Abstract

The double layer on transition metals, i.e., platinum, features chemical metal–solvent interactions and partially charged chemisorbed ions. Chemically adsorbed solvent molecules and ions are situated closer to the metal surface than electrostatically adsorbed ions. This effect is described tersely by the concept of an inner Helmholtz plane (IHP) in classical double layer models. The IHP concept is extended here in three aspects. First, a refined statistical treatment of solvent (water) molecules considers a continuous spectrum of orientational polarizable states, rather than a few representative states, and non-electrostatic, chemical metal–solvent interactions. Second, chemisorbed ions are partially charged, rather than being electroneutral or having integral charges as in the solution bulk, with the coverage determined by a generalized, energetically distributed adsorption isotherm. The surface dipole moment induced by partially charged, chemisorbed ions is considered. Third, considering different locations and properties of chemisorbed ions and solvent molecules, the IHP is divided into two planes, namely, an AIP (adsorbed ion plane) and ASP (adsorbed solvent plane). The model is used to study how the partially charged AIP and polarizable ASP lead to intriguing double-layer capacitance curves that are different from what the conventional Gouy–Chapman–Stern model describes. The model provides an alternative interpretation for recent capacitance data of Pt(111)–aqueous solution interfaces calculated from cyclic voltammetry. This revisit brings forth questions regarding the existence of a pure double-layer region at realistic Pt(111). The implications, limitations, and possible experimental confirmation of the present model are discussed.
Keywords: electric double layer, metal−water interactions, specific adsorption, partial charge transfer, water chemisorption, inner Helmholtz plane, pseudo-capacitance, double layer capacitance
Introduction
There is a wave of renewed interest in the electric double layer (EDL) structure in electrocatalysis, probably due to two reasons.1−4 On the one hand, the activity of many electrocatalytic reactions is revealed to be affected markedly by electrolyte cations5−9 and pH.10−13 There is growing consensus that demystifying these electrolyte effects requires fundamental knowledge of the EDL,3,14 as discussed in several review articles.14−17 On the other hand, a growing body of experimental data on the EDL of single crystals of transition metals, notably platinum, has been reported recently.4,18−23 These new data, though not fully consistent as detailed below, have already surprised us as they do not agree, even qualitatively, with classical double layer models.
A well-documented classical model is the Gouy–Chapman–Stern (GCS) model, which describes the EDL as a series combination of an inner layer and a diffuse layer.24 The GCS model has many variants, with detailed treatments of the diffuse layer and the inner layer.25−27 A hallmark of GCS-like models is the appearance of a minimum in the differential double-layer capacitance (Cdl) curve, named the Gouy–Chapman minimum, at a particular electrode potential, named the potential of zero charge (pzc), in dilute electrolyte solutions. Unless otherwise noted, pzc refers to potential of zero free charge. Moving away from the pzc in cathodic and anodic directions, Cdl grows and then decreases, resulting in a camel-shaped Cdl profile with two peaks and one valley. The two peaks usually signify overcrowding of counterions near the metal surface, which are usually asymmetric due to different sizes of cations and anions. The two peaks move closer and eventually melt into a single peak as the electrolyte solution gets more and more concentrated.28,29
The aptness of the GCS model has been corroborated on mercury-like electrodes, first by Gouy in the 1910s30 and then by Grahame in the 1940s and 1950s,31e.g., see Parsons’s review.32 In the presence of specific adsorption of anions, Grahame discriminated an inner Helmholtz plane (IHP) and an outer Helmholtz plane (OHP) in the inner layer.31 More than just a refinement of the Stern model, the Grahame model sowed the seeds of subsequent developments in the EDL theory, including the consideration of field-dependent orientation of water molecules,33−36 discreteness of adsorbed ions,37−40 and quantum effects of metal electrons,41−44 as discussed in a recent tutorial.27 Valette and Hamelin and coworkers have demonstrated that the GCS model is valid also for silver single crystals in nonspecifically adsorbing electrolytes.45−48
The idyll reaches an end for platinum. The Gouy–Chapman minimum is absent in the study of Pajkossy and Kolb where Cdl of Pt(111) in a KClO4 and HClO4 mixture electrolyte with a concentration down to 1 mM was measured.18,20,49 Instead, their data show a Cdl peak near the pzc in the anodic direction. Contrarily, Xue et al. observed Gouy–Chapman minima for Pt(111) in 0.05 M MeClO4 electrolytes (Me = Li, Na, K, Rb, Cs), however, at potentials ∼0.2 V below the pzc.23 More recently, Ojha et al. observed the Gouy–Chapman minimum for Pt(111) in 0.1 mM HClO4.21,22 In addition, they found that the Gouy–Chapman minimum turns into a maximum near the pzc when increasing the electrolyte solution above 1 mM, qualitatively consistent with previous observations of Pajkossy and Kolb.18,20,49
One should be aware of the discrepancy in the magnitude of Cdl of Pt(111) at the pzc, denoted Cdl, pzc, in different studies. Cdl, pzc is around 20 μF cm–2 in the study of Pajkossy and Kolb using an electrolyte concentration of 0.1 M,18,20,49 in the range of 10–30 μF cm–2 for 0.05 M NaOH and 0.10 M HClO4 in the study of Schouten et al.,50 around 30 μF cm–2 in the study of Xue et al. with an electrolyte concentration of 0.05 M,23 and around 100 μF cm–2 in the study of Ojha et al. with an electrolyte concentration of 0.005 M.21,22 Note that Cdl, pzc shall decrease with the electrolyte concentration, according to the GCS model. Surprisingly enough, recent values of Cdl, pzc in more dilute electrolyte solutions are even several times higher than previous values in more concentrated electrolyte solutions. Understanding the origin of the anomalously large Cdl, pzc is an open question on which several viewpoints already exist, which I am about to discuss, and that motivates this study.
To rationalize the large Cdl, pzc, Doblhoff-Dier and Koper introduced weak (around several hundred meV), attractive ion–surface interaction, of which possible origins are also discussed.51 This effectively increases the ion concentration in the double layer, shortens the Debye length, and magnifies the magnitude of Cdl, pzc. Recently, Schmickler rationalized the higher Cdl, pzc as a consequence of weak specific adsorption of anions.52 The introduction of anion adsorption can increase the overall capacitance, and decrease the Parsons–Zobel slope. However, this model cannot explain that the Parsons–Zobel slopes are not sensitive to the identity of electrolyte anions.22 More recently, Ojha, Doblhoff-Dier, and Koper modified their previous model by further adopting a two-state submodel for the first-layer water molecules,22 which was developed earlier by Le et al.53
This marked and surprising discrepancy
in Cdl, pzc might be ascribed to
different methods used to
determine Cdl. Pajkossy and Kolb extracted Cdl from electrochemical impedance spectroscopy
(EIS) data using the Frumkin–Melik–Gaikazyan (FMG) model.18,20,49 Xue et al. extracted Cdl from EIS data using an electric circuit model
involving empirical constant phase elements.23 Ojha et al. calculated Cdl from cyclic
voltammogram (CV) data.21,22 EIS is, in principle,
more accurate than CV in determining Cdl, because EIS can separate Cdl from possible
pseudo-capacitances of chemisorption. In the presence of chemisorption,
the FMG model expresses the total capacitance as
, where Cdl is
the pure double layer capacitance, Rad is the adsorption resistance, Cad is
the adsorption capacitance, σad is the coefficient
of Warburg diffusional impedance, and ω is the angular frequency.
Only when ω ≫ ωc = (CadRad)−1 can one approximate Ctotal(ω)
as Cdl. At low frequencies, one will find Ctotal(ω) ≈ Cdl + Cad. Earlier studies have
shown that ωc is around 1 MHz for hydrogen adsorption
at Pt(111) in 0.5 M perchloric acid54 and
above 10 kHz for hydroxyl adsorption at Pt(111) in 0.1 M perchloric
acid.50 A CV measurement with a scanning
rate of 50 mV s–1 over a potential range of 0.05
V consumes 1 s and, hence, has a characteristic frequency of 1 Hz,
which is far lower than the typical ωc of Pt(111)
electrode processes. Ojha et al. reported the capacitance calculated
from impedance at a single frequency of 18 Hz, which is also much
below ωc.22
The fact that the CV-derived capacitance is much higher than the EIS-derived capacitance could be attributed to pseudo-capacitances of chemisorption. Based on this assumption, a theory for the large CV-derived capacitance of Pt(111) in the double layer region is developed in this paper. I am about to show that chemisorption at a small percentage of platinum sites can explain the observed large capacitance.
The remainder of this paper is organized as follows. First, I present a short review of the developments on the IHP model after the pioneering works of Grahame and Parsons.55,56 Most of these important works that adequately reflect the depth, core, and beauty of electrochemical research are less known and only sparsely cited nowadays. This literature review also sets the baseline for the present model, which is not entirely new, but extends previous models in some aspects. Second, a detailed derivation of the model, including a water layer submodel and a chemisorption submodel, is presented. Third, the model is demystified for two cases. In the first case without chemisorbed ions, emphasis is put on understanding how metal–water interactions change the pzc and the double layer capacitance. For the second case with chemisorbed ions, emphasis is put on the consequence of chemisorption-induced surface dipole moment. The model is then used to interpret recent experimental data of Ojha et al.21,22 in the double layer region. This revisit questions current fundamental understanding of the EDL at Pt(111). Lastly, discussion on assumptions, implications, and possible extensions of the model is presented before the conclusion section.
Models of the IHP: Grahame–Parsons Model and beyond
Above analysis has turned the spotlight on the IHP in understanding the EDL at transition metals. Figure 1 summarizes three groups of IHP models. The Grahame–Parsons model views the IHP as an ionic layer with specifically adsorbed ions rigidly lining up.31,55,56 The potential drop from the metal surface to the IHP is composed of two parts. The first part is due to the metal surface charge, and the second part is the net charge of specifically adsorbed ions at the IHP. There is no electron transfer during specific ionic adsorption; hence, the specifically adsorbed ions retain the charge they have in the bulk solution. In the literature, e.g., ref (57), there is another Grahame–Parsons model, which describes the thermodynamics of specific ion adsorption, and determines the coverage of adsorbed ion as a function of the electrode potential. The Grahame–Parsons model of ion adsorption was later improved by the Alexeyev–Popov–Kolotyrkin model58 and the Vorotyntsev model.59
Figure 1.
Models of the inner Helmholtz plane (IHP). (a) Grahame–Parsons model views the IHP as an ionic layer where specifically adsorbed ions are rigidly aligned.31,55,56 (b) Many models are focused on the field-dependent behaviors of solvent molecules, most often water molecules, at the IHP.34,36,63 (c) Recent models consider the coexistence of solvent molecules and partially charged chemisorbed ions (A–ξ) at the IHP.72,73
Another group of IHP models focuses on solvent molecules at the IHP in the absence of specifically adsorbed ions, (Figure 1b); see an earlier review by Guidelli.60 The two-state model assumes that water molecules take either the hydrogen-up or hydrogen-down configuration, depending on the local electric field.34 The three-state model further considers formation of water dimers and interconversion between water dimers and monomers.36 Both models are founded purely on electrostatic considerations; thus, the average dipole moment of the water layer is zero when the metal surface has a zero net charge. In addition to electrostatic interactions, chemical interactions between water molecules and the metal surface were speculated in the 1970s.61 Damaskin and Frumkin considered both electrostatic and chemical metal–water interactions at the IHP.62 Electrostatic interactions are treated using the approach of Bockris, Devanathan, and Miller (BDM),63 which does not consider properly the configurational entropy. The fraction of chemisorbed water molecules was expressed phenomenologically as an exponential function of the electrode charge. Recently, we have seen a surge of atomistic simulations of water molecules near metal surfaces.53,64−68 These first-principles calculation results provide unbiased, accurate atomistic details, guiding the development of phenomenological models, as demonstrated in the work by Le et al.53
In most cases, the IHP is neither an ionic layer nor a solvent layer. Instead, it is a mixture of solvent molecules and chemisorbates. Chemisorbates, usually partially charged due to partial charge transfer, contribute an additional potential drop in the inner layer.69 Therefore, a refined model of the IHP should be a binary IHP model that considers co-existing solvent molecules and partially charged ions. Carnie and Chan developed a so-called civilized model of EDLs, which considers ions as hard spheres with embedded point charge (integral charges, rather than partial charges), and solvent molecules as hard spheres with embedded dipole moment in the 1980s.70 The Carnie–Chan model was used to calculate the dipole moment of specifically adsorbed ions at the IHP.71 More recently, the binary IHP model was further extended to consider potential-varying coverage of partially charged chemisorbates.72,73 More than just a further complication of a model, it brings forth nontrivial implications of chemisorption-induced surface dipole moment (μchem). It was revealed that μchem of adsorbed hydroxyl leads to a nonmonotonic surface charging behavior of Pt(111)–aqueous solution interfaces. Specifically, the free charge, not the total charge nor the metal charge,74 exhibits a negative–positive–negative transition with increasing electrode potential. This implies the occurrence of negative Cdl and more than one potential of zero free charge. Nonmonotonic surface charging behaviors of a polycrystalline Pt electrode were measured using the radioactive tracers method by Frumkin and co-workers75,76 and later in several experiments of laser-induced potential transients.77−79 The nonmonotonic surface charging behaviors were modeled earlier by assuming that the pzc grows linearly with the coverage of adsorbed oxygen and later by a phenomenological treatment in which the potential drop in the IHP is interpolated between two limiting states.76 The model developed in refs (72) and (73) and later modified in ref (80) provides a consistent treatment of chemisorption and surface charging behaviors as a function of electrode potential. It was also shown in ref (72) that orientational polarization of interfacial water molecules, described using the two state model, leads to a negative capacitive component to Cdl, an effect that was later elucidated in greater detail by Le et al.53
The IHP model that I am about to present in this paper improves over the previous one in refs (72) and (73) in three aspects. First, the present model abandons the BDM-type two-state water model and adopts a statistical treatment of water polarization with a continuum spectrum of orientational states, following the approach developed earlier in refs (81) and (82). The configurational entropy of water molecules at the IHP is considered. In addition to electrostatic interactions, chemical interactions with the metal are considered, as inspired by ab initio molecular dynamics (AIMD) results,53,66,67 but in a different manner compared to the Damaskin–Frumkin model. Second, chemisorption at the IHP is now energetically distributed in such a way that the equilibrium potential follows a distribution instead of taking a single value as usual. This allows treatment of chemisorption at different types of active sites, chemisorption of different species, and lateral interactions between chemisorbates in a unifying manner. Third, as suggested by recent detailed atomistic simulations, e.g., ref (83), the IHP is further divided into an adsorbed ion plane (AIP) and an adsorbed solvent plane (ASP). This refinement is necessary because the AIP and ASP have very different properties, which I am about to show in model parameterization.
Theoretical Methods
Physical Picture of the Double Layer
A schematic illustration of an electrocatalytic double layer is presented in Figure 2a, and a continuum model is depicted in Figure 2b. The metal electrode is assumed to be planarly uniform. Surface roughness is not considered here. The first-layer water molecules are usually highly polarized due to the high interfacial electric field and chemical interactions with the metal; see a recent review by Groß and Sakong.68 The ASP denotes the outer plane of adsorbed solvent molecules. Adsorbed ions, if any, are usually located closer to the metal surface than adsorbed solvent molecules, according to atomistic simulations, e.g., by Braunwarth et al.83 The central plane of adsorbed ions is denoted AIP. Different characteristic planes are used for depicting the ASP and AIP due to difference between ionic charge and dipolar charge. Nonspecifically adsorbed ions with their solvation shell reside in the diffuse layer stretching from the OHP toward the solution bulk. The regions between the metal and the AIP, the AIP and the ASP, and the ASP and the OHP are described as dielectric continua, parameterized with thicknesses δi and permittivities ϵi, with i = 1,2,3.
Figure 2.
(a) Schematic illustration of the electrocatalytic double layer with partially charged, chemisorbed ions, adsorbed solvent (water in this case) molecules, and nonspecifically adsorbed ions. (b) Continuum EDL model consists of an AIP denoting the central plane of chemisorbed ions, an ASP denoting the outer plane of adsorbed water molecules, and an OHP denoting the central plane of nonspecifically adsorbed ions. The regions between the metal and the AIP, the AIP and the ASP, and the ASP and the OHP are described as dielectric continua, parameterized with respective thicknesses δi and permittivities ϵi, with i = 1,2,3. (c) Overall potential difference between the metal bulk and the solution bulk can be divided into four components: χM denoting the potential drop at the metal surface due to electron distribution, Δϕad the potential change due to partially charged, chemisorbed ions, Δϕw the potential change due to orientational polarization of adsorbed water molecules, and Δϕfree the potential change due to excess ionic charge stored in the diffuse layer, namely, free surface charge.
Potential Differences in Different Components
As EDL is a grand canonical system with electrode potential being one of the independent variables, the overarching task of any EDL model is to calculate the electric potential distribution from the metal to the electrolyte solution. Figure 2c depicts components of the overall potential difference between metal bulk (ϕM) and solution bulk (ϕS). At the metal surface exists a large potential drop due to the metal electron tail, denoted χM, which is readily obtained from electronic structure calculations, e.g., in ref (84). There are, at least, three other potential differences, including one due to partially charged, chemisorbed ions, one due to adsorbed solvent (water in this work) molecules, and one due to the free surface charge density, σfree, defined as negative of the net ionic charge in the diffuse layer.80
The above decomposition scheme, following ref (85), is expressed as
| 1 |
Within a continuum picture, Δϕfree is written as
![]() |
2 |
where the first term denotes
the potential change over the diffuse layer, which is derived from
the Bikerman model considering finite ion size,28,73
is the Debye length, ϵb is the dielectric permittivity of the bulk solution, cb is the bulk ion concentration, R is
the gas constant, F is the Faraday constant, T is the temperature, and γ = 2cbdh3 with dh being
the diameter of hydrated ions. Size asymmetry of cations and anions
is not considered here, for the sake of an analytical expression of
Δϕfree as a function of σfree, and effects of size asymmetry on the Cdl are known already.86−89
The other two components Δϕad and Δϕw are determined from two submodels as introduced below.
Distributed Chemisorption Model
The amount of charge
carried by partially charged, chemisorbed ions at the AIP is calculated
as,
, where Nad is
the number density of adsorption sites, θA and ξA are the coverage
and net electron of a chemisorbed ion of type A, and e0 is the elementary charge.80 The parameter ξA is related to
the partial charge transfer coefficient, λA, via ξA = zA + λA, with zA being the initial charge of chemisorbed ions. A total charge
transfer process corresponds to, λA = – zA. Note in passing that
the net charge ξA and the partial
charge transfer coefficient λA are
different from the electrosorption valency that can be measured experimentally.85
Δϕad is then calculated
as the potential difference across an equivalent planar capacitor
with a charge amount of
on each side
| 3 |
A positive-valued ξA means negatively charged chemisorbed ions, which introduces a positive Δϕad. A chemisorption submodel is required to determine θA as a function of the electrode potential.
For chemisorption of hydrogen, H+ + e + * ↔ Had, the Frumkin isotherm gives90,91
| 4 |
where θH is the coverage of Had, θH, max is the maximum coverage, γH is the lateral interaction coefficient, β = F/RT, and EH0 is the equilibrium potential on the standard hydrogen electrode (SHE) scale. Recently, Hormann et al. has illustrated how to obtain EH and γH from density functional theory (DFT) calculations.92 The existence of lateral interactions stands in the way of obtaining an analytical solution for θH.
The Frumkin isotherm can be transformed into a series of Langmuir-type isotherms with a distributed EH0,
| 5 |
where fH(E) denotes the probability of an equilibrium potential at E.
The distribution in equilibrium potential can be ascribed to the existence of structural defects, lateral interactions that change the electrochemical potential for later adsorbates, and a potential-dependent distribution in the electrochemical potential of reactants, among others. I will go back to this in the discussion section.
As any distribution can be expanded as a series of normal distributions, EH0 is assumed to take the following general form,
| 6 |
where gk(E; μk, σk) are normal distributions of weight pk, expectations μk, and variances σk
| 7 |
and ∑kpk = 1 for the weight values.
Now, θH is expressed analytically as
| 8 |
Similarly, the coverage of adsorbed OH is written as
| 9 |
where fOH(E) is the distribution function of the equilibrium potential of OH adsorption, which has the same form as eq 6.
The pseudo-capacitance of chemisorption of Had and OHad is calculated as
| 10 |
with A = H (−) or OH (+).
Cautious readers may think a term (1 ± ξA) ( + for Had and – for OHad) should be added in eq 10 to account for the partial charge transfer. This is unnecessary because no matter how the electron is partitioned between the metal and the adsorbate, one electron is taken out from or moved into the electrode, if we separate the total surface charge into σfree and the charge involved in specific adsorption of ions. I will explain this point later when discussing the relationship between the total capacitance, CAad, and Cdl in eq 24.
The Frumkin isotherm can be regarded as a Langmuir isotherm with the equilibrium potential linearly varying with the adsorbate coverage, with the linear coefficient being the lateral interaction coefficient. In this sense, the distribution function expressed in eq 6 is just a generalization of the Frumkin isotherm. In fact, the linear coverage dependence of the equilibrium potential has no fundamental reason but is just a first approximation. Nonlinear effects have been observed in many cases such as the chemisorption of OH at Pt(111) in the study of Climent et al.91 The exact form of fA(E) corresponding to the Frumkin isotherm is unknown currently. However, I will extract fH(E) and fOH(E) from experimental CVs that are usually fitted using Frumkin isotherms.
Water Model with a Continuous Spectrum of Orientational States
The last term in eq 1 is the potential difference due to the average dipole moment of the first-layer water molecules,
| 11 |
where θw is the coverage of water molecules, μw is the dipole moment of a water molecule, and ⟨ cos αw⟩ is the statistical average of cosαw with αw being the angle between water dipoles and the local electric field, as depicted in Figure 3. Herein, a positive electric field points toward the bulk solution, and the direction of a dipole points from its negative end to its positive end.
Figure 3.

(a) αw is the angle between water dipoles and the local electric field. A positive electric field points toward the bulk solution, and the direction of a dipole points from its negative end to its positive end. Chemical metal–water interactions result in an energy difference between water of αw = 0° and that of αw = 180°, denoted δμ. (b) Shift in the potential of zero charge as a function of δμ̃ (the solid line: exact results, the dotted line: linear approximation). The calculation uses the following set of model parameters: Nad = 0.15 Å–2, θw = 0.5, ϵ2 = 3.25ϵ0, μw = 0.73 D, as discussed in the model parameterization section. Matlab script of this figure is provided in the Supporting Information of this paper.
In this model, αw could take any value between 0 and 180°, instead of two values in the Watts–Tobin model and the BDM model that were widely adopted in recent work.53,72 In this sense, the present treatment could be named an infinite-state water model.
Considering a canonical ensemble consisting of Nw water molecules at the IHP, at the mean-field level, the Hamiltonian of this canonical ensemble is written as
| 12 |
where αi is the angle between the ith water
and the local electric field
and δμ is the chemical interaction
energy difference between the case of αw = 0° and the case of αw = 180°; see the illustration in Figure 3a. The introduction of δμ accounts for
chemical interactions between water molecules and the metal surface.
Damaskin and Frumkin also considered chemical metal–water interactions.62 A key difference between this water model and
the Damaskin–Frumkin model is that a statistical ensemble approach
is adopted here.
The partition function of the water layer is written as93
| 13 |
where ψi is the longitude of the solid angle.
The Helmholtz free energy of the water layer is calculated as
| 14 |
The statistical average of cos(αi) is calculated as
![]() |
15 |
Surface Charging Formula
Combining eqs 1, 2, 3, and 11 leads to a formula from which one can solve for σfree as a function of ϕM,
![]() |
16 |
At the pzc, σfree = 0, eq 16 is thus simplified to
| 17 |
where the subscript 0 denotes the pzc condition.
If there is no chemisorbed ions at the pzc, eq 16 is further simplified to
| 18 |
where the dimensionless
is used.
The potential difference ϕM – ϕS at an electrode–electrolyte interface is not measurable without involving its counterpart at a reference electrode94 and must be differentiated from the electrode potential E in experiments.
Conversion from (ϕM– ϕS) to Electrode Potential
In potentiostatic experiments, one can control the potential of the working electrode (WE), namely, the metal in Figure 2a, with respect to a reference electrode (RE). The electrode potential is written as E = ϕM – ϕRE with ϕRE being the inner potential of the RE, up to an unimportant constant accounting for different work functions of WE and RE materials. The electrode potential could be rewritten as a difference between two potential differences, E = (ϕM – ϕS) – (ϕRE – ϕS). Here, it is assumed that the RE is put in the same electrolyte with the working electrode, and ϕRE – ϕS is the potential difference between the RE and the electrolyte solution. The reversible hydrogen electrode (RHE) belongs to this case. If the RHE is used as the RE, (ϕM – ϕS) becomes
| 19 |
where the subscript ‘RHE’ specifies the reference electrode. This equation can be related to the constant inner-potential DFT proposed recently by Melander et al.95
According to the Nernst equation, the potential difference of the RHE and that of the standard hydrogen electrode (SHE) are correlated as
| 20 |
where (ϕRE – ϕS)SHE is the value at the SHE.
Combined, eqs 19 and 20 lead to
| 21 |
given that
.
Substituting eq 21 into eq 16 gives
![]() |
22 |
where (ϕRE – ϕS)SHE – χM is, in general, a function of ESHE.
Double-Layer and Total Capacitance
In the presence of chemisorption, it is important to distinguish two types of differential capacitance. The differential double-layer capacitance, corresponding to the free surface charge, is defined as
| 23 |
The total surface charge σtotal, of which the change is measurable by counting electrons flowing through the external circuit, is the sum of σfree, corresponding to the excess ionic charge stored in the diffuse layer, and the charge involved by specific adsorption of ions, namely,74
with A = H (−) or OH (+). The differential total capacitance, corresponding to σtotal, is thus the sum of Cdl and pseudo-capacitances given in eq 10,
| 24 |
It is important to understand that specific adsorption of ions not only contributes to CAad but also influences Cdl via the resultant surface dipole moments.
Model Parameterization
The process of parameterizing a continuum EDL model is more or less a process connecting it to atomistic simulations using Kohn–Sham DFT, reactive force field molecular dynamics (MD), etc. The model contains four groups of parameters. The first group concerns the multilayer structure and dielectric properties of the EDL. The second and third group describe adsorbed water molecules and chemisorbed ions, respectively. The last group defines material properties of the metal and the electrolyte solution.
The first group of parameters includes ϵi and δi. Recent years have witnessed detailed DFT and MD studies of Pt(111) electrodes with chemisorbed H and OH. Le et al. showed that the density of Had peaks at ∼1 Å above the metal surface.67 Braunwarth et al. gave a bond length of 1.05 Å for adsorbed O on Pt(111).83 Therefore, I use δ1 = 1Å for the AIP. In the presence of an electron tail in the space between the metal and the AIP, ϵ1 should be higher than the vacuum permittivity ϵ0; otherwise, Cdl is bounded by CAIP = 8.85 μF cm–2. Herein, I assume ϵ1 = 10ϵ0 to account for the effects of electrons spilled from Pt(111). These studies also indicated that the adsorbed, first-layer water molecules are around 2.1 Å above the metal surface. Therefore, δ2, the thickness of the gap between the AIP and ASP, is approximated as δ2 = 1.1 Å. Since orientational (inertial, slow, low-frequency) polarization of water dipoles is explicitly considered in Δϕw, ϵ2 is the electronic (noninertial, fast, high-frequency) permittivity, which is known to be ϵ2 = 3.25 ϵ0 at room temperature.96 Le et al.67 used ϵ2 = 4ϵ0. An earlier estimate by Bockris et al. reads ϵ2 = 6ϵ0.63 For the region between the ASP and OHP, I used δ3 = 3.5 Å, a typical radius of hydrated ions, and ϵ3 = 78.5 ϵ0, the permittivity of bulk solution.
As for adsorbed water molecules, Le et al. determined a coverage
of ∼0.5.53 Hence, I use θw = 0.5. Schnur and Groß calculated the adsorption energy
of the H-down and H-up water at several solids using an ice-like water
bilayer model.97 The H-down and H-up configurations
correspond to αw = 180° and αw = 0°, respectively. Their data show that
δμ̃ ≈ 0 at Ru(0001) and δμ̃
≈ 1 at Au(111), Ag(111), and Pt(111). It is understood that
δμ̃ has the unit of kBT. In more recent AIMD simulations, a more detailed
picture of interfacial water molecules has emerged.53,66,98 At the Pt(111)–water interface, a
fraction of water molecules are bound strongly through their oxygen
atom with the metal so that the hydrogen atoms of the water molecule
are above the oxygen atom. This is called the O-down configuration,
corresponding to αw = 0° in this
model. A bit further away from the metal surface resides another type
of water molecules, which can be further divided into the H-down (αw = 180°) and H-up (αw = 0°) configurations. These recent AIMD results suggest δμ̃
< 0 for the present model. The intrinsic dipole moment of a water
molecule is 1.85 D. However, in order to reproduce the permittivity
of bulk water using the dipolar Poisson–Boltzmann theory, Abrashkin
et al. used an effective value of 4.86 D.99 Herein, I determine μw from the interfacial potential
change due to water orientation, which was determined to be −0.2
V for a clean, uncharged Pt(111) surface by Le et al.65 From eq 22, the present model determines the orientation-induced potential
change as
. Assuming δμ̃ = –
1, I thus determine
. If δμ̃ = – 2,
μw = 0.43 D. The lower bound of μw is 0.23 D. A lower effective value of μw accounts
for the unevenness of the first-layer water molecules as well as offset
effects by second-layer water molecules.
Malek and Eikerling calculated the work function change due to chemisorbed oxygen at Pt(111), from which they estimated the partial charge residing on each adsorbed oxygen.100 Using a vacuum permittivity, they determined that each adsorbed oxygen on Pt(111) carries a partial charge of ∼0.012e. Since ϵ1 = 10ϵ0 is used here, I should use ξOH = 0.12 to reproduce the same magnitude of work function change induced by chemisorbed ions. Chemisorbed H is assumed to be electroneutral, namely, ξH = 0. The present model does not need a single-valued equilibrium potential and lateral interaction coefficient. Instead, this model only needs fH(E) and fOH(E), which are determined from experimental CVs.
With Pt(111) as the metal, I have Nad = 0.15 Å–2 based on a lattice constant of 3.92 Å, and the potential of zero charge is Epzc = 0.29 VSHE.101 Given that water dipole orientation lowers the pzc by 0.2 eV, Epzc0 without considering this water dipole effect should be 0.49 VSHE. In accordance with experiments by Ojha et al.,21,22 I consider an electrolyte solution of 0.1 mM HClO4 + x mM LiClO4. The diameter of hydrated cations and anions is taken uniformly as dh = 7 Å.
The Case without Chemisorbed Ions
The central formula expressed in eq 22 contains the interplay between the diffuse layer, water molecules at the ASP, and partially charged chemisorbed ions at the AIP. To simplify the matter, I consider in this section a simple case without chemisorbed ions, namely, θA = 0 in eq 22. The focus is put on how chemical interactions between adsorbed water molecules and the metal influence the pzc and furthermore how field-dependent reorientation of water molecules changes the Cdl profile.
Influence of Metal–Water Interactions on the pzc
In the absence of chemisorbed ions, the pzc is obtained from eq 22 as
| 25 |
where the first term is the potential drop at the metal surface determined by the electron density profile at the interface, the second term is a constant of the SHE, and the third term represents the influence of water molecules at the ASP.
Epzc, SHE0 is defined as the value for the case of δμ̃ = 0, namely, the case in which metal–water interactions are the same for water molecules possessing αw = 0° and αw = 180°. This artificial case is hardly met. In the general case of δμ̃ ≠ 0, I obtain
| 26 |
where the second term on the right-hand side represents the pzc shift due to orientational polarization of adsorbed water molecules. It is important to emphasize that the pzc shift due to electronic polarization of adsorbed water molecules is included in Epzc, SHE0.
When |δμ̃| ≤ 1, Epzc, SHE is asymptotic to
| 27 |
Equation 27 implies that Epzc, SHE is greater (smaller) than Epzc, SHE0 when δμ̃ > 0 (<0), as shown in Figure 3b. For the case of δμ̃ < 0, the first-layer water molecules prefer the αw = 0° (H-up or O-down) configuration. For this case, the resulting potential rise due to the water dipole moment compensates partly for the potential drop at the metal surface, leading to a smaller Epzc, SHE. In the opposite case, the first-layer water molecules prefer the αw = 180° (H-down) configuration, resulting in an additional potential drop and thereby a higher Epzc, SHE.
Influence of Water Orientational Polarization on Cdl
In this section, I analyze how water orientational polarization affects the differential double-layer capacitance Cdl. In addition to the full model, two simplified models are introduced to differentiate multiple coupled effects. The simplified model without ion size effects is described by a variant of eq 22 in the limit of γ = 0,
| 28 |
In eq 28, I have neglected the chemisorption-induced dipole moment for this case without chemisorbed ions. The simplified model without water effects is expressed as a variant of eq 22 in the limit of θw = 0,
![]() |
29 |
In Figure 4, I compare the three models for the case of δμ̃ = 0 with the base set of model parameters. The model without water effects has a camel-shaped Cdl curve with two very broad peaks corresponding to overcrowding of counterions. The Cdl curve is symmetric around the pzc as I have neglected size asymmetry. Otherwise, the peak corresponding to smaller counterions will be higher, see, e.g., ref (27).
Figure 4.

Differential double layer capacitance as a function of the electrode potential for the case of δμ̃ = 0 for two effective values of water dipole moment: (a) μw = 0.73 D, (b) μw = 0.90 D. Epzc, SHE0 represents the potential of zero charge without water dipole effects, which is used as the electrode potential reference here. Three models, including the full model, a simplified model without ion size effects, and a simplified model without water effects. The electrolyte solution is 0.1 mM HClO4. δμ̃ = 0 and other model parameters have their base values. Matlab script of this figure is provided in the Supporting Information of this paper.
Upon an increment in the magnitude of σfree, adsorbed water molecules at the ASP could reduce the potential drop across the EDL via rearranging their orientations. Consequently, Cdl is increased due to water polarization effects, as seen in the comparison between the full model and the model without water effects. Equivalently, water molecules at the ASP contribute a negative capacitance in series with the diffuse layer capacitance, as pointed out earlier by Huang et al.72 and Le et al.53
The two peaks of the Cdl curve become sharper and move closer to the pzc, compared to their counterparts of the model without water effects. The two peaks signify saturation of water polarization effect, rather than overcrowding of counterions in previous modified Poisson–Boltzmann theories.28,29 Mathematically, it corresponds to ⟨ cos αw⟩ = 1 or – 1 depending on the sign of σfree. These two peaks are significantly elevated when μw increases from the base value of 0.73 to 0.9 D, as shown in Figure 4b.
Since the electrolyte solution is ultradilute, c0 = 0.1 mM, the ion size effect is unimportant near the pzc as the Cdl curves of the full model and the model without ion size effects overlap in the region of |ESHE – Epzc, SHE0| < 0.2 V. As σfree grows in magnitude and the EDL becomes more crowded with counterions, the ion size effect cannot be neglected any longer. Without considering the ion size effect, the two peaks are higher.
Figure 5a examines the influence of δμ̃ on the Cdl curve. As already discussed, the pzc, corresponding to the valley of the Cdl curve, shifts to more negative electrode potentials as δμ̃ goes negative. The Cdl curve becomes highly asymmetric with the left peak being much higher than the right one for δμ̃ = – 1. A negative δμ̃ means that in the absence of electrostatic interactions, water molecules prefer the configuration of αw = 0° (the O-down or H-up configurations). Therefore, these water molecules have a higher capacity to screen negative σfree, which intends to rotate water molecules toward αw = 180°, than positive σfree. Consequently, a higher Cdl peak is observed in the negatively charged region. Interestingly, the peak at the negatively charged surface almost disappears for δμ̃ = – 6. For the case of such a very negative δμ̃, the orientation of water molecules is dominated by the strong chemical interactions and is no longer tunable by the local electric field. Therefore, the Cdl curve resembles that of the model without water effects in Figure 4, except a horizontal shift due to δμ̃.
Figure 5.

Differential double layer capacitance as a function of electrode potential for (a) different values of δμ̃ normalized to thermal energy kBT when the electrolyte concentration is 0.1 mM and (b) different electrolyte concentrations when δμ̃ = – 1. Epzc, SHE0 represents the potential of zero charge without water dipole effects, which is used as the electrode potential reference here. Other model parameters have their base values. Matlab script of this figure is provided in the Supporting Information of this paper.
Figure 5b examines the concentration dependence of the Cdl curve while δμ̃ = – 1 is fixed. Cdl near the pzc is higher because the Debye length becomes smaller as the electrolyte concentration increases. In addition, the two peaks come closer in more concentrated electrolyte solution. These peaks are ascribed to saturation of water polarization. Let us denote as σfreec the critical magnitude of σfree leading to saturated water polarization. At higher concentrations, Cdl is higher and σfree can be achieved at a smaller potential difference away from the pzc.
The Case with Chemisorbed Ions
Effect of Chemisorption-Induced Dipole Moment
Now chemisorption is brought into the play, and its influence on the Cdl curve is examined. Let us consider the Pt(111)–0.1 M HClO4 interface. The CV of this interface is usually divided into a hydrogen adsorption region up to ∼0.4 VRHE, a hydroxyl adsorption region above ∼0.55 VRHE, and a double layer region in between. Climent et al. have conducted a comprehensive analysis of CVs of this interface at different temperatures using Frumkin isotherms.91 Herein, a revisit from a different view assuming distributed equilibrium potentials of Had and OHad is presented, with the focus put on the effect of chemisorption-induced dipole moment on Cdl.
Figure 6a shows a model fitting of the CV within the potential range of [0.15, 0.75]VRHE and a model-based decomposition of the total capacitance Ctot into Cdl and adsorption capacitances (CHad and COHad). The extracted distribution functions fH and fOH, defined in eq 6, are shown in Figure 6b. fH is composed of four Gaussian functions with preset mean values of 0, 0.1, 0.2, and 0.3 VRHE, and fOH is composed of two Gaussian functions with preset mean values of 0.7 and 0.8 VRHE. Twelve free parameters are therefore the variances σk and weights pk of these six distributions. Other EDL parameters have their base values. The basic idea of this approach resembles that of the distribution of relaxation times (DRT) proposed by von Schweidler in 1907; see a recent review on DRT.102 The present approach could be termed distribution of chemisorption energies (DCE). A better fitting of Ctot can be obtained if a greater number of Gaussian functions are included, which is not the purpose of this work. The mere purpose of this fitting is to confirm that a Frumkin isotherm with repulsive lateral interactions can be mapped into a series of Langmuir isotherms with a DCE.
Figure 6.

(a) Differential double-layer capacitance (Cdl), adsorption capacitance of Had and OHad (CHad, COHad), and their sum (Ctot) for the double layer of Pt(111)–0.1 M HClO4 interface. Experimental data, shown in circles, are taken from ref (22). (b) Model-based distributions of equilibrium potential of H and OH adsorption. (c) Comparison of differential double-layer capacitance for three values of the net electron number of OHad. The case of ξOH = 0 corresponds to the case without chemisorption-induced dipole moment, μOH. Matlab script of this figure is provided in the Supporting Information of this paper.
Figure 6a indicates that Cdl is a small percentage of Ctot. Figure 6c shows Cdl for the cases with and without Δϕad caused by chemisorption. Three values of the net electron number of OHad, ξOH, are compared, including the base value of 0.12, a zero value corresponding to the case of a complete charge transfer during chemisorption and hence a zero Δϕad, and also a higher value of 0.20 to increase the magnitude of Δϕad. When Δϕad = 0, the Cdl curve contains a higher peak at ∼0.2 VRHE, and a much depressed peak at potentials more positive than Epzc = 0.29 VSHE (0.346 VRHE for this case), similar to the results shown in Figure 5b. In comparison, Cdl drops above 0.5 VRHE due to a positive Δϕad when ξOH = 0.12. Particularly, negative Cdl is found when ξOH = 0.20.
Negative Cdl means that σfree decreases with increasing electrode potential, resulting in a nonmonotonic surface charging behavior. Chemisorption-induced nonmonotonic surface charging behaviors were revealed in a previous work72 and elucidated in a, hopefully, more accessible manner recently.80 The basic idea, as expressed in eq 1, is that negatively charged chemisorbed OH introduces a potential drop Δϕad, which is tantamount to increasing the pzc. As a result, σfree increases at a lower rate with the electrode potential, namely, a smaller Cdl, or decreases with the electrode potential, namely, a negative Cdl, depending on the magnitude of μOH.
Double Layer Region
Let us proceed to analyzing the recent experimental data by Ojha, Doblhoff-Dier, and Koper, focusing on the double layer region between 0.40 and 0.65 VRHE.22 In order to distance the pzc as much as possible from the hydrogen and hydroxyl adsorption/desorption regions, they delicately used 0.1 mM HClO4 as the base electrolyte solution with a pH of 4, where the pzc is 0.53 VRHE. They repeated the measurements on the Pt(111)–0.1 mM HClO4 interface three times, allowing us to examine how experimental uncertainties affect model results in Figure 7. Afterward, experimental data in the presence of different concentrations of LiClO4 are analyzed in Figure 8.
Figure 7.
(a–c) Model–experiment comparison for the electric double layer at Pt(111)–aqueous interfaces. The experiment on a Pt(111) electrode in 0.1 mM HClO4 was repeated three times.22 Model results are marked as solid lines and experimental data as circles. The total capacitance (Ctotal) is decomposed into the capacitance due to H adsorption, CHad, that is due to OH adsorption, COHad, and the pure double-layer capacitance, Cdl. (d–f) Model-based distributions of the equilibrium potential of H and OH adsorption. (g–i) Coverages of adsorbed H and OH. Matlab script of this figure is provided in the Supporting Information of this paper.
Figure 8.

(a) Model–experiment comparison in the presence of different concentrations of LiClO4 for the electric double layer at Pt(111) in 0.1 mM HClO4 + x mM LiClO4.22 Model results are marked as solid lines and experimental data as circles. For the sake of clarity, a vertical shift of 50 between different lines is used in panel (a). Coverages of adsorbed (b) Had and OHad, (c) the pure double-layer capacitance, Cdl, and (d) adsorption capacitances CHad and COHad. Matlab script of this figure is provided in the Supporting Information of this paper.
Figure 7a–c shows the model–experiment comparison for three pieces of capacitance data in 0.1 mM HClO4. The EDL parameters are fixed at their base values, and the free fitting parameters are the weight pk, expectations μk, and variances σk of distribution functions of Had and OHad. One Gaussian distribution for Had and two for OHad are sufficient in this narrow potential region. The fitted distributions are displayed in Figure 7d–f. The coverages of Had and OHad are shown in Figure 7g–i.
In Figure 7a–c, solid lines represent model results, and circles represent experimental data. The total capacitance (Ctotal) is further decomposed, with the aid of the model, into the capacitance due to H adsorption, CHad, that is due to OH adsorption, COHad, and the pure double-layer capacitance, Cdl. The measured capacitance Ctotal is contributed majorly by pseudo-capacitances CHad and COHad. Three measurements show a moderate difference in between: the terrace in the potential range of 0.55–0.6 VRHE, attributed to water adsorption in ref (22), almost disappears in the third measurement.
The present model attributes different profiles of Ctotal to variations in the distributed equilibrium potential of OHad, fOH, as shown in Figure 7d–f. The OH adsorption has a bimodal normal distribution and the one at lower potentials with p1 < 6% is responsible for the terrace of Ctotal in the potential range of 0.55–0.6 VRHE. The terrace appears when the two normal distributions are well resolved, as in Figure 7d,e, and disappears when they come closer and become connected, as in Figure 7f. The variations in fOH imply that the assumed distribution of equilibrium potential is likely affected by stochastic factors.
The coverages of adsorbed Had and OHad, denoted θH and θOH, are less than 5% in the double-layer region. However, such a small amount of adsorption results in adsorption capacitance several times higher than Cdl, which demonstrates the importance of and also difficulty in separating Cdl from adsorption capacitances for electrocatalytic double layers.
Experimental data in the presence of different concentrations of LiClO4 are analyzed in Figure 8. The experiments were conducted on Pt(111) in 0.1 mM HClO4 + x mM LiClO4 (x = 0.1,0.3, 0.6, 1.0).22 Experimental capacitance curves are shifted vertically in Figure 8a for visual clarity. The model results in solid lines neatly agree with experimental data in circles in the examined potential region. The fitted coverages of Had and OHad, the pure double layer capacitance, and pseudo-capacitances of chemisorption are shown in Figure 8b–d, respectively. The correspondence between line colors and ionic concentrations indicated in Figure 8a is retained in all subfigures. As for the pure double-layer capacitance Cdl in Figure 8c, a Gouy–Chapman minimum is found at 0.53 VRHE. The asymmetry around the pzc is caused by water adsorption. As shown in Figure 5a, a negative value of δμ̃, namely, the O-down configuration is more favorable, results in a more tilted increase in Cdl when the electrode potential moves away from the pzc in the more negative direction. The valley and the terrace of the measured Ctotal shifts to lower potentials with increasing concentration of LiClO4. In the present model, the shift in the capacitance minimum and terrace is associated with adsorption of OH, as shown in Figure 8d. Possible reasons for this are discussed in the next section.
Discussion
Is There a Pure Double Layer Region at Pt(111)?
In contrast with previous works,22,51,52 the present model maintains that voltammogram-derived capacitance of Pt(111) is majorly contributed by CHad and COHad from a small percentage of active sites, Figure 8. One of the implications of this claim is that there is no pure double-layer region for the Pt(111) electrode.
The occurrence of chemisorbed OHad in the so-called double layer potential region could have two causes. On the one hand, there may be structural defects occupying a small percentage of the solid surface.103,104 Adsorption of OH (H) occurs at lower (higher) potentials at these structural defects. In a study by Góez-Marín and Feliu,105 they experimentally investigated the relation between oxygen reduction activity and step density and extrapolated this relation to obtain the oxygen reduction activity of an ideally defect-free Pt(111). They found that the extrapolated activity of this ideal Pt(111) is lower than that of a real Pt(111). This comparison leads them to conclude that “This is true even for those electrodes that do not apparently show any step contribution signal in the blank voltammogram, pointing out that a real electrode has always surface defects.” Recent works on stepped Pt electrodes, such as Pt(553)106 and Pt(311),107 have shown low-potential water dissociation in the hydrogen adsorption/desorption region, and the onset potential moves toward the “double layer” region at higher pHs.
On the other hand, interfacial water molecules have an energy distribution, as seen in AIMD simulations,108 and higher-energy water molecules are decomposed and then adsorbed as OHad at lower potentials. Such a distribution, intrinsically stochastic to a certain extent, echoes the experimental uncertainties found in three repeated measurements on the Pt(111)–0.1 mM HClO4 interface. Model results in Figure 8 further indicate that a higher concentration of LiClO4 promotes earlier adsorption of OH. Berna et al. have adopted a similar idea to interpret chemisorption of OH at Pt(111).109 Specifically, they attributed the broad peak in the low potential region and the sharp peak in the high potential region to the dissociation of two kinds of water molecules with different electrochemical potentials. Their treatment could be regarded as a special case of this model where the distribution function is actually composed of two delta functions.
Currently, it is hard to detect adsorbed intermediates at such a low coverage. Nevertheless, Kondo et al. deduced from their in situ surface X-ray scattering data that a small amount of oxygen-containing species (unspecified) and/or ClO4– is adsorbed on Pt(111) immersed in 0.1 M HClO4 at 0.57 VRHE within the double-layer region.110 It has also been shown that the ClO4 concentration impacts the CV of Pt(111) in the double layer region.110,111 Though I have assumed Had and OHad as the chemisorbates when explaining the experimental data of Ojha et al., the existence of other chemisorbates such as ClO4– and organic impurities cannot be excluded for this moment.110,111
Limitations and Possible Extensions of the Model
The present double layer model belongs to GCS-type models. An essential feature of these models is the decomposition of the double layer into several subregions, which are then described as dielectric continua. I am aware of the harsh criticisms on such GCS-type models regarding drastic simplifications of the double layer structure and its parameterization procedure that arbitrariness may slip in.112 Nevertheless, it is undeniable that GCS-type models are simple yet effective in describing physicochemical properties of the double layer. Fortunately, accurate atomic details provided by careful first-principles calculations can help to avoid unphysical assumptions, omission of key factors, overparameterizaton and error cancellation, etc., in phenomenological modeling. Having said this, I stress that first-principles approaches have its own purpose, rather than just serving to elaborate phenomenological models.
Though the model results are obtained for a Pt(111) electrode in HClO4-based solution, I do not see principal difficulties in applying it to the alkaline regime, to other electrolyte solution with specifically adsorbing anions such as sulfate and halide ions and to stepped Pt electrodes. In these new scenarios, the model needs to be extended by including new specific adsorption and/or chemisorption processes and to be re-parameterized on a case-by-case basis.
The model can be extended in the following directions in the future. First, the locations of chemisorbed ions and the first water layer, which are fixed in the present work, can be allowed to change as a function of electrode potential. Similar considerations of relaxation of the metal–solution gap have been shown to be able to bring forth negative Cdl.113 Second, the model may also be extended to consider how electronic effects of chemisorbed ions and water molecules change χM.65,114 Third, a constant δμ̃ has been used so far, and it is meaningful to consider a potential-dependent change of δμ̃, caused by interactions between chemisorbed ions and water molecules on the one hand and orientational dependency of water binding energy on the other.115 Lastly, noncovalent interactions between chemisorbed ions and electrolyte species, key to understanding cation effects as shown in a recent study,116 are also important factors to be treated. Lastly, the model can also be extended to study other chemisorption processes, including those of electrolyte anions and organic impurities, at different types of active sites.
Conclusions
A modified double layer model has been formulated for metal–aqueous interfaces in electrocatalytic systems. The present model divides the inner Helmholtz plane (IHP) further into an AIP (adsorbed ion plane) and ASP (adsorbed solvent plane), accounting for different locations of chemisorbed ions and solvent molecules. The present model employs a statistical treatment of the first-layer water molecules with chemical interactions imposed by the metal and a distributed treatment of chemisorption of electrolyte species.
The model shows that the potential of zero charge, pzc for short, depends on metal–water chemical interactions. I introduce δμ to characterize the chemical interaction energy difference between water dipoles of αw = 0° and that of αw = 180°, where αw is the angle between water dipole direction and the local electric field. It is found that the pzc becomes higher (lower) when δμ > 0 (<0). In addition to a shift in the pzc, the first-layer water molecules also dramatically elevate the double-layer capacitance Cdl. Equivalently, it means that the first-layer water molecules contribute a negative capacitance component in series with other components, which is a known effect. However, the formula derived in this work to describe this effect, eq 26, is, to my knowledge, a new contribution. The model also shows that chemisorption-induced dipole moment could lead to a new valley of the capacitance curve and the capacitance in the valley could take negative values.
The model is then employed to interpret recent CV-based capacitance data of Pt(111)–aqueous solution interfaces. The explanation is based on the ansatz that the experimental data are not the pure double-layer capacitance but include pseudo-capacitances of chemisorption of electrolyte species. This ansatz resolves the discrepancy that the CV-based capacitance values are several times higher than the values obtained using electrochemical impedance spectroscopy. In addition, it constitutes an alternative angle of understanding stochastic uncertainties found in these experiments. The ansatz and the resulting explanation, though being appealing, must be taken with care, as it implies that there is no pure double layer region even for Pt(111), an aspect that is yet to be confirmed.
Before closing, three experiments to confirm this ansatz are sketched here. One can examine the frequency dependence of interfacial capacitance in the double layer region, especially, at the pzc. If it is a pure double-layer capacitance, there should not be any frequency dispersion below the characteristic frequency of EDL capacitance, which is expressed as fdl ∝ D/λD2 with D being the diffusion coefficient of electrolyte ions and λD being the Debye length, according to an exact analytical solution developed in ref (117). With typical values of D = 10–9 m2s–1 and λD = 1 – 10 nm, fdl has a range of 107 – 109 Hz. It is thus safe to claim that a pure double-layer capacitance should not depend on frequency below 1 MHz. In other words, shall one observe that the interfacial capacitance decreases with increasing frequency, it is very likely that the interfacial capacitance contains pseudo-capacitance(s). If it is difficult to obtain reliable high-frequency impedance data, a time-domain method that is easier to implement is to measure the current response to a potential step from the pzc by a small amount, say 20 mV. If it is pure double layer charging, the time constant, expressed as tdl ∝ 10LλD/D with L being the thickness of the diffusion layer,118 has a range of 1 – 10 ms using typical values of D = 10–9 m2s–1, L = 100 μm, and λD = 1 – 10 nm. In other words, shall one observe a current ramping with a time constant longer than 10 ms, it is very likely that other slower process(es) is (are) involved. The third experiment involves ultrafast CV measurements, which have been demonstrated on the order of megavolt per second,119 and then check the scan-rate dependency of the obtained capacitance profiles. This would require careful ohmic drop compensation and noise control.
Acknowledgments
This work is supported by the Alexander von Humboldt Foundation and the Helmholtz Young Investigators Group Leader Programme. I am grateful to Professors Gary Attard, Juan Feliu, Jianbo Zhang, Katharina Doblhoff-Dier, Ms. Lulu Zhang, and Mr. Jinwen Liu for helpful discussions.
Data Availability Statement
No new experimental data. The scripts used to calculate the model results are available from the author upon request.
Supporting Information Available
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/jacsau.2c00650.
Matlab scripts for reproducing results reported in this paper (PDF)
The author declares no competing financial interest.
Supplementary Material
References
- Steinmann S. N.; Seh Z. W. Understanding electrified interfaces. Nat. Rev. Mater. 2021, 6, 289–291. 10.1038/s41578-021-00303-1. [DOI] [Google Scholar]
- Sebastián-Pascual P.; Shao-Horn Y.; Escudero-Escribano M. Toward understanding the role of the electric double layer structure and electrolyte effects on well-defined interfaces for electrocatalysis. Curr. Opin. Electrochem. 2022, 32, 100918. 10.1016/j.coelec.2021.100918. [DOI] [Google Scholar]
- Shin S.-J.; Kim D. H.; Bae G.; Ringe S.; Choi H.; Lim H.-K.; Choi C. H.; Kim H. On the importance of the electric double layer structure in aqueous electrocatalysis. Nat. Commun. 2022, 13, 174. 10.1038/s41467-021-27909-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Haid R. W.; Ding X.; Sarpey T. K.; Bandarenka A. S.; Garlyyev B. Exploration of the electrical double-layer structure: Influence of electrolyte components on the double-layer capacitance and potential of maximum entropy. Curr. Opin. Electrochem. 2022, 32, 100882. 10.1016/j.coelec.2021.100882. [DOI] [Google Scholar]
- Kozawa A. Effects of anions and cations on oxygen reduction and oxygen evolution reactions on platinum electrodes. J. Electroanal. Chem. (1959) 1964, 8, 20–39. 10.1016/0022-0728(64)80035-8. [DOI] [Google Scholar]
- Strmcnik D.; Kodama K.; van der Vliet D.; Greeley J.; Stamenkovic V. R.; Marković N. M. The role of non-covalent interactions in electrocatalytic fuel-cell reactions on platinum. Nat. Chem. 2009, 1, 466–472. 10.1038/nchem.330. [DOI] [PubMed] [Google Scholar]
- Strmcnik D.; van der Vliet D. F.; Chang K. C.; Komanicky V.; Kodama K.; You H.; Stamenkovic V. R.; Marković N. M. Effects of Li+, K+, and Ba2+ Cations on the ORR at Model and High Surface Area Pt and Au Surfaces in Alkaline Solutions. J. Phys. Chem. Lett. 2011, 2, 2733–2736. 10.1021/jz201215u. [DOI] [Google Scholar]
- Suntivich J.; Perry E. E.; Gasteiger H. A.; Shao-Horn Y. The Influence of the Cation on the Oxygen Reduction and Evolution Activities of Oxide Surfaces in Alkaline Electrolyte. Electrocatalysis 2013, 4, 49–55. 10.1007/s12678-012-0118-x. [DOI] [Google Scholar]
- Resasco J.; Chen L. D.; Clark E.; Tsai C.; Hahn C.; Jaramillo T. F.; Chan K.; Bell A. T. Promoter Effects of Alkali Metal Cations on the Electrochemical Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2017, 139, 11277–11287. 10.1021/jacs.7b06765. [DOI] [PubMed] [Google Scholar]
- Li M. F.; Liao L. W.; Yuan D. F.; Mei D.; Chen Y.-X. pH effect on oxygen reduction reaction at Pt(111) electrode. Electrochim. Acta 2013, 110, 780–789. 10.1016/j.electacta.2013.04.096. [DOI] [Google Scholar]
- Briega-Martos V.; Herrero E.; Feliu J. M. Effect of pH and Water Structure on the Oxygen Reduction Reaction on platinum electrodes. Electrochim. Acta 2017, 241, 497–509. 10.1016/j.electacta.2017.04.162. [DOI] [Google Scholar]
- Ledezma-Yanez I.; Wallace W. D. Z.; Sebastián-Pascual P.; Climent V.; Feliu J. M.; Koper M. T. M. Interfacial water reorganization as a pH-dependent descriptor of the hydrogen evolution rate on platinum electrodes. Nat. Energy 2017, 2, 17031. 10.1038/nenergy.2017.31. [DOI] [Google Scholar]
- Rao R. R.; Huang B.; Katayama Y.; Hwang J.; Kawaguchi T.; Lunger J. R.; Peng J.; Zhang Y.; Morinaga A.; Zhou H.; You H.; Shao-Horn Y. pH- and Cation-Dependent Water Oxidation on Rutile RuO2(110). J. Phys. Chem. C 2021, 125, 8195–8207. 10.1021/acs.jpcc.1c00413. [DOI] [Google Scholar]
- Waegele M. M.; Gunathunge C. M.; Li J.; Li X. How cations affect the electric double layer and the rates and selectivity of electrocatalytic processes. J. Chem. Phys. 2019, 151, 160902. 10.1063/1.5124878. [DOI] [PubMed] [Google Scholar]
- Colic V.; Pohl M. D.; Scieszka D.; Bandarenka A. S. Influence of the electrolyte composition on the activity and selectivity of electrocatalytic centers. Catal. Today 2016, 262, 24–35. 10.1016/j.cattod.2015.08.003. [DOI] [Google Scholar]
- Moura de Salles Pupo M.; Kortlever R. Electrolyte Effects on the Electrochemical Reduction of CO2. ChemPhysChem 2019, 20, 2926–2935. 10.1002/cphc.201900680. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Deng B.; Huang M.; Zhao X.; Mou S.; Dong F. Interfacial Electrolyte Effects on Electrocatalytic CO2 Reduction. ACS Catal. 2022, 12, 331–362. 10.1021/acscatal.1c03501. [DOI] [Google Scholar]
- Pajkossy T.; Kolb D. M. Double layer capacitance of Pt(111) single crystal electrodes. Electrochim. Acta 2001, 46, 3063–3071. 10.1016/S0013-4686(01)00597-7. [DOI] [Google Scholar]
- Pajkossy T. Voltammetry and Impedance of Pt(111) Electrodes in Aqueous KClO4 Solutions. Zeitschrift für Physikalische Chemie 2003, 217, 351. 10.1524/zpch.217.4.351.20382. [DOI] [Google Scholar]
- Pajkossy T.; Kolb D. M. On the origin of the double layer capacitance maximum of Pt(111) single crystal electrodes. Electrochem. Commun. 2003, 5, 283–285. 10.1016/S1388-2481(03)00046-8. [DOI] [Google Scholar]
- Ojha K.; Arulmozhi N.; Aranzales D.; Koper M. T. M. Double Layer at the Pt(111)–Aqueous Electrolyte Interface: Potential of Zero Charge and Anomalous Gouy–Chapman Screening. Angew. Chem. 2020, 59, 711–715. 10.1002/anie.201911929. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ojha K.; Doblhoff-Dier K.; Koper Marc T. M. Double-layer structure of the Pt(111)–aqueous electrolyte interface. Proc. Natl. Acad. Sci. 2022, 119, e2116016119 10.1073/pnas.2116016119. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xue S.; Garlyyev B.; Auer A.; Kunze-Liebhäuser J.; Bandarenka A. S. How the Nature of the Alkali Metal Cations Influences the Double-Layer Capacitance of Cu, Au, and Pt Single-Crystal Electrodes. J. Phys. Chem. C 2020, 124, 12442–12447. 10.1021/acs.jpcc.0c01715. [DOI] [Google Scholar]
- Schmickler W.; Santos E., Interfacial electrochemistry. Springer-Verlag Berlin Heidelberg: New York, 2010; p XII, 270. [Google Scholar]
- Guidelli R.; Schmickler W. Recent developments in models for the interface between a metal and an aqueous solution. Electrochim. Acta 2000, 45, 2317–2338. 10.1016/S0013-4686(00)00335-2. [DOI] [Google Scholar]
- Damaskin B. B.; Petrii O. A. Historical development of theories of the electrochemical double layer. J. Solid State Electrochem. 2011, 15, 1317–1334. 10.1007/s10008-011-1294-y. [DOI] [Google Scholar]
- Zhang L. L.; Li C. K.; Huang J. A Beginners’ Guide to Modelling of Electric Double Layer Under Equilibrium, Nonequilibrium and AC Conditions. J. Electrochem. 2022, 28, 2108471. 10.13208/j.electrochem.210847. [DOI] [Google Scholar]
- Kornyshev A. A. Double-Layer in Ionic Liquids: Paradigm Change?. J. Phys. Chem. B 2007, 111, 5545–5557. 10.1021/jp067857o. [DOI] [PubMed] [Google Scholar]
- Budkov Y. A.; Kolesnikov A. L. Electric double layer theory for room temperature ionic liquids on charged electrodes: Milestones and prospects. Curr. Opin. Electrochem. 2022, 33, 100931. 10.1016/j.coelec.2021.100931. [DOI] [Google Scholar]
- Gouy G. Sur La Constitution De La Charge Electrique A La Surface D’Un Electrolyte. J. Phys. Theor. Appl. 1910, 9, 457–468. 10.1051/jphystap:019100090045700. [DOI] [Google Scholar]
- Grahame D. C. The Electrical Double Layer and the Theory of Electrocapillarity. Chem. Rev. 1947, 41, 441–501. 10.1021/cr60130a002. [DOI] [PubMed] [Google Scholar]
- Parsons R. The electrical double layer: recent experimental and theoretical developments. Chem. Rev. 1990, 90, 813–826. 10.1021/cr00103a008. [DOI] [Google Scholar]
- Mott N. F.; Watts-Tobin R. J. The interface between a metal and an electrolyte. Electrochim. Acta 1961, 4, 79–107. 10.1016/0013-4686(61)80009-1. [DOI] [Google Scholar]
- Watts-tobin R. J. The interface between a metal and an electrolytic solution. Philos. Mag. 1961, 6, 133–153. 10.1080/14786436108238358. [DOI] [Google Scholar]
- Macdonald J. R.; Barlow C. A. Theory of Double-Layer Differential Capacitance in Electrolytes. J. Chem. Phys. 1962, 36, 3062–3080. 10.1063/1.1732426. [DOI] [Google Scholar]
- Bockris J. O. M.; Habib M. A. Contributions of water dipoles to double layer properties: A three-state water model. Electrochim. Acta 1977, 22, 41–46. 10.1016/0013-4686(77)85051-2. [DOI] [Google Scholar]
- Levine S.; Bell G. M.; Calvert D. THE DISCRETENESS-OF-CHARGE EFFECT IN ELECTRIC DOUBLE LAYER THEORY. Can. J. Chem. 1962, 40, 518–538. 10.1139/v62-080. [DOI] [Google Scholar]
- Fawcett W. R.; Levine S. Discreteness-of-charge effects in electrode kinetics. J. Electroanal. Chem. Interfacial Electrochem. 1973, 43, 175–184. 10.1016/S0022-0728(73)80488-7. [DOI] [Google Scholar]
- Grahame D. C. Discreteness-of-charge-effects in the inner region of the electrical double layer. Zeitschrift für Elektrochemie, Berichte der Bunsengesellschaft für physikalische Chemie 1958, 62, 264–274. 10.1002/bbpc.19580620309. [DOI] [Google Scholar]
- Fawcett W. R. Fifty years of studies of double layer effects in electrode kinetics—a personal view. J. Solid State Electrochem. 2011, 15, 1347. 10.1007/s10008-011-1337-4. [DOI] [Google Scholar]
- Badiali J. P.; Rosinberg M. L.; Goodisman J. Contribution of the metal to the differential capacity of an ideally polarisable electrode. J. Electroanal. Chem. 1983, 143, 73–88. 10.1016/S0022-0728(83)80255-1. [DOI] [Google Scholar]
- Badiali J. P.; Rosinberg M. L.; Vericat F.; Blum L. A microscopic model for the liquid metal-ionic solution interface. J. Electroanal. Chem. 1983, 158, 253–267. 10.1016/S0022-0728(83)80611-1. [DOI] [Google Scholar]
- Schmickler W. A jellium-dipole model for the double layer. J. Electroanal. Chem. 1983, 150, 19–24. 10.1016/S0022-0728(83)80185-5. [DOI] [Google Scholar]
- Kornyshev A. A. Metal electrons in the double layer theory. Electrochim. Acta 1989, 34, 1829–1847. 10.1016/0013-4686(89)85070-4. [DOI] [Google Scholar]
- Valette G. Double layer on silver single-crystal electrodes in contact with electrolytes having anions which present a slight specific adsorption: Part I. The (110) face. J. Electroanal. Chem. Interfacial Electrochem. 1981, 122, 285–297. 10.1016/S0022-0728(81)80159-3. [DOI] [Google Scholar]
- Valette G. Double layer on silver single crystal electrodes in contact with electrolytes having anions which are slightly specifically adsorbed: Part II. The (100) face. J. Electroanal. Chem. Interfacial Electrochem. 1982, 138, 37–54. 10.1016/0022-0728(82)87126-X. [DOI] [Google Scholar]
- Valette G. Double layer on silver single crystal electrodes in contact with electrolytes having anions which are slightly specifically adsorbed: Part III. The (111) face. J. Electroanal. Chem. Interfacial Electrochem. 1989, 269, 191–203. 10.1016/0022-0728(89)80112-3. [DOI] [Google Scholar]
- Hamelin A.; Foresti M. L.; Guidelli R. Test of the Gouy-Chapman theory at a (111) silver single-crystal electrode. J. Electroanal. Chem. 1993, 346, 251–259. 10.1016/0022-0728(93)85016-A. [DOI] [Google Scholar]
- Pajkossy T.; Kolb D. M. Double layer capacitance of the platinum group metals in the double layer region. Electrochem. Commun. 2007, 9, 1171–1174. 10.1016/j.elecom.2007.01.002. [DOI] [Google Scholar]
- Schouten K. J. P.; van der Niet M. J. T. C.; Koper M. T. M. Impedance spectroscopy of H and OH adsorption on stepped single-crystal platinum electrodes in alkaline and acidic media. Phys. Chem. Chem. Phys. 2010, 12, 15217–15224. 10.1039/c0cp00104j. [DOI] [PubMed] [Google Scholar]
- Doblhoff-Dier K.; Koper M. T. M. Modeling the Gouy–Chapman Diffuse Capacitance with Attractive Ion–Surface Interaction. J. Phys. Chem. C 2021, 125, 16664–16673. 10.1021/acs.jpcc.1c02381. [DOI] [Google Scholar]
- Schmickler W. The Effect of Weak Adsorption on the Double Layer Capacitance. ChemElectroChem 2021, 8, 4218–4222. 10.1002/celc.202100842. [DOI] [Google Scholar]
- Le J.-B.; Fan Q.-Y.; Li J.-Q.; Cheng J. Molecular origin of negative component of Helmholtz capacitance at electrified Pt(111)/water interface. Sci. Adv. 2021, 6, eabb1219 10.1126/sciadv.abb1219. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sibert E.; Faure R.; Durand R. High frequency impedance measurements on Pt(111) in sulphuric and perchloric acids. J. Electroanal. Chem. 2001, 515, 71–81. 10.1016/S0022-0728(01)00639-8. [DOI] [Google Scholar]
- Grahame D. C. Components of Charge and Potential in the Non-diffuse Region of the Electrical Double Layer: Potassium Iodide Solutions in Contact with Mercury at 25°1. J. Am. Chem. Soc. 1958, 80, 4201–4210. 10.1021/ja01549a022. [DOI] [Google Scholar]
- Grahame D. C.; Parsons R. Components of Charge and Potential in the Inner Region of the Electrical Double Layer: Aqueous Potassium Chloride Solutions in Contact with Mercury at 25°. J. Am. Chem. Soc. 1961, 83, 1291–1296. 10.1021/ja01467a009. [DOI] [Google Scholar]
- Damaskin B.; Pankratova I.; Palm U.; Anni K.; Väärtnǒu M.; Salve M. Comparison of the ionic adsorption theories for the electrode/solution interface by computer simulation. J. Electroanal. Chem. Interfacial Electrochem. 1987, 234, 31–53. 10.1016/0022-0728(87)80160-2. [DOI] [Google Scholar]
- Kolotyrkin Y. M.; Alexeyev Y. V.; Popov Y. A. Double layer model taking into account the specific adsorption of ions: Application to the process of hydrogen adsorption on lead in the presence of iodide ion. J. Electroanal. Chem. Interfacial Electrochem. 1975, 62, 135–149. 10.1016/0022-0728(75)80032-5. [DOI] [Google Scholar]
- Vorotyntsev M. Chemisorption theory for charged species at electrodes in the model of an energetically homogeneous surface. J. Res. Inst. Catal., Hokkaido Univ. 1983, 30, 167–177. [Google Scholar]
- Guidelli R., Monolayer Models of Metal-Water Interphases and their use in the Interpretation of Differential Capacity Curves and of Organic Adsorption. In Trends in Interfacial Electrochemistry, Silva A. F., Ed. Springer Netherlands: Dordrecht, 1986; pp. 387–452, 10.1007/978-94-009-4694-1_15. [DOI] [Google Scholar]
- Trasatti S. Work function, electronegativity, and electrochemical behaviour of metals: II. Potentials of zero charge and “electrochemical” work functions. J. Electroanal. Chem. Interfacial Electrochem. 1971, 33, 351–378. 10.1016/S0022-0728(71)80123-7. [DOI] [Google Scholar]
- Damaskin B. B.; Frumkin A. N. Potentials of zero charge, interaction of metals with water and adsorption of organic substances—III. The role of the water dipoles in the structure of the dense part of the electric double layer. Electrochim. Acta 1974, 19, 173–176. 10.1016/0013-4686(74)85013-9. [DOI] [Google Scholar]
- Bockris J. O. M.; Devanathan M. A. V.; Müller K. On the structure of charged interfaces. Proc. Roy. Soc. A 1963, 274, 55–79. 10.1016/B978-1-4831-9831-6.50068-0. [DOI] [Google Scholar]
- Sakong S.; Forster-Tonigold K.; Groß A. The structure of water at a Pt(111) electrode and the potential of zero charge studied from first principles. J. Chem. Phys. 2016, 144, 194701. 10.1063/1.4948638. [DOI] [PubMed] [Google Scholar]
- Le J.; Iannuzzi M.; Cuesta A.; Cheng J. Determining Potentials of Zero Charge of Metal Electrodes versus the Standard Hydrogen Electrode from Density-Functional-Theory-Based Molecular Dynamics. Phys. Rev. Lett. 2017, 119, 016801 10.1103/PhysRevLett.119.016801. [DOI] [PubMed] [Google Scholar]
- Sakong S.; Groß A. The electric double layer at metal-water interfaces revisited based on a charge polarization scheme. J. Chem. Phys. 2018, 149, 084705 10.1063/1.5040056. [DOI] [PubMed] [Google Scholar]
- Le J.-B.; Chen A.; Li L.; Xiong J.-F.; Lan J.; Liu Y.-P.; Iannuzzi M.; Cheng J. Modeling Electrified Pt(111)-Had/Water Interfaces from Ab Initio Molecular Dynamics. JACS Au 2021, 1, 569–577. 10.1021/jacsau.1c00108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Groß A.; Sakong S. Ab Initio Simulations of Water/Metal Interfaces. Chem. Rev. 2022, 10746. 10.1021/acs.chemrev.1c00679. [DOI] [PubMed] [Google Scholar]
- Schmickler W.; Guidelli R. The partial charge transfer. Electrochim. Acta 2014, 127, 489–505. 10.1016/j.electacta.2014.02.057. [DOI] [Google Scholar]
- Carnie S. L.; Chan D. Y. C. The modelling of solvent structure in the electrical double layer. Adv. Colloid Interface Sci. 1982, 16, 81–100. 10.1016/0001-8686(82)85009-4. [DOI] [Google Scholar]
- Schmickler W.; Guidelli R. Ionic adsorption and the surface dipole potential. J. Electroanal. Chem. Interfacial Electrochem. 1987, 235, 387–392. 10.1016/0022-0728(87)85223-3. [DOI] [Google Scholar]
- Huang J.; Malek A.; Zhang J.; Eikerling M. H. Non-monotonic Surface Charging Behavior of Platinum: A Paradigm Change. J. Phys. Chem. C 2016, 120, 13587–13595. 10.1021/acs.jpcc.6b03930. [DOI] [Google Scholar]
- Huang J.; Zhou T.; Zhang J.; Eikerling M. Double layer of platinum electrodes: Non-monotonic surface charging phenomena and negative double layer capacitance. J. Chem. Phys. 2018, 148, 044704 10.1063/1.5010999. [DOI] [PubMed] [Google Scholar]
- Zhang M.-K.; Cai J.; Chen Y.-X. On the electrode charge at the metal/solution interface with specific adsorption. Curr. Opin. Electrochem. 2022, 36, 101161. 10.1016/j.coelec.2022.101161. [DOI] [Google Scholar]
- Frumkin A. N.; Petrii O. A. Potentials of zero total and zero free charge of platinum group metals. Electrochim. Acta 1975, 20, 347–359. 10.1016/0013-4686(75)90017-1. [DOI] [Google Scholar]
- Damaskin B. B.; Safonov V. A.; Petrii O. A. Model of two limiting states for describing the properties of the electric double layer in the absence of specific adsorption of ions. J. Electroanal. Chem. Interfacial Electrochem. 1989, 258, 13–25. 10.1016/0022-0728(89)85158-7. [DOI] [Google Scholar]
- Safonov V. A.; Krivenko A. G.; Choba M. A. Initial oxidation of silver electrode in weakly acidic sodium fluoride solutions: Potential shifts induced by laser heating. Electrochim. Acta 2008, 53, 4859–4866. 10.1016/j.electacta.2008.02.003. [DOI] [Google Scholar]
- Garcia-Araez N.; Climent V.; Feliu J. Potential-Dependent Water Orientation on Pt(111), Pt(100), and Pt(110), As Inferred from Laser-Pulsed Experiments. Electrostatic and Chemical Effects. J. Phys. Chem. C 2009, 113, 9290–9304. 10.1021/jp900792q. [DOI] [Google Scholar]
- Sarabia F. J.; Sebastián P.; Climent V.; Feliu J. M. New insights into the Pt(hkl)-alkaline solution interphases from the laser induced temperature jump method. J. Electroanal. Chem. 2020, 872, 114068. 10.1016/j.jelechem.2020.114068. [DOI] [Google Scholar]
- Huang J. Surface charging behaviors of electrocatalytic interfaces with partially charged chemisorbates. Curr. Opin. Electrochem. 2022, 33, 100938. 10.1016/j.coelec.2022.100938. [DOI] [Google Scholar]
- Gongadze E.; Iglič A. Decrease of permittivity of an electrolyte solution near a charged surface due to saturation and excluded volume effects. Bioelectrochemistry 2012, 87, 199–203. 10.1016/j.bioelechem.2011.12.001. [DOI] [PubMed] [Google Scholar]
- Levy A.; Andelman D.; Orland H. Dipolar Poisson-Boltzmann approach to ionic solutions: A mean field and loop expansion analysis. J. Chem. Phys. 2013, 139, 164909. 10.1063/1.4826103. [DOI] [PubMed] [Google Scholar]
- Braunwarth L.; Jung C.; Jacob T. Potential-Dependent Pt(111)/Water Interface: Tackling the Challenge of a Consistent Treatment of Electrochemical Interfaces**. ChemPhysChem 2022, n/a, e202200336 10.26434/chemrxiv-2022-xndvx. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kastlunger G.; Lindgren P.; Peterson A. A. Controlled-Potential Simulation of Elementary Electrochemical Reactions: Proton Discharge on Metal Surfaces. J. Phys. Chem. C 2018, 122, 12771–12781. 10.1021/acs.jpcc.8b02465. [DOI] [Google Scholar]
- Foresti M. L.; Innocenti M.; Forni F.; Guidelli R. Electrosorption Valency and Partial Charge Transfer in Halide and Sulfide Adsorption on Ag(111). Langmuir 1998, 14, 7008–7016. 10.1021/la980692t. [DOI] [Google Scholar]
- Zhang Y.; Huang J. Treatment of Ion-Size Asymmetry in Lattice-Gas Models for Electrical Double Layer. J. Phys. Chem. C 2018, 122, 28652–28664. 10.1021/acs.jpcc.8b08298. [DOI] [Google Scholar]
- Qing L.; Jiang J. Double-Edged Sword of Ion-Size Asymmetry in Energy Storage of Supercapacitors. J. Phys. Chem. Lett. 2022, 13, 1438–1445. 10.1021/acs.jpclett.1c03900. [DOI] [PubMed] [Google Scholar]
- Bossa G. V.; Caetano D. L. Z.; de Carvalho S. J.; May S. Differential capacitance of an electrical double layer with asymmetric ion sizes in the presence of hydration interactions. Electrochim. Acta 2019, 321, 134655. 10.1016/j.electacta.2019.134655. [DOI] [Google Scholar]
- Gongadze E.; Iglič A. Asymmetric size of ions and orientational ordering of water dipoles in electric double layer model - an analytical mean-field approach. Electrochim. Acta 2015, 178, 541–545. 10.1016/j.electacta.2015.07.179. [DOI] [Google Scholar]
- Garcia-Araez N.; Climent V.; Feliu J. M. Analysis of temperature effects on hydrogen and OH adsorption on Pt(111), Pt(100) and Pt(110) by means of Gibbs thermodynamics. J. Electroanal. Chem. 2010, 649, 69–82. 10.1016/j.jelechem.2010.01.024. [DOI] [Google Scholar]
- Climent V.; Góez R.; Orts J. M.; Feliu J. M. Thermodynamic Analysis of the Temperature Dependence of OH Adsorption on Pt(111) and Pt(100) Electrodes in Acidic Media in the Absence of Specific Anion Adsorption. J. Phys. Chem. B 2006, 110, 11344–11351. 10.1021/jp054948x. [DOI] [PubMed] [Google Scholar]
- Hörmann N. G.; Reuter K. Thermodynamic cyclic voltammograms: peak positions and shapes. J. Phys.: Condens. Matter 2021, 33, 264004. 10.1088/1361-648X/abf7a1. [DOI] [PubMed] [Google Scholar]
- Huang J. Hybrid density-potential functional theory of electric double layers. Electrochim. Acta 2021, 389, 138720. 10.1016/j.electacta.2021.138720. [DOI] [Google Scholar]
- Trasatti S. The “absolute” electrode potential—the end of the story. Electrochim. Acta 1990, 35, 269–271. 10.1016/0013-4686(90)85069-Y. [DOI] [Google Scholar]
- Melander M.; Wu T.; Honkala K. Constant inner potential DFT for modelling electrochemical systems under constant potential and bias. ChemRxiv 2022, This content is a preprint and has not been peer-reviewed. 10.26434/chemrxiv-2021-r621x-v2. [DOI] [Google Scholar]
- Putintsev N. M.; Putintsev D. N. High-frequency dielectric permittivity of water and its components. Russ. J. Phys. Chem. A 2011, 85, 1113–1118. 10.1134/S0036024411070272. [DOI] [Google Scholar]
- Schnur S.; Groß A. Properties of metal–water interfaces studied from first principles. New J. Phys. 2009, 11, 125003. 10.1088/1367-2630/11/12/125003. [DOI] [Google Scholar]
- Sakong S.; Groß A. Water structures on a Pt(111) electrode from ab initio molecular dynamic simulations for a variety of electrochemical conditions. Phys. Chem. Chem. Phys. 2020, 22, 10431–10437. 10.1039/C9CP06584A. [DOI] [PubMed] [Google Scholar]
- Abrashkin A.; Andelman D.; Orland H. Dipolar Poisson-Boltzmann Equation: Ions and Dipoles Close to Charge Interfaces. Phys. Rev. Lett. 2007, 99, 077801 10.1103/PhysRevLett.99.077801. [DOI] [PubMed] [Google Scholar]
- Malek A.; Eikerling M. H. Chemisorbed Oxygen at Pt(111): a DFT Study of Structural and Electronic Surface Properties. Electrocatalysis 2018, 9, 370–379. 10.1007/s12678-017-0436-0. [DOI] [Google Scholar]
- Martínez-Hincapié R.; Sebastián-Pascual P.; Climent V.; Feliu J. M. Exploring the interfacial neutral pH region of Pt(111) electrodes. Electrochem. Commun. 2015, 58, 62–64. 10.1016/j.elecom.2015.06.005. [DOI] [Google Scholar]
- Kobayashi K.; Suzuki T. S. Distribution of Relaxation Time Analysis for Non-ideal Immittance Spectrum: Discussion and Progress. J. Phys. Soc. Jpn. 2018, 87, 094002 10.7566/JPSJ.87.094002. [DOI] [Google Scholar]
- Fernández P. S.; Tereshchuk P.; Angelucci C. A.; Gomes J. F.; Garcia A. C.; Martins C. A.; Camara G. A.; Martins M. E.; Da Silva J. L. F.; Tremiliosi-Filho G. How do random superficial defects influence the electro-oxidation of glycerol on Pt(111) surfaces?. Phys. Chem. Chem. Phys. 2016, 18, 25582–25591. 10.1039/C6CP04768H. [DOI] [PubMed] [Google Scholar]
- Nie S.; Feibelman P. J.; Bartelt N. C.; Thürmer K. Pentagons and Heptagons in the First Water Layer on Pt(111). Phys. Rev. Lett. 2010, 105, 026102 10.1103/PhysRevLett.105.026102. [DOI] [PubMed] [Google Scholar]
- Góez-Marín A. M.; Feliu J. M. Oxygen reduction on nanostructured platinum surfaces in acidic media: Promoting effect of surface steps and ideal response of Pt(111). Catal. Today 2015, 244, 172–176. 10.1016/j.cattod.2014.05.009. [DOI] [Google Scholar]
- Chen X.; McCrum I. T.; Schwarz K. A.; Janik M. J.; Koper M. T. M. Co-adsorption of Cations as the Cause of the Apparent pH Dependence of Hydrogen Adsorption on a Stepped Platinum Single-Crystal Electrode. Angew. Chem. 2017, 56, 15025–15029. 10.1002/anie.201709455. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rizo R.; Fernández-Vidal J.; Hardwick L. J.; Attard G. A.; Vidal-Iglesias F. J.; Climent V.; Herrero E.; Feliu J. M. Investigating the presence of adsorbed species on Pt steps at low potentials. Nat. Commun. 2022, 13, 2550. 10.1038/s41467-022-30241-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Korpelin V.; Kiljunen T.; Melander M. M.; Caro M. A.; Kristoffersen H. H.; Mammen N.; Apaja V.; Honkala K. Addressing Dynamics at Catalytic Heterogeneous Interfaces with DFT-MD: Anomalous Temperature Distributions from Commonly Used Thermostats. J. Phys. Chem. Lett. 2022, 13, 2644–2652. 10.1021/acs.jpclett.2c00230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berná A.; Climent V.; Feliu J. M. New understanding of the nature of OH adsorption on Pt(111) electrodes. Electrochem. Commun. 2007, 9, 2789–2794. 10.1016/j.elecom.2007.09.018. [DOI] [Google Scholar]
- Kondo T.; Masuda T.; Aoki N.; Uosaki K. Potential-Dependent Structures and Potential-Induced Structure Changes at Pt(111) Single-Crystal Electrode/Sulfuric and Perchloric Acid Interfaces in the Potential Region between Hydrogen Underpotential Deposition and Surface Oxide Formation by In Situ Surface X-ray Scattering. J. Phys. Chem. C 2016, 120, 16118–16131. 10.1021/acs.jpcc.5b12766. [DOI] [Google Scholar]
- Attard G. A.; Brew A.; Hunter K.; Sharman J.; Wright E. Specific adsorption of perchlorate anions on Pt{hkl} single crystal electrodes. Phys. Chem. Chem. Phys. 2014, 16, 13689–13698. 10.1039/C4CP00564C. [DOI] [PubMed] [Google Scholar]
- Schmickler W. Double layer theory. J. Solid State Electrochem. 2020, 24, 2175–2176. 10.1007/s10008-020-04597-z. [DOI] [Google Scholar]
- Partenskii M. B.; Jordan P. C. Relaxing gap capacitor models of electrified interfaces. Am. J. Phys. 2010, 79, 103–110. [Google Scholar]
- Zhu J.-X.; Le J.-B.; Koper M. T. M.; Doblhoff-Dier K.; Cheng J. Effects of Adsorbed OH on Pt(100)/Water Interfacial Structures and Potential. J. Phys. Chem. C 2021, 125, 21571–21579. 10.1021/acs.jpcc.1c04895. [DOI] [Google Scholar]
- Clabaut P.; Staub R.; Galiana J.; Antonetti E.; Steinmann S. N. Water adlayers on noble metal surfaces: Insights from energy decomposition analysis. J. Chem. Phys. 2020, 153, 054703 10.1063/5.0013040. [DOI] [PubMed] [Google Scholar]
- Ovalle V. J.; Hsu Y.-S.; Agrawal N.; Janik M. J.; Waegele M. M. Correlating hydration free energy and specific adsorption of alkali metal cations during CO2 electroreduction on Au. Nat. Catal. 2022, 5, 624–632. 10.1038/s41929-022-00816-0. [DOI] [Google Scholar]
- Li C. K.; Huang J. Impedance Response of Electrochemical Interfaces: Part I. Exact Analytical Expressions for Ideally Polarizable Electrodes. J. Electrochem. Soc. 2021, 167, 166517. 10.1149/1945-7111/abd450. [DOI] [Google Scholar]
- Huang J. On obtaining double-layer capacitance and potential of zero charge from voltammetry. J. Electroanal. Chem. 2020, 870, 114243. 10.1016/j.jelechem.2020.114243. [DOI] [Google Scholar]
- Amatore C.; Maisonhaute E.; Simonneau G. Ohmic drop compensation in cyclic voltammetry at scan rates in the megavolt per second range: access to nanometric diffusion layers via transient electrochemistry. J. Electroanal. Chem. 2000, 486, 141–155. 10.1016/S0022-0728(00)00131-5. [DOI] [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Data Availability Statement
No new experimental data. The scripts used to calculate the model results are available from the author upon request.








