A number of observations related to interfacial electrostatics of polar liquids question the traditional assumption of dielectric theories that bulk dielectric properties can be continuously extended to the dividing surface separating the solute from the solvent. The deficiency of this approximation can be remedied by introducing local interface susceptibilities and the interface dielectric constant. Asymmetries of ionic hydration thermodynamics and of the mobility between cations and anions can be related to different propensities of the water molecules to orient their dipole toward and outward from solutes of opposite charges. This electrostatic asymmetry is reflected in different interface dielectric constants for cations and anions. The interface of water with neutral solutes is spontaneously polarized due to preferential water orientations in the interface. This phenomenon is responsible for a nonzero cavity potential directly related to a nonzero surface charge. This connection predicts that particles allowing a nonzero cavity potential must show mobility in an external electric field even if the net charge of the particle is zero. The theory predicts that a positive cavity potential and a positive surface charge translate to an effectively negative solute charge reported by mobility measurements. Passing of the cavity potential through a minimum found in simulations might be the origin of the maximum of mobility vs the ionic size observed experimentally. Finally, mobility of proteins in the field gradient (dielectrophoresis) is many orders of magnitude greater than predicted by the traditionally used Clausius-Mossotti equation. Two reasons contribute to this disagreement: (i) a failure of Maxwell’s electrostatics to describe the cavity-field susceptibility and (ii) the neglect of the protein permanent dipole by the Clausius-Mossotti equation. An analytical relation between the dielectrophoretic susceptibility and dielectric spectroscopy of solutions provides direct access to this parameter, confirming the failure of the Clausius-Mossotti equation in application to protein dielectrophresis.

All materials satisfy the Coulomb law at the microscopic level. It states that for any two particles carrying charges q1 and q2 and separated by the distance r12, the potential energy is written as (Gaussian units1)

U(r12)=q1q2r12.
(1)

A certain level of coarse-graining is typically introduced in condensed-matter calculations and in atomistic numerical simulations by assigning charges qj to composite particles: atoms or molecular groups. Such a formulation is still referred to as the microscopic form of the Coulomb law since all atomic coordinates rj are followed in the simulation trajectory. We will follow this adopted practice and designate this very detailed level of description as the microscopic form of the Coulomb law.

Most practical calculations applied to complex systems, solutions, and interfaces depart from this detailed microscopic picture by applying some level of averaging. A powerful approach, leading to equations of electrodynamics of materials,2 replaces the complete information about the charges and positions of the atoms with continuous scalar and vector fields. The main challenge here is how to incorporate specific properties of interfaces into the field equations. The traditional formulation of electrodynamics of continuous media2 assumes that bulk material properties can be extended all the way to the dividing surfaces separating different components of the inhomogeneous material. If one component is a liquid, this mixed substance is known as a solution. The focus of this discussion is on electrostatics of solutions.

If one is concerned only with the interaction of charges q1 and q2 in a condensed medium, the only parameters relevant for such an observation are the distance between the charges and the thermodynamic state of the material. When one fixes only the thermodynamic parameters of the system (temperature, pressure, etc.) allowing all possible microscopic configurations, one arrives at the free energy of the solution characterizing its thermodynamic state. If, in addition, the microscopic distance r12 is fixed while allowing all possible configurations consistent with this constraint, the result is the potential of mean force, which is a partial free energy depending on a single remaining microscopic coordinate r12. Application of polarization fields of classical electrostatics leads to the factor ϵ1 in Eq. (1),

U(r12)=q1q2ϵ(T,P)r12,
(2)

which is often called the screened potential. The dependence of the dielectric constant ϵ(T,P) on the thermodynamic state of the material is an indication that U(r12) carries the meaning of the free energy, i.e., the reversible work required to separate two charges to the infinite distance in the liquid. Within the dielectric theories, the same function ϵ(T,P) appears when measuring polarization of bulk dielectrics in dielectric experiments.3 

A similar procedure of reducing the system’s manifold of degrees of freedom is used in the calculations of the solvation free energy, i.e., the chemical potential of a solute at infinite dilution. Here, the polarization field of the medium is integrated out in the external interaction potential produced by the solute. One naturally wonders if the polarization of the interface induced by the solute field is described by the same rules as dielectric screening at large distances r12 in Eq. (2); i.e., whether the same dielectric constant can be used to construct the electrostatic response for both problems. The answer given to this question by electrostatic theories of dielectrics is affirmative. Here, we argue that there are a number of observations where departure from this answer is required and a separate susceptibility of the interface needs to be constructed. It is not reduced to either ϵ of bulk dielectric or to its algebraic manipulations.

The assumption of continuous electrodynamics that bulk properties apply to all components of the solution up to the dividing surfaces supposes that the spatial dimension of the interface is much smaller than the distance on which continuous electrostatic fields vary. It is this assumption that often fails in describing solutions of molecular solutes in molecular solvents, a situation typical for solution chemistry and biochemistry. In molecular systems, the field of solute charges and the response of the solvent in the interface often vary on the same length-scale, thus invalidating the continuum assumption. Alternative coarse-graining solutions, at least partially incorporating the microscopic structure of the solute-solvent interface into the equations for continuous fields, are required. We will discuss below how such extensions of the continuum picture can be realized for electrostatics of solutions and for the calculation of free energies and forces acting on particles dissolved in polar liquids. It turns out that the basic equations of electrostatics of continuous fields can be preserved at the expense of introducing local interface susceptibilities distinct from material bulk properties.

In applications to electrostatics of solutions, the common practice is to separate the molecular charges into those of the solutes and those of the solvent. Since solutes are brought into the liquid solvent to create the solution, the charges of the solutes, in accord with the standard formulations of electrostatics,1 are viewed as external (free) charges with the density ρ0(r)=jqjδ(rrj) characterizing the distribution of partial atomic charges qj placed at positions rj within a given solute molecule. The charges of all N0 solutes can be obtained from this charge distribution by translation and rotation transformations. Since we assume an ideal solution here, considering a single solute is sufficient for our purposes. For instance, a simple ion carrying the charge q can be placed at the origin of the coordinate system producing ρ0(r)=qδ(r) and drρ0(r)=q.

In addition to the solute charge density ρ0(r), there are atomic charges at the molecules of the solvent. Their distribution, regardless of the assignment of the charges to a specific molecule of the solvent, can be characterized by a scalar field of the instantaneous microscopic (subscript “m”) charge density ρm(r) representing a specific instantaneous configuration of the solvent. Thermal agitation alters ρm(r) making it a fluctuating field, but the Coulomb law applies to each configuration of the nuclei in the system. The instantaneous microscopic (subscript “m”) electrostatic potential is

ϕm(r)=ϕ0(r)+drρm(r)|rr|,
(3)

where ϕ0(r) is the electrostatic potential created by the solute charges ρ0(r).

The conservation of the total charge of the solvent requires that the fluctuating scalar field of the solvent charge density can be replaced by the divergence of the vector field P(r) known as the polarization density,2,4

ρm=P.
(4)

This equation indicates that there is a charge associated with any small volume in the liquid in which P(r) is nonuniform. Quoting from Feynman:4 “If there is a nonuniform polarization, its divergence gives the net density of charge appearing in the material. We emphasize that this is a perfectly real charge density: we call it ‘polarization charge’ only to remind ourselves how it got there.” The emphasis on “real” is a warning against a common misconception that ρm is not a physical charge. Indeed, because of complexities of describing the distribution of molecular charges in condensed materials, classical electrostatic theories avoid calculating ρm by formulating the problem in terms of free charges ρ0.1 Nevertheless, there is no fundamental distinction between ρ0 and ρm, and modern atomistic simulations treat both on equal footing.5 

Equation (4), which is based on the conservation of charge, in principle applies to both bulk materials and interfaces. However, for interfacial problems, the formulation is significantly simplified by separating the electrostatics arising from the slowly varying P in the bulk from fast changing liquid polarization in the interface. The microscopic potential ϕm then becomes a sum of the interface (surface), ϕs, and bulk, ϕb, contribution. As we show below, the bulk contribution is constant and thus vanishes in the electrostatic field Eb=ϕb.

Fast alteration of the polarization field in the interface can be accounted for by multiplying the bulk polarization field P with a function changing sharply from unity to zero in the interface. In the limit when the size of the interfacial region is significantly smaller than the characteristic size of the solute, a step function θΩ(r) can be applied. This is a Heaviside function, which is equal to zero inside the solute and is equal to unity inside the solvent. If θΩP is substituted instead of P in Eq. (4), one obtains

ρm(r)=θΩ(r)P+Pnδ(rrS).
(5)

Here, Pn is the projection of the polarization density field on the unit vector normal to the dividing surface at the point rS and directed outward from the dielectric and δ(r) is the delta-function. The instantaneous microscopic potential created by the solvent then becomes the sum of the potential of the bulk charges ϕb and of the surface charges ϕs,

ϕm=ϕ0+ϕb+ϕs.
(6)

The potential components are given by integrals over the surface S of the solute and over the volume Ω occupied by the solvent,

ϕs(r)=SdSPn(rS)|rrS|,ϕb(r)=Ωdrρm(r)|rr|.
(7)

As in Eq. (3), ρm(r) defines the microscopic (instantaneous) charge density in the bulk. The presence of the interface is responsible for the surface electrostatic potential ϕs, which fluctuates due to thermal fluctuations of Pn. The next step is to specify the observable statistical averages of the electrostatic potentials ϕs and ϕb.

Statistically averaged projection of the polarization field on the outward normal direction is identified in electrostatic theories with the surface charge density4 

σ=Pn,
(8)

where angular brackets denote an ensemble average. It is often stated that “the surface charge description accounts for the polarization of the entire solvent.”6 This result comes from the observation, following from the Helmholtz theorem,1 that the field produced by the bulk density ρm can be identified with the longitudinal component of the polarization field,7 

Eb=ϕb=4πPL.
(9)

Here, the “L” subscript denotes the longitudinal [irrotational, in contrast to solenoidal (transverse);1 see Fig. 1] projection of the vector field P. The main result here is that the field of the bulk can be identified with the local polarization field at the point where Eb is calculated. Since there is no polarization in the volume excluded from the liquid, Eb=0 inside the solute. Therefore, the bulk of the liquid produces no field inside the solute, and the electrostatic potential ϕb must be a constant. Only the knowledge of the surface charge is required to calculate the electrostatic field in a void.

FIG. 1.

Examples of the longitudinal (irrotational) (a) and solenoidal (transverse) (b) vector fields. The dot in (a) indicates the source of a divergent (longitudinal) field.

FIG. 1.

Examples of the longitudinal (irrotational) (a) and solenoidal (transverse) (b) vector fields. The dot in (a) indicates the source of a divergent (longitudinal) field.

Close modal

It is important to stress that this result is solely a consequence of the Coulomb law, combined with the requirement of expulsion of the liquid polarization from the solute volume. This conceptual framework puts emphasis on the interfacial liquid structure in formulating solution electrostatics. The bulk turns out to be irrelevant, and finding the electrostatic field in a void becomes an interfacial problem.8,9 We next explore this perspective with specific examples connecting the surface charge density with solvation and mobility of solutes in polar liquids.

The discussion presented above has shown that the problem of finding the statistically averaged electric field in a void,

E=ϕm,
(10)

can be fully formulated in terms of the surface charge density given by Eq. (8) and the density of external charges ρ0. The microscopic origin of σ is not specified: it might originate from molecular multipoles (dipoles, quadrupoles, etc.), induced dipoles due to electronic molecular polarizability, and surface hydrogen bonds. All these microscopic origins of the surface charge are averaged over the microscopic configurations to produce σ. As we show below, σ involves two components: (i) static surface charge due to spontaneous polarization of the interface and (ii) surface charge induced by external charges. The second component can be expressed in terms of the interface susceptibility or, alternatively, in terms of the interface dielectric constant. It is instructive to start the discussion with dielectric theories, where only the second component of the surface charge is considered and is calculated in terms of the dielectric constant of the bulk.

We start with the simplest configuration of an ion centered in the cavity created by its repulsive core [Fig. 2(a)]. The cavity radius is a and the charge q is placed at its geometrical center. In Maxwell’s electrostatics,1 the ion carrying the charge q polarizes the surrounding liquid to create the surface charge density

σM=PnM=ϵ14πϵE0n,E0n=qa2.
(11)

It carries a sign opposite to q, thus screening the electric field of the external charge E0=ϕ0 (“M” designates Maxwell’s electrostatics). When multiplied with the surface area S=4πa2, this uniformly distributed surface charge density yields the total charge of the interface (surface charge)

QM=σMS=q(1ϵ1).
(12)

The electrostatic potential in the liquid is a sum of the solute potential ϕ0=q/r and the surface potential ϕsM=QM/r, thus yielding the standard expression for the screened potential q/(ϵr) inside the liquid. In contrast to the decaying potential ϕsM in the liquid, a constant potential is created by the uniform surface charge inside the cavity

ϕc(q)=a1QM(q).
(13)
FIG. 2.

(a) A positive ion with the charge q and the cavity radius a immersed in a dielectric with the dielectric constant ϵ. The surface charge σ of the sign opposite to q screens the ion charge. (b) The dependence of the electrostatic potential in the cavity on the ionic charge for ion hydration.10 Two qualitative features found in simulations of ion hydration are indicated in the plot: (i) a nonzero offset potential ϕc(0)=ϕcst0 and (ii) a higher slope for anions compared to cations.

FIG. 2.

(a) A positive ion with the charge q and the cavity radius a immersed in a dielectric with the dielectric constant ϵ. The surface charge σ of the sign opposite to q screens the ion charge. (b) The dependence of the electrostatic potential in the cavity on the ionic charge for ion hydration.10 Two qualitative features found in simulations of ion hydration are indicated in the plot: (i) a nonzero offset potential ϕc(0)=ϕcst0 and (ii) a higher slope for anions compared to cations.

Close modal

The reversible work of bringing charge q into the void is the electrostatic component of the free energy of ion solvation. Since ϕc(q)q, the reversible work of charging the cavity is quadratic in q,

F0=12ϕc(q)q=q2/(2a)(1ϵ1).
(14)

This is the celebrated Born equation.11 The measurable free energy of bringing an ion from vacuum to solution includes, in addition to F0, a number of additional components12 (see below). However, as we discuss next, even the free energy of electrostatic solvation needs modification to account for the static electrostatic potential, independent of q, within the solute cavity.

The quadratic scaling with the ion charge F0q2 predicted by the Born equation has been confirmed with great accuracy by computer simulations of hydrated ions.10,13 Nevertheless, two deviations from the Born equation were identified: (i) a nonzero offset potential ϕc(0)=ϕcst0 at zero ionic charge and (ii) different slopes of ϕc(q) for anions and cations.10,13–15 According to simulations10,15–18 and from experimental hydration enthalpies,19 accounting for different slopes requires either ϵ+>ϵ or a+<a in Eq. (14) [Fig. 2(b)]. Here, the ± subscript is applied not to the charge of the ion, but instead to the surface charge in the interface: cations produce a negative screening charge (“” subscript) and anions create a positive screening charge (“+” subscript). Therefore, hydration of anions leads to a more compact and more polar hydration shell compared to hydration of cations.20,21

If preferential molecular orientations are caused by specifics of the solute-solvent and solvent-solvent interactions, a static (spontaneous) interfacial polarization must follow. For interfaces with water, a nonuniform orientational distribution is usually related to two sources: (i) surface hydrogen bonds viewed as local interactions not reducible to the total charge q carried by the solute and (ii) competition between water’s dipoles and quadrupoles to minimize their free energy in the interface, which is present even for the water-vapor interface.22,23 A nonuniform orientational distribution of interfacial multipoles leads to a nonzero normal polarization density component Pn and to a uniform surface charge density σst integrating to a nonzero total surface charge Qst=σstS. This uniform surface charge will produce a static cavity potential23–25 [Eq. (7) and Fig. 2(b)]

ϕcst=a1Qst.
(15)

With the account for the static potential, the Born solvation free energy F0 becomes only a part of the overall electrostatic free energy16,23,26 (also labeled as the “intrinsic” hydration free energy27)

ΔGq=qϕcst+F0.
(16)

The first term here is the energy of the charge q in the static potential of the interface ϕcst, while the second term is the free energy of polarizing the interface by charge q, which is quadratic in the charge, F0q2 [Eq. (14)].

Two issues related to the static cavity potential need to be clarified. First, one has to realize that the polarization field P is a composite quantity including both the molecular dipoles m (induced and permanent) and molecular quadrupoles Q (as well as higher multipoles),

P(r)=jmjδ(rrj)13jQjδ(rrj)+,
(17)

where

Q=12kqk[3rkrkrk2I]
(18)

is the traceless molecular quadrupole28 defined through partial atomic charges qk at the coordinates rk relative to the center of mass of the solvent molecule. The sum in Eq. (17) runs over the solvent molecules with coordinates rj, thus producing the vector field of the polarization density.1,28 Therefore, the static potential ϕcst is caused by both dipolar and quadrupolar molecular orientations in the interface.

Direct calculations for SPC/E (extended simple point charge) water have shown that ϕcst is dominated by quadrupolar polarization for small cavities.23 This contribution, decaying as a1, becomes insignificant29 for a>6Å (assuming zero potential assigned to the bulk). The constant ϕcst reported for large cavities is mostly the result of preferential dipolar orientations in the interface. Second, the cavity potential ϕcst=ϕs0 is only one part in the overall surface potential ϕv0=ϕvs+ϕs0, which represents the potential drop between the vapor and the potential inside the liquid cavity. In addition to the potential drop ϕs0 at the liquid-solute interface considered here, it contains the potential drop ϕvs at the vapor-liquid interface.23,25,30,31 The total potential drop ϕv0, which comes as an incomplete compensation between ϕvs and ϕs0, needs to be considered for comparing calculated solvation free energies with experimental free energies ΔGsolv for transferring ions from vacuum into the liquid. In addition to ϕv0, the total ΔGsolv is affected by the free energy of cavity formation and solute-solvent dispersion and induction interactions.12,31 The free energy of electrostatic solvation ΔGq in Eq. (16) is only one component of the more complex ΔGsolv.

The second term, F0, in Eq. (16) is the Born solvation free energy due to the surface charge induced by the field of the solute. If one assumes that the cavity radius is not affected by altering solute charge (large solutes),24 the asymmetry of the slope of the induced cavity potential between cations and anions [Fig. 2(b)] implies different propensities of water to orient its dipole inward or outward from the ion, depending on the direction of the external field. To account for this effect, one can assign different induced charge densities σ± to the positive (+) and negative () surface charge and corresponding interface dielectric constants ϵ±. Obviously, this change of perspective implies that ϵ± is not a material property anymore but instead is an interfacial parameter.

Two specific interface dielectric constants ϵ± are special cases of the interface dielectric constant ϵint characterizing the integral ability of the interface to be polarized by a probe charge placed inside a void.7 For a spherical cavity with the charge q at its center, the interface dielectric constant defines the induced surface charge, which adds to the static charge to make the total surface charge

Q=(σst+σ)S=Qstq(1ϵint1).
(19)

The corresponding cavity potential becomes

ϕc(q)=ϕcst(q/a)(1ϵint1).
(20)

The cavity potential is the sum of the static (first term) and induced (second term) components. The difference between the microscopically derived charge Q in Eq. (19) and the dielectric charge QM in Eq. (12) is caused by two effects: Qst0 and ϵintϵ. To distinguish between the static and induced surface charge, one can define the interface susceptibility

4πχintq=1ϵint1.
(21)

The superscript “q” here marks the response of the interface to the placement of the charge to the solute, in contrast to the dipolar susceptibility of the interface χintd considered below, which describes the response to a dipole placed into the cavity. The charge susceptibility is obviously related to the Born solvation energy by the following equation:

4πχintq=a(ϕcstϕc(q))/q=2aF0/q2.
(22)

When ions of opposite charge are concerned, one can assign ϵint to ϵ for cations and to ϵ+ for anions by assuming constant slopes of ϕc(q) for each ion charge [Fig. 2(b)]. The dielectric constants ϵ± can, therefore, be estimated by analyzing the interfacial polarization induced in the surrounding polar liquid by cations and anions. This analysis has been recently performed for charged fullerenes in SPC/E water.32 The result is ϵ+=19.7 and ϵ=13.5 for q=1.0 and q=1.0, respectively. These values qualitatively agree with a steeper slope of ϕc(q) for anions compared to cations [Fig. 2(b)]. One, therefore, generally expects for interfacial water,

ϵ<ϵ+.
(23)

Inequality (23) is illustrated in Fig. 3 where 4πχintq is calculated from Eq. (22) by using simulated16 electrostatic solvation free energies ΔGq and the cavity potentials ϕcst for a number of cations and anions in SPC (simple point charge) force-field water. The cavity radius was not separately established in that study, and the sum of the ionic, ri, and water, rw, Lennard-Jones (LJ) radii is used to estimate a=ri+rw. The distinction between χintq for cations and anions is qualitatively consistent with Eqs. (21) and (23). One, however, finds 4πχintq>1, which corresponds to a negative ϵint. This difficulty implies that a taken as a sum of LJ radii is not a reliable approximation.

FIG. 3.

4πχintq calculated according to Eqs. (16) and (22) with solvation free energies ΔGq and static cavity potentials ϕcst from molecular dynamics simulations of charged Lennard-Jones (LJ) solutes in SPC water.16 The cavity radius a=ri+rw was taken as the sum of LJ radii of the ion, ri, and of the water molecule, rw. Artificial ions were used in simulations: from Li+ to I+ to LJ solutes with a=6, 8, and 10 Å for cations. The same solutes with q=1 were used for anions.16 

FIG. 3.

4πχintq calculated according to Eqs. (16) and (22) with solvation free energies ΔGq and static cavity potentials ϕcst from molecular dynamics simulations of charged Lennard-Jones (LJ) solutes in SPC water.16 The cavity radius a=ri+rw was taken as the sum of LJ radii of the ion, ri, and of the water molecule, rw. Artificial ions were used in simulations: from Li+ to I+ to LJ solutes with a=6, 8, and 10 Å for cations. The same solutes with q=1 were used for anions.16 

Close modal

Solvation of small ions is affected not only by the interface susceptibility, but also by often a significant change of the structure of the hydration shell induced by the ion. This difficulty is dealt with by different formalisms to specify the effective size of the solute, which is often expressed in terms of the solute-solvent radial distribution function (RDF). The cavity size roughly coincides with the first peak of the solute-solvent RDF, but more complex expressions involving integrals of the RDF follow from perturbation theories.18 We do not consider these more special cases here focusing solely on the induced surface charge and the corresponding interface susceptibility.

Both dielectric constants, ϵ and ϵ+, calculated for fullerenes in water32 are substantially below the dielectric constant of bulk SPC/E water (71). This outcome, shared by a number of simulation studies,7,32–36 suggests suppression of the interfacial response in the direction normal to the interface. Dipoles in the interface, frustrated by the local fields and geometric constraints,37–40 do not develop the complete dielectric screening of the bulk material. One anticipates a general result,41–44 

ϵint<ϵ.
(24)

Combined with Eqs. (12) and (21), this inequality also implies that dielectric theories yield the upper limit for the magnitude of the induced screening charge,

4π|q|χintq<|QM|.
(25)

Microscopic interfaces will screen the ionic charge less effectively than prescribed by dielectric theories.

We now remove the ion from the void and consider the simplest configuration related to mobility in a uniform external field: an empty cavity in a uniformly polarized liquid (Fig. 4). Polarization of the liquid leads to a surface charge at the void. According to electrostatics of dielectrics,1,2 the surface of the cavity gains the surface charge density σ(θ)=σ1cosθ corresponding to the first-order, =1, term in the expansion of the surface density in Legendre polynomials P(cosθ). This dipolar surface charge density creates a cavity dipole opposite in the direction to the external field with the magnitude Mint=σ1Ω0, where Ω0=(4π/3)a3 is the cavity volume. If the uniform polarization P is created in the medium, Maxwell’s electrostatics predicts1,2

MintM=3Ω02ϵ+1P.
(26)

The total charges of the negative, q=πσ1a2, and positive, q+=πσ1a2, lobes of the surface charge density are equal in the magnitude and the total surface charge Q=q++q is zero.

FIG. 4.

Cavity in the dielectric uniformly polarized with the external field E0. The surface charge density is negative, σ(θ)=σcosθ, of the right half-sphere (0<θ<π/2) and positive, σ(θ)=σ+cosθ on the left half-sphere (π/2<θ<π). The two lobes produce the dipole moment Mint in the direction opposite to the external polarizing field. The uniform polarization P in the bulk is created by the field of external charges E0. For the geometry of a plane capacitor, the Maxwell field inside the capacitor is E=E0/ϵ.

FIG. 4.

Cavity in the dielectric uniformly polarized with the external field E0. The surface charge density is negative, σ(θ)=σcosθ, of the right half-sphere (0<θ<π/2) and positive, σ(θ)=σ+cosθ on the left half-sphere (π/2<θ<π). The two lobes produce the dipole moment Mint in the direction opposite to the external polarizing field. The uniform polarization P in the bulk is created by the field of external charges E0. For the geometry of a plane capacitor, the Maxwell field inside the capacitor is E=E0/ϵ.

Close modal

From the discussion of ion solvation, it is easy to realize possible pitfalls of this picture when applied to liquids with asymmetric distribution of molecular charges, such as water.22,45 If the propensities of the solvent multipoles to orient toward and away from the uncharged solute are different, this should result in different surface charge densities for the positive and negative lobes of the surface charge. For hydration of solutes of the nanometer size, such as proteins, one can anticipate a scenario in which the solute surface provides sites with strong hydrogen bonds with the interfacial water molecules. In that case, orientations of water dipoles pointing toward the solute will be more probable, and the negative lobe in Fig. 4 will be strongly depopulated. One would expect Q>0 for this scenario.

The phenomenology of ion solvation suggests a route to improve the standard picture in terms of charge-specific σ±>0 and ϵ±>1. If σ(θ)=σ+cosθ is assigned to the positive lobe (θπ/2) and σ(θ)=σcosθ is assigned to the negative lobe (θπ/2, Fig. 4), the total charge of the interface becomes

Q=14S(σ+σ).
(27)

Correspondingly, the dipole moment of the void is oriented opposite to the external field and carries the magnitude

Mint=Ω0σ¯,σ¯=12(σ++σ).
(28)

The standard case of Maxwell’s electrostatic in Eq. (26) is recovered when there is no preference for the surface dipoles to orient either inward or outward and σ+=σ=σ1.

Two scenarios can be anticipated for these interfacial multipoles. The static charge leads to the electrostatic force fEE0 acting on the solute (see below), which is linear in the external field.46 On the contrary, the induced surface charge densities σ± can differ as a consequence of different susceptibilities of the interface with respect to the external field pointing toward the solute and away from it, as is shown in Fig. 4. In this case, the induced surface charge

Q(ϵ1ϵ+1)E0
(29)

will be positive given inequality (23) and scale linearly with the applied field. Correspondingly, the resulting force will be quadratic in the electric field of external charges E0,

fE(ϵ1ϵ+1)E02.
(30)

For both static and induced surface charges, no gradient of the external field is required to cause the force, in contrast to the dielectrophoretic force discussed below. The result is that uncharged solutes will experience a dragging force in a uniform external field.

An important consequence of Eq. (15) is that the static surface charge and mobility of uncharged solutes can be evaluated from the static cavity potential often available from simulations.15,18,24,47 The energy eϕcst is substantial: from 9kcal/mol for a hollow cavity24 with the radius up to 15Å to 1324kcal/mol for an uncharged protein where all partial charges were set equal to zero.47 A number of simulations have shown that ϕcst is nearly constant at a>6 Å. This observation translates into Qst increasing approximately linearly with the cavity size [Eq. (15)]. Given that mobility in the field is inversely proportional to the particle radius [Eq. (39)], one gets a nearly constant contribution of the static surface charge to the mobility of the solute. This contribution is affected not only by the identity of the solvent, but also by the structure of the interface.

This is illustrated in Fig. 5, where Qst calculated from Eq. (15) is plotted against the cavity radius a=rmax identified with the position of the first maximum rmax of the solute-solvent RDF.18 Two sets of points in Fig. 5, obtained from molecular dynamics simulations of nonpolar solutes of changing size, identify different strengths of LJ attraction between a nonpolar solute and TIP3P water.48 The surface charge density σst0.01e/nm2 produced by these simulations is below the range of 0.020.4e/nm2 typically extracted from mobility measurements of air bubbles and hydrophobic oil drops.49,50 However, the laboratory data are affected by ionic adsorption,51,52 and corresponding modifications of the surface charge are not included in our analysis. Surface hydrogen bonds and release of dangling O–H bonds53,54 will additionally affect Qst.

FIG. 5.

Static charge Qst (in units of the elementary charge) calculated from Eq. (15) by using the cavity potential ϕcst inside uncharged Kihara solutes solvated in TIP3P water.18 The Kihara solute is represented by a hard-sphere core decorated at its surface with a Lennard-Jones (LJ) solute-solvent interaction potential. The two sets of points are the results of simulations at two values of the solute-solvent LJ energy listed in the plot. The cavity radius is associated with the position of the first maximum of the solute-solvent radial distribution function, a=rmax.

FIG. 5.

Static charge Qst (in units of the elementary charge) calculated from Eq. (15) by using the cavity potential ϕcst inside uncharged Kihara solutes solvated in TIP3P water.18 The Kihara solute is represented by a hard-sphere core decorated at its surface with a Lennard-Jones (LJ) solute-solvent interaction potential. The two sets of points are the results of simulations at two values of the solute-solvent LJ energy listed in the plot. The cavity radius is associated with the position of the first maximum of the solute-solvent radial distribution function, a=rmax.

Close modal

Consider a sample of polar liquid with N0 solutes carrying charge q each placed in a plane capacitor producing the vacuum field E0. If the field is along the z-axis of the laboratory frame, the z-component of the total force acting on the sample is

Fz=E0Vdr[N0ρ0+ρb],
(31)

where the volume V can be chosen in a way that there is no polarization field at the volume boundary. If the density of the bound charge is taken as above, ρb=[θΩP], then the total bound charge is zero and

Fz=N0qE0.
(32)

If the overall neutral electrolyte is considered instead, the total force on the sample is obviously zero.

Calculating the mobility of an ion in an external field requires a somewhat different procedure. Instead of calculating the force acting on the sample, one calculates the force acting on all charges within the spherical shear surface of radius R, which then becomes hydrodynamic or the Stokes radius. The force acting on this volume ΩR=(4π/3)R3 is obviously

fE=χcE0ΩRdr[ρ0+ρb].
(33)

Here, χc is the cavity susceptibility correcting the external field E0 to the cavity field inside the shear surface. By allowing two dividing surfaces for the polar liquid, at the cavity radius a and at the Stokes radius R, one obtains

fE=χcqeffE0,
(34)

where

qeff=q+4π[a2Pr(a)R2Pr(R)].
(35)

Here, Pr(r) is the radial projection of the polarization field estimated at a and R. The radial projection Pr=Pn at r = a is the negative of the surface charge density composed of the field-induced, σ, and static, σst, components [Eq. (19)]: Pn=σst+σ. If one assumes that the field-induced components of Pr(R) and Pr(a) are specified by ϵ and ϵint, respectively, one obtains

qeff=qQst+q(ϵ1ϵint1).
(36)

Alternatively, one gets

qeff=q/ϵQ,
(37)

where Q is the screening charge in Eq. (19).

The standard dielectric result qeff=q follows at ϵint=ϵ and Qst=0. In this limit, there is no divergence of the radial polarization since P=r2(r2Pr)/r and PrE0r2 (Fig. 6). According to the standard rules of electrostatics discussed above, there is no bound charge density ρb created by this polarization field, and there is no net bound charge between two spherical surfaces at r=a and r=R. A nonzero bound charge of the hydration shell comes from the static charge due to spontaneous polarization of the interface combined with polarization divergence caused by the solvent dipoles polarized differently between the cavity surface and the shear surface (ϵintϵ). The radial polarization projection Pr(r) is typically a decaying oscillatory function in the interface.55 

FIG. 6.

Schematic representation of the distance decay of the radial projection Pr(r) of the polarization field in the solvation shell of the ionic solute with the cavity radius a. The decay Prr2 predicted by dielectric theories (dashed line) leads to the zero volume bound charge ρb=P. Any profile of Pr(r) deviating from this law (solid line) will produce a nonzero bound charge given in terms of the interface dielectric constant ϵint. Reproduced with permission from M. Dinpajooh and D. V. Matyushov, Physica A 463, 366 (2016). Copyright 2016 Elsevier B.V.

FIG. 6.

Schematic representation of the distance decay of the radial projection Pr(r) of the polarization field in the solvation shell of the ionic solute with the cavity radius a. The decay Prr2 predicted by dielectric theories (dashed line) leads to the zero volume bound charge ρb=P. Any profile of Pr(r) deviating from this law (solid line) will produce a nonzero bound charge given in terms of the interface dielectric constant ϵint. Reproduced with permission from M. Dinpajooh and D. V. Matyushov, Physica A 463, 366 (2016). Copyright 2016 Elsevier B.V.

Close modal

The stationary drift of an ion is established by balancing the electrostatic force with the force of hydrodynamic friction56,57 with the friction constant

ζH=6πηR.
(38)

Here, the stick boundary conditions are assumed and η is the solvent viscosity. Equating the drag force in Eq. (34) with the friction force, one arrives at the Hückel equation for the product of ionic mobility μ and viscosity η (the Walden product58),

(6πR)μη=ϵχcqeff,
(39)

where μ refers to the absolute mobility (mobility at zero ionic strength) and qeff is given by Eq. (36).

Equation (39) can be further simplified by applying Maxwell’s cavity susceptibility,1,59

χcM=32ϵ+1.
(40)

Here, as above, the superscript “M” is used to specify Maxwell’s electrostatics. This form of χc, which should be used with care (see below), allows one to assume ϵχcM3/2 for ϵ1 in Eq. (39). When this approximation is adopted, 6π/(ϵχcM) becomes equal to 4π, which corresponds to the slip boundary conditions in the standard models of mobility not accounting for the cavity-field corrections.58 Equation (39) can be changed to

4πμη=χcχcM|qR(1+ϵ1ϵint1)ϕcstaR|.
(41)

The hydrodynamic radius R is likely to deviate upward from the cavity radius a, and one can anticipate that the induced surface charge density at the cavity radius should generally be different from the induced surface charge density at the shear surface. The extent of this deviation is hard to estimate even from microscopic considerations. For instance, the interface dielectric constant in Eq. (19) is in practice determined from simulation configurations by averaging the dipolar response over about three solvation shells.32,33 The cavity and Stokes radii fall within this coarse-graining range for most small ions, and induced charges at the cavity and shear surfaces are hard to distinguish.

An intriguing experimental observation much discussed in the past is the appearance of the mobility maximum as a function of the ionic radius found for both anions and cations.58 Theories of this effect have focused on modifying the friction component in the force balance. The overall friction ζ experienced by an ion can be separated57 into the hydrodynamic component given by Eq. (38) (due to solute-solvent repulsion) and the friction ζs arising from the solute-solvent soft interactions: ζ=ζH+ζs. The assumed additivity between the repulsive hydrodynamic and soft friction components is not, however, supported by computer simulations.60 This observation questions the validity of all theoretical formalisms based on the additivity assumption.57,61

If the soft component is described by the electrostatic interaction with a continuum solvent, the result is the dielectric friction,21,57,60–63

ζs=cτDq2ϵR3.
(42)

Here, the numerical constant c differs between Zwanzig62 and Hubbard-Onsager63 formulations and τD is the Debye relaxation time of the solvent. While these traditional theories considered dielectric friction originating from collective dielectric relaxation of the medium, the molecular model of ionic mobility by Bagchi and Biswas61 placed the emphasis on fast ballistic (single-particle) dynamics of water in the interface with the solute as the driving force for mobility of small ions. Calculations supporting this view were based on the mean-spherical approximation, which, similarly to Eq. (42), results in ζsq2 scaling for the friction component originating from soft solute-solvent interactions.

Tests of the predicted scaling ζsq2/R3 by numerical simulations have not been entirely consistent. While no maximum in conductivity for uncharged solutes was found in early simulations,21 more recent studies using low-charge solutes, q=0.1e, have shown the existence of a maximum in mobility as a function of the solute radius.64 These recent results, confirming earlier simulations with neutral solutes,65 suggest that the mobility maximum is not related to dielectric friction but, instead, is a result of optimization, at a certain size, of solute’s migration between the voids in the liquid. Therefore, if a nonlinear dependence of mobility on the ionic size is to be related to soft friction, it is not driven by the electrostatic solute-solvent interactions. Furthermore, the maximum in the experimental conductivity (for Cs+ among monovalent cations in water) turns out to coincide with the minimum of the activation enthalpy as a function of the ionic size.64 This experimental fact does not seem to allow an obvious explanation in terms of dielectric friction.

Along these lines, a connection between the mobility maximum, the residence time of water molecules in the hydration shell, and the entropy of solvation was also proposed.21 These simulations, therefore, suggest that conductivity reaches maximum for the most disordered hydration shell. This explanation might be just another reflection of the same physical reality of reaching optimum migration of ions between voids in the liquid when the hydration shells are most disordered and loose.64 Similarly, simulations at elevated pressure found the mobility to increase with increasing pressure,66 which was explained by increased disorder of hydration shells caused by breaking the network of hydrogen bonds induced by pressure. To summarize, all these studies have placed the focus on the structure of the hydration shell as the origin of the mobility maximum, in contrast to the dissipation mechanisms advocated by the theories focused on dielectric friction.

The present perspective is focused on microscopic screening in the interface and does not offer new insights into the mechanism of energy dissipation of the moving ion into the polar liquid. In contrast to previous theoretical studies concerned solely with the friction mechanisms, the focus here is on the electrostatic drag force acting on the ion. From this perspective, the theory predicts a linear dependence of the Walden product on the static cavity potential ϕcst [Eq. (41)]. The static cavity potential was indeed found to pass through a minimum16 at the ionic size roughly consistent with the observed conductivity maximum (for Cs+ among monovalent cations58,61). This result implies that 4πμη in Eq. (41) is expected to pass through a maximum, as observed for ionic conductivity.

The results of simulations16 for ϕcst are collected in Fig. 7. The range of ϕcst achieved in simulations is insufficient to explain the conductivity maximum for cations. Both ϵintϵ and χcχcM might need to be considered.7,67 In particular, χcχcM found in simulations67,68 and discussed below in connection with protein dielectrophoresis might be the amplification factor converting a weak maximum of ϕcst to a much more robust mobility maximum found for cations. Importantly, the term containing ϕcst in Eq. (41) is independent of the ionic charge and, therefore, would explain the appearance of the mobility maximum in simulation of solutes with a low or zero charge.64,65 Physically, the minimum of ϕcst should reflect the most orientationally disordered hydration shell, in line with previous qualitative explanations connecting the shell disorder with the mobility maximum.21,66

FIG. 7.

ϕcst from molecular dynamics simulations of uncharged Lennard-Jones (LJ) solutes in SPC water16 vs the cavity radius a taken as the sum of LJ solute and water radii. Solutes used in simulations are the same as in Fig. 3: from Li0 to I0 to LJ solutes with a=6, 8, and 10 Å.16 The vertical arrow marks the size of Cs+ for which the maximum of mobility is recorded experimentally.61 

FIG. 7.

ϕcst from molecular dynamics simulations of uncharged Lennard-Jones (LJ) solutes in SPC water16 vs the cavity radius a taken as the sum of LJ solute and water radii. Solutes used in simulations are the same as in Fig. 3: from Li0 to I0 to LJ solutes with a=6, 8, and 10 Å.16 The vertical arrow marks the size of Cs+ for which the maximum of mobility is recorded experimentally.61 

Close modal

Explaining the conductivity maximum for anions will require additional studies. For anions, ϕcst enters with the positive sign the right-hand side of Eq. (41), thus producing a minimum. Whether the combination of this minimum with the first 1/R term can produce a mobility maximum is hard to establish given that ϵint is likely to be a nonmonotonic function of the solute radius. Interfacial water molecules next to a negatively charged surface undergo a structural crossover releasing dangling O–H bonds pointing to the substrate with increasing the charge of the substrate.38,69 This transition is accompanied by a spike of ϵint, which decays on both sides of the crossover point.32 

Equation (41) resolves an apparent contradiction between simulations of voids and LJ solutes producing a positive surface charge15,18,24,47 and a formally negative charge assigned to nonpolar, nonionic particles in water (oil drops, air bubbles, etc.) based on their mobility.49,70 Since Qst enters the effective ionic charge interacting with the field with the negative sign, the static surface charge yields a negative contribution to qeff. The negative charge of nonpolar solutes has been traditionally ascribed to adsorption of hydroxide ions to the interface. An alternative mechanism discussed here requires static interfacial orientation of the water molecules pointing their dipoles toward the nonpolar component [Qst>0 in Eqs. (15) and (36)]. Whether hydroxide49 or bicarbonate51 ions alone, or their combination with Qst>0, can explain the mobility of nonionic particles in water requires further studies. The perspective presented here allows direct access to this problem through atomistic simulations: one needs to calculate the cavity potential from simulations to access the static surface charge and combine this input with the average total charge within the shear surface from preferentially adsorbed ions.

A somewhat idealized configuration of the void in a polar liquid considered above can now be extended to the configuration relevant to a number of problems appearing when solutions are placed in an external electric field. Consider the solution placed in a spatially nonuniform electric field E0(r), which creates the nonuniform Maxwell field E(r). For most practical problems, E0(r) can be viewed as uniform on the length-scale of the solute. We, therefore, drop the spatial dependence and put E0(r)=E0. If the solute carries its own dipole moment M0, the external field aligns the dipole, thus creating an average dipole moment along the field,

M0E=χ0Ω0E0.
(43)

Here, the solute dipolar susceptibility χ0 yields the average dipole induced by the field, which otherwise averages out to zero by free rotations in solution, M0=0. The angular brackets E denote an ensemble average in the presence of an external field, in contrast to denoting the average taken when the field is switched off.

The averages at zero field can be expressed in terms of material liquid properties and interfacial susceptibilities. Since the external fields are always significantly weaker than the internal microscopic fields, perturbation expansion in terms of the interaction energy ME0 between the external field and the sample dipole moment M is always a good approximation. The susceptibility χ0 follows directly from the first-order perturbation theory and is given by the following equation:

χ0=(3Ω0)1βδM0δM.
(44)

Here, δM0=M0M0 and δM=MM are deviations from average dipole moments, which can be assigned zero values in isotropic solutions; β=(kBT)1 is the inverse temperature.

Despite a simple algebraic form given by Eq. (44), the susceptibility χ0 is hard to calculate. The main conceptual and technical difficulty comes from the cross correlation between the solute dipole M0 and the solvent dipole Ms entering M=M0+Ms. This difficulty is overcome in mean-field theories of dielectrics by introducing the concept of the cavity field,3 which is the field produced by the dielectric inside the solute. The cavity-field susceptibility68χc=Ec/E0 is defined as the ratio of the field inside the solute cavity Ec to the external field. A significant simplification of the problem is achieved71 by realizing that the same susceptibility is the ratio of δM0δM in Eq. (44) and the self-correlation of the solute dipole (δM0)2,

χc=δM0δM(δM0)2.
(45)

This relation incorporates all cross correlations between the solute dipole and the dipole moments of the solvent molecules into a single parameter, which can be independently calculated from either the statistical theories of solutions or from numerical simulations.71 These calculations turned out to be critical for understanding dielectrophoresis of proteins: fluctuations of the protein-water interface, related to protein’s elastic flexibility, produce substantial deviations from the solution of the standard dielectric boundary-value problem1 leading to Eq. (40). More specifically, the dielectric solution assumes that thermal agitation can only alter orientations of the solvent dipoles in the interface. In contrast, the protein-water interface is affected by elastic fluctuations of the protein coupled to polarized hydration water. These elastic fluctuations enhance the dipolar solute-solvent correlations in Eq. (45), leading to χc significantly exceeding the dielectric result.71,72

The total ensemble-averaged dipole moment at the solute induced by the external field is the sum of the ensemble-averaged solute dipole M0E and the dipole moment of the interface considered above,

ME=M0E+MintE.
(46)

Here, in contrast to the case of a void shown in Fig. 4, we have added angular brackets to the interface dipole to stress that it is now determined by the combined effect of the external field polarizing the interface and the field of the solute. In analogy with Eq. (43), one can define the susceptibility of the interface as follows:

MintE=χintdΩ0E0,
(47)

where the superscript “d” refers to the dipolar symmetry of the problem.

The great utility of the cavity susceptibility χc is that it not only helps to eliminate many-particle cross correlations from the dipolar response of the solute dipole, but can also be applied to fully characterize the interface susceptibility in Eq. (47). It is given by the following expression:71 

4π3χintd=32(ϵ1)[χcχcL].
(48)

In this equation, χcL is the cavity susceptibility for a specific model of interfacial polarization known as the virtual Lorentz cavity.3 This is an imaginary cavity obtained by separating a closed volume of the liquid from the rest of the bulk. The field created in this separated volume by the bulk polarized by a uniform external field makes the Lorentz cavity field. Since there is no physical interface, there is no physical interfacial polarization and no surface charge density σ. The Lorentz cavity is, therefore, a useful physical limit corresponding to σ=0. The cavity-field susceptibility in the Lorentz limit is

χcL=ϵ+23ϵ.
(49)

It is clear that χcL significantly exceeds χcM in Eq. (40) when ϵ1: χcL/χcM(2/9)ϵ.

Equation (48) shows that the appearance of the interface dipole Mint is linked to the surface charge and is proportional to the deviation of the cavity susceptibility from the Lorentz limit representing σ=0. If one assumes that the rules of Maxwell’s electrostatics apply, one can use χcM from Eq. (40) in Eq. (48) to obtain

χintM=34πϵϵ12ϵ+1.
(50)

This outcome is consistent with Eq. (26): for the solution placed in the plane capacitor, P=(ϵ1)/(4πϵ)E0, and one arrives at Eq. (50) from Eq. (26). We now turn to the question of how these general results for a solute dipole in a polarized polar liquid apply to the problem of solution dielectrophoresis.73 

Dielectrophoresis represents the force acting on a particle in a solution polarized by a nonuniform electric field. The presence of the field gradient, which is required to produce the force acting on a dipole, distinguishes dielectrophoresis from ionic mobility considered so far. A nonzero average dipole M0E is created by the external field considered to be uniform on the size of the solute. As described above, the same external field polarizes the solute-solvent interface inducing the interface dipole MintE. Since both of these dipoles are proportional to the external field, the reversible work of orienting the solute dipole and of polarizing the interface is given by the free energy,1,2 which involves the factor of 1/2,

FDEP=12MEE0.
(51)

Correspondingly, since MEE0, the dielectrophoretic force is proportional to the gradient of the squared field,73–75 

fDEP=3ϵΩ08πKE2.
(52)

Here, it is assumed that the external field is produced by a plane capacitor with the Maxwell field given as E=E0/ϵ. Establishing the connection between E and E0 requires solving the dielectric boundary-value problem for more complex geometries of external field sources. Finally, K in Eq. (52) is the dielectrophoretic susceptibility,

K=4πϵ3[χ0+χintd],
(53)

which includes both the susceptibility χ0 of reorienting the solute dipole [Eq. (43)] and the susceptibility χdint of polarizing the interface [Eq. (47)].

The parameters in Eq. (52) are chosen in such a way that K becomes the Clausius-Mossotti factor when continuum electrostatics is used to calculate the induced dipole moments. When the solute dipole is neglected and χ0=0, K represents polarization of a void in a polar liquid. One obtains from Eqs. (50) and (53)

K=1ϵ1+2ϵ.
(54)

This result corresponds to a negative dielectrophoresis (0.5<K<0) when the particle is repelled from the region of a higher electric field.

The standard applications of the Clausius-Mossotti equation assume that the spherical solute carries its own dielectric constant ϵ0. This limit is easy to obtain from the result for a void by noting that all results of electrostatics are sensitive only to the ratio of two dielectric constants, ϵ/ϵ0, at the dielectric dividing surface.2 Substituting ϵ/ϵ0 in place of ϵ in Eq. (54), one arrives at the commonly used expression,73,74

K=ϵ0ϵϵ0+2ϵ.
(55)

The Clausius-Mossotti factor assumes that the material of the solute is polarizable but does not possess its own permanent dipole moment. In terms of properties of bulk materials, the assumption is that the solute is paraelectric. On the contrary, any material possessing a nonvanishing permanent dipole is characterized as a ferroelectric. When proteins are concerns, each protein molecule can be viewed as a ferroelectric domain, and neglecting χ0 in Eq. (53) is not justified. Dropping the permanent dipole of the solute applies only at sufficiently high frequencies when rotations of the dipole are dynamically frozen, and it does not have sufficient time to respond to an oscillating external field.76 As the solute size increases, one anticipates that the dipole moment M0 scales linearly with the linear dimension of the solute, and χ0M02/Ω0 scales inversely proportional to the linear dimension; e.g., χ0 changes as R01 with the solute radius R0. Since the interface susceptibility χindd in Eq. (53) is constant, it becomes the dominant factor in the dielectrophoretic response for sufficiently large solutes. This limit is, however, not reached for soluble globular proteins.

Proteins typically possess a large density of the surface charge produced by protonation/deprotonation of the water-exposed residues.77 The negative and positive charges mostly compensate each other78 to add to a typically negative overall protein charge at physiological conditions. The nonspherical shape of the molecule and an incomplete compensation between positive and negative charges lead to an overall large dipole moment of several hundreds of debye units when calculated relative to the protein center of mass.78–84 

When both the permanent dipole and interface components are included in the dielectrophoretic susceptibility, one obtains71 

K=χcLϵyp+[χcχcL](ϵyp+3ϵ2(ϵ1)).
(56)

Here, yp=ye+y0 is the effective polarity76,85 of the protein as measured by the combination of the variance of its dipole moment,

y0=(4πβ/9Ω0)(δM0)2,
(57)

and the polarity parameter ye quantifying the density of induced dipoles in the protein molecule; Ω0 is the protein volume. Since yey0 at sufficiently low frequencies, ypy0 is typically a very good approximation.

Large dipole moments of globular proteins78–84 are responsible for large values of the parameter yp in Eq. (56). For instance, molecular dynamics simulations of cytochrome c in TIP3P water72 have produced the protein dipole of M0240 D and yp67 at T=310K. Combined with χc1.6 from simulations, the susceptibility K becomes equal to 8×103. This result points to a dramatic failure of the Clausius-Mossotti equation,86 which limits the dielectrophoretic susceptibility by the condition 0.5<K<1 [Eq. (55)]. Given that yp1 for most proteins at sufficiently low frequencies of the external field, one can write a simplified equation for K,

K=ϵχcyp.
(58)

With the empirical value of the cavity field71,72χc1.3 [see Eq. (64)], one can convert Eq. (58) to a practical equation for the dielectrophoretic susceptibility of a single protein,

K=44ϵT0TM02Ω0.
(59)

Here, T0=300K, M0 is in debyes, and the protein volume Ω0 is in Å3.

Equation (58) gives the dielectrophoretic susceptibility at frequencies below the frequency of protein tumbling (10MHz), which determines rotational relaxation of protein’s permanent dipole moment (β-relaxation in dielectric spectroscopy of proteins78). All functions entering Eq. (56) become dependent on the frequency ω when an oscillatory external field is applied. The difficulty encountered with the direct extension of the static equation (58) to the dynamic domain is that the frequency dependence of χc is mostly unknown. One has to turn to alternative means to access K(ω). This connection is provided by dielectric spectroscopy of solutions discussed next.

Equation (58) suggests that the product ypχc is required to determine K in Eq. (52). The same product enters the dielectric increment Δϵ=ϵmixϵ of a low-concentration (ideal) solution compared to the dielectric constant of the solvent. For ideal solutions, the dielectric increment is a linear function of the volume fraction η0 occupied by the solutes. One can derive the following equation connecting dielectric measurements to dielectrophoresis:72 

K(ω)=Δϵ(ω)η02ϵ(ω)2+19ϵ(ω).
(60)

Here, both the dielectrophoretic susceptibility and the dielectric function are written as functions of the circular frequency ω of the oscillating external field. Correspondingly, Δϵ(ω) is the increment of the frequency-dependent dielectric function of the solution over that of the solvent, ϵ(ω).

The static dielectric constant of the protein solution is typically higher than that of the solvent because of a large dipole moment of the protein (dielectric increment).78,87 According to Eq. (60), the range of frequencies with Δϵ(ω)>0 corresponds to positive dielectrophoresis, K(ω)>0. However, the solution dielectric constant falls below the solvent value at the frequencies exceeding the frequency of overdamped rotational tumbling of the protein. A crossover to negative dielectrophoresis is, therefore, predicted by Eq. (60) for frequencies exceeding the crossover frequency ωc at which Δϵ(ωc)=0 and K(ωc)=0.

Note that the dielectrophoretic force is proportional to the gradient of E2 in the whole range of frequencies, both below ωc and above it. The crossover itself is merely a dynamic effect of dynamic freezing of protein dipole rotations at ω>ωc. A linear scaling with the field appears from the static polarization of the water dipoles in the interface producing Qst. This force does not require a field gradient and is present even for a uniform external field. The overall force acting on the protein is the sum of the linear and quadratic terms in the field,

fE=χcqeffE0+fDEP,
(61)

where qeff is the effective charge of the solute given by Eq. (36).

Figure 8 shows the frequency-dependent dielectrophoretic susceptibility calculated from Eq. (60) by using dielectric functions reported for solutions of the ribonuclease A protein.83 The dielectric function for water ϵ(ω) was taken from Ref. 88. The low-frequency K(ω) turns out to be in the same range of 104 as calculated from molecular dynamics simulations of cytochrome c72 and a number of other proteins.71 The crossover to negative dielectrophoresis at ω>ωc is also clearly observed.

FIG. 8.

Frequency-dependent dielectrophoretic susceptibility K(ν), ν=ω/(2π) calculated from Eq. (60) and dielectric data reported in Ref. 83 for the ribonuclease A protein at concentrations listed in the plot (mM, 298.15 K). ϵ(ω) for water is from Ref. 88 The dotted horizontal line marks K(νc)=0 reached at the crossover frequency νc=ωc/(2π).

FIG. 8.

Frequency-dependent dielectrophoretic susceptibility K(ν), ν=ω/(2π) calculated from Eq. (60) and dielectric data reported in Ref. 83 for the ribonuclease A protein at concentrations listed in the plot (mM, 298.15 K). ϵ(ω) for water is from Ref. 88 The dotted horizontal line marks K(νc)=0 reached at the crossover frequency νc=ωc/(2π).

Close modal

The parameter δ=Δϵ(0)/c, where c is the protein concentration in mg/cm3, is often used to determine the dipole moment of the protein from dielectric spectroscopy.73,78,84,87,89 Dielectric experiment provides access to the root mean square (rms) protein dipole,

Mrms=(δM0)2,
(62)

which becomes72 

Mrms=105Dmδ/χc.
(63)

Here, m is the molar mass of the protein in kilodaltons and the protein dipole is in units of debyes.

If fluctuations of the dipole moment due to elastic motions of the protein are neglected, pure rotations of the permanent dipole produce (δM0)2=M02 and M0=0. The rms dipole Mrms=M0 can be identified with the permanent dipole moment in this case. This is the standard application of Oncley’s formula79 used to estimate protein’s dipole moment from dielectric spectroscopy of solutions.84,90

Equation (63) is essentially identical to Oncley’s formula79 in which χc is replaced by 2b/9 with the empirical parameter b5.8.73,79 This empirical result leads to

χc1.3,
(64)

which turns out to be very close to χc1.21.6 calculated from Eq. (45) applied to molecular dynamics trajectories for hydrated globular proteins.71,72 This value of the cavity susceptibility is substantially above the prediction of Maxwell’s electrostatics χcM1.5ϵ1 [Eq. (40)] and also exceeds the Lorenz result χcL1/3 [Eq. (49)]. Consequently, both empirical and simulation evidence suggest that Maxwell’s cavity field is a poor approximation for this property when applied to hydrated proteins.68 Failure of Maxwell’s electrostatics was also documented for absorption of radiation by protein solutions, where the cavity susceptibility becomes a prominent parameter of the theory.76 The absorption of THz radiation follows the Lorentz cavity equation [Eq. (49)].

The Lorentz cavity field can be brought in a better agreement with χc reported by simulations if another recent calculation is taken into account. From the analysis of simulations, the dielectric constant of hydration shells of protein was found to be very low, ϵ2.5.34 If ϵ in Eq. (49) is replaced with ϵ/ϵ0 to account for polarization of the protein, one obtains

χc=ϵ+2ϵ03ϵ.
(65)

The simulation and empirical values of χc [Eq. (64)] are consistent with this equation at ϵ2.5 and ϵ03.5. One still has to recognize that a large χcχcM reported by simulations is likely related to phenomenology not accounted for by dielectric models. Elastic deformation of the protein coupled with hydration water is likely to be responsible for the cross correlations between the protein and water dipole moments in Eq. (45), leading to the reported values of χc.

The physical principle unifying different phenomena discussed here is that the local interfacial response of a polar liquid is distinct from the macroscopic dipolar response responsible for the long-distance dielectric screening and measured by the dielectric experiment. Studies of ion solvation have shown asymmetry of solvation thermodynamics between anions and cations and, in addition, the presence of spontaneous polarization of the interface around solutes carrying a zero charge. This observation translates to asymmetry in the local interfacial susceptibilities between anions and cations and to the presence of a nonzero surface charge, even in the case of uncharged solutes. Furthermore, mobility of ions is determined by the total effective charge qeff within the shear surface. In the limit of Maxwell’s electrostatics, the total charge becomes the ion charge q since there is no overall bound charge from the solvent within the shear surface. Alternatively, if the polarization response is nonuniform in the interface, an effective charge qeff [Eq. (36)] appears in the electrostatic force acting on the ion. The difference between qeff and q is due to the preferential molecular orientation in the interface (static charge) and different local polar susceptibilities at the cavity and shear surfaces.

Protein dielectrophoresis is another problem where dielectric calculations have been traditionally applied. The dielectrophoretic susceptibility of a large biopolymer is typically calculated from the Clausius-Mossotti equation, which carries two deficiencies: (i) it is based on Maxwell’s cavity susceptibility and (ii) it neglects the permanent dipole of the solute. The first deficiency might be related to elastic flexibility of the solute-solvent interface characteristic of biomolecules, which leads to much stronger cross correlations between the solute and solvent dipoles than anticipated by dielectric theories [Eq. (45)]. The second deficiency is most important since the large dipole moment of the solute (protein) produces orders-of-magnitude higher dielectrophoretic susceptibility than predicted by the Clausius-Mossotti equation. Proteins are, therefore, poised to show “gigantic” dielectrophoresis compared to the standard predictions.

This research was supported by the National Science Foundation (NSF) (No. CHE-1800243). The author is grateful to Mark Hayes for many discussions and for comments on the manuscript.

1.
J. D.
Jackson
,
Classical Electrodynamics
(
Wiley
,
New York
,
1999
).
2.
L. D.
Landau
and
E. M.
Lifshitz
,
Electrodynamics of Continuous Media
(
Pergamon
,
Oxford
,
1984
).
3.
C. J. F.
Bottcher
,
Theory of Electric Polarization, Vol. 1: Dielectrics in Static Fields
(
Elsevier
,
Amsterdam
,
1973
).
4.
R. P.
Feynman
,
R. B.
Leighton
, and
M.
Sands
,
The Feynman Lectures on Physics. Vol II: Mainly Electromagnetism and Matter
(
Addison-Wesley
,
Reading, MA
,
1964
).
5.
M. P.
Allen
and
D. J.
Tildesley
,
Computer Simulation of Liquids
(
Clarendon Press
,
Oxford
,
1996
).
6.
B.
Honig
and
A.
Nicholls
, “
Classical electrostatics in biology and chemistry
,”
Science
268
,
1144
1149
(
1995
).
7.
D. V.
Matyushov
, “
Electrostatics of liquid interfaces
,”
J. Chem. Phys.
140
,
224506
(
2014
).
8.
K.
Barros
and
E.
Luijten
, “
Dielectric effects in the self-assembly of binary colloidal aggregates
,”
Phys. Rev. Lett.
113
,
017801
(
2014
).
9.
M.
Shen
,
H.
Li
, and
M.
Olvera de la Cruz
, “
Surface polarization effects on ion-containing emulsions
,”
Phys. Rev. Lett.
119
,
138002
(
2017
).
10.
R. M.
Lynden-Bell
and
J. C.
Rasaiah
, “
From hydrophobic to hydrophilic behaviour: A simulation study of solvation entropy and free energy of simple solutes
,”
J. Chem. Phys.
107
,
1981
1991
(
1997
).
11.
M.
Born
, “
Volumen und hydratationswärme der ionen
,”
Z. Phys.
1
,
45
48
(
1920
).
12.
A.
Ben-Naim
,
Molecular Theory of Water and Aqueous Solutions: Understanding Water
(
World Scientific Publishing Co. Pte Ltd.
,
2014
).
13.
G.
Hummer
,
L. R.
Pratt
, and
A. E.
Garcia
, “
Molecular theories and simulation of ions and polar molecules in water
,”
J. Phys. Chem. A
102
,
7885
(
1998
).
14.
D. L.
Mobley
,
A. E.
Barber
,
C. J.
Fennell
, and
K. A.
Dill
, “
Charge asymmetries in hydration of polar solutes
,”
J. Phys. Chem. B
112
,
2405
2414
(
2008
).
15.
J. P.
Bardhan
,
P.
Jungwirth
, and
L.
Makowski
, “
Affine-response model of molecular solvation of ions: Accurate predictions of asymmetric charging free energies
,”
J. Chem. Phys.
137
,
124101
(
2012
).
16.
S.
Rajamani
,
T.
Ghosh
, and
S.
Garde
, “
Size dependent ion hydration, its asymmetry, and convergence to macroscopic behavior
,”
J. Chem. Phys.
120
,
4457
4466
(
2004
).
17.
A.
Mukhopadhyay
,
A. T.
Fenley
,
I. S.
Tolokh
, and
A. V.
Onufriev
, “
Charge hydration asymmetry: The basic principle and how to use it to test and improve water models
,”
J. Phys. Chem. B
116
,
9776
9783
(
2012
).
18.
M.
Dinpajooh
and
D. V.
Matyushov
, “
Free energy of ion hydration: Interface susceptibility and scaling with the ion size
,”
J. Chem. Phys.
143
,
044511
(
2015
).
19.
W. M.
Latimer
,
K. S.
Pitzer
, and
C. M.
Slansky
, “
The free energy of hydration of gaseous ions, and the absolute potential of the normal calomel electrode
,”
J. Chem. Phys.
7
,
108
111
(
1939
).
20.
P. G.
Kusalik
and
G. N.
Patey
, “
On the molecular theory of aqueous electrolyte solutions. II. Structural and thermodynamic properties of different models at infinite dilution
,”
J. Chem. Phys.
89
,
5843
5851
(
1988
).
21.
S.
Koneshan
,
J. C.
Rasaiah
,
R. M.
Lynden-Bell
, and
S. H.
Lee
, “
Solvent structure, dynamics, and ion mobility in aqueous solutions at 25°C
,”
J. Phys. Chem. B
102
,
4193
4204
(
1998
).
22.
J.
Frenkel
,
Kinetic Theory of Liquids
(
Dover
,
New York
,
1955
).
23.
L.
Horváth
,
T.
Beu
,
M.
Manghi
, and
J.
Palmeri
, “
The vapor-liquid interface potential of (multi)polar fluids and its influence on ion solvation
,”
J. Chem. Phys.
138
,
154702
(
2013
).
24.
H. S.
Ashbaugh
, “
Convergence of molecular and macroscopic continuum descriptions of ion hydration
,”
J. Phys. Chem. B
104
,
7235
(
2000
).
25.
T. L.
Beck
, “
The influence of water interfacial potentials on ion hydration in bulk water and near interfaces
,”
Chem. Phys. Lett.
561–562
,
1
13
(
2013
).
26.
R. C.
Remsing
and
J. D.
Weeks
, “
Role of local response in ion solvation: Born theory and beyond
,”
J. Phys. Chem. B
120
,
6238
6249
(
2016
).
27.
E.
Harder
and
B.
Roux
, “
On the origin of the electrostatic potential difference at a liquid-vacuum interface
,”
J. Chem. Phys.
129
,
234706
(
2008
).
28.
C. G.
Gray
and
K. E.
Gubbins
,
Theory of Molecular Liquids
(
Clarendon Press
,
Oxford
,
1984
).
29.
C. C.
Doyle
,
Y.
Shi
, and
T. L.
Beck
, “
The importance of the water molecular quadrupole for estimating interfacial potential shifts acting on ions near the liquid–vapor interface
,”
J. Phys. Chem. B
123
,
3348
3358
(
2019
).
30.
M.
Wilson
,
A.
Pohorille
, and
L. R.
Pratt
, “
Surface potential of the water liquid–vapor interface
,”
J. Chem. Phys.
88
,
3281
3285
(
1988
).
31.
M. D.
Baer
,
A. C.
Stern
,
Y.
Levin
,
D. J.
Tobias
, and
C. J.
Mundy
, “
Electrochemical surface potential due to classical point charge models drives anion adsorption to the air–water interface
,”
J. Phys. Chem. Lett.
3
,
1565
1570
(
2012
).
32.
S. M.
Sarhangi
,
M. M.
Waskasi
,
S. M.
Hashemianzadeh
, and
D.
Matyushov
, “
Effective dielectric constant of water at the interface with charged C60 fullerenes
,”
J. Phys. Chem. B
123
,
3135
3143
(
2019
).
33.
M.
Dinpajooh
and
D. V.
Matyushov
, “
Dielectric constant of water in the interface
,”
J. Chem. Phys.
145
,
014504
(
2016
).
34.
S.
Seyedi
and
D. V.
Matyushov
, “
Dipolar susceptibility of protein hydration shells
,”
Chem. Phys. Lett.
713
,
210
214
(
2018
).
35.
S.
Varghese
,
S. K.
Kannam
,
J. S.
Hansen
, and
S. P.
Sathian
, “
Effect of hydrogen bonds on the dielectric properties of interfacial water
,”
Langmuir
35
,
8159
8166
(
2019
).
36.
S.
Mondal
and
B.
Bagchi
, “
Water in carbon nanotubes: Pronounced anisotropy in dielectric dispersion and its microscopic origin
,”
J. Phys. Chem. Lett.
10
,
6287
6292
(
2019
).
37.
O.
Teschke
and
E. F.
de Souza
, “
Water molecule clusters measured at water/air interfaces using atomic force microscopy
,”
Phys. Chem. Chem. Phys.
7
,
3856
3865
(
2005
).
38.
J.-J.
Velasco-Velez
,
C. H.
Wu
,
T. A.
Pascal
,
L. F.
Wan
,
J.
Guo
,
D.
Prendergast
, and
M.
Salmeron
, “
Interfacial water. The structure of interfacial water on gold electrodes studied by x-ray absorption spectroscopy
,”
Science
346
,
831
834
(
2014
).
39.
Y.-C.
Wen
,
S.
Zha
,
X.
Liu
,
S.
Yang
,
P.
Guo
,
G.
Shi
,
H.
Fang
,
Y. R.
Shen
, and
C.
Tian
, “
Unveiling microscopic structures of charged water interfaces by surface-specific vibrational spectroscopy
,”
Phys. Rev. Lett.
116
,
016101
(
2016
).
40.
L.
Fumagalli
,
A.
Esfandiar
,
R.
Fabregas
,
S.
Hu
,
P.
Ares
,
A.
Janardanan
,
Q.
Yang
,
B.
Radha
,
T.
Taniguchi
,
K.
Watanabe
,
G.
Gomila
,
K. S.
Novoselov
, and
A. K.
Geim
, “
Anomalously low dielectric constant of confined water
,”
Science
360
,
1339
1342
(
2018
).
41.
P.
Debye
and
L.
Pauling
, “
The inter-ionic attraction theory of ionized solutes. IV. The influence of variation of dielectric constant on the limiting law for small concentrations
,”
J. Am. Chem. Soc.
47
,
2129
2134
(
1925
).
42.
B. E.
Conway
,
J. O.
Bockris
, and
I. A.
Ammar
, “
The dielectric constant of the solution in the diffuse and Helmholtz double layers at a charged interface in aqueous solution
,”
Trans. Farad. Soc.
47
,
756
766
(
1951
).
43.
L. S.
Palmer
,
A.
Cunliffe
, and
J. M.
Hough
, “
Dielectric constant of water films
,”
Nature
170
,
796
796
(
1952
).
44.
P. J.
Stiles
,
J. B.
Hubbard
, and
R. F.
Kayser
, “
Dielectric saturation and dielectric friction in electrolyte solutions
,”
J. Chem. Phys.
77
,
6189
6196
(
1982
).
45.
M. A.
Wilson
,
A.
Pohorille
, and
L. R.
Pratt
, “
Comment on “Study on the liquid–vapor interface of water. I. Simulation results of thermodynamic properties and orientational structure”
,”
J. Chem. Phys.
90
,
5211
5213
(
1989
).
46.
M.
Dinpajooh
and
D. V.
Matyushov
, “
Mobility of nanometer-size solutes in water driven by electric field
,”
Physica A
463
,
366
375
(
2016
).
47.
D. S.
Cerutti
,
N. A.
Baker
, and
J. A.
McCammon
, “
Solvent reaction field potential inside an uncharged globular protein: A bridge between implicit and explicit solvent models?
,”
J. Chem. Phys.
127
,
155101
(
2007
).
48.
W. L.
Jorgensen
,
J.
Chandrasekhar
,
J. D.
Madura
,
R. W.
Impey
, and
M. L.
Klein
, “
Comparison of simple potential functions for simulating liquid water
,”
J. Chem. Phys.
79
,
926
935
(
1983
).
49.
J. K.
Beattie
,
A. M.
Djerdjev
, and
G. G.
Warr
, “
The surface of neat water is basic
,”
Faraday Discuss.
141
,
31
39
(
2009
).
50.
J.-S.
Samson
,
R.
Scheu
,
N.
Smolentsev
,
S. W.
Rick
, and
S.
Roke
, “
Sum frequency spectroscopy of the hydrophobic nanodroplet/water interface: Absence of hydroxyl ion and dangling OH bond signatures
,”
Chem. Phys. Lett.
615
,
124
131
(
2014
).
51.
X.
Yan
,
M.
Delgado
,
J.
Aubry
,
O.
Gribelin
,
A.
Stocco
,
F.
Boisson-Da Cruz
,
J.
Bernard
, and
F.
Ganachaud
, “
Central role of bicarbonate anions in charging water/hydrophobic interfaces
,”
J. Phys. Chem. Lett.
9
,
96
103
(
2017
).
52.
S. E.
Sanders
,
H.
Vanselous
, and
P. B.
Petersen
, “
Water at surfaces with tunable surface chemistries
,”
J. Phys. Condens. Matter
30
,
113001
(
2018
).
53.
J. G.
Davis
,
B. M.
Rankin
,
K. P.
Gierszal
, and
D.
Ben-Amotz
, “
On the cooperative formation of non-hydrogen-bonded water at molecular hydrophobic interfaces
,”
Nat. Chem.
5
,
796
802
(
2013
).
54.
M.
Dinpajooh
and
D. V.
Matyushov
, “
Interfacial structural transition in hydration shells of a polarizable solute
,”
Phys. Rev. Lett.
114
,
207801
(
2015
).
55.
V.
Ballenegger
and
J.-P.
Hansen
, “
Dielectric permittivity profiles of confined polar liquids
,”
J. Chem. Phys.
122
,
114711
(
2005
).
56.
A.
Einstein
,
Investigations on the Theory of the Brownian Movement
(
BN Publishing
,
2011
).
57.
P.
Wolynes
, “
Dynamics of electrolyte solutions
,”
Annu. Rev. Phys. Chem.
31
,
345
376
(
1980
).
58.
R. L.
Kay
, “
The current state of our understanding of ionic mobilities
,”
Pure Appl. Chem.
63
,
1393
1399
(
1991
).
59.
L.
Onsager
, “
Electric moments of molecules in liquids
,”
J. Am. Chem. Soc.
58
,
1486
1493
(
1936
).
60.
J. C.
Rasaiah
and
R. M.
Lynden-Bell
, “
Computer simulation studies of the structure and dynamics of ions and non-polar solutes in water
,”
Philos. Trans. R. Soc. A
359
,
1545
1574
(
2001
).
61.
B.
Bagchi
and
R.
Biswas
, “
Ionic mobility and ultrafast solvation: Control of a slow phenomenon by fast dynamics
,”
Acc. Chem. Res.
31
,
181
187
(
1998
).
62.
R.
Zwanzig
, “
Dielectric friction on a moving ion
,”
J. Chem. Phys.
38
,
1603
1605
(
1963
).
63.
J.
Hubbard
and
L.
Onsager
, “
Dielectric dispersion and dielectric friction in electrolyte solutions. I
,”
J. Chem. Phys.
67
,
4850
4857
(
1977
).
64.
P.
Kumar
and
S.
Yashonath
, “
Ionic conductivity in aqueous electrolyte solutions: Insights from computer simulations
,”
J. Mol. Liq.
277
,
506
515
(
2019
).
65.
P. K.
Ghorai
and
S.
Yashonath
, “
The Stokes–Einstein relationship and the levitation effect: Size-dependent diffusion maximum in dense fluids and close-packed disordered solids
,”
J. Phys. Chem. B
109
,
5824
5835
(
2005
).
66.
M.
Druchok
and
M.
Holovko
, “
Ion hydration under pressure: A molecular dynamics study
,”
Z. Naturforsch.
68a
,
112
122
(
2013
).
67.
D. R.
Martin
and
D. V.
Matyushov
, “
Microscopic fields in liquid dielectrics
,”
J. Chem. Phys.
129
,
174508
(
2008
).
68.
D. R.
Martin
,
A. D.
Friesen
, and
D. V.
Matyushov
, “
Electric field inside a “Rossky cavity” in uniformly polarized water
,”
J. Chem. Phys.
135
,
084514
(
2011
).
69.
C.-Y.
Li
,
J.-B.
Le
,
Y.-H.
Wang
,
S.
Chen
,
Z.-L.
Yang
,
J.-F.
Li
,
J.
Cheng
, and
Z.-Q.
Tian
, “
In situ probing electrified interfacial water structures at atomically flat surfaces
,”
Nat. Mater.
18
,
697
701
(
2019
).
70.
S.
Wall
, “
The history of electrokinetic phenomena
,”
Curr. Opin. Colloid Interface Sci.
15
,
119
124
(
2010
).
71.
D. V.
Matyushov
, “
Dipolar response of hydrated proteins
,”
J. Chem. Phys.
136
,
085102
(
2012
).
72.
S. S.
Seyedi
and
D. V.
Matyushov
, “
Protein dielectrophoresis in solution
,”
J. Phys. Chem. B
122
,
9119
9127
(
2018
).
73.
R.
Pethig
,
Dielectrophoresis: Theory, Methodology and Biological Applications
(
Wiley
,
2017
).
74.
T. B.
Jones
,
Electromechanics of Particles
(
Cambridge University Press
,
Cambridge
,
1995
).
75.
R.
Pethig
, “
Review article—Dielectrophoresis: Status of the theory, technology, and applications
,”
Biomicrofluidics
4
,
022811
(
2010
).
76.
D. R.
Martin
and
D. V.
Matyushov
, “
Terahertz absorption of lysozyme in solution
,”
J. Chem. Phys.
146
,
084502
(
2017
).
77.
D. J.
Barlow
and
J. M.
Thornton
, “
Charge distribution in proteins
,”
Biopolymers
25
,
1717
1733
(
1986
).
78.
E. H.
Grant
,
R. J.
Sheppard
, and
G. P.
South
,
Dielectric Behaviour of Biological Molecules in Solution
(
Clarendon Press
,
Oxford
,
1978
).
79.
J. L.
Oncley
, “
The investigation of proteins by dielectric measurements
,”
Chem. Rev.
30
,
433
450
(
1942
).
80.
R.
Pethig
, “
Protein-water interactions determined by dielectric methods
,”
Annu. Rev. Phys. Chem.
43
,
177
205
(
1992
).
81.
J.
Antosiewicz
, “
Computation of the dipole moments of proteins
,”
Biophys. J.
69
,
1344
1354
(
1995
).
82.
S.
Takashima
, “
Electric dipole moment of globular proteins: Measurement and calculation with NMR and X-ray databases
,”
J. Non-Cryst. Solids
305
,
303
310
(
2002
).
83.
A.
Oleinikova
,
P.
Sasisanker
, and
H.
Weingärtner
, “
What can really be learned from dielectric spectroscopy of protein solutions? A case study of ribonuclease A
,”
J. Phys. Chem. B
108
,
8467
(
2004
).
84.
B. L.
Mellor
,
E.
Cruz Cortés
,
D. D.
Busath
, and
B. A.
Mazzeo
, “
Method for estimating the internal permittivity of proteins using dielectric spectroscopy
,”
J. Phys. Chem. B
115
,
2205
2213
(
2011
).
85.
G.
Stell
,
G. N.
Patey
, and
J. S.
Høye
, “Dielectric constants of fluid models: Statistical mechanical theory and its quantitative implementation,” in Advances in Chemical Physics, edited by I. Prigogine and S. A. Rice (Wiley, 1981), Vol. 48, pp. 183–328.
86.
R.
Pethig
, “
Limitations of the Clausius-Mossotti function used in dielectrophoresis and electrical impedance studies of biomacromolecules
,”
Electrophoresis
40
,
2575
2583
(
2019
).
87.
G. P.
South
and
E. H.
Grant
, “
Dielectric dispersion and dipole moment of myoglobin in water
,”
Proc. R. Soc. A
328
,
371
387
(
1972
).
88.
N. Q.
Vinh
,
M. S.
Sherwin
,
S. J.
Allen
,
D. K.
George
,
A. J.
Rahmani
, and
K. W.
Plaxco
, “
High-precision gigahertz-to-terahertz spectroscopy of aqueous salt solutions as a probe of the femtosecond-to-picosecond dynamics of liquid water
,”
J. Chem. Phys.
142
,
164502
(
2015
).
89.
Y. D.
Feldman
and
V. D.
Fedotov
, “
Dielectric relaxation, rotational diffusion and the heat denaturation transition in aqueous solutions of RNAse A
,”
Chem. Phys. Lett.
143
,
309
312
(
1988
).
90.
S.
Takashima
and
K.
Asami
, “
Calculation and measurement of the dipole moment of small proteins: Use of protein data base
,”
Biopolymers
33
,
59
68
(
1993
).