30
Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of Texas Austin, Texas and International Solvay Institutes Universitt Libre de Bruxelles Brussels, Belgium and STUART A. RICE Department of Chemistry and The James Franck Institute The University of Chicago Chicago, Illinois VOLUME 116 AN INTERSCIENCE '' PUBLICATION JOHN WILEY & SONS, INC. NEW YORK CHICHESTER WEINHEIM BRISBANE SINGAPORE TORONTO

Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

  • Upload
    others

  • View
    4

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

Advances in CHEMICAL PHYSICS

Edited by

I. PRIGOGINE

Center for Studies in Statistical Mechanics and Complex Systems The University of Texas

Austin, Texas and

International Solvay Institutes Universitt Libre de Bruxelles

Brussels, Belgium

and

STUART A. RICE

Department of Chemistry and

The James Franck Institute The University of Chicago

Chicago, Illinois

VOLUME 116

AN INTERSCIENCE '' PUBLICATION JOHN WILEY & SONS, INC.

NEW YORK CHICHESTER WEINHEIM BRISBANE SINGAPORE TORONTO

Page 2: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of
Page 3: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

ADVANCES IN CHEMICAL PHYSICS VOLUME 1 16

Page 4: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

EDITORIAL BOARD

BRUCE, J. BERNE, Department of Chemistry, Columbia University, New York, New York, U.S.A.

KURT BINDER, Institut fur Physik, Johannes Gutenberg-Universitat Mainz, Mainz, Germany

A. WELFORD CASTLEMAN, JR., Department of Chemistry, The Pennsylvania State University, University Park, Pennsylvania, U.S.A.

DAVID CHANDLER, Department of Chemistry, University of California, Berkeley, California, U.S.A.

M. S. CHILD, Department of Theoretical Chemistry, University of Oxford, Oxford, U.K.

WILLIAM T. COFFEY, Department of Microelectronics and Electrical Engineering, Trinity College, University of Dublin, Dublin, Ireland

F. FLEMING CRIM, Department of Chemistry, University of Wisconsin, Madison, Wisconsin, U.S.A.

ERNEST R. DAVIDSON, Department of Chemistry, Indiana University, Bloomington, Indiana, U.S.A.

GRAHAM R. FLEMING, Department of Chemistry, University of California, Berkeley, California, U.S.A.

KARL F. FREED, The James Franck Institute, The University of Chicago, Chicago, Illinois, U.S.A.

PIERRE GASPARD, Center for Nonlinear Phenomena and Complex Systems, Brussels, Belgium

ERIC J. HELLER, Institute for Theoretical Atomic and Molecular Physics, Harvard- Smithsonian Center for Astrophysics, Cambridge, Massachusetts, U.S.A.

ROBIN M. HOCHSTRASSER, Department of Chemistry, The University of Pennsylva- nia, Philadelphia, Pennsylvania, U.S.A.

R. KOSLOFF, The Fritz Haber Research Center for Molecular Dynamics and Depart- ment of Physical Chemistry, The Hebrew University of Jerusalem, Jerusalem, Israel

RUDOLPH A. MARCUS, Department of Chemistry, California Institute of Tech- nology, Pasadena, California, U.S.A.

G. NICOLIS, Center for Nonlinear Phenomena and Complex Systems, Universitk Libre de Bruxelles, Brussels, Belgium

THOMAS P. RUSSELL, Department of Polymer Science, University of Massachusetts, Amherst, Massachusetts.

DONALD G. TRUHLAR, Department of Chemistry, University of Minnesota, Minnea- polis, Minnesota, U.S.A.

JOHN D. WEEKS, Institute for Physical Science and Technology and Department of Chemistry, University of Maryland, College Park, Maryland, U.S.A.

PETER G. WOLYNES Department of Chemistry, University of California, San Diego, California. U.S.A.

Page 5: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

Advances in CHEMICAL PHYSICS

Edited by

I. PRIGOGINE

Center for Studies in Statistical Mechanics and Complex Systems The University of Texas

Austin, Texas and

International Solvay Institutes Universitt Libre de Bruxelles

Brussels, Belgium

and

STUART A. RICE

Department of Chemistry and

The James Franck Institute The University of Chicago

Chicago, Illinois

VOLUME 116

AN INTERSCIENCE '' PUBLICATION JOHN WILEY & SONS, INC.

NEW YORK CHICHESTER WEINHEIM BRISBANE SINGAPORE TORONTO

Page 6: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

This hook is printed on acid-free paper. @

Copyright 0 2001 by John Wiley & Sons, Inc. All rights reserved

Published simultaneously in Canada.

No part of this publication may be reproduced, stored in a retrieval system or transmitted in any form or by any means, electronic, mechanical, photocopying, recording, scanning or otherwise, except as permitted under Sections 107 or 108 of the 1976 United States Copyright Act, without either the prior written permission of the Publisher, or authorization through payment of the appropriate per-copy fee to the Copyright Clearance Center, 222 Rosewood Drive, Danvers, MA 01923, (978) 750-8400, fax (978) 750-4744. Requests to the Publisher for permission should be addressed to the Permissions Department, John Wiley & Sons, Inc., 605 Third Avenue, New York, NY 10158-0012, (212) 850-601 1, fax (212) 850-6008, E-Mail: [email protected].

For ordering and customer service, call 1 -800-CALL-WILEY

Library of Congress Catalog Number: 58-9935

ISBN 0-471-40541-8

Printed in the United States of America.

10 9 8 7 6 5 4 3 2 1

Page 7: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

CONTRIBUTORS TO VOLUME 116

BIMAN BAGCHI, Solid State and Structural Chemistry Unit, Indian Institute

SARIKA BHATTACHARYYA, Solid State and Structural Chemistry Unit,

MICHAEL E. CATES, Department of Physics and Astronomy, University of

J. W. HALLEY, School of Physics and Astronomy, University of Minnesota,

JOSEPH KLAFTER, School of Chemistry, Tel Aviv University, Tel Aviv, Israel

RALF METZLER, School of Chemistry, Tel Aviv University, Tel Aviv, Israel and Department of Physics, Massachusetts Institute of Technology, Cambridge, MA

D. L. PRICE, Department of Physics, University of Memphis, Memphis, TN WOLFFRAM SCHROER, Institut fur Anorganische und Physikalische Chemie,

PETER SOLLICH, Department of Mathematics, King’s College, University of

S. WALBRAN, Forschungszentrum Jiilich GmbH, Institut fuer Werkstoffe

PATRICK B. WARREN, Unilever Research Port Sunlight, Bebington, Wirral,

HERMANN WEINGARTNER, Physikalische Chemie 11, Ruhr-Universitat

of Science, Bangalore, India

Indian Institute of Science, Bangalore, India

Edinburgh, Edinburgh, United Kingdom

Minneapolis, MN

Universitat Bremen, Bremen, Germany

London, London, United Kingdom

und Verfahren der Energietechnik (IVW-3), Jiilich, Germany

United Kingdom

Bochum, Bochum, Germany

V

Page 8: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of
Page 9: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

INTRODUCTION

Few of us can any longer keep up with the flood of scientific literature, even in specialized subfields. Any attempt to do more and be broadly educated with respect to a large domain of science has the appearance of tilting at windmills. Yet the synthesis of ideas drawn from different subjects into new, powerful, general concepts is as valuable as ever, and the desire to remain educated persists in all scientists. This series, Advances in Chemical Physics, is devoted to helping the reader obtain general information about a wide variety of topics in chemical physics, a field that we interpret very broadly. Our intent is to have experts present comprehensive analyses of subjects of interest and to encourage the expression of individual points of view. We hope that this approach to the presentation of an overview of a subject will both stimulate new research and serve as a personalized learning text for beginners in a field.

I. PRIGOGINE STUART A. RICE

vii

Page 10: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of
Page 11: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

CONTENTS

CRITICALITY OF IONIC FLUIDS By Hermann Weingartner and Wolffram Schroer

MODE COUPLING THEORY APPROACH TO LIQUID STATE DYNAMICS

By Birnan Bagchi and Sarika Bhattacharyya

ANOMALOUS STOCHASTIC PROCESSES IN THE FRACTIONAL DYNAMICS FRAMEWORK: FOKKER-PLANCK EQUATION,

DISPERSIVE TRANSPORT, AND NON-EXPONENTIAL RELAXATION

By R a y Metzler and Joseph Klafter

MOMENT FREE ENERGIES FOR POLYDISPERSE SYSTEMS BJ Peter Sollich, Patrick B. Warren, and Michael E. Cares

CHEMICAL PHYSICS OF THE ELECTRODE-ELECTROLYTE INTERFACE

By J. W Halley, S. Walbran and D. L. Price

1

67

223

265

337

AUTHOR INDEX

SUBJECT INDEX

389

403

ix

Page 12: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of
Page 13: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

ADVANCES IN CHEMICAL PHYSICS VOLUME 116

Page 14: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of
Page 15: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

CRITICALITY OF IONIC FLUIDS

HERMANN WEINGARTNER

Physikalische Chemie II, Ruhr- Universitat Bochum, Bochum, Germany

WOLFFRAM SCHROER

Institut fur Anorganische und Physikalische Chemie, Universitiit Bremen, Bremen, Germany

CONTENTS

I. 11.

111.

IV.

A. B.

C. D. E.

Introduction Background Survey of Experimental Results

Liquid-Vapor Transitions in One-Component Ionic Fluids Liquid-Liquid Demixings in Binary Electrolyte Solutions I . Forces Driving the Phase Separation 2. Coexistence Curves 3. Scattering and Turbidity 4. Viscosity 5. Electrical Conductance 6. Interfacial Properties 7. The Ion Distribution Near Criticality Liquid-Vapor Transitions in Aqueous Electrolyte Solutions Liquid-Liquid Demixings in Multicomponent Systems Summarv

Theoretical Methods at the Mean-Field Level A. Models for Ionic Fluids B. Monte Carlo Simulations C. Analytical Theories of the Restricted Primitive Model

1. General Issues 2. Pairing Theories

D. Lattice Theories E. Beyond the Primitive Models

Advances in Chemical Physics, Volume 116, edited by I. Prigogine and Stuart A. Rice. ISBN 0-471-40541-8 0 2001 John Wiley & Sons, Inc.

1

Page 16: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

2 HERMANN WEINGARTNER AND WOLFFRAM SCHROER

F. G. Summary

A. The Restricted Primitive Model

Mean-Field Theories of Inhomogeneous Fluids and Fluctuations

V. Results from Mean-Field Theories

1. Critical Point and Coexistence Curve 2. The Ion Distribution of the RPM Near Criticality

B. Hard Spheres of Different Size and Charge C. Beyond Primitive Models D. Inhomogeneous Fluids and Fluctuations

I. Charge and Density Fluctuations 2. The Range of Validity of Mean-Field Theories 3. Interfacial Properties

VI. Theories of Critical Behavior E. Summary

A. Critical Phenomena and Range of Interactions B. Lattice Models C. Monte Carlo Simulations of Fluid Models D. Analytical Theories of Fluids

1. The Restricted Primitive Model 2. The Unrestricted Primitive Model 3. The Role of r p 4 Dependent Interactions

E. Crossover Theories and Tricriticality

References VII. Conclusions

I. INTRODUCTION

This chapter deals with critical phenomena in simple ionic fluids. Proto- typical ionic fluids, in the sense considered here, are molten salts and electro- lyte solutions. Ionic states occur, however, in many other systems as well; we quote, for example, metallic fluids or solutions of complex particles such as charged macromolecules, colloids, or micelles. Although for simple atomic and molecular fluids thermodynamic anomalies near critical points have been extensively studied for a century now [l], for a long time the work on ionic fluids remained scarce [2, 31. Reviewing the rudimentary information avail- able in 1990, Pitzer [4] noted fundamental differences in critical behavior be- tween ionic and nonionic fluids.

In preparing the present account, we have been impressed by how much the field has changed since Pitzer's review and a similar review published by us in 1995 [5]. Therefore, a comprehensive account of the present status of the field seems timely. Thus, we presume that the reader is aware of the fundamentals of critical phenomena, as described in many reviews [6-81 and monographs [9-121. Earlier reviews of one or another aspect of ionic criticality by Pitzer [4,13], Levelt Sengers and Given [14], Fisher [15,16], Stell [17,18], and ourselves [5] are notable.

Page 17: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

CRITICALITY OF IONIC FLUIDS 3

The organization of this review is as follows: In Section I1 we describe the theoretical and experimental background of the field. Section I11 reviews experimental work on the criticality of ionic fluids. Section IV presents the basic theoretical methods for describing ionic phase transitions at the mean- field level. Results obtained by these techniques are reviewed in Section V. Section VI reviews the theoretical work concerned with the nature of the critical point. The review closes in Section VII with a brief summary and outlook.

11. BACKGROUND

Liquid-vapor and liquid-liquid coexistence curves of nonionic fluids have an approximately cubic top in the temperature-density plane [6, 8, 91. Any analytical equation of state such as the van der Waals equation predicts a parabolic top, which simply follows from the analytical expansion of the free energy in powers of temperature and density. This anomaly is but one exam- ple for the universal nonanalytic behavior of fluid properties near critical points. In the 1960s and 1970s the theoretical concepts for describing this universal behavior have been developed to a mature state [lo], culminating in the renormalization group (RG) analysis of critical phenomena [ 11,19,20].

Nonanalytical divergences at the critical point result from fluctuations of the order parameter M , which is a measure of the dissimilarity of the co- existing phases. In pure fluids one identifies M with the density difference of the coexisting phases, for mixtures M is related to some concentration variable. The magnitude and spatial extend of these fluctuations is described by their correlation length 5, which upon approach to the critical point on an isochoric path diverges as

as T + 0,

where T = J(T - T,) /T , ( describes the temperature distance from the critical temperature T,, v is a universal exponent, and to a nonuniversal critical amplitude. The diverging fluctuations result in divergences of other thermo- physical properties which along specified paths are described by asymptotic power laws with universal exponents ,u and nonuniversal amplitudes XO:

Table I compiles the scaling laws of interest here. Only two exponents are independent, and the others are related by the exponent equalities given in Table I.

Page 18: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

4 HERMANN WEINGARTNER AND WOLFFRAM SCHROER

TABLE I Asymptotic Scaling Laws and Critical Exponents""

Exponent

Property Application Scaling Law Path' Mean-Field king

Correlation length Either Isothermal One-component

Osmotic Two-component

Heat capacity One-component Two-component

Pressure One-component Chemical potential Two-component Order parameter Either Viscosity Either Correlation function Either Scaling corrections A Either

compressibility

susceptibility

d = d , . X = X , 112 d = d, 1

x = x, d = d, 0 x = x, T = T, 3 T = T, coexistence 1 I2 d = d,,X = X , O(?) d = d , , X = X , 0

0.63 1.24

0.11

4.8

0.326 0.04 0.033 0.54

"The isothermal compressibility is made dimensionless by defining KT = KT/P,., where P, is the critical pressure. The osmotic susceptibility is defined as (aX,/api) , ,T. where Xi is the mole fraction of component i and pi its chemical potential. 'The following exponent equalities are valid: y = p(S - 1 ) ; 2 - 2 = p(S + I ) ; 7 = v(2 - v ) ; Dv = (2 - a). The last equation is only valid for king systems. ' X stands for a suitably defined composition variable. For details see text. dFor values of the exponents given with more decimal places see Ref. 23.

According to RG theory [11, 19, 201, universality rests on the spatial dimensionality D of the systems, the dimensionality n of the order parameter (here n = l), and the short-range nature of the interaction potential @(r) . In D = 3 , short-range means that 4(r ) decays as r-p withp 2 D + 2 - y 2 4.97 [21], where y 2 0.033 is the exponent of the correlation function g(r ) of the critical fluctuations [22] (cf. Table I). Then, the critical exponents map onto those of the Ising spin-1/2 model, which are known from RG cal- culations [23], series expansions [1 l , 12, 241 and simulations [25, 261. For insulating fluids with a leading term of @(r) oc rP6 (6-81 and for liquid metals [27-291 the experimental verification of Ising-like criticality is unques- tionable.

The bare Coulombic interaction ( p = 1) and interactions of charges with rotating dipoles ( p = 4) do not fall into this class, and it has been argued for a long time [30] that in this case one expects analytical ("classical") behavior. This implies that the system can be described by a mean-field Hamiltonian, in which the interaction of a particle is ascribed to the mean field of all other particles, thus ignoring local fluctuations [ 101. In real ionic fluids the

Page 19: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

CRITICALITY OF IONIC FLUIDS 5

screening of the effective electrostatic interactions to shorter range by counterions-that is, the so-called Debye shielding [3 1, 321-may, however, restore king behavior. Thus, one has to discriminate between mean-field behavior and (D=3, n = 1) Ising criticality. The critical exponents are compared in Table I.

We note that even short-range interactions may, however, allow a mean- field scenario, if the system has a tricritical point, where three phases are in equilibrium. A well-known example is the 3He-4He system, where a line of critical points of the fluid-superfluid transition meets the coexistence curve of the 3He-4He liquid-liquid transition at its critical point [33 I. In D = 3, tricriticality implies that mean-field theory is exact [ 1 13, independently from the range of interactions. Such a mechanism is quite natural in ternary systems. For one or two components it would require a further line of hidden phase transitions that meets the coexistence curve at or near its critical point.

Starting with a study on the liquid-vapor coexistence of ammonium chloride (NH,CI) [34], there have been repeated reports on classical ionic criticality [4], but none of these studies allows unambiguous conclusions [14]. In 1990 more decisive results were reported by Singh and Pitzer [35] , who observed a parabolic liquid-liquid coexistence curve of an electrolyte solution. This apparent classical behavior was the stimulus for most theoretical and experimental work reported here.

A prerequisite for understanding these phenomena is a proper description of the molecular forces driving criticality. This puts the problem at the very heart of traditional electrolyte theory. The “restrictive primitive model” (RPM) of equisized, charged hard spheres in a dielectric continuum forms the simplest model to deal with these issues [36, 371. The existence of a critical point of the RPM was for a long time taken for granted [38, 391, and it was proved in the 1970s by Monte Carlo (MC) simulations [40, 411 and statistical-mechanical theories [42]. By a comparative analysis of the theory of ionic and neutral fluids, Hafskjold and Stell [36] asserted in 1982 that the RPM shows Ising behavior. There are, however, problems in this regard [15], and a decisive RG analysis is still lacking. Some time ago, Fisher [15] and Stell [17] discussed the status of the theory, but did not agree on firm conclusions.

There are other scenarios for an apparent mean-field criticality [15, 171. The most likely one is crossover from asymptotic king behavior to mean- field behavior far from the critical point, where the critical fluctuations must vanish. For the vicinity of the critical point, Wegner [43] worked out an expansion for nonasymptotic corrections to scaling of the general form

x = xo ?. ( 1 + x, FA + x, Pa + . ’ .) , (3)

Page 20: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

6 HERMANN WEINGARTNER AND WOLFFRAM SCHROER

where A is the universal exponent for corrections to scaling (cf. Table I), and the Xi are nonuniversal correction amplitudes. In nonionic fluids, the lowest- order corrections of this expansion prove to be sufficient for describing the near-critical behavior [6-81, because the crossover region is wide and mean-field behavior is never attained [44,45]. An apparent mean-field beha- vior of ionic fluids may mean that the asymptotic range is more narrow, as was indeed observed in early work on the liquid-liquid phase transition in solutions of sodium in ammonia [46]. About 40 years ago, Ginzburg [47] derived a criterion for the temperature distance Fx from the critical point up to which mean-field theory remains self-consistent, thus allowing us to estimate the location of the crossover regime. Contemporary crossover the- ories incorporate Ginzburg’s ideas into an RG treatment, as discussed in Sec- tion V1.E.

111. SURVEY OF EXPERIMENTAL RESULTS

A. Liquid-Vapor Transitions in One-Component Ionic Fluids

The critical points of alkali halides such as NaCl are located at temperatures above 3000 K [48-501, while experimental data do not extend beyond 2000 K 1481. Critical point estimates are, however, often needed for comparative purposes. Matching results of molecular dynamics (MD) simulations to the available experimental data, Guissani and Guillot [5 1 ] developed an equation of state (EOS) for NaCl which predicts T, = 3300 K and a critical mass den- sity of dc=0.18g crnp3. Pitzer [13] recommended a lower critical density, but, as discussed later, some MC data used in his assessment are questionable.

Comparison with nonionic fluids is possible through the corresponding- states principle [37] . If, as usual, the reduced temperature T * is defined as the ratio of the thermal energy kBT to the depth of the potential, one finds for a symmetrical Coulomb system with charges q = z+e = Iz- le

where p = l /kBT.&o is the dielectric constant of the background. o = (o+ + o-)/2 is the mean diameter of the ions. The reduced density is defined via the excluded volume as

3 P* =Pa >

where p = p+ + p- = (N+ + N - ) / V is the total number density of the par- ticles. With EO = 1 and o = 0.276 nm for NaC1, one finds from Guissani and

Page 21: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

CRITICALITY OF IONIC FLUIDS 7

Guillot’s EOS TF N 0.058 and p: N 0.08. As discussed later in Section V.A, the best available MC results for the charge- and size-symmetrical hard- sphere ionic fluid yield [52, 531

T: = 0.048 - 0.05, pz = 0.07 - 0.08, (6)

in surprisingly good agreement with these values. Both Tr and pz are much lower than observed for simple nonionic fluids; for example TZ = 1.3 1 and pr N 0.32 for the Lennard-Jones fluid [54]. The low values are signatures for phase transitions driven by Coulombic interactions [37].

A direct observation of critical points of molten salts is possible, if specific interactions reduce T,. One such system is NH4C1, where decomposition in the gaseous phase into HC1 and NH3 lowers T, to about 1150 K [34]. Liquid NH4C1 is a highly conducting ionic melt even near criticality [55]. Guillot and Guissani [56, 571 performed an extensive theoretical analysis of the effect of this decomposition on the EOS. Buback and Franck [34] determined in 1972 the liquid-vapor coexistence curve of NH4Cl.

Guided by comparison with the lattice gas model, the order parameter of one-component systems is associated with the mass density d; that is, M = d= (d - d,)/d, (in real fluids a better choice would be a linear com- bination of density and entropy, which accounts for the liquid-vapor asymmetry [58]). Then, the scaling law in Table I predicts a straight line in a double-logarithmic plot of the density difference between the coexisting liquid and gaseous phases, dL-dc, vs. T-T, with a slope given by the exponent p. Figure 1 shows that for NH4CI such a representation yields p N 0.5, distinctly different from p 2 0.33, generally found for nonionic fluids, as exemplified in Fig. 1 by data for xenon and ammonia. Because, at the extreme conditions the critical point of NH4C1 could be reached only to within 10 K, it is easy to object against the observed mean-field behavior [59]. Nevertheless, Buback and Franck’s results indicate that ionic fluids may differ appreciably in their effective critical behavior from nonionic fluids.

1178K) [60]. BiC13 is, however, little conducting in the liquid phase, and it resembles nonionic rather than ionic fluids. This may warn that specific interactions may destroy the peculiar effects of ionic criticality. Likewise, water (T, = 647 K) shows Ising behavior [61]. At the critical point, autoprotolysis of water is enhanced by three orders of magnitude relative to ambient conditions [62].

For comparison, we briefly consider results for metallic fluids. Early experiments for sodium (T, g 2573 K) [63] indicate a parabolic coexistence curve, but the more accurate results for rubidium (T, 2017K), cesium (T, N 1924K) [27,28] and mercury (T, N 1751 K) [29] ensure an Ising-like

On the other hand, king behavior is observed for BiC13 (T,

Page 22: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

8 HERMANN WEINGARTNER AND WOLFFRAM SCHROER

- . . 1 10 1 00 1000

T-T, , K

Figure 1. Log-log plot of the density difference d L - dc between the liquid and vapor phase of ammonium chloride [34] and bismuth chloride [60] versus the temperature distance T - T, from the critical temperature T,. For comparison, data are also shown for xenon and ammonia. The slope of the straight line for NH&1 is f l = 0.5. The slopes of the other lines are

= 0.326. Redrawn with permission from M. Buback, Thesis, Karlsruhe 1969.

criticality. A rationale is that the electrons screen the Coulombic interactions between the cores to short range. Calculations of Maggs and Ashcroft [64] and Goldstein et al. [65] show the screened interactions to decay as rP6, just as in nonionic fluids.

Another interesting aspect of results for metallic fluids concerns the liquid-vapor asymmetry. The “law of the rectilinear diameter” implies that the diameter (& + &)/2 of the coexistence curve is a linear function of p. In contrast to nonionic fluids, where deviations from this law are detectable only in experiments of high accuracy [6-81, large departures occur in metallic fluids [27-291. This indicates substantial differences in the effective interaction potentials for the vapor and liquid phases [66, 671, obviously related to a transition from nonconducting to conducting states. We will come back to this issue, when considering results for electrolyte solutions, where large differences are found in the conductances of the coexisting phases [68] as well.

B. Liquid-Liquid Demixings in Binary Electrolyte Solutions

1. Forces Driving the Phase Separation

In discussing the results for NH4C1 of Buback and Franck, Friedman [69] suggested in 1972 that some electrolyte solutions with liquid-liquid demix-

Page 23: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

CRITICALITY OF IONIC FLUIDS 9

ing near room temperature, as discovered by him earlier [3], may be more suitable for studies of ionic criticality than molten salts. Although already ob- served in 1903 for Kz+S02 [2], for a long time the examples for liquid- liquid immiscibilities in electrolyte solutions remained rare [3, 701. One dif- ficulty in finding such systems arises from the interference of crystallization, driven by high melting points of salts. Low-melting salts with large organic cations and anions enable the systematic design of more suitable systems [7 1, 721. The use of some apparently “exotic” systems in studies of ionic critical- ity is dictated by this need for low-melting salts.

The presence of the solvent may introduce a wide spectrum of short-range ion-solvent interactions. An intriguing hypothesis was that mean-field-like criticality is restricted to systems, where Coulombic interactions prevail [5, 35, 721. The critical parameters (6) may serve as a criterion for the dominance of Coulombic interactions, because the corresponding states principle can be extended to solutions, if EO in Eq. (4) is interpreted as the dielectric constant E , of the solvent [37]. One should, however, appreciate that, owing to uncertainties in this approximation, and also in estimating the diameters of large, internally flexible ions, a mapping of experimental critical parameters onto the reduced variables defined by Eqs. (4) and ( 5 ) can only be done in an approximate way. A general criterion evolving from such a reasoning is that “Coulombic” immiscibilities occur at low ion densities in solvents of low dielectric constant.

Such immiscibilities were, for example, observed for aqueous solutions comprising salts with multivalent ions such as BaC12 [73] or U02S04 [74] at high temperatures, where E, is low. Because these systems were never applied in studies of ionic criticality, we restrict the discussion to salts with univalent ions.

For salts with univalent ions, Eq. (4) predicts critical points near room temperature for systems with E , P+ 5 [72]. Liquid-liquid immiscibilities in several electrolyte solutions are known to satisfy this criterion [5 , 7 1, 721. Note that these gaps do not necessarily possess an upper critical solution temperature (UCST). Theory can rationalize a lower critical solution tem- perature (LCST) as well, if the product E,~T decreases with increasing temperature.

The critical parameters (6) exclude immiscibilities in aqueous solutions near room temperature, where T * > 0.5. Nevertheless, liquid-liquid coexistence curves were found at such conditions for some tetraalkylammo- nium salts with large cations and large anions [75-771. In early debates of ionic criticality, this observation led to some confusion. Originally studied near the UCST [77], such gaps later proved to be closed loops with the LCST suppressed by crystallization [72,78]. In one case, the LCST could be reached [79, 801.

Page 24: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

10 HERMANN WEINGARTNER AND WOLFFRAM SCHROER

Due to the obvious non-Coulombic nature, the mechanism of these transitions was investigated in great detail [72,76,81], also with regard to an unusual pressure dependence of the miscibility gaps [78, 80, 821. All evidence suggests that the hydrophobic nature of the cation in an aqueous environment drives these immiscibilities, just as in many aqueous solutions of hydrophobic nonelectrolytes. This interpretation is confirmed by an analysis of the solution thermodynamics in terms of a semiempirical EOS [76] and more refined statistical mechanical calculations [& 11. Some well-studied aqueous solutions of weak organic electrolytes such as isobutyric acid + water [6,8] seem to represent such hydrophobic demixings as well.

In passing, we note that such hydrophobic interactions enforce cation- anion and cation-cation pairing [72,76,81]. In contrast to expectations from electrostatic arguments, cation-anion pairing increases with increasing ion size- for example, in the anion series F- < C1- < Br- < N03- < I - < C104-. In biophysical chemistry, this anion series is known as “Iyotropic series” or “Hofmeister series” [83 J, which governs salting-out effects and many other properties of biomolecules in aqueous solutions [84]. The tendency for hydrophobic phase separation increases along this anion series as well [72].

The dichotomy of Coulombic versus hydrophobic interactions in driving criticality was probably first clearly stated in Ref. 76. Becoming aware that mechanisms of the hydrophobic type also occur in some nonaqueous solvents of high cohesive energy density e.g., in ethylene glycol, formamide, etc., Weingartner et al. [72] created the term solvophobic dernixing, which is now widely in use for discriminating these phase equilibria from Coulombic transitions [5, 15, 171.

Another subtle case, where specific interactions may obscure the effects of Coulombic criticality, is ethylammonium nitrate (EtNH3N03) + 1-octanol (T, E 315 K) [85]. In contrast to all other known examples, the critical point is located in the salt-rich regime at a critical mole fraction of X, E 0.77. Electrical conductance data indicate strong ion pairing, presumably caused by a hydrogen bond between the cation and anion which stabilizes the pairs in excess to what is expected from the Coulombic interactions [85]. This warns that, beyond the Coulombic/solvophobic dichotomy widely discussed in the literature, additional mechanisms may affect the phase separation [5].

2. Coexistence Curves

Expansion of the asymptotic power law for the order parameter in terms of a Wegner series (3) leads to [43]

M = BoTP (1 + B1 fa + B , P + . ’ .) . (7)

Page 25: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

CRITICALITY OF IONIC FLUIDS 11

Note that the correction amplitudes Bj are not independent, but related to B1 through B2 rx Bt , B3 c( B:, and so on [43]. Moreover, their signs are well- defined: A converging series requires B1 > 0 followed by alternating signs for the higher terms [86]. One should therefore be careful when trying to assess the physical significance of the sign and magnitude of correction amplitudes obtained in fits with freely adjustable amplitudes.

It is usually stated that there is some ambiguity in the order parameter in Eq. (7), because, on thermodynamic or experimental grounds, many choices exist, and there is a priori no reason for selecting a preferred variable [8]. Experience for nonionic fluids shows that all reasonable choices (e.g., the mole fraction or volume fraction) yield the same value of p [6,8]. In contrast, the size of the asymptotic range, the numerical values of the Wegner corrections, and the behavior of the diameter of the coexistence curve depend on the choice of M . If one accepts, however, the principle of isomorphism between one- and two-component systems, one can argue that, among all variables of practical relevance, the volume fraction 4; of component i is the best choice [58]. In fact, in terms of 4;, a symmetrization of coexistence curves is often achieved [6, 81. According to these arguments, the mole fraction Xi is only suitable in data analysis, when the molar volumes of the components are of similar size. The refractive index n, which often forms the primary experimental quantity, or better the Lorenz-Lorentz function (n’ - l ) / (n2 + 2), may also form reasonable choices, because both are almost linearly related to 4i. However, other views can be found [8].

Figure 2 shows an example of the order parameter dependence of the coexistence curve data for tetra-n-butylammonium picrate (Bu4NPic) + l-dodecanol[87] plotted in terms of mole and volume fractions, respectively. If represented in the mole fraction scale, the phase diagram is highly skewed and located in the solvent-rich regime. As emphasized by Fisher [15], such highly skewed phase diagrams resemble those of polymer solutions in poor solvents [88], suggesting that we look for theoretical analogies. The volume fraction leads to a more symmetric phase diagram, and the asymptotic range becomes larger. No variable was found that generates a fully symmetric coexistence curve [87].

Table I1 summarizes the existing studies of ionic criticality and lists the critical parameters. In the following, we will focus on results for immis- cibilities which seem to be primarily driven by Coulombic interactions, as exemplified by Pitzer’s system n-hexyl-triethylammonium n-hexyl-triethyl- borate (HexEt3N+HexEt3B -) + diphenylether [35], solutions of Bu4NPic in alcohols [87], and solutions of Na in NH3 [46].

We note that some coexistence curves of ionic systems with pronounced non-Coulombic interactions were also investigated with great care. These include aqueous solutions of tetraalkylammonium salts [77,79] and solutions

Page 26: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

12 HERMANN WEINGARTNER AND WOLFFRAM SCHROER

0.00

lh 0.05

0.10

0 0.2 0.4 +, 0.6 0.8 1

0.00

OO lh 0.05 0

0 0

0.10 1 0 1 P I

0.2 0.4 0.6 0.8 1 xi 0

Figure 2. Liquid-liquid coexistence curve of the system tetra-n-butylammonium picrate + I-dodecanol [87] plotted (a) in terms of the volume fraction 4 , and (b) in terms of the mole fraction XI of the salt.

of EtNH,NO, in l-octanol [89]. In these cases, plain Ising behavior was found. The same is true for many phase-separating aqueous solutions of organic compounds with OH, amine, and COOH groups that are slightly conducting, as exemplified by isobutyric acid + water. These are, however, usually considered as nonionic fluids; for reviews and references see Refs. 6, 8, and 90. Plain Ising behavior was also found in other cases: ethyl-trimethyl- ammonium bromide (EtMe,NBr) + CHCI, doped with ethanol [91] to sup- press crystallization; tetra-n-butylammonium naphtylsulfonate (Bu4NNS) + toluene [92]; n-heptyl-tri-n-butylammonium dodecylsulfate (HepBu,- NDS)+ cyclohexane [92, 931. The observation of Ising behavior in the latter systems is interesting, because the solvents are fairly inert, but in total, these systems are little characterized, and the role of specific interactions remains unclear. When specific interactions drive the phase separation, results may bear little relevance for the problem of ionic criticality.

Turning now to phase transitions primarily driven by Coulombic forces, Pitzer’s group [94] reported in 1985 experiments for Bu4NPic + l-chloro- heptane with an UCST at T, S 414 K. The experiments indicate a parabolic coexistence curve, but their limited precision gave rise to doubts [14, 771. Singh and Pitzer were therefore quick to search for a more suited Coulombic system [71], and in 1990 they reported a system comprising the low-melting salt HexEt3NHexEt3B dissolved in diphenylether with T, S 3 17 K [35].

Page 27: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

CRITICALITY OF IONIC FLUIDS 13

TABLE I1 Survey of Work on Liquid-Liquid Phase Separations in Binary Electrolyte Solutions

System " T,,K' Composition Method '

HexEt3NHexEt3B + diphenylether 3 17.69

Bu4NPic + I-chloroheptane Bu4NPic + I-tetradecanol Bu4NPic + 1 -tridecanol

BujNPic + I-dodecanol

Bu,NPic + 1-undecanol Bu4NPic + 1 -decanol Bu4NPic + 2-propanol Bu4NPic + 1 ,Z-propanediol Bu4NPic + I ,4-butanediol

EtNH3N03 + 1-octanol

EtMe3NBr + chloroform' Bu4NNS + toluene HepBuiNDS + cyclohexane Pe4NBr + H 2 0 Bti3PrNI + H 2 0 (UCST) Bu3PrNI + H20(LCST) Na + NH3

311.18 294.82 308.69 3 12.49 309.50 4 14.4 35 I .09 341.73 342.35 341.79 34 I .85 335.91 334.8 331.99 334.38 326.53 31 8.29 3 18.29 322.19 332.43 33 1.90 315.19 314.94 313 302.3 1 " 313.56 298.84 319 356 404.90 345.72 33 1.72 23 I S.5 23 1.55

X , = 0.053 X,. = 0.049 X , = 0.049 X,. = 0.049 X , = 0.049 X , = 0.049 X,. = 0.085 MI,. = 0.272 w, = 0.298 M', = 0.284 w, = 0.298 w,. = 0.298 w,. = 0.295 w,. = 0.298 IV, = 0.304 w, = 0.304 WJ, = 0.309 w,. =0.315 w, = 0.350 w, = 0.410 IV, = 0.469 IV,. = 0.469

X,. = 0.766 X,. = 0.760 X , = 0.760 X,. = 0.766 X , = 0.053 X , =0.106 X,. = 0. I40 X , = 0.0305 vv' = 0.4371

X , = 0.0415

x, =0.773

M.',. = 0.4457

x,.=0.0415

C [35,71] C 1961 T 1951 T,S [951, D 11071 v [125] 111341

C [87], T,S [93], D [ I l l ] S,D [971 C [87], T,S [93], D [111] T [108-110] v [I221 C [87]. T,S 1931, D [ I 11 1 C [891

E [133]

C [87], T,S 1931, D [ I 111 C 1871, T,S 1931, D [ I I I] T [108-1l01 T [108-1101 E [I331 T,S,D [ I 161 C [891 S [I181 N [I171 V 11261 C,S P I 1 C S [921 C,S [92,11 I] C [771 C [ W , C,S,D,T I51 C P91, C,S,D,T [51 C 1461 N [112,113]

c 1941

T [l08-ll0]

T [l08-110]

Abbreviations: Me, methyl; Et, ethyl; Pi-, n-propyl; Bu, n-butyl; Pe, n-pentyl; He, n-hexyl; Hep, n-heptyl; Pic, picrate; NS, naphtylsulfonate; DS, dodecyl-sulfate. 'Some samples show a temperature drift with time, so that T, is only approximately given, or there are different samples with different T, values within a series of measurements. For details we refer to the original papers. 'Abbreviations: C, coexistence curve; S, static light scattering; D, dynamic light scattering; T, turbidity; N, coherent neutron scattering; V, shear viscosity; E, electrical conductance; I, interfacial phenomena (ellipsometry).

Deuterated 1-octanol. Sample contained 1 wt% ethanol.

Page 28: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

14 HERMANN WEINGARTNER AND WOLFFRAM SCHROER

cl c 0.3 ! I I I I

0.0001 0.0005 0.001 0.005 0.001

F- Figure 3. Effective exponent Peff of Pitzer’s system n-hexyl-triethylammonium n-hexyl-

triethylborate + diphenylether. Curves a and b are derived from Singh and Pitzer’s data presuming asymptotic mean-field behavior and asymptotic king behavior [35] , respectively. Curve c is derived from the data of Wiegand et al. [96].

Refractive indices of the coexisting phases were measured at T 2 2 x lop5. Figure 3 shows the effective exponent Peff = d In Mld In T , defined by the local slope in the double-logarithmic plot of M versus T. At f > lop4 the behavior is mean-field-like. At lower T, the data are consistent with mean-field behavior; but at the cost of large amplitudes of the Wegner corrections, they can also be forced to reflect asymptotic Ising behavior. We anticipate here that Sing and Pitzer’s results were immediately confirmed by the turbidity experiments of Zhang et al. [95].

Due to the extraordinary implications of these observations in these two studies, Pitzer’s system was reexamined with regard to the coexistence curve and to other critical properties. Wiegand et a]. [96] remeasured the refractive indices of the coexisting phases in the range lop4 < T < 0.04. The results disagree with Singh and Pitzer’s data, yielding p = (0.34 k 0.01) close to the king value. Crossover is not detectable. Figure 3 compares the effective exponents resulting from these data with Pitzer’s original results. This brings, of course, the status of Singh and Pitzer’s data into question.

We note that Pitzer’s value of T, E 317 K could not be reproduced in later work which, depending on the sample, yielded values between 288 and 309 K [95,96]. These differences seem to indicate a chemical instability of the salt, so that decomposition products displace T,. Weingartner et al. [72, 971 suggested that solutions comprising Bu4NPic as a low-melting salt may behave more reproducibly, because the salt is chemically more stable.

Extending earlier work [72], Schroer, Weingartner and coworkers [87] were able to observe the critical points of Bu4NPic in a homologous series of alcohols with dielectric constants between 16.8 (2-propanol) and 3.6

Page 29: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

CRITICALITY OF IONIC FLUIDS

8 4 -

Or 2 -

z - . I-"

0

15

2-alcanols 0

0 . I-alcanols =. . .

Coulombic limit . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

(1-tetradecanol). Figure 4 shows the critical temperatures as a function of the chain length n of the alkyl residues. For a given salt, Eq. (4) predicts the product T, . E , to be constant; one finds T, . E, E 800-1000. Figure 4 shows that, when increasing the apolar part of the molecule, one clearly moves from a distinctly non-Coulombic to an essentially Coulombic mechanism for phase separation.

Some of these systems were subsequently employed in precise studies of the coexistence curves [87]. Figure 5 shows the effective exponents Beff based on volume fractions derived from these data. When reducing the solvophobic character of the systems by increasing the chain length n of the alcohols, deviations of Beff from the Ising value become visible which show a trend toward the mean-field value.

The results substantiate earlier observations for the liquid-liquid phase transition of Na + NH3. This system shows a transition to metallic states in concentrated solutions; but in dilute solutions and near criticality, ionic states prevail [98], and the gross phase behavior seems to be in accordance with a Coulombic transition [37 ] . Crossover was found at T = [46], and it seems to be much more abrupt than in the picrate systems. However, much depends on the subtle details of the data evaluation. Das and Greer [99] could smoothly represent the data by a Wegner series.

Page 30: Advances in CHEMICAL PHYSICS · 2013-07-23 · Advances in CHEMICAL PHYSICS Edited by I. PRIGOGINE Center for Studies in Statistical Mechanics and Complex Systems The University of

16 HERMANN WEINGARTNER AND WOLFFRAM SCHROER

0.30

0.38

= cl“ 0.34

0.30

0.38 1

1-dodecanol 0.32 0.30 I

0.38 1 I

1-tetradecanol 0.32 0.30 I

0,001 0.01 0.1 - T

Figure 5. Effective exponents Peff of tetra-n-butylarnrnoniurn picrate dissolved in several alcohols [87] as a function of the reduced temperature.

The diameter ( M I + M 2 ) / 2 of the coexistence curve is represented by the series [loo, 1011

involving the exponent cz of the heat capacity. The linear term contains both a critical contribution and a regular contribution. Moreover, there is a spurious 28 contribution [8, 1021, when an improper order parameter is chosen in data evaluation. In a mean-field system with cz = 0 and p = 1/2 the diameter is “rectilinear.” For a long time, the weak diameter anomaly in nonionic systems was a matter of controversy, because the deviations from rectilinear behavior are small [6, 81, and the 2p contribution obscures the results.