The ability to simulate electrochemical reactions from first-principles has advanced significantly in recent years. Here, we discuss the atomistic interpretation of electrochemistry at three scales: from the electronic structure to elementary processes to constant-potential reactions. At each scale, we highlight the importance of the grand-canonical nature of the process and show that the grand-canonical energy is the natural thermodynamic state variable, which has the additional benefit of simplifying calculations. We show that atomic forces are the derivative of the grand-potential energy when the potential is fixed. We further examine the meaning of potential at the atomic scale and its link to the chemical potential and discuss the link between charge transfer and potential in several situations.

Electrochemical transformations are of increasing importance in the transition of society away from fossil fuels, such as for electricity storage or the conversion of renewable electricity into chemicals or fuels. Simultaneously, atomic-scale calculations are increasingly able to describe the electrochemical interface with greater levels of sophistication. Fundamentally, electrochemical transformations occur in systems that are open to the flow of electrons, from an external circuit, which allows the potential to be controlled at the point of reaction. Increasingly, atomistic and electronic-structure methods that simulate the controlled potential are being employed to describe such transformations. In this Perspective, we discuss the thermodynamic implications of switching from standard atomistic simulations, in which the number of electrons is fixed, to grand-canonical simulations, in which the system’s electrical potential is fixed. We will discuss the implications at three different scales: at the electronic structure level, at the atomistic level, and, finally, at the reaction thermodynamics level.

Electrochemical reactions are chemical transformations in which the electric potential of an electrode is controlled by an external circuit, such as a potentiostat. Here, we examine the concept of potential at the atomistic scale. Introductory textbooks often show the potential as dropping off smoothly from an electrode surface as it is screened by the solvent’s double layer. The electric potential is a quantity that is accessible—and fundamental—in electronic-structure calculations, so we performed some simple simulations to analyze the behavior of the potential near an electrode. In Fig. 1(b), we have plotted some traces, along the z axis, of the electrostatic potential throughout a periodic calculation of the type often used to simulate such reactions. (Calculation details are in the section titled Methods.) The first observation is that the potential is not single-valued, but varies dramatically with position in the cell, with strong variations, in particular, near atomic nuclei. This picture is quite different from that often shown in textbooks.

FIG. 1.

Traces of the potential in an electronic-structure calculation. (a) The atomic structure used, consisting of a water layer above a periodic Pt slab. A top view is also shown as an inset. (b) Vertical potential ϕ traces at three different (x, y) coordinates; these points (0, 1, and 2) are marked on (a). (c) xy-averaged potential traces, with two different charges (q1 and q2). (d) The difference in the two plots of (c). For this system, q1 and q2 represent systems with 0.46 and 0.51 excess electrons, respectively.

FIG. 1.

Traces of the potential in an electronic-structure calculation. (a) The atomic structure used, consisting of a water layer above a periodic Pt slab. A top view is also shown as an inset. (b) Vertical potential ϕ traces at three different (x, y) coordinates; these points (0, 1, and 2) are marked on (a). (c) xy-averaged potential traces, with two different charges (q1 and q2). (d) The difference in the two plots of (c). For this system, q1 and q2 represent systems with 0.46 and 0.51 excess electrons, respectively.

Close modal

In Fig. 1(c), we instead plot the xy average of the potential at each z coordinate to smooth out variations in-plane with the electrode surface. Here, we do this at two different charges—that is, with different numbers of electrons in each simulation—and we can see a slight difference in the two potential profiles. At both charges, we still observe large spatial variations in potential. If we examine the difference between these two profiles [Fig. 1(d)], we finally see something that resembles the smooth drop-off near the surface, as is often shown in textbooks. Thus, it is best not to think of the electric potential as a single-valued quantity at an electrode, but rather as a small—but significant—perturbation to the scalar potential.

While we often think of using charge (q) to control potential (ϕ), this can be difficult to assess in electrochemical simulations since ϕ is not single-valued, but rather varies in space. Instead, it is simpler to consider the electron’s chemical potential (μe), which is controlled by the number of electrons (Ne) in the system. The chemical potential—which is defined as the incremental change in the system’s free energy when electrons are added or removed—has a single value for the system and, thus, is unambiguous to define. The potentiostat controls μe by injecting or removing electrons from the system—that is, through current—which in static atomic configurations is manifested as changes in Ne. Since μe and Ne are conjugate variables, they cannot be individually varied—that is, if μe is specified, then we have no control over the number of electrons in the system, assuming that the other key variables (such as atomic positions, temperature, or system volume) are fixed.

In practice, to control the potential in atomistic simulations—whether we think in terms of μe or ϕ—the most straightforward approach is to use grand-canonical simulations in which Ne is a free variable used to set the potential to a target value. Grand-canonical density functional theory (GC-DFT)1–12 methods have emerged as tools to study potential-dependent reactions and processes at user-specified applied potentials. These methods add or subtract electrons to the simulation cell in order to achieve a desired potential and can therefore control the potential to any arbitrary precision. Since the electron's chemical potential is equivalent to the Fermi level against a reasonable reference state, this quantity can be accessed through the work function. Grand-canonical methods have been used to study the potential-dependence of a multitude of interfacial reactions, including hydrogen evolution,13–15 CO2 reduction,6,16–18 formic acid oxidation,19 and CO adsorption.20 In particular, these methods allow for the potential-dependent study of electrochemical reaction barriers—something otherwise accessible through computationally expensive extrapolation methods21—but at relatively low computational expense and with precise control over the potential at which the barrier is evaluated.

Here, we use the solvated jellium (SJ)8 method to calculate reaction energies and barriers at the constant applied potential, as illustrated in Fig. 2. This electronically grand-canonical method is implemented in the GPAW22,23 electronic structure code and allows for computationally efficient calculations at the constant applied potential. The countercharge (“jellium”) is spatially constrained in the implicit solvent, and a solvent exclusion region is included in the innermost region of the interface to avoid spurious interactions between the implicit and explicit solvents. In our view, the primary purpose of the implicit solvent is to screen the large fields that would otherwise be created by the charge imbalance; what little energetic effects of solvation that occur on the top side of the water tend to cancel out between stages of a reaction. Importantly, we find an approximately linear relationship between Ne and μe (shown in Fig. 2) at a fixed geometry, which leads to rapid convergence to a specified μe, often with a computational penalty of only ∼30% over the course of a trajectory. In this work, we control the potential to within a threshold of 0.01 V for structural optimizations and 0.001 V for normal mode analyses. Full computational details are included in the section titled Methods.

FIG. 2.

Potential-controlled simulations in the solvated jellium (SJ) method. (Fractional) electrons are added or removed in order to change the system’s work function, while a countercharge (“jellium”) is introduced in a region of implicit solvent, above the explicit solvent that sits in the reactive zone. In the left image, the jellium is shown by the dashed lines and the implicit solvent (in terms of the dielectric constant) is shown as the shaded region. The center image shows that the excess electrons localize at the surface of the simulated electrode and in the Stern layer (for reasonably small perturbations). The right image shows that the potential varies linearly with the applied charge.

FIG. 2.

Potential-controlled simulations in the solvated jellium (SJ) method. (Fractional) electrons are added or removed in order to change the system’s work function, while a countercharge (“jellium”) is introduced in a region of implicit solvent, above the explicit solvent that sits in the reactive zone. In the left image, the jellium is shown by the dashed lines and the implicit solvent (in terms of the dielectric constant) is shown as the shaded region. The center image shows that the excess electrons localize at the surface of the simulated electrode and in the Stern layer (for reasonably small perturbations). The right image shows that the potential varies linearly with the applied charge.

Close modal

A traditional electronic structure calculation takes place at fixed nuclear coordinates {xi} and with a fixed number Ne of electrons—typically, Ne matches the positive charge of the nuclei (or pseudopotentials) in the system. Other information input to the calculation includes the electronic temperature Te. These fixed quantities are the natural variables of the Helmholtz energy, Apot, and thus, the potential energy is calculated in this thermodynamic potential, whose fundamental equation is given as

dApot=SedTe+μedNeiFidxi,
(1)

where the electronic entropy Se is the conjugate variables of Te and Fi is the component of the force vector associated with nuclear coordinate xi. (Additional terms may be present if the system volume changes.) Here, we are using the subscript “pot” on Apot to clearly distinguish that this energy only includes the system at a static nuclear configuration (that is, the potential energy surface), as opposed to the system free energy including nuclear motion, which we will examine later in this work.

In typical electronic structure methods, such as DFT, the force can be acquired at minimal additional cost by employing the Hellmann–Feynman theorem.24 By differentiating Eq. (1), we find the relationship between the force and potential energy,

Fi=ApotxiNe,Te,{xji}.
(2)

In atomistic calculations, Apot as a function of {xi} is typically referred to as the “potential energy surface” (PES), and we get the expected relation that the force on any atom is the negative gradient of the PES at a constant number of electrons, electronic temperature, and position of all other nuclei. This relationship is the key to many atomistic methods. In stationary-point searches, the gradient of the potential energy surface is used to quickly converge to a local minimum or saddle point in Apot. The consistency of the force with the potential energy surface is also required to conserve energy in molecular dynamics simulations.

In electrochemical systems, the number of electrons, Ne, is a free variable that is used to control the electronic chemical potential, μe. Thermodynamics allows us to transform our thermodynamic potential by invoking a Legendre transform,

ΦpotApotμeNe,

where Φpot may be referred to as the grand-potential energy. This transformation allows us to re-express the fundamental equation with natural variables μe, Te, and {xi},

dΦpot=SedTeNedμeiFidxi.

Therefore, each force can equivalently be expressed as

Fi=Φpotxiμe,Te,{xji}.

That is, the forces are simultaneously the negative gradient of the potential energy (Apot) at constant charge and the grand-potential energy (Φpot) at constant potential.

The implications of this are significant for atomistic simulations. That is, all of our standard algorithms—such as for structure optimization, saddle point searches, or molecular dynamics—will work unmodified as long as the transformed energy Φpot is used when constant-potential simulations are employed. As we will see later, the same transformation (−μeNe) is required for thermodynamically consistent reaction energies. Thus, Φpot can be considered the natural energy associated with the grand-canonical ensemble—and it is the energy returned to the user by default in the SJ method.

In Fig. 3, we illustrate the force-equivalence using the results of constant-charge and constant-potential DFT calculations. In these calculations, we start from a given charge and potential condition (the black dot) and then calculate Apot along a constant-charge path and Φpot along a constant-potential path. Here, we clearly see that the two curves share a slope at the black dot—where the charge–potential conditions coincide—and we further see that this slope matches that of the force calculated analytically by the Hellmann–Feynman theorem. Thus, although there are two different potential energy surfaces (Apot and Φpot), the force at any point is valid for either path. This also implies that stationary points on the potential energy surface—local minima and saddle points—are stationary points in either ensemble.25 That is, the atomic coordinates that describe a local minimum or saddle point found in the canonical ensemble (Apot) are equivalently a local minimum or saddle point in the grand-canonical ensemble (Φpot).

FIG. 3.

The analytical force is the derivative of both the potential energy Apot at constant Ne and the grand-potential energy ΦpotApotNeμe at constant μe. Calculations show the displacement of an H atom rising between a Pt surface and an explicit water layer at indicated charge and potential conditions. At the black point, the two systems are identical, having a charge of +0.18 and a potential of 0 VSHE.

FIG. 3.

The analytical force is the derivative of both the potential energy Apot at constant Ne and the grand-potential energy ΦpotApotNeμe at constant μe. Calculations show the displacement of an H atom rising between a Pt surface and an explicit water layer at indicated charge and potential conditions. At the black point, the two systems are identical, having a charge of +0.18 and a potential of 0 VSHE.

Close modal

We will next turn to the topic of an elementary reaction, which takes place at an electrode surface. As the simplest example, consider a proton reacting to form an adsorbed hydrogen atom in the first step of the hydrogen evolution reaction. Conventionally, reactions such as this are written formally with an integer electron transfer, i.e., H+ + e + *H*, where * represents an adsorption site on the surface of the electrode. However, if the reaction takes place at a fixed potential (expressed in terms of μe), then we have no control over the number of electrons (Ne) in either the initial or final state: these are the free variables that set μe. That is, we should properly write

H++xe+*ϕH*(1x)+,
(3)

where x is not specified a priori, but is instead inferred a posteriori based on the value of Ne in potential-controlled calculations of the bare electrode * and the hydrogen-adsorbed electrode (H*). The (1−x)+ superscript refers not to the charge localized on the final state, but to the net charge difference between the two states. This non-integer electron transfer has gone by various names in the previous literature, including “electrosorption valency” and “partial charge transfer”.11,26–31

To be clear, this does not imply that electrons are not quantized—they certainly come in integer quantities. If x is found to be 0.9 electrons in the above reaction, we would physically expect the potentiostat to hold the potential (or work function) constant by injecting, on average, nine electrons every ten times this reaction occurred to an electrode that might have 1015 sites per cm2 of surface. However, if we examine a complete reaction—i.e., a full catalytic cycle—then we certainly should expect the total number of electrons transferred to be an integer. For example, if the above step is followed by a second protonation to liberate H2 gas, leaving the surface in its initial state, then exactly two electrons must be involved. This is also what we infer from grand-canonical electronic structure calculations since the surface will be in an identical state before and after the reaction.

Can this non-integer electron transfer be measured experimentally? Recently, a team (including authors of this work) showed strong agreement between grand-canonical electronic structure calculations and spectroscopic experiments that involved the protonation of a surface-tethered 4-mercaptobenzoic acid (4MBA).32 SJ calculations suggested that about −0.1 electrons should accompany this protonation at the constant potential. Experimentally, this is manifested by a potential-dependence of pKA since the protonation reaction

*4MBA+H++xeϕ*4MBAHx

involves a partial electron whose chemical potential is a function of ϕ, the applied potential. This implies that pKA should vary with potential as − xe/(kBT × ln 10), where e is the (positive) charge of an electron and kB is the Boltzmann constant. This gives a slope of about −1.7 V−1, which is very close to what was observed experimentally. The SJ calculations were also in close agreement with the related compound 2MBA,33 in which the charged group on the MBA is closer to the surface, resulting in a larger charge transfer of about −0.6 electrons upon protonation.

We next turn our attention to reaction barriers, calculated at a specified potential. Consider the initial state of the proton-discharge reaction [Eq. (3)]: the unit cell contains two subsystems, both of which are charged; the acidic water layer is electron-deficient and the electrode surface may have a net negative or positive charge depending on the potential. In the course of this elementary reaction, the proton discharges from the acidic water layer, adsorbing onto the surface, forming an adsorbed hydrogen atom. We show the reaction pathway for this reaction in Fig. 4. As we noted elsewhere,34 these calculations are adiabatic and do not introduce discontinuities in the charge transfer profiles; the electron does not “jump” in integer increments. Instead, the electron transfer occurs smoothly across the interior images in the NEB calculation. We recently reported15 that while the electron transfer process has a finite width along the reaction coordinate, the inflection point of the electron transfer profile tends to coincide with the saddle point over a wide range of potentials at least for this simple reaction. Here, 0.63 electrons are injected to keep the applied potential constant.

FIG. 4.

Reaction pathway (black circles) and charge transfer profile (red squares) for the Volmer step on Pt(111)/1ML calculated with the solvated jellium method at a constant potential of −0.1 VSHE. The free energy of the key states is shown in black.

FIG. 4.

Reaction pathway (black circles) and charge transfer profile (red squares) for the Volmer step on Pt(111)/1ML calculated with the solvated jellium method at a constant potential of −0.1 VSHE. The free energy of the key states is shown in black.

Close modal

Intuitively, one might expect the magnitude of charge transfer to be close to, but not precisely one, keeping in mind the discussion in the previous paragraphs. However, the charge transfer magnitude is significantly, and consistently, less than unity; grand-canonical methods and Bader charge analyses35 typically reveal charge separations on the order of 0.5–0.7 electrons.8,15,36–39 This fractional charge transfer was long thought to be an artifact of self-interaction40–47 in Kohn–Sham DFT48,49—the interaction of an electron (or basis function thereof) with itself via the electron density in the Hartree potential—but Chen et al.39 have recently suggested that this is a consequence of hybridization between solvated protons and the electron density of the electrode. Chen et al. reported that charge transfer is fractional at different levels of theory—including hybrid methods with fractional and full exact exchange—as long as the solvated proton is in close vicinity to the electrode. However, they reported that the charge transfer approaches unity when the proton is moved farther away from the electrode. The extent of charge transfer is thus dependent on the spatial separation of the two subsystems, and the energetics of hybridized states need not be identical to those of the bulk.

Here, we investigate the effect of the applied potential on the magnitude of charge transfer. Figure 5 shows the reaction pathway and electron-transfer profile for the Volmer reaction on Pt(111)/1ML over a potential window of 1 V. The magnitude of charge transfer decreases by 0.12 electrons when the potential is swept from ϕ = 0.2 VSHE to ϕ = −0.8 VSHE; this corresponds to an 18% reduction in charge transfer. Hence, the change in reaction energy of interfacial processes is not strictly linear with respect to the applied potential, but exhibits some curvature over wide potential ranges.

FIG. 5.

Reaction pathway (black circles) and charge transfer profile (red squares) for the Volmer step on Pt(111)/1ML calculated with the solvated jellium method at various fixed potentials. The extent of charge transfer is dependent on the applied potential.

FIG. 5.

Reaction pathway (black circles) and charge transfer profile (red squares) for the Volmer step on Pt(111)/1ML calculated with the solvated jellium method at various fixed potentials. The extent of charge transfer is dependent on the applied potential.

Close modal

The origin of this phenomenon is shown in Fig. 6, where the extent of electron transfer is plotted relative to the Pt–H+ distance. The data in Fig. 6 cover the range −0.8 ≤ ϕ ≤ 0.2 VSHE. First, we note that these potentials are more negative than the experimentally determined potential of zero charge of 0.3 VNHE,50 and a hexagonal H-down water geometry is more favorable than an H-up geometry. At ϕ = 0.3 VSHE, we find that the water reorients into a mixed H-down/H-up geometry. As the electrode becomes more negatively charged, electrostatic forces cause the water bilayer to approach the surface. When the distance between the two subsystems decreases, the degree of overlap increases, and the increased overlap between the electron density of the electrode and the solvated proton of the initial state decreases the magnitude of charge transfer over the course of the reaction. We should carefully note that the magnitude of such an effect can be expected to have been amplified by the periodic nature of the simulation cells employed and the lack of buffering ions; larger-size simulations will be necessary to explore this effect in more detail.

FIG. 6.

The magnitude of charge transfer depends on the spatial separation of the two subsystems, which in turn depends on the applied potential.

FIG. 6.

The magnitude of charge transfer depends on the spatial separation of the two subsystems, which in turn depends on the applied potential.

Close modal

In a recent study,15 we demonstrated that the potential-dependence of electrochemical reaction barriers can accurately be described with a curved scaling model developed from parabola intersections, as shown in Fig. 7 (top). In this model, the sole free parameter b corresponds to the reaction barrier at thermoneutral conditions, that is, the potential that makes ΔE = 0. The functional form of the scaling is then just the intersection of these parabolas, given by

E=0,ΔE<4b,(ΔE+4b)216b,4bΔE4b,ΔE,ΔE>4b.
(4)

This model not only captures the reaction barriers at intermediate driving forces but also predicts the asymptotic approach to activationless and barrierless discharge for high and low magnitude driving forces, respectively. Importantly, it also captures the curved nature between these limits and fits all these criteria with only a single adjustable parameter. The scaling is shown in Fig. 7 for the Volmer reaction on Pt(111)/1ML. The reaction barrier approaches the barrierless region for low-magnitude applied potentials (ΔEϕ > 4b) and the activation less region for high-magnitude driving forces (ΔEϕ < − 4b).

FIG. 7.

Scaling of the Volmer step on Pt(111)/1ML. Top: The overlapping-parabolas model, giving curvature to the barrier vs energy-change plot. Bottom: Scaling within this model using a single parameter b, which was regressed to 0.197 eV. Solid lines and discrete points represent the fit and calculated barriers, respectively. Note that the limiting cases are captured by the functional form in Eq. (4).

FIG. 7.

Scaling of the Volmer step on Pt(111)/1ML. Top: The overlapping-parabolas model, giving curvature to the barrier vs energy-change plot. Bottom: Scaling within this model using a single parameter b, which was regressed to 0.197 eV. Solid lines and discrete points represent the fit and calculated barriers, respectively. Note that the limiting cases are captured by the functional form in Eq. (4).

Close modal

Next, we turn our attention to reaction thermodynamics in which we focus on the calculation of ΔG for overall reactions, elementary steps, and barriers. We emphasize that from the standpoint of reactions, electrons are no different from other reactants or products—if they are reactants, they need to be supplied to the system, and if they are products, they must be removed, both at a specified free energy. However, unlike other reactants and products, electrons need not be supplied in integer quantities, as discussed above. See Fig. 8, showing an example of the reduction of gaseous CO to adsorbed CHO; in this case, CO, a proton, and x electrons must be supplied to the system. The value of x can be inferred from constant-potential electronic structure calculations.

FIG. 8.

Although electrons are added or removed from a system during catalysis, the thermodynamic treatment of them is not fundamentally different from other reactants or products, shown here for the reaction * + CO + H+ + xe* CHO.

FIG. 8.

Although electrons are added or removed from a system during catalysis, the thermodynamic treatment of them is not fundamentally different from other reactants or products, shown here for the reaction * + CO + H+ + xe* CHO.

Close modal

The computational hydrogen electrode (CHE)51,52 is often used in elementary reactions in order to access the free energy of a proton/electron pair from that of H2 gas. In the CHE scheme, when the chemical potential of a proton/electron pair is encountered (as μ[H+] + μ[e+]), it is substituted with the thermodynamically equivalent chemical potential of hydrogen gas (as 12μ[H2]), which is true at the reference potential and pH and can be thermodynamically adjusted to other potentials and pH’s. The CHE scheme is thus designed for the net energy change across elementary steps involving coupled integer proton/electron transfers and is silent on the treatment of reaction barriers.

When the electron is non-integer or not directly coupled with a proton, such a relation can still be employed in which the proton and electron are decoupled.15,53 We refer to this as the “decoupled computational electrode” (DCE) model,15 which decouples the proton and electron chemical potential while enforcing thermodynamic consistency. Here, the energy of any state along a reaction trajectory, be it a thermodynamically stable state or a saddle point, is referenced to the correct reference state—a proton solvated in the bulk of the electrolyte—and enforces thermodynamic consistency. For example, consider the reaction of a proton with an electrocatalytic surface in order to create the barrier of the Volmer reaction,

H++xe+*H*

(where the final state bears a net charge of (1 − x) + relative to the initial state; we omit this for brevity of nomenclature). The change in Gibbs free energy G associated with this partial reaction is given as

ΔGϕ=Gϕ[H*]Gϕ[*]μ[H+]xμϕ[e].

In the DCE scheme (like in the CHE scheme), the chemical potential of the proton is replaced by the equilibrium relation of the desired reference electrode. If we employ the reversible hydrogen electrode (RHE) scale (H++e12H2, which is in equilibrium at 0 VRHE at any pH and temperature), then

μ[H+]=12μ[H2]μ0V[e].

A simple substitution of this into the previous equation leads to

ΔGϕ=Gϕ[H*]Gϕ[*]12μ[H2]+eϕ+(1x)μϕ[e],

where we have used μ0V[e] − μϕ[e] = . The free energy of the initial and final states, at potential ϕ, is constructed from the potential energy calculations (Apot or Φpot to create Gϕ of ΩϕGϕNeμe). If the grand free energy Ωϕ is used, then the equation simplifies to

ΔGϕ=Ωϕ[H*]Ωϕ[*]12μ[H2]+eϕ.

Thus, if grand free energies Ωϕ are employed, which include the electronic correction, the DCE formulation is nearly identical to the traditional CHE approach, and becomes identical in the limit of integer electron transfer. This free energy difference is apparent in Fig. 4 as the difference between the state marked G and that marked GIS.

The electronic structure calculations were undertaken in the open-source code GPAW,22,23 employing the solvated jellium method as described by us in an earlier publication,8 and implemented in GPAW. Atomistic manipulations were carried out in ASE,54 the atomic simulation environment. All calculations used the RPBE60 exchange--correlation functional of Hammer, Hansen, and Nørskov, with the exception of those to create the potential traces of Fig. 1.

In simulations to create Fig. 1, a 3 × 3 × 3-atom periodic slab of Pt was used with a hexagonal water overlayer, as shown. The PBE55 exchange–correlation functional was used with Fermi–Dirac smearing of the electronic occupations at 0.1 eV and 4 × 4 × 1 k-point sampling. The solvent exclusion zone used van der Waals radii for all atoms, and the implicit solvent was only present on the explicit-solvent side of the slab, with a dipole correction present in the direction normal to the slab, to isolate the electronic perturbations to the solvated side of the slab. The implicit solvent was the simplified continuum solvent model of Held and Walter,56 with parameters of u0 = 0.180 eV, ɛ = 78.36, γ = 18.4 × 10−3 Pa ⋅ m, and T = 298.15 K being used, as described in their works and the GPAW documentation.

Saddle points were calculated with the scaled and dynamic34 implementation of the nudged elastic band (NEB)57–59 method. Thermodynamically stable intermediates and saddle points were optimized to within 0.03 eV/Å, and the convergence scaling of chains-of-states is based on a displacement metric with a scaling factor, αs = 1; full details are provided in Ref. 34. Structures used to deduce subtle trends, e.g., in Fig. 6, were optimized until the forces acting on all unconstrained atoms are below 0.01 eV/Å. Platinum electrodes were modeled as 3 × 3 × 3 Pt(111) surfaces with full or fractional hydrogen coverage in the fcc sites. The grid spacing is 0.15–0.18 Å, and the systems were sampled on a 4 × 4 × 1  k-point mesh. The bottom layer was fixed at the bulk-optimized lattice constant of 3.99 Å with the RPBE60 exchange–correlation functional. Free energy corrections are included in the harmonic limit at 298 K,52 

ΔGϕ=ΔEϕ+ΔZPE+0TCpdTTΔS,

where ΔEϕ is the potential energy at the constant potential extrapolated to 0 K and subsequent terms come from partition functions obtained in normal-mode analyses.

The authors thank Alexandar Zaslavsky, Xin Lian, and Matthew Shinkar, all of Brown University, for valuable technical discussions and acknowledge the support from the U.S. National Science Foundation under Award No. 1553365.

The authors have no conflicts to disclose.

Per Lindgren: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Investigation (equal); Methodology (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Georg Kastlunger: Conceptualization (equal); Formal analysis (equal); Methodology (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal). Andrew A. Peterson: Conceptualization (equal); Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Methodology (equal); Project administration (equal); Supervision (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

The data that support the findings of this study are available from the corresponding author upon reasonable request.

1.
A. Y.
Lozovoi
,
A.
Alavi
,
J.
Kohanoff
, and
R. M.
Lynden-Bell
, “
Ab initio simulation of charged slabs at constant chemical potential
,”
J. Chem. Phys.
115
,
1661
1669
(
2001
).
2.
M.
Otani
and
O.
Sugino
, “
First-principles calculations of charged surfaces and interfaces: A plane-wave nonrepeated slab approach
,”
Phys. Rev. B
73
,
115407
(
2006
).
3.
C. D.
Taylor
,
S. A.
Wasileski
,
J.-S.
Filhol
, and
M.
Neurock
, “
First principles reaction modeling of the electrochemical interface: Consideration and calculation of a tunable surface potential from atomic and electronic structure
,”
Phys. Rev. B
73
,
165402
(
2006
).
4.
R.
Jinnouchi
and
A. B.
Anderson
, “
Electronic structure calculations of liquid-solid interfaces: Combination of density functional theory and modified Poisson-Boltzmann theory
,”
Phys. Rev. B
77
,
245417
(
2008
).
5.
K.
Letchworth-Weaver
and
T. A.
Arias
, “
Joint density functional theory of the electrode-electrolyte interface: Application to fixed electrode potentials, interfacial capacitances, and potentials of zero charge
,”
Phys. Rev. B
86
,
075140
(
2012
).
6.
J. D.
Goodpaster
,
A. T.
Bell
, and
M.
Head-Gordon
, “
Identification of possible pathways for C–C bond formation during electrochemical reduction of CO2: New theoretical insights from an improved electrochemical model
,”
J. Phys. Chem. Lett.
7
,
1471
1477
(
2016
).
7.
R.
Sundararaman
,
W. A.
Goddard
, and
T. A.
Arias
, “
Grand canonical electronic density-functional theory: Algorithms and applications to electrochemistry
,”
J. Chem. Phys.
146
,
114104
(
2017
).
8.
G.
Kastlunger
,
P.
Lindgren
, and
A. A.
Peterson
, “
Controlled-potential simulation of elementary electrochemical reactions: Proton discharge on metal surfaces
,”
J. Phys. Chem. C
122
,
12771
12781
(
2018
).
9.
A.
Bouzid
and
A.
Pasquarello
, “
Atomic-scale simulation of electrochemical processes at electrode/water interfaces under referenced bias potential
,”
J. Phys. Chem. Lett.
9
,
1880
1884
(
2018
).
10.
M. M.
Melander
,
M. J.
Kuisma
,
T. E. K.
Christensen
, and
K.
Honkala
, “
Grand-canonical approach to density functional theory of electrocatalytic systems: Thermodynamics of solid-liquid interfaces at constant ion and electrode potentials
,”
J. Chem. Phys.
150
,
041706
(
2019
).
11.
N. G.
Hörmann
,
N.
Marzari
, and
K.
Reuter
, “
Electrosorption at metal surfaces from first principles
,”
npj Comput. Mater.
6
,
136
(
2020
).
12.
N. G.
Hörmann
and
K.
Reuter
, “
Thermodynamic cyclic voltammograms based on ab initio calculations: Ag(111) in halide-containing solutions
,”
J. Chem. Theory Comput.
17
,
1782
1794
(
2021
).
13.
Y.
Huang
,
R. J.
Nielsen
, and
W. A.
Goddard
, “
Reaction mechanism for the hydrogen evolution reaction on the basal plane sulfur vacancy site of MOS2 using grand canonical potential kinetics
,”
J. Am. Chem. Soc.
140
,
16773
16782
(
2018
).
14.
M.
Van den Bossche
,
E.
Skúlason
,
C.
Rose-Petruck
, and
H.
Jónsson
, “
Assessment of constant-potential implicit solvation calculations of electrochemical energy barriers for H2 evolution on Pt
,”
J. Phys. Chem. C
123
,
4116
4124
(
2019
).
15.
P.
Lindgren
,
G.
Kastlunger
, and
A. A.
Peterson
, “
A challenge to the g ∼ 0 interpretation of hydrogen evolution
,”
ACS Catal.
10
,
121
128
(
2020
).
16.
M. R.
Singh
,
J. D.
Goodpaster
,
A. Z.
Weber
,
M.
Head-Gordon
, and
A. T.
Bell
, “
Mechanistic insights into electrochemical reduction of CO2 over Ag using density functional theory and transport models
,”
Proc. Natl. Acad. Sci. U. S. A.
114
,
E8812
E8821
(
2017
).
17.
H.
Zhang
,
W. A.
Goddard
,
Q.
Lu
, and
M.-J.
Cheng
, “
The importance of grand-canonical quantum mechanical methods to describe the effect of electrode potential on the stability of intermediates involved in both electrochemical CO2 reduction and hydrogen evolution
,”
Phys. Chem. Chem. Phys.
20
,
2549
2557
(
2018
).
18.
G.
Kastlunger
,
L.
Wang
,
N.
Govindarajan
,
H. H.
Heenen
,
S.
Ringe
,
T.
Jaramillo
,
C.
Hahn
, and
K.
Chan
, “
Using pH dependence to understand mechanisms in electrochemical CO reduction
,”
ACS Catal.
12
,
4344
4357
(
2022
).
19.
S. N.
Steinmann
,
C.
Michel
,
R.
Schwiedernoch
,
J.-S.
Filhol
, and
P.
Sautet
, “
Modeling the HCOOH/CO2 electrocatalytic reaction: When details are key
,”
ChemPhysChem
16
,
2307
2311
(
2015
).
20.
R.
Sundararaman
,
M. C.
Figueiredo
,
M. T. M.
Koper
, and
K. A.
Schwarz
, “
Electrochemical capacitance of co-terminated Pt(111) dominated by the co–solvent gap
,”
J. Phys. Chem. Lett.
8
,
5344
5348
(
2017
).
21.
E.
Skúlason
,
V.
Tripkovic
,
M. E.
Björketun
,
S.
Gudmundsdóttir
,
G.
Karlberg
,
J.
Rossmeisl
,
T.
Bligaard
,
H.
Jónsson
, and
J. K.
Nørskov
, “
Modeling the electrochemical hydrogen oxidation and evolution reactions on the basis of density functional theory calculations
,”
J. Phys. Chem. C
114
,
18182
18197
(
2010
).
22.
J. J.
Mortensen
,
L. B.
Hansen
, and
K. W.
Jacobsen
, “
Real-space grid implementation of the projector augmented wave method
,”
Phys. Rev. B
71
,
35109
(
2005
).
23.
J.
Enkovaara
,
C.
Rostgaard
,
J. J.
Mortensen
,
J.
Chen
,
M.
Dułak
,
L.
Ferrighi
,
J.
Gavnholt
,
C.
Glinsvad
,
V.
Haikola
,
H. A.
Hansen
,
H. H.
Kristoffersen
,
M.
Kuisma
,
A. H.
Larsen
,
L.
Lehtovaara
,
M.
Ljungberg
,
O.
Lopez-Acevedo
,
P. G.
Moses
,
J.
Ojanen
,
T.
Olsen
,
V.
Petzold
,
N. A.
Romero
,
J.
Stausholm-Møller
,
M.
Strange
,
G. A.
Tritsaris
,
M.
Vanin
,
M.
Walter
,
B.
Hammer
,
H.
Häkkinen
,
G. K. H.
Madsen
,
R. M.
Nieminen
,
J. K.
Nørskov
,
M.
Puska
,
T. T.
Rantala
,
J.
Schiøtz
,
K. S.
Thygesen
, and
K. W.
Jacobsen
, “
Electronic structure calculations with GPAW: A real-space implementation of the projector augmented-wave method
,”
J. Phys.: Condens. Matter
22
,
253202
(
2010
).
24.
R. P.
Feynman
, “
Forces in molecules
,”
Phys. Rev.
56
,
340
343
(
1939
).
25.
C.
Bureau
and
G.
Lécayon
, “
On a modeling of voltage-application to metallic electrodes using density functional theory
,”
J. Chem. Phys.
106
,
8821
8829
(
1997
).
26.
J. W.
Schultze
and
K. J.
Vetter
, “
Experimental determination and interpretation of the electrosorption valency γ
,”
J. Electroanal. Chem.
44
,
63
81
(
1973
).
27.
W.
Schmickler
and
R.
Guidelli
, “
Ionic adsorption and the surface dipole potential
,”
J. Electroanal. Chem.
235
,
387
392
(
1987
).
28.
W.
Schmickler
and
D.
Henderson
, “
The interphase between jellium and a hard sphere electrolyte: Capacity-charge characteristics and dipole potentials
,”
J. Chem. Phys.
85
,
1650
1657
(
1985
).
29.
W.
Schmickler
, “
The surface dipole moment of species adsorbed from a solution
,”
J. Electroanal. Chem.
249
,
25
33
(
1988
).
30.
A. A.
Kornyshev
and
W.
Schmickler
, “
On the coverage dependence of the partial charge transfer coefficient
,”
J. Electroanal. Chem.
202
,
1
21
(
1986
).
31.
S.
Trasatti
and
R.
Parsons
, “
Interphases in systems of conducting phases (Recommendations 1985)
,”
Pure Appl. Chem.
58
,
437
(
1986
).
32.
A.
Ge
,
G.
Kastlunger
,
J.
Meng
,
P.
Lindgren
,
J.
Song
,
Q.
Liu
,
A.
Zaslavsky
,
T.
Lian
, and
A. A.
Peterson
, “
On the coupling of electron transfer to proton transfer at electrified interfaces
,”
J. Am. Chem. Soc.
142
,
11829
11834
(
2020
).
33.
C.
Ma
and
J. M.
Harris
, “
Surface-enhanced Raman spectroscopy investigation of the potential-dependent acid-base chemistry of silver-immobilized 2-mercaptobenzoic acid
,”
Langmuir
27
,
3527
3533
(
2011
).
34.
P.
Lindgren
,
G.
Kastlunger
, and
A. A.
Peterson
, “
Scaled and dynamic optimizations of nudged elastic bands
,”
J. Chem. Theory Comput.
15
,
5787
5793
(
2019
).
35.
G.
Henkelman
,
A.
Arnaldsson
, and
H.
Jónsson
, “
A fast and robust algorithm for Bader decomposition of charge density
,”
Comput. Mater. Sci.
36
,
354
360
(
2006
).
36.
M. E.
Björketun
,
V.
Tripkovic
,
E.
Skúlason
, and
J.
Rossmeisl
, “
Modeling of the symmetry factor of electrochemical proton discharge via the Volmer reaction
,”
Catal. Today
202
,
168
174
(
2013
).
37.
K.
Chan
and
J. K.
Nørskov
, “
Electrochemical barriers made simple
,”
J. Phys. Chem. Lett.
6
,
2663
2668
(
2015
).
38.
K.
Chan
and
J. K.
Nørskov
, “
Potential dependence of electrochemical barriers from ab initio calculations
,”
J. Phys. Chem. Lett.
7
,
1686
1690
(
2016
).
39.
L. D.
Chen
,
M.
Bajdich
,
J. M. P.
Martirez
,
C. M.
Krauter
,
J. A.
Gauthier
,
E. A.
Carter
,
A. C.
Luntz
,
K.
Chan
, and
J. K.
Nørskov
, “
Understanding the apparent fractional charge of protons in the aqueous electrochemical double layer
,”
Nat. Commun.
9
,
3202
(
2018
).
40.
J. P.
Perdew
and
A.
Zunger
, “
Self-interaction correction to density-functional approximations for many-electron systems
,”
Phys. Rev. B
23
,
5048
5079
(
1981
).
41.
P.
Mori-Sánchez
,
A. J.
Cohen
, and
W.
Yang
, “
Many-electron self-interaction error in approximate density functionals
,”
J. Chem. Phys.
125
,
201102
(
2006
).
42.
A.
Ruzsinszky
,
J. P.
Perdew
,
G. I.
Csonka
,
O. A.
Vydrov
, and
G. E.
Scuseria
, “
Density functionals that are one- and two- are not always many-electron self-interaction-free, as shown for H2+, He2+, LiH+, and Ne2+
,”
J. Chem. Phys.
126
,
104102
(
2007
).
43.
A. J.
Cohen
,
P.
Mori-Sánchez
, and
W.
Yang
, “
Insights into current limitations of density functional theory
,”
Science
321
,
792
794
(
2008
).
44.
R.
Haunschild
,
T. M.
Henderson
,
C. A.
Jiménez-Hoyos
, and
G. E.
Scuseria
, “
Many-electron self-interaction and spin polarization errors in local hybrid density functionals
,”
J. Chem. Phys.
133
,
134116
(
2010
).
45.
T.
Schmidt
and
S.
Kümmel
, “
One- and many-electron self-interaction error in local and global hybrid functionals
,”
Phys. Rev. B
93
,
165120
(
2016
).
46.
T. Z. H.
Gani
and
H. J.
Kulik
, “
Where does the density localize? Convergent behavior for global hybrids, range separation, and DFT+U
,”
J. Chem. Theory Comput.
12
,
5931
5945
(
2016
).
47.
Q.
Zhao
and
H. J.
Kulik
, “
Where does the density localize in the solid state? Divergent behavior for hybrids and DFT+U
,”
J. Chem. Theory Comput.
14
,
670
683
(
2018
).
48.
P.
Hohenberg
and
W.
Kohn
, “
Inhomogeneous electron gas
,”
Phys. Rev.
136
,
B864
B871
(
1964
).
49.
W.
Kohn
and
L. J.
Sham
, “
Self-consistent equations including exchange and correlation effects
,”
Phys. Rev.
140
,
A1133
A1138
(
1965
).
50.
K.
Ojha
,
N.
Arulmozhi
,
D.
Aranzales
, and
M. T. M.
Koper
, “
Double layer at the Pt(111)–aqueous electrolyte interface: Potential of zero charge and anomalous Gouy–Chapman screening
,”
Angew. Chem.
132
,
721
725
(
2020
).
51.
J. K.
Nørskov
,
J.
Rossmeisl
,
A.
Logadottir
,
L.
Lindqvist
,
J. R.
Kitchin
,
T.
Bligaard
, and
H.
Jónsson
, “
Origin of the overpotential for oxygen reduction at a fuel-cell cathode
,”
J. Phys. Chem. B
108
,
17886
17892
(
2004
).
52.
A. A.
Peterson
,
F.
Abild-Pedersen
,
F.
Studt
,
J.
Rossmeisl
, and
J. K.
Nørskov
, “
How copper catalyzes the electroreduction of carbon dioxide into hydrocarbon fuels
,”
Energy Environ. Sci.
3
,
1311
(
2010
).
53.
M. H.
Hansen
and
J.
Rossmeisl
, “
pH in grand canonical statistics of an electrochemical interface
,”
J. Phys. Chem. C
120
,
29135
29143
(
2016
).
54.
A.
Hjorth Larsen
,
J.
Jørgen Mortensen
,
J.
Blomqvist
,
I. E.
Castelli
,
R.
Christensen
,
M.
Dułak
,
J.
Friis
,
M. N.
Groves
,
B.
Hammer
,
C.
Hargus
,
E. D.
Hermes
,
P. C.
Jennings
,
P.
Bjerre Jensen
,
J.
Kermode
,
J. R.
Kitchin
,
E.
Leonhard Kolsbjerg
,
J.
Kubal
,
K.
Kaasbjerg
,
S.
Lysgaard
,
J.
Bergmann Maronsson
,
T.
Maxson
,
T.
Olsen
,
L.
Pastewka
,
A.
Peterson
,
C.
Rostgaard
,
J.
Schiøtz
,
O.
Schütt
,
M.
Strange
,
K. S.
Thygesen
,
T.
Vegge
,
L.
Vilhelmsen
,
M.
Walter
,
Z.
Zeng
, and
K. W.
Jacobsen
, “
The atomic simulation environment—a python library for working with atoms
,”
J. Phys.: Condens. Matter
29
,
273002
(
2017
).
55.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
, “
Generalized gradient approximation made simple
,”
Phys. Rev. Lett.
77
,
3865
3868
(
1996
).
56.
A.
Held
and
M.
Walter
, “
Simplified continuum solvent model with a smooth cavity based on volumetric data
,”
J. Chem. Phys.
141
,
174108
(
2014
).
57.
H.
Jónsson
,
G.
Mills
, and
K. W.
Jacobsen
, “
Nudged elastic band method for finding minimum energy path of transitions
,” in
Classical and Quantum Dynamics in Condensed Phase Simulations
, edited by
B.
Berne
,
G.
Ciccotti
and
D.
Coker
(
World Scientific
,
1998
), Vol. 385.
58.
G.
Henkelman
and
H.
Jónsson
, “
Improved tangent estimate in the nudged elastic band method for finding minimum energy paths and saddle points
,”
J. Chem. Phys.
113
,
9978
9985
(
2000
).
59.
G.
Henkelman
,
B. P.
Uberuaga
, and
H.
Jónsson
, “
A climbing image nudged elastic band method for finding saddle points and minimum energy paths
,”
J. Chem. Phys.
113
,
9901
9904
(
2000
).
60.
B.
Hammer
,
L. B.
Hansen
, and
J. K.
Nørskov
, “
Improved adsorption energetics within density-functional theory using revised Perdew-Burke-Ernzerhof functionals
,”
Phys. Rev. B
59
,
7413
7421
(
1999
).