A Generalized Debye-Hückel Theory of Electrolyte Solutions

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

A generalized Debye-Hückel theory of

electrolyte solutions
Cite as: AIP Advances 9, 015214 (2019); https://doi.org/10.1063/1.5081863
Submitted: 17 November 2018 . Accepted: 04 January 2019 . Published Online: 16 January 2019

Jinn-Liang Liu, and Chin-Lung Li

AIP Advances 9, 015214 (2019); https://doi.org/10.1063/1.5081863 9, 015214

© 2019 Author(s).
AIP Advances ARTICLE scitation.org/journal/adv

A generalized Debye-Hückel theory


of electrolyte solutions
Cite as: AIP Advances 9, 015214 (2019); doi: 10.1063/1.5081863
Submitted: 17 November 2018 • Accepted: 4 January 2019 •
Published Online: 16 January 2019

Jinn-Liang Liua) and Chin-Lung Li

AFFILIATIONS
Institute of Computational and Modeling Science, National Tsing Hua University, Hsinchu 300, Taiwan

a)
Electronic mail: jlliu@mx.nthu.edu.tw

ABSTRACT
We propose a generalized Debye-Hückel (DH) theory by using a recent Poisson-Fermi model that accounts for the steric, cor-
relation, and polarization effects of ions and water treated as nonuniform spheres with interstitial voids. The generalized DH
theory reduces to the classical one when these effects are not considered. The Debye length is also generalized to include the
steric effect. The new theory yields an electrolyte (analytical) equation of state for calculating the activity coefficient of aqueous
electrolyte solutions, which is of fundamental importance in thermodynamic modeling for a variety of chemical and biological
systems. Results obtained by the analytical equation well fit experimental data for eight 1:1 and six 2:1 electrolytes with only three
adjustable parameters. By contrast, the DH models extended from the classical DH theory such as the Pitzer model can have
several to great many parameters due to explosive combinations of wide ranges of composition, temperature, and pressure.
© 2019 Author(s). All article content, except where otherwise noted, is licensed under a Creative Commons Attribution (CC BY) license
(http://creativecommons.org/licenses/by/4.0/). https://doi.org/10.1063/1.5081863

I. INTRODUCTION coefficients in all types of binary and multi-component solu-


tions at variable concentration, temperature, and pressure
Thermodynamic modeling is of fundamental importance with only 3 adjustable parameters. The equation generalizes
in the study of chemical and biological systems.1–8 Since the classical DH equation from the Poisson-Fermi (PF) theory
Debye and Hückel (DH) proposed their theory in 19239 and developed in recent years.16–28
Hückel extended it to include Born energy effects in 1925,10 The Fermi-like distribution of ions of equal size with voids
a great variety of extended DH models have been devel- in ionic liquids was recently derived by Kornyshev18 using
oped for modeling aqueous or mixed-solvent solutions over Bikerman’s lattice model.29 The corresponding entropy func-
wide ranges of composition, temperature, and pressure.2,11–14 tional20 involves a volume fraction, which is empirical and
Despite these intense efforts, robust thermodynamic mod- unrealistically large to fit experimental data in some appli-
eling of electrolyte solutions still presents a difficult chal- cations.19 This entropy functional does not directly yield
lenge for extended DH models due to enormous amount of classical Boltzmann distributions as the volume of ions van-
parameters that need be adjusted carefully and often sub- ishes.24 On the other hand, the entropy functional proposed in
jectively.11–15 It is a frustrating despair (see, e.g., p. 301 in Ref. 27 treats all ions and water molecules as nonuniform
Ref. 1 and p. 11 in Ref. 7) that about 22,000 parameters12 need spheres with interstitial voids. The corresponding Fermi dis-
be extracted from the available experimental data for one tem- tribution reduces to Boltzmann distribution when the volume
perature for combinatorial solutions of the most important 28 of all different spheres vanishes.27 The PF functional does not
cations and 16 anions in salt chemistry by the Pitzer model,2 use any lattice models but simply uses the volume equation
which is the most widely used DH model with unmatched of all particles with voids and a steric potential22 defined by
precision for modeling electrolyte solutions.13 the voids. The steric potential can describe different steric
By contrast, we propose here an electrolyte equation of energies for different ions and water in size similar to differ-
state,14 which is for calculating single-ion and mean activity ent electric energies for different ions in charge, which are

AIP Advances 9, 015214 (2019); doi: 10.1063/1.5081863 9, 015214-1


© Author(s) 2019
AIP Advances ARTICLE scitation.org/journal/adv

especially relevant to determining selectivity of specific ions The radii of Ωi and the outer boundary of Ωsh are denoted by
by certain biological ion channels.23,24,26,27 RBorn
i
(ionic cavity radius32 ) and Rsh
i
, respectively.
The correlation length — a measure of how strongly The potential function φ(r) can be found by solving the PF
correlations between ions develop17,20,21 and how water equation25,28
molecules polarize in electrolyte solutions in response to an K+1
electric field22–28 — generalized from that in Ref. 30 is for   X
s l2c ∇2 − 1 ∇2 φ(r) = ρs (r) B qk Ck (r) in Ωs (2)
describing the correlation effect and is not empirical in con- k=1
trast to empirical ones in earlier works.17,20–28 It depends not
only on salt concentration but also on all sizes of ions and and the Laplace equation
water (in contrast to that of size independence in Ref. 30). − ∇2 φ(r) = 0 in Ωi ∪ Ωsh , (3)
The generalized DH equation — modeling how electrolyte
solutions behave differently from ideality — is an algebraic where  s =  w  0 ,  w is the dielectric constant of p bulk water
equation that is easily and efficiently evaluated in contrast to in Ωsh ∪ Ωs ,  0 is the vacuum permittivity, lc = lB lD /48 is
the model system of partial differential equations (PDEs) used a density-density correlation length independent of specific
in our prior works,25,28 which are computationally much more ionic radius,30 lB and lD are the Bjerrum and Debye lengths,
expensive. Nevertheless, both models have the same three respectively, K + 1 denotes water, qK+1 = 0, the concentra-
v

parameters for calculating not only single-ion but also mean tions Ck (r) = CBk exp −βk φ(r) + v0k Strc (r) in Ω are Fermi-like
activity coefficients of electrolyte solutions at variable tem- distributions for all k = 1, . . ., K + 1,22,27 CBk are constant bulk
perature (ranging from 298.15 to 573.15 K shown in Ref. 28) and concentrations in molarity (M), β k = qk /kB T, vk = 4πa3k /3 is the
pressure (from 1.01 to 80.5 bars28 ) or in higher concentration P
K+1

volume of a type k particle with radius ak , v0 = v /(K + 1)
k=1 k
range (up to 6 mol/kg28 ).
Γ(r)
is an average volume of all types of particles, Strc (r) = ln GammaB
II. THEORY in Ω is a (dimensionless) steric potential that defines a steric
−v
The activity coefficient γ i of an ion of species i in an energy v0k Strc (r)kB T of type k ion or water at r,22,27 Γ(r)
PK+1
aqueous electrolyte solution with a total of K species of ions = 1 − k=1 vk Ck (r) is a void fraction function, and ΓB = 1
describes deviation of the chemical potential of the ion from − K+1
P
v CB is a constant void fraction. Since the steric poten-
k=1 k k
ideality (γ i = 1).4 The excess chemical potential µiex = kB T ln γi tial takes particle volumes and voids into account, the shell
can be calculated by25,31 volume V sh of the shell domain Ωsh can be determined by the
v0 Ow Vsh −vw Ow
1 1 steric potential Strc = ln i
= ln i 25,28
, where
µiex = qi φ(0) − qi φ 0 (0), (1) sh vw Vsh CBK+1 Vsh ΓB
2 2 the occupant (coordination) number Ow
of water molecules
i
where kB is the Boltzmann constant, T is an absolute tempera- is given by experimental data.33 The shell radius Rsh is thus
i
ture, qi is the ionic charge of the hydrated ion (also denoted by determined and depends not only on Ow but also on the bulk
i
i), φ(r) is a reaction potential31 function of spatial variable r in void fraction ΓB , namely, on all salt and water bulk concentra-
the domain Ω = Ωi ∪ Ωsh ∪ Ωs shown in Fig. 1, Ωi is the spher- tions (CBk ).25,28
ical domain occupied by the ion i, Ωsh is the hydration shell The fourth-order PF equation (2) reduces to the clas-
domain of the ion, Ωs is the rest of solvent domain, 0 denotes sical second-order Poisson-Boltzmann (PB) equation when
the center (set to the origin) of the ion, and φ 0 (r) is a poten- lc = vk = 0 for all k,27 i.e., the correlation and steric effects
tial function when the solvent domain Ωs does not contain any are not considered. Eq. (2) was first proposed in Ref. 17 (with
ions at all with pure water only, i.e., when the solution is ideal. vk = 0 for all k) and subsequently derived in Refs. 20 and 27
(with vk ,0) from different perspectives of electrostatics. For
reducing complexity of using higher-order approximations in
implementation, we transform the fourth-order PDE (2) to the
following two second-order PDEs22
 
l2c ∇2 − 1 ψ(r) = ρs (r) in Ωs , (4)

s ∇2 φ(r) = ψ(r) in Ωs , (5)


where the extra unknown function ψ(r) is a density-like func-
tion as seen from (4) by setting lc = 0. The boundary and
interface conditions for φ(r) and ψ(r) in (3)–(5) are22
φ(r) = ψ(r) = 0 on ∂Ωs \∂Ωsh , (6)

FIG. 1. The model domain Ω is partitioned into the ion domain Ωi (with radius ψ(r) = −ρs (r) on ∂Ωsh ∩ ∂Ωs , (7)
RBorn
i
), the hydration shell domain Ωsh (with radius Rsh
i
), and the remaining solvent
domain Ωs . [φ(r)] = 0 on ∂Ωi ∪ (∂Ωsh ∩ ∂Ωs ), (8)

AIP Advances 9, 015214 (2019); doi: 10.1063/1.5081863 9, 015214-2


© Author(s) 2019
AIP Advances ARTICLE scitation.org/journal/adv

[∇φ(r) · n] = 0 on ∂Ωsh ∩ ∂Ωs , (9) where


λ21 − λ22
Θ= √ √ , (14)
[ (r)∇φ(r) · n] = i ∇φ ∗ (r) · n on ∂Ωi ,

(10) λ21 λ2 Rsh + 1 − λ22 λ1 Rsh +1
i i
 q     q 
where ∂ denotes the boundary of a domain, the jump function r = |r |, λ1 = 1 − 1 − 4l2c /l2DPF / 2l2c , and λ2 = 1 + 1 − 4l2c /l2DPF /
[φ(r)] = limrsh →r φ(rsh ) − limri →r φ(ri ) at r ∈ ∂Ωi with rsh ∈ Ωsh  
and ri ∈ Ωi ,  (r) =  s in Ωsh and  (r) =  i in Ωi ,  i =  ion  0 ,  ion is 2l2c . Note that limlc →0 λ1 = 1/l2DPF , limlc →0 λ2 = ∞, and
a dielectric constant in Ωi , n is an outward normal unit vector limlc →0 Θ = limCB →0 Θ = limlDPF →∞ Θ = 1.36 The linearized PF
1
at r ∈ ∂Ωi , and φ ∗ (r) = qi /(4πi |r-0|). Eqs. (2) and (3) avoid large potential φ PF (r) reduces to the linearized PB potential φ PB (r)
errors in a direct approximation of the delta function δ(r − 0) = qi e−r/lD /(4πs r) as in standard texts (e.g. Eq. (7.46) in Ref. 4)
in the singular charge qi δ(r − 0) of the solvated ion at the origin by taking limlc →0 φ PF (r) with vk = 0 for all k, Rsh i
= 0, and
0 by transforming the singular charge to the Green’s function r > 0.36
φ ∗ (r) on ∂Ωi in (10) as an approximation source of the electric As discussed in Ref. 37, since the solvation free energy
field produced by the solvated ion.34,35 of an ion i varies with salt concentrations, the Born energy
For simplicity, we consider a general binary (K = 2) q2i 1w − 1 /8π0 R0i in pure water (i.e. CBi = 0) with a constant
electrolyte Cz2 Az1 with the valences of the cation Cz1 + and Born radius R0i should change to depend on CBi ≥ 0. Equiva-
anion Az2 − being z1 and z2 , respectively. The first-order Tay-
lently, the Born radius RBorn in (13) is variable and we can model
lor approximation of the charge density functional ρs (φ(r)) in i

(2) with respect to the electric potential φ(r) yields36 it from R0i by a simple formula25,28
 B  1/2  B  3/2
−CB1 q1 RBorn = θR0i , θ = 1 + α1i Ci
B
+ α2i Ci + α3i Ci , (15)
ρs (φ(r)) ≈ [(q1 − q2 ) − Λq1 ]φ(r), (11) i
kB T
B
2 where Ci = CBi /M is a dimensionless bulk concentration and
f  g
where Λ = CB1 (v1 − v2 ) / ΓB v0 + v21 CB1 + v22 CB2 + v23 CB3 , which is
a dimensionless quantity corresponding to the steric potential α1i , α2i , and α3i are parameters for modifying the experimen-
Strc (r). Consequently, we obtain a generalized Debye length tal Born radius R0i to fit experimental activity coefficient γ i
1/2 that changes with the bulk concentration CBi of the ion. The
s kB T
lDPF = * B + (12) Born radii R0i given below are from Ref. 37 obtained from the
C ((1 − Λ)q 2−q q )
, 1 1 1 2 - experimental hydration Helmholtz free energies of those ions
given in Ref. 5. The three parameters in (15) have physical or
that reduces to the original Debye length lD 4 if v1 = v2 ,0 (two
mathematical meanings unlike many parameters in the Pitzer
ionic species having equal radius and thus Λ = 0) or v1 = v2 = v3
model.12,13,15 The first parameter α1i adjusts R0i and accounts
= 0 (all particles treated as volumeless points in standard texts
for the real thickness of the ionic atmosphere (Debye length),
for PB4 ).
which is proportional to the square root of the ionic strength
In Ref. 36, we analytically solved the linear PF PDEs (3), (4),
in the DH theory.4 The second α2i and third α3i parameters are
and (5) with (11) in a similar way as Debye and Hückel solved the
adjustments in the next orders of approximation beyond the
linear PB equation for a spherically symmetric system. How-
DH treatment of ionic atmosphere.28
ever, the spherical domain shown in Fig. 1 and the boundary  
and interface conditions in (6)–(10) are different from those The potential value φ 0 (0) = limCB →0 φ PF (0) = qi / 4πs R0i
1
of the standard method for the linear PB equation in physical by limCB →0 Θ = 1 and limCB →0 RBorn
i
= R0i . From (1) and (13), we
1 1
chemistry texts.4 The analysis consists of the following steps: thus have a generalized activity coefficient γiDHPF in
(i) The nonlinear term ρs (r) in (4) is linearized to the linear term
−s φ/l2DPF in (11) as that of Debye and Hückel. (ii) The corre- q2i
ln γiDHPF = * 1 − 1 + Θ − 1 +, (16)
sponding linear PDEs to (4) and (5) are then formulated into a 8πs kB T , RBorn R0 Rsh
system of eigenvalue problems with eigenfunctions (φ(r), ψ(r)) i i i -

and eigenvalues (λ1 , λ2 ), where the general solution of φ(r) is which satisfies the DH limiting law, i.e., γiDHPF = γiDH = 1
equal to that of Debye and Hückel in the solvent domain Ωs
for infinite dilute (ideal) solutions as CBi →0. The generalized
(not the entire domain) when lc = v1 = v2 = v3 = 0. (iii) A unique
 activity coefficient γiDHPF reduces to the classical DH activity
pair of eigenfunctions φ PF (r), ψ PF (r) is found under condi-
coefficient γiDH ,9 namely,
tions (6)–(10), where φ PF (r) is equal to that of Debye and Hückel
in Ωs when lc = v1 = v2 = v3 = 0. −q2i
The analytical potential function that we found36 is ln γiDH = (17)
8πs kB T(Ri + lD )
qi qi
4π s RBorn
+ (Θ − 1) in Ωi
4π s Rsh if RBorn = R0i (without considering Born energy effects), Rsh = Ri


i i

i i


 qi qi
+ (Θ − 1) in Ωsh (an effective ionic radius9 ), lDPF = lD (no steric effect), and


 4π s r 4π s Rsh

φ PF (r) =

i (13) lc = 0 (no correlation effect). The reduction shown in Ref. 36
√ √
2 − λ 2 (r−Ri ) −λ 2 e− λ1 (r−Ri ) 
sh sh 


qi  λ1 e

is by taking the limit of the last term in (16) as lc → 0, i.e.,


2
in Ωs ,

√ √

 4π s r  λ2 ( λ2 Rsh +1)−λ2 ( λ1 Rsh +1) 
 
sh = R +l .
−1
limlc →0 Θ−1
 
1 i 2 i
   Ri i D

AIP Advances 9, 015214 (2019); doi: 10.1063/1.5081863 9, 015214-3


© Author(s) 2019
AIP Advances ARTICLE scitation.org/journal/adv

FIG. 2. Single-ion activity coefficients of 1:1 electrolytes.


Comparison of DHPF results (curves) with experimental
data (symbols)38 on i = C+ (cation) and A− (anion) activity
coefficients γ i in various [CA] from 0 to 1.6 M.

III. RESULTS the curves in these figures are given in Table I from which
Results obtained by (16) and shown in Figs. 2 and 3 agree we observe that RBorn
i
deviates from R0i slightly. For example,
well with the experimental data by Wilczek-Vera et al.38 (also RCl− /RCl− = 1.065 ∼ 0.926 (not shown) with [CaCl2 ] = 0 ∼ 1.6 M
Born 0

used in Refs. 25, 28, and 37). The values of α1i , α2i and α3i for

TABLE I. Values of α1i , α2i , α3i in Eq. (15).

Fig.# i α1i α2i α3i

2a Li+ −0.006 −0.037 0.004


2a Cl− 0.052 −0.015 0
2b Li+ −0.006 −0.011 −0.004
2b Br− 0.026 −0.057 0.010
2c Na+ 0 0 0
2c F− 0.027 0 0
2d Na+ −0.045 0.009 −0.002
2d Cl− 0.063 −0.014 −0.002
2e Na+ −0.049 0.042 −0.013
2e Br− 0.071 −0.048 0.006
2f K+ 0.005 0.051 −0.015
2f F− 0.033 −0.028 0.003
2g K+ 0.031 0.022 −0.005
2g Cl− 0.020 −0.025 0.004
2h K+ 0.025 −0.062 0.018
2h Br− 0.001 0.082 0
3a Mg2+ −0.038 −0.010 −0.003
3a Cl− 0.078 −0.025 −0.004
3b Mg2+ −0.030 −0.032 0.002
3b Br− 0.091 −0.031 −0.007
3c Ca2+ −0.025 −0.014 −0.002
3c Cl− 0.116 −0.054 0.002
3d Ca2+ −0.053 −0.030 0.002
3d Br− 0.080 0.095 −0.037
3e Ba2+ 0.090 −0.108 0.022
FIG. 3. Single-ion activity coefficients of 2:1 electrolytes. Comparison of DHPF 3e Cl− 0.056 −0.005 0
results (curves) with experimental data (symbols)38 on i = C2+ (cation) and A− 3f Ba2+ −0.003 −0.042 0.006
(anion) activity coefficients γ i in various [CA2 ] from 0 to 1.5 M. 3f Br− 0.029 0.041 −0.019

AIP Advances 9, 015214 (2019); doi: 10.1063/1.5081863 9, 015214-4


© Author(s) 2019
AIP Advances ARTICLE scitation.org/journal/adv

for Fig. 3c, since the cavity radius RBornCl−


is an atomic mea- The activity coefficient γiDHPF is temperature and pres-
sure from the infinite singularity δ(r − 0) at the origin, i.e., sure dependent since Eq. (16) includes the temperature T and
φ PF (r) and thus γiDHPF are very sensitive to RBorn . On the other the voids in the void fraction ΓB in (12). For the mean activ-
i
hand, γiDHPF is not very sensitive to Rsh (Rsh = 5.123 ∼ 5.026 Å ity coefficients of either ternary (Fig. 428 ) or binary (Figs.
i Cl− 5 and 628 ) systems, we only need to adjust three parame-
for Fig. 3c), i.e., the fixed choice of Ow = 18 (an experimental
i ters for one cation (not all ions) denoted by i, i.e., θ(CBj ) = 1
value for Ca2+ in Ref. 33) for all curves is not critical but reason-
for all other j ions in (15) because the mean activity coeffi-
able.25 The error between the estimated Ow and its unknown
i cient is very sensitive to the value of θ(CBi ) ≈ 1. For exam-
true value can always be compensated by small adjustments of
ple, the mean activity coefficient γ NaCl in the ternary solution
RBorn
i
. Table I also shows the significant order of these param-
NaCl + MgCl2 with [MgCl2 ] from 0 to 6 mol/kg (Fig. 4B28 )
eters, i.e., α1i > α2i > α3i in general cases. We postulate is calculated by γNaCl = 21 (γNa+ + γCl− ) with θ(CBNa+ ) ≈ 1 (α1Na
+

that particular cases that violate this order are due to either + +
= 1.00527, α2Na = 0.00042, α3Na = 0.00019 in the row 4(b)
our fitting method28 or our choice of experimental data that
are subject to measurement errors. We shall investigate this of Table III28 ) and θ(CBCl− ) = θ(CB 2+ ) = 1 (without adjust-
Mg
issue more closely in the future with more data sets of same j j j
ing any corresponding parameters, i.e., α1 = α2 = α3 = 0
conditions from other sources to develop predictive DHPF for j = Cl− and Mg2+ 28 ). The solvation free energy of Na+ ion
models for experimental use. The values of other symbols are (or γNa+ ) is affected by Cl− and Mg2+ via correlation (lc in
aLi+ = 0.6, aNa+ = 0.95, aK+ = 1.33, aMg2+ = 0.65, aCa2+ = 0.99, (14) and (16)) and Debye screening (lDPF ) effects. In essence,
aBa2+ = 1.35, aF− = 1.36, aCl− = 1.81, aBr− = 1.95, aH2 O = 1.4 Å, we only need three parameters to fit experimental data for
R0Li+ = 1.3, R0Na+ = 1.618, R0K+ = 1.95, R0 2+ = 1.424, R0 2+ = 1.708, each activity-concentration curve for either single-ion or
Mg Ca
mean activities in all types of binary and multi-component
R0 2+ = 2.03, R0F− = 1.6, R0Cl− = 2.266, R0Br− = 2.47 Å,  w = 78.45, solutions.
Ba
 ion = 1, T = 298.15 K. The evaluation of (16) is purely algebraic and thus much
The generalized Debye length lDPF changes with the size, more efficient than expensive computations of the PDE sys-
valence, and concentration p of all particle species so does the tem (4)–(10) as performed in our earlier work.28 For example,
correlation length lc = lB lDPF /48. Nevertheless, the largest it took 0 (rounded) and 116 seconds of total run time on a
value of the ratio lDPF /lD (not shown) for all cases in Figs. 2 laptop with Intel Core i7-6500U CPU by the algebraic and
and 3 is only 1.028 indicating that the generalized Debye PDE models, respectively, to produce 12 γMg2+ values for the
length lDPF does not deviate from the classical (size inde- corresponding curve in Fig. 3a.
pendent) lD too much in the range of concentrations con-
sidered here. The eigenvalues λ1 and λ2 (not shown) are
real for all cases implying that the correlation length should
indeed be generalized to include nonuniform densities of all IV. CONCLUSION
species as proposed in Ref. 30. However, the present corre- In summary, we use the Poisson-Fermi model to gen-
lation length is not universal30 but size dependent in con- eralize the classical Debye-Hückel theory to account for the
trast to that in Ref. 30 derived from lD . Correlation lengths steric, correlation, and polarization effects in aqueous elec-
used in earlier works17,20,22–28 are all empirical constants that trolyte solutions. The generalization results in new physical
depend specifically on the size and/or valence of a particular models of electrostatic potential, Debye length, and single-
species of interest. The correlation length used here is not a ion or mean activity coefficients over wide ranges of con-
parameter. centration (0 to 6 mol/kg), composition (binary or mixtures),
We make some remarks on future applications of the gen- temperature (298.15 to 573.15 K), and pressure (1.01 to 80.5
eralized DH equation (16) to fit or study experimental data bars).
of various ionic activities as summarized in Refs. 2, 8, and Results obtained by the analytical formula of activity coef-
12–14 to name a few. As seen from Table I, our model requires ficient have shown to well fit experimental data with only
only three parameters α1i , α2i , and α3i based solely on the cru- three parameters in contrast to many parameters in classi-
cial Born radius RBorni
, which is unknown and depends in a cal DH-based models. We only need three parameters for one
very complex and complicated way on physical interactions ion (not all ions) for each activity-concentration curve to fit
between ions and water molecules at variable concentration, data for either single-ion or mean activities in all types of
composition, temperature, and pressure etc. For single-ion binary and multi-component solutions. These three param-
activity coefficients, the fitting curves by (16) in Figs. 2 and eters have physical meanings for approximating unknown
3 are almost identical to those in Figs. 2 and 3 in Ref. 28 Born energy that depends on the composition and condi-
with different values of α1i , α2i , and α3i (cf. Table III in Ref. 28). tions of electrolyte in a very complex way. The values of
Therefore, Eq. (16) can also be used to calculate mean activity these parameters show their significant order in the approx-
coefficients of electrolytes (as shown in Fig. 4 in Ref. 28) even imation in general cases. The formula is analytical and thus
at variable temperature and pressure (ranging from 298.15 to very efficient for calculations when compared with numer-
573.15 K and 1.01 to 80.5 bars in Figs. 5 and 628 ) or in higher ical computations of differential equations in our earlier
concentration range (up to 6 mol/kg in Figs. 4-628 ). work.

AIP Advances 9, 015214 (2019); doi: 10.1063/1.5081863 9, 015214-5


© Author(s) 2019
AIP Advances ARTICLE scitation.org/journal/adv

19
ACKNOWLEDGMENTS M. Z. Bazant, M. S. Kilic, B. D. Storey, and A. Ajdari, “Towards an under-
standing of induced-charge electrokinetics at large applied voltages in
This work was supported by the Ministry of Science and concentrated solutions,” Adv. Coll. Interf. Sci. 152, 48–88 (2009).
Technology, Taiwan (No. MOST 107-2115-M-007-017-MY2 to 20
M. Z. Bazant, B. D. Storey, and A. A. Kornyshev, “Double layer in ionic
J.-L.L. and No. MOST 107-2115-M-007-003 to C.-L.L.). liquids: Overscreening versus crowding,” Phys. Rev. Lett. 106, 046102
(2011).
21
B. D. Storey and M. Z. Bazant, “Effects of electrostatic correlations on
REFERENCES
electrokinetic phenomena,” Phys. Rev. E 86, 056303 (2012).
1 22
R. Robinson and R. Stokes, Electrolyte Solutions (Butterworths Scientific J.-L. Liu, “Numerical methods for the Poisson-Fermi equation in elec-
Publications, London, 1959); (Dover Publications, New York, 2002). trolytes,” J. Comput. Phys. 247, 88–99 (2013).
2 23
K. S. Pitzer, Thermodynamics (McGraw Hill, New York, 1995). J.-L. Liu and B. Eisenberg, “Correlated ions in a calcium channel model: A
3
B. Hille, Ionic Channels of Excitable Membranes (Sinauer Associates Inc., Poisson-Fermi theory,” J. Phys. Chem. B 117, 12051–12058 (2013).
24
Sunderland, MA, 2001). J.-L. Liu and B. Eisenberg, “Poisson-Nernst-Planck-Fermi theory for mod-
4
K. J. Laidler, J. H. Meiser, and B. C. Sanctuary, Physical Chemistry (Houghton eling biological ion channels,” J. Chem. Phys. 141, 22D532 (2014).
25
Mifflin Co., Boston, 2003). J.-L. Liu and B. Eisenberg, “Poisson-Fermi model of single ion activities in
5
W. R. Fawcett, Liquids, Solutions, and Interfaces: From Classical Macroscopic aqueous solutions,” Chem. Phys. Lett. 637, 1–6 (2015).
26
Descriptions to Modern Microscopic Details (Oxford University Press, New J.-L. Liu, H.-j. Hsieh, and B. Eisenberg, “Poisson-Fermi modeling of the ion
York, 2004). exchange mechanism of the sodium/calcium exchanger,” J. Phys. Chem. B
6
G. Lebon, D. Jou, and J. Casas-Vázquez, Understanding Non-equilibrium 120, 2658–2669 (2016).
27
Thermodynamics: Foundations, Applications, Frontiers (Springer, 2008). J.-L. Liu, D. Xie, and B. Eisenberg, “Poisson-Fermi formulation of nonlo-
7
W. Kunz, Specific Ion Effects (World Scientific, Singapore, 2010). cal electrostatics in electrolyte solutions,” Mol. Based Math. Biol. 5, 116–124
8
J. H. Vera and G. Wilczek-Vera, Classical Thermodynamics of Fluid Systems: (2017).
28
Principles and Applications (CRC Press, 2016). J.-L. Liu and B. Eisenberg, “Poisson-Fermi modeling of ion activities in
9
P. Debye and E. Hückel, “Zur theorie der elektrolyte. I. Gefrierpunkt- aqueous single and mixed electrolyte solutions at variable temperature,”
serniedrigung und verwandte erscheinunge (The theory of electrolytes. I. J. Chem. Phys. 148, 054501 (2018).
29
Lowering of freezing point and related phenomena),” Phys. Zeitschr. 24, J. J. Bikerman, “Structure and capacity of electrical double layer,” Philos.
185–206 (1923). Mag. 33, 384–397 (1942).
30
10
E. Hückel, “Zur theorie konzentrierterer wässeriger lösungen starker B. P. Lee and M. E. Fisher, “Density fluctuations in an electrolyte from
elektrolyte,” Phys. Z. 26, 93–147 (1925). generalized Debye-Hueckel theory,” Phys. Rev. Lett. 76, 2906 (1996).
31
11
W. Voigt et al., “Quality assurance in thermodynamic databases for perfor- D. Bashford and D. A. Case, “Generalized Born models of macromolecular
mance assessment studies in waste disposal,” Pure and Applied Chemistry solvation effects,” Annu. Rev. Phys. Chem. 51, 129–152 (2000).
32
79, 883–894 (2007). A. A. Rashin and B. Honig, “Reevaluation of the Born model of ion
12
W. Voigt, “Chemistry of salts in aqueous solutions: Applications, experi- hydration,” J. Phys. Chem. 89, 5588–5593 (1985).
33
ments, and theory,” Pure Appl. Chem. 83, 1015–1030 (2011). W. W. Rudolph and G. Irmer, “Hydration of the calcium(II) ion in an aque-
13
D. Rowland, E. Königsberger, G. Hefter, and P. M. May, “Aqueous elec- ous solution of common anions (ClO4 − , Cl− , Br− , and NO3 − ),” Dalton Trans.
trolyte solution modelling: Some limitations of the Pitzer equations,” Appl. 42, 3919 (2013).
34
Geochem. 55, 170 (2015). I.-L. Chern, J.-G. Liu, and W.-C. Wang, “Accurate evaluation of electro-
14
G. M. Kontogeorgis, B. Maribo-Mogensen, and K. Thomsen, “The Debye- statics for macromolecules in solution,” Methods Appl. Anal. 10, 309–328
Hückel theory and its importance in modeling electrolyte solutions,” Fluid (2003).
35
Phase Equil. 462, 130–152 (2018). W. Geng, S. Yu, and G. Wei, “Treatment of charge singularities in implicit
15
D. Fraenkel, “Simplified electrostatic model for the thermodynamic solvent models,” J. Chem. Phys. 127, 114106 (2007).
excess potentials of binary strong electrolyte solutions with size-dissimilar 36
C.-L. Li and J.-L. Liu, “Analysis of generalized Debye-Hückel equation
ions,” Mol. Phys. 108, 1435 (2010). from Poisson-Fermi theory,” arXiv:1808.02757, 2018.
16 37
I. Borukhov, D. Andelman, and H. Orland, “Steric effects in electrolytes: A M. Valiskó and D. Boda, “Unraveling the behavior of the individual ionic
modified Poisson-Boltzmann equation,” Phys. Rev. Lett. 79, 435–438 (1997). activity coefficients on the basis of the balance of ion-ion and ion-water
17 interactions,” J. Phys. Chem. B 119, 1546 (2015).
C. D. Santangelo, “Computing counterion densities at intermediate cou-
pling,” Phys. Rev. E 73, 041512 (2006). 38
G. Wilczek-Vera, E. Rodil, and J. H. Vera, “On the activity of ions and
18 the junction potential: Revised values for all data,” AIChE. J. 50, 445
A. A. Kornyshev, “Double-layer in ionic liquids: Paradigm change?,” J. Phys.
Chem. B 111, 5545–5557 (2007). (2004).

AIP Advances 9, 015214 (2019); doi: 10.1063/1.5081863 9, 015214-6


© Author(s) 2019

You might also like