D.C. Conductivity and Dielectric Properties in Silver Chloride, Revisited Joachim Maier†

David S. Mebane∗

September 28, 2009

Abstract The anomolous D.C. ionic conductivity of silver chloride at high temperatures is analyzed quantitatively with the aid of mean field models and classical Monte Carlo simulations. It is found that the conductivity anomaly in silver chloride along with the exponential increase of the dielectric constant at high temperatures may be explained by referring to the idea of a “soft lattice,” wherein silver ions are increasingly free to move about their regular lattice sites as the temperature rises. This wide ranging of silver ions leads to the formation of transient dipoles, which have the dual effect of increasing the dielectric constant and decreasing the free energy of the lattice.

Introduction The temperature dependence of the ionic conductivity in silver halides is well described by an intrinsic high-temperature regime characterized by Frenkel disorder (see Fig. 1, regime 3) and an extrinsic regime characterized by a dopant-compensated carrier concentration (see Fig. 1, regimes 1 and 2), with a certain degree of defect-dopant association which becomes increasingly important as the temperature falls (see Fig. 1, ∗ Max-Planck-Institut für Festkörperforschung, Stuttgart, DE. Support provided by the U.S. National Science Foundation International Research Fellowship Program, Grant No. 0701145. † Max-Planck-Institut für Festkörperforschung, Stuttgart, DE, E-mail: s.weigleinfkf.mpg.de

1

regime 2). In this paper, we will focus our attention exclusively on the high-temperature intrinsic region (Fig. 1, regime 3 and inset), directly below the transition point (the transition being either melting or sublattice melting), in which an anomalous superexponential behavior arises.

Figure 1: DC conductivity plot 1 for impure AgCl samples as measured by Abbink and Martin, 2 with inset focus on the high-temperature intrinsic regime taken from the work of Aboagye and Friauf. 3 In the main figure, region 1 refers to vacancy conductivity in a doped regime, where the influence of vacancy-dopant associates can be seen at low temperatures, region 2 refers shows the low-temperature situation in nominally purified samples, where low levels of impurities begin to dominate, and region 3 is the hightemperature intrinsic region. Inset: curve 1 is experimental data gathered by Aboagye and Friauf, curve 2 is the authors’ fit according to Debye-Hückel-Lidiard, and curve 3 is the ideal model. Phonon dispersion profiles point to a high deformability of the silver halides, 4–6

2

suggesting a “softness” of the lattice compared to the alkali halides. This softness – it has been variously interpreted as a high polarizability of Ag+ 5,7,8 or as an extraordinary ability of silver cations to polarize halides 9,10 – leads to interesting behavior in the DC conductivity. In silver iodide, a superionic transition is encountered, which can also be viewed as a melting of the silver sublattice. In AgCl and AgBr, there is no superionic transition, however, an anomalous increase in the conductivity over and above both the predictions of the ideal and ideal plus Debye-Hückel-Lidiard models was clearly shown by Aboagye and Friauf in both materials. 3 The interpretation of this anomaly as related to the increase in the number of Frenkel pairs was confirmed by the sodium-diffusion experiments of Batra and Slifkin. 11,12 Semi-empirical calculations of the Frenkel energy in AgCl by Catlow and co-workers suggested that the anomaly could be explained by a decrease in the energy of formation of Frenkel pairs with temperature, 13,14 a concept that was first presented in terms of the link between the Frenkel energy and the expansion of the crystal by Schmalzried. 15,16 Hainovsky and Maier 17,18 later showed that the conductivity anomaly in silver halides is well described by an excess chemical potential that follows a cube root law in the defect concentration. Their explanation is based on the assumption that the excess free energy of the more or less randomly distributed defects can be mimicked by the Madelung energy of an ordered defect lattice. The success of this formulation and its tenable link to Coulomb interactions suggests that the problem of defect Coulomb interactions in the silver halides deserves another look. Also weighing in is the as yet unexplained exponential increase in the dielectric constant of AgCl with temperature in the anomolous region. 19,20 A specific mechanism for this phenomenon has not been proposed. At the same time, the question arises: if dipoles are forming in the lattice, as is suggested by the increase in the dielectric constant, then what is the nature of the dipole, and how does it influence the coulombic interaction energy and the population of mobile Frenkel defects? We will explore this question in two parts. The first part is a necessary prelude to the purpose described above: an examination of the most frequently used methods for the mean-field estimation of the excess free energy due to Coulomb interactions among

3

lattice defects in ionic solids. We have written a Monte Carlo simulation focused on the situation in silver chloride: small concentrations of defects with formal charges interacting coulombically on a fixed lattice. The purpose of the simulation is to provide a fixed reference point for evaluation of theoretical methods for estimating Coulomb energies in solids. The most common model appearing in the literature is the Debye-Hückel-Lidiard theory (DHL). 21 This theory is essentially an application of the Debye-Hückel-Bjerrum pair theory for liquid electrolytes in the restricted primitive model (RPM) to the solid case. As DHL generally only considers clustering involving impurities, its expression in pure, intrinsically defective solids effectively becomes that of Debye-Hückel (DH) evaluated with an ion-pair separation derived from the lattice. We will therefore focus on comparisons between the simulation and the Debye-Hückel theory, along with the mean spherical approximation (MSA) 22–24 and the cube root law, in Part 1. These models have been chosen in large part for their ease of applicability and popularity in the context of solid-state defect modeling; the idea of comparing them to the lattice simulations is to determine whether it is worthwhile to try more sophisticated techniques, many of which are available in the literature. The principal conclusion, which has an immediate bearing on the following analysis in AgCl, is that the error inherent in applying central-ion, continuum theories to the lattice arises from the discrete nature of the lattice in the neighborhood of the central ion. This motivates the adaptation for the application part of an improved central-ion theory developed by Fisher and Levin, 26,27 in which nearest-neighbor interactions are not treated coulombically. Bound pairs of ions are instead considered as a single species: a central dipole, which influences the rest solution through coulombic forces. In Part 2 of the paper, we focus on the comparison of dielectric and conductivity data in AgCl. For this, we adapt two different theories to our purposes – the Onsager theory of the influence of dipoles on the dielectric constant in solution 25 and the FisherLevin mean-field theory of electrolytes mentioned above. The dipole we propose is a short-lived displacement of the silver ion from its regular lattice site beyond that found in the course of typical lattice vibrations. In order to

4

facilitate the analysis with a minimum of mathematical complexity, we treat this distribution of displacements as a single chemical species: a first-nearest-neighbor pairing of a silver interstitial with a silver vacancy. Such a pair is likely to be very short-lived, and may not have a well-defined energetic minimum in AgCl, particularly at temperatures close to the melting point. However, we show that such short-lived displacements may still affect the dielectric response and conductivity of the material.

1 Free-Ion Mean Field Models for Coulomb Interactions in AgCl In the RPM, particles are considered to be hard spheres of equal size in a background of uniform dielectric constant. This simple idea has been found to be a useful model for the study of electrolyte behavior in the context of the liquid. The goal of this work is to provide the beginnings of a solid-state analog to several early continuum Monte Carlo studies 28–30 that compared various mean-field models to the RPM away from the critical point. It is this type of comparison that should prove most useful in the evaluation of mean-field theories used in the establishment of solid-state defect models. Although no study has yet appeared strictly fitting this description, the group of Tétot has come closest, examining solid-gas oxygen equilibrium in an oxide with charged defects and Coulomb interactions over the course of several studies. 31–33 More recent Monte Carlo work has focused on the critical behavior of the RPM. 34–36 In the course of these efforts, some studies dedicated to the critical behavior of the lattice restricted primitive model have appeared. 37,38 This model is, as usually presented, a simple cubic lattice with an adjustable ratio between the size of the particle and the lattice spacing.

1.1 Simulation Details The most important aspect of the Monte Carlo simulation is its treatment of the lattice. Simulations involving charged particles restricted to a fixed lattice can be very efficient

5

– on the other hand, we must contend with the fact that the “particles” in our simulation are actually defects in a lattice full of real ions. The Tétot group simulations cited above, for example, treated Coulomb interactions between real ions, which is not practical for the low defect concentrations (some on the order of ppm) we are interested in here. Fortunately, a simulation treating only defects and relegating the perfect lattice to a configuration-independent background is easily justified based only upon the concept of the Frenkel energy. Consider the Frenkel defective solid at constant volume, temperature and number of particles, with only a configuration-independent change in energy and vibrational entropy associated with defect formation (the Frenkel energy), and effective Coulomb interactions between defects. The Helmholtz free energy of such a system is: F = F0 − kB T log

"

N

Ω(X,N)

X=1

i=0

∑ e−X µ0 /kBT ∑

e−Ei /kB T

#

(1)

where the index i refers to defect configurations and Ei the configuration-dependent (Coulomb) energy, µ0 is the Frenkel energy, N is the total number of cation sites and X the number of defect pairs. F0 is the free energy of the perfect lattice. Eq. (1) shows that we may calculate the equilibrium defect concentration at a given temperature and volume by means of a grand canonical-like simulation involving only defects. For an evaluation of the theory over a wide concentration and temperature range, however, it is more practical to fix the concentration of Frenkel pairs at the outset. Assuming N → ∞, we therefore replace the sum inside the logarithm in Eq. (1) with its maximum term in the outer sum:

f=

Ω(X,N) kB T F = f 0 + x µ0 − log ∑ e−Ei /kB T N N i=0

(2)

We may thereby concentrate on evaluation of internal energies for various values of T and x = X/N. We used the minimum image approximation for evaluating configuration energies. Use of the minimum image method is made possible by the relatively low concentrations; the criterion used to assess the validity of the minimum image approximation is 6

the dimensionless parameter Γ=



4π x 6a3

1/3 

q2 ε kB T



(3)

with a the distance between next-nearest-neighbor cations and anions in the perfect lattice, q the elementary charge and ε the dielectric constant. According to comparisons with the Ewald method for simulations of a one component plasma in a neutralizing background, Γ should be less than 10 for the minimum image method to provide a good approximation. 28,39 In our case, Γ ≤ 3.37. In order to further reduce any error due to the minimum image convention or the imposed periodicity germane to any finite-size simulation, we used a dodecahedral simulation box (as the Tétot group did 33 ). We can confirm that the dodecahedron produces much better estimate of the Madelung constant (calculated in our case by filling the cation lattice with defects, while restricting the placement of interstitials to form a zinc-blende structure). However, for the (much lower) defect concentrations reported below, this change of simulation geometry made only a negligible difference in the results, indicating that our use of the minimum image approximation here is entirely appropriate. Only single particle moves were made, using a random choice of particles. Due to the fact that there are twice as many interstitial sites as regular cation sites, it was ensured that the odds of choosing an interstitial was twice that of choosing a vacancy. A new configuration was retained or discarded based on the Metropolis criterion. Evaluation of a single state point consisted of running 32 independent Monte Carlo simulations. Given a random starting point for each simulation and assuming ergodicity, they may be considered to be independent samples. This enables the application of a confidence interval test for termination. The equilibration criterion was the onset of the fluctuation of the ensemble average (determined over short intervals) about the average calculated from the beginning of the equilibration run, after a minimum number of Monte Carlo steps (generally on the order of 104 ). The code was written in C++ using the SPRNG random number generator (version 4.0) 40,41 and the Blitz++ container class. 42 MPI was implemented using MPICH. 43,44

7

The simulation was tested by comparing the results of a simulation at 720 K, low concentrations and progressively larger distances of closest approach to the predictions of the Debye-Hückel limiting law:

udhll = −x

q2 κ ε

(4)

where

κ=



4π q2 x ε a3

1/2

(5)

is the inverse Debye length. Five simulation sizes (125, 216, 343, 512 and 729 defect pairs) were evaluated to an accuracy of 5% with a 95% confidence interval, and the results extrapolated via a linear law in the inverse simulation size to the thermodynamic limit (see Fig. 2). Table I displays the limiting values for x = 10−6 and x = 10−7 . The predictions of the simulation come closer to the theoretical as concentrations decrease and enforced separation distances increase, finally coming within 1.2% at x = 10−7 and third-nearest-neighbor separation. These results are reasonable and can be taken as an indication that the simulation is working properly, although it also foreshadows the strong differences between the solid and continuum cases we will explore below. x 10−6 10−7

1NN 1.76 5.42

2NN 1.68 5.35

3NN 1.67 5.25

DHLL 1.64 5.19

Table 1: Thermodynamic-limit values of the excess internal energy per cation site for x = 10−6 and x = 10−7 at first through third nearest neighbor spacing, compared with the Debye-Hückel limiting law. Values are listed on the top line in 10−9 eV, and on the lower line in 10−11 eV.

1.2 Mean-Field Theories The Debye-Hückel internal energy per cation site is:

udh = −x

q2 κ ε (1 + κ r)

where r is the distance of closest approach between two ions.

8

(6)

Figure 2: Extrapolations of the internal energy per cation site to the thermodynamic limit (abscissa is the inverse number of defect pairs in the simulation, 1/X) for hightemperature, low-concentration simulations at a) x = 10−6 and b) x = 10−7 and first through third nearest-neighbor separation distances. In each plot, the dash-dot line is the linear fit to 1NN, the dashed line is the fit to 2NN, and the dotted line the fit for 3NN separation distances. A horizontal line indicating the level of the Debye-Hückel limiting law is included in each plot for comparison. The MSA is a closed-form solution for the coulombic excess energy in the continuum derived using Ornstein-Zernike integral equation theory. 22–24 The internal energy per cation site in the MSA is: −x

2q2 B εr 9

(7)

where B=

1 + ξ − (1 + 2ξ )1/2 , ξ

ξ = κr

(8)

A successful approach to estimate the conductivity anomaly in the silver halides was provided by the cube root method. 17,18,45 The ansatz is a crude, ad hoc model that distributes the defects uniformly over the lattice, taking the form of a rocksaltstructured defect superlattice. The activity coefficient is then given by the charging process: kB T log γ± = 2

Z 1 ζ q2 f 0

εd



(9)

q2 f x1/3 εa

(10)

= UM x1/3 /ε

(11)

=

where f is the Madelung constant for the rocksalt structure, d = a/x1/3 is the mean separation distance of defects on the superlattice and UM the Madelung energy of the perfect lattice. This formalism neglects the excess configurational entropy, and can therefore only be considered a low-order estimate. But the superlattice configuration is a local minimum of the energy in configuration space, which we expect to come close to the actual energy at moderate temperatures. Indeed it was found that the excess energy is well-described by a cube root behavior. 18 Despite this success, as an ad hoc model, it is not adequate for the kind of detailed study undertaken in Part 2.

1.3 Results The first results are a simple, direct comparison between the theories and the simulation √ with no restrictions, and r = a 3/2. Ensemble averages for the internal energy on a two-dimensional 6×6 grid of state points spanning T = 500 − 720 K (in even divisions of 1/T ) and xt = 10−5 − 10−2 (logarithmically) were calculated. The lattice parameter and dielectric constant of low-temperature AgCl (5.54 Å and 12.5, respectively 20) were used. Simulation sizes of 27, 64, 125, 216 and 343 pairs were calculated to 5% accuracy on a 95% confidence interval, and extrapolated to the thermodynamic limit

10

using a linear law in 1/X. The difference is, unsurprisingly, large: an extreme overestimate (considering the excess energy to be negative) in all cases, as plotted in Fig. 3.

Figure 3: Comparison of simulated and DH internal energy per site versus temperature and for various concentrations, with no restrictions, for total cation vacancy site fractions of a) 1.0×10−5, 4.1×10−5 and 1.7×10−4, and b) 5.8×10−4, 2.9×10−3 and 8.0×10−3. Open symbols are the simulated values and closed are DH. It is likely that much of the overestimate reflected in Fig. 3 is due to the tendency for pairs to condense into bound pairs; to test this assumption, we extended the distance of √ closest approach to the second-nearest-neighbor position, or r = a 11/2. The results 11

are shown in Figs. 4 and 5. It shows a much smaller, but still significant overestimate.

Figure 4: Comparison of simulated and DH internal energy per site versus temperature and for various concentrations for second-nearest neighbor distance of closest approach, for total cation vacancy site fractions of a) 1.0×10−5, 4.1×10−5 and 1.7×10−4, and b) 5.8×10−4, 2.9×10−3 and 8.0×10−3. Open symbols are the simulated values and closed are DH. It is interesting that in both first- and second-nearest-neighbor cases, the percentage overestimate reaches a maximum with concentration and then begins to fall again. This is consistent with an explanation for the overestimate of the internal energy as due to the formation of bound pairs: at higher concentrations, the number of such bound pairs

12

Figure 5: The percentage of overestimate of DH and MSA internal energies per site versus temperature and for various concentrations for second-nearest neighbor distance of closest approach, for total cation vacancy site fractions of a) 1.0×10−5, 4.1×10−5 and 1.7×10−4, and b) 5.8×10−4, 2.9×10−3 and 8.0×10−3. Open symbols are DH and closed symbols MSA. naturally due to a lack of available sites, and the trapping effect is therefore not as severe. In Figure 6, the data for unrestricted and second-nearest-neighbor restricted (respectively) are plotted as concentration vs. excess energy per defect pair. The excess energy per pair is a rough estimate of the excess chemical potential (−kB T log γ± ).

13

The slope of a straight line fit to the log-log plot therefore specifies the approximate power law fit to the excess chemical potential at constant temperature. The situation is strikingly different between the unrestricted and second-nearest-neighbor cases: in the former, the slope is close to 1/3 at high temperatures but quickly drops, while in the restricted situation the slope lies much more stably in the range between 0.4 and 0.3. Since the difference between the two cases is the likely predominance of closely-bound species in the unrestricted simulation, this suggests that the cube-root formulation constitutes a good approximation of the excess free energy due to unbound defects.

2 Bound Pairs in AgCl 2.1 Fisher-Levin Theory Some years ago, Fisher and Levin extended Debye-Hückel beyond the Bjerrum correction for bound pairs by estimating the interaction between a bound pair dipole (which is considered to be neutral in the Bjerrum theory and in DHL) and the remaining free ions. 26,27 The resulting theory was quite successful in the prediction of critical points in the RPM. This approach has, in fact, been extended in a general way to the lattice case, 46 but the simplicity of the results in the continuum made it worth discovering whether or not they could be useful in the solid.

2.1.1 Theory In order to apply the continuum results to the solid, we have treated a next-nearest neighbor pair as a separate chemical species with its own formation energy, and adapted the estimation of the Bjerrum pair separation and the effective shell radius of the dipole to the lattice. This is somewhat analogous to Lidiard’s approach in the adaptation of Debye-Hückel-Bjerrum to the solid. We consider a binary, 1-1 solid electrolyte of the rocksalt structure with a Frenkel defect in the cation sublattice. The system has a fixed number of particles (anions and cations), along with a fixed volume and temperature. Fixing the number of particles, the volume and the crystal structure fixes the lattice parameter. In terms of the placement of 14

Figure 6: The excess free energy due to Coulomb interactions per defect pair plotted versus the defect site fraction, for a) unrestricted and b) next-nearest-neighbor distance of closest approach. The slopes of linear fit in the log-log formulation are given for each temperature. ions, we ignore distortion of the lattice due to defect formation (the energetic effects of which are accounted for in the Frenkel energy and the energy of formation of a bound pair). This approach ignores any change in volume due to defect formation. (The Gibbs energy of defect formation at constant pressure and the corresponding Helmholtz energy at constant volume are equal within a second order term in the actual change in volume with defect formation. 47)

15

As before, the Helmholtz free energy of the system may be written in terms of that of the perfect lattice and that due to defects:

F = F0 (N, T,V ) + Fd (N, T,V, Xt )

(12)

where N is the total number of cation sites and Xt the total number of Frenkel pairs. In a first approximation, the defect-dependent term can be split into a configurationindependent term, an ideal configurational entropy and a Debye-Hückel interaction term: fd = xt µ0 (T ) + kBT xt log

xt2 + fdh (xt , T ) 2

(13)

where fdh is the Debye-Hückel interaction, and we have taken into account 1 ≫ xt . Minimizing Eq. (13) with respect to xt yields the equation of state typically used in the analysis of intrinsic disorder in solid state systems. We now consider oppositely charged, first-nearest-neighbor pairs to have a distinct energy of formation. The resulting free energy is: 

p + x log x2 + (1 − p − x)log (1 − p − x) 8  + (2 − 8p − 9x)log (2 − 8p − 8x) + fdh (x, T ) + fdi (x, p, T ) (14)

fd = xµ0 + pµ1 + kB T p log

where µ1 is the formation energy for bound pairs, x is the number of free pairs per cation site, p is the number of bound pairs per cation site (such that x + p = xt ), and fdi is the dipole screening energy. The dipole screening energy has been derived in detail for the case of the liquid in Ref. 27 , using a mean-potential approximation in the manner of Debye and Hückel. The result, in terms of free energy per cation site, is:

fdi = −kB T

π pxa 4 (T ∗ )2 r2

+ O (κ r2 )

(15)

where r2 is radius of the hard-shell sphere enclosing the dipole pair and T ∗ is the

16

reduced temperature: T∗ =

kB T ε a q2

(16)

In writing Eq. (15) we took the nearest-neighbor interstitial-vacancy separation dis√ tance a 3/2 as both the Bjerrum pair separation distance and the ionic diameter. As in the continuum case, the choice of the effective sphere radius is somewhat arbitrary. Levin and Fisher used the average angular radius of the bisphere formed by the overlapping exclusion zones of the two ions comprising the dipole, yielding 1.162r1, with r1 the Bjerrum separation distance. In AgCl, this combination is almost exactly a. This is both convenient and close to optimal: in the lattice, the ion positions surrounding the dipole can be thought of as lying on the faces and edges of squares of successively increasing side length. The smallest, of length a/2, includes the dipole constituents themselves; the next largest, of side length 3a/2, includes a set of cation and interstitial sites at distances between 0.83a and 1.3a from the dipole center, several of which fall at 1.09a. Setting r2 = a, Eq. (15) becomes:

fdi = −kB T

π px 4 (T ∗ )2

(17)

With the above definition, along with that for the Debye-Hückel energy, we can find the equations of state for different thermodynamic conditions.

2.1.2 Simulation Results Since we are primarily interested here in examining mean-field representations of the Coulomb energy, it is convenient to work in an ensemble where the total number of Frenkel pairs is constant, and the bound pair is introduced as a separate species. The free energy is: kB T f = f0 + xt µ0 − log N

Xt

Ω(N,Xt ,P)

P=0

i=0

"

∑ e−P(µ1 −µ0 )/kBT ∑

17

e−Ei /kB T

#

(18)

Given µ1 − µ0 and Xt , a Monte Carlo simulation can directly estimate average populations of bound pairs and configurational energies. The simulation was modified to keep track of the number of next-nearest-neighbor pairs, the formation or dissolution of which was associated with a given energy. Clusters of more than one 1NN pair per ion were effectively prohibited (or strongly discouraged), as each ion was only considered by the simulation as participating in at most one bound pair. Formation of any additional pairs were evaluated at zero energy, meaning that electrostatic repulsion between any approaching ion and the like-signed member of the bound pair is very likely to prohibit higher-level clustering. The simulation kept track of the degree of higher-order clustering, and it was, in fact, found to remain negligibly low. After some trial-and-error, it was determined that a (temperature independent) bound pair formation energy (µ1 − µ0 ) of -0.31 eV resulted in a good turnover and movement of bound pairs, along with a sizeable population of them at higher defect concentrations. To find the theoretical equation of state matching the ensemble used in the simulation, we substitute Eq. (17) into Eq. (14), replace p with xt − x and minimize with respect to x. The result is:

µ0 − µ1 + kB T log

q2 κ 4x2 π (xt − x) − − kB T =0 xt − x ε (1 + κ r) 4 (T ∗ )2

(19)

where we have used the approximation 2 − 8xt − x ≈ 2 and r is the distance of closest approach for free ions, which is taken to be the second-nearest-neighbor distance r = √ a 11/2. This equation can be solved numerically for x at a given temperature and xt . The results for the average population of bound pairs are shown in Fig. 7. The continuum theory generally shows a greater tendency to form pairs; this effect increases with increasing overall concentration and decreasing temperature. At concentrations below 10−4 , the difference is negligible. As concentrations approach 1%, the pair population overestimate reaches as high as 45%. This tendency is almost certainly due to the fact that the DH theory, even at en-

18

Figure 7: Simulated calculations of the fraction of bound pairs as a function of temperature, compared with theoretical predictions, for total cation vacancy site fractions of a) 1.0×10−5, 4.1×10−5 and 1.7×10−4, and b) 5.8×10−4, 2.9×10−3 and 8.0×10−3. Open symbols are the simulation and closed are theoretical. forced 2NN separations, overestimates the excess energy substantially; the theoretical system is therefore driven more strongly toward the energy benefit of pair formation. If anything, one expects any error in the theoretical dipole screening energy to fall on the side of an overestimate, which would drive the results in the other direction. (In other words, one would expect the theory to predict a lower concentration of bound pairs if the deciding factor were the dipole screening energy alone.) This suggests that

19

the Fisher-Levin theory used in the solid is primarily limited by the shortcomings of the attached DH theory operating in the lattice context.

2.2 Application in Silver Chloride It has been assumed that the next-nearest-neighbor associate between interstitial and vacancy in AgCl is unstable. 19 Such bound Frenkel pairs are thermodynamically stable in some cases, for instance in SiO2 . 48 In AgCl, however, it might be assumed that it would be difficult to designate such a bound pair as a distinct state (as opposed to a quasi-continuous density of states) due to the softness of the lattice, as discussed in the introduction. A silver ion leaving its regular site in the rocksalt structure and entering a body-centered interstitial position (which is likely to be unstable even far removed from the vacancy in favor of a split interstitial configuration 49) would not, in this picture, find a single stable state, instead moving over a nearly continuous set of states. Nonetheless, the consideration of a distinct bound pair state in this manner might be thought of as a good first approximation to a more complicated reality, and as it greatly simplifies the analysis, we adopt this convention in the following. In terms of the defect chemistry, what we propose is the following quasi-chemical reaction: AgxAg + Vxi ⇋ Agi VAg

x

(20)

Since the associate in Eq. (20) is uncharged, it does not affect the usual defect chemical treatment except through the secondary effect of the associated dipole (investigated below) and the site balance at high concentrations. We may write the usual mass-action law as: 

Agi VAg

x 

∆S0 ∝ κ ∝ exp kB 



−∆H 0 exp kB T 



(21)

2.2.1 Dielectric Properties The change of the dielectric constant with temperature as measured by Smith 19 is shown in Fig. 8. It is clear that there is an exponential behavior at high temperatures.

20

We propose that this exponential increase is due to the increase in the concentration of bound Frenkel pair dipoles with temperature according to the mass action law of Eq. (21).

Figure 8: The change in the dielectric constant with temperature at 2.4 × 1010 Hz, as measured by G. Smith. 19 The frequency of the dielectric constant measurement deserves some discussion. Sevenich and Kliewer 19 (in whose paper Smith’s data was reprinted) list the frequency as 2.4 × 1010 Hz. This is a frequency likely to be well above the resonance frequencies of well-defined defect associates. However, we maintain that, given the undefined nature of the bound pair dipole considered here, reorientation times will be much faster – on the order of the vibrational frequencies of silver ions. To calculate the effect of the dipoles on the dielectric constant, we turn to Onsager’s theory of the dielectric constant due to rigid dipoles. 25 Onsager’s expression for the dielectric constant of a solution containing a dilute concentration of dipoles (assuming the polarizability of the dipole species to be zero) is: 25

ε = εs +

4π (m∗ )2 ρd 3kB T

21

(22)

where εs is the dielectric constant of the solvent, ρd is the concentration of dipoles and ∗

m =



3εs 2εs + 1

2

m

(23)

with m the dipole moment. We consider εs to be the dielectric constant of AgCl at low temperatures. For the experimental system, there exists no overall limitation (other than the number of cations in the lattice) on the number of Frenkel pairs that can form. Therefore, in order to make predictions about the experimental system, our treatment of Eq. (14) must change. We now minimize with respect to both x and p simultaneously, leading to the set of equations:

µ0 + kBT log

q2 κ x2 πp =0 − − kB T 2 ε (1 + κ r) 4 (T ∗ )2

µ1 + kB T log

πx p =0 − kB T 8 4 (T ∗ )2

(24) (25) (26)

where again we have used the fact that x, p ≪ 1. Eliminating p, we find:

µ0 + kBT log

  q2 κ x2 µ1 − x∆T =0 − − 8∆T exp − 2 ε (1 + κ r) kB T

(27)

where ∆T =

π kB T 4 (T ∗ )2

(28)

examining Eq. (27), we may write:   µ1 − x∆T µex = µdh − 8∆T exp − kB T

(29)

√ Taking equations (23), (22) and (25) along with the expression m = qa 3/2 leads to:

ε = εs +

4π q2 2kB Ta



3εs εs + 1

2

22

  µ1 − x∆T exp − kB T

(30)

We recognize that, since ∆T depends on ε , Eq. (30) can strictly be solved only intrinsically. However, for this particular case, x∆T ≪ µ1 over the temperatures and concentrations studied, and we may therefore ignore this term in the analysis. Eq. (30) (taking ∆T = 0) can be written in empirical form as:

ε = A+

−C B exp T T

(31)

where A = εs , B is related to the entropy of dipole formation and C to the energy of formation. The experimental data at high temperatures (above 500 ◦ C) is plotted in Arrhenius form in Fig. 9. Its adherence to Eq. (31) is quite good, as shown. The least-squares fitting yields A = εs = 14.5, B = 8.30 × 106, and C = 5.16 × 103, with the latter two corresponding to h1,ε = 0.444 eV and s1,ε = 2.01kB , where µ1 = h1 − T s1 .

Figure 9: Experimental data for dielectric permittivity in AgCl from 520-680 K, along with the model fit to the data based on Eq. (31).

2.2.2 Conductivity Given the energy and entropy of bound pair formation, Eq. (29) can be compared directly to the experimentally determined excess free energy. 3 The result is shown in Figure 10, as a comparison between the predictions of the theory and the calculations

23

of Aboagye and Friauf.

Figure 10: Excess free energy of Aboagye and Friauf 3 compared with the predictions of the Fisher-Levin theory. Bound pair energies calculated by applying the Onsager model to the dielectric constant data, as shown in Figure 9. The predictions of DebyeHückel are shown for comparison.

3 General Discussion Despite the short lifetime of the proposed bound Frenkel pair, it is quite possible to recognize how such an entity – which can form at any occupied silver site – could acquire a significant concentration with respect to the other defects at moderate to high temperatures. It is also possible to imagine how the average orientation of these constantly appearing and disappearing dipoles may respond to an applied electrical field, and how such species would interact with the surrounding (temporally more stable) population of unbound defects, thereby lowering the overall defect formation energy. It is perhaps important to point out that, in general, although the formation energies for such bound pairs are much lower than the formation energies of free pairs, the concentration of bound pairs will increase with increasing temperature (although the proportion of bound pairs with respect to the total number of defect pairs in the system will decline). The excess free energy calculation for the dipole theory shown in Figure 10 over-

24

shoots the estimate of Aboagye and Friauf at low temperatures, only to be overcome at the higher temperatures. This is not a surprising behavior, given that the excess free energy calculated in that work was assumed to be limited to the Debye-Hückel value at low temperatures: any addition to Debye-Hückel will appear to overestimate by comparison. At high temperatures, it is likely that some account of the change in the Frenkel energy as predicted by theoretical and semi-empirical calculations will have to be taken into account. It may also be the case that the upturn in the excess free energy with increasing temperature may sharpen somewhat if we were to consider the density of states associated with the bound pair, as opposed to lumping all of these states together in a single, degenerate energy level as we have here. Such an exercise may be the focus of future studies. The first part of this paper demonstrates that the DHL theory is, in the temperature and concentration range appropriate to the analysis of intrinsically defective AgCl, a significant underestimate (in terms of absolute magnitude) of the Coulomb interaction energy due to unbound defects. The mean spherical approximation is slightly better, but there is still a significant gap between the results of the theory and those of the simulation. Although we have not calculated activity coefficients as part of this study (such calculations can be made in the context of pseudo-grand canonical simulations based on Eq. (1)), the current results suggest that comparisons to cluster models, 50,51 which have depicted the DH approach as a fairly good estimate of the excess energy in the context of AgCl, 19,52 should be reexamined. The general problem of Coulomb interactions in solids – the lattice primitive model – has proven especially difficult to treat generally and concisely. What is required is a method that somehow captures the relative simplicity of the continuum mean-field theories while retaining the most important quantitative features of the lattice problem. It is worth pointing out that there does, in fact, seem to be a region in temperatureconcentration space in which continuum and lattice solutions converge (see Table 1). Furthermore, it would seem from the results in the first section, which show a great improvement in the predictions of the continuum theory moving from first to second nearest-neighbor distances of closest approach, that the area of greatest difficulty for

25

the application of central-ion continuum theories in the solid are at locations close to the central ion. This is intuitive (since it is at these close-in locations that the lattice and continuum will present the greatest contrast from the perspective of the central ion), and suggests that a Mott-Littleton type of approach – explicitly recognizing the lattice within a certain boundary but reverting to the continuum beyond that – might be a good strategy. With the lattice analysis restricted to the short range in such a way, there are many options that present themselves; perhaps the most general of these is the cluster variation method (CVM). If CVM, which is widely used and has a straightforward variational form, could be combined with a central-ion theory – or perhaps central-cluster theory in the spirit of both CVM and the Fisher-Levin approach, and as is sometimes done in shell calculations – this could provide a general framework for the solution of Coulomb interaction problems in the solid while easily incorporating multi-body non-classical interactions.

Conclusion The assumption that silver ions displace from their regular lattice positions to form “intrinsic bound pairs” is shown to be capable of explaining the exponential increase in the dielectric constant of AgCl with temperature. The same effect leads to a better understanding of the high temperature conductivity behavior in the pre-melting regime. In addition, a Monte Carlo simulation has shown that the continuum-based estimates of the Coulomb interaction free energy provided by the Debye-Hückel theory produces a serious underestimate of the actual Coulomb interaction energy of the solid.

Acknowledgements DSM wishes to thank Eugene Kotomin, Radha Banhatti, Eugene Heifets and Rotraut Merkle for helpful discussions. JM would like to dedicate this work to Prof. John O. “Josh” Thomas, for the occasion of his 65th birthday.

26

References [1] Main figure: Reprinted from The Journal of Physics and Chemistry of Solids, Vol. 27, H. C. Abbink and D. S. Martin, “Ionic Conductivity of Silver Chloride Containing Cadmium Chloride,” Pages 205–215, Copyright (1966), with permission from Elsevier. Inset: J. K. Aboagye and R. J. Friauf, Physical Review B, Vol. 11, Pages 1654–1664, 1975. Copyright (1975) by the American Physical Society. [2] H. C. Abbink and D. S. Martin, J. Phys. Chem. Solids, 1966, 27, 205. [3] J. K. Aboagye and R. J. Friauf, Phys. Rev. B, 1975, 11, 1654–1664. [4] B. Dorner, W. Vonderosten and W. Buhrer, J. Phys. C, 1976, 9, 723–732. [5] K. Fischer, H. Bilz, R. Haberkorn and W. Weber, Phys. Stat. Sol., 1972, 54, 285– 294. [6] W. G. Kleppmann and W. Weber, Phys. Rev. B, 1979, 20, 1669–1677. [7] R. C. Baetzold, C. R. A. Catlow, J. Corish, F. M. Healy, P. W. M. Jacobs, M. Leslie and Y. T. Tan, J. Phys. Chem. Solids, 1989, 50, 791–800. [8] M. Wilson, P. A. Madden and B. J. CostaCabral, J. Phys. Chem., 1996, 100, 1227–1237. [9] M. Bucher, Phys. Rev. B, 1984, 30, 947–956. [10] G. Mahan, Solid State Ionics, 1980, 1, 29–45. [11] A. P. Batra and L. M. Slifkin, Bull. Am. Phys. Soc., 1976, 21, 324–324. [12] A. P. Batra and L. M. Slifkin, Phys. Rev. B, 1975, 12, 3473–3475. [13] C. R. A. Catlow, J. Corish, J. H. Harding and P. W. M. Jacobs, Phil. Mag. A, 1987, 55, 481–498. [14] C. R. A. Catlow, J. Corish and P. W. M. Jacobs, J. Phys. C, 1979, 12, 3433–3445. [15] H. Schmalzried, Ber. Bun. Ges., 1980, 84, 120–124.

27

[16] H. Schmalzried, Z. Phys., 1959, 22, 199. [17] N. Hainovsky and J. Maier, Solid State Ionics, 1995, 76, 199–205. [18] N. Hainovsky and J. Maier, Phys. Rev. B, 1995, 51, 15789–15797. [19] R. A. Sevenich and K. L. Kliewer, J. Chem. Phys., 1968, 48, 3045. [20] J. Corish and P. W. M. Jacobs, J. Phys. Chem. Solids, 1972, 33, 1799. [21] A. B. Lidiard, Phys. Rev., 1954, 94, 29–37. [22] E. Waisman and J. L. Lebowitz, J. Chem. Phys., 1972, 56, 3086. [23] E. Waisman and J. L. Lebowitz, J. Chem. Phys., 1972, 56, 3093. [24] E. Waisman and J. L. Lebowitz, J. Chem. Phys., 1970, 52, 4307. [25] L. Onsager, J. Am. Chem. Soc., 1936, 58, 1486–1493. [26] M. E. Fisher and Y. Levin, Phys. Rev. Lett., 1993, 71, 3826–3829. [27] Y. Levin and M. E. Fisher, Physica A, 1996, 225, 164–220. [28] D. N. Card and J. P. Valleau, J. Chem. Phys., 1970, 52, 6232. [29] J. C. Rasaiah, J. P. Valleau and D. N. Card, J. Chem. Phys., 1972, 56, 248. [30] J. P. Valleau, L. K. Cohen and D. N. Card, J. Chem. Phys., 1980, 72, 5942–5954. [31] M. Benzakour, R. Tetot and G. Boureau, J. Phys. Chem. Solids, 1988, 49, 381– 384. [32] R. Tetot, M. Benzakour and G. Boureau, J. Phys. Chem. Solids, 1990, 51, 545– 550. [33] R. Tetot, B. Nacer and G. Boureau, J. Phys. Chem. Solids, 1993, 54, 517–525. [34] J. M. Caillol, D. Levesque and J. J. Weis, Phys. Rev. Lett., 1996, 77, 4039–4042. [35] G. Orkoulas and A. Z. Panagiotopoulos, J. Chem. Phys., 1994, 101, 1452–1459.

28

[36] J. P. Valleau, J. Chem. Phys., 1991, 95, 584–589. [37] A. Diehl and A. Z. Panagiotopoulos, J. Chem. Phys., 2006, 124, 194509. [38] A. Z. Panagiotopoulos and S. K. Kumar, Phys. Rev. Lett., 1999, 83, 2981–2984. [39] S. G. Brush, H. L. Sahlin and E. Teller, J. Chem. Phys., 1966, 45, 2102. [40] M. Mascagni and A. Srinivasan, ACM Trans. Math. Soft., 2000, 26, 436–461. [41] M. Mascagni and A. Srinivasan, ACM Trans. Math. Soft., 2000, 26, 618–619. [42] T. . L. . Veldhuizen, Computing in Object-Oriented Parallel Environments. Second International Symposium, ISCOPE 98 Proceedings., 1998, 223. [43] W. Gropp, E. Lusk, N. Doss and A. Skjellum, Parallel Computing, 1996, 22, 789–828. [44] W. D. Gropp and E. Lusk, User’s Guide for mpi h, a Portable Implementation of MPI, Mathematics and Computer Science Division, Argonne National Laboratory, 1996. [45] F. Zimmer, P. Ballone, J. Maier and M. Parrinello, J. Chem. Phys., 2000, 112, 6416–6423. [46] V. Kobelev, A. B. Kolomeisky and M. E. Fisher, J. Chem. Phys., 2002, 116, 7589– 7598. [47] C. R. A. Catlow, J. Corish, P. W. M. Jacobs and A. B. Lidiard, J. Phys. C, 1981, 14, L121–L125. [48] A. Shluger, M. Georgiev and N. Itoh, Philos. Mag. B, 1991, 63, 955–964. [49] D. J. Wilson, A. A. Sokol, S. A. French and C. R. A. Catlow, Phys. Rev. B, 2008, 77, 064115. [50] A. R. Allnatt and M. H. Cohen, J. Chem. Phys., 1964, 40, 1871. [51] A. R. Allnatt and M. H. Cohen, J. Chem. Phys., 1964, 40, 1860.

29

[52] A. R. Allnatt, J. Corish and P. W. M. Jacobs, J. Phys. Chem. Solids, 1975, 36, 1233–1235.

30

DC Conductivity and Dielectric Properties in Silver ... - Semantic Scholar

Sep 28, 2009 - theory of the influence of dipoles on the dielectric constant in solution25 and the Fisher- ...... DSM wishes to thank Eugene Kotomin, Radha Banhatti, Eugene Heifets and Rotraut. Merkle for helpful discussions. JM would like to dedicate this work to Prof. John O. “Josh” Thomas, for the occasion of his 65 th.

2MB Sizes 0 Downloads 282 Views

Recommend Documents

DC Conductivity and Dielectric Properties in Silver ... - Semantic Scholar
Sep 28, 2009 - pair theory for liquid electrolytes in the restricted primitive model (RPM) to the solid case. As DHL generally ... the context of solid-state defect modeling; the idea of comparing them to the lattice simulations is to ..... the side

Model of dissipative dielectric elastomers - Semantic Scholar
Feb 3, 2012 - View online: http://dx.doi.org/10.1063/1.3680878. View Table of Contents: ... As a result of their fast response time and high energy density, dielectric ... transducer is an electromechanical system with two degrees of freedom.

Dielectric and Conductivity Studies on Cobalt ...
vices, electrochromism, optical data storage, LB films, liquid crystals, and ... 2400 elemental analyser, which uses a combustion. Correspondence to M. R. ...

DC Proposal: Enriching Unstructured Media ... - Semantic Scholar
//apple.com/wwdc), for SemTech there is just the event website (http://sem- ... ple Inc.”, “iPhone OS”, and “Apple Worldwide Developers Conference”. Each.

DC Proposal: Enriching Unstructured Media ... - Semantic Scholar
Common search operations, for example searching for ... However, more advanced use cases, such as summaries or ..... A Tweet Consumers' Look At Twitter.

Dielectric and Conductivity Studies on Cobalt ...
vices, electrochromism, optical data storage, LB films, liquid crystals, and nonlinear ... room temperature (300°K) to 383°K under dynamic vacuum (10 2 Torr). The cell was ... carried out by using a package based on virtual in- strumentation (VI) .

Improved estimation of clutter properties in ... - Semantic Scholar
in speckled imagery, the statistical framework being the one that has provided users with the best models and tools for image processing and analysis. We focus ...

Improved estimation of clutter properties in ... - Semantic Scholar
0167-9473/02/$-see front matter c 2002 Elsevier Science B.V. All rights reserved. ... (See Ferrari and Cribari-Neto, 1998, for a comparison of computer-intensive and analytical ... degrees of homogeneity, and different models can be used to encompass

Microwave dielectric characterisation of 3D-printed ... - Semantic Scholar
Mar 4, 2016 - Right: BaTiO3/ABS polymer composite containing 50 wt% BaTiO3. .... (6) was generated specifically for the purpose of this illustration by reducing .... Crump, S. S. Apparatus and method for creating three-dimensional objects.

Dielectric anomalies of solid CO and N in the ... - Semantic Scholar
May 24, 1999 - h temperature associated with the onset of hysteresis or a crossover region. We also note that the thermal. w x resistivity data on pure N 10 also ...

Dielectric anomalies of solid CO and N in the ... - Semantic Scholar
May 24, 1999 - onset temperature T ,42 K in the b-phase of pure N have been observed. A simple ... These systems are believed to be analogous to the.

Local and Global Consistency Properties for ... - Semantic Scholar
A placement mechanism violates the priority of student i for position x if there exist .... Let x ∈ X. We call a linear order ≻x over ¯N a priority ordering for position type x. ...... Murat Sertel Center for Advanced Economic Studies Working Pa

Local and Global Consistency Properties for ... - Semantic Scholar
3For instance, Thomson's (2009, page 16) “Fundamental Definition” of consistency deals with a variable .... Ergin (2002) refers to the student-optimal stable mechanism ϕ≻ as the “best rule” and ...... American Mathematical Monthly 69, 9–

Structural and dielectric properties of crystalline and ...
Nov 27, 2004 - the electronic density of states (not shown) indicates that the gap is clean of .... set was optimized [16] in order to reproduce the correct energy ...

in chickpea - Semantic Scholar
Email :[email protected] exploitation of ... 1990) are simple and fast and have been employed widely for ... template DNA (10 ng/ l). Touchdown PCR.

in chickpea - Semantic Scholar
(USDA-ARS ,Washington state university,. Pullman ... products from ×California,USA,Sequi-GenGT) .... Table 1. List of polymorphic microsatellite markers. S.No.

Structural and Dielectric Properties of Polyaniline ...
Fig.3 Variation of Dielectric constant with temperature. This is due to the space charge accumulations at the structural interfaces of an inhomogeneous dielectric material. This resulting space charge produces distortions of the microscopic field, wh

Dielectric response of N2-Ar solid solutions in the ... - Semantic Scholar
High sensitivity dielectric measurements are reported for N2-Ar solid mixtures for 49x(N2) ..... with recent views of glass formation in terms of replica sym-.

Dielectric response of N2-Ar solid solutions in the ... - Semantic Scholar
hindrance that lead to a significant coupling of the center-of- .... 35 K, and (T) changed smoothly without any observable jump at T. Fig. 3. A distinct jump is ...

Structural and dielectric properties of Fe doped ZnO nanoparticles.pdf
direct band gap (3.37 eV) and a large exciton binding energy. of 60 meV .... Page 3 of 7. Structural and dielectric properties of Fe doped ZnO nanoparticles.pdf.

Timing properties of gene expression responses to ... - Semantic Scholar
Feb 7, 2009 - Computer Science Department, Stanford University. Stanford ... tosolic ribosomal genes are only repressed after mitochondrial ribosome is ..... gene expression profile for three methods: polynomials of degree 2 and 3, and.

Conditionals and Pseudo-Conditionals in ... - Semantic Scholar
The theorem statement is followed by three proof sentences that build a quite complex ..... It combines the proof planning technology from the field of automated.

Timing properties of gene expression responses to ... - Semantic Scholar
Feb 7, 2009 - namic analysis for determining the quality of time-series expression profiles. Nature Biotechnology 23:1503–1508. [9] Storey J, Xiao W, Leek J, ...

Some Properties of the Lemoine Point - Semantic Scholar
Jun 21, 2001 - Let A B C be the pedal triangle of an arbitrary point Z in the plane of a triangle. ABC, and consider the vector field F defined by F(Z) = ZA + ZB + ...