Surface excess is a quantitative measure of the amount of a substance that accumulates at the interface, in excess of what would be expected based on its bulk concentration
For a solute that accumulates at the surface, the surface excess is positive
For a solute that avoids the surface, the surface excess is negative
We can write something similar for Gibbs energy as well
We shall define the surface excess per unit area as Γiσ, the surface concentration
dγ=−i∑ΓidμiGibbs Isotherm
This equation can be interpreted as follows
If adding more of solute i (raising its chemical potential) decreases the surface tension, then Γi is positive and solute i will accumulate at the interface
If adding more of solute i (raising its chemical potential) increases the surface tension, then Γi is negative and solute i will avoid the interface
Similarly, we can express Γi for a solute as a rate of change of surface tension with respect to its chemical potential
Γsolute=(∂μsolute∂γ)
We also refer to Γsolute as the excess adsorption of a solute
The interface between two phase is usually not a surface with zero thickness, in the sense that the surface is composed of one layer of molecules from each phases only
The density profile is a measure of the distribution of molecules across the interfacial layer between two phases
It provides information on how the density of the molecules changes as a function of distance from the interface
The system under consideration can consist of either a liquid and a vapor phase, or two liquid phases
We define the location of the interface using the Gibbs Dividing Surface
This conceptual boundary is chosen to be at a location such that the excess adsorption of a solvent at the interface, denoted as Γsolvent=0
The mass balance condition can be defined mathematically as
∫−∞zG(ρ(z)−ρα)dz+∫zG∞(ρ(z)−ρβ)dz=0⎩⎪⎪⎪⎪⎨⎪⎪⎪⎪⎧ρ(z)ραρβzGdensity of the system at position zbulk density of phase αbulk density of phase β position of the Gibbs dividing surface
The equation says that the excess mass in the region where the density transitions from ρα to ρβ is zero
In other words, the Gibbs dividing surface is placed such that the total "extra" mass on one side of the surface cancels out the "missing" mass on the other side.
To find the precise location for this dividing surface, we use
zG=ρα−ρβ1∫−∞∞zdzdρdz
dzdρ is the gradient of the density across the interface, describing how the density changes with position.
The integral ∫−∞∞zdzdρdz gives a weighted average of the position z, where the weight is the density gradient. This effectively gives the center of mass of the density transition across the interface.
The factor ρα−ρβ1 normalizes this weighted average with respect to the difference in bulk densities of the two phases.
Thermal fluctuations at interfaces cause small undulations, or "waves," that modify the interface's shape and thickness
These fluctuations are particularly important in systems where surface tension and temperature play major roles in determining the interface's stability
Capillary Wave Theory (CWT) is a powerful framework used to describe these undulations and predict the behavior of fluctuating interfaces.
The term "capillary waves" refers to undulations on the surface of a fluid that arise due to thermal fluctuations
The wavelength λ of a capillary wave can vary from molecular scales (nanometers) to millimeters, with larger wavelengths corresponding to lower-energy fluctuations.
The wavevector is given by q=\frac{2\pi}
The position of the interface at any horizontal position x is denoted as h(x)
While a perfectly flat interface might have a sharp boundary, fluctuations smear out this boundary, leading to a finite interfacial thickness.
The thickness is given by the root mean square :⟨h(x)2⟩
Spectrum of Capillary Waves
A key feature of CWT is that it predicts a continuous spectrum of waves, each contributing to the overall undulation of the interface.
The energy associated with each wave depends on its wavelength λ and surface tension γ.
Eq∝γλ2kBT
Energy of a fluctuation decreases as the surface tension increases, and also decreases for longer wavelengths.
Free Energy and Capillary Waves
The free energy of an interface is minimized when it remains flat.
Thermal fluctuations increase the free energy by introducing small deviations from the planar state
For small perturbations from a flat interface, the Hamiltonian (energy function) for capillary waves is given by:
One of the most important results of CWT is that interface thickness increases with the system size
⟨∣h(x,y)∣2⟩=2πγkBTln(ξL)
⟨∣h(x,y)∣2⟩ is the mean-square thickness
When the its value is large,it indicates greater undulations and a thicker, more diffuse interface.
2πγkBT relates the magnitude of the thermal fluctuations to both temperature and surface tension.
Higher temperatures increase the fluctuations because more thermal energy is available to disturb the interface.
Higher surface tension dampens the fluctuations, as it acts to resist changes in surface area and restore the interface to its equilibrium state.
The logarithmic factor shows that the mean-square displacement grows as the system size L increases relative to the microscopic cutoff length ξ
L is the characteristic length scale of the system (e.g. the lateral size of the interface). The longer the lateral size, the greater the contribution of long-wavelength capillary waves to the overall fluctuations. In large systems, more fluctuations across different scales can occur, leading to a thicker interface.
ξ is a microscopic cutoff length, often taken as the molecular size or correlation length. It is needed to avoid the equation predicting unphysical behavior at very small scales (L<ξ). This cutoff represents a lower limit on the length scales over which continuum theories (like Capillary Wave Theory) are valid.
Gravitational Cut-off for Capillary Waves
While the interface thickness theoretically increases without bound with increasing system size, gravity imposes an upper limit on the amplitude of the fluctuations
This upper limit is governed by the capillary length λc, which is the length scale beyond which fluctuations are damped by gravitational forces
λc=gΔργ
λc (or κ−1) defines the largest wavelength of capillary waves that can be supported by the system before gravity starts to significantly damp them.
Δρ is the density difference between the two phases
Given that gravity imposes an upper limit, we must then modify our previous equation to
⟨∣h(x,y)∣2⟩=2πγkBTln(ξκ−1)
Thus, the largest fluctuations that can exist are of the order of κ−1, preventing unbounded growth of the interface thickness.
This equation accounts for the finite thickness of the interface, limited by gravity and molecular interactions.
The first term, 2κ∫(2H−H0)2dA, describes the energy cost due to deviations from the membrane's intrinsic curvature the membrane adopts without any external forces (H0).
The second term, κˉ∫KdA, represents the Gaussian curvature energy, which is concerned with the membrane's topological deformations (e.g. how the membrane bends and twists).
Symmetric Membrane and Fourier Transform
The bending energy of a symmetric membrane (with no spontaneous curvature, (H0=0) is given by
However, solving this equation directly is complex because of the spatial variations in h(x,y). To simplify the problem, we express the membrane's thickness function h(x,y) in Fourier space.
h(q)=∫h(x,y)eiq⋅(x,y)dxdy
In Fourier space, the Laplacian ∇2h(x,y) becomes multiplication by −q2, meaning that:
Fc=21∫∫κ(∇2h)2dxdy=2L2κq∑q4∣h(q)∣2
The bending energy depends on the amplitude of the fluctuations h(q) and increases rapidly with the wave vector q (because of the q4 term)
Short-wavelength fluctuations (large q) are much more costly in terms of energy than long-wavelength fluctuations (small q).
Thermal Fluctuations and Equipartition of Energy
Thermal energy causes the membrane to fluctuate, and each wave mode receives a share of the thermal energy due to the equipartition theorem
According to this principle, each mode contributes:
⟨∣h(q)∣2⟩=L2κq4kBT
The amplitude of the fluctuations decreases with q4, meaning that large wave vectors (small-wavelength fluctuations) are suppressed by the bending energy
Larger wavelengths (small q) dominate the fluctuations because they incur a smaller energy cost.
To compute the total mean-square thickness ⟨h2⟩ of the membrane, we sum (or integrate) the contributions from all wave modes. This can be done by integrating the mean-square amplitude ⟨∣h(q)∣2⟩ over all possible wave vectors:
The interface potential V(ℓ) is a function of the wetting layer thickness ℓ, where ℓ is the thickness of the liquid film on the solid substrate
The interface potential represents the free energy per unit area required to create and sustain a wetting layer of thickness ℓ.
In wetting theory, the wetting layer thickness ℓ(x) describes how a liquid spreads over a solid surface
The film starts with a thin layer, referred to as the precursor film, which has a thickness denoted by the adsorption thickness ( ℓa ). The precursor film spreads out uniformly across the surface and is much thinner than the height of the droplet.
As we move along the surface, ℓ(x) increases until it reaches the full droplet thickness.
The interface Potential depends on the relative size of ℓ and ℓa
For small deviations of ℓ from ℓa (i.e. when ℓ≈ℓa), the interface potential can be approximated as:
V(ℓ)≈γVS(1+(ℓ−ℓa)2)⎩⎪⎪⎨⎪⎪⎧γVSℓℓasurface tension between the solid and vapor current thickness of the wetting layeradsorption thickness (equilibrium value of ℓ )
For much larger wetting layers (ℓ≫ℓa), the interface potential is influenced by several factors, and the approximation becomes:
V(ℓ)≈VVDW(ℓ)+VCOUL(ℓ)+VCOEX(ℓ)+γLS+γLV⎩⎪⎪⎪⎪⎪⎪⎪⎨⎪⎪⎪⎪⎪⎪⎪⎧VVDW(ℓ)VCOUL(ℓ)VCOEX(ℓ)γLSγLVcontribution from van der Waals forcescontribution from Coulombic forcesshift from coexistence conditionssurface tension between the liquid and solidsurface tension between the liquid and vapor
This more general expression accounts for various types of forces acting on the wetting layer as its thickness increases
Contributions to the Interface Potential
Van der Waals forces act between the liquid and the substrate. The form of VVDW(ℓ) depends on whether the forces are non-retarded (dominant at smaller distances) or retarded (dominant at larger distances):
Non-retarded van der Waals forces: These are dominant when the wetting layer thickness is small ( ℓ≈10nm ) and can be approximated as:
VVDWnon-retarded(ℓ)≈ℓ2A⋯⋯A is a constant that characterizes the strength of force at shorter distance
Retarded van der Waals forces: For larger thicknesses ( ℓ>50nm ), retardation effects become important, and the potential decays more rapidly:
VVDWretarded(ℓ)≈ℓ3B⋯⋯B is a constant that characterizes the strength of the force at longer distance.
The transition from ℓ2A to ℓ3B occurs due to retardation effects, which weaken the interaction over larger distances as the speed of light influences how these forces propagate.
In systems where electrostatic interactions are present (e.g. ionic liquids, solutions with surface dissociation), Coulombic forces contribute to the interface potential.
These forces are screened by the presence of dissolved ions, and the screening length ( ℓc ) determines how far the electrostatic interactions extend.
The Coulombic potential is given by:
VCOUL(ℓ)≈σ0exp(ℓc−ℓ)⎩⎪⎪⎨⎪⎪⎧σ0ℓℓcsurface charge densitythickness of the wetting layerscreening length
Electrostatic forces decay exponentially as the distance ℓ increases, with the characteristic length scale ℓc determined by the screening effects (e.g., ionic strength in the solution).
The equilibrium wetting layer thickness ( lw ) corresponds to the point where the system finds mechanical equilibrium for the wetting layer ( i.e. V(ℓ) is minimized at that point )
dldV(ℓ)∣∣∣∣∣∣ℓw=0
Since V(ℓ)=ℓ2A+Δμℓ+γLS+γLV, performing the differentiation will give us
lw=(Δμ2A)1/3
As Δμ→0, the wetting layer thickness ℓw→∞, meaning that the system transitions into a regime of complete wetting where the liquid forms a continuous film on the surface.
The spreading coefficient S quantifies the driving force for the liquid to spread across the surface
S=ℓw2A+Δμℓw
If S>0, the liquid completely wets the surface.
If S<0, the liquid forms droplets on the surface (partial wetting).
S→0 as Δμ→0 and ℓw→∞
Moreover, wetting transition occurs when γVS=V(ℓ)
At this point, the interfacial potential equals the original solid-vapor surface tension, meaning the solid is effectively "replaced" by the liquid, leading to the formation of a thick liquid film on the surface.
The Interfacial Phase Diagram
However, ℓw does not increase smoothly as the temperature drop, indicating the existence of phase transitions
Complete wetting occurs when the liquid fully coats the surface, and the thickness of the film becomes large, often referred to as ℓw→∞
Prewetting is a first-order phase transition (involves a discontinuous jump in the film thickness) that can occur when the system forms a thin liquid film on the surface, even below the wetting transition temperature
The interfacial phase diagram is a tool used to describe the behavior of surfaces at different temperatures and compositions.
Prewetting Line: Marks the temperature and chemical potential where a thin film starts to form on the surface. Below this line, the surface is essentially "dry."
Wetting Transition Line: The temperature where the liquid completely wets the surface, forming a thick or infinite film.
Surface Critical Point: Where the distinction between phases at the surface vanishes, leading to critical fluctuations in the wetting film.
In the wetting of surfaces, the balance of forces at the three-phase contact line determines the shape and stability of droplets or thin films. These forces include:
Surface Tension (γ): The force acting along the surface of a liquid, which minimizes the surface area.
Line Tension (τ): The force acting along the three-phase contact line, contributing to the wetting behavior at very small scales.
Free Energy and Line Tension
The total free energy F of the system can be written in differential form, incorporating contributions from temperature (T), volume (V), chemical potential (μ), surface area (A), and the length of the three-phase contact line (L):
dF=−SdT−PdV+i∑μidNi+γdA+τdL
τdL represents the contribution of line tension to the system's free energy, where τ is the force per unit length along the three-phase contact line.
τ=(∂L∂F)Ni,T,V,A
Line tension length ( ξ ) is a characteristic length scale that relates the line tension ( τ ) to the surface tension ( γ ). It provides a measure of how significant line tension effects are compared to surface tension:
ξ=∣∣∣∣∣γτ∣∣∣∣∣with units of length (m)
ξ has units of meters (m), and it represents the distance over which line tension significantly affects the wetting behavior.
When L≤ξ, line tension effects are important.
For larger droplets, where L≫ξ, the influence of line tension becomes negligible, and surface tension dominates.
The theoretical value of line tension ( τtheor ) can be estimated using thermal energy ( kBT ) and a characteristic molecular size ( d )
Near Wetting Transitions
Near wetting transitions (when θ→0 and S→0), the influence of line tension becomes particularly significant
The characteristic length scale ( ξS ) associated with line tension near wetting transitions is given by the ratio of the line tension and spreading coefficient:
ξS=∣∣∣∣Sτ∣∣∣∣asθ→0
As θ→0, the ξS approaches the micrometer scale.
This suggests that as the contact angle decreases (near the point of complete wetting), line tension effects can become important over larger length scales
Contact angle dependence with line tension
The free energy change of the system is given by:
ΔF=−Sπr2+γVLπh2+τ2πr−γVLA
S=γVS−(γLS+γLV)
h is the height of the droplet
r is the radius of the droplet's base (contact line radius).
The Young-Laplace equation describes how the contact angle is influenced by surface tensions. However, when line tension is considered, a modified form is used:
The behavior of particles at interfaces depends on their size and the interfacial forces. In particular, the criterion for deformation is given by
R≥κ−1whereκ−1=gΔργ⎩⎪⎨⎪⎧γgΔρSurface tension of the liquidGravitational accelerationDensity difference
Particles with a radius R≥κ−1 deform the interface significantly.
The Bond number ( B ) determines if deformation can be ignored:
B=γΔρgR2≪1
If B≪1, deformation effects are negligible, especially for particles smaller than 1μm.
The adsorption behavior of particles at interfaces is size-dependent. The probability of a particle residing at a given height H is expressed as:
p∼exp(−3kBT4πR3ΔρgH)
Larger particles experience stronger gravitational forces and have lower adsorption probabilities at higher heights due to increased gravitational potential energy.
The interplay of gravitational, thermal, and interfacial forces determines droplets' equilibrium position and distribution.
As particle size increases, gravitational forces acting on the particle grow stronger. This causes larger particles to settle downward, while smaller particles are less affected by gravity.
Thermal energy counteracts gravitational forces, keeping smaller particles more evenly distributed across the interface.
The balance between these two forces sets the particle height ( H ) according to the equation:
gravityΔρgVH=kBTthermal energy
Thus, for very small particles (R<Rgrav≈20nm), thermal fluctuations dominate, preventing significant settling.
Thermodynamics
The total free energy for a particle at an interface includes contributions from interfacial tensions and line tension:
ΔF=(γc2−γc1)Ac2−γ12A12c+τL
γ12: Interfacial tension between phases 1 and 2.
γc1,γc2: Interfacial tensions between the particle and phases 1 and 2, respectively.
Ac2: Area of the particle in contact with phase 2.
A12c: Area of the liquid-liquid interface replaced by the particle.
τ: Line tension at the three-phase contact line.
L: Length of the three-phase contact line.
The stability of a particle is determined by the derivative of the free energy ΔF with respect to the contact angle θ. For the particle to remain stable, the free energy must exhibit a local minimum.
The disappearance of the local minimum occurs when the second derivative of the free energy:
dθ2d2ΔF∣∣∣∣θmin=0
Performing the derivative gives us:
cosθmin=(cosθ)1/3
This expression defines the critical contact angle ( θmin ) at which stability is lost.
The particle is stable at the interface only if its radius R exceeds a critical value Rmin, given by:
Rmin=γVLsinθmin(1−cosθmincosθ)τ
Smaller particles ( R<Rmin ) are unstable due to the dominating effects of line tension.
Larger particles ( R>Rmin ) achieve equilibrium as the energy landscape retains a local minimum.
Deformations depend on both aspect ratio (α) and the contact angle, where larger α have stronger quadrupolar effects and smaller α have weaker deformations.
Electrochemical interfaces involve the interaction between a conductive electrode and an electrolyte, where the redistribution of charges at the interface results in phenomena such as capacitance, adsorption, and electrochemical reactions.
Capacitance at these interfaces plays a central role in describing charge storage and dynamic response.
The electrode is modeled as an equivalent circuit consisting of:
A capacitance ( C ), representing charge storage at the interface.
A resistance ( Rr ), representing Faradaic processes or charge transfer resistance.
The total impedance ( Z ) of the system is given by:
Z=R+iωC+Rr11⎩⎪⎪⎨⎪⎪⎧RCRrωSeries ResistanceInterfacial capacitanceResistance for charge transfer (Faradaic processes)Angular frequency of the applied signal
In order for the measurement to be accurate, the capacitive contribution dominates:
ωC≫Rr1
Differential and Integral Capacitance
Capacitance at electrochemical interfaces can be described in terms of differential and integral capacitance:
Measures the rate of change of charge with respect to the applied potential
Differential capacitance is typically determined experimentally by measuring the charge response to small changes in potential.
Integral Capacitance ( K ):
K=ϕ−ϕσ=0σ⎩⎪⎨⎪⎧σϕϕσ=0Total ChargeApplied PotentialPotential at zero charge
Integral capacitance is typically calculated and provides a broader measure of charge storage over a potential range, integrating both adsorption and structural effects.
The relationship between C and K captures how the local capacitance changes with potential:
C=K+(ϕ−ϕσ=0)dϕdK
The term K represents the contribution from total stored charge, while (ϕ−ϕσ=0)dϕdK accounts for variations in storage behavior as potential changes.
When K changes slowly with potential, C≈K. But rapid variations in K lead to significant differences between C and K.
Electrocapillary Phenomena
Electrocapillary phenomena describe how surface tension ( γ ) at an electrode-electrolyte interface is influenced by surface charge ( σ ) and applied potential ( ϕ ).
The basic equation of electrocapillarity links changes in surface tension to surface charge and the chemical potential ( μi ) of adsorbing species:
dγ=−σdϕ−i∑Γidμi
If the solution composition remains constant ( dμi=0 ), the equation simplifies to the Lippmann Equation:
dγ=−σdϕorσ=−dϕdγ
This provides a direct relationship between surface charge and the slope of the electrocapillary curve:
The electrocapillary curve exhibits the following features:
Parabolic Shape: The curve peaks at the potential of zero charge ( ϕσ=0 ), where σ=0 and γ=γ0.
Charge Dependence:
For ϕ>ϕσ=0: Positive charge accumulates (Q>0).
For ϕ<ϕσ=0: Negative charge dominates (Q<0).
Capacitance can also be derived from the second derivative of surface tension with respect to potential:
C=−dϕ2d2γ
Differential capacitance quantifies the local response of surface tension to changes in potential.
Peaks in C occur at points of rapid surface charge redistribution.
The surface tension of a stationary liquid drop is influenced by gravitational and geometric factors.
γ=2ρgh2⋅1.64R+h1.64R⎩⎪⎪⎨⎪⎪⎧ρghRliquid densitygravitational accelerationheight of the liquid columnradius of curvature of the drop
The instantaneous current is proportional to the surface charge density and the rate of surface expansion ( dtdS ):
Iϕ=σdtdS
The averaged current over a drop period τ is:
Iˉ=τ1∫0τIϕdτ=τσ∫0τdS=τσSmax
Smax: Maximum surface area during drop formation.
Helmholtz Layer and Capacitance
The Helmholtz layer describes the compact, immobile layer of ions at an electrode-electrolyte interface
The electric field ( E ) across the Helmholtz layer is proportional to the surface charge density ( $sigma$ ):
E=ϵ4πσ
The potential drop ( ϕ0 ) across the Helmholtz layer is related to the field and the thickness of the layer ( d ):
ϕ0=Ed=ϵ4πσd
From this relationship, the surface charge density can be expressed as:
σ=4πdϵϕ0
Surface tension at the interface varies with potential (ϕ) according to the Lippmann equation:
σ=−dϕdγ
Integrating this relationship gives the total surface tension as:
γ=γ0−∫0ϕ0σdϕ=γ0−8πdϵϕ02
This leads to the inverted parabolic shape of the electrocapillary curve, where surface tension reaches a maximum at the potential of zero charge (ϕσ=0).
The differential capacitance of the Helmholtz layer, which quantifies its ability to store charge, is derived from the relationship between charge and potential:
C=dϕ0dσ
Substituting (σ=4πdϵϕ0) into this expression will give us the capacitance of the Helmholtz layer!!!
C=K=4πdϵ
Base on this relationship, we can understand a few important things
Charge Separation: The Helmholtz layer acts like a parallel plate capacitor, with ions on the electrolyte side and electrons/holes on the electrode side.
Electrocapillary Behavior: The parabolic dependence of surface tension on potential highlights the balance between charge density and potential drop.
Capacitance: Helmholtz capacitance is independent of the ionic concentration in the solution, as it only depends on ϵ and d.
The Gouy-Chapman theory provides allows us to understand the diffuse electrical double layer formed at the interface of a charged electrode and electrolyte.
Fundamental Assumptions
The theory is built upon the Poisson-Boltzmann approximation and assumes:
The energy of an ion is determined only by the electric field and screened by the solvent.
No fluctuations or chemical interactions between ions and the solvent.
The potential profile across the interface is governed by the Poisson Equation:
dx2d2ϕ=−ϵ4πρ,ρ=ec0i∑ziνiexp(−kBTWi)⎩⎪⎨⎪⎧ϕρWielectric potentialcharge density=eziϕ ; Work to move ion i to interface
The potential gradient at the interface:
(dxdϕ)2=ϵ8πkBTc0[i∑νiexp(−kBTeziϕ)−1]
Surface Charge Density
The surface charge density σ is given by:
σ=σ∗(c0)i∑νi[exp(−kBTeziϕ0)−1]
σ∗(c0) is a scaling factor related to ion concentration and Debye length and is given by
σ∗(c0)=e2πLBc0=(2π123)eLBrs31
LB is the Bjerrum Length represents the scale of electrostatic interactions
LB=ϵkBTe2
For a 1:1 electrolyte:
σ=2σ∗(c0)sinh(2kBTeϕ0)
Capacitance of the Double Layer
The capacitance ( C ) of the electrochemical interface is determined by the effective screening length ( λG )
C=4πλGϵ
λG is the modified Debye length, which accounts for the influence of surface potential ( ϕ0 ) and is defined as:
λG=cosh(2kBTeϕ0)λD
λD is the normy Debye length, representing the intrinsic screening length and relates to the Bjerrum Length (LB) and the bulk ionic concentration ( c0 )
λD=8πLBc01
Using the hyperbolic cosine identity cosh(2kBTeϕ0)=1+sinh2(2kBTeϕ0) and the expression for surface charge, the effective length ( λG ) can be rewritten in terms of ( σ ):
λG=1+4π2LB2λD2(eσ)2λD
If we define 2Ψ0=u0=kBT2eϕ0 and CD=4πλDϵ, we can rewrite C in a simpler manner because why not
C=CDcosh(2u0)
The Electrocappilary Equation
The surface tension is related to the surface potential (ϕ0) as:
γ=γ0−4kBT2πLBc0cosh(2kBTeϕ0)
At higher electrolyte concentrations, the surface tension varies significantly with potential.
The curves flatten as the electrolyte concentration decreases, consistent with reduced ionic screening.
Potential Profiles
The Gouy-Chapman theory describes the electric potential and charge distribution in the electrical double layer formed at the interface between an electrode and electrolyte.
As mentioned before, the dimensionless potential is defined as:
Ψ0=2kBTeϕ
Dimensionless distance, is expressed in terms of the Debye length
X=LDx
The potential profile accounts for both non-linearized and linearized scenarios:
Non-linearized Potential Profile:
Accounts for large values of Ψ0, with exponential decay modified by ionic correlations.
Ψ(Ψ0,X)=ln[exp(Ψ0)+1exp(Ψ0)−1exp(−X)+1]
This exact form considers the influence of stronger potentials.
Linearized Approximation (Weak Potentials):
Suitable for weak potentials where the ionic distribution is approximately symmetric.
When Ψ0≪1, the potential simplifies to:
Ψ(Ψ0,X)≈Ψ0exp(−X)
This describes a simple exponential decay.
The charge distribution in the double layer is governed by:
Graham theory is critical for understanding interfacial capacitance in systems where no specific ionic adsorption occurs. It combines the contributions of different capacitances, enabling the characterization of interfacial phenomena.
The surface charge density (σ) is:
σ=−Q2
A negative Q2 indicates the charge's role in balancing the counter-ions in the diffuse layer.
The potential difference at the interface is split into components:
By defining the total interfacial capacitance (C) as the inverse of dσdϕ0, the total capacitance (C) can be expressed as:
C(σ)1=CH(σ)1+Cd(σ,c0)1
Helmholtz Capacitance (CH): Represents the contribution from the compact layer directly adjacent to the electrode. It does not depend on electrolyte concentration, but can depend on the charge
Diffuse Capacitance (Cd): Captures the influence of the Gouy-Chapman diffuse layer. Use the equation derived earlier
Hence, the slowest (smallest) capacitance dominates the total response, making it easier to identify the limiting factor in the system.
The Parsons-Zobel plot represents the relationship between the inverse of total capacitance C1 and the inverse of the diffuse layer capacitance Cd1
The y-intercept is the reciprocal of the Helmholtz layer capacitance. This is significant as CH remains independent of the bulk ionic concentration.
¶ Mean Field Molecular Models of the Compact Layer
The mean field molecular model aims to describe the electrostatic and thermodynamic behavior of a charged compact layer, such as that observed in ionic or electronic systems.
This model provides insights into the distribution of particles (e.g. electrons, ions) and the resulting electric field, capacitance, and polarization under applied potentials.
By simplifying the complex interactions within the layer to a "two-state" system, the model focuses on the competition between up (↑) and down (↓) states of particles and their alignment with the applied electric field.
Assumptions
Two-State Model
The two-state model assumes particles in the system can exist in only two discrete energy states, labeled up (↑) and down (↓).
These states may correspond to physical orientations (dipole alignment in an electric field)
Mean Field Approximation
This approximation used to simplify many-body problems. In this context, rather than accounting for the detailed interactions between individual particles, the effect of all particles is averaged into a single, uniform field.
This significantly reduces computational complexity while retaining key macroscopic behaviors, especially for systems with large particle numbers where fluctuations around the mean are small.
It is particularly effective in studying systems like compact layers, where the field primarily depends on the net charge distribution.
The energy of each state is influenced by the applied electric field E=dϕ, and the population of these states depends on the Boltzmann distribution:
N↓N↑∝exp(−kBTU↑−U↓+2pE)⎩⎪⎨⎪⎧NiUipPopuation of state iEnergy of state iDipole Moment
The order parameter, r, quantifies the alignment of particles with the field:
r=NN↑−N↓=tanh{kBTdp(ϕ−ϕc)},
N=N↑+N↓ is the total particle density
ϕc is the critical potential and is defined as: ϕc=2pd[U↑−U↓].
The potential ϕ satisfies the self-consistent relation: ϕ=ε∞1[4πσd−4πpNtanh(kBTdp(ϕ−ϕc))],
Two-State Capacitance
The surface charge density, σ, is expressed as:
σ=4πdε∞ϕ+dpNr
The differential (Helmholtz) capacitance, CH, measures the sensitivity of surface charge to changes in potential:
CH=dϕdσ=C0+4πLeffcosh(kBTdp(ϕ−ϕc))1.
C0=4πdε∞ is the intrinsic (geometric) capacitance
Leff−1=4πNkBT(dp)2LBeff is the effective screening length
Our overlord ✨Kornyshev✨ have cited his own paper on equations that describe contribution of metallic electrons to capacitance in double-layer systems. The model incorporates both spatial charge distribution effects and dielectric properties to refine the classical Helmholtz picture.
The Helmholtz capacitance, ( KH ), is expressed as:
4πKH1≈a−zσ+ε(σ)d
a: Closest distance of approach of excess charge to the metal surface.
a(σ)=a(0)−A>0;B>0Aσ−Bσ2
zσ: Center of mass of the excess charge distribution:
ε(σ): Dielectric constant as a function of the surface charge density (σ).
This equation highlights the competition between a, which is fixed by physical constraints, and zσ, which is determined by the spatial distribution of charge.
The differential capacitance, ( CH ), includes dynamic factors:
4πCH1≈a+σdσda−z∗+ε∗(σ)d
z∗ : Modified center of mass for differential charge, dependent on (σ).
z∗=dσd(σzσ),ε∗(σ)1=dσd[ε(σ)σ].
dσda: Sensitivity of a to changes in surface charge density, accounting for ionic rearrangements.
ε∗(σ): Effective dielectric constant that incorporates charge density effects.
The surface electronic density, ( nσ(z) ), describes the deviation from the neutral state and is defined as:
∫−∞0dz[nσ(z)−n+(z)]+∫0∞dznσ(z)=−σ
nσ(z): Electronic profile influenced by the field.
n+(z): Neutral density.
This equation ensures charge conservation, connecting the surface charge ( σ ) with the electronic density profile. It forms the foundation for calculating the total energy ( E[n] ).
The energy functional ( E[n] ) captures the contributions of electrostatics, quantum mechanics, and surface interactions:
E[n]=WH[ρ]+T[n]+Exc[n]+Eps[n]
WH[ρ]: Hartree energy from electrostatic interactions.
WH[ρ]=21∫dzρ(z)ϕ(z),ρ(z)=n+(z)−n(z).
T[n]: Kinetic energy of the electronic profile.
T[n]=∫dz{0.3(3π)2/3n(z)5/3+72n(z)∣∇n(z)∣2}
Exc[n]: Exchange-correlation energy, describing electron interactions.
zˉ=−n+σ: Mean position of the electronic profile.
β(σ): Scaling function, describing decay.
This trial function connects the profile ( nσ(z) ) to physical parameters such as σ, providing a basis for energy calculations.
Total energy functional:
E[β]=WH(β)+T(β)+Exc(β)+Eps(β).
Center of mass of the excess charge:
zσ=−σ1∫dzz[nσ(z)−n0(z)].
Simplifying:
zσ=2zˉ+σn+[β(0)21−β(σ)21].
The center of mass ( zσ ) measures the average spatial position of the excess charge ( σ ) relative to the surface.
zσ=−σ1∫dzz[nσ(z)−n0(z)],
nσ(z): Actual charge density profile in the presence of surface charge ( σ ).
n0(z): Neutral charge density profile, serving as the reference.
This equation builds on the trial function approach, where nσ(z) is explicitly modeled to describe the spatial distribution of electrons. Using the trial function’s symmetry and decay properties, zσ simplifies to:
zσ=2zˉ+σn+[β(0)21−β(σ)21]≈z0+pσ+rσ2
zˉ=−n+σ represents the mean position of charge, derived from the trial function.
β(σ): The decay parameter, describing how rapidly nσ(z) diminishes spatially.
The first theoretical approach by Marčelja and Radić (1976) described water using an order parameter profile η(z), where:
g=g0+aη2(z)+c(∂z∂η(z))2
This expansion leads to exponentially decaying forces due to water’s correlation lengths. The derivative of the free energy functional gives pressure as:
ΔG=∫0L(g−g0)dz;P=∂L∂ΔG≈−4ϵ0e−L/ξ0
indicating exponentially decaying structural forces within short ranges.
Field Theory in Pure Water
In field theory, three fields are introduced: E (electric field), D=ϵ0E+P (displacement field), P (polarization density field).
The Hamiltonian is decomposed into electrostatic Hel and water correlation Hcorr contributions: Hel[D0,P1,P2]=∫V2ϵ01(D0−(P1+P2))2dr Hcorr[P1,P2]=∫V2ϵ01[K11P12+K12(∇⋅P1)2+K13(∇2P1)2]
The correlation function χ∥(k) and dielectric function ϵ(k) link these fields, yielding three correlation lengths:
λd1,λd2: exponential decay lengths
λo: oscillation wavelength.
Coupled Differential Equations for Water Correlations
Minimizing the functional between two surfaces gives:
The Ising model is applied to describe order-disorder transitions in adsorbed ionic or molecular layers at an electrode. It models interactions between nearest-neighbor adsorption sites using the Hamiltonian:
H=−i,j∑(JxSi,jSi+1,j+JySi,jSi,j+1)
Si,j=±1: represents the adsorption states.
Jx,Jy: Coupling constants for horizontal and vertical nearest-neighbor interactions.
These transitions occur between two specific configurations of adsorbed particles:
1. Ordered Phase
(1×1) Phase
Adsorbed particles occupy every site uniformly.
E1×1=E0−Jx−Jy,
2. Different Ordered Phase
(1×2) Phase
The phase involves alternating adsorption rows
E1×2=E0+Jx−Jy,
Using these, the coupling constants ( Jx ) and ( Jy ) can be expressed in terms of the observed energy difference, ΔE(σ):
Jx=2ΔE(σ),Jy=E0−21(E1×2+E1×1).
The critical temperature (T) dictates when these transitions occur between ordered and disordered phases.
sinh(kBTJx(σ))sinh(kBTJy)=1
Relating the coupling constants (Jx,Jy) to T. This in turns allows ΔE(σ) to connect T and σ:
Jy∣ΔE(σ)∣=F(JykBT),F(x)=xln(coth(x1)).
For ΔE(σ)≈A(σ−σ0), the phase boundary can be written out with a linear approximation:
σ=σ0+sgn(σ−σ0)AJyF(Jy2kBT).
This approximation is used to construct the T-σ phase diagram:
Disordered Region:
The ( 1×1 ) and ( 1×2 ) phases are separated by a disordered region ( D ), where neither phase dominates.
The width of the disorder region Δσ is:
Δσ≈A2JyF(Jy2kBT)≈valid when 2kBT≪Jy8AkBTexp(−kBTJy)
The Ising order parameter M is a quantitative measure of the surface reconstruction state. It distinguishes between reconstructed and unreconstructed configurations of the surface and is defined as:
M=2LxLy1⎝⎛i(even),j∑Si,j−i(odd),j∑Si,j⎠⎞
Lx and Ly represent the dimensions of the lattice.
Si,j is a spin-like variable denoting the state of the surface atoms. Positive and negative values of Si,j correspond to different surface orientations or configurations.
The value of M tells us what state the system is in
1. Reconstructed State
M=±1
Fully reconstructed surface with ordered domains
2. Unreconstructed State
M=0
Represents the absence of reconstruction, where surface atoms are randomly or uniformly distributed without domain formation.
M exhibits critical behavior near the reconstruction transition and scales with the reduced temperature ( T−Tc) ) or reduced charge ( σ−σc ):
M∼(T−Tc)β∼(σ−σc)β
Here:
Tc: Critical temperature at which reconstruction occurs.
σc: Critical charge density.
β=81: Critical exponent for the Ising model.
Roughening and Ising Lines in the Free Fermion Theory
The roughening transition of steps on a surface can be understood through the partition function of a single step that occasionally jumps between neighboring atomic rows
Zs=exp(−kBTLyηs)[1+2exp(−kBTWs)]Ly⎩⎪⎨⎪⎧ηsWsLyStep energy per unit lengthEnergy of effective kinks Length of the step along the y-axis
The free energy associated with the step is Fs=kBTlnZs. Setting Fs=0, the roughening temperature Tr is derived:
Tr≈2kBηsexp(kBTrWs)
For domain walls on an Ising surface, a similar expression describes the Ising transition temperature Ti
Ti≈2kBηwexp(kBTiWw){ηwWwIsing domain wall energy per unit lengthEnergy of effective kinks on domain wall
The step energy for reconstructed and unreconstructed surfaces depends on their configuration:
Energy of the lowest step of unreconstructed surface (1×2):
ηs(1×2)=E1×2−E1×1=ΔE
Energy of the lowest step of reconstructed surface (1×1):
ηs(1×1)=2E1×1−E1×2=−2ΔE
The energy of an Ising wall relates to the (1×1) step energy:
ηw=2ηs(1×1)=−ΔE
Kink energies on domain walls and steps are connected:
Ws≈2Ww
Full Phase Diagram
The roughening of the unreconstructed (1×1) state involves the formation of (1×2) steps, which destabilize the surface. The associated charge (σr(1)) is given by:
σr(1)=σ0+2AkBTexp(−kBT2W0)
W0 is the reference energy for kink formation, and A relates to the surface charge per area.
Higher temperatures promote step roughening by overcoming the energy barrier 2W0.
Conversely, the roughening of the reconstructed (1×2) state occurs through the introduction of (1×1) steps. The charge (σr(2)) for this transition is expressed as:
σr(2)=σ0−4AkBTexp(−kBT2W0)
Compared to the (1×1) roughening, this transition involves a stronger dependence on temperature due to the coefficient 4, emphasizing the greater energy required to disrupt the reconstructed state.
As the temperature increases further, the surface transitions from ordered to disordered configurations.
This transition involves breaking the long-range order in reconstructed domains, with the charge σI given by:
σI=σ0−2AkBTexp(−kBTW0)
Here, the smaller energy barrier W0 compared to roughening transitions indicates that the order-disorder transition is less energetically demanding.
By combining these transitions, the phase diagram is constructed with distinct boundaries corresponding to the roughening and Ising transitions.
The roughening transitions (σr(1)) and (σr(2)) define the limits of stability for the (1×1) and (1×2) states, while (σI) marks the transition to a disordered state.
The relationships Ww=W0,Ws=2W0 ensure consistency between the step kink energies and domain wall energies
Lattice Gas Model and Ionic Distributions in Bulk and at Interfaces
We can use a lattice gas model to describe ionic liquids by quantifying ionic distributions in bulk and at interfaces under an electric potential. The free energy functional, F, forms the core of this model. It is given by:
F=eΦ(N+−N−)+kBTln(N−N+−N−)!N+!N−!N!,
N+ and N− are the numbers of cations and anions, respectively, N is the total number of available sites
Φ is the electric potential.
Minimizing the free energy functional with respect to N+ and N− yields the equilibrium ion distributions, N+(Φ) and N−(Φ).
A central parameter γ, defined as γ=NNˉ, represents the ratio of total ions in the bulk to total available sites
γ=cmax2c0 ties bulk concentration (c0) to the maximum possible ionic concentration (cmax).
The distribution configurations for ions and holes are captured via the partition terms:
C~0 is the normalized capacitance factor: C~0=4πϵκ~=αC0.
α is the asymmetry parameter: α=[1+2γ(a−b)]−1 with a and b representing ionic properties and γ the ionic saturation factor.
The graph below presents C0C as a function of u, showing dependence on α:
As α increases in value, the symmetry increases
Diffusion and Kinetics of Ionic Exchange
The diffusion coefficient D is temperature-dependent and follows the Arrhenius equation:
D≈D0e−Ea/kBT,
The modified Nernst-Einstein relation incorporates ion-specific properties, expressed as:
ΛNE=kBTc{p+q2D++p−q2D−}
c : Concentration,
p+,p−: Ion probabilities,
The survival probability function C(t) determines the state dynamics of ions:
C(t)=⟨h(0)h(0)⟩⟨h(0)h(t)⟩,
h(t) represents the state of a molecule:
h(t)={10if the molecule remains in the same state,otherwise.
The lifetime τ of an ion is calculated as:
τ=∫0∞C(t)dt.
Kinetics of Exchange Process:
Free and Bound States: The survival probability function C(t) is studied for both free and bound states of ions.
Exponential Models: The data are fitted using exponential and biexponential decay models:
Ci(t)=aie−t/τf,i+(1−ai)e−t/τs,i,
where τf,i and τs,i represent fast and slow time constants.
Two-State System with Exchange Theory: Velocity Autocorrelation Function (VACF)
The two-state exchange model describes molecular motion in systems where ions or molecules transition between free and bound states. The dynamics are captured using the Velocity Autocorrelation Function (VACF), denoted as Kv(t):
Kv(t)=⟨v(0)⋅v(0)⟩⟨v(0)⋅v(t)⟩.
p1: Probability of the free state.
p2: Probability of the bound state.
The probability that a molecule remains in the same state over time is given by:
ci(t)=aie−t/τf,i+(1−ai)e−t/τs,i
ai: Fraction of ions in fast relaxation,
τf,i: Fast relaxation time,
τs,i: Slow relaxation time.
The derivative of ci(t), gi(t)=−dtdci(t), describes the decay rate of the survival probability.
The VACF is expressed in the s-domain (Laplace transform) as:
where the overall VACF is a weighted sum of the contributions from free and bound states.
Anomalous Capacitance
The anomalous capacitance phenomenon in conducting nanopores is explained using the concept of a "superionic state". In slit nanopores, the electrostatic potential is described by: