You are on page 1of 16

This article was downloaded by: [Michigan State University]

On: 28 February 2015, At: 01:46


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954
Registered office: Mortimer House, 37-41 Mortimer Street, London W1T 3JH, UK

Molecular Physics: An International


Journal at the Interface Between
Chemistry and Physics
Publication details, including instructions for authors and
subscription information:
http://www.tandfonline.com/loi/tmph20

Potential energy functions for


atomic solids
a a
John N. Murrell & Rachel E. Mottram
a
School of Chemistry and Molecular Science University of
Sussex, Falmer , Brighton, Sussex, NB1 9QJ, England
Published online: 23 Aug 2006.

To cite this article: John N. Murrell & Rachel E. Mottram (1990) Potential energy functions for
atomic solids, Molecular Physics: An International Journal at the Interface Between Chemistry
and Physics, 69:3, 571-585, DOI: 10.1080/00268979000100411

To link to this article: http://dx.doi.org/10.1080/00268979000100411

PLEASE SCROLL DOWN FOR ARTICLE

Taylor & Francis makes every effort to ensure the accuracy of all the information
(the “Content”) contained in the publications on our platform. However, Taylor
& Francis, our agents, and our licensors make no representations or warranties
whatsoever as to the accuracy, completeness, or suitability for any purpose of the
Content. Any opinions and views expressed in this publication are the opinions and
views of the authors, and are not the views of or endorsed by Taylor & Francis. The
accuracy of the Content should not be relied upon and should be independently
verified with primary sources of information. Taylor and Francis shall not be liable
for any losses, actions, claims, proceedings, demands, costs, expenses, damages,
and other liabilities whatsoever or howsoever caused arising directly or indirectly in
connection with, in relation to or arising out of the use of the Content.

This article may be used for research, teaching, and private study purposes. Any
substantial or systematic reproduction, redistribution, reselling, loan, sub-licensing,
systematic supply, or distribution in any form to anyone is expressly forbidden.
Terms & Conditions of access and use can be found at http://www.tandfonline.com/
page/terms-and-conditions
MOLECULAR PHYSICS,1990, VOL. 69, NO. 3, 571-585

Potential energy functions for atomic solids


by J O H N N. M U R R E L L and R A C H E L E. M O T T R A M
School of Chemistry and Molecular Science
University of Sussex, Falmer, Brighton, Sussex NB1 9Q J, England

(Received 31 Aufust 1989; accepted 7 November 1989)

A potential energy function has been proposed for atomic solids consisting
Downloaded by [Michigan State University] at 01:46 28 February 2015

of two-body and three-body terms which over its range of parameters encom-
passes all simple crystal structures. Phase diagrams have been constructed as a
function of a single parameter in the two-body term and one, of several, param-
eters in the three-body term, and these include as the most stable structures
hcp, fcc, bec, sc and diamond. The hcp and fcc lattices have been shown to
distort to lower symmetries with appropriate three-body parameters and these
distortions are in accord with known structures. A brief study has been made of
the appropriate parameters for silicon.

1. Introduction
Atomic solids can have a wide variety of crystal structures [1]. Typical examples
are the inert gas solids, which are face centred cubic (fcc), silicon, which has a
diamond structure, the alkali metals which have a body centred cubic (bcc) struc-
ture, and magnesium which has a hexagonal close packed (hcp) structure. Lower
symmetry structures are also found. For example, zinc has a distorted hcp structure.
High pressure or low temperature polymorphs are common. Lithium, for example,
converts to hcp below 78 K.
The question we wish to answer in this paper is whether there is a simple
potential function which, with appropriate choice of parameters, can reproduce the
structures and stabilities, ideally even the elastic constants of all such solids. This is
like asking whether there is a common functional form for the potentials of all
diatomic molecules. The answer to this is yes, within limits. For example, a Morse
function is broadly acceptable, but if greater accuracy is required for vibrational
spectroscopy, then more complicated functions are needed.
If a potential is found which reproduces the solid structures we can then
examine its more general validity. Will it also reproduce the properties of the liquid
state? Will it give the structures of defects and of the solid surface? Will it reproduce
the properties of microclusters? This is an ambitious programme that may not
succeed but there are a few indications that the task is not impossible. There is, for
example, a simple potential for silicon [2] that reproduces the main structural
features of the solid and liquid states, although it is reported that it gives a poor
account of the structures and stabilities of silicon clusters [3]. There is also a
potential deduced for lithium clusters that predicts correctly the bcc structure for
the solid [4], but it does less well for the liquid state [5] and is poor for the elastic
constants of the solid. Finally, we note that for the inert gases, at least, the task is
not too difficult because their potentials are dominated by pair interactions [6].
In both the silicon and lithium work referred to above, the potentials have been
constructed as a sum of two-body and three-body terms. This is notwithstanding the
0026-8976/90 $3.00 9 1990 Taylor & Francis Ltd
572 J . N . Murrell and R. E. Mottram

fact that for small dusters of these elements the higher order terms are certainly not
negligible [7, 8]. Indeed for lithium clusters there is clear evidence from calculations
that the terms in the many body expansion do not converge at low orders. However,
this does not mean that empirically chosen two-body plus three-body terms will not
work well; at least for the structurally most important parts of the potential energy
surface. If we were even to show that a two-body plus three-body potential could
not explain some structures, that would be an interesting conclusion.
Our general strategy therefore will be to seek a potential which is a sum of
two-body and three-body terms
v = Y Y v 2, + Ei j E L
i j>i >i k>j
Downloaded by [Michigan State University] at 01:46 28 February 2015

: z
i 9
z
j•i" "
--U + g j # i k ~zi , j Ir ~jkJ~ (1)

If all atoms in the solid are equivalent, the atomisation energy is the term in
brackets in this expression. If we divide up the solid into shells (p) of equivalent
atoms at distance rv neighbouring any one atom (1), which contain nv atoms, then
the atornisation energy is
1 1
Va = ~ X n, V]]'(r,) + -~ 2 nv 2 X 6o -xovv,v'a)t" r,,, r,j), (2)
p p p' j(p'~

where i is any one atom in p, j ranges over the atoms in p', and 6 u is zero if i = j but
unity otherwise. Note that the contribution from different shells (p' r p) occurs twice
in the second summation in (2).
For a monatomic solid the potential must be symmetric to exchange of atoms
(even if not all lattice sites are equivalent) and hence the three-body term must also
be symmetric to exchange. This means that it can be constructed from the integrity
basis for three variables [9]. This is formed by first making the transformation

= ,,/1/2 -j1/2/[o2/, (3)


Qa Lx/2/3 -~/1/6 -x/1/6j[_p3_]
where
Pa -- r i i - r~ (4)
r ~ being a reference length for the interatomic separation. We note that the func-
tions Qi are equivalent to the vibrational normal modes in internal coordinates for
an equilateral triangle of sides r ~ . The integrity basis then consists of the three terms
Q1, Q~ + Q~, Q~ -- 3QaQ2 (5)
and a totally symmetric three-body term is
Vta)(Qa, Q~ + Q~, Q~ - 3Q3 Qz2). (6)
We later make a more detailed examination of this.
In this paper we examine only solid structures with one lattice parameter; in
particular cubic diamond, sc, bec, hcp and fcc. Table 1 gives the number of neigh-
bours to any atom in these structures at distances up to and including atoms at two
lattice spacings, which is the distance at which we terminated our lattice
P o t e n t i a l e n e r g y f u n c t i o n s f o r a t o m i c solids 573

Table 1. Numbers of atoms in shells at distances from any atom up to two lattice spacings.
D = diamond, S = sc, B = bcc, F --- fcc, H = hop.
Lattice
distance D S B F H
1 4 6 8 12 12
2/~/3 -- -- 6 -- --
x/2 -- 12 -- 6 6
x/(8/3) 12 -- 12 -- 2
x/(ll/3) 12 -- 24 -- 12
~/3 -- 8 -- 24 18
2 -- 6 8 12 6
Downloaded by [Michigan State University] at 01:46 28 February 2015

s u m m a t i o n s t . F o r slowly decaying potentials this termination will of course lead to


errors in the total potential, but as we only treat exponentially decaying or r -6
decaying functions, this should not be serious.
T o go further and use the potentials either to fit exactly elastic properties of a
crystal or in c o m p u t e r simulations of properties of bulk phases, defects, or surfaces,
the question of a cut-off will have to be addressed m o r e fully. The ratio of three-
b o d y to t w o - b o d y terms increases as the cut-off distance of the potential is
increased. With the exponentially decaying functions p r o p o s e d in this paper there
should be no difficulty in obtaining convergence. Indeed for the w o r k presented here
we have s h o w n that there is no change in the basic physics by extending the cut-off
range; o u r conclusion that the potentials we use can reproduce all the simple crystal
forms is not dependent on the cut-off distance, and the qualitative form of the phase
diagrams shown in later figures does not appear to change.
Finally, we acknowledge that we are by no means the first to use three-body
potentials in studies of crystals and we refer in this paper to some specific examples.
A recent paper dealing with the ab initio calculation of m a n y b o d y potentials for
transition metals [10] makes reference to several earlier studies of the subject.

2. Two-body stabilities
In order to reduce our parameters to a bare minimum, we first examine the
t w o - b o d y stabilities of the lattices referred to above using a simple R y d b e r g pair
potential
V = O(1 + b(r - re) ) exp ( - b ( r - re)). (7)
Expressing the potential in units of the dissociation energy of the diatomic, D, and
distances in units of the equilibrium distance r e, this becomes a one-parameter
function
V / D = --(1 + a p ) exp ( - - a p ) , (8)

p = (r -- re)/r e , a = br e . (9)

t For brevity within the text we use the symbols D, S, B, H and F defined in this table.
574 J . N . Murrell and R. E. M o t t r a m

20-

15

10

t> o

-5 ~
Downloaded by [Michigan State University] at 01:46 28 February 2015

-10 -

--15

-20 '''"'"......."'"'"'" , '' ~ "'~


0.5 1.0 1.5 2.0
R
Figure 1. The total potential energy of the crystal, V, as a function of the interatomic
spacings in units of r e for the Rydberg two-body potential, a = 3-0. D ( - - - - - - ) ,
S ( - - - - ) , B ( .... ), F ( ). The H potential is very close to the F and has not been
included.

The R y d b e r g function has been shown to be a satisfactory function for most


diatomics and when the polynomial is extended up to cubic terms in 9 is in most
cases very accurate except in the long range limit [11].
F o r diatomic molecules the parameter a in (8) can be determined from the
h a r m o n i c force constant. Selected values are Li 2 (3.26), Si2 (4.21), a n d Mg2 (5-74),
and apart from the inert gases, for which (7) is n o t a g o o d potential, m o s t elements
which form atomic solids have values of a close to 4.0.
Figures 1 and 2 show the s u m m e d t w o - b o d y energies (from (2)) for the five
lattices under consideration as a function of the lattice spacing for a --- 3.0 and
a = 7.0. Table 2 gives values of the o p t i m u m interatomic distance (units of re) and
atomization energy (units of D) for this range of a. F o r a < 5.6, B is the m o s t stable,
and for a > 5-6, H is the m o s t stable a l t h o u g h the difference between H and F is

Table 2. Optimum lattice spacings and energies for different values of a.

- - Vml./D Rmin/re
a D S B F H D S B F H
3 10"62 12"99 20-25 18-89 19.57 0'66 0.73 0-68 0"72 0.71
4 7.64 9.88 14-42 13.75 14.09 0.71 0.78 0.76 0-80 0-79
5 4.82 7"67 10.33 10.14 10.25 0-80 0-84 0.85 0"88 0.88
6 3-24 6-03 8.24 8-25 8"28 0.89 0.89 0.91 0-94 0.94
7 2.59 4.96 7.19 7.30 7.31 0.95 0.93 0.94 0.97 0.97
2-00 3-00 4-00 6-00 6-00 I'00 1-00 1.00 1-00 1.00
Potential energy functions for atomic solids 575

20-

15-

10-

5-

~>
Downloaded by [Michigan State University] at 01:46 28 February 2015

0-

-5-

-10 -

-15

-2O
05 11o 115 210
R
Figure 2. Potentials for the solids for the Rydberg two-body potential, a = 7"0. Symbols as
in figure 1.

small and approaches zero as a approaches infinity. Elements for which the two-
body potential is dominant can therefore be expected to have B, F or H structures
and this appears to be the ease for most metals. The best value of a for D is 3-7 and
for S is 5-2; these giving optimum stabilities relative to B. The termination of the
lattice summations at twice the nearest neighbour separation will give greatest
errors for small a. F o r the D lattice at a = 3, for example, third neighbours produce
about 83 per cent of the binding energy from second neighbours.
An alternative two-body potential is one whose long range form is r -6. Two
functions which, in reduced variables, have just one adjustable parameter are the
(EXP-6)

V = (6/(Z - 6)) exp (Z(1 - p)) - (Z/(Z - 6))p-6 (io)

and the Lennard-Jones (Z - 6)

v = (6/(Z - 6~p -z - (Z/(Z - 6~p -6. (11)

For both of these Z determines the steepness of the repulsive wall but the range of
the attractive branch is independent of Z. It is therefore not surprising that both
potentials produce the same order of stability of the solids H > F > B > S > D, and
this order is independent of Z. Low values of Z in both cases are o p t i m u m for
structures B, S and D.
576 J . N . Murrell and R. E. Mottram

3. The three-body terms


A three-body function which has been used [12] for fitting the potential surfaces
of homonuclear triatomic molecules such as 0 3 is the product of a polynomial
in the integrity basis (5) and an exponentially decaying function of Q1, either
exp (-/~Q1) or (1 - tanh ~Q:). It would be unphysical for the three-body term to
decay more slowly than the two-body terms at large distances so the value of/~
needs to be related to a when the two-body term is (8). Several options have been
explored and the most satisfactory appears to be to choose a simple exponential
decay with/~ = a. The three-body term is therefore written
V r = (Co + C,Q1 + C2Q~ + C3(Q~ + Q])
+ C4Q ~ + CsQI(Q 2 + Q~) + C6(Q] - 3Q3 Q22)) exp ( - a Q , ) . (12)
Downloaded by [Michigan State University] at 01:46 28 February 2015

Higher polynomial terms could of course be considered but taking terms up to


cubic brings in both Q22 + Q32 and Q33 - 3Q3 Q22 which are the important ' s h a p e '
functions; they distinguish between triangles having the same perimeter (Q1) but
different angles. Contours of constant Q22 + Q] are circles in (Q2, Q3) space. For
isosceles triangles Q~ + Q] is positive for apex angles both smaller and greater than
60 ~ Q] - 3Q3 Q2, however, is positive for isosceles triangles with apex angle greater
than 60 ~ and negative for angles less than 60 ~.
The three-body terms were calculated for the shells listed in table 1 but were
restricted to triangles where the third side (r o in (2)) is less than or equal to 2. The
total numbers of non-zero terms were six times 56, 71, 252, 231 and 210 for D, S, B,
/4 and F respectively. The stabilities of the structures were first examined for
separate variation of each of the seven coefficients Ci in (12) for a range of a from 3
to 10. Figure 3 shows the most stable phase for variation of Co and a. For negative

r D

r o S
3

H B

B
3 4 5 6 8 9 10
a

Figure 3. Phase diagram using the Rydberg two-body potential for the variation of a and
C O(expressions (8) and (12)). In this and subsequent figures the areas showing the most
stable phases have been indicated as follows: D = diamond, S = sc, B = bcc, H = hop,
F = fcc.
Potential energy functions for atomic solids 577

L
J
2.5

i
2.0
P
B
1.5

o
I
1.0

F
Downloaded by [Michigan State University] at 01:46 28 February 2015

0.5

f i t
00l 3 4 5 6
a
7 8 9 i0

Figure 4. Phase diagram for variation of a and C 1 (other features as in figure 3).

values of Co the lattices contract to unphysical distances; this is a reflection of the


fact that the two-body terms are rather soft in the repulsive wall. For positive values
of C O all the lattices expand, due to the repulsive nature of all three-body terms.
Not surprisingly, as C o is increased one first moves into a region where S is the most
stable phase and then, at larger Co, D is the most stable. In these regions the lattice
distances are all typically 1-5 times larger than the two-body equilibrium distance
(re) and the binding energy is similar to, or less than, the two-body binding energy
(O).
At no point in this phase diagram is F the most stable phase although its energy
is never far away from that of the H structure.
The B structure which is the most stable phase at low a with just two-body
terms is rapidly replaced by H on increasing Co. However, it can be seen that there
is a second region where B is stable at high a with a binding energy about 2.5 D,
and a lattice spacing just slightly larger than r e . In this region B, H and F have very
similar energies.
For positive values of C~ the lattices contract to unphysical values. Figure 4
shows that for negative values there are large areas of stability for structures B and
F and a very small area for stable H. In all regions the lattice spacings are of the
order of r e and the binding energies several times larger than D.
Negative values of C 2 contract the lattices to unphysical values. Figure 5 shows
that for low values of a and positive values of C2 there are areas of stability for
structures S and D. However, for these parameters the lattices B, H and F have
expanded substantially, the potentials even becoming repulsive. The S lattice is
strongly contracted in this region but the D lattice is, like B, H and F, expanded.
F o r large negative values of Ca the lattices contract to unphysical distances but
for small negative values there are regions where either B or F are the most stable
phases; again B, H and F all have similar energies. It can be seen in figure 6 that
there is a small region at low a where S is the most stable phase. It is possible that
578 J . N . MurreU and R. E. Mottram

D
6

4
r l F

3
Downloaded by [Michigan State University] at 01:46 28 February 2015

3.0 3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0


a

Figure 5. Phase diagram for variation of a and C 2 (other features as in figure 3).

this is an artefact of our termination of two-body and three-body terms at twice the
lattice spacing, which is a poor approximation at low a.
The parameters (?4 and C5 show no special behaviour, paralleUing C1 and C3
respectively. For this reason there does not seem to br any case for including them
in the three-body term unless the intention is to make an accurate fit to some
experimental data which require a potential with many parameters.
Figure 7 is perhaps the most interesting of the phase diagrams as it shows large
regions of stability for four of the structures Positive values of C6 expand all the
lattices except D, which is contracted. At a = 4, C6 = 4, for example, the D lattice

r
d' ~H =,

: i I :

I i t ! ~ I ',
oL , t -.~ -L ~ i

4 5 a
~, to

Figure 6. Phase diagram for variation of a and C 3 (other features as in figure 3).
8V
6
Potential energy functions for atomic solids

!,
579

H !
2

o
0
1
B
-2
Downloaded by [Michigan State University] at 01:46 28 February 2015

}: iI
-4

--6 I '
3 4 5 6 a 7 8 9 I0

Figure 7. Phase diagram for variation of a and C 6 (other features as in figure 3).

spacing is 0-65 r~ and its binding energy is about 9 D. The F lattice is the least stable
in this region but both B and H are close in energy to D, For negative values of C6,
B, H and F, all become very stable. C6 gives the greatest discrimination between the
stabilities of the two close-packed structures H and F.
It can be seen from figures (3)-(7) that variation of a and one of the polynomial
coefficients can give a phase boundary between all pairs of solids except D with B
and F. Positive values of Co favour D and S and negative values of C6 favour F, and
indeed at a = 3, Co = 8, C6 = - 6 all five solids have about the same stability
although this is low (less than D) and occurs at large values of r~ (1.5r~ to 2ro).
Clearly the full variation of a and all the coefficients gives great flexibility to the
relative and absolute stabilities of the structures.
If the two-body terms are replaced by either of the p-6 potentials, (10) or (11),
the situation is not greatly changed, although the patterns of the phase diagrams are
altered in small details. The main difference is that the repulsive wall of the two-
body terms, determined by Z, and the range of the three-body terms, determined by
a, are not connected. For standard values of Z, say, 15 in (10) or 12 in (11), the
repulsive wall is quite steep so there is more resistance to contraction of the lattice
for attractive three-body terms than there is with the Rydberg potential.
Figures 8-12 show phase diagrams for the L-J potential (11) with Z = 12, as
functions of a and one coefficient in the three-body term. A significant difference
from figures 3-7 is the absence of B structures for low a in figures 8, 10 and 12; these
appeared in the corresponding figures 3, 5 and 7 mainly as a result of the two-body
terms.
It can be seen from figures 8-12 that all phase boundaries occur except D/B and
D/F; the same situation found before.
Finally, in this section, we turn to an asymptotically r - " three-body term. The
leading three body dispersion energy is given by the Axilrod-Teller triple dipole
term, normally written [13].
v(3 cos 0, cos 0b cos 0c + 1)/(r,,~r~crc~3, (13)
580 J . N . Murrell and R. E. Mottram

3
Downloaded by [Michigan State University] at 01:46 28 February 2015

0 , i i
3 4 5 6 7 8 9 I0
a

Figure 8. Phase diagram for variation of a and C o using the L-J two-body potential (11)
with Z = 12. Symbols as in figure 3).

where 0i are the angles of the three atom triangle and r o the internuclear distances.
The parameter v is negative for inert gas trimers and roughly proportional to the
product of the pair dispersion energy coefficients.
Transforming expression (13) to Q~ coordinates gives a very complicated expres-
sion which we will not quote. However, if this expression is expanded in inverse
powers of
QI = Q1 + x/3. (14)

B
1.5

1.0

0.5 L
/
i ) H
0.0
3 4 5 7 9 10
a

Figure 9. Phase diagram for variation of a and C 1 (other features as in figure 8).
Potential energy functions for atomic solids 581
7

H :
2
Downloaded by [Michigan State University] at 01:46 28 February 2015

3 4 5 a 6 7 8

Figure 10. Phase diagram for variation of a and C2 (other features as in figure 8).

The leading terms are


V(a) _ 192-9V(Q,
------T0 (1 _ 0-409(Q~
2 ) ( Q+, 0Qs
2 _ 19-48(Qs
3~Q,~_~3Q~Q~).). (15)

It can be confirmed that this reduces to expression (13) for equilateral triangles
(Q2 = Q3 = 0, cos 0 = 0-5, Q'I = x/3r) 9
With negative values of v and either of the p - 6 two-body potentials, structure F
becomes the most stable; for example, with the L-J function (11) and Z = 12, F is

-13~ 5 6 a ? 8 9 10

Figure 11. P h a s e d i a g r a m for v a r i a t i o n of a a n d C 3 ( o t h e r features as in figure 8).


582 J . N . Murrell and R. E. Mottram

8y
10

H
4

m
Downloaded by [Michigan State University] at 01:46 28 February 2015

-2
3 4 5 6 a 8 9 i0

Figure 12. Phase diagram for variation of a and C 6 (other features as in figure 8).

the most stable for v < 0-08. The Axilrod-Teller three-body dispersion term favours
the F structure over H for the inert gas solids but it appears to be insufficient to
explain the difference in stabilities between the two [6]. For positive values of v first
S, then D becomes the most stable. For example, with the L-J potential, Z = 12,
structure S becomes the most stable at 0-35 and D at 0.57. Structure B has no region
of stability for any value of v for any reasonable value of Z for either of the p - 6
potentials. However, positive values of v are not physically realistic and we therefore
conclude that the Axilrod-Teller potential is not generally going to be important for
the potentials of any but the inert gas solids.

4. Distorted structures
In this section we examine the question of what type of potential will lower the
symmetry of atomic solids from the structures we have been considering. We have
already mentioned, for example, that zinc has a distorted hcp structure [1]. There
are two types of situation to be considered: one is when the potential is analytic at
the high symmetry point and the other when it is non-analytic.
The latter is characterized by Jahn-Teller situations, there being spatial elec-
tronic degeneracy at the high symmetry configuration which is broken on reducing
the symmetry with, of necessity, a lowering of the potential. Jahn-Teller potentials
are non-analytic and can normally be expressed in the form
V = A +_ B 1/2, (16)

which in turn can be taken as the eigenvalues of a 2 x 2 matrix. This situation has
been treated for triatomic molecules (e.g. H3) using the type of potentials we have
already considered [14]. The complication is that both the diagonal and off-
diagonal terms of the 2 x 2 matrix must be written as sums of two-body and
three-body terms. We will not consider this complication in this paper.
Potential energy functions for atomic solids 583

X 3 triatomic molecules are also known where there is a distortion from the D3h
structure but this is not due to a Jahn-Teller situation; 0 3 is an example. In this
case the potential is analytic in D3h configurations and can be expressed by a
single-valued function of the type we have considered [12]. But, quite obviously,
one needs appropriate polynomial terms in the three-body term.
Although all atoms are equivalent in the hcp structure, and all atoms in a shell
about a chosen atom as origin are equivalent in respect of their two-body terms, all
atoms in a shell are not equivalent in their three-body terms. The twelve nearest
neighbours, for example, fall into two sets; six which make up a hexagon in the
(a, b)-plane and the three above and three below this plane. A similar situation
is found at the fourth neighbours (distance ~/3 from the origin) which are six in the
(a, b)-plane and 12 out of the plane. It follows that at the level of three-body terms
Downloaded by [Michigan State University] at 01:46 28 February 2015

there is no symmetry reason why the interatomic distances within the (a, b)-plane
should be the same as the interatomic distances out of this plane.
Table 3 gives the optimum distorted hop structures for selected parameters of
the three-body term; for simplicity the potential has been limited to summations
only over nearest neighbours. Only coefficients in the ranges shown in figures 3-7
have been considered. The distortion is defined by a parameter 2 which is the
multiplying factor for separation of the (a, b)-planes in the hcp lattice (2 = 1 for the
undistorted lattice). It has been confirmed that the stabilisation energy is quadratic
in the three-body coefficients for low values.
Distortions are smaller for high values of a than for low, as expected. The C 3 and
C6 terms give no distortion. Positive values of C O and C2, and negative values of C1
lead to a contraction in the interplanar spacing. However, the distorted hcp lattices
of Zn and Cd both show an expansion in the interplanar distance, their values of 2
being close to 1.14 [I]. This can be obtained with parameters a = 4, C o = - 0 - 1 1 ,
for example, or with the sets a = 4, C1 = - 1, C6 = 13, and a = 4, C 2 = 1, C6 = 7.
Although there is no symmetry reason why the cubic structures should distort,
as all atoms in a shell are equivalent in respect of their three-body terms, there is
also no reason given an appropriate set of coefficients why they should not distort.
Taking the fcc structure, for example; if we again confine our attention to first
neighbours, then the only three-body terms which differ from those in the hcp lattice
are those between the three atoms above and the three atoms below the (a, b)-plane.
The three-body triangles for these are all isosceles with obtuse apex angles of 120 ~
or 180 ~ (109 ~ or 146 ~ for the hcp structure) and hence the shape terms in the

Table 3. Distortion of the hep lattice for selected values of a (8) and C, (12). Rmi. is the
interatomic distance in the (a, b)-plane (units re) and Vmi, (units D) the optimum
potential. The distortion is measured by 2 the multiplying factor for the separation of
(a, b)-planes. Values for the optimum undistorted lattice (2 = 1) are also given.

Rmi n - Vmi n
a n C, 2 R,,i, --Vml, (2 = 1) (2 = 1)
3 0 0-5 0.84 1"37 3"495 1"30 3"443
10 0 0"5 0"92 1"10 3"899 1"07 3.815
3 1 --0"5 0"95 1"07 6-982 1-05 6"970
10 1 --0-5 0"99 I'01 6"095 1"01 6-092
3 2 2"0 0"96 0-97 4-863 0-96 4"852
10 2 2-0 1.00 1.00 5.912 1.00 5-912
584 J . N . Murrell and R. E. Mottram

three-body polynomial are not very different. Indeed we find the distortions of the
fcc structure have a very similar pattern to those given in table 3. It is interesting to
note that the normal (0t) form of crystalline Hg has a rhombohedral structure which
can be derived from fee by compression along a body diagonal [1].
However, we find no distortion along the lattice vectors of the fcc structure (the
4-fold axes of the nearest neighbour set) for any choice of the coefficients Ci so we
conclude that if there is a set of coefficients which distorts fcc, it will be to produce a
lattice of much lower symmetry.

5. A preliminary examination of solid silicon


We have already mentioned the work of Stillinger and Weber on solid silicon
Downloaded by [Michigan State University] at 01:46 28 February 2015

[2]. Their potential is a function of seven parameters chosen in an arbitrary manner


to ensure that the diamond structure was slightly more stable than the cubic struc-
tures and that computer simulation gave a melting point and liquid structure in
reasonable accord with experiment. Energy and distance scaling parameters were
used to reproduce the experimental lattice energy and spacing of the crystal. One
particular feature of their potential is that the two-body and three-body terms have
a cut-off at 3.77/k which is 1-6 times the interatomic spacing, so that only nearest
neighbours contribute to the energy of the diamond lattice. Other potentials have
been derived for silicon and some have been reviewed by Cowley [15], who con-
cludes that the Stillinger-Weber gives the best general picture and is appreciably
better than others in predicting the Raman absorption frequency of the solid.
The two-body Rydberg potential for the ground state of Si 2 has parameters
D = --3-242eV, r c = 2.246/~, a = 4.571. The interatomic distance in the solid is
2.35 A, or 1.05 re, and the atomization energy is 4.63 eV, or - 1 . 4 2 8 D. Of the three
regions in figures 3-7 where D is the most stable structure, that of figure 7 has a very
contracted lattice relative to the two-body potential whereas those in figures 3 and 5
are very expanded. Although there is no reason for the two-body term to be identi-
cal to the Si 2 potential, it is unlikely to be very different and we conclude therefore
that Co and C2 are likely to be the important parameters in the potential.
Co and C2 as a pair strongly influence the S/D phase boundary. This is shown,
for example, in figure 13 for a - - 4 . 5 . If C O = 0.25 and C 2 = 3.55 the diamond
structure has an energy - 1 - 3 7 D and an interatomic distance of 1.11 re; these are
close to the required values. Adding Ca = - 0 " 3 to this set gives a diamond struc-
ture with the correct energy but the interatomic distance is 1.09 r=. With C 3 = - 0 " 6
the interatomic distance is correct but the lattice is a little too stable, - 1 - 5 2 D . It
appears that with the Si2 ground state two-body potential it will be possible with
further optimization of the parameters to reproduce both the lattice energy and
distance; the elastic constants may of course be poor.
In the Stillinger-Weber potential the two-body potential has no relationship to
the Si2 potential and the atomization energy of the solid is approximately twice the
value of their empirical two-body term. Along the (Co, C2) S/I) boundary shown in
figure 13 the atomization energy for D runs from 0.87D (r = 1.25r,) to 1-37D
(r = 1-11re) and at the largest of these (C O = 0.25, C2 = 3.55) the B, H a n d F
structures have energies 1.13D, 1-20D and 1-16D respectively. The Stillinger-
Weber ratios of the atomization energies are roughly consistent with these.
At the present level of analysis there would seem to be no difficulty in finding a
reasonable potential of the type we have examined for solid silicon that reproduces
Potential energy functions for atomic solids 585

4
i

i i
D
/

2
Downloaded by [Michigan State University] at 01:46 28 February 2015

!
i
i
o
o 3 Co 4 5 6

Figure 13. Phase diagram for variation of C Oand C 2 with the Rydberg potential, a = 4.2.

the atomization energy, interatomic distance and behaviour on melting (the melting
point and the increase in density on melting). A more critical test will be whether a
potential can be found that reproduces the elastic constants. The advantage of our
approach is that it will be possible to set up analytical relationships between the
parameters of our potentials (the two-body parameter and the three-body
coefficients), and the elements of the force constant matrices from each neighbour
shell. This is the procedure used already by Cowley [15] for the StiUinger-Weber
potential and is based on very extensive analyses of the relationship between the
elastic constants of the solid and potential for the lattice [16]. This extension of our
work is now in progress.

References
Eli WELLS,A. F., 1984, Structural Inorganic Chemistry, 5th edition (Clarendon Press).
[2] STILLINGER,F. H., and WEBER,T. A., 1985, Phys. Rev. B, 31, 5262.
[3] MISTRIOTIS,A. D., FLYTZANIS,N., and FARANTOS,S. C., 1980, Phys. Rev. B, 39, 1212.
[4] MURRELL,J. N., 1989, Int. J. quant. Chem. (in the press).
[5] MADDEN,P., and ALLISON,C. M., 1989 (private communication).
[6] NIEBEL, K. F., and VENABLES,J. A., 1976, Rare Gas Solids, Vol. 1, edited by M. L. Klein
and J. A. Venables (Academic Press), Chap. 9.
[7] RAGHAVACHARI,K., 1986, J. chem. Phys., 84, 5672.
[8] Ko~os, W., NIEVES,F., and NOVARO,O., 1976, Chem. Phys. Lett., 123, 321.
[9] WEYL,H., 1946, The Classical Theory of Groups (Princeton U.P.).
[10] MORIARTY,J. A., 1988, Phys. Rev. B, 38, 3194.
[11] MURRELL,J. N., and SORreL,K. S., 1974, J. chem. Soc. Faraday Trans. II, 70, 1552.
[12] MURRELL,J. N., CARa~R, S., FARANTOS,S. C., HUXLEY,P., and VARANDAS,A. J. C., 1984,
Molecular Potential Energy Functions (Wiley).
[13] AXILROD,B. M., and TELLER, E., 1943, J. chem. Phys., 11, 299.
[14] MURRELL,J. N., and VARANDAS,A. J. C., 1986, Molec. Phys., 57, 415.
[15] COWLEY,E. R., 1988, Phys. Rev. Lett., 23, 2379.
[16] HERMAN,F., 1959, J. Phys. Chem. Solids, 8, 405.

You might also like