You are on page 1of 12

F. Zahid et al.

: Static and Dynamic Properties of Liquid Less Simple Metals

987

phys. stat. sol. (b) 215, 987 (1999)


Subject classification: 61.25.Mv; 65.50.+m; 66.20+d; S1; S4

Investigations of the Static and Dynamic Properties


of Liquid Less Simple Metals
F. Zahid (a), G.M. Bhuiyan1 ) (b), S. Sultana (a), M.A. Khaleque (a),
R.I.M.A. Rashid (a), and S.M.M. Rahman (c)
(a) Department of Physics, University of Dhaka, Dhaka-1000, Bangladesh
(b) Technische Universitat Chemnitz, Institut fur Physik, D-09107 Chemnitz, Germany
(c) Department of Physics, Sultan Qaboos University, P.O. Box 36, Postal Code 123,
Al-Khod, Muscat, Oman
(Received August 21, 1998; in revised form June 9, 1999)
Static structure factors, Sq, thermodynamic properties and shear viscosities for liquid Zn, Cd, Hg,
In, Tl, Sn, Pb, Sb, and Bi are investigated systematically. A model pseudotential which combines
the sp- and d-band contributions is employed to derive interionic interactions. The liquid state is
described by using the thermodynamically self-consistent variational modified hypernetted chain
(VMHNC) integral equation theory, and also by using the linearised version of the Weeks-Chandler-Andersen (LWCA) thermodynamic perturbation theory of liquids. Results of calculations for
all of the above properties are found to be in good agreement with available experimental data.
The calculated results also reveal qualitatively that there is a relationship between the rate of
change of viscosity with temperature and the depth of the potential well concerned.

1. Introduction
Elements lying near the bottom of the 3d, 4d and 5d series (for example, Zn, Cd, Pb,
Hg, Sn, In etc.) have completely filled d bands. Nonetheless, their physical properties
are significantly influenced [1 to 3] by the d bands via sd mixing (sometimes referred to
as sd hybridization). This is the main reason to call these elements as the less simple
metals. Therefore, a complete description of these systems requires a model that can
take into account both sp- and d-band effects in the interionic interactions.
Recently Bretonnet and Silbert (BS) [4] have proposed a model to describe interionic interactions, primarily, for liquid transition metals. This model treats sp and d
bands separately within the well established pseudopotential formalism. The sp band is
described via the empty core model; the d-band contribution is derived from the dband scattering phase shift by using the inverse scattering approach. The resulting model pseudopotential thus reduces to a simple local form which appears similar to that of
the Heine-Aberenkov model [5]. The main difference between these two is that the
core term in the former one is a function of distance and in the latter it is a constant.
The BS model is simple to handle numerically. Moreover, the local form permits one to
extend this model to other liquid metals for which effects of sd hybridization are significant. The BS model has also proved to be successful for liquid transition metal calculations [6, 7]. Note that the norm conserving non-local pseudopotentials are, in principle,
1

) Corresponding author: e-mail: physics@du.bangla.net

988

F. Zahid et al.

to be preferred for accurate predictions. But there are evidences that local pseudopotentials describe physical properties, in some cases, even better [1] than the former
ones.
In the theory of liquid metals the knowledge of static structure factor, Sq, or its
fourier transform, gr, is the essential prerequisite for a complete description of static,
thermodynamic as well as transport properties. The study of viscosity is important for
the metallurgical and industrial purposes. For example, the rates of many industrially
important reactions are limited by the diffusion of the reactant species. Moreover,
knowledge of viscosity is required in the theoretical determination of critical cooling
rates for glass formation.
In the present work static structure factors, specific heat at constant pressure and
shear viscosities are investigated in the vicinity of the melting temperatures. The temperature dependence of viscosity has also been examined. Two separate liquid state
theories are used to carry out the structural calculations. One is the thermodynamically
self-consistent variational modified hypernetted chain (VMHNC) [8] integral equation
theory and another is the linearised version of the Weeks-Chandler-Andersen (LWCA)
[9] thermodynamic perturbation theory of liquids. Although huge bank of experimental
[10 to 13] as well as theoretical [14] data are available in the literatures, nobody until
now performed systematic calculations both for static and transport properties on the
same footing. In addition it is also tempting to extend the BS model to the study of
liquid simple metals.
Recent experimental data [13] as well as theoretical calculations [14, 15] suggest that
hard sphere (HS) liquids describe some physical properties of simple liquid metals very
well. Recently, Itami and Sugimura [16] reported that HS results of shear viscosity for
liquid Sn agree better with the corresponding space laboratory data [17]. Furthermore,
HS analytic expressions for the above physical properties are well documented [13 to
15]. Therefore, it is natural to choose a HS reference as a first approximation in the
perturbative calculations.
The layout of this paper is as follows. In Section 2 we briefly describe the derivation
of the effective interionic interaction. The outline of the VMHNC and LWCA theories
of liquids is presented in the same section. This section also includes formulas for shear
viscosity calculations. In Section 3 results of calculations are presented and discussed.
Some concluding remarks are presented in the same section.

2. Theories
2.1 Effective pair potentials
The local pseudopotential may be constructed by the superposition of sp- and d-band
contributions as [4]
8 2
 r 
P
>
<
Bm exp
if r < Rc ;
ma
1
wr m1
>
: Z
r
if r > Rc ;
where a, Rc and Z stand for softness parameter, core radius and the effective s-electron
occupancy number, respectively. The term inside the core is obtained from d-band scattering phase shift by using an inverse scattering approach. This term takes care of the

Static and Dynamic Properties of Liquid Less Simple Metals

989

d-band effect. The term outside the core is the bare Coulomb interaction between an
electron and an ion. The coefficient Bm depends on a, Rc and Z [18]. The unscreened
form factor of (1) may be written as
"
# 

B1 J1
8B2 J2
4pZ
3

n cos qRc
wq 4pna

2
q2
1 a2 q2 2 1 4a2 q2
with


  

Rc
Rc
sin qRc
1 m2 a2 q2 1 m2 a2 q2
Jm 2 exp
ma
ma
maq



Rc
2
1 m2 a2 q2 cos qRc ;
ma

where n denotes ionic number density. The effective interionic potential is given by
0
1
1

Z@
2
FN q sin qr dqA ;
4
vr
1
r
p
0

where
FN

q2 wq2
4 pZ n

!2 



1
1
1
:
Eq 1 Gq

In (5), Eq denotes the dielectric function and Gq takes care of the local field corrections; these functions are taken from Ichimaru-Utsumi [19].
2.2 The VMHNC theory
The variational modified hypernetted chain (VMHNC) theory, originally proposed by
Rosenfeld [8], belongs to a new generation of fairly accurate integral equation theories
of liquids. Like most other integral equation theories the VMHNC solves the OrnsteinZernike (OZ) equation with a closure relation
gr exp hr cr bvr Br ;

where b is the inverse of temperature times Boltzmann constant. The bridge function
Br is approximated by using the analytic solution of the Percus-Yevick equation for
HS namely Br BHS
PY r; y. The packing fraction y is the variation parameter which is
determined by minimising the VMHNC configurational Helmholtz free energy. Once y
is fixed for a specific thermodynamic state, Sq or pair correlation function gr can be
evaluated.
2.3 The LWCA theory
The starting point of the linearised version of the Weeks-Chandler-Andersen (LWCA)
theory [9] is the WCA thermodynamic perturbation theory [20]. In the WCA the socalled blip function is defined as
Pr Ys r fexp bvr exp bvs rg ;

990

F. Zahid et al.

where vr and vs r denote the soft sphere and HS potentials, respectively. Ys is the
cavity function. The function r2 Pr gives two sharp tooth-shaped nonlinear features
when plotted as a function of r. In the LWCA theory these are approximated by rightangle triangles (for details see Ref. [9]). Fourier transform of Pr is then expanded in
terms of Bessel functions. By use of the thermodynamic condition that for an effective
HS diameter Pq vanishes, one gets [9]


2 bsv0 s S 2
bvs ln
:
8
bsv0 s S 2
The solution of (8) gives the effective value of the HS diameter s.
2.4 Shear viscosity
If liquid metals are approximated by HS liquids then the shear viscosity may be written
as [14, 21]
h0 hh hs hk ;

where the first term on the right is the contribution of the hard part of the potential,
1

hh hh hh ;
1

hh
2

hh

10



kTy
8y
1 gy Dy ;
2gy
5

11

96y2 gy kT
;
5p
W

12

with

Dy W
and
W

1

5zs
8nMgy

2
6
41

3
4W
7
5
5zs
4W
nMgy


  
4pkT 1=2 6y 2=3
:
M
pn

13

14

The contribution of the soft part of the potential is


4pMn2
hs
30zs


@v
4 @v

gr dr ;
@r2 r @r

15

where
z2s

4pMn

 2

@ v 2 @v
gr dr :
r

@r2
r @r
2

1=3

r>6y=pn

16

991

Static and Dynamic Properties of Liquid Less Simple Metals

The kinetic energy contribution to the shear viscosity may be written as




8ygy
5kT 1
5

:
hk
5zs
8gy W
4nMgy

17

The final expression is


h Cs y h0 ;

18

where Cs y is the scaling correction term which takes into account the effects of multiple scattering.

3. Results and Discussion


The model pseudopotential used in this work has three parameters; Rc , a and Z. These
are chosen in the following way. Values of Rc are generally determined by fitting to the
physical properties of the system of interest, for example, bulk modulus [22], electrical
restivity [23] or structural data [22]. Since we are interested in investigating not only
structural properties but also transport properties, values of Rc are taken from [23] that
involves both structure and transport. Regarding the choice of Z, we follow the concept
that the effect of hybridization between s and d electrons can be approximately accounted for by changing the relative occupancy of the s and d bands [24]. Here by
integrating the partial s-density of states resulting from the self-consistent band structure calculations [25] it is found that a value of Z  1:3 is the most reasonable value for
the elemental systems; this is also supported by the self-consistent calculations by Moriarty [26]. The value is also in conformity with the effective charge transfer noted in
the augmented-spherical-waves (ASW) calculations [27]. Here it is worth noting that
during the progress of our calculations we have observed that for Z 1:3 the VMHNC
theory which employs the full potential (i.e. both repulsive and attractive parts) gives
the best fits to the experimental Sq. Here we may also mention that, in principle,
there is no restriction as such to use different but suitable values of Z for different
systems [7] provided self-consistent calculations of charge transfer support these values.
Finally, the values of the softness parameter a are determined by following the proceTa b l e 1
Input temperature T, ionic number density n, core radius Rc, softness parameter a, and
effective HS diameter s
system

T (K)

n 
A3

Rc (a.u.)

a (a.u.)

s 
A

Zn
Cd
Hg
In
Tl
Sn
Pb
Sb
Bi

723
623
523
433
588
523
613
933
573

0:0637
0:0428
0:0386
0:0369
0:0332
0:0353
0:0310
0:0320
0:0289

1:27
1:23
0:92
1:32
1:13
1:30
1:47
1:06
1:49

0:285
0:253
0:170
0:278
0:218
0:273
0:307
0:193
0:317

2:46
2:70
2:59
2:78
2:82
2:75
2:98
2:86
2:92

992

F. Zahid et al.

Fig. 1. Effective pair potentials for a) liquid Zn, Tl, Sn, Pb, and Sb; b) liquid Cd, Hg, In and Bi

dure of Bhuiyan et al. [7]. The values of the parameters are listed in Table 1. Resulting
effective pair potentials are illustrated in Fig. 1. From the figure it appears that the
depth of the potential well is the minimum for Zn and the maximum for Sb; depths for
all the rests lie in between these two limits. On the other hand the distance of the
position of the first minimum is found to be the smallest for Zn and the largest for Pb.

Fig. 2. Sq for a) liquid Pb, In, Zn, Cd and Hg; b) liquid Tl, Bi, Sb and Sn. Solid lines denote the
VMHNC results, dotted lines the LWCA, and discrete circles and triangles the experimental data

Static and Dynamic Properties of Liquid Less Simple Metals

993

Liquid structures for Zn, Cd, Hg, In, Tl, Sn, Pb, Sb and Bi near their melting temperatures are calculated by using the VMHNC as well as LWCA theories of liquids.
Values of the input temperatures and densities are taken from experiments [13].
We have used the Gillan's algorithm [28] for numerical solution of the Ornstein-Zernike equation and the closure relation (6). In all cases we have found that a 1024
points grid with step size Dr 0:06 
A is sufficient for the present calculations. Results
of the VMHNC calculations for Sq are presented in Fig. 2. From the figure we notice
that the average profiles of the structure factors for liquid Pb, In, Zn, Cd, Bi are reasonably good and for the rest are fair when compared with the corresponding experimental values [13]. But the values of the principal peak are somewhat smaller than
experimental data for all systems except for Hg, Bi and Sb. For the latter systems the
peak values are of equal height. It is also noticed that the positions of the first peak are
slightly shifted toward large q relative to the experimental data for Hg, Sb and Sn.
Except these discrepancies at the level of the principal peak, overall results in terms of
position and height of the peaks, and phase of oscillations are in good agreement. The
small humps that appear in the high angle side of the main peak of Bi and Sn are not
reproduced by the present theoretical calculations. We believe, the fine details of these
deviations may be interpreted in terms of a combined approach considering (i) the
many-body effects, (ii) the packing consideration of the constituent atoms and (iii) the
ion-core polarizibility inherent to all non-simple metals. In addition, the obvious limitations of the potentials used in the VMHNC might also be partly responsible for the
observed discrepancies. It is believed that particularly the position and height of the
principal peak lie at the heart of a delicate balance between the repulsive and attractive
parts of the interionic interactions. Here, it may be relevant to note that the accuracy of
the VMHNC calculations for liquid simple and transition metals is found to be comparable to that of the computer simulations [29 to 31].
The values of the effective HS diameters, s, calculated by using the LWCA theory
are presented in Table 1. It is seen that the value of s is the largest for Pb and the
smallest for Zn. This is consistent with the relative distance of positions of the principal
minima of their potential profiles which have been described above. The LWCA results
of calculations for Sq are illustrated in Fig. 2 (dotted lines). It appears that the principal peak values, in this case, are somewhat smaller than those of the VMHNC data
except for Zn. For the latter the 1st and the 2nd peak values are found to be slightly
larger. With this exception the qualitative agreement of the LWCA results with
VMHNC as well as with experimental data is reasonably good. Following this agreement we are tempted to extend our calculations for the thermodynamic as well as transport properties of the systems concerned.
The calculated values for specific heat at constant pressure, and entropies are presented in Table 2. From the table it is noticed that the calculated values for entropies
are somewhat larger than the observed data [10] for all systems except for liquid Zn,
for the latter it is found to be slightly smaller. The difference between the theory and
experiment is found to be the largest for liquid Hg and the smallest for Zn. Table 2
also shows that the theoretical values for specific heat, CP , are slightly larger than corresponding experimental data [10] for all systems except for Tl, Sb and Bi. For the
latter systems the values are slightly smaller. The discrepancy between the theory and
experiment is seen to be more visible for Zn. However, the difference between the
theoretical and experimental values lies within 1 to 14% of the experimental data. It

994

F. Zahid et al.
Ta b l e 2
Specific heat capacity CP and total entropy Stot. Both are divided by NkB, N being the
total number of atoms
system

CP

Zn
Cd
Hg
In
Tl
Sn
Pb
Sb
Bi

Stot

theory

experiment

theory

experiment

4:41
3:86
3:24
3:65
3:47
3:44
3:78
3:47
3:38

3:88
3:82

3:52
3:63
3:33
3:48
3:78
3:56

8:41
10:67
12:84
10:79
12:64
11:67
12:06
12:59
13:00

9:00
9:70
8:3
9:10
10:90
9:70
11:10
11:70
11:20

may be relevant to mention here that, since thermodynamic properties are global in
nature, these figures may lie even within the experimental errors. This suggests that the
overall qualitative agreement is reasonably good.
Now, we turn to the results of calculations for shear viscosities of the systems concerned. The theoretical values are presented in Table 3 and are compared with the
corresponding experimental data [11, 12]. It is noticed that the major contribution to
the viscosity arises, unlike noble and transition metals [32], from the soft core part of
the potential for all systems except for Zn. The hard core part contribution is significant in this case but the kinetic energy contribution is negligibly small. This nature of
Ta b l e 3
Calculated values of various contributions to shear viscosities (in cP); notations are defined in the text. The second row for each system refers to the calculations without
d-band contribution in the potentials. Experimental data are from Ref. [11], bracketed
values from [12]
1

system

hh

hh

hs

hk

h0

h theo:

h expt:

Zn

0:0379
0:0565
0:0222
0:0196
0:0153
0:0074
0:0135
0:0150
0:0193
0:0126
0:0135
0:0156
0:0236
0:0232
0:0212
0:0091
0:0163
0:0188

1:1382
1:6850
0:7201
0:4842
0:4349
0:1306
0:4558
0:4301
0:5550
0:1988
0:3987
0:3047
0:7334
0:5778
0:5263
0:1218
0:4530
0:4474

0:6974
1:2801
1:2364
1:0333
1:5418
1:5404
1:0602
0:8541
1:2358
1:0210
1:0306
0:7716
1:1840
0:9294
0:9104
0:7999
1:0792
0:8454

0:0477
0:0575
0:0315
0:0318
0:0273
0:0226
0:0203
0:0231
0:0309
0:0327
0:0219
0:0301
0:0341
0:0365
0:0339
0:0276
0:0272
0:0314

1:9212
3:0793
2:0102
1:5690
2:0194
1:7011
1:5519
1:3223
1:8409
1:2652
1:4647
1:1221
1:9749
1:5671
1:4917
0:9585
1:5757
1:3432

3:46
5:54
2:41
1:88
2:02
1:70
1:78
1:52
2:02
1:40
1:61
1:23
2:38
1:89
1:64
1:05
1:73
1:47

3.85 (3.50)

Cd
Hg
In
Tl
Sn
Pb
Sb
Bi

2.28 (.)
2.10 (2.04)
1.89 (1.80)
2.64 (2.64)
1.85 (1.81)
2.65 (2.61)
1.22 (1.43)
1.80 (1.63)

Static and Dynamic Properties of Liquid Less Simple Metals

995

our results is broadly supported by other theoretical calculations [14] for liquid simple
metals. For Zn the major contribution arises from the hard core part of the potential;
we shall discuss the possible cause of this later. Rahman and Bhuiyan [3] have calculated shear viscosities at different temperatures for Cd, In and Pb. They have used the
Fermi sphere model potentials and the Gibbs-Bogoliubov variational scheme, and have
found the values to be 2.8, 2.0 and 2.4 (cP) for Cd, In and Pb, respectively, at temperatures of the present work. In [33] the values of viscosity are found to be 2.1 for Pb.
Since the theory of viscosity employs the full potential for its evaluation, and the
potential involves both sp- and d-band effects, it is interesting to examine how results
of viscosity are affected by the absence of d bands. To this end, we have performed
calculations without d-band contributions in the potentials, i.e., with the empty core
potential alone. These results are also presented in Table 3. From the table it is noticed
that results calculated with d band predict experimental data much better than those
calculated without d band, for all systems, and the difference between the two sets of
values is significantly large.
Various theoretical values for h are reported by different authors [12 to 14]. As far as
overall agreement with experimental data is concerned, results of the present calculations are found to be the best for the concerned systems. The values of the scaling
correction term, Cs y, used in the above calculations are taken from [33].
In this paper we have reported on the temperature dependence of the shear viscosity.
To do this accurately, within the present formalism, we indeed need temperature dependent potentials. In our case, potentials do not depend explicitely on temperature. But to
have a feeling about the temperature dependence of pair interaction, we have examined the effects of the temperature dependence of density on potentials. In this case we
have observed that, above the melting point (but not very far away from it), the change
in density with increase of temperarure is small. Consequently the resulting change in
potentials is not adequate for LWCA theory to produce distinctly different HS diameters at different temperatures. Protopapas et al. (PAP) [34] in their original work on
transport properties of liquid metals proposed an empirical formula to evaluate temperature dependence where the packing fraction at melting has been assumed to be
0.472 for all systems investigated. Itami
and Sugimura (IS) [16] proposed another formula for temperature dependent hard sphere diameter (HSD). They
also choose a convenient value of HSD
at melting but suggested that it should
be determined theoretically by using
the WCA type of methods. Following

Fig. 3. Hard sphere diameter, s, versus temperature curves for liquid Zn, Cd, Hg, In,
Pb, Tl, Sb, Sn and Bi obtained by using the
formula of Ref. [34]

996

F. Zahid et al.
Fig. 4. Shear viscosity as a function of temperature for liquid Zn, Cd, Hg, In, Pb, Tl, Sb,
Sn and Bi

this suggestion we have first determined


the packing fraction near melting temperature by using the LWCA theory and
then follow the procedure of [34]. It is
worth noting that both IS and PAP formulas show similar temperature dependent curves for HSD but these differ
slightly in magnitude [32]. The temperature dependence of HSD is illustrated in
Fig. 3. It is seen that s versus T curves
exihibit, roughly speaking, a similar rate
of change. But a close look allows one to divide these into three classes. The first class,
for example Hg, has a rapid rate of decrement of HSD with increase of temperature.
For the second class, for example, for Cd, Sb and Zn, the rate of change of decrement
is very slow. For the third class (i.e. for the rest) the rate of change lies in between
these two limits. The resulting change in shear viscosity is illustrated in Fig. 4. Here the
trend of the rate of change is different from that of the HSD. For Zn the rate of
change of viscosity with temperature is found to be the greatest, and it is the least for
Sb; for all other systems the rate of change lies in between these two extreme values.
From the potential profile it is seen that the depth of the potential is the largest for Sb
and the smallest for Zn. The same trend between the rate of change of viscosity and
the depth of potentials also holds for others (for example, for Cd, Tl, Bi and Hg). So,
qualitatively, it appears that liquid simple metals having deeper potential wells show a
slow rate of change of viscosity with temperature and vice versa. Atoms in a liquid
metal with a deeper potential well should have, in principle, a higher correlation which
can be one of the possible causes responsible for the slow rate, and vice versa. But the
complete understanding of this trend from the microscopic point of view needs further
work along this line.

4. Conclusions
Static structure factors for less simple liquid metals are calculated by using the Bretonnet and Silbert (BS) model in conjuction with the VMHNC as well as LWCA theory of
liquids. The overall agreement with experimental data is found to be reasonably good.
But, as far as static structure is concerned, potentials for liquid Bi, Sn, Sb and Hg need
improvements. At this stage it is not clear whether any improved form of parameterization would lead to improvement in the static structure results.
Thermodynamic properties, namely the specific heat and entropy, and one transport
property, the shear viscosity, are also investigated by using the same potentials. Reasonably good agreement with experiment has been found. For liquid Zn we have found
exceptional results for both thermodynamic and transport properties. When we look at

Static and Dynamic Properties of Liquid Less Simple Metals

997

the potential profiles (see Fig. 1) it is seen that for Zn the potential well is very shallow.
It is also seen that the value of the position r of the first minimum is the smallest, and
oscillations are almost out of phase relative to the potentials of the rest. Therefore, we
assume that the cause of discrepancy for Zn is due to the unusual features of its potential. Now the question arises why the results of Sq for Zn are in good agreement.
This may be understood in the following way. It is well known that the static structure
of liquid metals is mostly determined by the repulsive part of the potential. The effects
of the attractive part and the part of the Friedel oscillations on Sq are not significant.
The results of our calculations show that the soft part contribution of the potential to
the viscosity predominates. This result, thus, emphasizes that the attractive part of the
potential plays an important role in describing, unlike static structure, the transport
properties.
Calculations of viscosity, in the present work, have employed the full potentials that
involve both sp- and d-band contributions (see Section 2.4). The results for viscosity,
calculated within the framework of our parameterization, are found to be much better
in agreement when d-band contribution is included than those obtained without d
bands. This result, we believe, clearly demonstrates the effects of the d band on the
viscosity of the less simple liquid metals. However, our results for shear viscosities
(near melting temperatures) may be considered as the best, in term of agreement with
experiment, among all values so far reported in the literature.
The temperature dependence of viscosity shows a qualitative trend that liquid metals
with deep potential well have slow rate of change of viscosity with temperature and
vice versa.
Transport properties are very sensitive to profile of the potential. Therefore, the quality of our results suggests that the BS model potentials have right features to represent
the real metallic systems, and can be used as a starting point for realistic calculations.
Moreover, the quality of our results for the shear viscosity, and the amount of differences between the results with and without d-band contributions also suggest that, to
describe physical properties of the liquid less simple metals quantitatively, one must
include d-band effects properly in the interionic interactions.
Acknowledgement One of us (G.M.B.) is thankful to Alexander von Humboldt foundation for financial support and hospitality in Germany where the last part of the work
has been completed.

References
[1] C. Fiolhais, J.P. Perdew, S.Q. Armster, J.M. MacLeran, and M. Brajczewska, Phys. Rev. B 51,
14001 (1995).
[2] Y. Ari, Y. Shirakawa, S. Tamaki, M. Saito, and Y. Waseda, Phys. Chem. Liq. 35, 253 (1998).
[3] S.M.M. Rahman and L.B. Bhuiyan, Phys. Rev. B 33, 7243 (1986).
[4] J. L. Bretonnet and M. Silbert, Phys. Chem. Liq. 24, 169 (1992).
[5] V. Heine and I. Abarenkov, Phil. Mag. 9, 451 (1964).
[6] G.M. Bhuiyan, J.L. Bretonnet, L.E. Gonzalez, and M. Silbert, J. Phys.: Condensed Matter 4,
7651 (1992).
[7] G.M. Bhuiyan, J.L. Bretonnet, and M. Silbert, J. Non-Cryst. Solids 156/158, 145 (1993).
[8] Y. Rosenfeld, J. Stat. Phys. 42, 437 (1986).
[9] A. Mayer, M. Silbert, and W.H. Young, Chem. Phys. 49, 47 (1980).
[10] R. Hultgren et. al., Selected Values of Thermodynamic Properties of Elements, American
Society for Metals, Ohio 1973.

998

F. Zahid et al.: Static and Dynamic Properties of Liquid Less Simple Metals

[11] E.A. Brandes, Smithells Metals Reference Book, 6th ed., Butterworth, London 1983.
[12] T. Iida and R.I.L. Guthrie, The Physical Properties of Liquid Metals, Clarendon Press, Oxford
1993.
[13] Y. Waseda, The Structure of Non-Crystalline Materials, McGraw-Hill Publ. Co., New York
1980.
[14] M. Shimoji, Liquid Metals, Academic Press, London 1977.
[15] W. H. Young, Rep. Progr. Phys. 55, 1769 (1992).
[16] T. Itami and K. Sugimura, Phys. Chem. Liq. 29, 31 (1995).
[17] G. Frohberg, in: Material Science in Space, Eds. B. Feuerbacher, H. Hamacher, and R.J. Neumann, Springer-Verlag, Berlin 1986 (p. 425).
[18] J.L Bretonnet, G.M. Bhuiyan, and M. Silbert, J. Phys.: Condensed Matter 4, 5359 (1992).
[19] S. Ichimaru and K. Utsumi, Phys. Rev. B 24, 7385 (1981).
[20] H.C. Andersen, C. Chandler, and J.D. Weeks, Adv. Chem. Phys. 34, 105 (1976).
[21] S.A. Rice and A.R. Allnatt, J. Chem. Phys. 34, 2144 (1961).
[22] C. Regnault, Z. Phys. B (Condensed Matter) 76, 179 (1989).
Ch. Hausleitner, G. Khal, and J. Hafner, J. Phys.: Condensed Matter 3, 1589 (1991).
[23] W.A. Harrison, Electronic Structure and the Properties of Solids: The Physics of the Chemical
Bond, W.H. Freeman Co., San Francisco 1980.
[24] J. M. Wills and W. A. Harrison, Phys. Rev. B 28, 4363 (1983).
[25] I. Moruzzi, J. F. Janak, and A. R. Williams, Calculated Electronic Properties of the Elements,
Pergamon Press, New York 1978.
[26] J.A. Moriarty, Phys. Rev. B 42, 1609 (1990).
[27] J. Kuebler, private communication, 1994.
[28] M.J. Gillan, Mol. Phys. 39, 839 (1979)
[29] L.E. Gonzalez, D.J. Gonzalez, and M. Silbert, Physica B 168, 39 (1991).
[30] M.A. Khaleque, G.M. Bhuiyan, R.I.M.A. Rashid, and S.M.M. Rahman, Phys. Chem. Liq. 30, 9
(1995).
[31] G.M. Bhuiyan, M. Silbert, and M.J. Stott, Phys. Rev. B 53, 636 (1996).
[32] G. M. Bhuiyan, A. Rahman, M.A. Khaleque, R.I.M.A. Rashid, and S.M.M. Rahman, Phys.
Chem. Liq., in press.
[33] R.N. Joarder and R.V.G. Rao, phys. stat. sol. (b) 116, 299 (1983).
[34] P. Protopapas, H.C. Andersen, and N.A.D. Parlee, J. Chem. Phys. 59, 15 (1973).

You might also like