A theoretical exploration of the influencing factors for surface potential*
Liu Xin-Mina),b), Yang Ganga)†, Li Hanga)‡, Tian Ruia), Li Ruia), Ding Wu-Quana), Yuan Ruob)
Chongqing Key Laboratory of Soil Multi-scale Interfacial Process, College of Resources and Environment, Southwest University, Chongqing 400715, China
School of Chemistry and Chemical Engineering, Southwest University, Chongqing 400715, China

Corresponding author. E-mail: theobiochem@gmail.com

Corresponding author. E-mail: lihangswu@163.com

*Project supported by the National Natural Science Foundation of China (Grant Nos. 41371249, 41201223, and 41101223) and the Fundamental Research Funds for the Central Universities, China (Grant No. XDJK2015C059).

Abstract

Although it has been widely used to probe the interfacial property, dynamics, and reactivity, the surface potential remains intractable for directly being measured, especially for charged particles in aqueous solutions. This paper presents that the surface potential is strongly dependent on the Hofmeister effect, and the theory including ion polarization and ionic correlation shows significant improvement compared with the classical theory. Ion polarization causes a strong Hofmeister effect and further dramatic decrease to surface potential, especially at low concentration; in contrast, ionic correlation that is closely associated with potential decay distance overestimates surface potential and plays an increasing role at higher ionic concentrations. Contributions of ion polarization and ionic correlation are respectively assessed, and a critical point is detected where their contributions can be exactly counteracted. Ionic correlation can be almost neglected at low ionic concentrations, while ion polarization, albeit less important at high concentrations, should be considered across the entire concentration range. The results thus obtained are applicable to other interfacial processes.

Keyword: 82.65.+r; 82.33.Pt; 82.60.Qr; 05.70.Np; Hofmeister effect; ionic activity; interfacial process; potential distribution
1. Introduction

Surface potential is at the Stern plane of charged particles (Fig. 1) and has been used to probe the interfacial properties, dynamics, and reactivity. A number of methods based on zeta (ξ ) potential, [1, 2] negative adsorption, [3] positive adsorption, [4, 5] and pH indicator[6] have been used to obtain surface potential. It has important implications for a variety of nano/colloidal systems such as clay minerals, proteins, and membranes. Notwithstanding, for these charged particles in aqueous solutions, the direct measurement of surface potential remains a difficult task, and hence zeta (ξ ) potential that obviously has a small value is often used to represent surface potential[79] (see Fig. 1). On the other hand, not a few theories have been developed to predict surface potential, such as Poisson– Boltzmann theory, [4, 5, 10] self-consistent field theory, [11] classical density functional theory, [1214] and integral-equation theory.[15]

Generally, the theories currently used to predict surface potential overlook Hofmeister (or specific ion) effects. Specific ion effects, which were first observed about 120 years ago, have been found to be ubiquitous and play a central role in physical, chemical, biological, and colloidal sciences.[1618] A variety of experimental phenomena can be explained by use of ion specificities, such as DNA conformation and stability, [16, 1921] activity coefficient, [22] interfacial tension of electrolyte solutions, [23, 24] colloidal forces, [25, 26] and buffer.[27, 28] Through the effects of interactions between ions and charged surfaces, ions and ions as well as ions and solvent molecules, the specific ion effect exhibits pronounced influences on the distributions of ions at the interface and in bulk solutions, [29] which further alters the results of surface potential. Therefore, it is no wonder that classical theories that neglect Hofmeister effects are unlikely to make a sound prediction.[2830] The surface potential expression with the inclusion of specific ion effects due to the enhanced ion polarization in the diffuse layer has recently been derived in our laboratory; [31, 32] however, the physical meanings of the ion polarization and ionic correlation in this expression remain ambiguous, and such investigations are of necessity in order to promote our understanding of surface potential; meanwhile, insightful clues can be provided towards its direct measurement.

In the present work, the respective contributions of ion polarization and ionic correlation to surface potential are also evaluated for the various colloidal systems and a wide range of ionic concentrations. Strong ionic correlation between condensed counterions can lead to attractive interaction among colloidal particles[33] as well as charge reversal.[34] On the other hand, ionic correlation at high salt concentrations can reduce the counterion concentration in the diffuse layer.[35] This adds the difficulty in measuring the surface potential because ionic correlation usually works in a collaborative manner with ion polarization. We are not aware which is more important for the surface potential determination, ion polarization or ionic correlation. How these two factors behave at different ionic concentrations also remains unclear to us. In the meantime, surface potential decays gradually in the electric double layer (Fig. 1), and the relationship between the potential distribution and decay distance is explored, which helps to understand the difference from zeta (ξ ) potential. To address the above issues, the rest of this paper is organized as follows. In Section 2, we show the theoretical surface potential expressions with only Coulomb force as well as correction by ionic correlation, ion polarization, and both factors. In Section 3, the presence of strong ion polarization is evidenced, and then the respective contributions of ionic polarization and ionic correlations are evaluated. In Section 4 the main conclusions are drawn from the present study.

Fig. 1. Counterion distribution, relative positions of surface, and zeta (ξ ) potentials as well as schematic representation of these potentials, each as a function of the potential decay distance that starts from the charged surface and extends towards the bulk solutions (thick solid curve).

2. Surface potential theories

In this section, the various surface potential theories that are developed on the basis of the Poisson– Boltzmann (PB) equation are elaborated. For a given system, ion polarization and ionic correlation usually function collaboratively, and are intractable to distinguish for the experimental researchers.

2.1. Ion– surface Coulomb interaction

For a binary electrolyte system, ions in the diffuse layer should abide by the Boltzmann distribution,

where (in units: mol· L− 1) is the ionic concentration in bulk solution, φ (x) (in unit V) is the potential in the electric double layer, and Zi is the ionic valence, F (in units: C· mol− 1) is the Faraday constant, R (in unit: J· mol− 1· K− 1) is the gas constant, and T (in unit: K) is the absolute temperature.

The mean ion concentration in the diffuse layer equals

where Ni (mol· g− 1) is the amount of the i ion that interacts with the charged particle surface, S (dm2· g− 1) is the surface area of the charged particle, and κ (dm− 1) is the Debye– Hü ckel parameter.

In our previous work, [36] the potential distribution φ (x) of the binary electrolyte system containing monovalent and divalent counterions (e.g., NaCl + CaCl2) was derived on the basis of the nonlinear Poisson– Boltzmann (PB) equation,

where

In Eq. (4), φ 0 represents the potential at the original plane of the diffuse layer (or Stern plane), and here is regarded as surface potential (see Fig. 1).

Introducing the potential distribution function into Eq. (1) and then combining Eq. (2), we arrived at

with

In Eq. (5), ε is the dielectric constant of water (8.9× 10− 10 C2· J− 1· dm− 1), F (C· mol− 1) is the Faraday constant, m, also referred to as in some papers, [37] is written as , where and refer to the concentrations of univalent and bivalent ions, respectively.[31]

Then the following expressions can be derived for the binary system (i and j)

As a result, the surface potential of the classical theory can be given as,

Equation (8) is the derivative of the original PB equation where only Coulomb interaction occurs between ions and surface.

2.2. Ionic correlation

It is apparent that ionic correlation affects the ion-exchange process and hence this factor should be taken into account:[38, 39]

where μ i0 is the standard chemical potential, while , , and are respectively the ionic concentration, ionic activity, and ionic activity coefficient in the bulk solution. That is, the RT ln term results from ionic correlation in a non-ideal system.[38] The ionic activity is the effective ionic concentration in solution and there are several empirical formulae of calculating the ionic activity and ionic activity coefficient.[40]

The chemical potential in the diffuse layer can be written as

where , ã i, and are respectively the concentration, activity, and activity coefficient in the diffuse layer.

According to the previous studies, [31, 37, 41] ionic activity represents a function of the Coulomb energy , and the activity coefficient in the diffuse layer equals[37]

When the ion exchange process reaches equilibrium, we have , and so the combination of Eqs. (9)– (11) leads to

In consideration of ionic correlation, the PB expression is modified into

Expressions similar to Eq. (7) can be derived as

where

As a result, surface potential with the inclusion of ionic correlation is

It can be seen from the above descriptions that the ionic activity in solution has been taken into account by considering ion size and ionic dispersion potentials of ion– ion and ion– interface interactions.[28]

2.3. Ion polarization

As is well known, strong electric fields (108 V· m− 1 ∼ 109 V· m− 1) often exist at the surfaces of colloidal particles.[31, 32, 4250] As reported previously, the strongly polarizable ions can be adsorbed on the air– water and oil– water interfaces thus affecting the interfacial tensions.[23] For the negatively charged colloidal particles, the ions at their interfaces should be polarized more significantly, [51, 52] and the addition of polarization energy is absolutely needed. Therefore equation (14) is converted into the following equations:[29]

The above equation can be further transformed into[53]

with

where and refer to the mean dipole moment in the diffuse layer and electric field strength, respectively; φ 1/κ is the potential at x = 1/κ (the upper limit of the electric double layer) and converged to the null point, as verified by the potential distributions in Section 3.

The exponential term in Eq. (18) equals

Then the coefficients β i and β j are

The coefficients β i and β j represent functions of the dipole moment difference of the i and j ions. Based on Eqs. (18) and (20), further correction to the surface potential by considering ion polarization turns Eq. (16) into

At low ionic concentrations, the effect of ionic activity can be neglected, that is, should be approximately equal to given below

The specific ion effects due to ion polarization are embodied with coefficients β i and β j, which are sensitive to the choice of ion pairs because ion polarizabilities in the diffuse layer are probably different.[29] Note that refers to the surface potential corrected with ion polarization but not with ionic correlation. The PB equation cannot be resolved analytically for the mixed electrolyte solutions (e.g., NaCl + CaCl2) in the presence of the polarization effect.[54] As an alternative, we consider the polarization effect on the basis of the mean energy that is obtained directly from the PB equation.

3. Results and discussion
3.1. Ion polarization in experimental observations

The coefficients β Na and β Ca can be derived from Na+ – Ca2+ exchange equilibrium experiment, [43] and their empirical expressions are and , where I is the ionic strength.[31] The difference between β Na and β Ca reflects the polarization difference between Ca2+ and Na+ . From the empirical expressions and Eq. (21), the experimental polarization difference between Ca2+ and Na+ in the electric double layer can be

On the other hand, according to the classical theory, the polarization effect for a given ion and the polarization difference between Ca2+ and Na+ are estimated by

where

ε 0 is the dielectric constant in a vacuum and ε i = 1 + 4π α i/Vi, [55] is the effective ion polarizability in the solution, and ε i and ε w are the relative dielectric functions of the i ion and water, respectively, Vi is the ionic volume, and α i is the static polarizability of the i ion, with α Ca = 0.4692 Å 3 and α Na = 0.139 Å 3, respectively.[55, 56] According to the Ca2+ /Na+ exchange equilibrium results with different ionic strengths, [43]φ 0 can be calculated by use of Eq. (22). The potential at the distance 1/κ (φ 1/κ ) is determined through the potential distribution of Eq. (3).[36]

Figure 2 shows the polarization differences between Ca2+ and Na+ ions, obtained by the experimental electric double layer and classical theory. Owing to the neglect of ion polarization, the classical theory assumes that the dipole moments of Ca2+ and Na+ ions are close to each other and in addition, their difference does not show an observable change with the increase of ionic strength. These are apparently in discrepancy with the experimental observations. On the contrary, the electron configuration analysis is perfectly in line with the experimental observations in that there exists a remarkable difference between Ca2+ and Na+ dipole moments and the polarization effect of the former should be considerably larger (Fig. 2). As mentioned above, the strong electric fields (108 V· m− 1 ∼ 109 V· m− 1) often exist at the interfaces of colloidal particles and in many other systems. Under such strong electric fields, the dipole moments of interacting counterions should be significantly affected and exhibit a considerable increase, and the increment differences between two different counterions can be substantial: e.g., Ca2+ and Na+ investigated in this work. As compared with Na+ , the electron clouds of Ca2+ are obviously softer and accordingly Ca2+ corresponds to a much stronger polarization effect when interacted with the negatively charged surface, which should be responsible for arousing the Hofmeister effects among the various ions. In addition, the polarization difference between Ca2+ and Na+ shows an increase with the decrease of ionic strength, and this trend becomes more conspicuous at relatively small ionic strengths (I ≤ 0.04 mol· L− 1, see Fig. 2). Neither these experimental results nor electron-configuration results can be properly explained by the classical theory. The classical theory seriously underestimates the effective polarization of counterions. As a matter of fact, the various counterions near the charged surfaces should be polarized to a different degree, depending on their electron configurations.

Fig. 2. Polarization differences between the Ca2+ and Na+ ions (, in Debye), each as a function of the root square of ionic strength , in mol1/2· L− 1/2) in solution.

It is worth noting that the water should be also polarized by a strong electric field, [57] but which is obviously less than the charge dipole such as Ca2+ and Na+ ions; [5860] that is, the water solvent has a limited influence on the interactions between ions and surface charges that are mainly responsible for specific ion effects. Moreover, the influence of the ionic solution has already been included in the surface potential expression by ionic correlation (for the details, see Subsection 2.2.).

3.2. Influence of ion polarization on surface potential

Si4+ and Al3+ are the principal constituents for most clay minerals, and isomorphous substitutions with metal ions of almost the same radii create permanent negative charges, e.g., usually the substitutions of Mg2+ /Al3+ for montmorillonite. It seems very difficult for lattice cations to be replaced by ions from the surfaces and bulk solutions, especially Ca2+ and Na+ that have substantially larger radii than Si4+ and Al3+ . Table 1 lists the surface potentials of montmorillonite (mean size: 1480 nm) and rutile titania (TiO2, mean size: 25 nm), [32] that have been calculated by the various schemes in Section 2, i.e., the classical theory ( with only Coulomb interaction) as well as those with the addition of ion polarization , ionic correlation , and both effects . It shows that inclusion of only ion polarization can achieve comparable results to those with considering both ion polarization and ionic correlation ; instead, the ionic correlation in bulk solutions seems to play a minor role, because the values show severe deviations from those of whereas they are close to the data of the classical theory . A good example is montmorillonite with I (ionic strength) = 2.5 mmol· L− 1, where , , , and are equal to − 201.1, − 207.5, − 106.3, and − 109.7 mV, respectively. It can be found that the inclusion of ion polarization in and causes a dramatic decrease to surface potential and results in only approximately half the value from the classical theory . That is, the ion polarization should be of vital importance in all the presented cases (Table 1). A careful analysis of ionic correlation will be performed later.

Polarization and gathering of counterions in proximity to the negatively charged surfaces cause an enhancement of the shielding effect on the surface charges, which further leads to a substantial decrease to surface potential. It is in good agreement with the predicted result of ion polarization. Instead, few researchers[55, 56, 6163] insist that Hofmeister effects should result from the dispersion energy and ionic steric effect that are known to be significant only in the case of relatively high ion concentration. This seems to be a discrepancy with the fact that Hofmeister effects are more evident at low ionic concentrations due to the more pronounced enhancement of ion polarization in the electric double layer. With inclusion of ion polarization, the surface potentials ( and ) show an obvious decrease as compared with those from the classical theory (see Table 1).

Table 1. Values of ionic strength (I) of solution containing Na+ and Ca2+ ions and surface potentials (φ 0) for montmorillonite and rutile titania (TiO2) suspended in these solutions that are obtained by the various schemes in Section 2. Units of ionic strength and surface potential are mmol· L− 1 and mV, respectively.

Figure 3 shows the variations of surface potentials for illite, i.e., , , , and with the root square of ionic strength , [43] corresponding to a much wider range of ionic concentrations than those shown in Table 1. Being in accordance with the results of montmorillonite and rutile titania (TiO2), the surface potentials of illite decrease with the increase of ionic strengths, and especially, a more pronounced change rate on the surface potentials has been detected at relatively low ionic strengths. The logarithm functions can be established between the calculated surface potentials and root square of ionic strengths as follows:

The correlation coefficients (R2) given in Eqs. (27)– (30) indicate that these fitted functions describe reasonably the relationships between surface potentials (, , , and ) and ionic strengths (I). The correlation coefficients from the classical theory with ion polarization correction are close to 1.0, which demonstrates the fine prediction results for the two surface potentials ( and ), while those relating to the ion correlation correction ( and ) are somewhat less satisfactory, which may be caused by the charge oscillation at high ion concentration. Nonetheless, the changed trends of these two surface potentials (Eqs. (28) and (30)) are correctly described and also in agreement with the other two cases (Eqs. (27) and (29)). According to the versus results in Table 1 and Fig. 3, the surface potentials will be altered by 20.0% ∼ 50.0% when ion polarization is included. For surfaces with variable charges, charge and surface potential may be reversed at high ionic strengths; [56, 64] however, the montmorillonite and illite systems under study possess permanent charges resulting from isomorphous substitutions, and the surface charge densities will not be easily altered by ion strength, polarization, and other factors. Accordingly, charge oscillation may take place occasionally, while charge reversion is unlikely to occur.

The contribution of ion polarization can be evaluated by the difference between Eqs. (30) and (28) and expressed as

Therefore, ion polarization in the electric double layer strongly affects the surface potential and its contribution is quantitatively evaluated. The surface potential decays gradually within ionic solution, and then the potential distribution φ (x) can be described by the nonlinear PB equation.[36]

Fig. 3. Variations of surface potentials (φ 0, in unit V) of illite in ionic solutions with root square of ionic strength (, in mol1/2· L− 1/2).

The surface potential decays gradually within ionic solution, and the potential distribution φ (x) does not need to additionally consider the ion polarization because this factor has already been accounted for during the calculation of surface potential. In order to investigate the relationship between potential distribution and ionic concentration, two ionic strengths (0.01 mol· L− 1 and 0.6 mol· L− 1) are selected for the NaCl + CaCl2 mixed solution . The various surface potentials are calculated by use of Eqs. (27)– (30), and the corresponding variations of potential distribution φ (x) with decay distance x which are derived from Eq. (3) are plotted in Fig. 4. The ϕ (x)H and φ (x)HI potentials descend quickly with the increase of decay distance x, because the counterions with a larger distance away from the charged surface are dramatically less polarized. As a result, Hofmeister effects due to ion polarization show a pronounced decrease with the increase of decay distance (x).

The φ (x)H and φ (x)HI potentials approach to the null point at x = 2.40 nm for I = 0.6 mol· L− 1 whereas at x = 13.3 nm for I = 0.01 mol· L− 1, indicating an apparently faster decay rate at higher ion concentrations. In addition, at a given ionic strength, the φ (x)H and φ (x)HI potentials that include ion polarization vanish at the identical distances as the φ (x)I potential. This demonstrates that the forces generated by the surface charges are independent of ion polarization. At low concentrations (e.g., I = 0.01 mol· L− 1 in Fig. 4(a)), ion polarization makes an important contribution to the surface potential and potential distribution, and the inclusion of only ion polarization in the electric double layer is already sufficient to obtain reasonable results as indicated by the comparable φ (x)H and φ (x)HI data (Fig. 4), which accord with the surface potential data shown in Table 1 and Fig. 3. However, at relatively high concentrations (e.g., I = 0.6 mol· L− 1 in Fig. 4(b)), the φ (x)H results seem to deviate significantly from those of φ (x)HI, and it suggests that only the ion polarization effect cannot correctly account for the surface potential nor potential distribution, which will be discussed in the following.

3.3. Influence of ionic correlation on surface potential

As indicated in Fig. 3, surface potentials including ionic correlation in bulk solution change inversely with the increase of ionic strength. Ionic correlation affects the distribution of counterions near the charged surface and further the surface potentials. At low ionic concentrations, surface potentials that include only ionic correlation have obvious deviations from those of but are rather close to the classical surface potentials . This indicates that at low ion concentrations, ionic correlation plays a minor role in determining the surface potential. However, when ion concentrations gradually increase, the values begin to diverge from those of , and instead the and results become gradually closer to each other, implying a strengthening role played by ionic correlation.[35] In a manner similar to Eq. (31), the contribution of ionic correlation to surface potential can be assessed by the difference between Eqs. (30) and (29), and expressed mathematically as

It can be found that at high ion concentrations, ionic correlation affects surface potentials even more apparently than ion polarization (Fig. 3). For example, at I = 0.6 mol· L− 1, and . The former neglects ionic correlation and results in an obviously larger deviation from . Accordingly, ionic correlation is a factor that greatly affects surface potential, especially at high ion concentrations. At the same time, it is found that even at such high ion concentrations, ion polarization cannot be ignored ( for I = 0.6 mol· L− 1), which is quite different from the effect of ionic correlation at low ion concentrations. In addition, this gives a good explanation to the very slight influence of ionic correlation on the data shown in Table 1, for all the investigated cases therein are of rather low ion concentration.

Figure 4 shows that the potential distribution with considering the ionic correlation [φ (x)I] becomes obviously less with the increase of decay distance (x), and the decay rates of φ (x)I are more dramatic at higher ion concentrations. As discussed in Subsection 3.2, at a given ionic strength, the φ (x)I, φ (x)H, and φ (x)HI potentials vanish at the same distances that are strongly concentration-dependent, that is, these potentials are equal to zero at x = 2.40 nm for I = 0.6 mol· L− 1 while x = 13.3 nm for I = 0.01 mol· L− 1, respectively. This is assertive evidence for the association of concentration-associated ionic correlation for surface potential with the potential distribution. Accordingly, ionic correlation in bulk solution should be included when determining the surface potential and potential distribution.

At rather low ionic strengths (I ≤ 0.01 mol· L− 1), the polarization effect predominates during the determination of surface potential and potential distribution, whereas the effect of ionic correlation can be almost neglected (Figs. 3 and 4). With the increase of ionic strength, each of these two effects plays a definite role and it is their synergy that results in the accurate surface potential and potential distribution. Compared with the referenced data, the polarization effect is underestimated whereas ionic correlation overestimates surface potential. Therefore, there should exist a critical point on the curve where these two factors can be exactly counteracted, and for the illite system, the critical ionic strength (Ic) equals 0.16 mol· L− 1 (see Fig. 5), that is, at I < 0.16 mol· L− 1, ; at I = 0.16 mol· L− 1, , and at I > 0.16 mol· L− 1, . The presence of critical ionic strengths (Ic) can be safely extended to other systems albeit the exact values may change accordingly. The ion polarization and ionic correlation make a larger contribution to the surface potential at very low and high concentrations, respectively (Fig. 5). What should be noteworthy is that the effect of ionic correlation can be almost neglected at rather low ion concentrations, while ion polarization, although it makes less of a contribution to the ionic correlation at high concentrations, is still indispensable for the surface potential and potential distribution [φ (x)HI] and hence should be considered across the entire concentration range (see Fig. 5).

Fig. 4. Variations of the potential distribution (∣ φ (x)∣ , in V) with decay distance (x, in nm) at two different values of ionic strength (I): (a) 0.01 mol· L− 1 and (b) 0.6 mol· L− 1.

Fig. 5. Illustration of region divisions of surface potentials (, in V) of illite according to the relative contributions of ion polarization and ionic correlation.

4. Conclusions

The surface potential has been widely used to characterize the interfacial property, dynamics, and reactivity, while it is currently still a difficult task for the direct measurement. In this work, a number of significant pending problems appearing in the recently proposed surface potential expression are addressed, thus providing insightful clues for the direct measurement and striding towards the understanding of surface potential and further interfacial behaviors.

The various theories of surface potentials are elaborated, including the classical theory and recently developed one covering both ion polarization and ionic correlation. The ion polarization that is responsible for Hofmeister effects may be distinct for the various ions, and the polarization difference between two selected ions shows an obvious increase with the decrease of ion concentration. These cannot be explained properly by the classical theory and so the predicted surface potentials deviate substantially from those from the recently developed theory. The polarization effect decreases with the increase of ion concentration and causes the surface potential and potential distribution to decrease dramatically, especially at low ionic concentrations. Unlike the ion polarization that underestimates the surface potential, ionic correlation overestimates the surface potential and plays an increasing role with the increase of ion concentration. Surface potential decays quickly in ionic solution, and the null points of decay distances are identical for the various theories but different from the various ionic concentrations, thus providing good support of the fact that ionic correlation should be associated closely with surface potential.

The respective contributions of ion polarization and ionic correlation to the surface potential are evaluated for different systems, and the results indicate that these two factors are more significant at low and high ionic concentrations. There exists a critical point on each of the surface potential curves where the contributions of these two factors can be counteracted exactly. For an illite system, the critical ionic strength (Ic) equals 0.16 mol· L− 1. The ionic correlation can be almost neglected at low ionic concentrations, while ion polarization also plays a definite role at high ionic concentrations albeit less than ionic correlation should be considered in calculating the surface potential and potential distribution across the entire concentration range. In addition, the apparent difference between surface and zeta (ξ ) potentials can be viewed through the analysis of potential distributions, which clearly shows that zeta (ξ ) potential is obviously less than surface potential and the confusion of these two conceptions can lead to ridiculous results. These conclusions should be applicable to other interfacial processes.

Reference
1 Somasundaran P, Markovic B, Krishnakumar S and Yu X 1997 Hand book of Surface and Colloid Chemistry New York CRC Press [Cited within:1]
2 Ghosal S 2002 Anal. Chem. 74 4198 DOI:10.1021/ac025630t [Cited within:1] [JCR: 5.695]
3 Zhang X, Grimes B A, Wang J C, Lacki K M and Liapis A I 2004 J. Colloid Interface Sci. 273 22 DOI:10.1016/j.jcis.2003.10.015 [Cited within:1] [JCR: 3.172]
4 Li H, Qing C L, Wei S Q and Jiang X J 2004 J. Colloid Interface Sci. 275 172 DOI:10.1016/j.jcis.2003.12.055 [Cited within:2]
5 Hou J and Li H 2009 Soil Sci. Soc. Am. J. 73 1658 DOI:10.2136/sssaj2008.0017 [Cited within:2] [JCR: 1.821]
6 Kallay N, Preočanin T and Ivšić T 2007 J. Colloid Interface Sci. 309 21 DOI:10.1016/j.jcis.2006.10.075 [Cited within:1]
7 Li H, Wei S, Qing C and Yang J 2003 J. Colloid Interface Sci. 258 40 DOI:10.1016/S0021-9797(02)00077-2 [Cited within:1] [JCR: 3.172]
8 Begum S 2003 J. Colloid Interface Sci. 261 399 DOI:10.1016/S0021-9797(03)00076-6 [Cited within:1] [JCR: 3.172]
9 Zarzycki P, Rosso K M, Chatman S, Preočanin T, Kallay N and Piasecki W 2010 Croat. Chem. Acta 83 457 [Cited within:1] [JCR: 0.614]
10 Borukhov I, Andelman D and Orland H 1997 Phys. Rev. Lett. 79 435 DOI:10.1103/PhysRevLett.79.435 [Cited within:1] [JCR: 7.943]
11 Tong Z Y, Zhu Y J and Tong C H 2014 Chin. Phys. B 23 038202 DOI:10.1088/1674-1056/23/3/038202 [Cited within:1] [JCR: 1.148] [CJCR: 1.2429]
12 Peng B and Yu Y X 2009 J. Chem. Phys. 131 134703 DOI:10.1063/1.3243873 [Cited within:1]
13 Wang K, Yu Y X and Gao G H 2008 J. Chem. Phys. 128 185101 DOI:10.1063/1.2918342 [Cited within:1] [JCR: 1.578]
14 Sun Z L, Kang Y S, Kang Y M and Song J M 2014 Chin. Phys. B 23 126401 DOI:10.1088/1674-1056/23/12/126401 [Cited within:1] [JCR: 1.148] [CJCR: 1.2429]
15 Yu Y X, Wu J Z and Gao G H 2004 J. Chem. Phys. 120 7223 DOI:10.1063/1.1676121 [Cited within:1] [JCR: 1.578]
16 Kunz W, Henle J and Ninham B 2004 Curr. Opin. Colloid Interface Sci. 9 19 DOI:10.1016/j.cocis.2004.05.005 [Cited within:2] [JCR: 6.629]
17 Levin Y, Santos A P D and Diehl A 2009 Phys. Rev. Lett. 103 257802 DOI:10.1103/PhysRevLett.103.257802 [Cited within:1] [JCR: 7.943]
18 Nostro P L and Ninham B W 2012 Chem. Rev. 112 2286 DOI:10.1021/cr200271j [Cited within:1] [JCR: 41.298]
19 Kunz W, Lo Nostro P and Ninham B 2004 Curr. Opin. Colloid Interface Sci. 9 1 DOI:10.1016/j.cocis.2004.05.004 [Cited within:1] [JCR: 6.629]
20 Fujimoto S and Yu Y X 2010 Chin. Phys. B 19 088701 DOI:10.1088/1674-1056/19/8/088701 [Cited within:1] [JCR: 1.148] [CJCR: 1.2429]
21 Yu Y X and Fujimoto S 2013 Sci. China: Chem. 56 1735 DOI:10.1007/s11426-013-4959-9 [Cited within:1] [JCR: 1.199] [CJCR: 0.4146]
22 Lu J F, Yu Y X and Li Y G 1993 Fluid Phase Equilib. 85 81 DOI:10.1016/0378-3812(93)80006-9 [Cited within:1]
23 dos Santos A P and Levin Y 2013 Faraday Discuss. 160 75 DOI:10.1039/C2FD20067H [Cited within:2] [JCR: 3.821]
24 Yu Y X, Gao G H and Li Y G 2000 Fluid Phase Equilib. 173 23 DOI:10.1016/S0378-3812(00)00396-4 [Cited within:1]
25 Pashley R M, McGuiggan P M, Ninham B W, Brady J and Evans D F 1986 J. Phys. Chem. 90 1637 DOI:10.1021/j100399a037 [Cited within:1]
26 Tsao Y H, Evans D F, Rand R P and Parsegian V A 1993 Langmuir 9 233 DOI:10.1021/la00025a045 [Cited within:1] [JCR: 4.187]
27 Salis A, Cristina Pinna M, Bilaničová D, Monduzzi M, Nostro P L and Ninham B W 2006 J. Phys. Chem. B 110 2949 DOI:10.1021/jp0546296 [Cited within:1] [JCR: 3.607]
28 Boström M, Craig V S J, Albion R, Williams D R M and Ninham B W 2003 J. Phys. Chem. B 107 2875 DOI:10.1021/jp026804d [Cited within:3] [JCR: 3.607]
29 Liu X, Li H, Du W, Tian R, Li R and Jiang X 2013 J. Phys. Chem. C 117 6245 DOI:10.1021/jp312682u [Cited within:3] [JCR: 4.814]
30 Parsons D F, Boström M, Nostro P L and Ninham B W 2011 Phys. Chem. Chem. Phys. 13 12352 DOI:10.1039/c1cp20538b [Cited within:1] [JCR: 3.829]
31 Li H, Hou J, Liu X and Wu L 2011 Soil Sci. Soc. Am. J. 75 2128 DOI:10.2136/sssaj2010.0301 [Cited within:5] [JCR: 1.821]
32 Liu X, Li H, Li R, Tian R and Xu C 2013 Analyst 138 1122 DOI:10.1039/C2AN36069A [Cited within:3] [JCR: 3.969]
33 Levin Y 2002 Rep. Prog. Phys. 65 1577 DOI:10.1088/0034-4885/65/11/201 [Cited within:1] [JCR: 13.232]
34 Patey G N 1980 J. Chem. Phys. 72 5763 DOI:10.1063/1.438997 [Cited within:1] [JCR: 1.578]
35 Liu X, Li H, Li R, Tian R and Xu C 2012 Commun. Theor. Phys. 58 437 DOI:10.1088/0253-6102/58/3/20 [Cited within:2]
36 Liu X, Li H, Li R and Tian R 2013 Surf. Sci. 607 197 DOI:10.1016/j.susc.2012.09.008 [Cited within:3] [JCR: 1.838]
37 Li H and Wu L S 2007 Soil Sci. Soc. Am. J. 71 1694 DOI:10.2136/sssaj2007.0004 [Cited within:3] [JCR: 1.821]
38 Li H and Wu L S 2007 New J. Phys. 9 357 DOI:10.1088/1367-2630/9/10/357 [Cited within:2] [JCR: 4.063]
39 Yokoyama H and Yamatera H 1975 Bull. Chem. Soc. Jpn. 48 1770 DOI:10.1246/bcsj.48.1770 [Cited within:1]
40 Davies C W 1962 Ion Association Washington D. C. Butterworths [Cited within:1]
41 Liu X, Li H, Li R, Tian R and Hou J 2012 J. Soils Sed. 12 1019 DOI:10.1007/s11368-012-0517-x [Cited within:1]
42 Boroudjerdi H, Kim Y W, Naji A, Netz R R, Schlagberger X and Serr A 2005 Phys. Rep. 416 129 DOI:10.1016/j.physrep.2005.06.006 [Cited within:1] [JCR: 22.929]
43 Bolt G H 1955 Soil Sci. 79 267 DOI:10.1097/00010694-195504000-00004 [Cited within:3] [JCR: 1.051]
44 Mullet M, Fievet P, Reggiani J C and Pagetti J 1997 J. Membr. Sci. 123 255 DOI:10.1016/S0376-7388(96)00220-7 [Cited within:1]
45 Mao M, Fornasiero D, Ralston J, Smart R S C and Sobieraj S 1994 Colloids Surf. A 85 37 DOI:10.1016/0927-7757(94)02745-5 [Cited within:1]
46 Bowen W R and Mukhtar H 1993 Colloids Surf. A 81 93 DOI:10.1016/0927-7757(93)80237-9 [Cited within:1]
47 Smit W and Holten C L M 1980 J. Colloid Interface Sci. 78 1 DOI:10.1016/0021-9797(80)90489-0 [Cited within:1] [JCR: 3.172]
48 Barber J 1980 Biochim. Biophys. Acta Bioenerg. 594 253 DOI:10.1016/0304-4173(80)90003-8 [Cited within:1]
49 Kinraide T B and Wang P 2010 J. Exp. Bot. 61 2507 DOI:10.1093/jxb/erq082 [Cited within:1] [JCR: 5.242]
50 Dobrzynska I, Skrzydlewska E and Figaszewski Z 2006 Bioelectrochem. 69 142 DOI:10.1016/j.bioelechem.2006.01.004 [Cited within:1]
51 Manning G S 1981 J. Phys. Chem. 85 1506 DOI:10.1021/j150611a011 [Cited within:1] [JCR: 1.578]
52 Stellwagen E and Stellwagen N C 2003 Biophys. J. 84 1855 DOI:10.1016/S0006-3495(03)74993-5 [Cited within:1] [JCR: 3.668]
53 Liu X, Li H, Li R, Xie D, Ni J and Wu L 2014 Sci. Rep. 4 5047 DOI:10.1038/srep05047 [Cited within:1] [JCR: 15.333]
54 Frydel D 2011 J. Chem. Phys. 134 234704 DOI:10.1063/1.3598476 [Cited within:1] [JCR: 1.578]
55 Parsons D F and Ninham B W 2010 Langmuir 26 1816 DOI:10.1021/la902533x [Cited within:3] [JCR: 4.187]
56 Parsons D F and Ninham B W 2010 Langmuir 26 6430 DOI:10.1021/la9041265 [Cited within:3] [JCR: 4.187]
57 Schwierz N, Horinek D and Netz R R 2010 Langmuir 26 7370 DOI:10.1021/la904397v [Cited within:1] [JCR: 4.187]
58 Grahame D C 1950 J. Chem. Phys. 18 903 DOI:10.1063/1.1747807 [Cited within:1] [JCR: 1.578]
59 Joshi R P, Qian J, Schoenbach K H and Schamiloglu E 2004 J. Appl. Phys. 96 3617 DOI:10.1063/1.1789274 [Cited within:1] [JCR: 0.71]
60 Gavryushov S and Linse P 2003 J. Phys. Chem. B 107 7135 DOI:10.1021/jp030035w [Cited within:1] [JCR: 3.607]
61 Parsons D F and Ninham B W 2011 Colloids Surf. A 383 2 DOI:10.1016/j.colsurfa.2010.12.025 [Cited within:1]
62 Kosmulski M 2002 Langmuir 18 785 DOI:10.1021/la0155653 [Cited within:1] [JCR: 4.187]
63 Johnson S B, Scales P J and Healy T W 1999 Langmuir 15 2836 DOI:10.1021/la980875f [Cited within:1] [JCR: 4.187]
64 Wang K, Yu Y X and Gao G H 2004 Phys. Rev. E 70 011912 DOI:10.1103/PhysRevE.70.011912 [Cited within:1] [JCR: 2.313]