
investigation of the 2 ~ possible orderings ~~'~of an A-B alloy on a large lattice of
N points that would be impossible in a completely first-principles approach.

1.~7j. W. D. Connelly and A. R. Williams, Phys. Rev. B. 27, 5169 I1983).


~.~8A. Zunger, S.-H. Wei, L. G. Ferreira, and J. G. Bernard, Phys. Rev. Lett., 65, 353 (1990).
~~9Z. W. Lu, S.-H. Wei, A. Zunger, S. Frota-Pessoa. and L. G. Ferreira. Phys. Rev. B, 44. 512 (1991).
T R A N S I T I O N - M E T A L S AND THEIR ALLOYS 71

An important feature of the cluster methods is that the CVM provides an elegant
and relatively simple way to calculate the configurational entropy contributions to
the free energy, and thus provides a way to include many-body cluster interactions,
temperature, and volume effects into the determination of phase diagrams. The appli-
cation of the CVM-type approaches to alloys systems ~ ~ ~ < has been quite successl-hl.
The approximations involved in the CVM and CPA approaches are different.
These differences are apparent in the treatment of the random alloy, which has lead
to some controversy in the literature. In principle, both approaches treat the random
alloy by averaging over all possible configurations of A ~_,B,. To be random, the site
occupancies should be uncorrelated, but this does not mean that the physical
properties of individual sites are independent of their local environment. ~'~ In the
single-site CPA, this dependence of the properties on the local environment is
neglected, while it is included in the CVM calculations: a cluster CPA, however,
does include them. For systems with strong short-range order and bonding, these
local correlations may be important in determining the properties of the alloy.
As already mentioned above, the CVM and CPA approaches are applicable to
situations where there is an underlying common lattice. For many systems of interest
no such lattice exists. In principle, a CVM-type approach could be undertaken
for each possible crystal structure type, but this procedure will quickly become
infeasible. Thus, for many alloy systems, calculations for sets of plausible crystal
structures is still the most appropriate approach, especially when there is strong
compound formation. The choice of structures considered is often guided by
experiment and the types of patterns discussed elsewhere in this article.

16. DENSITY FUNCTIONAL THEORY

Although this article is not about electronic structure calculations p e r se. the results
of (first-principles) calculations play an essential role in developing an under-
standing of the properties of alloys, and so a brief discussion is warranted. Density
functional theory, ~~5~'6~'7 in" its local spin density approximation ~6s~' (LSDA)

.60 M. Sluiter, D. de Fontaine, X. Q. Guo, R. Podlouckv, and A. J. Freeman, Phys. Rev. B, 42, 10460
(1990).
~6~ j. M. Sanchez, J. P. Stark, and V. L. Moruzzi, Phys. Rev. B, 44, 5411 t1991).
~,2 A. Pasturel, C. Colinet, A. T. Paxton, and M. van Schilfgaarde. J. Phv,s.. Condens. Matter, 4, 945
(1992).
~,3 R. Magri, S.-H. Wei, and A. Zunger, Phys. Rev. B. 42, 11388 / 1990~.
1~4 E Hohenberg and W. Kohn, Phys. Rev., 136, B864 t 19641.
~6s N. D. Mermin, Phys. Re~:, 137, A 1441 I 1965).
~66W. Kohn and L. J. Sham, Phys. Rev.. 140. A 1133 t 1965 i.
le~7A. K. Rajagopal in Advances in Chemic'al Phvsic.s. Vol. 41. I. Prigogine and S. A. Rice. eds. {Wiley,
New York, 1980), p. 59: and reference.,, therein.
~e,s U. yon Barth and L. Hedin, J. Phys. C. 5, 1629/1972 ~.
~"~ O. Gunnarson, B. I. Lundqvist, and S. Lundqvist, Solid State Commun.. 11. 149 (1972).
72 R.E. WATSON AND M. WEINERT

and (more recently) generalized gradient approximation (GGA), 17~has been the
main underlying basis for first-principles electronic structure calculations. The
basis of the theory is the Hohenberg-Kohn ~~65 theorem, and its generalization
to magnetic systems, 167"168"169that the ground state energy of a many-body system
is a unique functional of the charge and spin densities and is a minimum for the
true ground state densities. Kohn and Sham ~66 showed that this functional could
be written as

E[n] - T~[n] + U[n] + E,.c[n, m]. (16)

The classical Hartree (Coulomb) energy, U, for electrons and nuclei of charge Zv
is given by

(17)

External fields other that the nuclear charges that couple to the charge and/or
spin densities can be included by adding in terms of the form ~d]" n(r)Vext(~').
The "kinetic" energy term Ts corresponds to the kinetic energy of a set of non-
interacting particles with the same density n"

, (l)
T , [ n ] - Z ~ d ~ ~,.a(]') -~.V 2 ~,.a(]'), (18)
icr

where the density for spin a is given in terms of the single-particle wave functions
~o by

nG(]- ) = ~ VT,~(~-)V,~(~-). (19)


io"

The spin density m(]') is simply the difference between the spin-up and -down
densities, m (~') - n 1"(~) - n +(~r), and the total density is given by n (~') = n 1"(~') +
n+(~'). The exchange-correlation energy, E,c, is a functional of both the charge
and spin density. In the LSDA, the unknown form is approximated locally by the
exchange-correlation energy appropriate to a homogeneous electron gas of the

170j. R Perdew, K. Burke, and M. Ernzerhof, Phys. Rex: Len.. 77. 3865 (1996): and references therein.
TRANSITION-METALS AND THEIR ALLOYS 73

166,167.168
same density:

ELDA
xc [ n , m ] = f d]'n(]')exc[n(]'),m(]')]. (20)

For the GGA, 17~E,c has a more complicated form and also depends on IVn(]')l.
4/3
(The local density form of the exchange energy is e, o~ (n~/3 + n~ )/n .) From this
functional form, and the relationship between the densities and the single-
particle wave functions, a set of self-consistent equations can be derived. ~66These
Schr6dinger-like single-particle equations,

1 ~ o }
(2~)

have a spin-dependent effective exchange-correlation potential given by

+
~,.(~r) = ~E'c[n' m] + f i E , , [ n, m ] (22)
&7(~) - fim(~')

This set of single-particle Kohn-Sham equations would, in principle, give the


exact solution to the many-body problem if the exact functional form of the
exchange-correlation potential were known. Unfortunately, the exact form is not
known, otherwise the complete solution to the underlying many-body problem
would also be known. Thus, approximations such as the LSDA or GGA must be
made. The types of properties that can be determined from calculations include
moments and spin densities, and properties derived from investigating the vari-
ation of the total energy with respect to some parameter, e.g., structural properties.
Formally, the eigenvalues and eigenfunctions obtained from the single-particle
equations do not have any physical meeting, but are simply a device to solve the
underlying self-consistent problem. However, from a heuristic point of view (and
with some formal justification166167171), the eigenvalues and eigenfunctions can
be used as approximations to the quasi-particle band structures of the real materials.
There are a number of deficiencies in the LSDA and GGA. For example, the
coupling of the spin and lattice is neglected. Thus, while the relative spin ordering
(ferro- versus anti-ferromagnetic) can be obtained, the spin direction relative to
the lattice is indeterminate. By including spin-orbit effects, the direction of the

171 G. E. W. Bauer, Phys. Rev. B, 27, 5912 (1983).


74 R.E. WATSON AND M. WEINERT

magnetic moment and magnetic anisotropies can be calculated. While spin-orbit


will also induce an orbital moment, Hund's second rule (orbital angular momen-
tum) effects are not correctly handled within the LSDA or GGA. In fact, it can
be shown ~7-' that any approximation that depends only on the charge and spin
density must fail to reproduce multiplet structure correctly, although a current-
density functional theory in principle could. Although various attempts to include
orbital effects in solids have been made, at present none has been theoretically
justified with anything resembling rigor. Even with a number of well documented
shortcomings, calculations using these approximations have been surprisingly
successful in describing a variety of alloy properties. For example, structural
properties such as c/a ratios and internal atomic positions are generally given to
within the experimental uncertainty, although the LDA generally underestimates
the volume. Moreover, the differences in heats of formation are given to a level
where the calculations can make meaningful quantitative predictions of phase
stability.
To solve the density functional equations, there are many possible methods;
which one is used is generally a matter of personal preference and should not
alter the physics. Most methods solve these equations variationally using a basis
set, but non-variational methods based on multiple-scattering techniques are also
in common use. Another important dividing line among calculational methods is
how the core electrons are treated. Although generally chemically inactive, the
core electrons do have an effect on the valence wave functions, both through their
contributions to V, and 14,, and through the orthogonality requirements that deter-
mine, for example, the number and placement of nodes. In all-electron calcula-
tions, the core states wave functions are explicitly (albeit approximately) treated.
In pseudopotential calculations, the core electrons are removed from the problem
and their contributions are approximated by a pseudopotential: although the
valence pseudo-wave functions do not have the correct nodal structure, modern
pseudo-wave functions are constructed to agree with the all-electron ones in the
region of space important to bonding. In all LSDA/GGA methods, including
pseudopotential methods, the core density is necessary for the calculation of
accurate spin-dependent exchange-correlation potentials and energies. Since these
173
functionals are nonlinear, neglect of the core density can result in gross errors.
Many of the alloys of interest are close-packed metals. Moreover, the electronic
structure is dominated by the behavior of the electrons near the atoms, where the
density and potential are rather spherical in shape. This observation lead to the
so-called "muffin-tin'" and "'atomic sphere" approximations that ignore the non-
spherical contributions to the density and potential. While these approximations
can be quite adequate for many problems, and important physical results can be

~7_,M. Weinert, R. E. Watson, and G. W. Fernando Ito be published).


~7.~M. Weinert, S. Bliigel, and P. D. Johnson. Phv.~. Rev. Lett., 71. 4097 i 1993t.
TRANSITION-METALS AND THEIR ALLOYS 75

obtained from calculations with varying degrees of sophistication, recent expe-


rience has shown that there are cases where the full anisotropic shape of the
density and potential ("full-potential") can have important consequences. For
example, despite being close-packed, a full-potential treatment was required to
accurately obtain the large charge redistributions in the gold-alkalis (cf. Sec. 14).
Similarly, the accuracy of calculations for structures such as the Frank-Kasper
phases that have sites with low symmetry benefit greatly from full-potential treat-
ments. As a general rule, one would prefer to solve any problem with a method
that is as accurate as possible simply to avoid uncertainties, especially when
differences in energy -0.01 eV/atom can be important.

17. PHASE STABILITY OF BINARY ALLOYS: Z R - A L

The phase diagrams of binary alloys vary significantly. For the ones of most
interest here, there are generally one or more ordered phases. These phases may
have broad concentration ranges or may be sharp (line compounds). In some cases,
ordered phases may only exist at high or low temperatures. In between the ordered
phases, two-phase regions exist. Determining the relative stability of different
phases may be quite complicated experimentally and theoretically, depending on
the alloy constituents. To illustrate complexity and what information can be
obtained theoretically, we use the Zr-A1 system as an example.
Of all the transition-metal aluminide binary phase diagrams, that of the Zr-A1
system, shown schematically in Fi,,~ . . 40, is one of the more complicated ones ~:a
There are ten reported compounds at various concentrations, all of which exper-
imentally are found to have extremely narrow concentration ranges, only on the
order of 1%. Of these phases, two (ZrsA1, and ZrsAI4) are high-temperature phases
and a number of the others decompose into other phases at high temperatures.
The phase diagram, which consists mainly of two-phase regions, is also rich in
175
a structural sense: The various compounds are found in hexagonal, orthorhom-
bic, tetragonal, and cubic structures, with four of the phases representing proto-
type structures (see Table 2). Because of these inherent structural differences, it
is not possible to describe the different alloys as simple decorations of one
underlying lattice.
The Zr-A1 system is a rather severe test for electronic structure theory since
it must describe not only the competition among different crystal structures at
a given concentration, but also the competition among phases with different

174D. J. McPherson and M. Hansen, D'ans. Am. Soc. Metals, 46, 354 (1954): M. P6stchke and
K. Schubert, Z. Metallkd., 53, 549 (1962): J. Murray, A. Peruzzi, and J. P. Abriata, J. Phase Equil.
13, 277 (1992): A. Peruzzi, J. Nucl. Mat., 186, 89 (1992).
175p. Villars and L. D. Calvert, Pearson's Handbook of Ci3stallographic Data for Intermetallic Phases,
Second Ed. (American Society for Metals. Metals Park, OH, 1991).
76 R.E. WATSON AND M. WEINERT

TABLE 2. PHASES. CRYSTAL STRUCTURES. AND CALCULATED HEATS


OF FORMATION ~ FOR ZR-AL ALLOYS. {AFTER REF. [64].}

Crystal Class Structure ~'pe AH {eV/atom)

ZrA13 tI 16 ZrAI, -0.50


tI8 TiAI3 -0.48
cP4 Cu3Au -0.47
ZrAI2 hPl 2 MgZn: -0.57
cF24 MgCu: -0.56
ti24 Ga,Hf -0.53
Zr2A13 oF40 Zr,AI~ -0.53
ZrAI oC8 CrB -0.46
cP2 CsC1 -0.29
tP2 CuAu-I -0.45
ZrsA14 hP 18 Ga.~Ti~ -0.42
Zr4AI3 hP7 Zr~AI3 -0.42
Zr3A12 tP20 Zr~AI, -0.40
Zr~A13 hPl 6 Mn,Si~ -0.35
ti32 W~Si~ -0.37
Zr,AI hP6 InNi, -0.37
tI 12 AI,Cu -0.31
Zr3AI cP4 Cu3Au -0.31
cF16 BiF~ -0.15
cP8 A 15 -0.26

2000 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

<
1800 X ~ xxxxxx ~ /
,'7
/

1600 L

1400
ii IllI';j
1200
E
y.
1000

800

6 0 0 . . . . . . . . . . . . ~. . . . _ . 1 . . . . . . .

0 10 2o 3o 4o 50 60 70 80 90 100
Atomic Percent Zr

FIG. 40. A phase diagram for Zr-AI based on those in Refs. [6. 174].
TRANSITION-METALS AND THEIR ALLOYS 77

stoichiometries. From the calculated ~4 heats of formation AH for all the reported
phases, as well as a number of other plausible competing ones, many of the
observed features of the phase diagram can be understood in simple terms.
At each concentration, a number of different plausible structures are consid-
ered. While this approach might seem less systematic than, e.g., the cluster vari-
ation method (CVM), ~5~~s4 these types of methods can only be applied to cases
where the different configurations can be expressed as different decorations of
the s a m e underlying lattice. Since this is not the case for the A1-Zr alloys, the
"usual suspects"' approach (greatly enlarged to a wider group of suspects) is the
more appropriate one.
The calculated heats of formation ~4 for a number of different ordered alloys at
different concentrations are given in Table 2 and the results are plotted in Fig. 41.

0.0

0.1 Experiment
0 Lo~T phases
[] High T phases BiF,

0.2

AI5
CsCl
0.3
CuAI~ <) /C) Cu~Au
/

Nln,Si; [] ,"
W,Si, ~'~) InN'i,
0.4 , 9 _

Ga~Ti, [j~ Zr.AI;


CuAuI ~,,'/
Cu.Au ~ CrB @
TiAI: /
z

0.5 ZrAI3 /
GazHf ~ , d Zr:AI
MgCu: @/ MgZn.

.6 ~ ' ' ' ' ' ' . . . . ' ' '


A1 0.25 0.50 0.75 Zr

All_~Zrx

FIG. 41. Calculated heats of formation for the Zr-A1 alloys: experimental data are from Ref. [ 176].
The size of the symbols corresponds to approximately 0.01 eV.
78 R.E. WATSON AND M. WEINERT

Also shown in Fig. 41 are the few existing calorimetric measurements ~v6~Tvfor
this system. The agreement between these data and the calculations is good and
the lowest energy structures are consistent with the observed phases. Note, how-
ever, that the spread in available experimental values is rather large compared to
the calculated differences among crystal structures at a given concentration. Such
discrepancies among different experimentally derived values is common, with
the result that the experimental heats are often incompatible with the observed
phase diagram, i.e., the heats of observed phases often lie above the tie lines
connecting other phases. Although theory obviously is not perfect, a major advan-
tage is that it provides a consistently determined set of values, and can usually
distinguish among close lying phases.
The largest calculated AH is for ZrA12 in the hexagonal Laves structure, con-
sistent with the pronounced skewing of the phase diagram shown in Fig. 40. As
discussed above, AH for the ordered compound must lie below the line connecting
the two competing phases in order for it to be stable relative to a two-phase
mixture of phases with different compositions. In Fig. 41, the tie lines from ZrA12
(hexagonal Laves structure) to Zr2A1 (InNi2) and from ZrA1 (CrB) to Zr~A1
(Cu3Au) are shown. Although these lines (and the other tie lines in this region)
are almost collinear, there is some curvature so that the observed low-temperature
phases are predicted to be stable. The two observed high-temperature phases,
ZrsA14 and ZrsA13, both clearly lie above the tie lines but by an amount that is
small on an absolute scale (-300 K), as is necessary if they are to be high-
temperature phases.
Since heats of the low-energy phases fall all on a nearly common line, only
small changes in stoichiometry might be expected to change the energy of a phase
relative to its neighbors and then rise above the tie lines. Thus, one might expect
that the concentration ranges for the different line compounds are quite narrow
and that the phase diagram consists mainly of two-phase regions, in agreement
with experiment. Further supporting evidence comes from the observation its that
the existence of certain phases such as ZrAI depends strongly on the experimental
conditions. The large number of observed phases is in some sense accidental in
that slight changes in the electronic structure will cause AH of certain phases to
shift, causing phases to disappear. (A decrease in binding energy of a phase will
cause it to disappear, while an increase in binding may suppress neighboring
phases.) Supporting this argument is the fact that the phase diagrams of the
isoelectronic systems 6 Ti-A1 and Hf-AI, although roughly similar to Zr-AI, both
have fewer observed phases. The results for the structural parameters such as lattice
constants, c/a ratios, and internal coordinates are generally in good agreement

~76 S. V. Meschel and O. J. Kleppa, J. Alloys Compounds. 191. 111 (1993).


~77
O. J. Kleppa, J. Phase Equilibria. 15, 240 (1994).
~Ts R. J. Kematick and H. F. Franzen, J. Solid State Chem.. 54. 226 (1984).
TRANSITION-METALSAND THEIR ALLOYS 79

with the existing experimental data, ~:5 although overall little is known experi-
mentally, especially concerning the internal coordinates.
A1-Zr bonding is relatively strong, thus favoring structures with unlike neigh-
bors. The intrinsic size, as measured by the elemental volumes, of Zr is signifi-
cantly (-50%) larger than AI. The standard close-packed structures (fcc, bcc, hcp)
give a poor filling of space and would require the AI-A1, Zr-AI, and Zr-Zr bond-
lengths to all be similar. Thus, crystal structures that have sites with different
sizes should be preferred. As discussed in Sec. 7, arguments based on the size
mismatch can be used to rationalize the occurrence of the CrB, rather the CuAu-I,
structure for ZrA1. Although size effects as such are difficult to quantify based
on electronic structure calculations, the consequences of the size differences can
be seen" The changes in bonding between the CrB and CuAu-I structures cause
the Fermi level of the CuAu-I phase to be about 0.19 eV higher than for CrB. A
simple estimate of the change in the sum of one-electron energies yields a
difference in binding of about 0.02 eV/atom, in almost perfect agreement with
the full calculation.
Size is also an issue in a number of other structures in the Zr-A1 system,
including the existence of the Laves phase for ZrAI: with its 12- and 16-fold
coordinated sites. The small difference in energy between the cubic and hexagonal
Laves is related to band filling and to the additional freedom associated with
relaxation of the c l a ratio and internal parameters, analogously to the fcc-hcp
case. Similarly, for ZrAl~ the antiphase tl16 (D0_>) structure at the ideal c/a ratio
has a heat of formation of about 0.007 eV/atom less bound than those of the
TiAI3 and Cu3Au structures. (The calculated difference in AH between the Cu~Au
and ZrA13 structures is about 0.027 eV/atom, in excellent agreement with the
value ~79 of 0.023 eV/atom extracted from studies of mechanical alloying.) While
the stabilization of these antiphase structures has been attributed to purely elec-
tronic effects, ~~ the relaxations induced by size effects play an essential role in
modifying the electronic structure.

18. TERNARYALLOYS" TRANSITION-METALALUXlINIDES

A more complicated situation occurs when the alloy is composed of more than
two elements. Now the number of possible competing phases increases dramat-
ically, both in the number of different structures possible for a given stoichiometry
and for the number of competing binaries (and higher order) phases at completely
different stoichiometries. Even with the exponentially increasing number of pos-
sibilities, it is still possible to calculate useful information.

179p. B. Desch, R. B. Schwarz. and P. Nash. d. Less Common Metals, 168, 69 (1991).
~80A. E. Carlsson and P. J. Meschter, J. Matel: Res., 4. 1060 (1989).
80 R.E. WATSON AND M. WEINERT

N MN M
FIG. 42. The triangle for the phase diagram of the ternary N-M-A system with the M2NA phase
indicated. The lines, passing through this phase, connect pairs of phases which may compete as two-
phase mixtures. The dotted triangle is a possible three-phase mixture.

Let us limit our discussion to the competition of ordered ternary structures


with two- and three-phase mixtures of binary and elemental phases. In general, the
composition of a ternary alloy can be represented by a triangle similar to Fig. 42.
The tie line construction for determining the stability of a binary alloy needs to
be generalized. For a ternary alloy on the line connecting the two binaries AI_xB~
and B~_,C, has a composition A~_~_x~B,~_~,, .. ,~_,,C~,. The heat of formation for
such a two-phase mixture is given by

A H ( a ) = (1 - a)AH(A~_,Bx) + aAH(B~_,C,). (23)

For a ternary alloy to be stable relative to the two-phase mixture at the same
composition, it must have a heat of formation more binding than M-/(cz). This
equation is the simple analogy of the binary alloy case. In Fig. 42, the position
of the ternary M2NA and some principal lines connecting competing pairs of
phases are shown. The relative lengths of these lines are simply related to a and,
hence to the heats of formation for the two-phase mixture.
In addition to these two-phase mixtures, three-phase mixtures of the elements
and binaries (and/or ternaries) can also occur. The requirement for a possible three-
phase mixture to occur with the correct stoichiometry is that the ternary phase be
TRANSITION-METALS AND THEIR ALLOYS 81

FIG. 43. The BiF3 (D03) structure for a M,NA alloy. The M atoms form a simple cubic lattice, with
the N and A atoms at alternating body-centered positions.

contained inside the triangle formed by connecting the three other phases; although
the example shown has all three vertices on different edges, this is not required.
Note that two-phase mixtures are simply special cases of a three-phase mixture
where the ternary lies on an edge of the triangle connecting the binary phases.
As a example of multi-component alloys systems, we consider the ternary
transition-metal aluminides. The bonding of aluminum with the transition-metals
is of scientific and technological concern. Despite this interest, the properties of
ternary aluminides have received relatively little theoretical attention. The occur-
rence of ordered ternaries may be inferred from inspection of the binary transition-
metal alloys" As discussed earlier, strong line compound formation usually arises
when the transition-metals are from the opposite ends of the row. In addition,
differences in atomic size encourage the elemental constituents to reside on their
respective sublattices. The same is true for the transition-metals when aluminum
becomes the third alloy constituent. Roughly half the ordered ternary aluminides
form in the antiphase Heusler BiF3 structure with a M2NA1 composition. This
structure is a bcc lattice with the M transition metals on the cube comers and the
N and A1 atoms alternately filling the body center positions in the x, y, and z
directions, cf. Fig. 43.
The reported and calculated ls~ occurrences of these transition metal aluminides
are given in Table 3. The phases reported suggest a pattern of occurrence and this
view is supported by consideration of compounds where Ga and In replace A1.
The observed compounds of Ga, whose size is similar to A1, lie in the same squares

~8~R. E. Watson, M. Weinert, and M. Alatalo, Phys. Rev. B, 57, 12134 11998).
82 R . E . WATSON AND M. W E I N E R T

TABLE 3. REPORTED AND CALCULATEI) 1~1 O(CURRE.X('ES ()F T H t - M : N A L CoMPOtXDS ix THE B I F 3


STRUCTURE" CALCL'LATED AND REPORTEI) TO EXIST lot: REPORTED/No C.aLCtLATIO.X I@):
CALCULATED TO EXIST. BUT NOT REPORTED (,~)" CAI_CULATEI) NOT TO EXIST AND NOT REPORTED
(X): AND CALCULATED NOT TO EXIST. BUT REPORTFD (@)

M/N Sc Y La Ti Zr Hf V Nb Ta Cr Mo W Mn Fe Ni

Cu O ~ O ,J Z 0
Ag O
Au O @ 9 x ~ x x x O
Ni O x 9 9 ~ 9 9 ,,.j ~ X ~._/
Pd O x "~ 9 ~ x x
Pt ,'-h

Co ~ ~' ~' ~" ~"


Rh x + ~ x x O
Ir o
Fe x 9 x 9 C c: 9 o 9 |
Ru x ~ ~ .iv, ~. X
Os

Mn 0

of the table as do those of A1. In, which is measurably larger, has compounds
limited to the Sc and Ti blocks of the N elements. Comparison of the elemental
volumes indicate that the compounds occur when the N atom and metalloid have
volumes which are within 30-40c~ of one another. Volume, however, is not the
sole factor controlling the appearance of these phases since the individual columns
of the table are not all "yes" or all "'no" to such phase occurrences.
The test for the occurrence of the ordered ternary involves the competition with
mixtures of elemental and binary phases, thereby requiring estimates of A/-/for
diverse competing binary phases such as discussed for the Zr-A1 system. The
calculated M-/for large numbers of ternaries ~s~ in" the BiFs structure are all (except
for Au2MoA1) negative, implying that they are stable relative to a three-phase mix
of the elemental solids. Of the systems which have been reported, the calculated
estimates indicate that all but one (to be discussed later) occur. Of the unreported
systems, some 9 are predicted to exist while the other 18 should be unstable. The
advantage of having the calculations is the ability to discern which of the unreported
phases should not occur, as opposed to those that are yet to be observed.
In Fig. 44, the competition between pairs of phases with Fe2ZrA1 is seen. The
zM-/is reasonably large, but binaries with significantly larger heats for both Zr-A1
and Fe-A1 occur. (Both ~ ZrA12 and Zr2Al~ have IM-/] > 0.5 eV/atom and are not
shown in the figure.) If only two-phase mixtures were considered, Fe2ZrA1 would
be predicted to be stable, even with the large binary heats. This result reiterates
TRANSITION-METALS AND THEIR ALLOYS 83

A! Fe
0.0 O.. 9
Zr . . ,::.
. . . . . . . . . ,,
/
-:..:. ,,
.::. ,,

m0.1 :::::. 1 i

::::::. 11
:::::-- i"
" "--.-.-...5.... i,,1
~- --0.2
FeZr ~ .... '";:.: ...... 9 Fe~Ai
m ....... >'~,-"J "....iii-~ Fe,Zr
<~ ---0.3
O--.1"1"
Zr,A! ,1
.11. F e ~Z~r,~ ~ Fe~A!
---0.4
ZrA! 9 FeA!
--0.5
relative stoichiometry

Fie. 44. The heats for Fe,ZrA1 in the BiF~ structure and competing phases. The positions and
lengths of the lines connecting the heats of observed (O) and unobserved/hypothetical phases (9
correspond to those shown in Fig. 42. The heat for the most bound three-phase mixture is indicated
by an open square.

the fact that the magnitude of a AH by itself does not determine the occurrence
of a phase" competing phases must always be considered. There are, however,
several three-phase mixtures that have greater binding than the ternary, with the
largest heat for the FeA1-Fe2Zr-ZrAI2 mixture, thus suppressing--though only
barelymthe formation of the ordered BiF3 phase.
With the combined theoretical and experimental results, a pattern begins to
appear, i.e., the stable ternary phases are largely associated with having the N
atom from the Ti column of the periodic table and the M from the Fe column.
The tendency for Rh and Pd (and one suspects Ir and Pt), to not form phases
outside of the Ti group is likely associated with the fact that the binary aluminides
177
of these four elements have substantially larger heats than do the other transition-
metal elements; it is competition from these binaries which suppresses the terna-
ries. In addition, to the 9 as yet unobserved but predicted phases, there is the
suggestion that if the 5d elements were considered, on the order of 20 new stable
phases might be found.
Of the phases which have been observed, all but one, Fe2NiA1, is predicted to
be stable. This discrepancy is troubling since the ordered ternary is predicted to
have a heat of formation (-0.1 1 eV/atom) significantly smaller in magnitude than
a large number of two- and three-phase mixtures. While magnetic effects could
be important, the calculated Fe and Ni moments are not far from the elemental
moments and so this is unlikely to be the explanation. (If there were a systematic
error in the magnetic energy resulting from the local density approximation, this
84 R.E. WATSONAND M. WEINERT

error would be expected to be larger for those systems whose magnetic moments
differ most between the alloy and the elemental metals.) The calculated heats of
the binary Ni-A1 compounds are already large enough to suppress the ternary,
but since these are in good agreement ~82 with experiment, a reference energy
problem is also unlikely. Similarly, the Fe-A1 heats, which are also in reasonable
agreement with experiment, ~77~3~s4 would suppress the ternary. Perhaps more
puzzling is the fact that the calculated structural a n d magnetic properties of
Fe2NiA1 are in excellent agreement with the experimental ls5 values: The ratio of
the volume of the compound to the sum of the volumes of the elemental metals
(Vegard's law) is 0.934 experimentally, while the calculated value is 0.935, and
the experimentally determined low-temperature magnetic moment is 4.25/JB per
formula unit, compared to the calculated value of 4.26/JB.
While it is possible that the calculations are simply unable to reproduce the
binding of this system, this explanation seems unlikely when the w h o l e set of
calculations is considered and the fact that other compounds with these elements
yield M4 values in reasonable agreement with experiment. Moreover, because
Fe-Ni alloys have small heats, it is unlikely that the ternary would have a heat
significantly larger than the calculated one. Another possibility is that the exper-
imentally observed phase is metastable. Since this region of the Fe-Ni-A1 phase
diagram is a bcc field, one expects a bcc-based structure, ordered, disordered, or
. . . . 186
possibly with short-range order. From the experimental side, d~sungmshlng
between BiF3 or CsC1 phases also may be hard, especially in the presence of
short-range order. Further investigations, especially experimental, are needed to
resolve this question.
These types of investigations of ternary (and higher order) alloys can be extended
to include the competition among other ternary phases. The transition-metal
aluminides discussed are somewhat special in that the ternaries are generally well
ordered; often there is site disorder. In any case, understanding the properties of
the multicomponent alloy in order to understand the phase diagram requires
detailed knowledge of the binary alloys. This information is often difficult to
obtain experimentally, and thus there is a definite need for theory. On the other
hand, the large number of competing binaries, etc., make the theoretical deter-
mination of the required heats a major computational undertaking.

~s2K. Rzyman, Z. Moser, R. E. Watson. and M. Weinert, J. Phase Equilibria, 17, 173 (1996) 19,
106 (1998).
183E D. Desai, J. Phys. Chem. Ref Data, 16. 109 (1987).
184
R. Hultgren, E D. Desai, D. T. Hawkins, M. Gleiser. and K. K. Kelley, Selected ~llues o(
Thermodynamic Properties of Binary Alloys iAmerican Society for Metals, Materials Park, OH,
1973).
185
K. H. J. Buschow and P. G. ,,'an Engen. J. Ma,~. Ma,~. Mater:, 25, 90 11981).
1s6R. Marazza, R. Ferro, and G. Rambaldi. J. Less Common Metals. 39, 341 11975).
TRANSITION-METALS AND THEIR ALLOYS 85

19. ENTROPYCONTRIBUTIONS

The thermodynamical quantity determining the phase stability is the Gibbs free
energy G(P, T), i.e., the free energy at constant pressure and temperature. At T = 0,
this reduces to the enthalpy. Thus, AH determines the phase stability at low
temperatures, but at higher temperatures entropy effects may play an important
role in determining phase stability. There are a number of contributions to the
entropy: electronic, vibrational, and configurational.
Electronic entropy arises in the finite temperature generalization ~'~ of density
functional theory, and takes the form

Sel - - k e E [ f , ln f , + (1 - f,)ln(1 - f i)], (24)


i

where the sum is over all electronic states i with (Fermi-Dirac) occupation numbers
lg7
f. These types of terms are simple to include in calculations and are necessary
in order to maintain the variational nature of the density functional equations.
Although the electronic entropy terms are not large, the differences among phases
can be a large fraction of the total entropy differences, j'~s The effect of electronic
entropy on the relative stability of the hcp and bcc phases of Ti was already
shown in Fig. 13.
The vibrational modes of the crystal are an important contribution to the free
energy of the system: the magnitude of the vibrational entropy is generally larger
than the electronic one, as manifested in the temperature dependence of the
specific heat. (The differences in vibrational entropies among phases are compa-
rable in magnitude to the electronic entropy differences.) The vibrational free
energy can be written in terms of the phonons (when they can be defined) as

hi_o; )
F,i b - ksT ~ In 2sinh2kBT , (25)
i

where the sum is over all (possibly temperature-dependent) phonon frequencies 60,.
This form follows simply from the calculation of the partition function, assuming
the vibrational energy of the system for a set of phonon occupation numbers {hi} is

E, ib = [{ n i }] = z(l)
i
n/+ ~ h o~i. (26)

18v M. Weinert and J. W. Davenport. Phys. Rev. B. 45. 13709 t1992~.


18s R. E. Watson and M. Weinert. Phys. Rev. B. 30. 1641 t1984t.
86 R.E. WATSONAND M. WEINERT

The entropy is then given by

3
Svib - - -~--~Fvib (27)

h COg"] /'1COg T 3 oai h COl ~.


= .28.
I

The high-temperature stabilization of the bcc lattice is the prototypical example


of the importance of vibrational contributions. Zener argued 34 that the more open
bcc structure is softer, having a low-frequency transverse phonon mode that favors
phonon excitation. Thus, the phonons would contribute a larger entropy term for
the bcc lattice than the closed-packed structures. Friede135 instead proposed that the
larger vibrational entropy in the bcc lattice arose from a lower Einstein frequency,
as might be expected from the smaller number of nearest neighbors. Basing the
estimate of the entropy difference between the structures on the ratio of nearest
neighbors in the fcc/hcp (12) and bcc (8) structures, gives

3 3
ASbcc-cp -- 5 In ~. = 0.6 ks/atom.

For a temperature of order 1000K, the contribution to the free energy is about
0.05 eV/atom, which is significant on the scale of the energy differences between
the bcc and fcc/hcp phases.
As discussed in Sec. 5, many of the transition metals are mechanically unstable
in their high-temperature phases, i.e., some of the phonon modes have ~ < 0. In
this case, the simple form of the vibrational free energy given by Eq. (25) is no
longer appropriate since the energy of a given configuration is not given by Eq. (26).
A complete solution to this problem is still lacking. One approach that should in
principle work is to calculate the energies for different configurations, calculate
the partition function, and then obtain the free energy. In Fig. 45 the calculated
energy for the frozen phonon corresponding to the unstable N-point T~ mode in
Ti is shown for two different electronic temperatures. Although this mode is
unstable for the ideal bcc structure, a minimum with respect to the displacement
does occur; the position and energy of the minimum depends on the electronic
entropy contributions. This mode is highly anharmonic, as seen in the righthand
side of the figure. These strong anharmonic terms will renormalize the frequencies
via phonon-phonon interactions. It is also possible to make a quasiharmonic-type
approximation about the minimum of the unstable phonons, but this too will require
the inclusion of anharmonic terms in order to account for the shape of the potential
energy surface.
TRANSITION-METALS AND THEIR ALLOYS 87

25 ,, , ,, . . . . .

-- L
E
J -5
0
0
9
2 o
-10
>
(D 0
g -15
0
u.I -25 9 T= 315K
0 T= 1260 K
-20 t

-50 -25
0.00 0.05 o. 10 0.000 0.005 0.010
6/abcc (6/abet) 2

FIo. 45. Calculated energies of the frozen phonon as a function of the displacement 5corresponding
to the N-point T~ mode in Ti for two different electronic temperatures. The lefthand panel gives the
energies as a function of the displacement. The righthand panel .,,ho~vs the anhannonic contributions:
a harmonic system would be represented by a horizontal line.

The last major entropy contribution is configurational entropy, i.e., the contri-
bution arising from having many different atomic orderings (configurations) close
in energy. It is this term that is largely responsible for the disorder of alloys at high
temperature. For an ideal mixture of A and B, the entropy is easy to derive. For
a lattice with N sites and a concentration A ~_,B,, the number of different possible
configurations is

N~
N, = (29)
(xN)!(N-xN)!

The entropy is given by S , - INN,. Using the standard approximation In v! = v


In v - 3; gives the entropy of mixing for an ideal binary, as

S
= - k s [ x l n x + (1 - x ) l n ( 1 - x ) ] . (30)

For x - ~, the configuration entropy per atom is -0.69 kR, which is comparable
to the Friedel value for the difference in vibrational entropy discussed earlier. Thus,
configurational entropy is expected to be reasonably large.
The configurational entropy for alloys can be calculated in different ways.
Given a set of effective cluster interactions, the CVM method provides a proce-
dure to include the entropy contributions in the free energy. For a binary alloy,
however, the N~ configurations of Eq. (29), where each lattice site is occupied by
either an A or B atom, are not the only possible ones: configurations with 11o
atoms on a site are also possible. The importance of these configurations will depend
on the energy needed to create a vacancy: for some systems (such as NiAI) vacancy
88 R.E. WATSON AND M. WEINERT

formation is the dominant mechanism for accommodating non-stoichiometric


concentrations. Thus, vacancies often must be treated as a third "atom" type,
greatly increasing the number of ECIs needed to describe the system. Since in
general the relative importance of vacancies is not known beforehand, vacancies
should be included in CVM-type approaches from the beginning.
To deal with systems with varying stoichiometry, it is often useful to expand
about an ordered phase, instead of the global expansion of the cluster method.
In this approach, the starting point is some ordered phase. The energies of the
possible intrinsic defects--antisites and vacancies--are calculated. If the concen-
tration of defects is low and the interaction energies among defects are small
compared to the defect formation energies, then the problem can be cast in terms
of the energies of the isolated defects, the stoichiometry of the alloy, and the
temperature.
To treat this problem, consider alloys around the concentration A ~_~oB~. Let
there be two sublattices, c~ and /3, that at the ideal concentration xo would be
occupied by A and B atoms, respectively. (The generalization to multiple sublat-
tices is straightforward and will not be discussed explicitly.) Rather than fixing
the number of atoms of each type, the number of atoms will be variable, i.e., we
are working in the grand canonical ensemble. Then the grand potential is given by

= U - T S - PaNA - PBNB, (31)

where p, is a chemical potential for atom type i. The number of A and B atoms
(NA, NB) are

Na = (1 _ xo)N _ N ,a, - N B + a N~ (32)

N B = xoN - N~. - N ~A + N B , a (33)

a
where N,a, is the number of vacancies on the a sublattice, N8 is the number of
B atoms on the o~ sublattice (antisite), and the other terms have similar meanings.
The internal energy U of the configuration is

v - N + + + + d, (34)

where e0 is the energy of the ideal alloy (per atom) and e~ is the difference in
energy between the crystal with a single defect of type D on sublattice )'and the
TRANSITION-METALS AND THEIR ALLOYS 89

ideal crystal. These energies can all be determined from first-principles (supercell)
calculations. The last major ingredient is an expression for the configurational
entropy, which can be determined by simple counting arguments similar to the
ones used to obtain Eq. (30):

S (x~N)'
= ~ln . "~ ~ . (353
NaVN
. B ' ( x.c ~ N N, - Ns) w.

Starting from these equations and the calculated defect energies, the thermo-
dynamical properties of the alloy can be derived. First, minimizing f~ with respect
o~
to Ni yields the concentration of defects

ct
(~ +,ua)/kBT
a e
r = (1 - x o ) (Ea ct (36)
+~4)/k~ T (EB+~A-~B)./kBT
l+e +e

a
(EB - u 4 - ~B)"'kB T
a e
c 8 = (1 - % ) ce . (37)
(E + u a ) / k B T (~B+~la-tdB)/kB T
l+e +e

The composition x of the alloy is then given in terms of the defect concentrations
by

1-x ( 1 - x , , ) - c ~a- c s +
c~ C~
A
(38)

which differs from the ideal concentration. To fully specify the problem, the
Gibbs condition (P = 0) relating the chemical potentials and the free energy

r = ~ G = l.liN i = U- TS (39)
P.T z

is imposed. Thus, given a set of defect energies E,, these equations can be solved
(numerically) for ~, for any given choice of x and T. This approach for including
configuration entropy contributions to the free energy has been successfully
applied to a number of different systems, including FeA1,~s9~9~PdA1, NiAI, ~9~
Ni3Sb, 191 and TiA1.192
90 R.E. WATSON AND M. WEINERT

0.5

Zr
> o.o

:"=
E -0.5 AG(T,x)
O 50 K
Q..
1300 K
-1.0
.
o
I

E P" AI
x: -1.5
O

-2.0
0.70 0.75 0.80
Zr.AII_x

FIG. 46. Calculated chemical potentials tk)r Zr and Ai as a function of concentration around Zr~AI
at low and high temperatures. The change in Gibbs free energy (relative to the elemental metals) is
also shown.

To illustrate the effect of configurational entropy, we return to the Zr-A1 alloys


discussed earlier. In Fig. 46, the chemical potentials for Zr and AI in the Zr3A1
structure as a function of concentration are given. Since the chemical potential
p; is the energy needed to increase the number of atoms i by one, the chemical
potentials are a measure of the contribution of each atom type to the heat of
formation (=AG (T = 0)). Several points are worth noting. First, Zr and AI have
significantly different contributions to the binding, with AI contributing more than
Zr in this concentration range, as expected for the minority atom. Making attri-
butions of relative contributions to a heat of formation requires such an analysis
and depends on having information about the intrinsic defects: the standard
approach of simply dividing the heats equally is obviously incorrect. Secondly, at
low temperatures the chemical potentials show a jump at the ideal concentration x,.
Although perhaps surprising at first, this behavior must occur: At x,, both sub-
lattices are filled. To increase, for example, the AI concentration requires putting
A1 on the Zr sublattice: conversely, to decrease the A1 concentration, AI is removed
from the A1 sublattice. (Similar types of processes for the Zr are also involved.)
Thus, changing the concentration away from the ideal stoichiometry x, involves

~s~j. Mayer, C. Elsfisser, and M. Ffihnle. Phys. Star. Sol. (t7) 191,283 (1995).
190C. L. Fu, Phys. Rex: B, 52. 3151 (1995).
~'~ G. Bester, B. Meyer, and M. F~hnle, Phys, Re~: B, 57, R11019 (1998).
1~2C. Woodward, S. Kajihara, and L. H. Yang. Phys. Re~: B. 57. 13459 (1998).
-- -- o U ~ ~q
ca
-zf -zf
E
~>
ca
= ~ =~ 8 ,.., N
9~ . - ~-~ <. o
~,~ .~ o= ~ '-- d ~
~ .~
E
N,--
,.
r
e,O ,~
.,-.,
Ca
9
-a "~ ---
0 ~ < ~ .
E
t.,
~,.~ ~, ~ g~ ~ ~._
- ~.g~ ~o ~o 9d ' ~ <
>7 ~t ....:.
= ~,- 8.g E'~ I
< -<,
=a. ca
~ 0 ~ .~ 0 ~- ~. ~
+ +
e->
...a s <1 cq c-i
~ ~ = = ."d
0 ca
II II ii ",~, ca ~ o ~ ~ ca
ca
E
o .- ~.~ ~ .~.Z
u 9~ E8 .-~ "~ ~ o~ .. o o ~J
o-- <i eq -- =d
mt
z ~, <. = ,~- o <m
<
=
~' ~ ' 2 - ~ ~ ~ o
~-~ .,--~ .,-~
.,.~ .,-.~ .-..,6
o
9 -~ ~ .,-, .,.~ <
~ ~ b b
Ca ~ "-" <
~D ~D
.,-~ ~.=~ > ~ ~ ~ U
I:~ .,,,~ ~ .,,,~
o L~ ~4 ~
~ N o ~ ~ -g >" ~ ~' :{
,..-,
o
92 R.E. WATSON AND M. WEINERT

---0.25 / /

E
o --0.30
> // . . . . . '
0
II
n
~" ---0.35
_

T=1300 K
-/- ..
- / / /

--0.40
0.60 0.65 0.70 0.75 0.80
ZrxAI,_x

FIG. 47. The calculated free energies {including configurational entropy only) for the Zr_,AI and
Zr3AI phases as a function of concentration for different temperatures. The tie lines (short-long dashed
line) connecting the T = OK energies of adjacent phases are also shov~n.

With the calculated chemical potentials, it is a simple matter to get the Gibbs
free energy as a function of concentration. The calculated free energies for Zr2A1
and Zr3A1 are shown in Fig. 47. For low temperatures, the free-energy curves for
both systems lie above the tie lines connecting the different phases. This result
suggests that both alloys are line compounds with no appreciable concentration
width, consistent with the earlier discussion based on the z3d-/values alone. For
both alloys, the free energies for the Al-rich systems are m o r e binding than the
alloys at the ideal concentrations. This picture differs from the naive expectation
(often seen in books) where the energy increases on either side of x,,, but is the
more general case. These calculated results can be rationalized by noting that
ZrA1 in the fcc-based CuAu-I can be obtained from the fcc-like Zr3AI in the
Cu3Au structure by A1 antisites" ZrA1 in this structure has a significantly larger
heat than Zr3A1.
At higher temperatures (Fig. 47), the free energy becomes more binding, but
the changes are not uniform. For both alloys the changes on the Al-rich side are
larger, reflecting the larger number of configurations possible when a minority
atom replaces a majority one. These changes in free energy can change the concen-
tration widths, or even make phases appear and disappear. We again emphasize
that the change in the free energy for a given phase will not by itself determine the
stability of the phase" instead it is the competition with other phases that is
important. Likewise, the comparison among phases must be done at the same
temperature. Thus, the fact that the high-temperature free energies for the Al-rich
alloys drop below the T = 0 tie lines does not mean that these alloys now have
large concentration widths. The results for these two systems suggest that these
are still line compounds with a two-phase region between them. If the competing
bcc and hcp Zr phases are considered, the Zr3A1 phase may then actually be
TRANSITION-METALS AND THEIR ALLOYS 93

suppressed at these temperatures. ~93As seen in the figure, configurational entropy


can be a significant contribution to the free energy, but to determine the temper-
ature dependence of phase stability, vibrational and electronic entropy need to
be included, especially if the competing structures differ significantly in their
elastic or electronic properties.
This approach for including configurational entropy is quite general, and can
be applied to alloys of transition-metals and metalloids of diverse crystal struc-
tures. The applications to the aluminides have arisen from technological concerns
and from questions concerning the relative importance of antisites and vacancies,
but it has clear relevance to the transition-metal oxides, nitrides, and carbides, where
vacancies are an essential feature.

VII. Magnetic Moments and Site Volumes

Magnetism in transition-metals alloys is a broad topic. The formation of magnetic


moments, 194 their sizes and their alignments, are fundamental to the properties
of many technologically important materials. This section will be concerned with
some trends in moment behavior and their relationship to the volumes of the
atomic sites at which they reside. Three experimental samplings of moment values
will be encountered here, the first being saturation magnetic moments. Except
for ferromagnets with but a single type of magnetic site, this does not provide a
measure of individual site moments, but it does provide a bound on other sam-
plings. The second is neutron diffraction for which there is not always a suitable
sample. The third to be invoked is the measurement of the hyperfine field, Hhf,
at the atom's nucleus on the site in question. The primary source of such a field
is due 195 to the electron contact spin density at the nucleus and involves s (and
relativistic P~/E) electrons which are the only electrons to have density there. The
major contribution arises from the exchange polarization of the core s electrons,
e.g., the ls, 2s, and 3s of Fe, by the magnetized valence electrons. The next
contribution comes from any net s spin density in the occupied valence bands
and any exchange polarization of the paired s-orbital character in those bands.
To a good first approximation Hhf may be related to the moment ~ at the site by
Hhf = A~. However, the constant of proportionality, A, does depend on details of
the magnetization density at the site and on the environment in which the site
finds itself. The choice of A will be an issue in determining the magnetic moment
values of the compounds to be considered here.

194M. Weinert and S. Blfigel, Magnetic Multilavers, L. H. Bennett and R. E. Watson, eds. (World
Scientific, Singapore, 1994), p. 51: and references therein.
~95A. J. Freeman and R. E. Watson, Magnetism, 2A, G. T. Rado and H. Suhl. eds. (Academic Press,
New York, 1965), p. 167; R. E. Watson and A. J. Freeman. Phys. Rev., 123, 2027 (1961).
94 R . E . W A T S O N A N D M. W E I N E R T

3.5 , , , , , , , , , , e, , . , 3.5 r r 1 T T ~ 1 T T ~ 1 v r v t

3.0 Fe ,N: 9 "~.(1


~4- Pr:Fe _N~
41,
Nd.Fe :B 9149
2.5 2.5

= 2.0 2.(I bcc Fe


E 9 FerN l A 9 9 -0
E
.'- 1.5 1.5 9 O JP ~ Nd.Fe

1.0
Y l .()
Fe ctMn

0.5 0.5 L

/
0.0 ~ ........ t ().0 , , ~ , , ,, . . . . l , 1 ,
11 12 13 14 11 12 13
Volume (~3) Volume (A3)

FIG. 48. M a g n e t i c m o m e n t versus W i g n e r - S e i t z cell v o l u m e for bcc and fcc Fe. for ct M n . and for
Fe sites in a n u m b e r o f c o m p o u n d s . T h e bcc and fcc Fe m o m e n t s are calculated: all others are interred
f r o m neutron diffraction e x p e r i m e n t s or (for Fe~,N:) f r o m hyperfine field m e a s u r e m e n t s , fcc Fe has
a h i g h - s p i n and l o w - s p i n transition and over a range o f v o l u m e t d a s h e d line), the h i g h - s p i n state is
metastable.

20. FE AND MN

The magnetic moment of bcc iron decreases ]~e'~: under hydrostatic pressure, as
does its unit cell volume. This trend, as obtained by all-electron full-potential
calculations over a wide range in volume, appears in the first panel of Fig. 48.
The slope (dM/dV) at the observed bcc Fe lattice volume is in agreement with
experiment. The ability to vary the moment can be exploited when fabricating
compounds, and has proven of technological relevance: One desirable property
is a large moment or, strictly speaking, large moment density in a magnet.
As has already been noted (Sect. 10 and Table 1 ), the a Mn structure is closely
related to the Frank-Kasper phases. It has four sites, one of which is 12-fold
coordinated, one 14-fold, and two 16-fold. The volumes of their Wigner-Seitz cells
increase measurably with increasing coordination. The system is ]gs antiferromag-
netic with some non-collinear components to the spin density. The single-crystal

n96 e.g., E. I. K o n d o r s k i i and V. L. Sedov. Soviet Phys. JETP. 11. 561 (1960).
~9v
e.g., J. S. K o u v e l a n d R. H. W i l s o n . J. Appl. Phys.. 32. 435 (1961).
~s T. Y a m a d a . N. K u n i t o m i . Y. N a k a i . D. E. C o x . and G. S h i r a n e . J. Phys. Soc. Japan. 2 8 . 6 1 5 (1970).
TRANSITION-METALS AND THEIR ALLOYS 95

neutron diffraction results ~9~ are plotted in the figure. There is almost an order of
magnitude increase in moment on going from the 12- to the 16-fold sites, accom-
panying a 15% increase in site volume. It is the ability of Mn to magnetically
accommodate itself to sites of measurably different sizes that allows it reside in
a crystal structure appropriate to a two or more component alloy system. The
trend seen in the figure, however, should not be attributed to volume alone. The
16-fold sites are (0,0,12,4) Frank-Kasper environments having four 6-fold, - 7 2 ~
disclination bond lines, with the four associated nearest neighbors lying in par-
ticularly close approach and sharing six common nearest neighbors. In addition
to having fewer neighbors, the 12-fold site has no 6-fold connected nearest
neighbors of this type. These differences can be expected to have measurable
effects on the bonding and moments at the two classes of sites.

21. THE IRON NITRIDES

For magnetic recording, there is a need for environmentally stable particulate


materials with high saturation magnetization and high coercivity. Fe particles do
quite well, but tend to oxidize, significantly degrading their magnetic properties.
The search for other materials has led to investigations of the iron nitrides, three
of which are represented in Fig. 48. This shows some Fe sites to have unusually
large moments for metallically bonded iron9
As was noted in Sect. 11, N when alloyed with transition-metals is generally
found in a (sometimes distorted) octahedral (0,6,0,0) environment. In the case
of Fe4N, the Fe atoms lie on an fcc lattice with the nitrogens lying at the body
centers. The smaller nearest-neighbor Fe sites (3c) are at the faces of the unit
cell, and the more distant, larger volume (la) Fe sites are at the corners. This
material was measured to have a saturation moment ~'~ of 2.2 p~ per atom, almost
identical to that of bcc Fe. Employing this value when refining his neutron diffrac-
tion data, Fraser obtained 2~) moments of 2.01_+0.04 and 2.98_+0.12 YB that are
plotted in the figure. Subsequent average saturation moment measurements
2{)1
yielded smaller values. The most rigorous spin-polarized LDA calcula-
9 ~ 0 "~ "~O~ "}O.1
nons ............. reproduce Fraser's value for the large moment Fe and are some 0.2
/,re larger than his value for the small moment site.

~'~ C. Guillaud and H. Creveaux, Compt. Rend., 222. 1170 t 1946}.


20o B. C. Fraser, Phys. Ret:, 112, 571 (1958).
2{}~X. Bao, R. M. Metzger, and W. D. Doyle, J. Appl. Phys.. 73. 6734 r 1993} and references therein.
-~~ R. Coehoom, G. H. O. Daalderop, and H. J. F. Jansen Phys. Rev. B, 48, 3830 (1993) and references
therein.
2o3 G. W. Fernando, R. E. Watson, M. Weinert, A. N. Kocharian, A. Ratnaweera, and K. Tennakone,
Phys. Ret: B, 61,375 (2000).
2~j. M. D. Coey, J. Appl. Phys., 76, 6632 (1994) and references therein.
96 R.E. WATSON AND M. WEINERT

Fe-ll

FIG. 49. The crystal structure of Fe~6N_,. Note the three different Fe sites and the compressed
octahedral environment around the nitrogen.

Hyperfine fields have been measured 2~176 for Fe4Ni. While there is some scatter
in the sets of results, it appears that when compared with Fraser's moment values,
the scale factor A is almost identical for the two sites. What is more, it is some
20% smaller than that for bcc Fe, i.e., for a given hyperfine field the moment is
20% larger in the compound. Because the more modem measurements of satu-
ration magnetization of the compound are 20+10% smaller than the value
employed by Fraser, a similar reduction in the weighted average value of the
moments plotted in the figure is expected. It is not obvious from Fraser's analysis
that the individual moments would scale in this way, although such scaling
appears to be within his stated error bars. One implication is that the LDA calcu-
lations have overestimated the magnetic moments. Another is that the average A
for the compound is close to that of bcc Fe. There is, however, no reason to
expect the bcc elemental metal and the fcc-based nitride to have essentially the
same A.
FeI6N2 is a metastable phase discovered by Jack. 2~ It has a body-centered-
tetragonal structure (Fig. 49) consisting of a bcc lattice of Fe into which a N is
inserted on every other (001) cell edge, and the structure then allowed to relax.
The Fe II (4e) sites, which lie along the edge, are closest to the nitrogen and have
the smallest Wigner-Seitz cell volumes. The symmetry at this site is not obviously
bcc-like. The bcc-like Fe III (8h) and the Fe I (4d) sites lie increasingly farther
away, with increasing site volumes.

205 K. H. Jack, Proc. Roy. Soc. London, Set. A. 208, 200 (1951).
TRANSITION-METALS AND THEIR ALLOYS 97

This phase is difficult to fabricate and is usually done with thin film techniques.
The result is a multiphase material, often with as little as half the sample being
Fe16N2. This has complicated experimental estimates of the average, as well as
the individual, Fe moments. Average values ranging from 2.2 to 3.5/,tB have been
published; the most reasonable choices appear to range between 2.3 and 2.9 ~B-
Granted the problems with samples, experimental estimates of the individual site
moments in Fe16N2 have, to date, inevitably involved hyperfine field measure-
ments using the M6ssbauer effect and ~VFe NMR. The NMR values :~ plotted in
Fig. 48 were obtained using the traditional FerN A factor, yielding an average
moment of 2.9 ~e- Going to an A based on the updated Fe4N saturation magne-
tization data lowers this average value, as well as the plotted points in the figure
(and those for FerN) by something on the order of 20%. M6ssbauer data for
Fe~6N2 interpreted employing 2~ an average of the A values for bcc Fe and the
traditional Fe4N value yield an average moment of 2.57 ~8. This value is consistent
with the NMR one, granted the different choice of A factors; the large site
moments are 3.5 and 3.05 /.tB from the two investigations, respectively. Full-
potential LDA results 2~176 yield an average moment of 2.37 Pe and a large site
moment of 2.82/,ts. Introduction of a Stoner-Hubbard type enhancement factor
yields essentially the same results for U = 1 eV, a value considered appropriate
to bcc Fe. 2~ If one insists that larger moments than these exist in this system,
strong correlation effects of unspecificed origin must be invoked.
Calculations offer insight on another matter. The introduction of nitrogen
increases the lattice volume and the magnetic moments, but does the proximity
of a nitrogen to an Fe site inhibit the moment? Calculations 2~176 for the FeI6N2
lattice with the N removed yield a slightly increased average moment associated
with a moment increase at the nearest lying Fe II, a smaller increase at the next
closest Fe III, and a reduction at the large moment Fe I site. The situation is more
straightforward for the Fe4N system because the removal of the N leaves an
expanded fcc Fe for which all sites are equivalent and must have the same
moment. Recalling that the nitride values are, on average, probably some 20%
too high, and shifting the values of the two sites downward to account for this,
the large moment site is (within the uncertainties) in line with the calculated
elemental moment curves, while the small moment site lies well below. This
suggests that the latter has its moment substantially reduced by its proximity to
the nitrogen, though its very proximity reduces the Fe site volume. Although the
numerical uncertainties are greater, the Fe~6N2 moments, upon being lowered,
probably display a similar pattern of behavior--an apparent reduction due to
nitrogen bonding for the near neighbors, with the moment of the large site lying
at, or only modestly above, the elemental curves.

2O6y. D, Zhang, J. I. Budnick, W. A. Hines, M. Q. Huang, and W. E. Wallace, Phys. Re~: B, 54, 51 (1996).
2or j. F. Janak, Phys. Rev. B, 16, 255 (1977).
98 R.E. WATSON AND M. WEINERT

As it now stands, the case of the iron nitrides is beset with sample and experi-
mental problems. Fraser's pioneering investigation predates modern neutron dif-
fraction techniques, the availability of theoretical form factors, and the newer
saturation magnetization measurements. A new study would be welcome. Such
a neutron diffraction study of Fe4 N might go far in clearing up the properties of
this compound, though it would likely offer only modest benefit to understanding
the 8:1 compound. There remains the issue of the scale factor A, its overall value
for some compound, and whether it can vary significantly from site to site. If the
bonding between N and neighboring Fe has significantly Fe s-p contributions,
there could well be significant variation in valance contributions to the hyperfine
fields, and hence A, from site to site.
There is the possibility that LDA theory may be in some difficulty with these
compounds. Normally full-potential spin-polarized treatments quite accurately
reproduce the experimental moments of Fe and its compounds. If the experimen-
tally derived values are taken seriously, then the predictions for the nitrides likely
lie outside the normal range of errors with the predictions for the 4:1 compound
lying high and/or with those for the 8:1 being low. However, since the experimental
values are not direct measurements, it is difficult to come to any firm conclusions.
Although Fe~6N2 may not have a "giant" average moment of the order of 3 ,us,
it is a stronger ferromagnet than is bcc iron. Efforts will continue, perhaps by
the inclusion of additional alloy components _,lls to improve on its usefulness as
a ferromagnetic material. Central to its properties, and those of Fe4N, is the
dependence of magnetic moments on site volumes.

22. THE RARE-EARTH--TRANSITION-METAL HARD MAGNETS

For a good "hard" permanent magnet, one would like a large magnetic anisotropy
and a large maximum energy product, (BH)m~,, which is the maximum of the
product of the applied field and the magnetic induction in the second quadrant
of the hysteresis loop. As a rule, the larger the energy product, the less magnetic
material is required for a given application. In the rare-earth-3d hard magnets,
the interaction of the aspherical rare-earth 4f charge density with the crystal
field 2~ is largely responsible for the anisotropy, while the 3d moments are impor-
tant to the energy product. These magnets have Frank-Kasper-like strucures (see
Table 1 and Sect. 10) with the transition-metals in 12-, 13-, and 14-fold coordi-
nated sites, with the rare earths typically in much larger 20-fold sites, and with
the rare earths connected to each other and to the larger 3d sites by 6-fold,-72 ~
disclination lines.

208
e.g., see M. Takahashi, H. Takahashi. H. Nashi. H. Shoji. T. Wakiyama, and M. Kuwabara, J. Appl.
Phys., 79, 5564 (1996).
TRANSITION-METALS AND THEIR ALLOYS 99

The first magnet of this type to be utilized was SmCos. Its uses ranged from
consumer headsets to magnets for wigglers and undulators at synchrotron light
sources. Unfortunately, its two components are in short supply and myth has it that
General Motor's management asked its researchers to find alternatives employing
more available and cheaper materials. Whether the tale is true or not, its researchers
and workers elsewhere succeeded. Three of these materials are represented in the
righthand panel of Fig. 48. The moments are from neutron diffraction studies at
4 K for Pr-,Fel7N,,.821~and for Nd:Fel4B 2~ and, at room temperature, f o r NdeFel7. 21'
The Curie temperature for the latter is not much above room temperature and the
reduced moments obtained for this system are presumed to be due to this effect.
The Pr compound involves investigating the effect of inserting N and expanding
the lattice. The 2"17 structure has an octahedral vacancy site that accommodates
3 atoms per formula unit, and was almost filled in the sample. The nitrogens lie
closest to the Fe (18f) sites, which have the smallest site volumes and moments
among the four Fe sites. Comparison with M6ssbauer hyperfine field data for the
unnitrided Pr2Fe~7 suggests that there is an increase in Fe moments with an increase
in lattice volume due to the nitriding. As for effects due to nitrogen proximity,
there is little hint in the plotted Pr_,Fe~7N2 s results of Fe moment loss on the scale
suggested by the iron nitrides.
The large volume, large moment site in Nd_~Fej4B, like the largest sites in the 2 1 7
compounds, is a 14-fold coordinated (0,0,12,2) Frank-Kasper site, the only such one
in each of the compounds. All the others sites are 12- or 13-fold coordinated. Also,
these 14-fold sites are relatively isolated from the B or N metalloids (when such
metalloids are present) and are involved in rare-earth-3d disclination lines. The
results plotted for Nd_,Fe~4B do present some difficulties: The two smallest volume
sites (the 12-fold Fe III 16k and the 13-fold Fe II 16k Fe sites) are reported to
have much larger moments than do the next two. This behavior represents a real
break in the trend of having increasing moments with increasing volume. However,
as Herbst's Table IX shows, ~5 various hyperfine field-based estimates e~~ indicate the
moments of the first four (or five) sites to be roughly the same. Either the A factors

:0~ The orientations of the rare-earth site crystal fields in these magnets appear to be associated with
the 6-fold - 7 2 ~ disclination lines of ,ahich a number pass through a site lsee Sect. 10 and Table l).
The "'major" disclination lines, which define the crystal field axis. are those involving some rare-
earth-rare-earth nearest neighbor pairs te.g., those in the basal plane for Nd,Fe~.~B. and parallel to
the c axis for SmCo~ and Nd2Fe~). At low temperatures, the ea.,,y axis of magnetization lk)r rare earths
with prolate 4f shells is parallel to these lines and. for oblate .,,hells. normal to them: the sign of the
crystal field is equivalent to having positively charged nearest neighbors on the major disclination
line. There are some reported exceptions to this rule. Several cases, involving Nd and Ho that couple
weakly to a crystal field, have canted magnetic structures at lov, temperatures and go over to the
above rule at elevated temperatures.
_~0j. K. Stalick, J. A. Gotaas, S. F. Chene. J. Cullen, and A. E. Clark. Matep: Lett.. 12. 93 11991 ).
21~ D. Givord, H. S. Li, and F. Tasset, J. Appl. Phys., 57, 4100 (1985).
2~2j. F. Herbst, J. J. Croat, R. W. Lee. and W. B. Yelon. J. Appl. Phys.. 53. 250 i19821.
100 R.E. WATSON AND M. WEINERT

vary substantially from site to site or the neutron diffraction-based moments are
wrong. The latter may be more likely, but the former, if true, more interesting.
Taken as a whole, Fig. 48 displays clear trends showing magnetic moments
tracking site volumes. Of the systems considered, only a Mn ranges from having
12- to 16-fold sites, and it displays a substantial variation in magnetic moment.
The presence of near-lying metalloids tends to reduce 3d site volumes, and the
presence of 6-fold disclination lines tends to be correlated with the volume of a
site. The extent to which 6-fold bond lines do contribute to the moments in a Mn
and in the compounds, and to what extent metalloid-3d bonding affects the
moments have yet to be disentangled. Nevertheless, site volumes appear to be
important to the magnitude of a site magnetic moment.

VIII. Band Gaps in Metallically Bonded Materials

Despite a general expectation (or prejudice) that alloys of transition-metals with


other metallic elements also should be metallic, there are exceptions. The semi-
conducting gold-alkalis, considered in Sect. 14, offer an admittedly extreme case
of metals becoming semiconductors upon compound formation. In other systems,
strong "correlation" effects are invoked to rationalize nonmetallic behavior. The
reason for discussing these systems here is that factors encountered earlier in this
article, such as the relative positions of d-band centers of gravity and the relative
sizes of constituent atoms, have significant impact on whether or not a (pseudo)gap
forms. In addition, this topic highlights the close relationship between the crystal
structure and electronic properties.

23. FESI AND OTHER MATERIALS

FeSi is a metallically bonded compound which, early on, was observed to have
unusual properties, such as a magnetic susceptibility that displays a maximum at
a temperature of 500 K. Jaccarino and coworkers concluded ~14 that the system has
a narrow semiconducting band gap with high d-band densities of states to either
side that could be involved in thermal excitations responsible for the susceptibility.
Since then FeSi has been the subject of many experiments. 2~2~6217218 From some
of the more recent photoemission work, it was argued that FeSi has a pseudogap,

2~3H. M. van Nort, D. B, de Mooij, and K. H. J. Buschow. J. Less Common Metals. 115. 155 (1986):
H. Onodera, H. Yamauchi, M. Yamada, H. Yamamoto, M. Sagawa, and S. Hirosawa, J. Mag. Mag.
Mater., 68, 15 (1987): and R. Fruchart, P. L'H6ritier, E Dalmas de R6otier, D. Fruchart, P. Wolfers,
J. M. D. Coey, L. E Ferreira, R. Guillen, P. Vulliet, and A. Yaouanc, J. Phys. F, 17, 483 (1987).
2~4 V. Jaccarino, G. K. Wertheim, J. H. Wernick, L. R. Walker. and S. Arajas, Phys. Rev., 160, 476 (1967).
2~ K. Tajima, Y. Endoh, J. E. Fischer, and G. Shirane, Phys. Re~: B, 38, 6954 (1988).
TRANSITION-METALS AND THEIR ALLOYS 101

k..,...... FeSi ".


%~ ,*" ..... ~1 *
9 :II" ......
"** 9 9149149 t

' 9149

~ wj~ 9149149149149149
9149 i
:8 1 ' ;''''~l"

_~ i .:..,;~11].
e'T

0o9 0~149 i

9 ...... 9..-i.
. ~ iJ
_~ . _ ~ .~149 ~176 I
X F M - R

FIG. 50. The calculated electron energy bands for FeSi along several high symmetry lines.

i.e., small hole pockets above and small electron pockets below the Fermi energy:
this behavior may arise from surface states in the gap. The general consensus is
that FeSi does have a gap which is less than 0.1 eV. which is small indeed.
FeSi forms in an eight-atom primitive cubic (cP8) structure of which it is the
prototype. Both atoms have an internal coordinate not determined by symmetry
that sets the positions of a given atom type in the cubic unit cell. Mattheiss and
219
Hamann calculated direct and indirect band gaps of 0.13 and 0.11 eV, respectively.
Subsequent calculations --~ are essentially in agreement. The FeSi bands along
high symmetry lines are plotted in Fig. 50. These are the results of unpublished
calculations of the present authors where the lattice constant and the internal
atomic coordinates were variationally determined. The plots give some suggestion
of the sizes of the gaps. However, the band extrema do not reside on such high
symmetry lines because mirror symmetries, which commonly occur at such lines,
do not exist in the space group T 4 appropriate to the FeSi structure. Scalar relativistic,
i.e., without spin-orbit coupling, estimates yield direct and indirect gaps of 0.15
and 0.12 eV, respectively, that become 0.12 and about 0.08 eV with the inclusion
of spin-orbit coupling. Larger band gaps are obtained for RuSi (0.25 and 0.16 eV)
and OsSi (0.36 and 0.26 eV) with spin-orbit coupling included.

2~6Z. Schlesinger, Z. Fisk, H.-T. Zhang, M. B. Maple, J. F. DiTusa, and G. Aeppli, Phys. Rev. Lett.,
71, 1748 (1993).
2~v K. Breuer, S. Messerli, D. Purdie, M. Gamier, M. Hengsberger, Y. Baer, and M. Mihalik, Phys.
Rev. B, 56, R7061 (1997).
2.s T. Susaki, T. Mizokawa, A. Fujimori, A. Ohno, T. Tonogai, and H. Takagi, Phys. Rev. B, 58, 1197
(1998).
219 L. F. Mattheiss and D. R. Hamann, Phys. Re~: B. 47, 13114 (1993 ~.
220 C. Fu, M. R C. M. Krijn, and S. Doniach, Phys. Rev. B, 49, R2219 {1994).
221 V. R. Galakhov, E. Z. Kurmaev, V. M. Cherashenkov, Y. M. Yarmoshenko, S. M. Shamin, A. V.
Postnikov, M. Neumann, Z. W. Lu, B. M. Klein, and Z.-R Shi, J. Phys. C, 7. 5529 {1995).
102 R.E. WATSON AND M. WEINERT

There are several other groups of transition-metal and main group alloys which
display somewhat wider band gaps, and two representatives have been reported
experimentally. Basov et al. observed 222 that RuAI2 is infrared active, indicating
the presence of a gap. There exist a number of calculations for RuA12 in the
22:~.224.225.226.227
observed Si,Ti structure - as well as in the other observed low-
temperature MoSi2 structure.-'-'B-~e4 (The 5d counterpart. OsAl,, also forms in the
9 ~'~6

MoSi2 structure.) With one exception-- that differs as to where the maximum in
the valence bands occurs, the calculations are in essential agreement" an indirect
bandgap of around a quarter of an eV and a direct gap in excess of 1 eV for the
Si2Ti structure. Assuming that direct transitions control the infrared activity, the
calculations appear to be in accord with the experimental results.

24. TRANSITION-METAL ALUMINIDES AND THE B~F~ STRUCTURE

Nishino and coworkers reported 22s semiconductor-like behavior in the temperature-


dependence of the conductivity of Fe2VAI, although they also observed photoe-
mission indicative of a metallic edge at the Fermi level. Fe2VA1 forms in the BiF3
structure, the most common ordered structure for ternary transition-metal alu-
minides. With composition M2NA, the A atom may be a main group element such
as A1, Si, Ge, Ga, or In, and M and N are transition-metals, with M usually from
the righthand half of the row and N from the left. The structure (Fig. 43) is a bcc
lattice with the M atoms at the body corners, and the N and A atoms alternating
in the body centers along the x, v, and - directions. This structure tends not to
form ~s~ if the elemental atomic volumes of the A and N atoms differ by more
than about 30%. With two transition-metals involved, substituting one transition-
metal for another from the same column of the periodic table can change the
separation of the centers of gravity, E,~,of the constituent d-bands, thus affecting
any gap at the Fermi level. Given the bcc lattice on which the structure lies,
interactions between second as well as first nearest neighbors are of importance.
Similar substitutions of metalloids or transition-metals of differing size will affect
the lattice constant and cause changes in these interactions. As will be noted, the
resulting changes in hybridization can and do affect whether a system has a gap

222 D. N. Basov, F. S. Pierce, P. Volkov, S. J. Poon, and T. Timusk. Phys. Rev. Lett., 73, 1865 (1994).
223 M. Springborg and R. Fischer, J. Phys. Condensed Matter, 10, 701 (1998).
224 M. Weinert and R. E. Watson, Phys. Re~: B, 58, 9732 (1998).
=s D. Nguyen Manh, G. Trambly de Laissardiere, J. P. Julien, D. Mayou, and F. Cyrot-Lackmann,
Solid State Comm., 82, 329 (1992).
226 S. E. Burkov and S. N. Rashkeev, Solid State Comm., 92, 525 (1994).
22v S. Ogut and K. M. Rabe, Phys. Rev. B, 54, R8297 (1996).
22Sy. Nishino, M. Kato, S. Asano, K. Soda, M. Hayasaki, and U. Mizutani, Phys. Rev. Lett., 79, 1909
(1997).
TRANSITION-METALS AND THEIR ALLOYS 103

Fe2NbAI ~~,~ i 1
>
oo 4

4--'
~ i

o
4-'
09
"6 8
C

I --1

-10 -5 0 5 V _~ X
Energy (eV)

FIG. 51. Calculated densities of states and band structure.,, along F to X (along which the valence
and conduction band extrema lie) for Fe,NbA1
_ and for Fe,VAI in the BiF..structure. The Fermi levels
define the zeroes of energy.

or a pseudogap. Thus, the relative sizes of the constituents affects whether a given
phase may occur and the hybridization important to gap formation.
The Fe2VA1 system, reported -~-'~ to be a semiconductor, has an even number of
valence electrons in the unit cell, consistent with having filled bands and the possi-
bility of a gap. The densities of states -~24for this compound and for Fe_~NbA1 are
shown in Fig. 51. The AI s-electron character for both compounds is concentrated
in the low-lying peak below the bulk of the bands. Although the AI s hybridization
may be significant to the energetics of the alloy formation, the AI site densities of
states, in the vicinities of the gaps are dominated by p and d character, with most
of the d-like density arising from tailing of transition-metal wave function character
onto the A1 site. The most important feature in the densities of states tbr the present
discussion is the deep hollow around the Fermi level.
The bands responsible for defining the (pseudo)gaps in the densities of states
near the Fermi levels (shown in Fig. 51) are along the V to X (=2n'(001)/a) line.
(Although the (001) and (010) directions are symmetry equivalent, the particular
choice for the X point has the advantage that the conventional choice of d,:_,_-
and d~:_~_,~orbitals are correctly symmetry-adapted.) In the case of Fe2NbA1, there
is a gap in the spectrum, with the minimum direct gap o f - 0 . 1 6 eV at the X point.
For Fe2VA1, slight changes in the relative positions of the bands cause the con-
duction band minimum at X to fall below the valence band maximum, which is
now at F. Thus, Fe2VA1 is a semimetal with small hole pockets around F and small
electron pockets at X, consistent with the metallic character of the photoemission
104 R.E. WATSON AND M. WEINERT

spectrum 228 and with the semiconductor-like conductivity because of the small
number of carriers. The bands for Fe2VA1. including its semimetallic nature, are
in agreement with other band calculations.- ....... The readjustment of the bands
at X and the shift relative to F on going from Fe:NbAI to Fe:VA1 is even greater
for their Ru counterparts, leading to larger electron and hole pockets. While
Fe2NbA1 and the Ru counterparts have not yet been reported to occur, these alloys
are predicted ~8~ to be stable relative to competing binary alloys.
Whether the gap intersects the Fermi level depends on the number of valence
electrons. Starting from RuzNbA1 and increasing the charge of Ru by 1 and decreas-
ing that of Nb by 2 (to maintain the total number of electrons) yields Rh2YA1.
This alloy is calculated to have a pseudogap. However, the e,~ of Rh and Y are
well separated and a situation similar to that seen in Fig. 22 arises, namely the
Fermi level, and with it the pseudogap, fall in a low density of states region
between the high d-band peaks. (The relative sizes of Y and A1 make this phase
unlikely to occur, but Rh2ScA1 or Co2ScA1 might.) One can also interchange the
charge of the metalloid with that of the N (or M) atom transition-metal(s) as well,
obtaining, for example. FeeTiSi and Ru2TiSi. The former has calculated direct
and indirect band gaps of 0.35 and 0.31 eV. while for Ru:TiSi they are of the
order of 0.01 eV. Such small values make the estimates a close call, but if Ru~TiSi
can be fabricated, this phase might have interesting properties.
The bands along F-X for an M2NAI alloy in the BiF3 structure are sketched
in the righthand part of Fig. 52. While the bands represent a case like Fe2NbA1
with a gap, the important band structure features of all of these alloys are similar.
The wave function character of the bands at X near the Fermi level are also labeled
in the figure. Naively, the t~ -e~ convention might seem a natural way to discuss
-,r ,

the bonding in this cubic system, but such a division is inadequate for under-
standing the formation of gaps in the BiF3 structure. The individual members of
the t2~ o r eg manifolds belong to different irreducible representations along A; it
is these irreducible representations that determine the allowed hybridizations and
interactions, and hence the (pseudo) gaps. (If these alloys were highly localized
systems with little hybridization among the sites a description in terms of t~ -e
or t2-e states, depending on the site. might be appropriate. This limit, however,
is not the correct one for these alloys.)
Just below the Fermi level for M2NA1 in Fig. 52 are a doubly degenerate (dashed
line) band and a singly degenerate (dotted) band. At X. the singly degenerate
band is a mixture of M (Fe) d,~, and p: states, while the doubly degenerate state
is a mixture of M (d,:, d,:, p,, p,), N (p, p,), and A1 (p,, p,) orbitals. The flat band
above the Fermi level is composed of M (Fe) d,:-,-" orbitals, and the band marked
with the heavy solid line is composed entirely of N (Nb) d~:-,: orbitals at X

229 D. J. Singh and I. I. Mazin, Phys. Rev. B, 57, 10598 ~1998).


230 R. Weht and W, E. Pickett, Phys. Re~: B, 58, 3855 11998).
TRANSITION-METALS AND THEIR ALLOYS 105

M2N 2 \ ~ ~ o M2NAI

........... "
..... \\" \k\j,. //-.i '~-i - -~'
1

>
- ~"' N x2-y 2
................................. M: x 2 - y ~
" / ............ " ----- / M, N. AI
r T /
I ..~. ...... . / ..... M: xy, z
LIJ
r/" . .... '1 % ......
-1

-2
~ . ---'- " " I
, . . . . . . .

X F A X

FIG. 52. Schematic of the bands along F-X of the BiF, structure for the transition-metal aluminide
M2NA1 and for M2N 2, with parameters appropriate for M = Fe and N= Nb. Dashed (dotted) lines
represent doubly (singly) degenerate bands, and the heavy solid line represents the N d~:,: band (at X)
discussed in the text. Note that M:N: is actually in the CsCI structure bands originating from the
R-M line of the CsC1 structure are marked by circles. The bands for M,N, have been shifted to align
the energy of the flat band at F with the similar band of M:NAI.

because, by symmetry, there is no coupling between the d,:-,: orbitals on the M


and N sublattices. From inspection of a large number of band structures for different
alloys in the BiF3 structure, the relative position of this N d,-_,: state at X is found
to be the major factor in determining whether or not a gap forms.
Around the Fermi level there is very little A1 character. Because of this depletion
of AI character, it is not obvious, at first glance, that AI plays any significant role
in the formation of the gaps. However, calculations which simply ignore the A1
atoms (M:N in the CaF2 structure) show that the gaps are destroyed (as well as
shifting the Fermi level), demonstrating that the A1 atoms do play an important,
if indirect, role. We will return to this issue shortly.
The origin of the (pseudo)gaps in the density of states for the BiF3 structure
can be understood using simple tight-binding arguments. In this structure, there
are four atoms per unit cell and three symmetry inequivalent sites. As discussed
above, states with d,-'-,-" character around the Fermi level play an important role
in determining the existence of the (pseudo)gaps. Along F - X , all d,-'-,: orbitals
belong to either the A: or A~ irreducible representations" conversely, no other s,
p, or d orbitals on any of the atoms belong to these irreducible representations.
For a general M2NA alloy, the symmetry at X dictates that the d,'--,: orbitals on
the M sublattice do not interact with those on the N-A sublattice. Thus, the four
bands at X derived from the d,-'-,." orbitals split into two sets of bonding and
antibonding levels. The tight-binding energies 23~ of the states derived from the
106 R . E . WATSON AND M. W E I N E R T

orbitals centered on the M sublattice are

.~l . .~/.~I
EI t ( X ) - e,l + 3 k',t ac~, (43)

3,1 . .~IM
E~ s ( X ) - e,t - 3 V , u ~, (44)

and for orbitals on the N-A sublattice, the energies are

E~'-;
-
.V

(X) - e~t +~ ___e,t +


A

--
A

,.)
.\"

+ (3 gja,~) " "I1/2 (45)

(z
In these equations, the usual notation is used: Ej is the tight-binding on-site
c4J
energy parameter for the d electrons of atom o~, and V,~ao < 0 is a second neighbor
interatomic potential matrix element between atoms o~ and/3. (The small Vjja terms
have been neglected.) The state corresponding to E I t ( X ) is below the bottom of
Fig. 52, the energy of the flat band just above the Fermi level corresponds to E~'(X),
and the state labeled "' N:x -~- x': "" corresponds to one of the pair of E~' ~a(X).
Note, the energies given in Eqs. (43-45) do not correspond 2s: to the energy of
the zone center e~ level in the bcc lattice of the elemental system.
For the case that the A atom is AI. the intrinsic A1 d levels are much higher
in energy than for the transition-metal, and the interaction between the d states

z31 j. C. Slater and G. F. Koster. Phv.s. Rev.. 94. 1498 (1954).


_~3_~From a cursory glance at the band structure of M.NAI or M.N.. one might inter that the flat band.
which is doubly degenerate at F. should be associated v<ith the energy of the zone center e,, level of
the elemental system. This expectation, how'ever, is wrong. To understand why. let all the atoms in
the BiF; structure be of type M. The system is then simply a supercell of the bcc structure, with a
one-to-one correspondence between bands in the BiF: and bcc structures. Although the d,:,- orbitals
nominally belong to the e,, manifold, the energies given in Eqs. (43-45). including E]~(X)._ differ from
the zone center bcc e~, e n e r g y

~t 16/r V ~t w
E , , ( F ) = E~ + .--7 ,j_+3~'L<,<,~.

" I ~t ' , I
In this equation, the tight-binding parameter ~. ,i,i,~ is again a second-neighbor interaction and V,t,~,~ is
a nearest-neighbor interaction. The simple explanation for the differences is that the flat band and
the one that is degenerate with it at 1- are bands from the outer part of the bcc Brillouin zone" the
differences in energy are related to the dispersion of the bcc bands. As an aside, there are no Vj,~,~
contributions to the band energies in Eqs. (43-45) because the phase factors associated with V,1,t,~
vanish at X: elsewhere along A these terms do contribute to the energies.
TRANSITION-METALS AND THEIR ALLOYS 107

on the N and A1 sites is vanishingly small. The effective lack of d orbitals on AI


. ,v-.v .~t .v I
can be expressed as [v,~,~ /(e,~ - e j ) --->O. In this case the two eigenvalues at
Al
X of Eq. (45) are simply a very high-lying state at E,I (not shown) and the one,
.v
marked with a heavy line in Fig. 52, at e a . Hence, the position of the N d,-'-,~-
state at X for M e N AI is determined not by any transition-metal d and AI p
hybridization, but rather by the lack of N-A1 d-d hybridization. This result,
unexpected perhaps, is essential in order to understand the formation of the gap
in these systems.
Clearly transition-metal-A1 hybridization must play a role in the formation of
the gap. To better understand the effect of AI on the formation of the (pseudo)gaps
in M 2 N A1 alloys, it is instructive to consider the case A - N, i.e., the transition-
metal binary M N (or equivalently, M~Ne) alloy in the CsC1 structure. These bands
are shown in the lefthand part of Fig. 52 for the A line of the BiF3 structure. Since
M2N2 in the BiF3 structure is a supercell of MN, half of the bands in Fig. 52 along
F-X of the BiF3 cell are from the F-X line in the CsCI simple cubic Brillouin
zone and the other half come from the R(= 2 Jr(~~~ ~)/a) to M(=2n'(~~O)/a) line;
these latter states are marked with circles in the figure. States aris-ing from the
different lines in the CsC1 Brillouin zone do not interact.
MeN2 does not have a gap because there are two bands with large dispersion
passing through the gap region of M e N A I . These bands are a doubly degenerate
band (dashed line) coming from the R-M line dispersing upward, and a downward
dispersing band (heavy solid line). As before, the band marked with the heavy
solid line has N d,:-,: character at X, and is the counterpart of the "N:xe-v 2'' band
of M2NA1 shown in the righthand part of the figure. Because there are now d,~'-,:
orbitals on both N sites, the energy of this band at X is given by (cf. Eq. (45))
N . NN NN
Ej + 3 V aj ~, which is 13V,t~t~1lower than the corresponding band of MeNAI. This
large reduction in the dispersion for the aluminide arising from the missing transition-
metal-A1 d-d hybridization is a necessary condition if a gap is to form.
The upwardly dispersing doubly degenerate band (dashed line with circles)
that destroys the pseudogap in MeN2 arises from the R-M line of the CsC1
Brillouin zone and crosses doubly degenerate bands originating from F-X of the
CsC1 zone. When one of the N atoms is replaced by A1. the artificial translational
symmetry of the supercell is broken. Now all the doubly degenerate levels belong
to the same irreducible representation and can hybridize, causing the band crossings
to change into anticrossings. The accompanying level repulsions between the
doubly degenerate bands determine the relative energy of the valence band max-
ima at F' and X. The positions of these doubly degenerate states are dependent
on the A1 p and transition-metal d hybridization since the A1 p,, and transition-
metal d:.,.:: orbitals all belong to this irreducible representation. For the large
potential difference corresponding to removing the A1 atom entirely, i.e., the CaF2
structure, the hybridization among the bands and the level repulsions are so large
that the pseudogap structure is destroyed.
108 R . E . W A T S O N A N D M. W E I N E R T

This analysis confirms that the relative positions of these different bands at X
are the best indicators of whether a gap is formed in the MeNAI alloys. In the
case of Fe2VA1, the Vd,e_,~_ band dips far enough below the F e d : _ , : to result
in semimetallic behavior. For the Ru compounds, the Nd,_,~ band at X is even
farther below EF. From the tight-binding results, we can make some further
observations and predictions. To get a gap, the energy of the Nd,:_,e band (Ex(X))
should be as high as possible relative to the energy (E]t)_ of the flat M d,__,2
__ N .~I . M.~! .
band" E N E ~ = e a - ( e,t - 3 v,tao~ should be as positive as possible, recalling
MM
that Vaao < 0. This requirement can be satisfied by separating the d-bands farther
and/or decreasing the interaction between the M atoms. This latter condition
suggests that larger N atoms (Nb vs. V), which will expand the lattice and reduce
the M - M interactions, are more likely to form gaps. This prediction is borne out
by the calculations for both M = Fe and Ru. Moreover, if the lattice is expanded
by about 6%, FezVAI becomes an indirect gap semiconductor. On the other hand,
MM
the Vaa o terms for the Ru compounds are significantly larger (-60%) than those
of the Fe alloys" the magnitude of these terms is the major reason that gaps do
not form in either Ru~VAI or Ru~NbA1.
The other term, the relative position of the transition-metal d levels, depends
on a number of parameters, including the specific elements and the volume. From
the calculated band structures, the Nb d levels are found to be higher than the V
d levels in both the Fe and Ru alloys. Since the d-bands in the 4d row are lower
than those in the 5d row, the analysis predicts that replacing Nb with Ta should
Ta Fe, Ru
increase the tendency to form a gap by increasing e,t - E,t As a test of this
Ta Ru
hypothesis, calculations were done for Ru2TaA1 and it was found that E,~ - E a
Na Ru l/Ru-Ru
is about 0.35 eV larger than E,~ - Ea , while -Ja~, is approximately the same.
The net result is that the pseudogap in the density of states at E F for RueTaA1 is
significantly deeper than for Ru2NbA1. The detailed reason for this change is that
the N d,: ,.2 state for Ru2TaA1 is above the doubly degenerate state at X, whereas
in Ru2NbA1 it is below. RueTaA1 is still a semimetal, with the N d _,_,~ state at
X -0.28 eV below the states at F. The density of states and band structure around
the Fermi level, however, are rather similar to those of Fe:VAI, suggesting that
Ru2TaA1 might have properties very similar to those of Fe:VA1.
From the results of many calculations, it would appear that hybridization band
gaps and semimetallic pseudogaps around the Fermi level may not be all that
uncommon, though few gaps, if any, may be as small as that of FeSi. Factors
encountered earlier in this article, such as the relative positions of d-band centers
of gravity and the relative sizes of constituent atoms, have significant impact on
whether a gap or a pseudogap forms. Finally, there is some suggestion from FeSi
that, contrary to normal experience, the local density approximation tends to
overestimate the band gaps, although the errors appear modest. This difference
in behavior reflects the difference in the nature of the gap compared to, for
example, the more usual s-p gaps in semiconductors: For these systems, the wave
TRANSITION-METALS AND THEIR ALLOYS 109

function character of the states on both sides of the gap are the same (d-like).
For this class of system, at least, it does not appear essential to introduce additional
correlation effects in order to account for the occurrence of band gaps.

IX. Epilogue

This article has been an eclectic review of the transition-metals and their alloys.
One underlying goal has been to make connections among parameters of long
standing--e.g., atomic size, valence electron count, electronegativities, and
charge transfer--and the results of modern calculations. Although such quantities
are often difficult to define rigorously and their physical basis may be somewhat
suspect, they provide an intuitively simple and appealing approach for discussing
bonding of alloys, and are firmly entrenched in the language of the field. Some
of the topics covered strayed farther afield, but with the hope that important
lessons could be learned about transition-metals: one such lesson is that many
alloys, contrary to the impression one might get from many textbooks, form in
other than the fcc, bcc, and hcp structures. Often the general issues raised, more
than the particular results, determined whether a topic was discussed.
We have considered Wigner and Seitz's "'giant computing" machine throughout
this article. Although we have yet to attain their machine, modern calculations
do yield insights into the complexities of transition-metal alloy bonding which
would be difficult or impossible to obtain from experiment alone. When done
with care, such calculations can yield estimates of thermodynamic quantities to
an accuracy comparable to experiment, making semiquantitative constructs of
phase diagrams possible.
Transition-metals and their alloys show a rich variety of properties, some of
which make them of technological interest. Beyond the practical interest, these
materials continue to be the focus of basic research. Some areas of current research
were discussed here, while others such as phase transitions, correlation effects,
and transition-metal-oxides were only touched upon in this article. Although it is
difficult to predict what will be the future "'hot" topics in condensed matter physics,
it is a safe bet to believe that many of them will involve the transition-metals, and
that understanding them will require a combination of experiments, simple models,
and detailed calculations.

Acknowledgments

This article is dedicated to the memory of Morris Perlman, a marvelous scientist


and collaborator. We also wish to thank Larry Bennett, Gayanath Fernando, and
Gtinter Schneider for valuable discussion and suggestions. The work at Brookhaven
110 R . E . WATSON AND M. W E I N E R T

National Laboratory was supported by the Division of Materials Sciences, U.S.


Department of Energy, under Contract No. DE-AC02-98CH 10886.

Appendix
Population Analyses: Standard and Site Specific

If a wave function gt is expanded in terms of a basis set. 0;, i = 1.... N, by

Ill" - E ci~i" (46)


1

the corresponding density is

Ili llt - ~ c; O; Oj c j . (47)


i.J

Integrating over all space, one obtains the total charge, Q, contained in this wave
function

Q = y_c]Sijc j. (48)
i.j

where Sij is the overlap integral between the ith and the jth basis functions.
Making use of the reality of Q. the Mulliken population ~ Q, associated with the
ith basis function may be derived:

Q= E R e [ ~ _ c [ S i j c j l - E Q i , (49)
i j i

Qi- ERecTSijcj 9 (50)


J

This procedure divides the overlap charge equally between basis functions i
andj. L6wdin proposed ~4~dealing with overlap by transforming over to an orthogonal
basis set. A modified Mulliken scheme ~~9may be derived where the overlap charge
is divided in proportion to the diagonal contributions"

2s;;Ic;I -~
O- R e c ; S i j c j = - Z Q i, (51)
i

0; - jY~s;;I C,- ~- -"Jr-


- -S j j C) -
~Re ~;S,jcj (52)

Summing over all the occupied wave functions yields the total populations.
-. ~ . ~ ~ ~
- ~ ~ ~ ~ ~-~ hr~
9 ,.-- "~_, ,_.
"~T_. r
0 I:::: ""
(1.,) .,--, c~
9~ ~ ..a ~ ~ ~
>.
.._~ ~'~
. ,-..,
--, IZ.
9 ~ ~ "~_~ *-~ ~
>., =,~
.,i
9
.,-.,
<
- - ~ o
.-- ~ 9
.,--, 41. .--' ,~
..~ >.~ ~ I
"- = ~'~ .>_
.-<
"~-.~ -~t~
,.~ ~:~ ~ ,.~= ~ . ~ ",~. 9 "i. -.4 ,.I:=
.'~ II
"a ~ II
c~
II II "-L ~ = o
II
~- .o ~ . = ~ _ ~ " 0
<
a'~b .m ~ ~ ~-~
c~
E~
.,~
9"~ ,.~ ~ "T- " ~ ~.,..~ ~ --
~ ~ ~ ..-- ~ ~ Z
9~ ~ ~-~
112 R.E. WATSON AND M. WEINERT

(A similar expression holds for ~ . ) The on- and off-site (tail) contributions are
then

f~_g :~ + f tg)."
,/~an - aL., . (59)
W).

and

p .f

/-~tail _ F A g' ), -- F.xg).


I"A W~. " (60)

and similarly for ~


With this division, the wave functions within the spheres are separated into
on-site, A = L only, and tail contributions in the manner similar to Eq. (54), except
the basis is allowed to correctly respond to the actual potential at the site in the
solid, including accounting for changes in s, p, and d counts. In this scheme, all
angular momenta A , L are attributed to the tails, consistent with chemical
intuition. Squaring the wave function (for simplicity the o~.~terms are omitted,
though they are included in any calculation) yields on-site, tailing, and overlap
contributions to the charge at the site, namely

on tail over
q~. = q,~ +q,~ +q,~ . (61)

q~n_ Zl z t e, (62)
.u

_ [~tailXpAO
on ,
qx~ ~Z~2 Re(,.,, ) (63)
,u

q).~.il - y~lff( " ''l _, {64)


/a

This last expression was used in the estimates appearing in Fig. 37, and such
tail terms plus overlap terms apportioned using the modified Mulliken scheme,
were subtracted off in Fig. 36.
The traditional and the site-specific population analyses offer different per-
spectives of the bonding behavior. Both depend on how the overlap charge is
apportioned and on the choice of basis set. However, it is the authors' experience
that the site-specific results are rather insensitive to these choices, and signifi-
cantly less sensitive than the traditional population analyses.

Das könnte Ihnen auch gefallen