ArticlePDF Available

Electrolytes at charged interfaces: Pair integral equation approximations for model 2–2 electrolytes

AIP Publishing
The Journal of Chemical Physics
Authors:

Abstract and Figures

The structure and thermodynamics for model 2–2 electrolytes at a charged interface have been determined by the so-called “pair” approximation of integral equation theory. In addition to Coulombic interactions, the potential models for the ion–ion and ion–wall interactions employ “soft” continuous potentials rather than “hard”-sphere or “hard”-wall potentials. The solvent is modeled as a structureless dielectric continuum at 25 °C. The structure is calculated using the inhomogeneous Ornstein–Zernike relation, together with the hypernetted chain closure and two choices for the functional relationship between the singlet and pair correlation functions. Both the interfacial density profile and the inhomogeneous pair correlation functions are calculated. Some thermodynamic properties of these systems are also evaluated. The results of the pair approximation are compared with the so-called “singlet” approximation, selected computer simulation results, Gouy–Chapman–Stern predictions, and experimental data. While qualitative agreement is generally found between the two levels of integral equation approximation, measurable quantitative improvements exist for both structural and thermodynamic predictions in the pair approximation. © 2001 American Institute of Physics.
Content may be subject to copyright.
Electrolytes at charged interfaces: Pair integral equation approximations
for model 22 electrolytes
Andrew C. Eaton
School of Chemistry, University of Sydney, NSW 2006 Australia
A. D. J. Haymet
a)
Department of Chemistry, University of Houston, Houston, Texas 77204
Received 19 December 2000; accepted 5 April 2001
The structure and thermodynamics for model 22 electrolytes at a charged interface have been
determined by the so-called ‘‘pair’’ approximation of integral equation theory. In addition to
Coulombic interactions, the potential models for the ionion and ionwall interactions employ
‘‘soft’’ continuous potentials rather than ‘‘hard’’-sphere or ‘‘hard’’-wall potentials. The solvent is
modeled as a structureless dielectric continuum at 25°C. The structure is calculated using the
inhomogeneous OrnsteinZernike relation, together with the hypernetted chain closure and two
choices for the functional relationship between the singlet and pair correlation functions. Both the
interfacial density profile and the inhomogeneous pair correlation functions are calculated. Some
thermodynamic properties of these systems are also evaluated. The results of the pair approximation
are compared with the so-called ‘‘singlet’’ approximation, selected computer simulation results,
GouyChapmanStern predictions, and experimental data. While qualitative agreement is generally
found between the two levels of integral equation approximation, measurable quantitative
improvements exist for both structural and thermodynamic predictions in the pair approximation.
© 2001 American Institute of Physics. DOI: 10.1063/1.1375141
I. THE ELECTRICAL DOUBLE LAYER
The region where an electrolyte solution meets a charged
solid surface, resulting in a charge separation at the bound-
ary, is often referred to as the electrical double layer,
1–5
since
order is induced in the liquids next to the charged surface.
This phenomenon occurs in many naturally-occurring and
electrochemical systems, such as colloids, micelles, mem-
branes and the electrodeelectrolyte interface, and can have
profound effects on the chemistry and physics of such
systems.
6–11
In this paper we solve the ‘‘pair’’ approximation for the
electrical double layer for symmetric models of 22 electro-
lytes, using ‘‘soft’’ continuous potentials. In this approach,
we solve directly for the inhomogeneous ion-wall and ion
ion correlations functions. The aqueous solvent is approxi-
mated by a structureless dielectric continuum of fixed dielec-
tric constant. We require as input only the structure of bulk
electrolytes from work by Duh and Haymet.
12
We calculate
concentration profiles, inhomogeneous correlation functions,
and a range of thermodynamic quantities, in order to make
comparison with the ‘‘singlet’’ approximation for the same
electrolyte model reported earlier by us.
13
The calculation of the structure of an electrolyte near a
planar surface, both charged and uncharged, is the principal
topic considered in this paper. To show the value of this
paper, we display, in Fig. 1, the normalized density profiles
calculated using AHNC the pair approximation discussed be-
low are compared with simulation data based on the charged
hard-sphere/charged hard-wall CHS/CHWpotential model.
Allowing for the different potential functions used, this com-
parison shows good agreement. It is worth noting once again
the dramatic effect the surface has on the structure, since all
of the functions shown are unity in the absence of the sur-
face. Absent from this figure are any data from the singlet
approximation, which calculates the density profile for an
electrolyte but assumes that ionion correlations are unaf-
fected by the surface. For this case, the singlet approximation
is unable to provide a solution. Hence the need for the kind
of pair approximations discussed in this paper.
The pair approximation, as applied to electrolytes next to
charged surfaces, was first detailed by Henderson and
Plischke
14–16
for an electrolyte next to a single surface, and
by Kjellander and co-workers
17–21
for an electrolyte between
two surfaces. More recently, the pair approximation has been
further applied in wider contexts by Kjellander and
co-workers.
22–26
Work on the charged surface/electrolyte in-
terface has largely used the charged hard-sphere/charged
hard-wall potential model.
An excellent review of theoretical contributions to the
understanding of the electrical double layer up to 1995 has
been presented by Attard.
27
A compact and informative deri-
vation of both pair and singlet approximations has also been
presented recently by our group.
28
This paper is organized as follows: The equations used
in this work are presented in Sec. II, with numerical details
in Sec. III. Our structural and thermodynamic predictions are
collected in Secs. IV and V. In Sec. VI we present our dis-
cussion and conclusions.
a
Author to whom correspondence should be addressed.
JOURNAL OF CHEMICAL PHYSICS VOLUME 114, NUMBER 24 22 JUNE 2001
109380021-9606/2001/114(24)/10938/10/$18.00 © 2001 American Institute of Physics
Downloaded 16 Jan 2004 to 129.171.128.66. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
II. SUMMARY OF PAIR EQUATIONS AND INPUT
The pair approximation that we have solved can be sum-
marized in terms of a definition and two equations: i the
inhomogeneous OrnsteinZernike OZ relation, which de-
fines the direct correlation function c(r) from the total cor-
relation function h(r); iithe hypernetted chain HNC clo-
sure, an approximate relation between the direct and total
correlation functions; and iii a density equation, a func-
tional relationship between the inhomogeneous correlation
functions and the singlet density profile. We emphasize that
the exact OZ relation serves purely as a definition of the
direct correlation function.
In previous works, the LovettMouBuffWertheim
LMBW equation,
29
TriezenbergZwanzig TZ equation
30
which is equivalent to the LMBW equation, or the Yvon
BornGreen YBG equation
31
have been employed as the
functional relation between the pair and singlet correlation
functions. We have used the LMBW equation in this work.
Additionally, we also use an alternative relation called the
AHNC approximation. This equation has been previously
derived in a somewhat lengthy manner by Kjellander and
co-workers.
17
More recently, our group has shown that it
may be obtained by minimization of an approximate free
energy functional for the grand potential.
28
The equations for
the pair approximations that we have solved in this work are
summarized below.
A. The inhomogeneous OZ relation
For a fluid of
species next to an infinite planar solid
surface with an external potential applied normal to the sur-
face, denoted henceforth as the charged ‘‘wall,’’ the inhomo-
geneous OZ relation between the total and direct correlation
functions is
h
wij
z
1
,z
2
,s
12
c
wij
z
1
,z
2
,s
12
k1
0
dz
3
wk
z
3
dx
3
dy
3
c
wik
z
1
,z
3
,s
13
h
wkj
z
3
,z
2
,s
23
. 1
Figure 2 depicts the cylindrical geometry used in Eq. 1.
The wall lies in the xy-plane. Particles 1 and 2 are at per-
pendicular distances z
1
and z
2
from the wall and are sepa-
rated by a distance R
12
R. Since this system is cylindrically
symmetric, it is useful to define the length
s s
12
x
1
x
2
2
y
1
y
2
2
. 2
By Pythagoras’ theorem
s
2
R
2
z
1
z
2
2
. 3
The inhomogeneous pair correlation function between
the wall w and species i and j is
g
wij
z
1
,z
2
,s
12
h
wij
z
1
,z
2
,s
12
1. 4
The bulk number density of species i and number density of
species i at a perpendicular distance z from the surface are
i
B
and
wi
(z), respectively. This is one of the system of
equations solved numerically in this work. This equation has
been called the OZ2 relation in the literature of inhomoge-
neous fluids.
16
We choose to retain the redundant subscript
w in our equation to explicitly distinguish the inhomoge-
neous correlation functions from their bulk counterparts
g
ij
(r) and c
ij
(r).
Using the 2D Fourier Hankel transformation
32,33
de-
fined by
f
ˆ
k
2
0
dssf
s
J
0
ks
,
5
f
s
1
2
0
dkkf
ˆ
k
J
0
ks
,
Equation 1 can be written in Fourier space as
h
ˆ
wij
z
1
,z
2
,k
cˆ
wij
z
1
,z
2
,k
k1
v
0
dz
3
wk
z
3
cˆ
wik
z
1
,z
3
,k
h
ˆ
wkj
z
3
,z
2
,k
6
FIG. 1. a Cationic g
w
(z) and b anionic g
w
(z) normalized concentra-
tion profiles for a 0.05 M 22 electrolyte at a surface charge density of
8.65
Ccm
2
: pair using AHNC approximation solid line; pair using
LMBW approximation dashed line; and the MC results circles of Torrie
and Valleau Ref. 46, using a CHS/CHW potential.
FIG. 2. The geometry used in the pair approximation. The charged wall lies
in the xy-plane. Particles 1 and 2 are at perpendicular distances z
1
and z
2
from the flat surface and are separated by a distance R
12
. We define s
12
using the relation s
12
2
(x
1
x
2
)
2
(y
1
y
2
)
2
, where x
i
and y
i
are x and y
coordinates of particle i.
10939J. Chem. Phys., Vol. 114, No. 24, 22 June 2001 Electrolytes at charged interfaces
Downloaded 16 Jan 2004 to 129.171.128.66. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
where J
0
is the zeroth order Bessel function. We now only
need to perform the one-dimensional integrals in Eq. 6.
Due to the long-range nature of direct and total inhomo-
geneous correlation functions, an arbitrary decomposition of
the direct and total correlation functions into short-range
parts denoted by the superscript ‘‘s’’ and long-range parts,
q and
, respectively, must be introduced to renormalize the
equations. The decomposition is
h h
s
q,
c c
s
, 7
h
s
c
s
.
Both
and q are usually chosen to have convenient analyti-
cal forms that yield well-behaved, short-ranged forms for c
s
and h
s
. We introduce
as a convenient variable to be used in
conjunction with c
s
to solve the pair approximation. Two
choices of
and q will be discussed in the next section. The
numerical details for solving the OrnsteinZernike equation
are described in Ref. 34. Note that we do not include the
asymptotic form discussed by Ulander and Kjellander,
21
which may be important at very high concentrations.
B. The HNC closure
The HNC closure may be derived by discarding high
order entropy contributions from the functional expansion of
the free energy.
28
The HNC closure in real space, also pre-
sented as Eq. 17 by Bush et al.,
28
is
h
wij
z
1
,z
2
,s
12
1 exp
u
ij
R
12
h
wij
z
1
,z
2
,s
12
c
wij
z
1
,z
2
,s
12
, 8
where u
ij
(R
12
) is the ionion potential for the interparticle
separation R
12
,
1
kT, k is the Boltzmann constant, and
T is the absolute temperature. Using the definitions in Eq.
7, Eq. 8 can be written as
c
wij
s
z
1
,z
2
,s
12
exp
u
ij
R
12
wij
z
1
,z
2
,s
12
q
wij
z
1
,z
2
,s
12
wij
z
1
,z
2
,s
12
1
wij
z
1
,z
2
,s
12
q
wij
z
1
,z
2
,s
12
. 9
Two analytical choices of
and q have been used in the
literature by Henderson and Plischke,
14
and by Kjellander
and co-workers.
19
A natural choice for
is setting it equal to
the negative dimensionless Coulombic part of the ionion
potential.
19
The Coulombic part of the potential has the form,
ij
R
12
q
i
q
j
R
12
, 10
where q
i
is the charge on species i, and the solvent is treated
as a structureless dielectric continuum of dielectric constant
4
r
0
, with
0
the permittivity of free space and
r
78.358 is the relative permittivity chosen to model water at
the temperature 25 °C and density 0.997 gcm
3
. In this work
the wall is assumed to have the same relative permittivity as
the solvent, and, thus, no image effects are considered.
In order to avoid numerical difficulties arising from
this choice of the long-range potential for small values of
z
1
z
2
, Henderson
15
proposed a slightly altered form of
,
wij
z
1
,z
2
,s
12
⫽⫺
ij
z
1
,z
2
,s
12
1 e
[s
12
2
(z
1
z
2
)
2
]
1/2
3
, 11
where is chosen such that, for the most distant point R
N
from the wall analyzed, e
[R
N
2
(z
1
z
2
)
2
]
1/2
is arbitrarily
small. Similarly, q is defined as
q
wij
z
1
,z
2
,s
12
wij
z
1
,z
2
,s
12
e
[s
12
2
(z
1
z
2
)
2
]
1/2
,
12
where the inverse Debye length
1
is given by
2
4
i1
v
i
B
(q
i
2
/d), and d is the hydrated ion size. We
note that the Hankel transforms of 1/
s
2
b
2
1/2
and
exp(a
s
2
b
2
1/2
)/
s
2
b
2
1/2
are 2
exp(
b
k)/k and
2
exp(b
k
2
a
2
1/2
)/
k
2
a
2
1/2
, respectively. We have
adopted the decomposition scheme of Henderson
15
for the
results reported in this paper. Results based upon the decom-
position scheme of Kjellander
19
are discussed in Ref. 34.
C. Density equations: The AHNC approximation
The OZ and HNC equations make up only two of the
three equations required to solve a fully self-consistent pair
approximation. A third equation may be derived by minimi-
zation of an approximate free energy functional for the grand
potential, known as the AHNC approximation.
28
One version of the AHNC density equation, obtained
from Bush et al.,
28
is
ln
wi
z
1
i
B
⫽⫺
q
i
z
1
␤␾
wi
LJ
z
1
1
2
1 c
wii
z
1
,z
1
,0
i
(ex)
j 1
v
0
dz
2
wj
z
2
q
i
q
j
z
1
z
2
2
0
r
cˆ
wij
*
z
1
,z
2
,0
0
ds
12
s
12
h
wij
2
z
1
,z
2
,s
12
, 13
where
i
(ex)
1
2
1 c
ii
0
兲兲
2
j 1
v
j
B
0
drr
2
c
ij
*
r
1
2
h
ij
r
2
, 14
c
ij
*
r
c
ij
r
q
i
q
j
4
0
r
r
, 15
c
wij
*
z
1
,z
2
,s
12
c
wij
z
1
,z
2
,s
12
q
i
q
j
4
0
r
s
12
2
z
1
z
2
2
1/2
, 16
10940 J. Chem. Phys., Vol. 114, No. 24, 22 June 2001 A. C. Eaton and A. D. J. Haymet
Downloaded 16 Jan 2004 to 129.171.128.66. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
q
i
z
1
⫽⫺
␲␴
z
1
2
0
r
q
i
0
2
q
i
j 1
v
0
dz
2
q
j
wj
z
2
z
1
z
2
2
0
r
, 17
and
0
0
⫽⫺
0
r
j 1
v
0
dz
2
q
j
z
2
wj
z
2
. 18
Defining t
wi
(z) ln
g
wi
(z)
␤␾
wi
LJ
(z), Eq. 13can be written
with t
wi
(z) as the subject,
t
wi
z
⫽⫺
1
2
q
i
0
1
2
1 c
wii
z
1
,z
1
,0
i
(ex)
j 1
v
0
dz
2
wj
z
2
q
i
q
j
z
1
z
2
2
0
r
cˆ
wij
*
z
1
,z
2
,0
0
ds
12
s
12
h
wij
2
z
1
,z
2
,s
12
.
19
D. Density equations: The LMBW equation
The LovettMouBuffWertheim LMBW equation is
a popular choice
15,22
to complement the OZ and HNC equa-
tions, but it does not lead to a consistent set of equations in
the sense of being a minimum of an approximate free energy
surface.
28
This equation is derived from balancing the gradi-
ent of the density with the force of the external potential and
the forces within the double layer.
35,36,27
There are a number
of formally equivalent versions of the LMBW equation.
28
The version we have used in numerical work is
ln
wi
z
1
兲兲
z
1
u
wi
z
1
z
1
j 1
v
0
dz
2
⳵␳
wj
z
2
z
2
q
i
q
j
z
1
z
2
2
0
r
cˆ
wij
*
z
1
,z
2
,0
. 20
Defining u
wi
(z)
wi
LJ
(z)
wi
(z), where
wi
(z)
⫽⫺
␲␴
z
/2
0
r
, with t
wi
(z) as the subject, Eq. 20 be-
comes
t
wi
z
1
z
1
wi
z
1
z
1
j 1
v
0
dz
2
⳵␳
wj
z
2
z
2
q
i
q
j
z
1
z
2
2
0
r
cˆ
wij
*
z
1
,z
2
,0
. 21
E. Model potentials
The model two-component bulk electrolyte used in our
study was introduced by Rossky and Friedman and
co-workers,
37,38
and is the same as that used earlier by us to
study the singlet approximation.
13
This model ionic solution
is that of equal-sized, spherically symmetric, ‘‘soft’’ ions,
interacting via the potential energy,
u
ij
r
kB
d
ij
d
ij
r
9
q
i
q
j
r
. 22
The parameters are chosen to be B5377.75
q
i
q
j
Å K, and
d
ij
2.8428 Å, the same as those used by Rossky and
co-workers.
37
This potential for a 22 electrolyte is dis-
played in Fig. 3a兲共solid lines.
The short-range wallion potential employed in these
calculations is a Lennard-Jones LJ 9-3 potential, obtained
by integrating a LJ 12-6 potential over the half-space of wall
atoms, as described by Steele,
39
wi
z
2
3
wi
2
15
d
wi
z
9
d
wi
z
3
, 23
where
wi
is the depth of the LJ 12-6 potential energy well
and d
wi
is the LJ 12-6 distance parameter. The value of d
wi
is chosen to be 4.2 Å, the hydrated ion size d, which corre-
sponds to the location of the minimum in the bulk potential
energy. The 9-3 well depth
min
is adjusted by the parameter
wi
. This parameter is chosen so that, at zero surface poten-
tial, the normalized concentration profile is approximately
equal to one for all z d/2. Consequently, the value of the LJ
9-3 reduced well depth
min
/kT is chosen to be 0.372 84 for
all calculations presented. Our short-range wallion potential
is shown in Fig. 3b兲共solid line. In the double layer litera-
ture, the majority of calculations that we compare with, par-
ticularly any Monte Carlo MC data we present, use a
‘‘hard-wall’’ potential, as shown in Fig. 3b兲共dashed line.
F. Surface thermodynamics
Once ionic concentration profiles have been determined,
thermodynamic quantities can be calculated. Direct compari-
son of the profiles with experimental data is not yet possible,
but it is possible to compare some thermodynamic quantities.
We collect here the formulas for the surface thermodynamics
properties for a two component electrolyte near a wall at a
surface potential
0
and a surface charge density
.
We define the normalized concentration profile g
wi
(z)
by
FIG. 3. a Bulk dimensionless potential energy u
ij
(r)/kT for the model
22 electrolyte solid lines at T25 °C. The dashed line is a restricted
primitive model RPM potential. b Short-range LJ 9-3 wallion dimen-
sionless potential energy
wi
(z)/kT solid line at T 25 °C. The dashed
line is a ‘‘hard-wall’’ potential. Note the definition of
min
and that the
vertical scale in b is one order of magnitude less than in a.
10941J. Chem. Phys., Vol. 114, No. 24, 22 June 2001 Electrolytes at charged interfaces
Downloaded 16 Jan 2004 to 129.171.128.66. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
wi
z
/
i
B
g
wi
z
1 h
wi
z
, 24
where h
wi
(z) is the singlet wallion total correlation func-
tion. The cationic
and anionic
surface excess can be
expressed as integrals over the respective singlet total corre-
lation functions h
w
(z) and h
w
(z),
i
0
dzh
wi
z
, i⫽⫹, . 25
The surface charge density
, used as input for our cal-
culations, can be calculated across the interfacial region,
⫽⫺
i⫽⫹,
q
i
i
B
i
. 26
Combining Eqs. 17, 18, and 26, the mean electrostatic
potential may be rewritten as
z
0
r
i1
v
q
i
z
dt
z t
wi
t
. 27
Electrocapillary curves are plots of the surface free energy
change
as the charge on the electrode is increased. The
change in surface free energy is calculated from the equation,
⫽⫺
0
0
d
. 28
The electrical double layer is comprised of separated layers
of charge, analogous to a conventional electrical capacitor. It
is a region where charge can be stored by varying the surface
potential. The differential capacitance C is defined to be
C
d
d
0
, 29
and has been calculated by performing a numerical differen-
tiation on a plot of
against
0
.
G. Contact theorem
The contact theorem provides a consistency check for
the accuracy of approximate theories. For a charged hard-
wall, the contact theorem connects the values of the ionic
densities at contact
wi
(d/2) with the bulk osmotic pressure
p and the surface charge density
. For a soft short-range
wallion potential the ‘‘contact’’ theorem for an exact
theory is
40,41
kT
i⫽⫹,
wi
0
p
2
␲␴
2
i⫽⫹,
0
dz
wi
z
d
wi
z
dz
. 30
For the HNC approximation, an alternative contact condition
may be derived,
42,43
kT
i⫽⫹,
wi
0
1
2
kT
i⫽⫹,
i
B
1
2
␲␴
2
i⫽⫹,
0
dz
wi
z
d
wi
z
dz
, 31
where the isothermal compressibility
is calculated using
kT
i⫽⫹,
i
B
1
kT
i,j⫽⫹,
i
B
j
B
drc
ij
*
r
. 32
Henderson
15
stated that pair calculations using the LMBW
equation ‘‘appear to satisfy’’ Eq. 31. Recalling that the
right-hand side of either contact theorem is necessarily zero
for our potential, we have found numerically that the AHNC
approximation satisfies the exact contact theorem as it
must.
28
In its dimensionless form, numerically we find it
holds to within 10
1
for the majority of our calculations. The
pair approximation with the LMBW equation does not, for
the majority of calculations, satisfy the dimensionless form
of either Eq. 30 or 31 to within 10
1
.
III. NUMERICAL CONSIDERATIONS
The pair approximation must be solved numerically. The
input to our calculation is: i the short-range potential en-
ergy between the wall and the ions
wi
(z); ii the short-
range ionion potential energy
ij
(r); iii the temperature
T; ivthe bulk electrolyte concentration c
i
; vthe structure
of the bulk model electrolyte, contained in the direct and
total correlation functions, c
ij
(r) and h
ij
(r), respectively,
for all distinct pairs of species, ij⫽⫹⫹, ⫹⫺, and ⫺⫺,as
a function of distance r; and vi the surface charge density
.
The equations are solved iteratively using a solver NK-
SOL; the advantages of using such a solver are discussed by
Booth.
44
For all calculations, the temperature T 25°C. The
bulk correlation functions c
ij
(r) and h
ij
(r) are based on cal-
culations made by Duh and Haymet,
12
using the HNC clo-
sure. Discrete grids are used for the normal and radial direc-
tions. Integrals were calculated using the trapezoidal method
of integration.
In the z-direction, a nonlinear grid z
i
was used, where
each grid point is defined by z
i
exp
(ib)*
Z
a, where a
real, b integer, and
Z real were chosen to ensure that
density profiles are numerically zero at the first point z
1
and
within a few percent of the bulk density at a cutoff, z
cut
,
from the wall. The grid in the z-direction was broken into
two regions. The first section represented the grid between
the wall and an arbitrarily chosen cutoff, z
cut
, beyond which
the density of ions was assumed to be that of the bulk. A
second region was selected from z
cut
to z
max
, a point arbi-
trarily chosen to represent infinity in integrals calculated in
the z-direction. Any inhomogeneous correlation function
which involves an ion located between z
cut
and z
max
is as-
sumed to have a value of the bulk correlation function of the
same ionion separation R. In general, a total of approxi-
mately 50 to 80 points were used in the z-direction, with
about 4575 used between the wall and z
cut
.
The Hankel transforms expressed by Eq. 5 can be
approximated
45
by
10942 J. Chem. Phys., Vol. 114, No. 24, 22 June 2001 A. C. Eaton and A. D. J. Haymet
Downloaded 16 Jan 2004 to 129.171.128.66. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
f
ˆ
k
m
4
k
N
2
j 1
N
f
s
j
J
0
k
m
s
j
J
1
2
k
M
s
j
,
33
f
s
j
1
s
N
2
m1
N
f
ˆ
k
m
J
0
k
m
s
j
J
1
2
k
m
s
N
,
where J
1
is the first order Bessel function. The points in the
radial direction (k
1
¯ k
N
and s
1
¯ s
N
) are determined by
the conditions that J
0
(k
m
s
N
) J
0
(k
N
s
j
) 0 and s
N
is suit-
ably chosen so that all functions f(s
i
) approach zero as i
approaches N. For the majority of calculations, N 90, with
s
N
30 Å. The numerical details of the pair approximation
calculations are further discussed in Ref. 34.
IV. RESULTS: ION CONCENTRATION PROFILES
A. Comparison of AHNC approximation and
simulations
To our knowledge, no simulation data exist for the real-
istic ‘‘soft’’ potentials we have studied with pair and singlet
theory. As a guide, we have chosen to make comparisons
with the Monte Carlo MC data of Torrie and Valleau.
46,47
In their simulations, Torrie and Valleau use a different set of
potentials than us, namely charged hard-spheres e.g., the
potential in Fig. 3a, dashed linesfor the model electrolyte,
and a short-range hard-wall potential, as shown in Fig. 3b
dashed line.
Figure 4 shows the cationic and anionic normalized con-
centration profiles for 0.5 M 22 electrolyte at a surface
charge density of 21.29
Ccm
2
. The normalized concen-
tration profiles for the singlet approximation have been in-
cluded for comparison. The corresponding surface potential
for each of the integral equation approximations is listed
below in brackets. The circles are the digitized MC computer
simulation results. The solid lines are the pair results using
the AHNC ( 99 mV) approximation. The long-dashed lines
are the singlet approximation ( 100 mV).
Taking into consideration the differences in the potential
functions, the approximate integral equation theories predict
qualitatively the same behavior as the MC data. The pair
approximations show quantitative agreement with MC data,
particularly for the anionic profile, and a measurable im-
provement over the singlet approximation.
For divalent electrolytes at lower concentrations it is no
longer possible to obtain singlet solutions at the surface
charge densities used in the MC simulations. In Fig. 1 above
we compare the AHNC approximation solid lines,
69.6 mV) with the MC results circles for 22 electrolytes
at 0.05 M and a surface charge density of 8.65
Ccm
2
.
We also show the pair using the LMBW approximation
dashed lines, 70.3 mV) for comparison. Allowing for the
differences in potential models, the pair approximations pre-
dict qualitatively the same behavior as the MC data.
B. Comparison of AHNC and singlet approximations
We now consider comparisons of normalized concentra-
tion profiles for the singlet approximation, using our ‘‘soft’’
potential model, and the pair approximation using the AHNC
density equation. In Fig. 5, comparisons for cationic and an-
ionic normalized concentration profiles are made between
pair results using the AHNC approximation solid lines and
the singlet approximation dashed lines, for 22 electrolytes
at 0.5 M with surface charge densities of 0, 10 and
30
Ccm
2
. A surface charge density of 30
Ccm
2
represents the upper limit for obtaining solutions for the sin-
glet approximation.
With no surface charge on the wall, the profiles of the
singlet and pair approximations are similar, with the pair
approximation result showing slightly less structure. At a
surface charge density of 10
Ccm
2
, there is very little
difference in the two profiles, though, as the magnitude of
the surface charge increases, the singlet approximation pre-
dicts significantly more structure than the pair approxima-
tion, with secondary peaks developing in both the cationic
and anionic profiles at a surface charge density of
30
Ccm
2
. While both integral equations carry approxi-
mations, and hence conclusions drawn from differences in
their results must be made with some caution, the differences
in density profiles would seem to be a consequence of the
combination of approximations in the singlet plus HNC
approximation,
28
namely that the inhomogeneous correlation
FIG. 4. a Cationic g
w
(z) and b anionic g
w
(z) normalized concentra-
tion profiles for a 0.5 M 22 electrolyte at a surface charge density of
21.29
Ccm
2
: pair using AHNC approximation solid line; singlet
long-dashed line; and the MC results circles of Torrie and Valleau Ref.
46, using a CHS/CHW potential.
FIG. 5. a Cationic g
w
(z) and b anionic g
w
(z) normalized concentra-
tion profiles for a 0.5 M 22 electrolyte at surface charge densities of 0,
10, and 30
Ccm
2
: pair using AHNC approximation solid line;and
singlet approximation long-dashed line.
10943J. Chem. Phys., Vol. 114, No. 24, 22 June 2001 Electrolytes at charged interfaces
Downloaded 16 Jan 2004 to 129.171.128.66. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
functions are equivalent to their bulk counterparts. The sin-
glet approximation exaggerates the density profiles since the
charged liquid is not allowed to relax from its bulk structure
even very close to the charged surface. Indeed, for all con-
centrations studied, the general trend is that, for increasing
surface charge, the pair approximation predicts less ionic
density structure than the singlet approximation.
C. Results: Concentration dependence
We now study the effect of changing the concentration
for 22 electrolytes, with a surface charge density of
30
Ccm
2
. We have obtained solutions for concentra-
tions over 3 orders of magnitude in concentration, at 25 °C.
The values of the Debye length
1
and the calculated sur-
face potential are indicated in brackets: 0.005 M 21.49 Å,
162.7 mV), 0.05 M 6.80 Å, 149.6 mV), 0.5 M 2.15 Å,
130.5 mV) and 5.0 M 0.68 Å, 92.6 mV). In Fig. 6, the
cationic and anionic normalized density profiles are shown.
The major feature of the counterion profiles is the rapidly
increasing peak as the concentration is decreased. The an-
ionic profile does not show the same rapidly increasing peak
height. Despite the cationic peak height rapidly rising, the
surface potential only increases by moderate amounts. For
the 5.0 M concentration, strong oscillations are seen in the
density profiles. As the concentrations is lowered to 0.5 M,
the profiles only oscillate weakly. No oscillations are seen
for concentrations below 0.5 M.
V. RESULTS: ELECTROSTATICS AND SURFACE
THERMODYNAMICS
Although we have shown that significant differences be-
tween the pair and singlet approximations can occur in the
structure of the double layer, the impact of this finding would
be somewhat lessened were these differences not to translate
into noticeable changes in the electrostatics and thermody-
namics of the double layer. The quantities in this section
currently represent the strongest links between theory and
experiment, and their evaluation and comparison is vital to
an understanding of the double layer and theoretical models
that are applied to it.
A. Mean electrostatic potential
The surface charge density
is used as input for the pair
calculations. For a 0.5 M 22 electrolyte, the dependence of
surface potential on surface charge density is shown in Fig.
7. The pair results using the AHNC solid line and LMBW
dashed line approximations are compared with the singlet
approximation long-dashed line. There is no simulation or
experimental data for comparison to our knowledge.
Comparing the pair AHNCand singlet approximations,
it can be seen that there are two distinct regions. As the
charge is increased from zero to about 10
Ccm
2
, there is
little difference among any of the integral equation approxi-
mations. After 10
Ccm
2
, the singlet approximation pre-
dicts substantially less surface potential for the same surface
charge density.
The two pair approximations show similar behavior.
While the curve for the LMBW approximation predicts
slightly higher surface potentials than for the AHNC ap-
proximation, they remain similar over the entire range. An
interesting feature of both pair approximations is the linear
nature of the functions at surface charge densities greater
than 10
Ccm
2
.
Figure 8 shows the change in the mean electrostatic po-
tential of the AHNC approximation, as the surface charge
density is increased by 10
Ccm
2
from 10 to
80
Ccm
2
, indicated by alternating solid and dashed
lines. This corresponds to an increase in the surface potential
from 54.6 to 311.8 mV. At a low charge on the wall, the
mean electrostatic potential has a very shallow minimum at a
z value near 1.4d. As the wall charge increases to
40
Ccm
2
, the minimum becomes deeper and shifts to
about z 0.9d. As the charge on the wall continues to in-
crease up to 80
Ccm
2
, there is very little change in the
overall functional form of the mean electrostatic potential.
B. Ionic surface excess
Figure 9 shows the cationic surface excess charge den-
sity as a function of surface charge density. The solid line
FIG. 6. a Cationic g
w
(z) and b anionic g
w
(z) normalized concentra-
tion profiles for a wall with a surface charge density of 30
Ccm
2
for a
5.0 M 22 electrolyte dotteddashed line; 0.5 M 22 electrolyte solid
line; 0.05 M 22 electrolyte long-dashed line; and 0.005 M 22 electro-
lyte dashed line.
FIG. 7. The surface potential as a function of surface charge density for a
0.5 M 22 electrolyte: pair using the AHNC approximation solid line; pair
using the LMBW approximation dashed line; and the singlet approxima-
tion long-dashed line.
10944 J. Chem. Phys., Vol. 114, No. 24, 22 June 2001 A. C. Eaton and A. D. J. Haymet
Downloaded 16 Jan 2004 to 129.171.128.66. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
represents the pair approximation with the AHNC density
equation for a 0.5 M 22 electrolyte. The long-dashed line
shows the singlet approximation. The circles represent the
experimental data of Harrison, Randles and Schriffrin
48
for a
0.5 M aqueous solution of magnesium sulfate at 25°C, ob-
tained from electrocapillary measurements on a mercury-
aqueous electrolyte interface. To our knowledge, no simula-
tion data exist for relevant ‘‘soft’’ potentials. Merely for a
simple comparison, the GouyChapmanStern approxima-
tion GCS兲共for hard potentials
13
and is the short-dashed
line. The results of the LMBW approximation show no dis-
cernible difference from those of the AHNC approximation,
so they are not included here.
For a negative electrode, the curve of the AHNC ap-
proximation agrees well with both the experimental data and
the curves of the other theoretical approaches. At positive
surface charges, the results from the AHNC approximation
are quantitatively and qualitatively different from the exist-
ing experimental data. The pair approximation predicts little
change in the cationic surface excess charge density as the
wall becomes more positively charged. This is a direct result
of the lack of change in the coion density profile as the wall
gains charge, whereas the experimental results indicate sub-
stantial increases in the coion density for these surface
charges. The pair approximation for our ‘‘soft’’ potentials
does qualitatively predict similar behavior to the GCS ap-
proximation for hard potentials. The AHNC approximation
results also differ markedly from the results of the singlet
approximation, which may seem in this case to predict the
behavior of the experimental data better, possibly due to a
cancellation of errors. This unexpected result is probably an
indication of the limitations of the models that are being
used. Our models fail to account for the nature of the solvent
and the important role that it plays. The size of the magne-
sium ion relative to the sulfate ion may also be a factor in the
differences that have been seen, and the nature of the solid
surface, including any image interactions, cannot be ignored.
These will be explored in future work, together with connec-
tions to the work of Greberg and Kjellander.
26
C. Electrocapillary curves
The change in surface free energy resulting from charg-
ing the interface can be calculated by the numerical integra-
tion of the surface charge density as a function of the surface
potential, as shown in Eq. 28 above. In Fig. 10, the surface
free energy is plotted as a function of surface potential for
0.5 M 22 electrolyte, using the AHNC approximation. For
comparison, the 0.5 M 22 electrolyte predictions for the
singlet approximation are shown. The digitized experimental
data circles of Grahame
49
for NaCl at 18 °C is only in-
cluded to show the order of magnitude of this quantity in the
absence of experimental data for a 0.5 M 22 electrolyte.
Figure 10 indicates that the predictions of the pair approxi-
mation are in good agreement with those of the singlet ap-
proximation. This indicates that the surface free energy is
relatively insensitive to the differences in the fluid structure
near the solid surface.
D. Differential capacitance
Given the charge separation between the interface and
the electrolyte solution, the double layer may be viewed as a
FIG. 8. The change in the mean electrostatic potential, as a function of
distance from the wall, as the surface charge density is increased from 10
to 80
Ccm
2
, in increments of 10
Ccm
2
, for the AHNC approxi-
mation.
FIG. 9. The cationic surface excess charge density as a function of surface
charge density, using the AHNC approximation solid line, for a 0.5 M 22
electrolyte. Also shown are the singlet approximation long-dashed line,
and the GouyChapmanStern approximation GCS兲共short-dashed line.
For comparison, the experimental data of Harrison, Randles and Schriffrin
circles are shown for a 0.5 M aqueous solution of magnesium sulfate at
25 °C.
FIG. 10. The surface free energy is plotted as a function of surface potential
for a 0.5 M 22 electrolyte, using the AHNC approximation solid line. For
comparison, the 0.5 M 22 electrolyte predictions for the singlet approxi-
mation long-dashed line are shown. The digitized experimental data
circles of Grahame Ref. 49 for the 11 electrolyte NaCl at 18 °C is
included merely to show the order of magnitude of this quantity in the
absence of experimental data for a 0.5 M 22 electrolyte.
10945J. Chem. Phys., Vol. 114, No. 24, 22 June 2001 Electrolytes at charged interfaces
Downloaded 16 Jan 2004 to 129.171.128.66. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
parallel plate capacitor. This phenomenon has drawn much
interest recently.
50,51
The differential capacitance is calcu-
lated using Eq. 29. Figure 11 shows the differential capaci-
tance as a function of the surface charge density for a 0.5 M
22 electrolyte for the pair AHNC approximation solid
line. Also shown are our earlier results the singlet approxi-
mation long-dashed line. No simulation data are available
for soft potentials. To provide some basis for comparison, we
include hard potential based GCS predictions dotted
dashed line and digitized data from Teran et al.
52
short-
dashed line, calculated for the RPM/charged hard wall
model within the hypernetted chain/mean spherical approxi-
mation singlet approximation. The results of the latter two
are discussed at greater length in our previous paper.
13
Note
that nothing in this figure alters the fact that the pair approxi-
mation is believed to have the best accuracy, the singlet next
in accuracy, and the GCS last. The predictions of the LMBW
approximations and the AHNC approximations do not differ
noticeably and so the LMBW results have not been included
here.
It can be seen that for soft potentials, the pair and singlet
approximation vary in their functional form, with qualita-
tively different behavior. The singlet approximation predicts
a rapidly increasing capacitance as the wall is charged. Con-
versely, the pair approximation shows a slowly increasing
capacitance as the interface becomes more charged, almost
to the point of being constant at higher charges. The capaci-
tance is dependent on the surface potentials of the pair and
singlet approximations as a function of the surface charge
density, but, unlike the electrocapillary curves, is very sensi-
tive to the differences in the density profiles of the pair and
singlet approximations, most particularly, the lack of change
in the coion density profile of the pair approximation.
VI. CONCLUSIONS
In this paper, we have calculated the structure, electro-
statics and thermodynamics of electrolytes using two pair
approximations, the LMBW and AHNC approximations.
The density profiles and thermodynamics produced by both
approximations were found to be similar for all electrolyte
concentrations considered. The pair approximations predict
ionic density profiles that are qualitatively, and frequently
quantitatively, in good agreement with MC simulation data,
whereas the singlet approximation generally predicts concen-
tration profiles with more structure than the pair approxima-
tion as the surface charge is increased.
Integral equations provide some interesting insights into
electrostatics and thermodynamics. For example, the surface
free energy is found to be relatively unaffected by the differ-
ences in the density profiles predicted by the different inte-
gral equations. The differential capacitance is extremely sen-
sitive to the type of integral equation used. The electrostatic
potentials show some differences between the singlet and
pair approximations. The ionic surface excess plot showed a
noticeable difference in the behavior of the pair and singlet
approximations. The experimental data for this quantity was
significantly different from the predictions of both integral
equations. This highlights some of the shortcomings of the
simple models we have used, in particular, the neglect of
discrete solvent molecules, the simplification of same ionic
radii and the lack of a realistic solid surface. These are not
new considerations, but are being revisited with renewed
vigor after some neglect.
53–58
The successful application of
the pair approximations to electrolytes next to a charged sur-
face, opens the door for their use on more complex systems,
such as those with polarizable potentials like the central
force models of water.
59
ACKNOWLEDGMENTS
This research was supported in part by the Australian
Research Council ARC兲共Grant No. A29530010, and in
part by the Welch Foundation Grant No. E1429.Weac-
knowledge gratefully many helpful conversations on this
topic with Dr. Stephan Marc
˘
elja ANU and Dr. Roland
Kjellander Gothenburg. Computational aspects of this re-
search have also been supported by ANUSF and the
NSWCPC. A.D.J.H. thanks Dr. Grant Goodyear and Dr. Jo-
han Ulander for a critical reading of the manuscript.
1
H. Helmholtz, Ann. Phys. Leipzig 7, 337 1879.
2
J. S. Rowlinson, ‘‘Development of theories of inhomogeneous fluids,’’ in
Fundamentals of Inhomogeneous Fluids, edited by D. Henderson Marcel
Dekker, New York, 1992, Chap. 1.
3
G. Gouy, J. Phys. Paris, Colloq. 9,4571910.
4
D. L. Chapman, Philos. Mag. 25, 475 1913.
5
O. Stern, Z. Elektrochem. Angew. Phys. Chem. 30, 508 1924.
6
N. T. Skipper, M. V. Smalley, G. D. Williams, A. K. Soper, and C. H.
Thompson, J. Chem. Phys. 99, 14201 1995.
7
J. Israelachvili and H. Wennerstro
¨
m, Nature London 379,2191996.
8
J. P. Quirk and S. Marc
ˇ
elja, Langmuir 13, 6241 1997.
9
M. Ullner, C. E. Woodward, and B. Jo
¨
nsson, J. Chem. Phys. 105, 2056
1996.
10
G. Toikka and R. A. Hayes, J. Colloid Interface Sci. 191,1021997.
11
M. J. Weaver, J. Phys. Chem. 100, 13079 1996.
12
D. M. Duh and A. D. J. Haymet, J. Chem. Phys. 97, 7716 1992.
13
M. J. Booth, A. C. Eaton, and A. D. J. Haymet, J. Chem. Phys. 103,417
1995.
14
M. Plischke and D. Henderson, J. Chem. Phys. 88, 2712 1988.
15
D. Henderson and M. Plischke, J. Phys. Chem. 92, 7177 1988.
16
D. Henderson, ‘‘Integral equation theories for inhomogeneous fluids,’’ in
Fundamentals of Inhomogeneous Fluids, edited by D. Henderson Marcel
Dekker, New York, 1992, Chap. 4.
17
R. Kjellander and S. Marc
˘
elja, J. Chem. Phys. 82, 2122 1985.
18
R. Kjellander and S. Marc
˘
elja, Chem. Phys. Lett. 127,4021986.
FIG. 11. The differential capacitance as a function of the surface charge
density for a 0.5 M 22 electrolyte for the pair AHNC approximation
solid line. Also shown is the singlet approximation long-dashed line,
GCS predictions dotteddash line and digitized data from Teran et al.
Ref. 52兲共short-dashed line, using a CHS/CHW potential with a singlet
approximation.
10946 J. Chem. Phys., Vol. 114, No. 24, 22 June 2001 A. C. Eaton and A. D. J. Haymet
Downloaded 16 Jan 2004 to 129.171.128.66. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
19
R. Kjellander, J. Chem. Phys. 88, 7129 1988, and references therein.
20
R. Kjellander, T. Åkesson, B. Jo
¨
nsson, and S. Marc
ˇ
elja, J. Chem. Phys.
97,14241992.
21
R. Kjellander and J. Ulander, J. Phys. Paris, Colloq. 10, 431 2000.
22
H. Greberg and R. Kjellander, Mol. Phys. 83,7891994.
23
J. Ennis, S. Marc
ˇ
elja, and R. Kjellander, Electrochim. Acta 41, 2115
1996.
24
H. Greberg, R. Kjellander, and T. Åkesson, Mol. Phys. 87, 407 1996.
25
H. Greberg, R. Kjellander, and T. Åkesson, Mol. Phys. 92,351997.
26
R. Kjellander and H. Greberg, J. Electroanal. Chem. 462, 273 1999.
27
P. Attard, Advances in Chemical Physics, edited by I. Prigogine and S. A.
Rice Wiley, New York, 1996, Vol. XCII.
28
M. R. Bush, M. J. Booth, A. Haymet, and A. Schlijper, Mol. Phys. 95, 601
1998.
29
J. Quintana, D. Henderson, and M. Plischke, J. Phys. Chem. 93,4304
1989.
30
D. G. Triezenberg and R. Zwanzig, Phys. Rev. Lett. 28, 1183 1972.
31
J. P. Hansen and I. R. McDonald, Theory of Simple Liquids Academic,
London, 1986.
32
R. V. Churchill, Operational Mathematics, 3rd ed. McGrawHill, New
York, 1972.
33
H. Bateman, Tables of Integral Transforms, edited by A. Erde
`
lyi
McGrawHill, New York, 1954, Vol. II.
34
A. C. Eaton, ‘‘Electrolytes at Charged Interfaces Integral Equation Ap-
proach,’’ Ph.D. thesis, School of Chemistry, University of Sydney, Aus-
tralia, 2000.
35
R. Lovett, C. Y. Mou, and F. P. Buff, J. Chem. Phys. 65, 570 1976.
36
M. S. Wertheim, J. Chem. Phys. 65, 2377 1976.
37
P. J. Rossky, J. B. Duduwicz, B. L. Tembe, and H. L. Friedman, J. Chem.
Phys. 73, 3372 1980.
38
R. Bacquet and P. J. Rossky, J. Chem. Phys. 79, 1419 1983.
39
W. A. Steele, The Interaction of Gases with Solid Surfaces Pergamon,
Oxford, 1974.
40
D. Henderson, L. Blum, and J. L. Lebowitz, J. Electroanal. Chem. 102,
315 1979.
41
S. L. Carnie and D. Y. C. Chan, J. Chem. Phys. 74, 1293 1981.
42
S. L. Carnie, D. Y. C. Chan, D. J. Mitchell, and B. W. Ninham, J. Chem.
Phys. 74, 1472 1981.
43
L. Blum and D. Henderson, ‘‘Statistical mechanics of electrolytes at in-
terfaces,’’ in Fundamentals of Inhomogeneous Fluids, edited by D. Hend-
erson Marcel Dekker, New York, 1992, Chap. 6.
44
M. J. Booth, A. G. Schlijper, L. E. Scales, and A. D. J. Haymet, Comput.
Phys. Commun. 119, 122 1999.
45
F. Lado, J. Chem. Phys. 49,30921968.
46
G. M. Torrie and J. P. Valleau, J. Phys. Chem. 86,32511982.
47
S. L. Carnie and G. M. Torrie, Adv. Chem. Phys. 56,1411984.
48
J. A. Harrison, J. E. B. Randles, and D. J. Schiffrin, J. Electroanal. Chem.
25,1971970.
49
D. C. Grahame, Chem. Rev. 41,4411947. The concentration of NaCl is
not specified. It appears to be a 1.0 M solution.
50
D. Boda, D. Henderson, and K.-Y. Chan, J. Chem. Phys. 110,5346
1999.
51
D. Boda and D. Henderson, J. Chem. Phys. 112, 8934 2000.
52
L. Mier y Tera
´
n, E.
´
az-Herrera, M. Lozada-Cassou, and D. Henderson,
J. Phys. Chem. 92, 6408 1988.
53
G. To
´
th, E. Spohr, and K. Heinzinger, Chem. Phys. 200,3471995.
54
D. Boda, K.-Y. Chan, and D. Henderson, J. Chem. Phys. 109,7362
1998.
55
J. Forsman, B. Jo
¨
nsson, and C. E. Woodward, J. Phys. Chem. 100, 15005
1996.
56
M. F. Toney et al., Nature London 368, 444 1994.
57
E.
´
az-Herrera and F. Forstmann, J. Chem. Phys. 102, 9005 1995.
58
M. F. Toney et al., Surf. Sci. 335, 326 1995.
59
M. J. Booth, D.-M. Duh, and A. D. J. Haymet, J. Chem. Phys. 101, 7925
1994, and references therein.
10947J. Chem. Phys., Vol. 114, No. 24, 22 June 2001 Electrolytes at charged interfaces
Downloaded 16 Jan 2004 to 129.171.128.66. Redistribution subject to AIP license or copyright, see http://ojps.aip.org/jcpo/jcpcr.jsp
Article
Classical density functional theory is combined with the extended primitive model and solvent primitive model to investigate how differential capacitance Cd of electrical double-layer formed inside cylindrical pore is influenced by solvent granularity, bulk concentration, counter-ion diameter, and neutral and non-hard sphere (HS) potential utail_αβ(r) between ion species, between solvent and counter-ion (and co-ion). Several main conclusions are drawn. (i) The Cd−surface charge strength (|σ|) curve generally rises as a result of consideration of the solvent granularity. (ii) An interionic attractive utail_αβ(r) helps in raising the Cd curve whereas a repulsive utail_αβ(r) tends to lower the Cd curve. Moreover, a secondary Cd peak appears when |σ| increases to an appropriately large value, and becomes increasingly evident with the strength of the attractive utail_αβ(r) and eventually disappears as the attractive utail_αβ(r)reduces in attraction or changes into repulsion at all. (iii) The Cd − |σ| curve is raised by increasing the repulsion between the solvent and counter-ion; however, both counter-ion diameter and bulk concentration cause obvious changes of the Cd − |σ∗| curve morphology. Concretely speaking, the Cd peak height and the peak position rise with the counter-ion size decreasing. Moreover, changing the utail_αβ(r) between solvent and counter-ion from attraction to repulsion facilitates a transition from bell-shaped curve to camel-shaped curve. (iv) Effects of all of the factors considered in influencing the Cd curve diminish increasingly with |σ|. All of the above observations can be explained by considering the interionic depletion potential induced by the solvent and its changes with the system parameters.
Article
This chapter describes theoretical approaches of the classical statistical mechanics and modern computer simulation methods to study the microscopic structure and thermodynamic and electrical properties of liquid electrolyte solutions at solid surfaces. Theoretical approaches include integral equations and density functional methods. We analyze their development to reach an accurate description of the interfacial structure and thermodynamics of nonuniform electrolyte solutions. The computer simulation techniques are reviewed. Methodological aspects concerned with the key issue of simulations, namely, accounting for the long-range electrostatic interactions, are given. Finally, we present some selected results for simple models and next for some real and experimentally important systems. Some prospects for future studies are discussed in the summary.
Article
Inhomogeneous correlation functions for model ‘ soft’ 2–2 and 1–1 electrolytes at a charged interface have been determined by the so–called ‘pair’ approximation of integral equation theory. The solvent is modeled as a structureless dielectric continuum at 25°C. The wall–ion–ion structure is calculated using the inhomogeneous Ornstein–Zernike relation, together with the hypernetted chain closure, and one of two choices for the functional relationship between the singlet and pair correlation functions. Both the interfacial density profiles and the inhomogeneous pair correlation functions are calculated. For most cases, the inhomogeneous pair correlation functions near the interface vary significantly from the homogeneous pair correlation functions. This deviation generally becomes stronger as the charge on the surface increases, and the deviation generally extends out further from the interface as the surface charge increases. The density profiles predicted by the pair approximations generally show less structure than the singlet approximation density profiles, whereas the inhomogeneous pair correlation functions generally predict more structure than would be expected by simply assuming bulk pair correlation functions. The density profiles and inhomogeneous correlation functions are also found to agree qualitatively with previous simulations which used ‘charged hard–sphere/charged hard—wall’ potentials.
Article
Full-text available
A study has been made of the electrical 'double layer' structure of molten salts, in particular a model of molten potassium chloride, using an integral equation approximation. This is in contrast to most statistical mechanical treatments of the double layer, which have concentrated on aqueous elecrolyte solutions. The results are compared with the output of computer simulations. In addition to the structural information contained in the density profiles, the calculations yielded charge profiles, the mean electrostatic potential and the double layer capacitance.
Article
Full-text available
The hypernetted chain/mean spherical approximation (HNC/MSA) integral equation for a totally asymmetric primitive model electrolyte around a spherical macroparticle is obtained and solved numerically in the case of size-asymmetric systems. The ensuing radial distribution functions show a very good agreement when compared to our Monte Carlo and molecular-dynamics simulations for spherical geometry and with respect to previous anisotropic reference HNC calculations in the planar limit. We report an analysis of the potential versus charge relationship, radial distribution functions, mean electrostatic potential, and cumulative reduced charge for representative examples of 1:1 and 2:2 salts with a size-asymmetry ratio of 2. Our results are collated with those of the modified Gouy-Chapman (MGC) and unequal radius modified Gouy-Chapman (URMGC) theories and with those of HNC/MSA in the restricted primitive model (RPM) to assess the importance of size-asymmetry effects. One of the most striking characteristics found is that, contrary to the general belief, away from the point of zero charge the properties of an asymmetric electrical double layer (EDL) are not those corresponding to a symmetric electrolyte with the size and charge of the counterion, i.e., counterions do not always dominate. This behavior suggests the existence of a new phenomenology in the EDL that genuinely belongs to a more realistic size-asymmetric model where steric correlations are taken into account consistently. Such novel features cannot be described by traditional mean-field theories such as MGC, URMGC, or even by enhanced formalisms, such as HNC/MSA, if they are based on the RPM.
Article
Full-text available
The spatial distribution of water molecules at solid-electrolyte interfaces has received extensive theoretical study, due to the importance of this interface in electrochemistry and other sciences. Such studies suggest that adjacent to the interface water is arranged in several layers, that the molecular arrangements in the inner layer is similar to bulk water, and that the inner-layer molecules have an oxygen-up (oxygen-down) average orientation for negative (positive) electrode charge (or, equivalently, potential). However, little of this has been verified by experimental measurements. In this paper we report surface X-ray scattering measurements of the water distribution perpendicular to a Ag(111)-electrolyte interface in 0.1M NaF at two potentials: +0.52 and -0.23 V from the potential of zero charge (PZC) on the electrode. We find that, first, the water is ordered in layers extending about three molecular diameters from the electrode. Second, the extent of ordering and the distance between the electrode and first water layer depend on potential, the latter being consistent with an oxygen-up (oxygen-down) average molecular orientation for negative (positive) electrode potential. Third, the inner water layer contains 1.55 × 1015 (at -0.23 V) and 2.6 × 1015 (at +0.52 V) water molecules per cm-2, remarkably more than expected from the bulk water density (i.e., ˜ 1.15 × 1015cm-2). Such a large compression shows that the molecular arrangements in the inner layer are significantly different from bulk, which has not been anticipated in current models of charged, aqueous interfaces. We give a qualitative explanation of this large density as resulting from the strong electric field at the charged Ag(111) electrode and present a tentative model of the molecular arrangements.
Book
The Advances in Chemical Physics series provides the chemical physics and physical chemistry fields with a forum for critical, authoritative evaluations of advances in every area of the discipline. Filled with cutting-edge research reported in a cohesive manner not found elsewhere in the literature, each volume of the Advances in Chemical Physics series serves as the perfect supplement to any advanced graduate class devoted to the study of chemical physics.
Chapter
Useful theories of inhomogeneous fluids can be obtained either from the singlet Ornstein-Zernike (OZ1) equation (Henderson et al, 1976).
Article
We derive a statistical mechanical expression for the surface tension of a liquid-vapor interface. It contains the density profile and the direct correlation function in the vicinity of the interface. When a local Ornstein-Zernike approximation is made on the direct correlation function, the surface-tension theory of Fisk and Widom is obtained.
Article
An exact formula for the contact value of the density of a system of charged hard spheres near a charged hard wall is obtained by means of a general statistical mechanical argument. In addition, a formula for the contact value of the charge profile in the limit of large field is obtained. Comparison with the corresponding expressions in the Poisson-Boltzmann theory of Gouy and Chapman shows that these latter expressions become exact for large fields, independent of the density of the hard spheres.
Article
The integral equation theory for a model 2–2 electrolyte is studied in detail. In this model electrolyte, the ions are assumed to be the same size, and interact via a continuous potential energy which behaves as the Coulomb potential at large distances and an inverse ninth power repulsion at short distances. The ions are embedded in a dielectric continuum of fixed dielectric constant, here taken to be 78.3 &egr;0 in order to model water at 25 °C. The bridge function for this model is studied as a function of concentration (a) for six proposed closures, and (b) via ‘‘exact’’ inversion of data from computer simulations. A proposed closure derived from examination of the inverted bridge function yields predictions in good agreement with computer simulations. We emphasize the importance of choosing an ‘‘optimized’’ long-range potential, as opposed to the traditional Coulomb choice. A simple functional form for the bridge function results from this optimized choice of long-range potential.
Article
The structure of the hard sphere dipolar liquid and the electrolyte with added hard sphere ions near a charged hard planar electrode has been investigated with the reference hypernetted chain integral equation (RHNC). We find a decrease of the dielectric function &egr; near the wall, a decrease of &egr; in the fluid due to saturation, a field dependent change in the dipole density near the wall, and a decrease of the ion density near the electrode due to solvation. Related to the demixing instability of the ion-dipole mixture, the ions suddenly concentrate near the surface at higher fields and lead to a sharp increase in the differential capacitance. Also electrostriction in the pure dipolar liquid with a field in the bulk is considered.
Article
The accuracy of the HNC approximation and modifications based on a simple approximation for the sum of bridge graphs has been studied for a charged soft sphere model for dilute aqueous 2-2 electrolytes. The accuracy is judged based on extensive Monte Carlo (MC) simulation for three representative concentrations: 0.005, 0.0625, and 0.2 M. The MC simulations find monotonically increasing like-charge correlation functions in agreement with the modified HNC results but in contrast with the HNC prediction of a local maximum at low concentration. Nevertheless, HNC provides the best overall quantitative match with MC except at very low concentration; at 5 mM the modifications are a significant improvement. For unlike charge correlations, there is good agreement among all methods; at the ionic contact peak, however, HNC underestimates the peak height and the modifications provide improved accuracy.
Article
In the primitive model of the electric double layer an electrolyte near a charged surface is modeled by an assembly of charged hard spheres in a medium of dielectric constant &egr;. We solve numerically the inhomogeneous Ornstein–Zernike equation for the pair correlation functions in the hypernetted chain and mean spherical approximations together with the Lovett–Mou–Buff–Wertheim equation for the density of the ions near a charged surface. For highly charged surfaces the profiles display layering and charge reversal. The density profiles and diffuse layer potential are generally in excellent agreement with Monte Carlo data. The pair correlation functions at small separation are significantly different from correlation functions in the bulk and at larger separation have the well-known power law decay (r−3) as function of separation when both hard spheres are kept at constant distance from the surface.
Article
The aim of this paper is twofold: (i) to clarify the definition of the mechanical (hydrostatic) stress tensor and the Maxwell (electrostatic) stress tensor in nonuniform electrolytes or Coulombic systems, and (ii) to derive contact conditions for an electrolyte at a charged hard wall to include (a) ion size effects, (b) imaging effects at the wall, and (c) effects due to the molecular nature of the solvent (the civilized model). The last result is valid for a general nonlinear dielectric whose molecules can have any electrostatic multipole interactions.