Download as pdf or txt
Download as pdf or txt
You are on page 1of 26

ANNUAL

REVIEWS Further
Ann. Rev. Phys. Chern. 1980. 31:131-56 Quick links to online content
Copyright C> 1980 by Annual Reviews Inc. All rights reserved

CONDUCTION IN FUSED .2703

SALTS AND SALT -METAL


SOLUTIONS
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

Norman H. Nachtrieb
Department of Chemistry, The University of Chicago, Chicago, Illinois 60464
by University of Sussex on 01/04/13. For personal use only.

INTRODUCTION

The field of molten salt chemistry is an important and growing one, and
is of interest at both fundamental and technical levels. In their efforts to
understand the equilibrium and transport properties of molten salts,
workers face all of the problems inherent to the liquid state, for which
there still exists no fully satisfactory theory. At the technical level, the
vast electrometallurgical industry depends upon ionic transport in
molten salts, whose Faradaic efficiency is a nontrivial factor in the
energy budgets of nations.
The field does not stand in isolation from other areas of research but
has important interfaces, if not areas of overlap, with certain interests in
liquid metals, the glassy state, and even nonpolar liquids and aqueous
electrolyte solutions. The transport of charge in an electric field is one
of the most characteristic properties of molten salts; it is related to other
transport properties, such as diffusivity and viscosity, and a clearer
understanding of this relationship is slowly emerging. Any survey of the
status of the electrical conductance in molten salts must necessarily
touch upon other transport and equilibrium properties of molten salts.
"Good" salts, of which the alkali halides and nitrates are typical, are
completely ionized in the solid and liquid states, having specific elec­
trical conductivities in the latter that lie in the range 1-10 ohm-I cm -1.
Other salts, such as the aluminum and mercuric halides, are very poor
conductors and behave more as molecular compounds in both the solid
and liquid states. However, aluminum and mercuric halides are Lewis
acids and readily form complex ions such as MX:- with halide ion
donors. Aluminum chloride, for example, forms completely ionized salts
such as NaAlCl4 or KAl2Cl7 with molten alkali halides. The pure

131
0066-426Xj80j1101-0131$01.00
132 NACHTRIEB

mercuric halides (HgCI2, HgBr2 and HgI2) have specific conductivities


'
of 3x 10-5, 1.5 X 10-4, and 3x 10-2 ohm-I em-I, respectively, at their
melting points, and what ionic conductivity they have is thought to be
due to autoionization equilibria of the sort
2HgX2 HgX+ + HgX;- .
=

This review lays no claim to comprehensiveness, but singles out topics


that appear to the writer to merit further attention and to offer the
prospect of a deeper insight into the mechanisms of charge transport in
high temperature ionic media. Among these are the use of high pres­
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

sures to test models of the conduction process and to demonstrate the


existence of a transition between the insulator and conductor behavior
of "poor" ionic salts. Another kind of transition-glassy to ionic-is
by University of Sussex on 01/04/13. For personal use only.

exhibited by certain salt mixtures and is receiving much attention. The


fields of molten salts and liquid metals converge on the common ground
of solutions that exhibit their complete or partial miscibility; their
history and current status are reviewed in the context of the metal­
nonmetal transition.
A number of excellent books survey the entire field of molten salt
chemistry, and though some are now old, they are far from out of date.
These include the English translation of DeIimarskii & Markov's Elec­
trochemistry of Fused Salts (1), Sundheim's Fused Salts (2), and Blander's
Molten Salt Chemistry (3). More recent are Lumsden's Ther­
modynamics of Molten Salt Mixtures (4), Janz's Molten Salts Handbook
(5), and Mamantov's Molten Salts: Characterization and Analysis (6). Of
still more recent date are the Proceedings of the International Symposium
on Molten Salts, edited by Braunstein et al (7), and the series of volumes
by Braunstein, Mamantov & Smith entitled Advances in Molten Salt
Chemistry (8). Sutp is the state of development of the field of fused salts
that thermodynamic, spectroscopic, structural, and transport studies are
highly i'nteractive, a:qd the references cited above are an indispensable
introduction to the sub� ect.

ELECTRICAL CONDUCTIVITY
Molten salts obey Ohm's Law, and apparent deviations from it are due,
not to variations in the mobilities of ions with the impressed electric
field, but to D:0n-ohmic effects that may occur at electrodes. These have
their origin in oxidation or reduction reactions, either reversible or
irreversible, that cause local gradients in the concentrations of ions
(concentration polarization) and provide the conditions necessary for
polarography and related non-ohmic electrode processes. The electric
field impressed upon ions in measurements of the electrical conductance
CONDUCTION IN FUSED SALTS 133

or in electrolyses is typically on the order of one volt cm -\, whereas the


mean square fluctuation of the local field at an ion due to thermal
motion is around 107 volt cm - \, a wholly negligible perturbation on the
equilibrium state of the system. Ions drift in a field-biased random walk,
and the close relationship of conductance and diffusion is evident.
Nonlinear field effects are not observed experimentally, and the specific
conductivity, (1, is the modulus of the current density:

J=(1E. l.

C1 is a scalar for isotropic


Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

A second rank tensor for anisotropic solids,


liquids, with units of ohm -\ cm -\; (1 is a direct function of the number
density of ions, and hence, for one-component molten salts, it is a
function of pressure and temperature. To normalize the electrical con­
by University of Sussex on 01/04/13. For personal use only.

ductance to a fixed number of ions (in the case of completely dissocia­


ted salts), it is usual to define the molar conductance, I\m' by
2.
where Vm is the molar volume of the salt.
Experimentally, it is found that both the specific and molar conduc­
tivities of completely dissociated molten salts conform well to the
Arrhenius Law for thermally activated processes

1\ =A exp( Ej RT),
-
3.
where A and E are parameters. This is also the form of equation that
describes certain transport processes in crystalline solids, notably diffu­
sion, where it sometimes holds over many orders of magnitude. The
vacancy model of diffusion in crystalline solids is firmly established, but
for many years it was uncritically adapted to various "hole theories" of
liquids that are now largely displaced by models that recognize the
highly cooperative nature of particle motions in liquids. This uncritical
use of the hole theory applies to various one- or two-event models of
transport properties of liquids that are still proposed.
The failure of hole theories for liquids, however, is no reason to
discard the thermodynamic formalism of the activated state. It is an
experimental observation that conductance and other transport processes
are thermally activated, in the sense that they are a function of tempera­
ture or the mean kinetic energy. It is important to bear in mind that the
activation energy, entropy, free energy, volume, etc are ensemble aver­
ages over many configurations of a system and have no necessary
connection with the dynamics of a single particle in isolation from its
environment. With this understanding, there is no reason to put quota­
tion marks around the activation energy or to refer to it as a parameter.
It is a bulk property of a system that is at equilibrium or very nearly so,
134 NACHTRIEB

and a1l of the machinery of thermodynamics may be employed with


confidence. Once more the caveat: The thermodynamic quantities as­
sociated with transport properties are macroscopic system quantities,
averages over the dynamical properties of ensembles of a many-particle
system, and have no direct relation to the motion of a single particle.
It is not surprising that the results of careful studies of the electrical
conductance of "good" ionic salts show no temperature-dependence of
the activation energy. This is because the range of temperatures accessi­
ble to study are fairly narrow, ordinarily several hundreds of degrees
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

Celsius. A temperature-dependence of the activation energy would be


expected if the temperature range were broad, because of changes in the
potential function with density. Departures from Arrhenius behavior are
observed, even over short temperature intervals, in glass-forming sys­
by University of Sussex on 01/04/13. For personal use only.

tems where there are pronounced configurational changes. It would


seem that studies of the temperature- and pressure-dependence of
electrical conductance offer little prospect of direct insight into the
fundamental mechanism of the process, in the light of the foregoing;
however, although this is probably true, it is not to say that such
measurements do not have great value. The value of such measurements
lies in the establishment of accurate values, against which the validity of
various approximations of statistical mechanical theories may be judged.

EXPERIMENTAL CONSIDERATIONS

There are primarily two methods available for measuring electrical


conductance: (a) bridge measurements of the resistance and (b) 4-probe
techniques, in which the potential drop is measured between a pair of
electrodes in a constant current established by a second pair of elec­
trodes. The latter technique is widely used in mobility measurements on
metals and semiconductors but has been little used for molten salts.
Treiber & T6dheide (9) employed the 4-probe technique, as well as the
bridge method, to measure the conductance of BiCl3 above its critical
point. This technique is free of possible electrode polarization effects,
and probably should be more widely used.
Far more commonly the conductance is calculated from the imped­
ance, as measured with some form of ac bridge. The apparent resistance
may show a frequency dispersion, attributable to the finite capacitance
of the conductance cell or to reactions occurring at the electrodes. These
may be largely eliminated in careful work; Robbins & Braunstein (10)
discuss precautions to observe. Depending upon the temperature range
and the corrosivity of the salt, cells have been made from Pyrex (II, 12),
quartz (13, 14), sapphire (15), or Lucalox (16), and electrodes from
CONDUCTION IN FUSED SALTS 135

platinum (II, 12, 13), stainless steel (15, 16), or even the melt itself
(9). For high pressure studies it is essential to separate the pressure­
transmitting fluid from the molten salt. For relatively low temperatures
(�300°C) and pressures (� 1 kb), where silicone oil may be used, this
has been accomplished ( 1 7, 18) by means of a well-fitting floating
piston. At higher temperatures and pressures, when argon is employed
in gas pressure systems, the conductance cell is usually provided with a
long capillary section between the gas/salt interface and the electrodes
to delay diffusion of argon into the measurement region. Oeaver &
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

Smedley (19) and Todheide (20) have employed this technique. Buback
& Franck (21) avoided the problem in their measurement of the conduc­
tance of molten NH4Cl from the critical point to the triple point by
using a sealed cell and balancing the known vapor pressure of the salt
by University of Sussex on 01/04/13. For personal use only.

along the vapor/liquid coexistence line with an external pressure of


argon gas.

HIGH PRESSURE STUDIES

In this section we review the manner in which the electrical conduc­


tance of molten salts varies with temperature and the indispensable role
that high pressure studies have played in distinguishing the behavior of
salts that are completely dissociated from those that are molecular in
character. For the latter" a continuous transition from insulator to ionic
conductor occurs as a function of density above the critical tempera­
ture. High pressure studies have also contributed to an understanding of
another kind of transition-the continuous change from the liquid state
to the glassy state for mixtures of salts that may themselves be good
ionic conductors. These studies have been carried out more or less
simultaneously over the past fifteen years and have now attained a
reasonably good level of understanding. Chronologically, the work on
molecular liquids began first, but it is easier to understand them in the
light of the studies on the alkali metal nitrates, which are completely
ionized.
The equivalent conductance of all of the alkali metal nitrates follows
the Arrhenius Law, with activation energies defined by
4.

( E/)p is in the range, 3.21-3.86 kcal mole-I, but there is no systematic


trend with cation radius. Cleaver et al (22) measured the pressure­
dependence to I kb along isotherms up to 450°C. However, the activa­
tion volume, defined by Eq. 5, showed a strong regular increase from
LiN03 to CsN03: 0.5, 3.8, 7.0, 7.8, and 8.3 em3 mole-I, respectively, for
136 NACHTRIEB

LiN03 , NaN03 , KN03, RbN03, and CsN03 at I atm:


�V 1\ = - RT( a In I\/ap)r. 5.
The proper way to assess the effect of temperature on' the electrical
conductance 'is at constant density, rather than at constant pressure. It is
simplest to carry out measurements at constant pressure, and by'means
of the equation of state, to convert them to isochoric conditions. It is
readily shown that the isobaric and isochoric activation energies are
rigorously related:
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

(E1\ ) p= (E1\) V + (Oilf3)(T� V 1\)' 6.


where a and f3 are the isothermal coefficients of expansion and com­
pressibility, respectively. Using values of a and f3 at I atm pressure,
Cleaver et al obtained results for (EI\)v shown in Table lao
by University of Sussex on 01/04/13. For personal use only.

Todheide et al (23) extended these measurements to 5.5 kb for


NaN03 ,KN03 ,RbN03,and CsN03 ,with results that are in substan­
tial agreement with those of Cleaver et al (See Table I b). Although the
values for (E1\)p and �V 1\ agree quite well, there are serious differences
for (£I\)v, the isochoric activation energies. Probably the (£1\) V values
of Todheide et al are more reliable, because 'they used PVT data
measured to 9 kb by Owens (24) (rather than atmospheric pressure
values of a and f3 for the first three nitrates)· and the PVT data of
Bannard (25) to 1 kb for CsN03 •

Table la Activation energies and volume for alkali nitrates a

3
;Salt Ep(ca1 mole·-I) �v1\ (em mole-I) Ev(ca1 mole-I)

LiN03 3400 0.5 3290


NaN03 3210 3.8 2110
KN03 3610 7.0 1680
RbN03 3860 7.8 1640
CsN03 3660 8.3 1040

a After Barton, Oeaver & Hills (22).

Table Ib Activation energies and volume for alkali nitrates a

3
Salt Ep (caI'mole ·-1 ) �V1\ (em mole-I) Ev(ca1 mole-I)

LiN03 3610 0.7 3490


NaN03 3200 3.7 2 100
KN03 3710 6.3 1890
RbN03 3990 7.2 1790
CsN03 3660 8.3 1700

"After Quist, Wurflinger & TOdheide (23).


· CONDUCTION IN FUSED SALTS 137

Schichtharle, Todheide & Franck (26) extended. measurements of the


electrical conductance of the alkali nitrates (Li':"'Rb) to 12 kb in a
piston-cylinder apparatus and observed that the activation energy (both
isobaric and isochoric) still remains temperature-independent; however,
the activation volume, which is independent of pressure at lower pres­
sures,. begins to show a small pressure-dependence. This pressure­
LlV1\ became more pronounced in studies to 55 kb by
dependence. of
Pi1z& Todheide (27). For NaN03, KN03, and RbN03 the activation
volume decreases' with increasing pressure, the effect being more· pro­
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

nounced at a given pressure the larger the radius of the cation (see
Table 2). The effect becomes apparent at intermediate pressures, setting
in at about 20 kb for NaN03, 10 kb for KN03, and 5 kb for RbN03•
Correspondingly, the activation energies, both ( Ef ) v and (EI\)p, in­
by University of Sussex on 01/04/13. For personal use only.

crease with pressure. The sharp change in (EI\)v is centered on the


pressure at which the activation volume decrease is most marked and
suggests that a change in the conduction mechanism is occurring. Only
LiN03 shows no change in activation volume up to 30 kb, although for
this salt, as for all, the isochoric activation energy increases with
increasing pressure, from atmospheric pressure up.
Todheide et al (23) attempted to interpret their data in terms of
various free volume models that have been proposed for transport
properties of liquids. The Cohen-Turnbull (28) model, proposed origi­
nally for the coefficient of self-diffusion, formulates this quantity as

D=Ar,exp ( -k/vf), 7.

where vf is the· free volume per particle (vf= �/No)' and· �= V - Vo'
V is the molar volume; v" is the molar volume at the 'glass transition
temperature, �,. where all particle mobility is "frozen." Under the
Table 2 Activation volumes of alkali nitrates versus pressure a

�v1\ (cm3 mole-I)

P(kbar) LiN03 NaN03 KN03 RbN03 CsN03

0 0.7 3.7 6.4 7.2 8.3


5 3.6 6.3 1.0
10 3.5 5.1 3.5
15 1.4 3.4 5.4 3.4
20 1.4 3.2 3.3
25 2.9
30 1.4 2.4
40 2.0;
50 1.8

a After Pilz & TOdheide (27).


138 NACHTRIEB

assumption that the Nernst-Einstein relation is valid,


I\. =(z2 F2)(D+ +D _ )/RT, 8.
the Cohen-Turnbull relation may be written
I\. =BT- & exp( -k/vf). 9.
The free volume in the Cohen-Turnbull model is given by
vf=o:vm(T - To) - f3vp(P-Po), 10.
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

where vm and vp are the mean volume per particle over the temperature
range T- To and the pressure range P-Po' respectively. Equations 9
and 10 lead to the expression for the logarithm of the molar conduc­
tance:
by University of Sussex on 01/04/13. For personal use only.

II.

The activation volume follows o n differentiating Eq. 1 1 with respect to


pressure at constant temperature:
�Vl\ =RTk{Jvp/[ avm(T- To)-{JvAP-Po) y. 12.

Equation 10 is not in agreement with the experimental observations of


Todheide et aI, because it predicts that at constant temperature the
activation volume should increase with increasing pressure, approach
infinity when the terms in the denominator cancel, and thereafter
decrease with further increase of pressure. In fact, the experimental
results show little pressure-dependence for LlV1\ up to moderately high
pressures and a decrease in the very high pressure region.
TOdheide also examined the Adam & Gibbs configurational entropy
model for its applicability to the high pressure conductance data.
According to Adam & Gibbs (29) the configurational entropy, Sc' of a
system goes to zero at the glass transition temperature, To, where all
transport properties vanish:
I\. =A exp( - B/TSc). 13.
The configurational entropy model describes the entropy by
Sc=LlCpIn(T/To)' 14.
where LlCp is the difference in the heat capacity of the liquid and glassy
system at T= To. Equations 13 and 14 require that
In I\. =ln A - C/[Tln(T/TJ J. 15.
TOdheide et al (23) find that Eq. 15 does not give the required linear
relationship for isobars of In/\ versus (Tln(T/To)-1 when the values of,
Angell & Helphrey (30) for To are used.
CONDUCTION IN FUSED SALTS 139

Similarly, the Adam-Gibbs model fails for the isochoric temperature­


dependence of In 1\, which requires
( E(.)v =( B/S1)(Sc+ T(aSc/aT)v )' 16.
If T=
o
evaluated. With these (reasonable) assumptions:
17.
The independence of (EI\)v on T found experimentally for the alkali
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

nitrates at all pressures then requires


(Cl (E)v/ClT)v=O=BT[ (Cl2SjaT2 )v- (2/SJ(ClSjaT)�]. 18.
This is possible only if (ClSc/aT)v=O, and correspondingly, (ClTo/ap)=
by University of Sussex on 01/04/13. For personal use only.

(To/T)(ap/aT)v' The values of (aTo/ap) evaluated under this conclu­


sion are about 15-20°C kb-1 and are about twice the values found by
Angell et al (31).
The configurational entropy model requires for the isothermal pres­
sure-dependence of the conductance:
�VI\=-(B/Sn(aSc/ap)y. 19.
Because �V1\ is independent of pressure (to at least 20 kb for NaN03 ),
differentiation of Eq. 19 leads to
(a2SjClP2 )T= (2/Sc>(ClSc/ClP)�. 20.
The solution to this differential equation for constant temperature
requires that
SJP)=�Cpln{T/To{P» = (a+bp)-l. 21.
This in tum requires Sc to vanish only at infinite pressure. It thus
appears that free volume models and those that involve the loss of
configurational entropy at the glass transition temperature are unable to
account for the observed high pressure results on alkali metal nitrates.
Although such models are not in accord with experiment, Todheide et al
(23) found that the Furth hole theory of liquids gives predictions that
are in qualitative agreement. Furth (32) proposed that density fluctua­
tions create voids of varying size in the continuum liquid, and he
developed a statistical mechanical theory for the distribution of hole
size. The probability that a hole has a radius between rand r+dr is
given by Furth as
oo
W(r)dr=Cdr exp ( -Eh/kT)dqdp,
Ia 22.

where the integral is over all coordinates of classical phase space and is
140 NACHTRIEB

normalized by the coefficient C. The energy associated with the creation


of a hole is equated with the work done against the surface tension and
against the confining pressure of the liquid:
Eh=47Tr2(J+4/3(7Tr3)(p-po); 23.
and the average volume of the voids is
24 .
In Furth's development, Po is the vapor pressure of the liquid and P is
the external pressure. At the boiling point, P=Po' and the integration of
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

W(r) and vh simplifies, leading to


vh=O.68(kT/(J)3/2 . 25.
by University of Sussex on 01/04/13. For personal use only.

This is the equation usually cited in connection with Furth's model,


even though it is applied to liquids far below their normal boiling
points. There are inconsistencies in the Furth model, including the use
of macroscopic properties such as surface tension and vapor pressure at
the molecular level, of course.
Putting these inconsistencies aside, Todheide found it nevertheless
interesting that Eq. 25 yields values for the average void volume and
radius that are in fair agreement with the experimental activation
volume and ionic radius for NaN03, as shown in Table 3. In the
thermodynamic context of the transition state theory of thermally
activated processes (33), the activation volume is taken to be the sum of
the hole volume �d the ''jump volume": ilV1\ = � + Jj:t._Todheide's
calcula�ons find Vh>ilVI\' so that Ji�l<O. Moreover, (aVh/aT )p>O
and (avh/aph<O, which is reasonable, although they are numerically
too large. Because ilV1\ is positive and nearly independent of pressure
up to 20 kb for NaN03, Jj:J: is a strongly decreasing function of
temperature.
Table 3 Activation parameters for the hole theory (NaNo)a

T("K) P(bar) av/\ (em3 mole-I) Vh b Vh C Jjb Jjc 'h b 'h C aGh b aGh C

(em3 mole-I) (em3 mole-I) (cmxI08) (kcal. mole-I)

623 4.0 8.39 4.07 -4.39 -0.07 1.49 1.17 4.64 3.96
673 3.7 9.67 4.38 -5.97 -0.68 1.57 1.20 5.00 4.23
723 1 3.5 11.1 4.65 -7.6 -1.15 1.64 1.23 5.38 4.48
673 2000 3.7 7.93 4.04 -4.23 -0.34 1.47 1.17 4.76 4.17
673 5000 3.7 6.32 3.63 -2.62 +0.07 1.36 1.13 4.52 4.10

"After Quist, wiirflinger & Todheide (23)


bCalculated from vh=O.68(kTja)3/2
"Calculated from 7kT=3(P-P.)Vh+ 9.06(vh)a3/2
CONDUCTION IN FUSED SALTS 141

The agreement between experiment and the Fiirth model may be


somewhat improved with a few modifications. If the pressure in Eq. 23
is taken to be the internal pressure of the liquid (P is on the order of
lO4 atmospheres) augmented by the external pressure, both terms in Eq.
23 must be included in integrating the phase integral to obtain W(r).
This leads (32) to the approximate expression:

26.
In addition, if the expression for the molar conductance is taken to be
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

of the form (33),


1\ =APexp( -t1GhlRT), 27.
where l:!.G h is the free energy of hole formation, the preexponential is
by University of Sussex on 01/04/13. For personal use only.

frequency-dependent. Logarithmic differentiation of Eq. 27 then leads


to

28.
or to

29.
The last term replaces the activation "jump volume," which is conceptu­
ally vague, by the pressure-dependence of the mean vibrational
frequency. It may be further simplified by means of the Gruneisen
constant (34),
y= -(aln vlaln Vh,

so that

l:!.V/\ = Vh-yPRT. 30.

The value of y is about 1.5 and


P=21.4x lO-6 bar-I, so that for
NaN03 at 673 K, V} should be on the order of
- 1.6 cm3 mole-I.
The results af solving Eq. 26 for Vh, using P;=T(al{3) for the
internal pressure, are shown in Table 3, where they, are compared' with
the results using Eq. 25. The hole volume shows a small increase with
temperature and a small decrease with pressur.�, as. seems reasonable,
and the values are only slightly larger than the ionic radius for Na + •

The activation "jump volume" is on the order of unity and negative, as


expected when regarded as a pressure- or temperature-dependence of
the vibrational frequency.
Just how much credence should be attached to the interpretation of
the conductance data by means of Fiirth's model is difficult to say. It is
in reasonable agreement with Todheide's data for NaN03, which has
142 NACHTRIEB

been the most carefully studied of all salts. Furth's model should be
tested against the other alkali nitrates and extended to the alkali halides.
Both Eq. 25 and 26 are approximations. In principle, the model provides
for a continuous distribution of void radii from zero to infinity and all
possible particle displacements, weighted by a distribution function. The
displacements that occur over distances less than the particle diameter
are included in principle, but are not "jumps" in the sense in which the
hole theory is usually viewed. Rather, they are displacements that occur
because of the highly cooperative motions of many particles. This is
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

implicit, rather than explicit, in the Furth model. In this sense, the
significance of the mean hole radius is that of an ensemble or time
average over the root mean square displacement. Corresponding mean­
ings attach to the other thermodynamic parameters of the model.
by University of Sussex on 01/04/13. For personal use only.

Unfortunately, Eq. 25 ignores the internal pressure of the liquid and is


applicable strictly at the boiling point. Equation 26 makes the ap­
proximation that all voids have the same radius, destroying the most
attractive feature of the model. Such is the fate of all models that
attempt to describe a dynamical many-body process in terms of simple
one- or two-event steps: They fail because of the mathematical ap­
proximations needed to obtain a numerical result.
A recent study by Cleaver (35) reports the results of high pressure
conductance measurements on a variety of alkali metal salts LiCI03,
LiCI04, NaN02, Na2Cr207, KCI03, K 2Cr207, KSCN and AgN03 from
I to 1000 bar. The data appear to be qualitatively in accord with the
trends in activation volume found in T6dheide's work, but the range of
pressure is too narrow to draw quantitative conclusions. A trend toward
decreasing LlVA with increasing temperature is observed, and also to an
increase with increasing cation radius. The influence of the anion on the
magnitude of LlVA follows the sequence: NO; < NO; < CIO; < CI04-
< Cr20;- , qualitatively in the order of ion size.

Conductance of Weak Salts


The conductance of incompletely dissociated salts is considerably more
complex than that of the alkali metal nitrates. At atmospheric pressure,
they do not conform to the Arrhenius Law, but show negative devia­
tions. Aten's (36) measurements in 1909 of the specific conductance of
bismuth trichloride from 250 to 350°C showed distinct curvature when
In /\ was plotted versus 1/ T, but its relatively high molar conductance
2
at the melting point (/\ = 28 ohm -1 cm mole -1) classified it as appre­
ciably ionized. Hevesy (37) found that the specific conductance of HgI2
decreases with temperature above its melting point, and Klemm & Biltz
(38) found similar behavior for InCl3 and InBr3. These three salts were
CONDUCTION IN FUSED SALTS 143

thought to comprise a unique class of anomalous conductance until


Grantham & Yosim (39) extended measurements to higher temperatures
on BiCI3, BiBr3, and BiI3 and observed their molar conductances to
pass through maxima at 460, 460, and 6 10°C, respectively. They conjec­
tured that the cause of the decrease in conductance is due to an increase
in covalent bonding, accompanying the decrease in density with increas­
ing temperature. They speculated that the phenomenon of a negative
temperature coefficient of conductance might not be unusual and fol­
lowed this study with a series of measurements (40) on molten CuI,
HgCI2, HgBr2, SnCI2, In13, Zn12, ZnCI2, TICI, TlBr, TlI, and Hgl2 to
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

temperatures as high as 1200°C. The first six were found to have


conductance maxima (HgI2' like InCl 2 and InBr2' is beyond its con­
ductance maximum at the melting point), and they concluded that this
by University of Sussex on 01/04/13. For personal use only.

behavior is general to all fluids at sufficiently high temperature and


diminished density. Even TlI shows a very slight departure from ex­
ponential temperature-dependence from 448 to 1060°C, and Hafner &
Nachtrieb (41) found downfield chemical shifts for 23sTI in NMR
studies on molten TICI, TlBr, and TlI, which they attributed to increas­
ing covalency with increasing temperature. Levy & Danford (42) showed
that the distance of closest approach decreases in the sequence (crystal
> melt> vapor) for cations and anions of even good salts, reflecting the
trend toward increasing covalent bonding with increase in temperature
and decrease in density. (In the gas phase, even good salts exist as
neutral or polymeric molecules and, of course, have no measurable
conductance.)
Darnell, McCollum & Yosim (43) extended the measurements on
BiCI3, BiBr3' and Bil3 to high pressures and observed that their specific
conductances conformed to the Arrhenius Law for temperatures up to
886°C at a pressure of 5.4 kb, in marked contrast to the nonArrhenius
behavior at atmospheric pressure, with activation energies of 3.68, 4.62,
and 1 1 kcal mole-I, respectively. Grantham & Yosim (44) measured the
specific conductance of BiCl3 and HgCl2 over the entire liquid range,
through the critical temperature, using the density data of Johnson &
Cubicciotti (45) and Cubicciotti et al (46) to calculate their molar
conductances. For both salts the molar conductances along the liquid­
vapor coexistence lines rise from their values at the melting point, pass
through a maximum, and decrease by over an order of magnitude to the
critical point (905°C for BiCI3). Studies by Darnell & McCollum (47)
on HgCl2 and Hgl2 to 20.5 kb show that both behave as strong
electrolytes at elevated pressure, with activation energies of 8.46 and
2.53 kcal mole-I, respectively, at 5.4 kb. All of these studies confirm
Grantham & Yosim's suggestion that salts such as these are largely
144 NACHTRIEB

ionized at high pressures but are molecular liquids at their critical


points.
Cleaver & Smedley (48) measured the specific conductance of HgCI2,
HgBr2' and HgI2 over the pressure range, 1- 1000 bar, in the tempera­
ture interval, 260-370°C for the purpose of deducing their activation
volumes. In contrast to fully ionized salts such as the alkali nitrates,
they found the activation volume to be negative and an order of
magnitude larger in numerical value: �V,,= - RT( a ln o/aP h= -50,
- 65, and - 102 cm3 mole- I, respectively, for HgCI2, HgBr2, and
HgCI2. Their findings were interpreted in terms of autoionization
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

equilibria of the kind


2HgX2=HgX+ + HgX;-,

which have been observed in aqueous solution by Clarke & Woodward


by University of Sussex on 01/04/13. For personal use only.

(49) and by Janz & James (50) in molten mercuric halide/alkali halide
solutions. The volume change accompanying the autoionization reaction
may be expressed by

AVo" = AV:+AV; , 31.


where �V� is
VHgX+ + VHgX, - 2VHgX"
all calculated from bond lengths and van der Waals radii. The value of
AV: is small, on the order of I cm3 mole-I, and negligible by compari­
son with AV;°; �V;0 is attributed to the electrostriction of the solvent
salt, due to the formation 'of solvation sheaths around ions. The stan­
dard partial molar free energy of solvation is calculated from the Born
equation,

�Go = -Ne2z2/2r(I-I/e), 32.


whose pressure derivative is just At:;° :
�V2 =(a�Go /ap)T= L [Ne2z2/2r2(1-I/e)(ar/aph
ions

33.

The derivatives in Eq. 33 were estimated from isothermal compressihili­


ties and the Clausis-Mosotti equation. Using several methods to esti­
mate the radii of the ionic species, Cleaver & Smedley found values for
. A�r that- are in reasonably good agreement with the experimentally
obtained values of 2AV". Their large negative magnitudes show that
-pressures in the range of 5- 10 kb can be expected to shift the autoion­
ization equilibria strongly toward complete ionization. On the basis of
their analysis, Cleaver & Smedley propose a simple relationship between
CONDUCTION IN FUSED SALTS 145

the activation volume and compressibility:


o ln { -!!J.V,J/oT=olnPTloT. 34.
3
For HgBrz these coefficients are 3.3x 10- K-1 and 2.7x 10- K-1, 3
respectively.
Treiber & Todheide (51) extended conductance studies on weak
electrolyte salts to 1200°C and 4 kb in further measurements on BiCI3•
They confirmed the earlier studies of Grantham & Yosim and of ;
Darnell et aI, concluding that the salt is less than I % ionized at its
critical point. Assuming, as in aqueous solutions, that the principal ionic
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

equilibrium is
2BiCl3 =BiCI! + BiCI; ,
they set an upper limit on the molar conductance for the fully ionized
by University of Sussex on 01/04/13. For personal use only.

salt from measured diffusion coefficients by means of the Nernst­


Einstein relation:
35.
With this as a guide, they selected the standard state for complete
ionization to be 400° C and 5.4 kb, where /\ =55 ohm -I cm2 mole -I.
Relative to this, the degree of ionization at other pressures and tempera­
tures is given by a = /\11\0. This assumes that the effects of tempera­
ture and pressure on the mobilities of completely ionized salts are small
compared with their effect upon the autoionization equilibrium.:..:....an
assumption fully borne out by their extensive studies on the pressure
and temperature effects on alkali nitrates. Their measurements of the
degree of ionization of BiCl3 ranged over more than five orders of
magnitude; (J increased very strongly with increasing density at constant
temperature and only weakly with rising temperature at constant den­
sity. Having a, they calculated ionization constants that ranged from 00
(a= 1) at 400°C and 5.4 kb to 5.3x 10-10 at lOOO°C and a density of 0.5
g-cm-3 (corresponding to 60 kb). Their studies, culminating the earlier
work and remarkable insight of Grantham & Yosim, make clear that
there is a continuous transition from covalent to ionic bonding in
"weak" salts that is a function of density. This transition is believed to
be very general, as the researches of Holzapfel & Franck (52) on the
dissociation of water and other polar liquids have brilliantly demon­
strated.
It is of much interest to know how general is this density-governed
transition from the covalent to ionic state. Ionic conductance is the
most direct experimental approach to this question. Poorly ionizing salts
have unambiguously been shown to be continuously transformable into
strongly dissociating salts. Whether strong electrolytes may be trans­
formed continuously into insulators by sufficient reduction in density
146 NACHTRIEB

requires measurements that approach or exceed their critical points to


answer. Possible salts for such studies are very limited. The alkali
halides would be the best candidates, but McGonigal (53) has estimated
their critical points to lie above 3000 K, and alkali nitrates and other
oxy-salts decompose at much lower temperatures. Hensel (54) has
unequivocally shown the metal-insulator transition in fluid mercury
above its critical temperature, but even among metals there are few that
have critical points within the range of laboratory study.
Buback & Franck (55) selected NH4Cl as a strong electrolyte salt
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

having a critical temperature (Tc = 882°C) and critical pressure (Pc =


1685 kb) within access of laboratory study. They measured the specific
conductance from the triple point (520°C), where the specific conduc­
tance is 2.07 ohm-I cm-I to 850°C (<7=3.18 ohm-I cm-I) along the
by University of Sussex on 01/04/13. For personal use only.

liquid-vapor coexistence line and found no deviation from that expected


for complete ionic dissociation. Because the data are not isochoric, the
density-dependence was determined from PVT data and the molar
conductance was expressed by

36.

the form of which best fits the density-dependent conductance of


molten sodium chloride. Even though these measurements extend to
within 32°C of the triple point (corresponding to T/ T.: = 0.97), the molar
conductance fits Eq. 36 to within experimental error, and Bruback &
Franck concluded that even at its critical point molten NH40 has too
high a density to induce the ionic/covalent transition. This increases the
interest in studies on the other alkali halides to temperatures in their
liquid range where the transition might be observed.

Ionic Salt-Glass Transition


Salts of alkali and alkaline earth metal cations with the network-forming
anions (silicates, borates, and phosphates) and melts of alkali halides in
salts such as BeF2 or ZnCl2 show pronounced deviations from Arrhenius
behavior in their electrical conductance and other transport properties.
In addition, the chemical composition plays a profound role and is not
understood in fundamental terms. Moynihan (56) has discussed the
"mixed alkali effect" in glass-forming melts. This effect refers to the
nonadditivity of conductance or diffusivity of monovalent cations in
anion network melts, such as are shown by the system XNa20-(l­
X)K20-3Si02• The usual equilibrium properties of the system (e.g.
molar volume, heat capacity, elastic moduli, etc) show little deviation
from additivity, but the electrical conductance exhibits a temperature­
dependent minimum as deep as five orders of magnitude below the
CONDUCTION IN FUSED SALTS 147

expected additivity line. The effect is much more pronounced the lower
the temperature becomes, and is usually characterized by the parameter,
'Ta = (J
add/ (J or T/\ = /\add/ /\, where (Jadd and /\add denote the specific
and molar conductances calculated as the sum of the conductances of
the pure components weighted by mole fraction, and (J and /\ are the
observed values. The value of T may range from unity to very large
numbers; it is never less than unity, although it approaches this value
for glass-forming molten salts at high temperatures, and infinity at
lower temperatures as the vitreous state sets in. Theories of the monova­
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

lent cation effect are in a qualitative state, and fall into two classes: (a)
those that propose attractive interactions between different cations that
lead to a decrease in their mobilities [Laity & Moynihan (57, 58),
Mazurin (59), Hendrickson & Bray (60)] and (b) structure/mechanistic
by University of Sussex on 01/04/13. For personal use only.

theories [Ravaine & Soquet (61) and Tomozawa (62)] that treat network
glasses as weak electrolyte salts, whose dissociation constants are strong
functions of ion concentrations. Ionic interactions may account for the
relatively small alkali cation effect in the induced glass-forming salts
such as Ca(N03h mixed alkali nitrate melts, but not in network-forming
melts.
Theories that attempt to account for the nonArrhenius behavior of
transport properties in viscous fluids and glasses are based upon the
concept that these processes cease at the glass transition temperature,
To. The Cohen-Turnbull model (28) emphasizes the free volume of a
system and its disappearance at the glass transition temperature (see Eq.
7-12). A density fluctuation that expands a void to some critical volume
and permits an ion to ''jump'' therein is the underlying concept of the
elementary transport act. It fits the data of Angell, Pollard & Strauss
(63) on Ca(N03h-KN03 melts quite well to 3 kb, but like all hole
models, it has been criticized on grounds that hard sphere potential
molecular dynamics calculations [Alder & Einwohner (64)] provide no
evidence for elementary displacements of molecular dimensions. The
results of Angell et al fit the Adam & Gibbs (29) configuration entropy
model very well if To is assumed to have a linear pressure dependence of
8 K kb I. Macedo et al (65, 66) have recently developed an environ­
-

mental relaxation model that ascribes the nonArrhenius behavior of


transport processes to the temperature-dependent distribution of char­
acteristic relaxation times. Islam et al (67) have compared the free­
volume, configuration entropy, and environmental relaxation models for
their applicability to melts of MnCl2 with tetra-n-butylammonium
halides and hydrous melts of pairs of the salts, Zn(N03h, Cu(N03h,
Cd(N03b NiCl2, and CoCI2• The environmental relaxation model
best fits the temperature- and concentration-dependence of these melts,
but it is a 5-parameter relation to /\!
148 NACHTRIEB

Metal-Molten Salt Solutions


Another kind of transition to be seen in certain molten salt solutions is
that between the ionic and metallic states. The true solution nature of
molten mixtures of the alkali metals with their alkali h alides was
discovered by Bredig, Johnson & Smith (68) in 1955, when they pub­
lished the phase diagram of the Na-NaCI system. In this and subsequent
determinations of the phase diagrams of all alkali metal-alkali halide
systems, B redig et al (69- 74) found mutual solubility of salt and metal
in all systems and determined the consolute temperatures, above which
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

miscibility is complete. They m easured the specific conductance in the


salt-rich regions of the Na-NaCl, Na-NaBr, K-KCl, and K-KBr sys­
tems, showing unambiguously the presence of a high mobility carrier
which they recognized as electronic. Subsequently, Dworkin, Bronstein
by University of Sussex on 01/04/13. For personal use only.

& Bredig (75) extended studies to the salt-rich region of the solutions of
Ca, Sr, La, and Nd metals in their molten chlorides, finding electronic
conduction in each. They (76, 77) then measured the specific conduc­
tance at 700°C of liquid potassium metal containing dissolved KF, KBr,
and KI-in the case of KBr at 740°C. The consolute temperature of the
K-KBr system is 728°C, and for this system the course of the specific
conductanc , e
salt-rich region all solutions b ehave as semiconductors, whereas in the
metallic region the dissolved salt behaves as an impurity that lowers
the metallic conductance. Above the consolute temperature for the
K-KBr system there is a continuous transition from ionic to semicon­
ducting to metallic conductance. Hsu & Nachtrieb (78, 79 ) have mea­
sured the specific conductance in the Cs-CsCI system, for which there is
complete miscibility; for this system, also, the transition from semicon­
ductor to metallic is continuous, with a broad range of concentrations
centered at 40 mole % Cs, where the temperature coefficient of conduc­
tance changes from positive to negative.
Heus & Egan (80, 81) used an electrochemical polarization method to
demonstrate and measure both electron and hole conduction in the
LiCI-KCI eutectic and in NaCl, using cells of the type

TaILiCI-KCl(eut) / PbCli5% wt)


LiCI-KCl(eut)
/ Pb

for the first system. Under conditions that eliminated convection, they
measured the electronic or hole current as a function of the applied
EMF, which is given by
i=RT/ FG { o;ef[ exp{EF/RT)-l] +Ohef[ l-exp ( - EF/RT)]}. 37.
where E is the applied EMF (volts ), F is the Faraday, G is the cell
CONDUCTION IN FUSED SALTS 149

constant, and o;ef and oi,ef are, respectively, the electron and hole
specific conductivities in the melt equilibrated with the reference elec­
e
trode. When E» RT/ F, the second term of Eq. 37 is negligible, and (J; j
is obtained as the slope of i versus E plots. Similarly, when E» RT/F,
-

(Ji,ef is obtained. In either case, the applied voltage is less than the
decomposition potential of the electrolyte. The electron-hole activities
(mole fractions) follow the mass action law (ae X ah =constant),
and in NaCI at 830°C are equal when ae=ah= 10-8 and 0e=(Jh=
10-8 ohm 1 cm -1. Along the solubility line at 830°C, where aNa= I
-
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

and XNa=0.032, Bedig's results and those of Heus & Egan for (J agree
very well. This ingenious method should be more widely used.
There are no very satisfactory theories for the electron mobility in
metal-molten salt solutions that span the full range of the phase dia­
by University of Sussex on 01/04/13. For personal use only.

gram. Probably Ziman's (82) theory of liquid metals is the best ap­
proach, where the h alide ion is treated as an impurity-scattering center,
but it fails [Wilson (83)] to give the correct sign of the temperature
coefficient of conductance in the dilute metal region because of the
extreme approximations that must be made. Several models that treat
the electrostatic interaction of the halide ion with the metal ion cores in
liquid alkali metals have been proposed [Gellings et al (84, 85) and
Thompson (86)], but there have been no applications to transport
properties.
The dilute metal region is probably the most interesting and theoreti­
cally the most difficult. At very low concentrations the electron is
trapped, possibly in an F-center-like state as Pitzer (87) has proposed,
and conductance may be accomplished by a "hopping" mechanism
[Rice (88)] from one localized site to another. As the electron concentra­
tion is increased, the tight-binding approximation is no longer applica­
ble. It must ultimately be replaced by a band description, but the
intermediate states are still obscure. In this region the theories of the
electronic states of disordered systems proposed by Mott (89), Cohen
(90) and others may point the way. In these theories the one-electron
bands are broadened, and electrons in the low energy tails of the band
are localized. Their contribution to conduction is thought to occur by a
thermally-activated hopping process. As the electron concentration is
increased, the bands further broaden and eventually overlap, and the
localized electrons become delocalized (metallic conduction) in the
empty band.
This review would be incomplete without reference to the interesting
studies of the Marburg group, where Hensel et al (91, 92) have shown
that a deep minimum occurs in the specific conductance in the liquid
Cs-Au metallic system. At the stoichiometric composition of CsAu, the
1 50 NACHTRIEB

specific conductance drops over three orders of magnitude from its


value for liquid cesium to 3 ohm J cm- J at 600°C. The alloy CsAu at
-

this composition is an ionic salt (Cs+ Au-), and these studies show the
existence of the nonmetal-molten salt transition in metallic alloy sys­
tems.

MOLECULAR DYNAMICS CALCULATIONS


One of the most important developments in the chemistry and physics
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

of molten salts during the past decade has been the use of computer­
based "experiments" for determining the equilibrium and transport
properties of dense fluids. Their value lies not so much in the quantita­
tive agreement with "real" experiments, although this is necessary and
by University of Sussex on 01/04/13. For personal use only.

good for many systems, as in the insight they provide into the local
structure of ionic melts and the time evolution of the fluctuations
around the local configurations of ions in real space and in velocity
space. Progressing from the introduction of the Monte Carlo method by
Metropolis (93) through the studies of hard sphere dense fluids by Alder
et al (94) and the definitive work by Rahman (95) on liquid argon, we
have witnessed the extension of computer-simUlation experiments in
recent years to one-component plasmas by McDonald et al (96) and to
molten salts by Singer (97), Woodcock & Singer (98), Lewis & Singer
(99), Sundheim (100), and others. Because of the long-range character of
the Coulomb interactions which dominate the potential in molten salts
and the multipoles induced by ions, this extension to molten salts
represents a major advance in the reduction of a many-body problem to
one of tractable dimensions.
Two general complementary approaches are possible: the Monte
Carlo method, in which ensemble averages over eqUilibrium configura­
tions of a small system of ions are taken; and the molecular dynamics
approach, in which the classical equations of motion are numerically
integrated over small time-steps. The two approaches are related through
the fundamental postulate of statistical mechanics that equates the
ensemble average of a dynamical variable of a conservative system with
its long-time average. Both procedures lead, in principle, to the same
results for the equilibrium properties of a system, but the evolution of
the time-dependent properties is necessarily inaccessible to the Monte
Carlo method. Molecular dynamics yields both equilibrium properties
and linear transport coefficients and receives the greater emphasis in
this review. Both methods require an acceptably accurate potential
function as the primary input, and the accuracy with which various
macroscopic properties are calculated depends upon its form in detail.
CONDUCTION IN FUSED SALTS 151

The dynamical state of the i th particle in a system of N particles is


given at each instant by
m i ( d 2 rJdt2 ) = Fi = - d<P(rpr2 • • • rN ) /dr, 3S.
where ri is the position vector of the ith particle, F; is the net force
acting upon it, and <P(rl ' r2 " . rN ) is the many-body potential of the
system at the particle site. Because there are 3 Cartesian coordinates, the
problem is to solve 3 N-coupled second-order differential equations
simultaneously. The cost of computation and the limitations on speed
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

and memory of even very large machines restricts the size of the system
to about N = 103 ions, and more commonly it consists of l OS cations
and lOS anions ( N = 6 3 ). The ions are confined to a cube having
periodic boundaries that restore escaping particles through an opposite
by University of Sussex on 01/04/13. For personal use only.

face, and the duration of an experiment is on the order of 10- \ 2 s, with


elementary time intervals of about 10- 14 s. The forces, velocities, and
positions of each particle are calculated by means of finite difference
algorithms based on Eq. 38 and refined by predictor-corrector cycles to
arbitrary accuracy before storage in memory for later processing. At the
start of an experiment the system is allowed to run through a number of
preliminary steps to establish thermal equilibrium, and when this is
attained the particle energies are scaled to put the system at some
predetermined temperature.
The ideal many-body potential is a hypersurface of N dimensions,
which may be represented as a sum of contributions from 2-body,
3-body, . . . N-body interactions:
N N N N N
<P(rl > r , , · rN ) = t L L <Pij / ! L L L <Pijk
+ 1 3
2 j
joFi

N N N
+ " . l/p! L L L 39.
j P
joFi p oF i , j , ... k
Such a generalized form for <I>(r\ , r " . rN) is mathematically cumber­
2
some, although it could be extracted from the virial form of the
equation of state. The usual practice, however, is to define an effective
pair potential, <Pejj ( rij, p, T ), fitted by means of the equation of state of
dense gases or crystals to a potential function of Born-Mayer-Huggins
form:
<Pejj= QIQ2irij + BCijexp {( ai+aj- fiJ/A }
40.
152 NACHTRIEB

The approximation to the many-body potential is then

�(rl ' r2 · · · rN ' p , T ) = t ( � * fPeff(rij' T) )


p, , 41 .

N= i

where the brackets denote either a time or ensemble>average, The terms


in fPe!! correspond to Coulomb, overlap repulsion, ion-induced dipole,
and induced dipole-induced dipole interactions. The parameters usually
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

chosen for alkali halides are those of Tosi & Fumi ( 101), who obtained
them as a best empirical fit from the equations of state of the rock salt
structure alkali halides. The equilibrium properties of the molten salts,
such as the internal energy at the melting point, heat of fusion, and pair
by University of Sussex on 01/04/13. For personal use only.

distribution function, are in generally good agreement with experiment,


although there appear to be small systematic underestimates of the
positions ,of the first and second maxima in the radial distribution
function from both MC and MD calculations, as compared with X-ray .
and neutron diffraction experiments.
The number of molecular dynamics calculations yielding the linear
coefficients of transport properties is much smaller. They depend upon
the time averages of various autocorrelation functions of dynamical
variables of equilibrium liquids: velocity, force, current, kinetic energy,
or shear tensor for diffusion, electrical conductance, thermal conduc­
tance, or viscosity. Their origin lies in Einstein's theory of Brownian
motion, and many subsequent contributors to the regression-fluctuation
theories of liquids. Zwanzig's ( 1 02) review of statistical mechanics
provides an excellent summary of the contributions of Born and Green,
Kirkwood, Kubo, and Callen, among many others, to transport processes
in liquids� The self-diffusion coefficient is given by the time integral
over the velocity autocorrelation function

D = I /3 itdT« v(O)' V( T » , 42 .
o

and the specific conductance as the corresponding time integral over the
current autocorrelation function

43 .

Because a molten salt contains both positive ions (i) and negative ions ..
( j ), the current autocorrelation function contains both like ion- and
cross-correlation ternlS. Thus, the current autocorrelation function be-
CONDUCTION IN FUSED SALTS , ,153

comes

« /(0)·/( T ) > = e2 ( ( ; Z; V;(O) - 7 zj v/O))


x(;z;v;(T) - 7 Zj V/ T )) ). 44.

Hansen & McDonald ( 103) carried out MD calculations on a system


of 2 16 particles that 'simulated molten NaCI at 1267 K with a molar
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

volume of 41.9 cm3• They found 0 = 3.2 ohm- I em- I, in reasonably


good agreement with the experimental value at this temperature, 3.7
ohm- I em - I .
Combining Eq. 42 and 43 leads to the Nemst-Einstein relation,
by University of Sussex on 01/04/13. For personal use only.

modified by the factor ( 1 - 6 ), where 6 arises from the cross-correlation


term and the factor i because N refers to the total number of ions,
rather than the number of "molecules" :
45 .
Hansen & McDonald found the value a = 0 . 19, compared with -exp�ri­
mental values 0. 15 for RbCI, 0. 18 for NaCl; - and 0.23 : for Cstl. It is
interesting that 6 is a positive quantity, because this means that there
are ionic motions that contribute to diffusion but do not contribute to
charge transport.' This could be the case if, at high temperatures, there is
a tendency toward the' formation of neutral ion pairs or clusters.
Indeed, - this is the picture of the local structure Woodcock ( 104)
"observed" in a model constructed from molecular dynamics calcula­
tions: on molten KCI. There is a distinct tendency toward ion pairs and
the 'formation of clusters aHhe melting point ( 1066 K). There are also
: large voids in the model · structure, which at first sight might be thought
to supporfthe hole theory of liquids. But the time evolution of the local
structure gives no evidence of single ion ')umps" into these voids;
rather, there is a highly cooperative motion of many particles, pairs, and
clusters. The observation of ion pairs and clusters for these salts is
completely in accord with the brilliant suggestion of Grantham &
yosim (39) that all salts might show a tendency toward eovalent
bonding when the temperature is sufficiently high or' the density suffi­
ciently low.
Lewis & Singer (99) carried out a molecular dynamics study of
molten NaCI from 1000 to 1738 K (bp), and obtained remarkably good
agreement for the thermodynamic quantities Vm , U p , PT' and Cv llSing
the Tosi-Fumi parameters. The specific conductance was not evaluated,
15 4 NACHTRIEB

Table 4 Diffusion coefficients in molten NaCI

T ("K) Vm (cm3 mole - I ) D+ X lOs cm2 sec- I D_ X lOs cm2 sec- I

Calc." Exp.b Calc." Exp .b

1073 39. 1 4 8.1 7.3 7.5 5.8


1 100 39.5 1 8.5 8.0 7.9 6.3
1200 40.88 10.2 10.5 9.4 8.4
1 300 42.27 12.2 13.2 1 1.3 10.7
1 400 43.70 14.6 16.1 1 3.5 13.1
1500 45. 1 7 17.4 19.1 16.1
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

15.7
1600 46.71 20.6 19.1
1 738 49.00 26.0 24.0

• Molecular dynamics calculations after Lewis & Singer (99).


by University of Sussex on 01/04/13. For personal use only.

b
Experimentai data after Bockris. Richards & Nanis (105).

although it could have been obtained from the current autocorrelation


function as described above. Instead, they calculated the self-diffusion
coefficients for Na+ and Cl- by means of Eq. 42 or its equivalent:
D; = lim <I r; ( t ) - r; (OW > . 46 .
t->oo

Table 4 shows a comparison with the experimental results reported by


Bockris, Richards & Nanis (105). The agreement is excellent-certainly
within the limits of experimental error.

Literature Cited

1. Delimarskii, I. K., Markov, B. F. Bunsenges. Phys. Chern . 77: 540


1961. Electrochemistry of Fused Salts, 10. Robbins, G. D., Braunstein, J. 1 969.
Trans!. A. Peiper!. Washington DC: See Ref. 6, pp. 443- 77
Sigma 1 1. DeNoojier, B. 1 965. PhD thesis.
2. Sundheim, B. R. 1 964. Fused Salts. Univ. Amsterdam, The Netherlands
New York: McGraw-Hill 12. Angell, C. A., Pollard, L. J., Strauss,
3. Blander, M. 1964. Molten Salt Chem­ W. 1 965. J. Chem . Phys. 43:2899
istry . New York: Interscience 1 3. Buback, M., Franck, E. U . 1 973. Ber.
4. Lumsden, J. 1966. Thermodynamics Bunsenges. Phys. Chern. 77: 1074
of Molten Salt Mixtures. New York: 1 4. Jaffe, J., van Artsdalen, E. 1956. J.
Academic Phys. Chem . 60:8, 1 1 25
5. Janz, G. J. 1 967. Molten Salts 15. Bronstein, H. R., Bredig, M. A. 1 958.
Handbook. New York: Academic J. A m . Chern. Soc. 80:2077
6. Mamantov, G. 1969. Molten Salts : 16. Hsu, c., Nachtrieb, N. H. 1980. J.
Characterization and Analysis. New Phys. Chern . In press
York: Dekker 1 7. Hamann, S. D., Strauss, W. 1 955.
7. Pemsler, J. P., Brau:nstein, J., Nobe, Trans. Faraday Soc. 5 1 : 1684
K., Morris, D. R., Richards, N. E. 18. Ellis, A. J. 1 959. J. Chem . Soc. 2299
1976. Proc. Int. Symp . Molten Salts. 1 9. Cleaver, B., Smedley, S. I. 1 970.
Princeton, NJ: Electrochemical Soc. Trans. Faraday Soc. 66: 1 1 1 5
8. Braunstein, J., Mamantov, G., Smith, 20. Todheide, K. 1 976. See Ref. 7, p. 20
G. P. 1971, 1973, 1 975. A dvances in 21. Buback, M., Franck, E. U. 1 973. Ber.
Molten Salt Chemistry, Vols. 1-3. Bunsenges. Phys. Chern. 77: 1074
New York: Plenum 22. Barton, A. F. M., Cleaver, B., Hills,
9. Treiber, G., Todheide, K. 1 973. Ber. G. J. 1 967. Trans. Faraday Soc.
CONDUCTION IN FUSED SALTS 155

63:208 1 97 1 . J. Chern . Phys. 55: 1 16


23. Quist, A. S., Wurflinger, A., 48. Cleaver, B., Smedley, S. I. 1 970.
Todheide, K. 1 972. Ber. Bunsenges. Trans. Farad� Soc. 66: 1 1 15
Phys. Chern . 76:652 49. Clarke, J. H. R., Woodward, L. A.
24. Owens, B. B. 1 966. J. Chern . Phys. 1965. Trans. Faraday Soc. 6 1 :207
44: 39 1 & 50. Janz, G. J., James, D. W. 1963. J.
25. Bannard, J . E . 1969. PhD thesis. Chern . Phys. 38:905
Univ. Southampton, England 51. Treiber, G., Todheide, K. 1 973. Ber.
26. Schichtharle, G., Todheide, K., Bunsenges. Phys. Chern . 77: 540
Franck, E. U. 1972. Ber. Bunsenges. 52. Holzapfel, W., Franck, E. U. 1 969.
Phys. Chern . 76: 1 168 Ber. Bunsenges. Phys. Chern . 70: 1 105
27. Pilz, Y., Todheide, K. 1 973. Ber. 53. McGonigal, P. J. 1963. J. Phys.
Bunsenges. Phys. Chern . 77:540 Chem . 67: 1931
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

28. Cohen, M. H., Turnbull, D. 1959 J. 54. Hensel, F. 1973. Electrons in Fluids,
Chern . Phys. 3 1 : 1 164 ed. J. Jortner, p. 355. Berlin: Springer
29. Adam, G., Gibbs, J. H. 1965. J. 55. Buback, M., Franck, E. U. 1973. Ber.
Chern . Phys. 43: 139 Bunsenges. Phys. Chern . 77: 1074
30. Angell, C. A., Helphrey, D. B. 1971. 56. Moynihan, C. T. 1979. J. Electro­
J. Chern . Phys. 75:230
by University of Sussex on 01/04/13. For personal use only.

chern . Soc. 126:2144


31. Angell, C. A., Pollard, L. J., Strauss, 57. Moynihan, C. T., Laity, R. W. 1 964.
W. 1 969. J. Chern . Phys. 50:2694 J. Phys. Chern . 68 :3312
32. FUrth, R. 1941. Proc. Carnbridge 58. Moynihan, C. T. 197 1 . Ionic Interac­
Phi/os. Soc . 37:252 tions, ed. S. Petrucci, 1 :26 1 . New
33. Glasstone, S., Laidler, K. J., Eyring, York: Academic
H . 1941. The Theory of Rate 59. Mazurin, O. Y. 1 965. The Structure
Processes, p. 470. New York: Mc­ of Glass, 4:5. New York: Con­
Graw-Hill sultants Bureau
34. See any text on solid state physics, 60. Hendrickson, J. R., Bray, P. J. 1 972.
e.g. Slater, J. C. 1939. Introduction to J. Phys. Chern . Glasses 1 3 : 107
Chernical Physics, p. 2 1 7. New York: 61. Ravaine, D., Soquet, J. L. 1977. J.
McGraw Hill Phys. Chern . Glasses 1 8 : 27
35. Cleaver, B. 1978. J. Chern . Soc. 62. Tomozawa, M. 1 977. Treatise on
Faraday Trans. II 77:2735 Materials Science and Technology, ed.
36. Aten, A. H. W. 1909. Z. Phys. Chern . M. Tomozawa, R. W. Doremus,
66:641 12:283. New York: Academic
37. Hevesy, G. Y. 192 1 . Kgl. Danske 63. Angell, C. A., Pollard, L. J., Strauss,
Videnskab. Selskab. Mat. Fys. Medd. W. 1969. J. Chern . Phys. 50:2694
3: 1 3 64. Alder, B. J., Einwohner, T. 1965. J.
38. Kle=, W., Biltz, W . 1926. Z. A norg. Chem . Phys. 43:3399
AI/g. Chern . 152:225 65. Si=ons, J. H., Macedo, P. B. 1 97 1 .
39. Grantham, L. P., Yosim, S. 1963. J. J. Chem . Phys. 54: 1 325
Phys. Chern . 67:2506 66. Howard, R., Simmons, J. H.,
40. Grantham, L. F., Yosim, S. 1966. J. Macedo, P. B. 1976. J; Chern . Phys.
Chern . Phys. 45 : 1 192 64:439
41. Hafner, S., Nachtrieb, N. H. 1964. J. 67. Islam, N., Kumar, S., Singh, K. P.
Chern . Phys. 40:2891 1979. J. Chem . Soc. Faraday Trans I
42. Levy, H. A., Danford, M. D. 1 964. 75 : 1830
Molten Salts, ed. M. Blander, p. 109. 68. Bredig, M., Johnson, J. W., Smith,
New York: Interscience W. T. Jr. 1955. J. Am. Chem . Soc.
43. Darnell, A. J., McCollum, W. A., 77:307
Yosim, S. 1 969. J. Phys. Chern. 69. Bredig, M., Bronstein, H. R., Smith,
73:41 16 W. T. Jr. 1955. J. Am. Chem . Soc .
44. Grantham, L. F., Yosim, S. 1968. J. 77: 1454
Phys. Chern. 72: 762 70. Bronstein, H. R., Bredig, M. 1958. J.
45. Johnson, J. W., Cubicciotti, D. 1 964. A rn . Chern . Soc. 80:2077
J. Phys. Chern . 68:2235 71. Bredig, M., Bronstein, H. R. 1960. J.
46. Johnson, J. W., Silva, W. J., Cubic­ Phys. Chern . 64: 64
ciotti, D. 1966. J. Phys. Chern . 72. Bredig, M., Johnson, J. W. 1960. J.
70: 1 1 69 Phys. Chern . 64: 1 899
47. Darnell, A. J., McCollum, W. A. 73. Bronstein, H. R., Bredig, M. 196 1 . J.
156 NACHTRIEB

Phys. Chem. 65: 1220 90. Cohen, M. H. 1970. J. Non-Cryst.


74. Bredig, M. 1964. Molten Salt Chem­ Solids 4:391
istry , ed. M. Blander, p. 367. New 91. Hoshino, H., Schmutzler, R. W.,
York: Wiley Hensel, F. 1975. Phys. Lett. A : 5 1 :7
75. Dworkin, A. S., Bronstein, H. It, 92. Schmutzler, R. W., Hoshino, H.,
Bredig, M. A. 1961. Discuss. Faraday Fischer, R., Hensel, F. 1 976. Ber.
Soc. 32:188 Bunsenges. Phys. Chem. 80: 107
76. Bronstein, H. R., Dworkin, A. S., 93. Metropolis, N., Rosenbluth, A. W.,
Bredig, M. A. 1961. J. Chern. Phys. Rosenbluth, M. N., Teller, A. H.,
34: 1 843 Teller, E. 1953. J. Chem . Phys. 2 1 :87
77. Bronstein, H. R., Dworkin, A. S., 94. Alder, B. J., Gass, D. M., Wain­
Bredig, M. A. 1962. J. Chem. Phys. wright, T. E. 1970. J. Chem. Phys.
37:677 53:3813
Annu. Rev. Phys. Chem. 1980.31:131-156. Downloaded from www.annualreviews.org

78. Hsu, C., Nachtrieb, N. H. 1980. J. 95. Rahman, A. 1964. Phys. Rev. A
Phys. Chem. In press 136:405
79. Hsu, c., Nachtrieb, N. H. 1980. J. 96. Ciccotti, G., Jacucci, G., McDonald,
Phys. Chem. In press I. R. 1976. Phys. Rev. A . 13:426
80. Heus, R. J., Egan, J. J. 1973. J. Phys. 97. Singer, K., McDonald, I. R. 1970. Q.
by University of Sussex on 01/04/13. For personal use only.

Chem . 77: 1989 Rev. Chem . Soc. 24:238


81. Heus, R . J., Egan, J . J . 1976. See 98. Woodcock, L. V., Singer, K. 197 1 .
Ref. 7, p. 523 Trans. Faraday Soc. 67: 12
82. Ziman, J. 1961. Phi/os. Mag. 6: 1013 99. Lewis, J. W. E., Singer, K. 1974. J.
83. Wilson, E. G. 1963. Phys. Rev. Lett. Chem. Soc. Faraday Trans. 1/ 7 1 :41
10:432 1 00. Sundheim, B. R. 1979. Chem. Phys.
84. Gellings, P. J., Huiskamp, G. B., van Lett. 60:427
Broek, E. G. 1972. J. Chem. Soc. A , 101. Tosi, M. P., Fumi, F. G. 1964. J.
p . 151 Phys. Chem. Solids 25:45
85. Gellings, P. J., van der Scheer, A., 102. Zwanzig, R. 1965. A nn . Rev. Phys.
. Caspers, W. J. 1973. J. Chem. Soc. Chem . 16:67
A, p. 531 103. Hansen, J. P., McDonald, I. R. 1975.
86. Thompson, R. 1972. J. lnorg. Nucl. Phys. Rev. A 1 1 :2 1 1
Chem . 34:2513 1 04. Woodcock, L. V. 1971. Nature Phys.
87. 1962 J. Am. Chem . Soc.
Pitzer, K. S. Sci. 232:63
84:2025 105. Bockris, J. O'M., Richards, S. R.,
88. Rice, S. A. 1961. Discuss. Faraday Nanis, L. 1965. J. . Phys. Chem.
Soc. 32: 181 69:627
89. Mott, N. F. 1967. Ado. Phys. 16:49

You might also like