Xiaoyu Zhao†
,
Koichi Jeremiah Aoki,
Jingyuan Chen* and
Toyohiko Nishiumi
Department of Applied Physics, University of Fukui, Bunkyo 3-9-1, Fukui 910-0017, Japan. E-mail: jchen@u-fukui.ac.jp; Fax: +81 776 27 8753
First published on 14th November 2014
Electric double layer capacitance at platinum electrodes is controlled by dipole moments of the solvent in the diffuse layer rather than that by ionic distribution, being different from that at mercury electrodes. The controlling step is found by comparing capacitance vs. electrode potential curves in ionic solutions with those in deionized latex suspensions. The curves do not involve a valley shape of Gouy–Chapman (GC)-Stern's type until ionic concentrations are less than 0.05 mM, because measured capacitance is controlled by the inner layer. The valley shape at low concentrations can be measured in deionized sulfonic latex suspensions, whose conductance is brought about by the ionic latex particles rather than the dissociated hydrogen ions. An expression for the capacitance by the ionic latex suspension is derived, which is demonstrated to be the same form of the potential dependence as for mono-valence ions. Ac-impedance data are obtained at parallel polycrystalline platinum wires without an insulating shield. The valley shape is found, which is analyzed by the inverse plot of the capacitance against the hyperbolic cosine of the dimensionless applied potential. The linearity of the plots seems to support the GC-theory, but the capacitance values are much larger than those calculated from the GC-theory. The extra amount can be attributed quantitatively to the orientation of solvent molecules by combining Debye's theory with the GC-theory.
1/Cd = 1/CH + 1/CD | (1) |
CH = εrε0/x2 | (2) |
CD = (2εrε0F2c/RT)1/2![]() | (3) |
The potential variation of the GC theory allows us to separate CH or CD from the observed capacitance. A possible technique is to plot of 1/Cd against 1/cosh(Fϕ/2RT) to yield a line with an intercept, 1/CH, according to the combination of eqn (1) and (3). The GC-term at ϕ = 0 is (2ε0εrF2c/RT)1/2, of which numerical values are 7.2, 2.3, and 0.7 μF cm−2 for c = 1, 0.1 and 0.01 mM of aqueous solutions. In contrast, values of CH (eqn (2)) range from 1.0 to 1.7 μF cm−2 for x2 (0.3–0.5 nm) at εr = 6 for the saturated dielectric constant of water.31 Values of 1/CD are smaller than those of 1/CH at conventionally used concentrations of electrolyte so that CD cannot be extracted by the plot. In order to determine CD-component, it is necessary to use concentrations at least less than 0.01 mM. Such low concentration yields 1 MΩ resistance in the 1 × 1 × 1 cm3 cube of the solution, which blocks accurate measurements of impedance. Indeed, an attempt of the accurate determination has not yet been reported, to our knowledge.
A strategy of overcoming the difficulty of the determination is to use suspensions of ion-incorporated microspheres, called ionic latex particles. Counterions are partially dissociated from the salt incorporated on the latex particles and are distributed near the particles even after the suspensions are deionized sufficiently.32–35 For example, acidified polystyrene-sulfonate latex (PSS) suspensions dissociate hydrogen ion from the –SO3H moiety in salt solutions as a strong acid. When it is deionized, only 1% of hydrogen ions are dissociated, as has been observed by voltammetry.36 The conductance of latex suspensions is brought about mainly with the Brownian motion of the multi-anion latex particles rather than counter ions (H+),37 because the ionic conductivity is proportional to the square of the charge number of one particle.38 Consequently, the PSS suspensions including 0.01 mM hydrogen ion can keep the conductance equivalent to 1 mM solution of hydrochloric acid.39 The suspension is expected to reveal such small values of CD that 1/CD > 1/CH.
This report is directed to finding the ionic concentrations at which the component of CD can be extracted from the observed capacitance. The extraction technique is to plot of 1/Cd against 1/cosh(Fϕ/2RT), which is expected to show linearity with the intercept of 1/CH. Our concern is whether experimental values of the slope, (2ε0εrF2c/RT)−1/2, agree with the theoretical ones or not. The deionized PSS suspensions are used here in order to sustain accuracy of the impedance data against high solution resistance. Reference electrodes such as Ag|AgCl cannot be used here because they increase ionic concentrations by leakage of ions.40 Therefore we use a two-electrode system composed of two parallel platinum wires.36,39,41–43 Before applying eqn (3) to the data of the latex suspensions, we examine the validity of the proportionality of CD to cosh(Fϕ/2RT) by solving the Poisson–Boltzmann equation for the suspensions.
The synthesized particles had uniform diameter, 3.33 ± 0.05 μm, determined by an optical microscope, VMS-1900 (Scalar, Tokyo), in the wet state. The amount of hydrogen ion on the latex particle was determined by titration with NaOH under monitoring of the conductivity of the suspension.36 The turning point of the titration curve allowed us to determine the number of the sulfonate moieties per particle, n = 6.7 × 108.39 Concentration of the PSS suspension was determined by drying a give volume (3 cm3) of an aliquot of a stock PSS suspension in a vacuum oven, by weighing the dried aliquot, and by dividing the weight by the weight of one particle to yield the number of the particles, where the weight of one particle was evaluated from the product of the volume of one particle by the density (1.05 g cm−3) of PSS.
The potentiostat was Compactstat (Ivium, Netherlands), equipping a lock-in amplifier. Applied alternating voltage was 10 mV in amplitude. The potentiostat for voltammetry at extremely low scan rates was HECS 972 (Huso, Kawasaki), controlled with a homemade software. Solution was deaerated by nitrogen gas for 15 min before each voltammetric run. Delay of the potentiostat was examined by the same process as in the previous report.41 Phase sensitivity was estimated by use of a series combination of a film capacitor 0.2 μF and a carbon resistance 1 MΩ. When the ratio of out-of-phase component to in-phase one was less than 0.001, the out-of-phase was observed to be underestimated.
The ac impedance of the deionized suspension between the two wire electrodes was obtained at zero dc voltage in the two-electrode system. Fig. 1 shows Nyquist plots obtained in hydrochloric acid. All the plots fell on each line, suggesting only the participation in the double layer impedance rather than faradaic impedance. The lines support the validity of the constant phase element.44–47 The extrapolation of the line to Z2 = 0 for infinite frequency yields the solution resistance, Rs.48 The method of determining Rs by the two wire electrodes has been demonstrated to be valid.41 The values of the resistances of HCl were inversely proportional of concentrations of HCl. Therefore the evaluation of Rs should be correct. Fig. 1 also shows the Nyquist plot for the PSS suspension in which the concentration of free hydrogen ion was 0.01 mM. The value of Rs in the suspension was close to that of 1 mM HCl. Consequently the conductance of the suspension is not brought about by the free hydrogen ion but by the condensed charge of the latex particles.39
The ac-current, I, in the electric double layer capacitance, Cd, responding to the ac voltage, V = Voexp(iωt), is given by the time-derivative of the double layer charge, q, i.e., I = dq/dt, where i is the imaginary unit, ω is the angular velocity and Vo is the ac-amplitude. When the capacitance is frequency-dependent, the current is expressed by
![]() | (4) |
The first term on the right hand side is iωCdV, belonging to the out-of-phase. In contrast the second term includes no imaginary number, and hence belongs to the in-phase, i.e. a resistive component. Since the current in eqn (4) is a sum of the real and the imaginary currents, the equivalent circuit can be represented as a parallel combination of the out-of-phase (defined as Cp) and the resistance (1/(∂Cd/∂t) = Rp). Then the double layer capacitance, Cd, conventionally used is given by
iωCd = 1/Rp + iωCp | (5) |
The present measurement is made at two identical wire electrodes in the two-electrode system. The total impedance is a series combination of the solution resistance, Rs, and two parallel combinations of Cp and Rp, each representing Cd at one Pt|solution interface, as is shown in Fig. 2. This equivalent circuit allows us to evaluate the frequency-dependent capacitance. The frequency-dependence of Rp ought to vary Z1 with frequency. Therefore the plots in Fig. 1 are tilted from a vertical line.
We obtain the expressions for the in phase and the out of phase components from the equivalent circuit in Fig. 2, as follows
![]() | (6) |
The explicit forms of Cp and Rp are given by
![]() | (7) |
Values of Cp and Rp were determined from Z1 and Z2 at each frequency by use of eqn (7).
Fig. 3 shows logarithmic variations of Cp with logarithm of frequency (f = 2πω) in the PSS suspension and HCl solutions.
![]() | ||
Fig. 3 Variation of Cp in (a) 1 mM HCl, (c) 0.05 mM HCl and (d) PSS suspension (including 0.01 mM H+) with logarithmic frequency. |
Values of Cp show linear relations except for the plot at the low concentration (0.05 mM) of HCl. Low ionic concentrations increase the solution resistance, which degrades accuracy of the impedance data. The linearity suggests
Cp = (Cp)1Hzfk | (8) |
The frequency-dispersion is thought to be caused by surface roughness.50–53 Since crystalline surface always contains a number of defects in order to stabilize entropically the surface energy, the frequency-dispersion is necessarily measured in the capacitance.
We evaluated Rp from values of Z1 and Z2 through eqn (7). Logarithmic values of Rp for 1 mM HCl and PSS are plotted against logarithmic frequency in Fig. 4. They fall on each line, of which slopes are −0.69 and −0.76, respectively. Therefore Rp can be expressed by k1/f0.69 and k2/f0.76, respectively. Since 1/Rp = ∂Cp/∂t = (∂Cp/∂f)(∂f/∂t) according to eqn (4), the slopes of logCp vs. log
f in Fig. 3 can provide ∂Cp/∂f, which may yield the dependence of Rp on f with the help of f = ω/2π = 1/2πt. The values of the slopes in Fig. 3 are (a) −0.29 and (d) −0.21, which yield the slopes ∂Rp/∂f = 0.29 − 1 = −0.71 and 0.21 − 1 = −0.79 for Fig. 4(a) and (d), respectively. They are close to those obtained by the slope in Fig. 3. Therefore, the relation 1/Rp = ∂Cp/∂t is valid numerically. Rp at extremely low frequency, i.e. at DC voltage tends to infinity. Then the equivalent in Fig. 2 is represented by Cp–Rs–Cp in series. Consequently no dc current flows for dc potential in a polarized potential domain.
![]() | ||
Fig. 4 Logarithmic variations of Rp in (a) 1 mM HCl and (d) PSS suspension with logarithmic frequency. |
The frequency-dispersion is not included in the GC theory, but occurs at the electrode surface itself as a subject of CH. Since it is far from a subject of ionic distribution, we regard it as a priori fact here.
![]() | ||
Fig. 5 Variations of (Cp)1.5Hz with the dc-potential obtained in (a) 1 mM HCl, (b) 0.2 mM HCl, (c) 0.05 HCl and (d) PSS suspension. |
The simulation study has pointed out that the ionic contribution of the capacitance should be smaller than that predicted by the GC theory because of hard core exclusion of solvents, ionic correlations and electrostatic repulsion.54 Application of this concept could allow us to extract the ionic contribution from the observed capacitance more easily than the prediction. Our experimental results, however, suggest the difficulty of the extraction.
The second variable satisfying CH > CD is the ac frequency. Fig. 6 shows variations of Cp with E at some frequencies in (A) 0.05 mM HCl and (B) the PSS suspension. The Cp values of the HCl solution for f > 30 Hz did not vary with the potential, whereas those for f < 10 Hz showed a valley shape, as is shown in Fig. 6(A). According to eqn (8), Cp is evaluated to be smaller at higher frequency, and so CH becomes smaller. Then the measured values of Cd is mainly determined by CH rather than CD through eqn (1). Therefore the potential dependence is extinguished at high frequency.
![]() | ||
Fig. 6 Dependence of Cp on the dc-potential in (A) HCl 0.05 mM and (B) PSS suspension at f = (a) 1.6, (b) 3.2, (c) 10, (d) 100 and (e) 3120 Hz. |
In contrast, the PSS suspension exhibited valley shapes even for high frequency (f < 10 kHz). Therefore the potential dependence can be obtained accurately by use of the deionized PSS suspension. The potential of the bottom of the valley may correspond to point of zero charge (PZC). The PZC here is at 0 V because of the use of the symmetric two-electrode system. No valley was obtained in the three-electrode system because of leakage of ions from a reference electrode.40 In other words, it is difficult to determine values of the PZC on a reference electrode potential scale. Reported values of the PZC may be at a bottom of valley caused by faradaic processes.47
The valleys in Fig. 6 might be caused by erroneous data analysis such as in subtraction of Rs near E = 0 V. A technique of detecting the valley without any data analysis is cyclic voltammetry. Cyclic voltammetry was made in the PSS suspension with the extremely low ionic concentration at very low scan rates in the two-electrode system in order to exhibit a valley shape. Usage of a reference electrode increased ionic strength not to exhibit any valley. Fig. 7 shows the slow scan voltammogram in the PSS suspension. A valley shape appeared at E = 0 V for iterative scans in both the anodic and the cathodic scan only at the scan rates as slowly as 1 mV s−1. Therefore it should be caused by the capacitance. The appearance of the valley in voltammograms may not have been reported yet, to our knowledge. No valley was observed in HCl solution.
![]() | ||
Fig. 7 Voltammetry in the PSS suspension at the scan rate 1 mV s−1. The scan started at E = 0 V in the positive direction. |
We attempt to separate Cp into CH- and CD-components by use of the plots of 1/Cp against 1/cosh(Fϕ/2RT) with an expectation of exhibiting linearity. We will demonstrate in Section 3.4 that CD for the PSS suspension can be expressed by the same potential dependence as in eqn (3). It is necessary to make relation between ϕ in the GC-theory with experimentally controllable voltages, E. The predicted distribution of the potential in solution, ϕS, is illustrated in Fig. 8, where superscripts, A and C, mean the anode and the cathode, respectively. ϕE is the potential of the electrode, ϕS0 is the potential at the electrode on the solution side, and ϕSx is the potential at x2. The voltage, ϕAE − ϕAS0, is the same value as ϕCE − ϕCS0, because the two electrodes are the common material (Pt). Therefore, the applied voltage is ϕAE − ϕCE, which is equal to ϕAS0 − ϕCS0. It can be written in terms of the sum of the three voltages.
E = ϕAS0 − ϕCS0 = (ϕAS0 − ϕASx) + (ϕASx − ϕCSx) + (ϕCSx − ϕCS0) | (9) |
The middle term represents the sum of the IR-drop and the voltage by the ion distributions due to the GC-theory between the two electrodes. Since our data of Cp do not contain any IR-drop, the middle term equals 2ϕ in the GC-theory. Then eqn (9) is reduced to
E = (ϕAS0 − ϕASx) + 2ϕ + (ϕCSx − ϕCS0) | (10) |
All the observed capacitance vs. potential curves were approximately symmetric with respect to E = 0 V. The symmetry implies that the ϕ = [E − (ϕAS0 − ϕASx) − (ϕCS0 − ϕCSx)]/2 at the anode should be the same as that at the cathode. We now define the dimensionless unknown variable as
r = −(ϕAS0 − ϕASx + ϕCSx − ϕCS0)/2ϕ | (11) |
The meaning of r is the ratio of the voltage in the Helmholtz layer to the externally controlled voltage. Then eqn (10) is reduced to E = 2ϕ(1 − r), and eqn (3) can be rewritten as
![]() | (12) |
Fig. 9 shows the plot of 1/Cp against 1/cosh(EF/4(1 − r)RT) in the PSS suspension for f = 10 Hz at some values of r. With an increase in r, the fitting curves by a quadratic equation vary from a concave (a) to a convex (c). A line is realized for r = 0.4 ± 0.2, indicating that a quarter of the applied voltage contributes to determination of x2-potential. The linear variation was found also for 0.05 mM HCl solutions although it contains large errors.
![]() | ||
Fig. 9 Variations of Cp−1 with cosh(EF/4(1 − r)RT) of the PSS suspension at f = 10 Hz for r = (a) 0.0, (b) 0.3 and (c) 0.6. |
The linearity in Fig. 9 seems to support the GC theory. However, the slope increased with an increase in the frequency for f < 10 Hz, as shown in Fig. 10 for (A) HCl solution and (B) PSS suspension. This is inconsistent with eqn (12) for the slope. The average values of the slopes of curves (a)–(c) in Fig. 10(A) and (B) are 0.63 F−1 m2, and 0.9 F−1 m2, respectively. On the other hands, the calculated values of (2ε0εrF2c/RT)−1/2 are 61 F−1 m2 and 140 F−1 m2, respectively, at c = 0.05 mM and 0.01 mM for εr = 78. The experimental values are two orders of the magnitude smaller than the calculated ones. The disagreement indicates that capacitances other than CD should be involved in parallel to CD. The GC-capacitance results from an excess amount of accumulated ions by the electric field, but does not include any contribution from solvent molecules. In Section 3.5, we will take into account a contribution of solvent molecules by the electric field, like the inner layer capacitance.
The capacitance (eqn (3)) by the Gouy–Chapman theory has some limitations in finite size of ions and saturation of dielectric constants. We first discuss the effect of finite size by use of the lattice-gas model by Kornyshev.15 Since the lattice-gas model includes a distribution of ions on the lattice, there is no possibility of overlapping ions. The capacitances with the overlap, CD, has been given for CD0 = (2ε0εrF2c/RT)1/2 by15
![]() | (13) |
![]() | ||
Fig. 11 Variations of eqn (13) for γ = (a) 0.0, (b) 0.01, (c) 0.1 and (d) 0.2. Circles are the data in Fig. 6(a). |
Saturation of dielectric constants decreases CD through εr in eqn (3). The value of CD0 for εr = 78 of bulk water and is εr = 6 (ref. 31) are, respectively, 7.2 and 2.0 μF cm−2at c = 1 mM, calculated from eqn (3) at ϕ = 0. The dielectric saturation has no effect on the shape of capacitance vs. potential curves.
(–SO3H)n ↔ (–SO3H)n−z + (–SO3−)z + zH+ | (14) |
The deionized suspension including N latex particles in a unit volume contains Nz dissociated hydrogen ions. A planar electrode is inserted in the suspension to apply voltage. Let the inner potential in solution at a location x from the electrode be ϕ(x). The number density of hydrogen ion in equilibrium with the potential in the thermal bath is given by the Boltzmann distribution, n+ = Nzexp(−eϕ/kBT), where e is the elementary charge, and kB is the Boltzmann constant. If the latex particle were to have the charge −ze in the form of a point charge, the Boltzmann distribution would be expressed by n− = N
exp(zeϕ/kBT). The charge −ze is, however, dispersed within the latex sphere 3.3 μm in diameter or 19 μm3 in volume so that the average concentration is z/19 μm3 or 0.6 mM. This low concentration makes zSO3− ions fluctuated independently in a thermal bath. As a result, the Boltzmann factor is exp(eϕ/kBT) rather than exp(zeϕ/kBT). The role of the latex is to encompass z anions in the sphere without taking a point charge. Then the number density of the anions is given by n− = N
exp(eϕ/kBT). The charge density at x is expressed by
ρ = Nze[exp(−eϕ/kBT) − (1/z)exp(eϕ/kBT)] | (15) |
The second term in the bracket can be rewritten as (1/z)exp(eϕ/kBT) = exp[(eϕ − kBT lnz)/kBT]. This indicates that the electrostatic energy, eϕ, is stabilized by the amount of the entropic term, kBT
lnz, owing to the confinement of z sulfonic ions within the particle.
When we introduce the following potential shifted by the entropic term
ϕz = ϕ − (kBT/2e)ln![]() | (16) |
Eqn (15) is reduced to
![]() | (17) |
Then the Poisson equation is given by
Multiplying the terms on the both hand sides by dϕz/dx, and integrating the resulting equation on the boundary conditions of ϕz = 0, dϕz/dx = 0 at x → ∞, we have
![]() | (18) |
The left hand side in eqn (18) is the same as (dϕ/dx)2. The surface charge density at the electrode is given by
![]() | (19) |
The capacitance by the ion distribution is expressed by
![]() | (20) |
The potential dependence in eqn (20) is the same as that (eqn (3)) by Gouy–Chapman equation. When we express the concentration of the dissociated hydrogen ions as molarity, cH,ltx, = Nz/NA, the pre-cosh term in eqn (20) is F(2cH,ltxε0εr/zRT)1/2, where NA is the Avogadro constant. Only the difference between eqn (3) and (20) lies in the expression for the concentration. Consequently the capacitance in the suspension at ϕ0 = 0 is smaller than that in acid by z−1/2.
It is of interest to consider the difference between the length of the ionic atmosphere, λ, of ordinary multi-valence electrolytes and that of the PSS latex. When the potential decay is expressed by the form of exp(−κx) for a constant κ, the length can be defined as κ−1 = λ. We will obtain the exponential form of the solution of the Poisson–Boltzmann equation. The ordinary electrolyte with the z-charged cation and z-charged anion obeys the following Poisson-Boltzmann equation
d2ϕ/dx2 = (2zeN/ε0εr)sinh(zeϕ/kBT) | (21) |
When a value of zeϕ/kBT is close to zero, the above equation is approximated as
d2ϕ/dx2 = λz–z−2ϕ | (22) |
![]() | (23) |
Since a solution of eqn (22) is exp(−x/λz–z), λz–z is nothing but the length of the ionic atmosphere. The length is inversely proportional to z. On the other hands, the length of the latex suspension can be obtained from the Taylor expansion of the term on the right hand side in eqn (17), which yields −(2cz1/2F2/RT)ϕz. Then the length for the latex suspension is expressed by
![]() | (24) |
The numerical value of the length by eqn (24) for cH,ltx = 0.01 mM and z = 6.7 × 106 is 1.9 nm, which is much smaller than the diameter of the PSS particle (3.3 μm). Therefore the ionic atmosphere is formed within the particle. Since –SO3− can move in the particle through the reaction –SO3H ↔ –SO3− + H+ like a free anion, it is reasonable to generate ionic concentration gradient within the particle as a particle approaches the electrode. Fig. 12 shows an illustration of (A) the distribution of H+ and Cl− and (B) that of H+ and –SO3− in the PSS latex for z = 3 when the number of ions of HCl is the same as that of PSS. The anion –SO3− is driven by the electric field in confinement of the sphere of the particle, as shown by the translation of –SO3− in the direction of arrows. As a result, an ionic distribution is formed within the particle, and then the ionic length varies from λl–l to λltx. We compare the length in the PSS suspension with that in HCl at a common concentration of H+. The ratio, λl–l/λltx, is z 1/4 (=51) for z = 6.7 × 106, indicating that the PSS suspension has stronger ionic intensity than HCl by 51 times. If all the negative charges on the one PSS particle were to be condensed like a point z-electron charge, the ratio λl–l/λz–z = z, being of the order of 106 might correspond to the ionic length of the order of femto m.
![]() | (25) |
We evaluate more accurately CD-slv. Our model is a laminar of solvent monolayer sheets through which the electric field is applied with the ionic distribution of e−x/λ. The solvent molecule has the dipole moment μ and the length l along the direction of the dipole. Since dipoles are oriented by the electric field which varies with the distance from the electrode, those closer to the electrode are more oriented than those far from the electrode. As a result, the closer to the electrode is a layer, the larger is its capacitance. Let the molecules in the n-th layer from x2 have the orientation angle, θn, from the direction normal to the electrode surface. According to the concept of the inner layer capacitance (eqn (2)), the capacitance of one molecule in the n-th layer is given by43
Cn = ε0/l![]() ![]() | (26) |
![]() | (27) |
A well-known relation between the angle with the electric field is the Debye equation,55 given by
cos![]() | (28) |
Here, the function L is defined by
L(x) = coth(x) − 1/x |
Inserting eqn (26) and (28) into eqn (27) yields
![]() | (29) |
We use Gouy–Chapman's equation for dϕ/dx at small values of eϕ/kBT in the x-coordinate when the anode and the cathode are set at x = w and −w, respectively. We assume that the potential distribution is symmetry with respect to x = 0. Therefore the boundary conditions in Fig. 8 can be reduced to ϕ(w − x2) = ϕ2 and ϕ(0) = 0. The solution of the Poisson–Boltzmann equation for small values of Fϕ/RT is given by
ε0εr(d2ϕ/dx2) = 2F2cϕ/RT | (30) |
The boundary value problem for large x has the following approximate equation:
ϕ = ϕ2e(x−w+x2)/λ | (31) |
Inserting eqn (31) into eqn (29) and replacing the summation by an integral yields
![]() | (32) |
![]() | (33) |
CD-slv = 3ε0kBT/μϕ2 = 6ε0kBT/μ(ϕASx − ϕCSx) | (34) |
This capacitance does not vary directly with the ionic concentration but vary through ϕ2 which depends on the ionic concentration.
A typical value of CD-slv in water (μ = 1.84 Debye) is 18 μF cm−2 at ϕ2 = 0.1 V. It is larger than the value for the GC-theory by 30 times. Since the density of ions is much smaller than that of solvent, the capacitance caused by ions ought to be much smaller than that by the orientation of solvent dipoles. Therefore it is necessary to take into account the contribution of the orientation of the dipole in the capacitance of the diffuse layer, in addition to the Gouy–Chapman's term,
![]() | (35) |
Since the orientation is brought about by the electric field which is relaxed with the ionic distribution, it is associated with cosh(EF/4(1 − r)RT). Then, the capacitance of the diffuse layer is represented as
CD = (CD-ion + CD-slv)cosh(EF/4(1 − r)RT) | (36) |
The measured capacitance is mainly controlled by CD-slv rather than CD-ion. This prediction is seen in the dependence of the slopes of 1/Cp vs. 1/cosh[EF(1 − r)/RT] on frequency through Fig. 10, like the frequency-dependence of the inner layer capacitances in Fig. 3.
Eqn (36) implies a parallel combination of CD-ion and CD-slv, which is in series with CH, as shown in Fig. 13(A). Since CD-slv depends on the frequency as the inner capacitance does, the measurement of CD-slv is necessarily associated with the apparent capacitance, RpD-slv, in the form of the inverse proportionality of the frequency, as for the apparent resistance, RpH, in the inner layer. Consequently, the equivalent circuit including the frequency-dependence takes a complicated form as shown in Fig. 13(B). Values of RpH and RpD-slv at low frequency are large enough to be negligible. Then the circuit in Fig. 13(B) tends to that in Fig. 13(A).
It would be interesting to extract CD,slv or CD,ion from Cp. A controllable variable of the extraction is the ionic concentration included in eqn (35). Usage of lower concentrations of PSS suspensions might allow us to evaluate CD,ion, which could be realized by the latex particles with larger values of the sulfonic moiety (n > 6.7 × 108) per particle. Unfortunately, our synthetic technique is not enough at present. Therefore, it is quite difficult to extract the component of GC capacitance from experimentally obtained capacitance.
It is not easy to extract the potential-dependent from the observed double layer capacitance, because values of 1/CD are smaller than those of 1/CH. The condition of 1/CD > 1/CH can be attained by decreasing ionic concentrations less than 0.05 mM. This concentration domain is incompatible with accurate impedance measurements in the context of high solution resistance. The difficulty can be solved by use of deionized latex suspensions, which can support conductivity with low free ionic concentration.
The valley shape appears in cyclic voltammograms in the PSS suspension at very slow scan rates. This is an evidence of the valley shape in capacitive currents. The bottoms of valley in both the capacitance-potential curves and the voltammograms correspond to a PZC. Any reference electrode cannot be used for detection of a valley because a reference electrode increases ionic concentration. Therefore it is difficult to determine PZC value vs. NHE.
Footnote |
† Tianjin Key Laboratory of Marine Resources and Chemistry, Tianjin University of Science and Technology, China. |
This journal is © The Royal Society of Chemistry 2014 |