1. Introduction
This paper offers an overview of the Landau and Lifshitz approach for the analysis of hydrodynamic fluctuations [
1], and its potential to be used as a tool for obtaining meaningful and important physical results. The starting point for the analysis is the fundamental law of entropy production, and therefore, its foundations are quite general. The Landau and Lifshitz theory of hydrodynamic fluctuations is very helpful in studying stochastic properties in viscous liquids. An intriguing example is the derivation of the Stokes–Einstein diffusion coefficient by integrating the fluctuating viscous stresses acting on the surface of a spherical particle, as demonstrated by Zwanzig [
2] and later by Fox and Uhlenbeck [
3]. The general approach can be extended beyond fluid momentum and heat transfer. It was used to offer a possible explanation for the attractive interactions between like-charged particles in colloid crystals [
4], the surface forces in thin liquid films due to surface capillary waves [
5], or ionic concentration-fluctuation correlations in electric double layers [
6].
The Landau and Lifshitz method can be adapted to examine the properties of solutions, including electrolytes. This was demonstrated by Donev et al. [
7] and Peraud et al. [
8], who studied the fluctuations in such systems in detail. Their main focus was on the non-equilibrium, gradient-driven transport in fluid mixtures and hence requires a numerical approach. Still, they offered results for the equilibrium properties of solutions such as structure factors and the corresponding radial distribution functions in the Debye and Huckel limit. The focus of the present review is on the general analytical methodology that allows for the derivation of thermodynamic equilibrium properties of electrolyte solutions such as osmotic pressure, radial distribution functions, and interaction energies between ions (potentials of mean force). The effect of the electrostatics enters the analysis via the Nernst–Planck equations [
9] for the fluxes in combination with the Poisson equation that relates the electrostatic potential to the charge in the fluid [
10]. We provide a detailed pathway for deriving the Debye and Huckel mean field theory of strong electrolytes [
11,
12]), starting from the general relationship for the entropy production. All assumptions are clearly outlined, which helps to better understand the limitations of the Debye and Huckel model. This analysis is less general than the one offered by Peraud et al. [
8], as it does not include non-equilibrium effects. Abandoning some of the simplifications (dilute solutions, small fluctuation amplitudes, solvent structural molecular contributions, etc.) may in principle lead to a more general theory. The complexity of the calculations, however, would increase and obtaining analytical results may again become impossible.
3. Fluctuation Correlations and Thermodynamics of Binary Symmetric Electrolyte Solutions
A binary electrolyte solution is actually a ternary system that consists of positive ions, negative ions, and solvent. However, if the concentrations of both positive and negative ions (
and
) is low in comparison to that of the solvent
c, then each ionic species presents an ideal solution, and all results from the previous section are applicable to each of them. Our analysis focuses on symmetric electrolytes where the ions carry the same charge numbers,
z, but with opposite signs. The presence of the charges in the solution, however, creates an additional difficulty associated with the electric fields and electrostatic interactions between the charged components. The celebrated theory of Debye and Huckel [
11] was the first successful attempt to overcome these difficulties and is still included in most texts that are dedicated to the physics and chemistry of electrolyte solutions. The fluctuation correlation analysis, described here, offers a different starting point to approach the same problem. It reveals some important physical insights and provides strategies for solving similar problems in a relatively straightforward way.
3.1. Balance Equations
We start with balancing the mass of dissolved electrolytes as well as the potential in the solution. The mass balance Nernst–Planck equations read
The coefficients
and
are the hydrodynamic mobilities of the positive and negative ions, while
e is the elementary charge. The electrostatic potential
is related to the charge density
by means of the Poisson equation [
10]
where
is the dielectric constant in vacuo, and
is the relative dielectric permittivity.
In order to account for the fluctuations in the solution, we define the fluxes, the concentrations, and the local potential as
and
where
and
are the diffusion coefficients for the positive and negative ions,
and
are the concentration fluctuations for the positive and negative ions, and
c is the uniform average concentration in units of number per volume. The resultant local electrostatic potential fluctuation is
.
The balance equations for all fluctuating quantities become
The charge fluctuation is
where
Rearranging the above equations yields
The parameter
is the Debye screening parameter (inverse length) [
11], which is a measure of potential magnitude reduction with distance around an ion in the solution. A significant simplification of Equation (
33) can be achieved by setting
, and noting that for symmetric electrolytes, the overall concentrations are
. This seems like a very strong limitation, but it does not affect the majority of the final results. Combining the first two equations in (
33) leads to an equation for the charge density fluctuation
3.2. Charge Fluctuation Correlations and Correlation Energy
We will use the Fourier transform technique [
16] to express the charge and flux fluctuations as
and
The inverse transforms are
and
where
is the wavevector and
is the frequency.
The ionic fluxes and their correlations are of special interest. For low ionic concentrations, the diffusion flux of each component does not depend on any variable pertinent to the other solutes (see Equations (
20)). Hence, the correlations between the fluxes for different species is zero, or
Equations (
38) and (
39) allow for the derivation of the important relationship (see
Appendix A)
Note that the right-hand sides for the first two equations in (
40) are the same irrespective of whether the positive or negative ionic fluxes are correlated.
Equations (
35)–(
38) allow Fourier transforming the whole Equation (
34)
Solving for the charge density
yields
The charge correlation fluctuation correlation is
Using Equations (
40) and (
43), the identity
leads to
The electrostatic energy per unit volume of the solution is (see
Figure 1)
Using the Fourier transform of Equation (
30)
Then, the correlation of the potential and charge fluctuations in Equation (
45) becomes (see also Equation (
43)).
Assuming
, and inverting (
47) with respect to
and
(see
Appendix B) leads to the following result for the electrostatic energy (
45)
Expanding the exponential leads to
The first term on the right-hand side is the self-energy of the ions, while the second term accounts for the electrostatic correlation energy [
14]
3.3. Thermodynamic Relationships
3.3.1. Free Energy and Chemical Potential
The focus of this section is on the effect of electrolyte solution non-ideality, which is represented by the correlation energy (
50), on its thermodynamic properties. We start with the general thermodynamic relationship between the internal energy
E and the Helmholtz free energy
A [
12,
14]
Substituting the correlation internal energy (
50) into the above equation and integrating over the temperature yields
Let
be the number of pairs of positive and negative ions, and
is the total number of water molecules. Since we are considering a dilute electrolyte solution, it is reasonable to assume that the total volume is
, where
is the molecular volume of water. Then, Equation (
52) becomes
Differentiating Equation (
53) with respect to
allows obtaining the chemical potential of the solvent (i.e., the water)
After some rearrangements, we finally obtain
where
is the total chemical potential and
is the ideal part defined by [
12,
14]
is the standard chemical potential.
3.3.2. Osmotic Pressure
The osmotic pressure difference between the electrolyte solution and the pure water solvent can be derived from the chemical equilibrium relationship
Rearranging (
57) yields
where
is the osmotic pressure of the electrolyte solution. For dilute electrolytes,
and (
58) simplifies to
The first term on the right-hand side of Equation (
59) corresponds to the ideal part of the solution osmotic pressure, while the second term accounts for the charge correlation effects. This result for the osmotic pressure also follows from the traditional Debye–Huckel approach [
12,
14].
3.3.3. Radial Distribution Functions and Pair Interaction Energies
The starting point for determining the radial distribution functions are the concentration-fluctuation correlation functions
,
, and
, where the indices + and − correspond to the two ionic species (positive and negative) in the solution. Combining Equations (
29) and (
30) leads to
Transforming Equation (
60) and solving for the concentration fluctuation yields
Equation (
61) can be used to find the Fourier transform if the concentration-fluctuations correlations for positive–positive negative–negative and positive–negative ionic combinations. These are then used to obtain the radial distribution functions and the potentials of mean force for all ionic interactions in the mean field (i.e., Debye–Huckel) limit (see also Ref. [
8]).
3.4. Positive–Positive Ionic Correlations
The Fourier transform of the positive–positive ions concentrations fluctuation correlation is derived from the first Equation (
61). After averaging, the result is
Inverting Equation (
62) (see Appendices B and C) and setting
leads to the following result for the concentration-fluctuation correlation in real space
where
is the radial distribution function for the positive ions. Comparing (
64) to the formal definition
leads to the Debye and Huckel [
11] result for the electrostatic energy of interaction between the ions in the solution
The energy is positive, which indicates a repulsive interaction.
3.5. Negative–Negative Ionic Correlations
Using the second Equation (
61), we obtain the concentration negative–negative ion concentration-fluctuation correlation in the form
Note that the sign in front of the last two terms on the right is positive, which is in contrast to Equation (
62). Following the same recipe as for the positive–positive ion interactions above (see
Appendix C), we derive
The radial distribution function is
and the interaction energy reads [
11]
which is also positive (i.e., repulsive). Hence, the radial distribution functions and interaction energies for ions that are both positive or both negative are identical.
3.6. Positive–Negative Ionic Correlations
The positive–negative ion concentration-fluctuation correlation is also derived from Equation (
61). The result becomes
The correlation between the positive and negative ionic fluxes is zero because of Equation (
39). In addition,
. The inversion of Equation (
71) at
yields
The radial distribution function and pair interaction energy between oppositely charged ions are
and [
11]
The sign of the interaction energy is negative, which is in agreement with the fact that positive and negative ions attract each other.
4. Conclusions
The Landau–Lifshitz approach for the treatment of hydrodynamic fluctuations is based on the fundamental expression for entropy production (
4). Its generalization to include transport processes (such as diffusion) is straightforward. The method offers simple expressions for the correlation of the fluctuating fluxes. The flux fluctuations are related to the concentration fluctuations, which in turn account for the effects of interactions. Proper averaging allows us to derive results that are pertinent to systems in thermodynamic equilibrium. The procedure is relatively simple, and the main mathematical techniques are the forward and inverse Fourier transforms, and more specifically the transforms of various Dirac delta functions.
The focus of this paper is on electrolyte solutions. We show how the Landau and Lifshitz method can be used to derive the Debye and Huckel theory of strong, dilute electrolytes. This is accomplished without formally solving the Poisson equation of electrostatics (
25). Even the Debye and Huckel assumption of low potential and the subsequent linearization of the Boltzmann distribution of the local charge
is not explicitly applied. Instead, we use the relationship between charge and potential fluctuations (
30) to write the transport Equation (
29) in terms of concentration fluctuations. The equilibrium state, in our analysis, is established by the integration (averaging) over the entire frequency spectrum in Fourier space. The assumptions
,
,
, and
are equivalent to the Debye and Huckel low potential approximation
. However, the fluctuation correlation analysis clearly demonstrates that the low potential approximation is implied by the small charge density fluctuations, which are a direct consequence of the small concentration fluctuations. This is also true for the Debye and Huckel analysis [
11], although there, the argument starts with the low potential approximation. Another curious feature of the method is that although it is examining the fluctuations in the system, the end result is a mean field theory, which is analogous to the Debye and Huckel result.
The correlation energy (
45) and the osmotic pressure, Equations (
58) and (
59), are entirely determined by the charge density fluctuation correlations given by Equations (
43) and (
44). The pair energy of interaction positive–positive, negative–negative, and neagtive–positive ions depends on the charge density fluctuation correlations but also includes contributions from the fluctuation correlations of respective ionic fluxes with those of the charge density (see Equations (
62), (
67) and (
71)). The terms that are proportional to the positive–positive and negative–negative flux fluctuation correlations in (
62) and (
67) do not contribute at all to the pair interaction energy but are responsible for the
term in Equations (
62) and (
67). There is no such term in Equation (
71) because the diffusion fluxes for the positive and negative ions are independent and uncorrelated (see Equations (
39) and (
40)).
The assumption that all diffusion coefficients are the same () can be abandoned, which makes all derivations and calculations significantly more lengthy and tedious. The final results, however, for the osmotic pressure, interaction energies and radial distribution functions are not affected. This is not surprising, since thermodynamic properties cannot depend on dissipative coefficients such as the diffusivities.