Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

A new perspective on aqueous electrolyte solutions

Gerhard Schwaab * and Simone Pezzotti *
Department of Physical Chemistry II, Ruhr-University Bochum, Bochum, Germany. E-mail: gerhard.schwaab@rub.de; simone.pezzotti@ens.psl.eu; Tel: +49 234 3224256

Received 13th May 2025 , Accepted 10th July 2025

First published on 15th July 2025


Abstract

Aqueous electrolyte solutions are central to many natural phenomena and industrial applications leading to continuous development of increasingly complex analytical models. These are based on an atomistic description of electrostatic interactions between ions, along with mean-field approaches for the dielectric response of water. Despite many achievements, such concepts often fall short in quantitatively describing scenarios where ion–ion correlations and specific solvation effects become relevant, particularly in concentrated electrolyte solutions. Here, we propose a shift in perspective, by introducing a statistical, coarse-grained approach to describe the average thermodynamic properties of aqueous electrolyte solutions. This method eliminates the need to define ion pairs or ion complexes and does not require any prior knowledge on specific solvation. We base our concept on separating the solution into a spherical observation volume whose size and average composition are uniquely determined by the solution parameters, and its environment, which consists of the remaining solution. This separation allows us to express the volume–environment interaction in terms of a generalized multipole expansion, i.e. in a convenient, additive way. We applied this approach to 135 electrolytes including some notoriously complex species, such as LiCl or ZnCl2 over their full solubility ranges. This paves the road toward understanding super-saturated and water-in-salt solutions and electrolyte nucleation.


1 Introduction

A variety of technological challenges such as the understanding of water in salt electrolytes (WISE),1–4 the development of advanced battery and energy storage technologies,5–9 the recycling of desalination brines10 and a save operation of deep-sea boreholes11 require a thorough microscopic understanding of concentrated electrolyte solutions. However, existing heuristic descriptions of their average excess thermodynamic properties have been derived by extrapolating from the diluted regime, leaving a gap of knowledge for concentrated solutions.12

From a physicochemical perspective, the osmotic and average activity coefficients of electrolytes, ϕ and ln[thin space (1/6-em)]γ±, respectively, determine the excess thermodynamic functions of the corresponding aqueous and non-aqueous solutions.13,14 Debye and Hückel were the first to describe electrolyte and water properties in dilute electrolyte solutions as a function of ion concentration and electrolyte composition.15 The theory has later been extended to higher concentrations by Bjerrum, Glueckauf, McMillan, and Mayer (see review by Vaslow16). Friedman, Pitzer, and coworkers extended the description to more complex electrolyte mixtures such as seawater.17 A special issue of the Journal of Fluid Phase Equilibria18 celebrates the 100th anniversary of the findings by Debye and Hückel and provides a summary of recent developments on the topic. In that issue, Simonin and Bernard19 compare several simple activity models including Debye–Hückel, the mean spherical approximation, and the Pitzer approach. Earlier, Khan et al.20 compared four physical descriptions of ln[thin space (1/6-em)]γ±. While the Pitzer and Bromley approach accurately describes the activity coefficients of 1[thin space (1/6-em)]:[thin space (1/6-em)]1 electrolytes, the method fails for several important 1[thin space (1/6-em)]:[thin space (1/6-em)]2 electrolytes, such as CaCl2 and MgCl2. A recent review by Held12 provides a critical comparison of the different excess Gibbs energy parametrizations up to high electrolyte concentrations.

Common to all these descriptions is a series expansion of ln[thin space (1/6-em)]γ± in the molality, the molarity, or the ionic strength framework starting from the limit of infinite dilution. In most cases, the dilute limit is described by a Debye–Hückel term. Each series expansion takes into account individual ion–ion interactions over the full volume in the configuration integral. The behavior at high molalities is in general derived by assuming ion-specific (hydrated) ion radii which are fitted to reproduce the experimental activity coefficients.17,19 The ability of ions to form different types of ion pairs or, more general, ion complexes makes an evaluation at high concentrations difficult.

Recently, a combined X-ray and simulation approach21–23 showed that at high concentrations, beyond the so-called Kirkwood-transition, ion–ion-correlations increase due to clustering, and the Debye length is no longer the characteristic length-scale. The critical concentration for the Kirkwood transition depends strongly on the electrolyte composition.

In spite of these complexities, it is very surprising that many experimental observables of aqueous electrolyte solutions, such as the effective molar extinction coefficients24 or the average apparent molar volume, show a nearly linear mol fraction dependency.25 This simple behavior suggests that, on a macroscopic level, most of the water-mediated ion–ion interactions partially compensate and lead to average interactions that can be described by simple analytical functions.

In the following, we focus on two-component solutions composed of a single electrolyte and water. We demonstrate that the separation of the solution into a well-chosen, uniquely defined, probe volume and its environment leads to a generalized multipole description of the excess interaction energy of electrolyte solutions. The number of required expansion orders is small: even for complex electrolytes such as ZnCl2 and LiCl three components are sufficient to describe the experimental data. When integrated, the resulting equation yields an analytical form of the osmotic coefficient ϕ and, thus, the water activity aw. The set of equations is applied to a total number of 135 electrolytes. The dependency of the fit parameters on electrolyte composition is discussed.

2 Introducing the statistical approach

The mean activity coefficient expresses the excess chemical potential of the electrolyte in units of RT. Activity coefficients, γ±, as a function of molality, mB, are reported for a large number of salts in two books by Lobo.26 To describe γ± based on a microscopic picture, our idea is to split the total electrolyte system into a spherical “observation” volume (d) with radius, Rd, and the remaining environment (e). The excess chemical potential is hence a function of the interaction free energy, Ude, between the two (see Fig. 1). We choose the observation volume to contain exactly one electrolyte unit (ν+ cations and ν anions) and the stoichiometric amount of water (i.e., the number of water molecules per ν+ cations and ν anions)
 
Nw = xw/xB(1)
where xB and xw are the electrolyte and water mol fractions, respectively. This definition is uniquely determined by the molality of the solution and its density. The spherical shape is adopted due to the isotropic nature of bulk electrolyte solutions.

image file: d5cp01781e-f1.tif
Fig. 1 Schematics of the coarse-grained approach used to determine the interaction energy in aqueous electrolyte solutions. (A) The solution is separated into a uniquely defined central “observation volume” with radius, Rd (black line), and its environment. All distances R = nRRh are measured in units of the hydrated electrolyte radius, Rh (dashed circle at the center), so that nd = Rd/Rh identifies the volume's radius in this coordinate system. In a hypothetical solution without ion–ion interactions, the charge distribution inside and outside the volume is uniform. When switching on ion–ion interactions, the solution averaged charge distribution inside the central volume produces an electrostatic potential in the local environment which attracts charges of the opposite sign. With increasing electrolyte concentration, Rd decreases, ion–ion interactions become more complex, and the interaction energy increases. (B) Example how, at high concentrations, dipole, quadrupole, and octupole interactions contribute to a complex interaction pattern between the observation volume and its environment. Please note that the charge distributions inside and outside the observation volume include both the ionic and the water contributions.

This observation volume is illustrated in Fig. 1A. The left scheme pictures a hypothetical solution without ion–ion interactions, where the charge distribution inside and outside the probe volume is uniform. In the low concentration limit (center frame), the presence of ion–ion interactions causes a non-homogeneous charge distribution that polarizes the surrounding environment. As discussed thereafter, this is well described by dipolar interactions (see Fig. 1B). With increasing concentration, the radius of the probe volume decreases due to its stoichiometric definition, while the charge distribution inside and outside the volume becomes increasingly complex and requires higher-order multipole interactions.

It is important to note that our observation volume is a purely statistical entity, defined to be charge neutral and containing on average the smallest possible, stoichiometric number of water and ions. Hence, its size, charge, and average composition are unrelated to the important spatial features, such as inhomogeneities in ion distribution and ion pairing, and the consequences of these on the water network observed, e.g., from simulations or diffraction experiments. However, the heterogeneity in the distribution of such configurations contributes, in our approach, to the instantaneous charge distribution inside the observation volume, due to both ions and water. This is a key feature of our approach, since water actively contributes to the charge distribution within the volume. This contribution is expected to be essential at high concentration, where specific ion solvation and water network arrangements dominate the free energy changes. These effects are, hence, naturally included in our charge distributions, which can be complex for strongly interacting ions where long-range ion–ion correlations and ion-clustering become important (see Fig. 1B).

The resulting charge distribution generates an electrostatic potential in the environment of our observation volume. The average interaction of this potential with the charge distributions outside the observation volume (which also depends on both ions and water), averaged over all possible configurations explored by the system, determines the interaction energy Ude. We choose to describe such an electrostatic potential at a distance, R, outside the volume as a multipole expansion in spherical coordinates. We express this distance, R = nRRh, in units of the hydration radius, Rh, of the electrolyte. It is determined by the composition of the solution, its density at constant temperature and pressure, and an effective number Nh of hydration water that depends on the electrolyte and the expansion order (see Appendix for details). This choice allows us to conveniently express the interactions in terms of a dimensionless distance unit, nR, removing the dependence on concentration and molar volume.

In a second step, we express the angular dependency of the effective charge at a distance R > Rd in terms of spherical harmonics (see Appendix for details). Thus, for each multipole term of order l, the integration over the angles leads to a contribution Ude,l(nR) = Ulwl(nR) where Ul is the lth order interaction energy. The weighting function, wl(nR), describes the radius dependency of the interaction strength. The total interaction energy of the observation volume with its environment is given by

 
image file: d5cp01781e-t1.tif(2)

Please note that the integration starts at the volume boundary (nd = Rd/Rh), which is solely determined by the composition of the solution and the hydration shell size of the electrolyte and is always positive.

The weighting function wl(nR) must satisfy the following conditions: (a) it must take into account the volume effect, i.e., the increasing number of charges interacting with the probe volume (b) the interaction strength due to the screening between the volume surface and nR must decay rapidly enough for eqn (2) to be integrable (c) it must take into account the fact that the charge distribution outside the probe volume is not purely random due to charge–charge interactions and hydrogen bonding, which limits the screening efficiency. (d) Different interaction orders require different screening lengths (e) all orders must vanish towards infinite dilution (i.e. nd → ∞), (f) its integrated form must be consistent with the experimentally observed activity coefficients. We found heuristically (see Appendix for details) that a weighting function of the form

 
image file: d5cp01781e-t2.tif(3)
fulfills all these requirements and allows for a good representation of the activity coefficients for 135 electrolytes in their whole solubility range. By adopting this, the integration yields
 
image file: d5cp01781e-t3.tif(4)
where we have converted the microscopic interaction energy Ude to molar quantities and normalized by the thermal energy RT. In addition, we have made use of the fact that
 
nd3 = (Rd/Rh)3 = xh/xB(5)
defines a ratio of the mol fraction of the hydrated electrolyte, xh and xB. This quantity is inversely proportional to the ratio of the observation volume and volume of the hydrated electrolyte, and, thus, to their mol fraction ratios (see Appendix for details). Dl, xlh, and λl are electrolyte dependent fit parameters. Dl characterizes the depth of the lth order interaction energy profile. The parameter xlh describes the crossover from the dilute to concentrated solutions. For xB > xlh, the lth order contribution becomes positive, indicating unfavorable (endothermic) interaction. We found in our analysis (see figures below and Tables 1 and 2) that solely the dipolar (l = 1) contribution requires in some cases x1h < 1. In all fits where quadrupole and octupole contributions become relevant, xl>1h = 1 leads to a satisfying description of the experimental data.

Table 1 Fit parameters for ln[thin space (1/6-em)]γ, part 1
Salt x h,Dipole D Dipole λ Dipole D Quadrupole λ Quadrupole D Octupole λ Octupole
NH4Br 0.348(26) 1.453(7) 0.517(12)
NH4Cl 0.370(9) 1.586(3) 0.537(4)
(NH4)2HPO4 1 8.866(25) 0.5422(12)
NH4NO3 1 2.840(23) 0.5591(22) 5.2(5) 1.91(6)
NH4ClO4 1 3.234(26) 0.576(3)
(NH4)2SO4 1 6.10(3) 0.4969(31)
NH4SCN 1 2.025(8) 0.5076(17)
BaBr2 0.05334(20) 2.287(5) 0.5449(16)
BaCl2 0.0836(15) 2.607(8) 0.527(3)
BaOH2 0.071(20) 3.27(18) 0.586(17)
BaI2 0.02945(9) 1.848(8) 0.5648(27)
Ba(NO3)2 1 6.72(10) 0.516(4)
Ba(ClO4)2 0.04917(11) 2.122(7) 0.5292(29)
CdBr2 0.926(18) 9.843(15) 0.4500(13)
CdCl2 0.034(7) 4.47(24) 0.597(12) 70(10) 1.354(13)
CdI2 0.0125(21) 5.08(26) 0.629(8) 190(20) 1.314(10)
Cd(NO3)2 0.0089(3) 1.387(16) 0.6115(19) 53(2) 1.2504(21)
Cd(ClO4)2 0.000886(7) 0.5475(14) 0.6692(11) 178(3) 1.1321(17)
CdSO4 0.0127(7) 3.69(5) 0.522(11) 67(7) 1.062(24)
CaBr2 0.00217(8) 0.788(14) 0.635(6) 90(5) 1.116(13)
CaCl2 0.000328(25) 0.364(17) 0.830(16) 1100(200) 1.282(17) 1600(200) 2.654(15)
CaI2 0.00160(5) 0.697(7) 0.647(4) 112(6) 1.125(7)
Ca(NO3)2 0.027(3) 1.87(4) 0.539(12) 17(3) 1.16(3)
Ca(ClO4)2 0.00166(6) 0.713(9) 0.645(6) 111(7) 1.129(11)
CsAc 0.03935(7) 0.7497(20) 0.5874(29)
CsBrO3 1 3.35(3) 0.5796(23)
CsBr 0.463(5) 2.1458(20) 0.5738(14)
CsClO3 1 3.37(4) 0.5806(26)
CsCl 0.398(5) 2.0339(18) 0.5742(17) 1000(1000) 6.8(9)
CsF 0.06227(15) 0.9032(14) 0.5902(16)
CsOH 0.0468(5) 0.7673(19) 0.5792(31)
CsI 1 2.582(8) 0.5438(13)
CsNO3 1 4.61(9) 0.655(7)
CsClO4 1 4.12(6) 0.602(3)
Cs2SO4 0.0145(22) 1.90(10) 0.627(5) 63(6) 1.301(12)
ChBr 0.540(7) 3.134(3) 0.6140(19)
ChCl 0.2243(14) 2.1110(20) 0.6269(19)
CrCl3 0.03002(21) 3.358(8) 0.557(5)
Cr(NO3)3 0.03592(31) 3.504(9) 0.535(6)
Cr2(SO4)3 0.080(5) 10.85(4) 0.508(13)
CoBr2 0.02628(16) 1.904(16) 0.609(4)
CoCl2 0.03964(10) 2.126(8) 0.5664(22)
CoI2 0.02031(23) 1.902(32) 0.632(6)
Co(NO3)2 0.00772(27) 1.252(16) 0.598(5) 36(3) 1.131(14)
Co(ClO4)2 0.001004(18) 0.555(5) 0.6455(19) 123(2) 1.092(4)
CoSO4 1 9.97(12) 0.454(4)
CuBr2 0.0334(6) 1.926(4) 0.5565(27) 130(20) 2.56(9)
CuCl2 0.050(10) 2.34(7) 0.548(16) 27(1) 1.81(16)
Cu(NO3)2 0.0130(9) 1.429(31) 0.574(7) 21(3) 1.097(27)
Cu(ClO4)2 0.001184(13) 0.595(3) 0.6439(17) 114(2) 1.097(3)
CuSO4 0.468(30) 9.35(4) 0.451(5)
Gdn2CO3 1 8.388(17) 0.6014(9)
GdnBr 1 2.38(6) 0.531(6) 2.83(3) 1.220(21)
GdnCl 1 1.84(12) 0.486(12) 2.92(10) 1.002(22)
GdnI 1 1.86(8) 0.486(9) 3.08(5) 1.136(22)
GdnNO3 1 3.34(9) 0.583(6) 15(4) 1.80(11)
GdnClO4 1 2.1(8) 0.51(6) 4.58(23) 1.09(16)
HBr 0.00272(4) 0.2921(27) 0.6857(24) 37.7(7) 1.190(6)
HCl 0.009(8) 0.36(24) 0.66(4) 7(1) 1.03(21)
HF 0.011(4) 5.5(4) 0.522(19) 140(40) 1.189(16)
HI 0.02000(8) 0.5701(7) 0.6002(15) −90(10) 2.93(7)
HNO3 0.07324(17) 0.8772(24) 0.5546(26)
HClO4 0.0381(4) 0.7332(7) 0.5889(19) −52(2) 2.527(26)
FeCl2 0.04413(20) 2.271(5) 0.5601(17)
k2succ 0.195(19) 2.583(15) 0.468(9)
LiAc 0.08644(31) 1.0147(19) 0.5774(20)
LiBr 0.0363(9) 0.788(6) 0.617(9) 800(300) 6.1(4) −60(10) 2.90(15)
LiClO3 0.0333(6) 1.07(12) 0.374(30)
LiCl 0.00261(14) 0.325(6) 0.753(11) 86(9) 1.319(14) 164(9) 3.61(6)
LiTFSI 0.0026 0.32 0.75 76(4) 1.27(3) 129(2) 3.1(1)
LiF 1 2.362(25) 0.523(4)


Table 2 Fit parameters for ln[thin space (1/6-em)]γ, part 2
Salt x h,Dipole D Dipole λ Dipole D Quadrupole λ Quadrupole D Octupole λ Octupole
LiOH 0.014(5) 0.78(10) 0.73(4) 40(20) 1.42(4)
LiI 0.001457(28) 0.2372(17) 0.729(4) 86(3) 1.254(6)
LiNO3 0.0353(10) 0.785(6) 0.6099(14) 3.93(26) 1.393(5)
LiClO4 0.00438(18) 0.316(8) 0.631(5) 14.7(8) 1.091(17)
Li2SO4 0.197(5) 3.536(9) 0.523(5)
MgBr2 0.00197(9) 0.742(18) 0.625(5) 79(3) 1.086(12)
MgCl2 0.00197(29) 0.74(5) 0.613(14) 75(7) 1.07(3)
MgI2 0.00144(17) 0.611(25) 0.647(11) 98(4) 1.096(22)
Mg(NO3)2 0.0034(8) 0.89(9) 0.544(20) 34(2) 0.98(5)
Mg(ClO4)2 0.01902(6) 1.510(5) 0.5686(23)
MgSO4 0.0060(12) 3.04(14) 0.572(18) 130(30) 1.120(29)
na2succ 0.0700(14) 2.144(5) 0.500(4)
KAc 0.04351(8) 0.7879(20) 0.5904(23)
KBrO3 1 3.49(8) 0.599(6)
KBr 0.2433(28) 1.4285(21) 0.5513(21)
KClO3 1 3.67(12) 0.624(10)
KCl 0.2797(31) 1.5296(14) 0.5557(21)
KF 0.0152(4) 0.605(4) 0.6420(23) 14.8(6) 1.235(6)
K2HPO4 0.45(6) 4.78(12) 0.518(6)
KH2PO4 1 4.62(8) 0.655(7)
KOH 0.0476(5) 0.945(18) 0.714(8)
KI 0.1878(19) 1.2326(20) 0.5417(21)
KNO3 1 2.92(7) 0.564(4) 5.28(15) 1.336(32)
KClO4 1 3.37(4) 0.5806(26)
K2SO4 1 6.18(10) 0.498(5)
KSCN 0.606(23) 1.739(4) 0.514(4)
RbAc 0.04085(5) 0.7767(13) 0.5922(14)
RbBrO3 1 3.02(4) 0.5672(30)
RbBr 0.441(11) 1.809(4) 0.5446(29)
RbClO3 1 2.91(31) 0.547(26)
RbCl 0.354(5) 1.7027(30) 0.5507(29)
RbF 0.0207(26) 0.717(27) 0.647(7) 18(2) 1.458(19)
RbI 0.402(7) 1.8067(26) 0.5518(23)
RbNO3 1 2.98(5) 0.5676(31) 4.65(4) 1.247(17)
RbClO4 1 3.883(28) 0.5968(17)
Rb2SO4 1 5.365(11) 0.4824(10)
AgNO3 1 2.93(5) 0.564(4) 5.20(3) 1.265(10)
NaAc 0.05354(13) 0.8300(13) 0.5769(19)
NaBrO3 1 2.863(24) 0.580(4)
NaBr 0.019(4) 0.54(8) 0.590(16) 5.3(5) 1.02(8)
NaClO3 1 2.015(8) 0.5114(17)
NaCl 0.0132(4) 0.548(7) 0.631(5) 14.5(9) 1.208(13)
NaF 0.49(4) 1.89(3) 0.551(5)
NaFo 0.1622(27) 1.1535(25) 0.547(4)
NaHCO3 1 2.602(21) 0.5446(28)
NaH2PO4 1 3.985(29) 0.644(4)
Na2HPO4 1 6.56(7) 0.537(4)
NaOH 0.00609(29) 0.476(8) 0.724(12) 50(6) 1.347(20) 240(30) 5.04(21)
NaI 0.00851(27) 0.472(5) 0.651(4) 19(1) 1.224(11) 3000(3000) 7.1(6)
NaNO3 0.000160(12) 0.0763(27) 0.859(12) 560(50) 1.261(15) 710(60) 2.321(26)
NaClO4 0.0358(16) 0.794(9) 0.6119(25) 8.9(4) 1.307(5)
Na2SO4 1 5.937(28) 0.5004(19)
NaSCN 0.0780(8) 0.957(7) 0.587(9)
SrBr2 0.001878(17) 0.7420(20) 0.6433(10) 108(1) 1.1316(15)
SrCl2 0.00382(4) 1.0243(29) 0.6354(10) 80(1) 1.1635(20)
SrI2 0.001399(10) 0.6480(14) 0.6468(8) 119(1) 1.1206(12)
Sr(NO3)2 0.038(9) 1.93(7) 0.503(19) 15(4) 1.10(4)
Sr(ClO4)2 0.0065(5) 1.15(3) 0.597(8) 30(5) 1.083(31)
ZnBr2 0.00339(25) 1.021(22) 0.678(9) 160(20) 1.323(7)
ZnCl2 0.00179(5) 0.830(7) 0.672(4) 191(7) 1.212(5) 243(9) 3.047(8)
ZnF2 0.148(8) 4.89(7) 0.5924(19)
ZnI2 0.0050(11) 1.10(7) 0.672(24) 130(30) 1.429(22)
Zn(NO3)2 0.0084(4) 1.18(3) 0.575(7) 23(2) 1.055(27)
Zn(ClO4)2 0.00094(13) 0.50(3) 0.609(8) 78(3) 1.004(19)
ZnSO4 1 10.126(24) 0.4191(9) −400(60) 3.05(6)


Fig. 2 shows example fits of individual electrolytes. For CsBr, which does not display strong cation–anion interactions, the dipole expansion represents the activity coefficient sufficiently well over the full data range. When increasing the cation–anion interaction strength, such as for the cases of LiCl and ZnCl2, complex ion–ion correlations occur in the solution.


image file: d5cp01781e-f2.tif
Fig. 2 Example fits of ln[thin space (1/6-em)]γ± for different electrolytes using the model description from eqn (4). The insets show the fit residuals.

In previous molecular dynamics and X-ray studies, these were shown to result in spatial inhomogeneities due to ion clustering.21–23 In our framework, the effect of these spatial inhomogeneities is encoded in the charge distributions within and outside the probe volume. With increasing ion clustering, the charge distributions become more complex, resulting in larger contributions from quadrupolar and octupolar terms in our model (see eqn (4)). Accordingly, the quadrupole and octupole terms of our statistical model capture the more complex behavior of LiCl and ZnCl2 solutions over the whole mole fraction range. The number of contributing multipole components provides a quantification of the impact of (multi-body) ion–ion correlations: complex ion–ion correlations require stronger contributions from an increased number of higher-order multipole expansion terms.

3 Discussion

Summarizing the results above, we have demonstrated that a multipole expansion of the interaction between the central observation volume and the surrounding solution is suitable to represent ln[thin space (1/6-em)]γ± (see eqn (4)). Each expansion exponent, λl, reflects the combination of the thermally averaged interaction between the multipole inside the observation volume and the charge distribution in its environment (Fig. 3). The interaction spatially decays as image file: d5cp01781e-t5.tif. The dipole–dipole interaction term (l = 1 in eqn (4)) compares well with the simplest form of the Debye–Hückel law (ln[thin space (1/6-em)]γ±mB1/2) for dilute solutions where xmB when choosing λDipole = 0.5 (see Fig. 3C).
image file: d5cp01781e-f3.tif
Fig. 3 Comparison of different electrolyte solutions. (A) Examples of the dipolar part of ln[thin space (1/6-em)]γ± for different electrolytes. The crossover from negative to positive values marks the transition from dipole–dipole to more complex interaction forms. 1/xB corresponds to the size of the probe volume. (B) Depth of the dipolar part of the potential versus exponent. Electrolytes with xDipoleh = 1 are shown with open markers. (C) Exponent λDipoleversus 1/xDipoleh, of the hydrated electrolyte. The dashed line represents the coefficient expected for the equivalent Debye–Hückel limiting law. (D) Amplitude DDipole for electrolytes with xDipoleh < 1. The data were fitted with a power law of the form image file: d5cp01781e-t4.tif with D0 = 2.81(8) and κ = 0.42(1). The lines display the power law for 1[thin space (1/6-em)]:[thin space (1/6-em)]1 (blue) and 2[thin space (1/6-em)]:[thin space (1/6-em)]1 and 1[thin space (1/6-em)]:[thin space (1/6-em)]2 (yellow) electrolytes. A few electrolytes with exceptional properties are labeled separately.

Fig. 3 shows the results of the dipolar contribution for 135 electrolytes. (See fit parameters in Tables 1 and 2 and the classification of the investigated electrolytes in Table 3 in the Appendix.) In panel (A), we show the dipolar contribution to ln[thin space (1/6-em)]γ± for four electrolytes with increasing interaction strength from CsI to LiCl. The crossover point xDipoleh (dashed vertical line) marks a boundary: dipole–dipole interaction becomes unfavorable, i.e., endothermic, for higher concentrations, and complex (i.e., quadrupole and octupole) interactions become more important. The radius of the observation volume where this happens is determined by 1/xDipoleh. With increasing ion–ion and ion–water interactions, the crossover shifts toward lower concentrations. This is in qualitative agreement with findings from a combined SAXS and simulation study (Fig. 3 in (ref. 22)) showing that a crossover from Debye–Hückel to more complex behavior occurs at lower concentrations for more strongly interacting ions.

Table 3 Electrolytes classified according to Fig. 3
Electrolyte class Members
1[thin space (1/6-em)]:[thin space (1/6-em)]1, xh < 1 ChBr, ChCl, CsAc, CsBr, CsCl, CsF, CsOH, HBr, HClO4, HF, HI, HNO3, KAc, KBr, KCl, KF, KI, KOH, KSCN, LiAc, LiBr, LiCl, LiClO3, LiClO4, LiI, LiNO3, LiOH, NaAc, NaBr, NaCl, NaClO4, NaF, NaFo, NaI, NaNO3, NaOH, NaSCN, NH4Br, NH4Cl, RbAc, RbBr, RbCl, RbF, RbI
1[thin space (1/6-em)]:[thin space (1/6-em)]1, xh = 1 AgNO3, CsBrO3, CsClO3, CsClO4, CsI, CsNO3, GdnBr, GdnCl, GdnClO4, GdnI, GdnNO3, HCl, KBrO3, KClO3, KClO4, KH2PO4, KNO3, LiF, NaBrO3, NaClO3, NaH2PO4, NaHCO3, NH4ClO4, NH4NO3, NH4SCN, RbBrO3, RbClO3, RbClO4, RbNO3
2[thin space (1/6-em)]:[thin space (1/6-em)]1 and 1[thin space (1/6-em)]:[thin space (1/6-em)]2, xh < 1 BaBr2, BaCl2, BaOH2, BaI2, Ba(ClO4)2, CdBr2, CdCl2, CdI2, Cd(NO3)2, Cd(ClO4)2, CaBr2, CaCl2, CaI2, Ca(NO3)2, Ca(ClO4)2, CoBr2, CoCl2, CoI2, Co(NO3)2, Co(ClO4)2, CuBr2, CuCl2, Cu(NO3)2, Cu(ClO4)2, FeCl2, K2HPO4, MgBr2, MgCl2, MgI2, Mg(NO3)2, Mg(ClO4)2, SrBr2, SrCl2, SrI2, Sr(NO3)2, Sr(ClO4)2, ZnBr2, ZnCl2, ZnF2, ZnI2, Zn(NO3)2, Zn(ClO4)2
2[thin space (1/6-em)]:[thin space (1/6-em)]1 and 1[thin space (1/6-em)]:[thin space (1/6-em)]2, xh = 1 Ba(NO3)2, Gdn2CO3, K2SO4, Na2HPO4, Na2SO4, (NH4)2HPO4, (NH4)2SO4, Rb2SO4
2[thin space (1/6-em)]:[thin space (1/6-em)]2, xh < 1 CdSO4, CuSO4, MgSO4, ZnSO4
2[thin space (1/6-em)]:[thin space (1/6-em)]2, xh = 1 CoSO4
3[thin space (1/6-em)]:[thin space (1/6-em)]1, xh < 1 CrCl3, Cr(NO3)3
3[thin space (1/6-em)]:[thin space (1/6-em)]2, xh < 1 Cr2(SO4)3)


Panel (B) shows the depth of the dipolar contribution to ln[thin space (1/6-em)]γ±versus its exponent for the different 1[thin space (1/6-em)]:[thin space (1/6-em)]1 to 3[thin space (1/6-em)]:[thin space (1/6-em)]2 electrolytes. We display electrolytes with xDipoleh = 1, i.e., where no crossover is observed, with open markers. This class contains both electrolytes that are well-described by the dipolar term, only, as well as special cases such as many guanidinium and sulfate salts for which quadrupolar contributions are important, despite no crossover being observed in the dipolar term. The vertical line at λDipole = 0.5 marks the value corresponding to the Debye–Hückel (DH) law. Electrolytes with large DDipole typically show exponents close to the DH-value. A comparison to panels (A) and (C) shows that DDipole is determined by both the strength of the electrolyte and the limiting concentration, where higher-order terms become more prominent. A larger exponent indicates an increased importance of thermal averaging effects (remember that thermal averaging changes the 1/R3-interaction of close dipoles to a 1/R6-dependency, when the effective interaction strength is much smaller than kT).

Panel (C) supports this picture. Here, we observe that (a) electrolytes with smaller 1/xDipoleh and (b) electrolytes with higher charge show exponents closer to the DH value. Electrolytes where higher order contributions dominate at low concentrations (i.e., xDipoleh is small) or which only weakly interact with each other and with water (e.g., 1[thin space (1/6-em)]:[thin space (1/6-em)]1, blue, vs. 2[thin space (1/6-em)]:[thin space (1/6-em)]1 and 1[thin space (1/6-em)]:[thin space (1/6-em)]2, yellow) tend to have larger exponents.

Panel (D) in Fig. 3 shows the interaction strength, DDipole, as function of 1/xh for 97 electrolytes. We observe a power law of the form DDipole = D0Im(1/xDipoleh)κ with D0 = 2.81(8) and κ = 0.42(1), where image file: d5cp01781e-t6.tif is the ionic strength of the electrolyte in our molecular frame. The proportionality of the interaction strength to Im is in agreement with the ionic strength dependency of the interaction energy as described by Debye and Hückel (see Appendix for a detailed comparison). We noticed a few exceptional cases which are, however, beyond the scope of our discussion.

Another interesting property of eqn (3) resulting from our model is that the weighting function (eqn (3)) is directly related to the microscopic radius-dependent partition function, Zl. The weighting function originates from the excess charge, Δq(nR), generated by the potential of the central observation volume at the given position. By assuming a Van’t Hoff-like equation for each multipolar expansion term, we can relate each Δql(nR) to a microscopic osmotic pressure Πl(nR)

 
image file: d5cp01781e-t7.tif(6)

Since the angular dependency of the charge difference is described by spherical harmonics, this radius dependency describes the changes along the major axes (e.g., along the dipole axis, for dipole–dipole interaction, see Fig. 1B). Integration of eqn (3) including the pre-factor Ul yields

 
image file: d5cp01781e-t8.tif(7)
so that λl and Ul determine the general shape and the sharpness of the partition function, respectively.

Fig. 4A, top shows the contributions to the partition function for LiCl using Ul = 1 for better comparability. The long-range interaction shows a gradual transition when hitting the crossover (dashed blue), where dipole–dipole interaction is replaced by higher-order, more complex interactions. In contrast, the quadrupole (yellow) and octupole (green) terms show a sharp decay at nR = 1.


image file: d5cp01781e-f4.tif
Fig. 4 (A) Microscopic insights from our new description: radius dependency of the contributions to the partition function (top) and the osmotic pressure (bottom) for LiCl. For clarity, we used the potential depth Ul = 1. The long-range dipole interaction shows a crossover from positive to negative values at nDipoleR = 7.3 (dashed blue) where nDipoleR = (1/xDipoleh)1/3. Purple: nR values for 1 m and 0.1 m solutions. (B) Example global fits of the osmotic and activity coefficients for LiCl (top) and ZnCl2 (bottom) using eqn (4) and (8). The insets show the fit residuals.

The corresponding osmotic pressures are shown in the bottom part of Fig. 4A. The dipole contribution is weak and only attractive (i.e., negative) at long distances. In contrast, the quadrupole and octupole terms are short-range and still attractive at the highest possible concentrations.

In addition, the description of ln[thin space (1/6-em)]γ± in eqn (4) allows us to derive an analytical form of the osmotic coefficient. The Gibbs–Duhem equation27 allows to convert ln[thin space (1/6-em)]γ± to the osmotic coefficient

 
image file: d5cp01781e-t9.tif(8)
which is closely related to the water activity
 
image file: d5cp01781e-t10.tif(9)
with n0 as moles of water in 1 kg of the solvent. When we describe molality as mB = n0xB/(1 − xB) and use eqn (4) to integrate eqn (8) we obtain
 
image file: d5cp01781e-t11.tif(10)
for the lth order contribution to the osmotic coefficient. Here, BxB(1 + λl, 0) is the incomplete Euler Beta function28 and Φ(xB, 2, 1 + λl) is the Lerch transcendent29 (see Appendix for details on the integration procedure and the special functions).

These analytical descriptions (eqn (4) and (8)) allow a direct global fit of experimental values of electrolyte activity, osmotic coefficient, and water activity and, therefore, simplify the retrieval of excess thermodynamic properties using different measurement techniques. Fig. 4B shows example fits for LiCl and ZnCl2 together with their fit residuals (insets). Please note that the oscillatory behavior in the residuals is an artifact, since the published data have been fitted by a combination of Debye–Hückel, Pitzer, and polynomial terms. In case of ZnCl2 Goldberg30 required 13 and 8 coefficients to reproduce ln[thin space (1/6-em)]γ± and the osmotic coefficient, respectively, while we need a total of seven identical fit parameters with a clear physical meaning (σlnγ± = 0.00983 and σϕ = 0.00589 vs. σlnγ± = 0.00747 and σϕ = 0.00684 in the work by Goldberg) for both physical properties. We have published the data sets used for the fit as well as the fit parameters and predicted values for ln[thin space (1/6-em)]γ±, ϕ, and aw separately in machine-readable form.31

Finally, we tested if the generalized multipole expansion approach applies to “water in salt (WISE)” electrolytes such as lithium-bis-(trifluoromethanesulfonyl)-imide (LiTFSI), see Fig. 5. Water activity data for this compound were reported recently by Zhigalenok et al.32 for intermediate to high LiTFSI concentrations. Due to a lack of low-concentration data, we fixed the dipolar set of parameters to those of LiCl to obtain a numerically stable fit. As shown by the residuals (black dots) in Fig. 5, our model is well suited to represent the water activity of LiTFSI over the full concentration range. As expected for systems with strongly inhomogeneous charge distributions,2 we must include the dipole plus quadrupole plus octupole terms.


image file: d5cp01781e-f5.tif
Fig. 5 Blue points: water activity data, ln[thin space (1/6-em)]aw (Zhigalenok et al.32) of the “water in salt” electrolyte LiTFSI as function of electrolyte mole fraction. Blue line: fit of ln[thin space (1/6-em)]aw based on our multipole expansion, which requires dipole plus quadrupole plus octupole terms. Due to the lack of low-concentration data, we used the LiCl parameter for the dipolar term to obtain a stable fit. Black dots: fit residuals.

4 Conclusions

We have demonstrated above a novel statistical approach to model and understand the excess thermodynamic functions, i.e., (ln[thin space (1/6-em)]γ±, osmotic coefficient, ϕ, and water activity, aw) of aqueous electrolyte solutions and applied it to 135 electrolytes. The foundation of our coarse-grained approach is a generalized multipole expansion which describes the interaction of all charges outside a purely statistical, stoichiometrically defined observation volume with the potential originating from the charge distribution inside. A key feature is that charge distributions inside and outside the observation volume depend both on ions and water. This allows us to take into account the properties of the hydration water network as a function of concentration beyond continuum model descriptions that reduce water's influence to the dielectric constant.

We summarize the key features of our description in Fig. 6: with increasing concentration, the charge distribution within our observation volume captures the emergence of spatial inhomogeneities, such as ion clustering. This becomes more complex with increasing concentration from panel (A) to (B), while still obeying, on average, the constraints of being charge-neutral by containing one electrolyte unit and the stoichiometric amount of water. The increased complexity of the charge distribution results in larger higher-order contributions to our multipole expansion.


image file: d5cp01781e-f6.tif
Fig. 6 Our simplifying statistical approach efficiently describes the thermodynamic properties of complex electrolyte solutions: (A) Weakly interacting ions or dilute solutions (e.g., CsBr, see Fig. 2) show a more homogeneous charge distribution (anions in red, cations in blue). Top: Illustration of typical instantaneous charge distributions within the observation volume. Center: Qualitative probability distribution of the instantaneous observation volume net charge. Regardless of the distribution, the average observation volume charge is zero by construction. Bottom: Illustration of instantaneous ion distributions inside and outside a selection of observation volumes (yellow spheres). The dipole term in our model is sufficient to describe the thermodynamics of these cases. (B) Same set of illustrations for complex electrolytes at high concentrations where ion clustering starts to become important (e.g., LiCl, ZnCl2, see Fig. 2). Even for these cases, the quadrupole and octupole terms in our model fully capture the effect of the increasing structural heterogeneity on the thermodynamics.

Even for the most complex studied electrolyte solutions, the 3rd-order expansion, including dipole, quadrupole, and octupole contributions, is found sufficient to describe the full set of thermodynamic properties in the whole solubility range. This is a substantial simplification compared to existing heuristic descriptions.

In addition, as summarized in Fig. 6, our statistical perspective allows us to collapse the effect of the complex structural heterogeneity found in experimental data and simulations for strongly interacting ions and concentrated solutions into the quadrupolar and octupolar terms of our thermodynamic functions. For example, recent studies suggest that a Kirkwood transition, i.e., the concentration at which the Debye length is no longer the characteristic length scale in an electrolyte solution, happens at lower concentrations for strongly interacting electrolytes due to ion clustering.22,23 This observation is in qualitative agreement with our findings in Fig. 2 and 8A, where quadrupole and octupole terms start to contribute significantly at lower concentrations for strongly interacting electrolytes.

Our universal approach is not restricted to binary electrolyte solutions but is generalizable to electrolyte mixtures and solutions in general. This process requires an additional summation over all possible (neutral) solute combinations in the observation volume and their interaction with the solute mixture in the environment. The change of perspective, we propose, offers new insights into the average thermodynamic properties of electrolyte solutions and their connection to local structural heterogeneity and ion complexes as they are characterized by state-of-the-art spectroscopic and simulation techniques. We hope these findings will contribute to the understanding of concentrated electrolyte solutions, which are relevant in many biological and electrochemical applications in today's society.

Author contributions

G. S.: conceptualization, data curation, formal analysis, investigation, methodology, software, visualization, writing – original draft, writing – review & editing. S. P.: conceptualization, writing – original draft, writing – review & editing.

Conflicts of interest

There are no conflicts of interest to declare.

Data availability

Data for this article, including the activity data used for the model fitting, the fit parameters as well as the resulting recommended activity coefficients, osmotic coefficients and water actvities are available at TUDOData (RESOLV-data) at https://doi.org/10.17877/RESOLV-2024-M4WGMZTG, 2025.

Appendices

Heuristic discovery of eqn (3) and (4)

Fig. 7 shows ln[thin space (1/6-em)]γ± for cesium bromide and sodium bromide as a function of molality (A), ionic strength (B), and the log-transformed coordinate rB = ln(1/xB) = ln(1 + Nw) (C). As defined above, Nw is the number of water molecules per electrolyte unit (i.e., per ν anions and ν+ cations). While ln[thin space (1/6-em)]γ± shows the well-known complex mol fraction and ionic strength dependency which is typically described by Debye–Hückel or Pitzer parameters, in the log-transformed coordinate system (panel (C)) ln[thin space (1/6-em)]γ± disappears for rB → ∞ and shows a curve similar to a Morse potential. Indeed, in this coordinate system ln[thin space (1/6-em)]γ± can be well represented over the full concentration range when using a small number (l ≤ 3) of fit functions of the form
 
image file: d5cp01781e-t12.tif(11)

image file: d5cp01781e-f7.tif
Fig. 7 Logarithm of the activity coefficient (ln[thin space (1/6-em)]γ±) for cesium bromide and sodium bromide in different coordinate systems: (A) molality, (B) ionic strength, and (C) rB = ln(1/xB). The dashed curve shows the Debye–Hückel limiting law for a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 electrolyte in the different coordinate systems.

A back-transform to mol fraction units recovers eqn (4).

This simple functional form led us to the conclusion that a (spherical) charge-neutral observation volume containing ν anions, ν+ cations, and Nw water molecules has a special meaning because it represents the statistical average over all possible configurations of that size. In addition, the question arose how the dipole and higher multipole moments of this observation droplet interact with the (again statistically averaged) excess charge distributions in the environment (eqn (2) and (21)). After removal of the angular dependencies, we chose the weight function such that the full integral recovered eqn (3).

Definitions of distances

We have introduced the hydrated electrolyte radius, Rh, in the main text without giving further specifications. We define this radius from density measurements26 and the composition of the solution as
 
image file: d5cp01781e-t13.tif(12)
where Nh is the hydration shell size (i.e., the number of water molecules per ν+ cations and ν anions) where the solvated ion complexes start to repel each other, and NA is Avogadro's constant. In cases where we could not obtain Ndipoleh from the activity data, we fixed its value to Ndipoleh = 0, see Tables 1–3. This value corresponds to a (hypothetical) contact ion pair without a solvation shell around the ions as a reference unit and yields xh = 1. In general, Nh will be different for different orders of the multipole expansion. Heuristically, we found that we could represent the electrolyte activities well when allowing NDipoleh ≥ 0 (this value corresponds to xDipoleh ≤ 1) while keeping Nh = 0 (i.e., xh = 1) for the quadrupole and octupole contributions. Please note that xh is a fit parameter which results from a macroscopic average over a distribution of microscopic states. We have neither information on the higher moments of this distribution nor the number of contact ion pairs in the solution. Supplementary techniques will have to provide this information.

The average apparent molar volume in the solution is given by [small phi, Greek, macron]sol = [M with combining macron]sol/ρsol where [M with combining macron]sol = xMB + (1 − x)Mw and ρsol are the average molar mass and the density of the solution, respectively.

Similarly, the radius of the “observation volume” is defined by

 
image file: d5cp01781e-t14.tif(13)
where Nw = xw/xB is the number of water molecules per electrolyte unit.

Spherical harmonics and special functions

Spherical harmonics and their application to electrolyte solutions. According to Chapter 5, Volume 2 (The classical theory of fields) in the Landau–Lifshitz series of textbooks33 the action of a field of a set of charges inside our spherical probe volume with radius Rd on a charge q at a distance Rq > Rd from the center of the probe volume can be developed as multipole expansion
 
image file: d5cp01781e-t15.tif(14)
where the lth term of the potential is given by
 
image file: d5cp01781e-t16.tif(15)

Here, l ≥ 0 and m are integer numbers, and Θ and Φ are the angles that describe the direction of Rq with respect to the coordinate system. The z-axis of the coordinate system is typically defined by the dipole moment of the distribution. The multipole moments

 
image file: d5cp01781e-t17.tif(16)
originate from the distribution of the set of charges, ea at the distances raRd and the angles θa and ϕa inside our probe volume. For simplicity, we have used point charges. However, this approach is easily transferable to systems described by charge densities, such as liquids. The moment l = 0 corresponds to the net charge in the probe volume and, therefore, vanishes. l = 1, 2, 3 correspond to the dipole, quadrupole, and octupole moments, respectively.

The spherical harmonics Ylm(Θ,Φ) are defined as

 
image file: d5cp01781e-t18.tif(17)
where image file: d5cp01781e-t19.tif and Pml is the associated Legendre polynomial. The spherical harmonics are orthonormal with respect to integration over the unit sphere:
 
image file: d5cp01781e-t20.tif(18)

We take advantage of this property by expressing the angular dependency of the net charge distribution, Δρ(R0,Θ,Φ)dR0 on a sphere with radius R0 = nRRh > Rd and infinitesimal thickness dnRRh as

 
image file: d5cp01781e-t21.tif(19)

The total interaction energy, Ude, describing the interaction of the observation volume with its environment, is given by integrating the product

 
dUde = ϕ(R0,Θ,Φρ(R0,Θ,Φ)(20)
over the volume outside the probe volume. An integration yields
 
image file: d5cp01781e-t22.tif(21)
which is an equivalent form of eqn (2).

The incomplete Euler Beta function. The incomplete Euler Beta function, Bz(a,b), belongs to the family of generalized hypergeometric functions. The particular case described in this manuscript (a = 1 + λ > 0) yields the following integral equation:28
 
image file: d5cp01781e-t23.tif(22)

Using the parameters from eqn (10), a = 1 + λl and b = 0, we obtain

 
image file: d5cp01781e-t24.tif(23)

The Lerch transcendent. The Lerch transcendent, Φ(z,s,a)29 is another function out of the family of generalized hypergeometric functions. For the case described here (s = 2, a = 1 + λl), it is characterized by the infinite sum
 
image file: d5cp01781e-t25.tif(24)

Using the parameters from eqn (10), s = 2 and a = 1 + λl, we obtain

 
image file: d5cp01781e-t26.tif(25)

Step by step integration of eqn (8). To integrate eqn (8) we describe all parameters in the mol fraction frame. Molality is in mol fraction units, described as
 
image file: d5cp01781e-t27.tif(26)

In mol fraction units, ln[thin space (1/6-em)]γ± is given by eqn (4). An evaluation leads to

 
image file: d5cp01781e-t28.tif(27)

Combining eqn (8), (26) and (27) yields

 
image file: d5cp01781e-t29.tif(28)
which can be integrated yielding eqn (10)

The Debye–Hückel term: dipole–dipole interaction. In the following we investigate in detail the long-range dipole–dipole interaction leading to the equivalent of the Debye–Hückel term in the classical theory of electrolyte solutions. The minimum hydration shell radius Rh as defined in eqn (12) defines a minimum dipole moment μ0 = qeffRh inside the observation volume. The composition of the electrolyte determines qeff which is expected to be unity for 1[thin space (1/6-em)]:[thin space (1/6-em)]1 electrolytes.

In the dilute case, we expect that the average distance between anions and cations will be much larger than Rh. Therefore, we estimate the average distance between anion and cation along the z-direction for a given dipole orientation leading to a positive dipole moment along the z-direction of μz = q(z+z) = qΔz with z+z > 0. To simplify this discussion, we assume a 1[thin space (1/6-em)]:[thin space (1/6-em)]1 electrolyte. Anion and cation can take any position within the observation volume so that Δz ranges from 0 to 2Rd. However, the number of ways to realize a dipole with a given strength depends strongly on Δz. For example, there is only a single configuration with Δz = 2Rd while there are many ways to realize configurations with ΔzRd. A detailed analysis of the resulting probability distribution shows that the effective distance Δzeff = 0.5Rd yielding μeff = 0.5qRd = 0.5μ0Rd/Rh. E.g. we expect this equation to hold for other electrolytes (e.g. 1[thin space (1/6-em)]:[thin space (1/6-em)]2, 2[thin space (1/6-em)]:[thin space (1/6-em)]1, 2[thin space (1/6-em)]:[thin space (1/6-em)]2) when replacing q by an effective charge qeff. For symmetry reasons, this dipole must be located at the center of the observation volume. When an external field Ez along the z-axis is applied, the probabilities Pp and Pa for an orientation parallel and antiparallel to the electric field, respectively, are different and can be described by Boltzmann's equation. When we assume that image file: d5cp01781e-t30.tif it follows from a linear Taylor expansion that

 
image file: d5cp01781e-t31.tif(29)

We explore next the dipole induced effect on an ion with charge q along the z-axis outside the observation volume. The interaction energy U of the ion at a position |z| > Rd outside the observation volume with μeff at the center of the observation volume and oriented along the z-axis is given by

 
image file: d5cp01781e-t32.tif(30)

For a positive charge U(z) is positive above the plane z = 0 and negative below. According to Boltzmann statistics, this leads to a depletion of positive charges above the plane and an enrichment below the plane. The expected charge difference between the upper and lower half-plane at the same z is given by Δq = q(EU(z)/kTEU(−z)/kT). Integration along the positive z-direction assuming |U(z)| ≪ kT yields Δqq2μeff/Rd. Since μeffRd and the electric field at the center the observation volume Ez ∝ Δq/Rd, it follows that the interaction energy Uμ,Ez between the dipole inside the observation volume and the electric field generated by the displacement from a single ion in the environment is inversely proportional to the radius of the observation volume. Negative charges will be enriched above the plane and depleted below the plane. The resulting electric fields are additive. The total effect of ν+ cations and ν anions will therefore be proportional to

 
image file: d5cp01781e-t33.tif(31)

To relate our description to standard Debye–Hückel theory, we introduce the molecular level ionic strength where image file: d5cp01781e-t34.tif and assume that image file: d5cp01781e-t35.tif and

 
image file: d5cp01781e-t36.tif(32)

This finding is not restricted to 1[thin space (1/6-em)]:[thin space (1/6-em)]1 electrolytes anymore.

Up to now, we have investigated the interaction of our central observation volume with a single dipole along the z-axis. To come to a more realistic scenario, we use a second coarse-graining step: we assume that the environment outside the central observation volume consists of dipoles with average dipole moment μ0 as defined in eqn (32). The central dipole μc interacts with the surrounding dipoles μs,ivia dipole–dipole interaction. While perfectly aligned dipoles at a distance R close to each other show an interaction energy that scales like

 
image file: d5cp01781e-t37.tif(33)
the thermally averaged dipole–dipole interaction energy scales like 1/R6. In general, the number of interaction partners will grow as 4πR2dR with increasing distance R from the central dipole.

If we express the distance R = nRRh between the dipoles in units of Rh (see eqn (12)) with proportionality constant nR, and normalize the interaction energy by kT, we obtain

 
image file: d5cp01781e-t38.tif(34)
where we have introduced the characteristic length
 
image file: d5cp01781e-t39.tif(35)
with the reference and where Rh is defined by eqn (12). The characteristic length and the Debye length λD are related to each other by image file: d5cp01781e-t40.tif. m0 = 1 mol kg−1 is introduced formally to yield the proper dimension and corresponds to a hypothetical 1-molal electrolyte solution with no interaction between ions.

Concentration dependency of the pre-factor. Both λc and Rh depend on a parameter combination that is concentration dependent. In the following, we will show that for most typical electrolytes, the concentration dependency is small. We will first focus on the factor image file: d5cp01781e-t41.tif. We approximate the dielectric constant as ideal combination of the real parts of the molar susceptibilities, χs and χw of the solute at infinite dilution and water, respectively:
 
image file: d5cp01781e-t42.tif(36)
where cs and cw are the concentrations of solute and water in the solution, respectively, and V0 = 1 L is the reference volume for concentration measurements. If we further use the average molar volume in the solution
 
[small phi, Greek, macron]sol = ϕw + x(ϕVϕw)(37)
where ϕw and ϕV are the apparent molal volumes of water and electrolyte, respectively, we obtain
 
image file: d5cp01781e-t43.tif(38)
where Δχ = χsχw is the difference in the real part molar susceptibilities of the electrolyte and water, and ΔϕV = ϕVϕw the difference in their apparent molar volumes. At room temperature, the dielectric constant of water εw and its density ρ are approximately 80 and 1000 g L−1. This yields χw ≈ 1.4 so that image file: d5cp01781e-t44.tif. In cases where the term linear in x in the denominator becomes noticeable, we can use a Taylor expansion of eqn (37). If the linear term becomes important before the higher order interactions prevail, we would expect an additional contribution in ln[thin space (1/6-em)]γ± which scales like x × xδ[thin space (1/6-em)]ln(x) = x1+δ[thin space (1/6-em)]ln(x) which is a member of this family of functions with a different exponent. A similar argument holds for the concentration dependency of Rh2. In conclusion, in our electrolyte picture, the vivid discussion on how much the change in dielectric constant determines the change in ln[thin space (1/6-em)]γ± is meaningless, since a change in dielectric constant and a higher order expansion are equivalent in our description.

Acknowledgements

This work was supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) under Germany's Excellence Strategy EXC-2033 390677874 RESOLV. We thank Christoph Held for providing activity coefficients in an electronically readable form and Martina Havenith for her longtime support of the project. We appreciate helpful discussions with Mohammadhasan Dinpajooh (Hadi) concerning the heterogeneity of complex electrolyte solutions. SP acknowledges funding by the European Research Council (ERC, ELECTROPHOBIC, Grant Agreement No. 101077129).

Notes and references

  1. L. Suo, O. Borodin, T. Gao, M. Olguin, J. Ho, X. Fan, C. Luo, C. Wang and K. Xu, Science, 2015, 350, 938–943 CrossRef CAS PubMed.
  2. J. Lim, K. Park, H. Lee, J. Kim, K. Kwak and M. Cho, J. Am. Chem. Soc., 2018, 140, 15661–15667 CrossRef CAS PubMed.
  3. J. Brown and A. Grimaud, Nat. Energy, 2022, 7, 126–127 CrossRef.
  4. M. Zhou, Z. Bo and K. K. Ostrikov, Phys. Chem. Chem. Phys., 2022, 24, 20674–20688 RSC.
  5. O. Borodin, J. Self, K. A. Persson, C. Wang and K. Xu, Joule, 2020, 4, 69–100 CrossRef CAS.
  6. L. Chen, J. Zhang, Q. Li, J. Vatamanu, X. Ji, T. P. Pollard, C. Cui, S. Hou, J. Chen, C. Yang, L. Ma, M. S. Ding, M. Garaga, S. Greenbaum, H.-S. Lee, O. Borodin, K. Xu and C. Wang, ACS Energy Lett., 2020, 5, 968–974 CrossRef CAS.
  7. M. Li, C. Wang, Z. Chen, K. Xu and J. Lu, Chem. Rev., 2020, 120, 6783–6819 CrossRef CAS PubMed.
  8. Y. S. Meng, V. Srinivasan and K. Xu, Science, 2022, 378, eabq3750 CrossRef CAS PubMed.
  9. R. Tiwari, D. Kumar, D. K. Verma, K. Parwati, P. Ranjan, R. Rai, S. Krishnamoorthi and R. Khan, J. Energy Storage, 2024, 81, 110361 CrossRef.
  10. C. Zhang, Y. Shi, L. Shi, H. Li, R. Li, S. Hong, S. Zhuo, T. Zhang and P. Wang, Nat. Commun., 2021, 12, 998 CrossRef CAS PubMed.
  11. F. Keshavarzi, A. Rasoolzadeh and K. Nasrifar, Ind. Eng. Chem. Res., 2024, 63, 3780–3796 CrossRef CAS.
  12. C. Held, J. Chem. Eng. Data, 2020, 65, 5073–5082 CrossRef CAS.
  13. W. J. Hamer and Y. Wu, J. Phys. Chem. Ref. Data, 1972, 1, 1047–1100 CrossRef CAS.
  14. K. S. Pitzer, Activity Coefficients in Electrolyte Solutions, CRC Press, 2018 Search PubMed.
  15. P. Debye and E. Hückel, Phys. Z., 1923, 24, 305 CAS.
  16. F. Vaslow, Thermodynamics of solutions of electrolytes, 1972 Search PubMed.
  17. K. S. Pitzer and J. M. Simonson, J. Phys. Chem., 1986, 90, 3005–3009 CrossRef CAS.
  18. C. Held and X. Liang, Fluid Phase Equilib., 2023, 575, 113931 CrossRef.
  19. J.-P. Simonin and O. Bernard, Fluid Phase Equilib., 2023, 571, 113805 CrossRef CAS.
  20. M. N. Khan, P. Warrier, C. J. Peters and C. A. Koh, J. Nat. Gas Sci. Eng., 2016, 35, 1355–1361 CrossRef CAS.
  21. D. Dhakal, D. M. Driscoll, N. Govind, A. G. Stack, N. Rampal, G. Schenter, C. J. Mundy, T. T. Fister, J. L. Fulton, M. Balasubramanian and G. T. Seidler, Phys. Chem. Chem. Phys., 2023, 25, 22650–22661 RSC.
  22. M. Dinpajooh, E. Biasin, E. T. Nienhuis, S. T. Mergelsberg, C. J. Benmore, G. K. Schenter, J. L. Fulton, S. M. Kathmann and C. J. Mundy, J. Chem. Phys., 2024, 161, 151102 CrossRef CAS PubMed.
  23. M. Dinpajooh, N. N. Intan, T. T. Duignan, E. Biasin, J. L. Fulton, S. M. Kathmann, G. K. Schenter and C. J. Mundy, J. Chem. Phys., 2024, 161, 230901 CrossRef CAS PubMed.
  24. G. Schwaab, F. Sebastiani and M. Havenith, Angew. Chem., Int. Ed., 2019, 58, 3000–3013 CrossRef CAS PubMed.
  25. F. J. Millero, Chem. Rev., 1971, 71, 147–176 CrossRef CAS.
  26. V. Lobo and J. Quaresma, Handbook of electrolyte solutions (Physical sciences data), Elsevier, Amsterdam, 1990, 1989 (ISBN 0-444-988847-5). xii+ 1168 pp. (Part A), xii+ 1186 pp. (Part B). Price Dfl. 1400.00 Search PubMed.
  27. R. Stokes, Activity coefficients in electrolyte solutions, CRC Press, 2018, pp. 1–28 Search PubMed.
  28. W. Research, Incomplete Beta Function, https://functions.wolfram.com/GammaBetaErf/Beta3/, 2023, https://reference.wolfram.com/language/ref/LerchPhi.html.
  29. W. Research, LerchPhi, https://reference.wolfram.com/language/ref/LerchPhi.html, 2023, https://reference.wolfram.com/language/ref/LerchPhi.html.
  30. R. N. Goldberg, J. Phys. Chem. Ref. Data, 1981, 10, 1–56 CrossRef CAS.
  31. C. Held and G. Schwaab, Experimental data on mean ionic activity coefficients and fit model representation for: A New Perspective on Aqueous Electrolyte Solutions, TUDOdata, 2025 DOI:10.17877/RESOLV-2024-M4WGMZTG.
  32. Y. Zhigalenok, S. Abdimomyn, M. Levi, N. Shpigel, M. Ryabicheva, M. Lepikhin, A. Galeyeva and F. Malchik, J. Mater. Chem. A, 2024, 12, 33855–33869 RSC.
  33. L. D. Landau, The classical theory of fields, Elsevier, 2013, vol. 2 Search PubMed.

Footnotes

In memoriam Herrmann Weingärtner.
Present address: Laboratoire CPCV, Département de Chimie, Ecole Normale Supérieure, PSL University, Sorbonne University, CNRS, 75005 Paris, France.

This journal is © the Owner Societies 2025
Click here to see how this site uses Cookies. View our privacy policy here.