Water adsorption energy, Eads, is a key physical quantity in sustainable chemical technologies such as (photo)electrocatalytic water splitting, water desalination, and water harvesting. In many of these applications, the electrode surface is operated outside the point (potential) of zero charge, which attracts counter-ions to form the electric double layer and controls the surface properties. Here, by applying density functional theory-based finite-field molecular dynamics simulations, we have studied the effect of water adsorption energy Eads on surface acidity and the Helmholtz capacitance of BiVO4 as an example of metal oxide electrodes with weakly chemisorbed water. This allows us to establish the effect of Eads on the coordination number, the H-bond network, and the orientation of chemisorbed water by comparing an oxide series composed of BiVO4, TiO2, and SnO2. In particular, it is found that a positive correlation exists between the degree of asymmetry ΔCH in the Helmholtz capacitance and the strength of Eads. This correlation is verified and extended further to graphene-like systems with physisorbed water, where the electric double layers (EDLs) are controlled by electronic charge rather than proton charge as in the oxide series. Therefore, this work reveals a general relationship between water adsorption energy Eads and EDLs, which is relevant to both electrochemical reactivity and the electrowetting of aqueous interfaces.

Water molecules adsorb on electrode surfaces when the solid and liquid phases come into contact with each other.1 It has been shown that the adsorbed water molecules play an important role in many sustainable chemical technologies, such as (photo)electrocatalytic water splitting,2 water desalination,3 and water harvesting.4 The water adsorption energy Eads is a key physical quantity that can reflect the interaction between electrolyte and electrode materials.5 Therefore, unraveling the relationship between Eads and the surface properties of electrode materials is of paramount importance.

When in contact with liquid electrolyte under working conditions, the electrode materials will attract counter-ions from the solution and form the electric double layers (EDLs) that strongly modulate the interfacial reactivity and wetting. EDLs can be developed either at metal electrodes beyond the potential of zero charge6 or at metal-oxide electrodes outside the point of zero charge. In the latter case, the metal-oxide surfaces are positively or negatively electrified by the adsorption of protons or hydroxyl groups, which depend on the pH of the environment.7 

The well-known classical Gouy–Chapman–Stern (GCS) model has been used to describe the structure of EDLs at oxide surfaces, in which EDLs consist of a space charge layer, a Helmholtz layer, and a Gouy–Chapman layer.1,8 The Helmholtz layer is formed by the interfacial proton charges and counter-ions from the electrolyte, which spans for 3–5 Å. The microscopic structure, in particular the fully solvated surface and the interaction and orientation of adsorbed molecular or dissociated water molecules, determines the electrochemical performances, i.e., the thermodynamics, kinetics, and product selectivity of electrode materials. Therefore, the description of the relationship between water adsorption energy Eads and EDLs is crucial.9–13 

Thanks to the development of advanced in situ experimental technologies and theoretical studies, it was not until recently that the structure and properties of interfacial EDLs started to be revealed. For example, electrochemical x-ray photoelectron spectroscopy (EC-XPS) applies to the characterization of the surface structure of the electrode,14 sum-frequency vibrational spectroscopy retrieves the vibrational spectra of the bonded water layer and the ion diffuse layer,15, in situ Raman spectroscopy investigates the interfacial water on single-crystal surfaces,16,17 the dual scale atomic force microscopy (AFM) spectroscopy technique provides direct access to the facet-dependent local surface charge density and the potential drop across the diffuse part of the EDL,18 and operando scanning electrochemical probe microscopy can correlate electrochemical activity with changes in surface properties.19 Moreover, various computational methods, e.g., quantum mechanics/implicit solvation model (QM/ISM),20,21 quantum mechanics/molecular mechanical (QM/MM),22,23 density functional theory-based molecular dynamics simulation (DFTMD),24–27 and machine learning-accelerated MD simulations,28–30 have been employed to simulate the electrochemical interfaces and describe the microscopic structure and the dynamics of water and ions near the interfaces.

Despite the tremendous experimental and theoretical studies that have been devoted to this topic, a molecular-level understanding of the effect of water adsorption energy Eads on microscopic structure and Helmholtz capacitance within EDLs remains elusive, owing to its complexity and difficulty in being probed with sufficient resolution.31–33 Recently, by combining the modern theory of polarization and the constant electric displacement D Hamiltonian developed by Stengel, Spaldin, and Vanderbilt,34 the finite-field DFTMD method has become an attractive technique for modeling EDLs at electrified solid–liquid interfaces.35–38 With the same amount but opposite types of proton charges, the two slabs of the simulated oxide surface are charged at a fixed chemical composition of the supercell. In the finite-field DFTMD simulations, the average Helmholtz capacitance CH is obtained based on the supercell polarization Pz.39 Note that the ensemble averaged supercell polarization Pz converges rapidly by the application of constant electric displacement simulations, which makes the computation of CH accessible within the timescale of DFTMD simulation.36 

In the previous work on applying finite-field DFTMD for modeling EDLs at both rutile TiO2(110)/NaCl36 and SnO2(110)/NaCl37 interfaces, it was found that water chemisorption is a key factor in accounting for the asymmetric EDLs seen at aqueous interfaces of oxide. To further quantify the relationships between Eads, molecular structure, and Helmholtz capacitance of EDLs on metal oxides, one of the most promising n-type oxide bismuth vanadates, BiVO4, is selected here as an example of an oxide electrode with weakly chemisorbed water. BiVO4 has been applied to photoelectrochemical water splitting and has garnered considerable attention due to its low-cost, stability, nontoxicity, ideal bandgap (∼2.4 eV), and high electron–hole separation efficiencies (over 70%).40,41 In addition, the Eads of BiVO4 was reported to be −0.5 eV,42 which is much smaller than that of TiO2(110) (−0.87 eV43) and SnO2(110) (−1.5 eV37). Moreover, it can be seen as a system with weakly chemisorbed water.

In the following, we first validate the DFTMD simulations of BiVO4(010)/NaCl electrolyte interfaces by computing the corresponding surface pKas and comparing them to experiments and the results of TiO2(110) and SnO2(110) surfaces. It is followed by the structural and dynamical analyses of water orientation, the local coordination environment, and counter-ions in the protonic double layers of BiVO4(010). Finally, we link the degree of asymmetry in the computed Helmholtz capacitance and the adsorption energy of water, which shows a surprising positive correlation at both metal oxide and graphene-like surfaces.

The acidity constants and EDLs at BiVO4(010)/NaCl electrolyte interfaces are calculated at a fixed chemical composition of the supercell. The BiVO4(010) surface was modeled by symmetric periodic slabs of eight-layers with lateral dimensions of a 2 × 2 surface cell. Note that the slab size was tested by the convergence calculations of Eads and bandgap (see Table S1 of the supplementary material). The two symmetric slabs were separated by an electrolyte space of 30.5 Å, giving an orthorhombic supercell of dimensions 10.39 × 10.18 × 53.25 Å3. A molality of 2.7 mol/kg (the concentration of the electrolyte used in industry) for NaCl electrolyte was simulated in our systems. The electrolyte space between the BiVO4(010) slabs was filled with 102 water molecules, 5 Na+ and 5 Cl in each unit cell.

All the DFTMD simulations are carried out using the freely available program package CP2K/Quickstep44,45 with the PBE functional.46 Goedecker–Teter–Hutter (GTH) pseudopotentials are employed to represent the core electrons.47,48 The atomic basis sets for the valence electrons were the standard short-ranged double-ζ basis functions, with one set of polarization functions (DZVP) optimized for molecular systems.49 The plane wave cutoff for electron density expansion is set at 400 Ry. The target accuracy for the SCF convergence was 3 × 10−6 a.u. The NVT ensemble is used for MD propagation with a time step of 0.5 fs, and the equilibrium temperature for all the simulations is kept to 330 K by using the Nose–Hoover thermostat.

For further details on acidity calculations and finite-field DFTMD simulations of BiVO4(010)-NaCl interfaces, see the supplementary material.

The classical MD simulations of the graphene/NaCl interfaces were conducted using the LAMMPS software package.50 The electrode material is composed of two graphene layers with a surface area of 51.1 × 49.2 Å2. The two electrodes were separated by a vacuum region of 13.2 Å, giving the total length of the simulated box to be 70 Å. The NaCl electrolyte is comprised of 4100 water molecules and 74 pairs of NaCl ions, giving a concentration of 1 M. The computational details of classical MD simulations are available in the supplementary material.

When the BiVO4 electrode is immersed in an aqueous electrolyte, the protonation or deprotonation of surface groups depends on their pKa values and pH in the solution. There is a special pH at which the concentration of H+ on the surface equals that of OH, i.e., the point of zero charge (PZC).62 PZC can be measured using the potentiometric titration method and electrochemical in situ spectroscopy technologies.53,63 However, the intrinsic pKa values of individual surface groups are typically unable to be characterized by experiments. Herein, we first calculate the acidity constants of the individual adsorption sites on the BiVOa(010)/NaCl interface from DFTMD simulations and free energy perturbation theory by applying harmonic restraining potentials VrH to all O–H bonds (see Table S2 of the supplementary material).64,65

At the BiVO4(010) surface, there are two types of surface sites capable of protonation or deprotonation (see Fig. 1): the hydroxylated sevenfold coordinated Bi groups and the bridging oxygen Obr connecting Bi and V groups,66 
(1)
(2)
where KH1 and KH2 are the deprotonation equilibrium constants, and the corresponding acidity constants pKa1 and pKa2 on the BiVO4(010) surface can be obtained from the expression of pKa = −log KH.
FIG. 1.

DFTMD models and radial distribution functions (RDFs) of the BiVO4(010)/NaCl interface used in the full reaction scheme of surface acidity calculations. Models for the Bi7fOH2/Bi7fOH groups (a) and BiVObrH+/BiVObr group (b). RDFs between Bi7fOH2/Bi7fOH groups and neighboring protons from the solvent (c) and RDFs between BiVObrH+/BiVObr groups and neighboring protons (d). Bi, V, O, H, Na, Cl, and the target proton are distinguished in purple, gray, red, white, blue, cyan, and black, respectively. The molecules involving protonation/deprotonation reactions are highlighted by a ball and stick representation. The numbers shown in (c) and (d) are the corresponding H-bond coordination numbers.

FIG. 1.

DFTMD models and radial distribution functions (RDFs) of the BiVO4(010)/NaCl interface used in the full reaction scheme of surface acidity calculations. Models for the Bi7fOH2/Bi7fOH groups (a) and BiVObrH+/BiVObr group (b). RDFs between Bi7fOH2/Bi7fOH groups and neighboring protons from the solvent (c) and RDFs between BiVObrH+/BiVObr groups and neighboring protons (d). Bi, V, O, H, Na, Cl, and the target proton are distinguished in purple, gray, red, white, blue, cyan, and black, respectively. The molecules involving protonation/deprotonation reactions are highlighted by a ball and stick representation. The numbers shown in (c) and (d) are the corresponding H-bond coordination numbers.

Close modal
Adding reactions 1 and 2 leads to the PZC, while subtracting reaction 2 from reaction 1 causes the dissociation of surface adsorbed water,
(3)
(4)
where the equilibrium constants KPZC = [H+]PZC2 = KH1KH2 and Kd = KH1/KH2, and the dissociation free energy of Eq. (4) can be calculated from the expression of ΔAdiss = −kBT ln Kd. Accordingly, the relations among pKa1, pKa2, PZC, and ΔAdiss are derived as follows:
(5)
(6)

In this work, we have used the full reaction scheme in the calculation of the surface acidity constants developed previously,65,67,68 as illustrated in Fig. 1. The deprotonation of the Bi7fOH2 group was coupled to a simultaneous protonation of H2O in the bulk electrolyte of the same inhomogeneous model system [Fig. 1(a), Reaction S2 of the supplementary material]. Similarly, the deprotonation process of a BiVObrH+ group was accompanied by another protonation of H2O in the bulk electrolyte [Fig. 1(b), Reaction S3 of the supplementary material]. The reverse of this reaction indicates a bridging oxygen site receiving a proton from the acidic bulk electrolyte. The corresponding deprotonation free energies ΔA1 and ΔA2 were calculated from a coupling parameter integral of the corresponding vertical energy gap (see Figs. S1 and S2 of the supplementary material). The pKa values we found for the two active surface hydroxyl groups are 6.2 and 0.8, respectively, giving the PZC value of 3.5 (Table I), in excellent agreement with the experimental characterization 2.5–4.551–53 and previous theoretical works.66 

TABLE I.

Comparison between computed pKa values for pKa1, pKa2, PZC, experimental values PZCexp., ΔpKa, the dissociation free energy of adsorbed water ΔAdiss (eV) and the water adsorption energy Eads (eV) on the BiVO4(010), TiO2(110), and SnO2(110) surfaces.

SystempKa1pKa2PZCPZCexp.ΔpKaΔAdissEadsa
BiVO4(010)/NaCl sol. 6.2 0.8 3.5 2.5–4.551–53  5.4 0.32 −0.58 
TiO2(110)/H2O43,54 6.3 2.4 4.4 4.5–6.055–57  3.9 0.23 −0.87 
SnO2(110)/H2O37,58 4.4 4.7 4.6 4.1–5.559–61  −0.3 −0.02 −1.5 
SystempKa1pKa2PZCPZCexp.ΔpKaΔAdissEadsa
BiVO4(010)/NaCl sol. 6.2 0.8 3.5 2.5–4.551–53  5.4 0.32 −0.58 
TiO2(110)/H2O43,54 6.3 2.4 4.4 4.5–6.055–57  3.9 0.23 −0.87 
SnO2(110)/H2O37,58 4.4 4.7 4.6 4.1–5.559–61  −0.3 −0.02 −1.5 
a

The Eads of BiVO4, TiO2, and SnO2 were calculated for one monolayer of water.

Since the H-bond network has a significant effect on pKa values, both the coordination numbers and the length of the H-bond can lead to a vital variation in the picture of surface acidity. To further interpret the effect of water adsorption energy Eads on pKa shifts, the RDFs between oxygen atoms of surface groups and hydrogen atoms in electrolyte are shown in Fig. 1. The comparison of H-bond coordination numbers of BiVO4 with those of TiO2(110) and SnO2(110) surfaces is shown in Table II. Note that the H-bond coordination numbers are obtained by integrating the first peaks of RDFs in Figs. 1(c) and 1(d), and the length of the H-bond for oxygen atoms on the BiVO4 surface is determined by the corresponding position of the first peak. The Eads of BiVO4(010) surface was calculated to −0.58 eV by one monolayer of adsorbed molecular water (Table I).

TABLE II.

Variation of the H-bond coordination numbers of protonated and deprotonated groups on the BiVO4(010), TiO2(110), and SnO2(110) surfaces. M refers to the active metal site on the oxide surface.

SystemMOH2MOHM2OH+M2O
BiVO4(010)/NaCl sol. 0.97 2.0 0.0 1.0 
TiO2(110)/H2O43  0.6 1.9 1.0 1.1 
SnO2(110)/H2O58  0.3 2.0 1.0 1.5 
SystemMOH2MOHM2OH+M2O
BiVO4(010)/NaCl sol. 0.97 2.0 0.0 1.0 
TiO2(110)/H2O43  0.6 1.9 1.0 1.1 
SnO2(110)/H2O58  0.3 2.0 1.0 1.5 

It is known that the surface metal–O bond strength is manifested by the corresponding bond-length.69 Because of the weakest interaction between the adsorbed water molecules and BiVO4 surface sites, it is found that the bond length of Bi7f and oxygen in the adsorbed water molecules is 2.6 Å, which is significantly larger than that of TiO2(110) and SnO2(110) surfaces (∼2.0 Å).43 On the other hand, the interaction between adsorbed water and solvent water in an electrolyte increases when the strength of water adsorption weakens. It is indeed found to have a higher H-bond coordination number of 0.97 for the Bi7fOH2 group than that of the TiOH2 (0.6) and SnOH2 (0.3) groups (Table II). In the case of the deprotonated group, it is found that the length of the H-bond of Bi7fOH [first peak of the blue line in Fig. 1(c)] is significantly smaller than that of Bi7fOH2 [first peak of the blue dashed line in Fig. 1(c)], while the length of the H-bond of TiOH is the same as that of TiOH2.43 One would expect that a shorter H-bond (stronger H-bonding) formed by Bi7fOH makes the deprotonated structure more stable, and more H-bonds (stronger H-bonding) formed on the Bi7fOH2 site make the protonated state more stable. Indeed, these two effects appear to cancel out each other and leave the resulting pKa1 of the BiVO4 surface similar to that of TiO2 (Table I). Meanwhile, it was found that BiVObrH+ has a significantly lower coordination number 0.0 than that of the Ti2ObrH+ (1.0) and Sn2ObrH+ (1.0) groups (Table II). This difference is much bigger than the lowering of the coordination number for deprotonated state BiVObr in the oxide series. Therefore, the protonated state of BiVObrH+ is much less stable, and the resulting pKa2 of the BiVO4 surface is the smallest.

Based on the values of pKa, we obtained the dissociation free energy of adsorbed water ΔAdiss on the BiVO4(010) surface according to Eq. (6), giving a value of 0.32 eV. There have always been different views on the dissociation of chemisorbed water at the BiVO4 interface. Previous theoretical studies had not reported or observed the dissociation of adsorbed water on the BiVO4(010) surface until Wang and Galli found that the dissociation of water does occur on the BiVO4(010) surface using ambient pressure resonant photoemission spectroscopy (AP-ResPES) measurements and first-principles simulations.2 It is worth noting that water dissociation on the pristine rutile TiO2(110) surface has been controversial for both the experiment and theory, with the dissociation free energy of adsorbed water ranging from 0.23 eV54 to 0.45 eV70 with density functional theory (DFT) studies. Compared to TiO2, BiVO4 has a slightly higher ΔAdiss (see Table I). Therefore, it is plausible for the spontaneous dissociation of adsorbed water on the BiVO4 surface with a sufficient timescale, although we did not observe it during the course of our DFTMD simulations at PZC.

Among rutile TiO2(110) and SnO2(110) and BiVO4(010) surfaces, the SnO2(110) surface has the strongest water adsorption energy of −1.5 eV with DFT studies37 (see Table I). Therefore, one would expect water molecules to have a stronger tendency for dissociative adsorption and ΔAdiss to be smaller. This is indeed the case, as shown in Table I.

In previous reports, it has been known that the local surface potentials and photoelectrochemical performance depend strongly on the pH of the ambient electrolyte40,53 and the interaction of H2O with the BiVO4 surface.2,71 To further understand how the chemisorbed water response to surface potential affects the molecular structure and capacitance of EDLs, we next simulate the electrified BiVO4/NaCl interface with finite-field DFTMD simulations (Fig. S3 of the supplementary material).

At low pH, the BiVO4 surface is positively charged by adsorbing protons, and the surface charge density σ = 30.3 μC/cm2. At high pH, the BiVO4 surface is negatively charged by the desorption proton from terminal Bi7fOH2, and σ = −30.3 μC/cm2. Detailed descriptions of the computational setup are given in the supplementary material. When looking at the structure of EDLs, it is found that the electrified BiVO4/NaCl interfaces are very different from those of rutile oxide TiO2 and SnO2, due to weaker water adsorption energy.

As we have seen in metal oxide systems, the adsorbed water molecules point toward the electrolyte solution [Fig. 2(a),
FIG. 2.

Comparison of the probability distribution of the angle θ between the bisector of adsorbed water molecules and the surface normal for (a) BiVO4(010)/NaCl, (b) TiO2/NaCl, and (c) SnO2/NaCl interfaces at different surface charge densities σ. The probability distribution for TiO2/NaCl and SnO2/NaCl interfaces is taken from Refs. 36 and 37, respectively.

FIG. 2.

Comparison of the probability distribution of the angle θ between the bisector of adsorbed water molecules and the surface normal for (a) BiVO4(010)/NaCl, (b) TiO2/NaCl, and (c) SnO2/NaCl interfaces at different surface charge densities σ. The probability distribution for TiO2/NaCl and SnO2/NaCl interfaces is taken from Refs. 36 and 37, respectively.

Close modal
], and the angle θ between the water dipole and surface normal is counted as 61° at the PZC. In response to electric field E in the protonic and deprotonic EDLs, the values of angle θ down-shifted and up-shifted, respectively, are consistent with previous DFTMD studies.36,37 However, the up-shifted values of θ show large differences at the negatively charged surface when comparing BiVO4, TiO2, and SnO2 [Figs. 2(a)2(c)] systems.

At BiVO4(010)/NaCl interfaces, the dynamics of adsorption and desorption processes can be presented by the time evolution of the Bi–Ow distance between surface Bi and oxygen in water, as shown in Figs. 3(a)3(c). When the Bi–Ow distance goes from ∼5 to ∼2.6 Å, it indicates the adsorption process of second layer water. When the Bi–Ow distance goes from ∼2.6 to ∼5 Å, it indicates the desorption process of the first layer of water. Interestingly, it is found that the dynamics of adsorption and desorption processes have a big difference between the first layer of water and the second layer of water at different charged BiVO4(010)/NaCl interfaces.

FIG. 3.

(a)–(c) Dynamics of the adsorption and desorption processes of first layer water and second layer water at different charged BiVO4(010)/NaCl interfaces. The corresponding DFTMD snapshots at different surface charge densities are described in the inset, respectively. (d) Variation of Bi7f-O RDFs between surface Bi7f sites and oxygen atoms of electrolyte water at different charged BiVO4(010)/NaCl interfaces. The numbers next to the first peak of the RDFs are the coordination numbers.

FIG. 3.

(a)–(c) Dynamics of the adsorption and desorption processes of first layer water and second layer water at different charged BiVO4(010)/NaCl interfaces. The corresponding DFTMD snapshots at different surface charge densities are described in the inset, respectively. (d) Variation of Bi7f-O RDFs between surface Bi7f sites and oxygen atoms of electrolyte water at different charged BiVO4(010)/NaCl interfaces. The numbers next to the first peak of the RDFs are the coordination numbers.

Close modal

At PZC, the adsorption process [black dashed line in Fig. 3(a)] is accompanied by the desorption process [gray dashed line in Fig. 3(a)] between the second layer of water and the first layer of water. Furthermore, it is observed that more than one water molecule adsorbed to the same Bi7f site at the positively charged BiVO4/NaCl interface [Fig. 3(b)]. That means the adsorption of first layer water is enhanced by the electric field in EDL at low pH. However, the adsorbed first layer of water partly desorbed into the near-surface at the negatively charged BiVO4/NaCl interface [Fig. 3(c)]. The results indicate that the adsorption of first layer water is weakened by the electric field in EDL at high pH.

The RDFs between Bi7f sites and oxygen atoms of electrolyte water at different charged BiVO4(010)/NaCl interfaces are described in Fig. 3(d). The average number of adsorbed water molecules for each Bi7f site is obtained by integrating the first peak of the Bi7f–O RDFs, giving 1.0, 1.3, and 0.6 at PZC, respectively, at positively charged and negatively charged interfaces. That means the amount of adsorbed water at the BiVO4 surface depends on its surface charge density, while it remains unchanged at the TiO2 and SnO2 surfaces. Although the adsorption and desorption processes can happen dynamically on the negatively charged side of TiO2, the number of adsorbed water is the same as in the cases of PZC and the positively charged side.36 These results give us a molecular-level understanding of the structure and dynamics of metal oxide electrodes with weakly chemisorbed water, which is different from the oxides with strong water adsorption energy.

It has been shown that the mechanism of the water oxidation reaction on the BiVO4 surface depends on the pH of the electrolyte.72 At low pH, the water oxidation reaction is initiated mainly through the dehydrogenation of adsorbed water molecules, which involves the H2O·+ radical cation as an intermediate.73 At high pH, the reaction is initiated through the oxidation of OH (OH/OH·).74 Based on observations here, the Bi7f–(OH2)2 structure may be more favorable to the formation of H2O·+ in acidic conditions. However, the desorption of terminally adsorbed water is a disadvantage to generate H2O·+ at alkaline conditions. This leaves the opportunity for Bi7fOH to combine hole and formate OH· at high pH. This is in line with the role of negative surface charge in hole selectivity reported by recent experiments75,76 and the ionized oxygen vacancy proposed by theoretical calculations.77 

In addition, such a weak interaction between interfacial water and BiVO4(010) surface groups gives the compensated counterions in the electrolyte the chance to have a direct and strong interaction with surface sites. Therefore, the inner sphere coordination formed at the BiVO4(010)/NaCl surface [see Figs. 4(a) and 4(b)]. The width of the Helmholtz layer dEDL is estimated at about 2.4 Å at low pH and 2.7 Å at high pH, respectively [Fig. 4(c)]. Note that the values of dEDL were obtained from the surface normal projected distance between the counterions and the nuclei of Obr sites. In addition, it is found that the adsorbed Cl on the Bi7f site at a positively charged surface [Fig. 4(d)] makes the width of dEDL smaller than that of the negatively charged surface. In the reported synchronous illumination XPS (SI-XPS) study, the Cl concentrates and bonds to Bi on the surface, which can facilitate H2O molecule activation as well as promote carrier charge separation and accelerate the water oxidation reaction.78,79 This may be explained by the adsorption of Cl, which shortens the width of dEDL and has a positive effect on Helmholtz’s capacitance, as seen here.

FIG. 4.

(a) and (b) The Helmholtz layer models of the electrified BiVO4(010)/NaCl interface at σ = 30.3 and −30.3 μC/cm2. (c) Dynamics of the width of the Helmholtz layer dEDL at σ = 30.3 and −30.3 μC/cm2. The values of dEDL were estimated from the average distance between the counterions and the nuclei of Obr sites, which corresponds to the width of the dashed black lines in Figs. 4(a) and 4(b). (d) Dynamics of the bond length of adsorbed Cl and surface site Bi7f at the positively charged BiVO4(010)/NaCl interface (σ = 30.3 μC/cm2). It is worth noting that the positions of Obr and Bi7f are not on the same plane along the z direction, and the bond length of Bi7f–Cl is different from the value of dEDL.

FIG. 4.

(a) and (b) The Helmholtz layer models of the electrified BiVO4(010)/NaCl interface at σ = 30.3 and −30.3 μC/cm2. (c) Dynamics of the width of the Helmholtz layer dEDL at σ = 30.3 and −30.3 μC/cm2. The values of dEDL were estimated from the average distance between the counterions and the nuclei of Obr sites, which corresponds to the width of the dashed black lines in Figs. 4(a) and 4(b). (d) Dynamics of the bond length of adsorbed Cl and surface site Bi7f at the positively charged BiVO4(010)/NaCl interface (σ = 30.3 μC/cm2). It is worth noting that the positions of Obr and Bi7f are not on the same plane along the z direction, and the bond length of Bi7f–Cl is different from the value of dEDL.

Close modal
Based on the finite-field DFTMD simulations, the calculated results of the average Helmholtz capacitance CH, protonic Helmholtz capacitances CH+, and deprotonic Helmholtz capacitance CH at the BiVO4/NaCl interfaces are listed in Table III. The average CH was calculated from the expression below based on the supercell polarization Pz,39 
(7)
where Lz is the dimension of the supercell perpendicular to the interface, q is the introduced surface proton charge, and A is the area of the x and y cross sections. Pz indicates the ensemble averaged supercell polarization, obtained from the average total dipole moment Mz as shown in Fig. S4 of the supplementary material. Accordingly, the value of CH for the BiVO4/NaCl interface is calculated to be 77 μF/cm2. Based on the shift of macro-averaged electrostatic potential (Δϕ) with respect to the PZC (see Fig. S5 of the supplementary material), the CH+ and CH are calculated to be 74 μF/cm2 and 77 μF/cm2, respectively (Table III).
TABLE III.

Comparison of average Helmholtz capacitance CH (μF/cm2), the Helmholtz capacitance of the positively charged surface CH+ (μF/cm2), the Helmholtz capacitance of the negatively charged surface CH (μF/cm2), the Helmholtz capacitance asymmetry ΔCHCH = CH - CH+), and the asymmetry factor γ in the Helmholtz capacitance for BiVO4(010)/NaCl, TiO2/NaCl, and SnO2/NaCl interfaces.

SystemσCHCH+CHΔCHγ
BiVO4(010)/NaCl sol. 30 77 74 77 0.04 
TiO2(110)/NaCl sol.36  38 72 59 85 26 0.36 
SnO2(110)/NaCl sol.37  40 73 58 101 43 0.59 
SystemσCHCH+CHΔCHγ
BiVO4(010)/NaCl sol. 30 77 74 77 0.04 
TiO2(110)/NaCl sol.36  38 72 59 85 26 0.36 
SnO2(110)/NaCl sol.37  40 73 58 101 43 0.59 

When looking at Table III and Fig. 2 together, one can see the shift in the orientation peak of water at positively charged surface Δθ+ is the largest in the case of BiVO4, which correlates well with CH+ (BiVO4) being the highest in the series. On the other hand, the shift in the orientation peak of water at negatively charged surface Δθ is the smallest in the case of BiVO4, which also correlates with the lowest capacitance value CH(BiVO4) among these three oxide systems. However, this trend is not clear when just comparing TiO2 and SnO2. This means this qualitative picture based on water orientation alone is not sufficient, and the incorporation of the different water adsorption strengths and their effects on polarization is needed in order to be quantitative.

In previous work, it has been shown that water chemisorption would explain asymmetric EDLs seen at aqueous interfaces of oxide.36,37 In this work, it is found that EDLs are indeed nearly symmetric on the BiVO4(010) surface, which possesses the weakest chemisorption of water molecules (Table III). When looking at the capacitance asymmetry of EDLs in metal oxide systems, a conspicuous monotonic increase is found in ΔCH when increasing the water adsorption energy of Eads, as shown in Fig. 5(c).

To verify this relationship between the capacitance asymmetry ΔCH and Eads at other types of aqueous interfaces, i.e., the metallic electrode with weakly physisorbed water, the graphene/NaCl electrolyte interface was selected and simulated with classical force fields and constant-potential MD (see the supplementary material and Fig. S6).80 It is worth noting that the nature of EDL at the metallic interface is quite different from that of the (insulating/semiconducting) oxide interface; the former is formed by the electronic charge [Fig. 5(b)], while the latter is formed by the proton charge [Fig. 5(a)].

FIG. 5.

(a) and (b) Schematic representation of the EDLs at the oxide/electrolyte interface and graphene/electrolyte interface. (c) The capacitance asymmetry ΔCH for BiVO4(010)/NaCl, TiO2/NaCl, and SnO2/NaCl surfaces as a function of water adsorption energy Eads (R2 = 0.91). The capacitances for TiO2 and SnO2 systems were taken from Refs. 36 and 37, respectively. (d) The capacitance asymmetry ΔCH for graphene/NaCl interfaces as a function of water adsorption energy Eads from classical MD simulations (R2 = 0.94). The model of the two-layer graphene/electrolyte interface used in classical MD simulations is shown in Fig. S6. The right y-axis corresponds to the asymmetry factor γ [Eq. (8)] in the Helmholtz capacitance. The absolute values of Eads are taken from these figures.

FIG. 5.

(a) and (b) Schematic representation of the EDLs at the oxide/electrolyte interface and graphene/electrolyte interface. (c) The capacitance asymmetry ΔCH for BiVO4(010)/NaCl, TiO2/NaCl, and SnO2/NaCl surfaces as a function of water adsorption energy Eads (R2 = 0.91). The capacitances for TiO2 and SnO2 systems were taken from Refs. 36 and 37, respectively. (d) The capacitance asymmetry ΔCH for graphene/NaCl interfaces as a function of water adsorption energy Eads from classical MD simulations (R2 = 0.94). The model of the two-layer graphene/electrolyte interface used in classical MD simulations is shown in Fig. S6. The right y-axis corresponds to the asymmetry factor γ [Eq. (8)] in the Helmholtz capacitance. The absolute values of Eads are taken from these figures.

Close modal

The adsorption energy was calculated from a single water molecule with a two-leg structure adsorbed on the surface of a two-layer graphene electrode (see Fig. S7 of the supplementary material). The strength of the water adsorption energy Eads of graphene electrodes was changed by adjusting the energy parameter ɛ of the Lennard-Jones potential between water molecules and graphene. The Eads in the classical MD simulations with energy parameters of 0.3ɛ, 0.7ɛ, ɛ, 1.5ɛ, and 2.0ɛ were calculated as 0.05, 0.08, 0.10, 0.14, and 0.17 eV, respectively. In the classical MD simulations, the Helmholtz capacitance (CH) is calculated from the definition of CH = dσ/d(Δϕ), and Δϕ is the electrostatic potential difference between the top graphene layer and bulk electrolyte, as shown in Fig. S8 of the supplementary material. The calculated CH+ and CH for electrified graphene/NaCl electrolyte interfaces with different adsorption capabilities are described in Fig. S9 of the supplementary material.

Based on the classical MD simulations, it is found that the ΔCH shows a similar monotonic increment as a function of Eads at electrified graphene/NaCl electrolyte interfaces [Fig. 5(d)]. This result indicates that the positive correlation that we found on oxide surfaces is also seen in graphene-like systems.

To further compare these two types of electrode surfaces, we introduce here a metric called the asymmetry factor γ,
(8)

The values of the asymmetry factor γ are shown on the right side of Figs. 5(c) and 5(d). Despite the fact that the absolute values of CH are dramatically different between metal-oxide and graphene-like surfaces, the asymmetry factor γ shows similar values. This indicates that the relationship between Eads and γ can be rather general across different types of electrode materials, although different scaling factors apply depending on the nature of EDLs. Furthermore, it is worth noting that the water adsorption energy at the graphene surface, as determined from highly accurate electronic structure calculations, is about 0.1 eV,81 which corresponds to a high value in the asymmetry factor γ of 0.25 in our quantification. This can be the physical reason behind the observed asymmetric response of the contact angle in the electrowetting experiment recently performed on this type of material.82 

In conclusion, by applying finite-field DFTMD simulations, the effect of water adsorption energy Eads on surface acidity and Helmholtz capacitance has been studied by selecting BiVO4 as an example of metal oxide electrodes with weakly chemisorbed water. Based on an oxide series composed of BiVO4, TiO2, and SnO2, we investigate the effect of Eads on the coordination number, the length of the H-bond, and the orientation of chemisorbed water. In particular, it is found that a positive correlation exists between the degree of asymmetry in the Helmholtz capacitance and the strength of the Eads. This correlation is further verified and extended to graphene-like systems with physisorbed water.

Overall, these observations in this work provide a general molecular picture of an electric double layer with different Eads. Our findings likely lay the foundation for future investigations of water reactivity and wetting in electrified semiconducting metal oxides and carbonaceous materials.

See the supplementary material for the following additional data: the convergence tests for the slab size, harmonic restraining potentials used in the acidity calculations, calculations of the deprotonation free energy ΔA, finite-field DFTMD simulations of electrified BiVO4(010)–NaCl interfaces, and classical MD simulations of electrified graphene–NaCl electrolyte interfaces.

This project received funding from the European Research Council (ERC) under the European Union’s Horizon 2020 Research and Innovation Program (Grant Agreement No. 949012). Mei Jia greatly appreciates the financial support from the Natural Science Foundation of Henan Province (Grant No. 242300420420570), the Key Scientific Research Projects of Colleges and Universities in Henan Province (Grant No. 24A150031), and the International Scientific and Technological Cooperation Projects in Henan Province (Grant No. 232102520020). We thank S. Han for the discussions and inputs during the initiation of this project.

The authors have no conflicts to disclose.

Mei Jia: Data curation (equal); Formal analysis (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Writing – original draft (equal); Writing – review & editing (equal). Junyi Wang: Data curation (supporting); Formal analysis (supporting); Investigation (supporting); Visualization (supporting). Qixiang Liu: Formal analysis (supporting); Writing – review & editing (supporting). Xiaohui Yang: Data curation (supporting); Formal analysis (supporting); Investigation (supporting); Writing – review & editing (supporting). Chao Zhang: Conceptualization (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Project administration (equal); Resources (equal); Supervision (equal); Writing – original draft (equal); Writing – review & editing (equal).

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

1.
W.
Stumm
,
Chemistry of the Solid-Water Interface: Processes at the Mineral-Water and Particle-Water Interface in Natural Systems
(
John Wiley & Son, Inc.
,
1992
).
2.
W.
Wang
,
M.
Favaro
,
E.
Chen
,
L.
Trotochaud
,
H.
Bluhm
,
K.-S.
Choi
,
R.
van de Krol
,
D. E.
Starr
, and
G.
Galli
, “
Influence of excess charge on water adsorption on the BiVO4(010) surface
,”
J. Am. Chem. Soc.
144
,
17173
17185
(
2022
).
3.
V. G.
Gude
,
N.
Nirmalakhandan
, and
S.
Deng
, “
Renewable and sustainable approaches for desalination
,”
Renewable Sustainable Energy Rev.
14
,
2641
2654
(
2010
).
4.
B.
Wang
,
X.
Zhou
,
Z.
Guo
, and
W.
Liu
, “
Recent advances in atmosphere water harvesting: Design principle, materials, devices, and applications
,”
Nano Today
40
,
101283
(
2021
).
5.
E.
Barry
,
R.
Burns
,
W.
Chen
,
G. X.
De Hoe
,
J. M. M.
De Oca
,
J. J.
de Pablo
,
J.
Dombrowski
,
J. W.
Elam
,
A. M.
Felts
,
G.
Galli
,
J.
Hack
,
Q.
He
,
X.
He
,
E.
Hoenig
,
A.
Iscen
,
B.
Kash
,
H. H.
Kung
,
N. H. C.
Lewis
,
C.
Liu
,
X.
Ma
,
A.
Mane
,
A. B. F.
Martinson
,
K. L.
Mulfort
,
J.
Murphy
,
K.
Mølhave
,
P.
Nealey
,
Y.
Qiao
,
V.
Rozyyev
,
G. C.
Schatz
,
S. J.
Sibener
,
D.
Talapin
,
D. M.
Tiede
,
M. V.
Tirrell
,
A.
Tokmakoff
,
G. A.
Voth
,
Z.
Wang
,
Z.
Ye
,
M.
Yesibolati
,
N. J.
Zaluzec
, and
S. B.
Darling
, “
Advanced materials for energy-water systems: The central role of water/solid interfaces in adsorption, reactivity, and transport
,”
Chem. Rev.
121
,
9450
9501
(
2021
).
6.
A.
Groß
and
S.
Sakong
, “
Ab initio simulations of water/metal interfaces
,”
Chem. Rev.
122
,
10746
10776
(
2022
).
7.
L.
Knijff
,
M.
Jia
, and
C.
Zhang
, “
Electric double layer at the metal-oxide/electrolyte interface
,” in
Encyclopedia of Solid-Liquid Interfaces
, 1st ed., edited by
K.
Wandelt
and
G.
Bussetti
(
Elsevier
,
Oxford
,
2024
), pp.
567
575
.
8.
A. J.
Nozik
and
R.
Memming
, “
Physical chemistry of semiconductor−liquid interfaces
,”
J. Phys. Chem.
100
,
13061
13078
(
1996
).
9.
J. L.
Bañuelos
,
E.
Borguet
,
G. E.
Brown
,
R. T.
Cygan
,
J. J.
DeYoreo
,
P. M.
Dove
,
M. P.
Gaigeot
,
F. M.
Geiger
,
J. M.
Gibbs
,
V. H.
Grassian
,
A. G.
Ilgen
,
Y.-S.
Jun
,
N. e.
Kabengi
,
L.
Katz
,
J. D.
Kubicki
,
J.
Lützenkirchen
,
C. V.
Putnis
,
R. C.
Remsing
,
K. M.
Rosso
,
G.
Rother
,
M.
Sulpizi
,
M.
Villalobos
, and
H.
Zhang
, “
Oxide– and silicate–water interfaces and their roles in technology and the environment
,”
Chem. Rev.
123
,
6413
6544
(
2023
).
10.
C.
Zhang
,
J.
Cheng
,
Y.
Chen
,
M. K. Y.
Chan
,
Q.
Cai
,
R. P.
Carvalho
,
C. F. N.
Marchiori
,
D.
Brandell
,
C. M.
Araujo
,
M.
Chen
,
X.
Ji
,
G.
Feng
,
K.
Goloviznina
,
A.
Serva
,
M.
Salanne
,
T.
Mandai
,
T.
Hosaka
,
M.
Alhanash
,
P.
Johansson
,
Y.-Z.
Qiu
,
H.
Xiao
,
M.
Eikerling
,
R.
Jinnouchi
,
M. M.
Melander
,
G.
Kastlunger
,
A.
Bouzid
,
A.
Pasquarello
,
S.-J.
Shin
,
M. M.
Kim
,
H.
Kim
,
K.
Schwarz
, and
R.
Sundararaman
, “
2023 Roadmap on molecular modelling of electrochemical energy materials
,”
J. Phys.: Energy
5
,
041501
(
2023
).
11.
J.
Wu
, “
Understanding the electric double-layer structure, capacitance, and charging dynamics
,”
Chem. Rev.
122
,
10821
10859
(
2022
).
12.
L.
Zeng
,
J.
Peng
,
J.
Zhang
,
X.
Tan
,
X.
Ji
,
S.
Li
, and
G.
Feng
, “
Molecular dynamics simulations of electrochemical interfaces
,”
J. Chem. Phys.
159
,
091001
(
2023
).
13.
T.
Pham
,
Y.
Ping
, and
G.
Galli
, “
Modelling heterogeneous interfaces for solar water splitting
,”
Nat. Mater.
16
,
401
408
(
2017
).
14.
S.
Louisia
,
M. T.
Koper
, and
R. V.
Mom
, “
Prospects for electrochemical x-ray photoelectron spectroscopy as a powerful electrochemical interface characterization technique
,”
Curr. Opin. Electrochem.
45
,
101462
(
2024
).
15.
Y.
Hsiao
,
T.-H.
Chou
,
A.
Patra
, and
Y.-C.
Wen
, “
Momentum-dependent sum-frequency vibrational spectroscopy of bonded interface layer at charged water interfaces
,”
Sci. Adv.
9
,
eadg2823
(
2023
).
16.
C.-Y.
Li
,
J.-B.
Le
,
Y.-H.
Wang
,
S.
Chen
,
Z.-L.
Yang
,
J.-F.
Li
,
J.
Cheng
, and
Z.-Q.
Tian
, “
In situ probing electrified interfacial water structures at atomically flat surfaces
,”
Nat. Mater.
18
,
697
(
2019
).
17.
Y.-H.
Wang
,
S.
Zheng
,
W.-M.
Yang
,
R.-Y.
Zhou
,
Q.-F.
He
,
P.
Radjenovic
,
J.-C.
Dong
,
S.
Li
,
J.
Zheng
,
Z.-L.
Yang
et al, “
In situ Raman spectroscopy reveals the structure and dissociation of interfacial water
,”
Nature
600
,
81
85
(
2021
).
18.
S.
Su
,
I.
Siretanu
,
D.
van den Ende
,
B.
Mei
,
G.
Mul
, and
F.
Mugele
, “
Facet-dependent surface charge and hydration of semiconducting nanoparticles at variable ph
,”
Adv. Mater.
33
,
2106229
(
2021
).
19.
C.
Santana Santos
,
B. N.
Jaato
,
I.
Sanjuán
,
W.
Schuhmann
, and
C.
Andronescu
, “
Operando scanning electrochemical probe microscopy during electrocatalysis
,”
Chem. Rev.
123
,
4972
5019
(
2023
).
20.
J.
Haruyama
,
T.
Ikeshoji
, and
M.
Otani
, “
Electrode potential from density functional theory calculations combined with implicit solvation theory
,”
Phys. Rev. Mater.
2
,
095801
(
2018
).
21.
G.
Jeanmairet
,
B.
Rotenberg
,
D.
Borgis
, and
M.
Salanne
, “
Study of a water-graphene capacitor with molecular density functional theory
,”
J. Chem. Phys.
151
,
124111
(
2019
).
22.
H.-K.
Lim
,
H.
Lee
, and
H.
Kim
, “
A seamless grid-based interface for mean-field QM/MM coupled with efficient solvation free energy calculations
,”
J. Chem. Theory Comput.
12
,
5088
5099
(
2016
).
23.
H.-K.
Lim
,
Y.
Kwon
,
H. S.
Kim
,
J.
Jeon
,
Y.-H.
Kim
,
J.-A.
Lim
,
B.-S.
Kim
,
J.
Choi
, and
H.
Kim
, “
Insight into the microenvironments of the metal–ionic liquid interface during electrochemical CO2 reduction
,”
ACS Catal.
8
,
2420
2427
(
2018
).
24.
J.
Cheng
and
M.
Sprik
, “
The electric double layer at a rutile TiO2 water interface modelled using density functional theory based molecular dynamics simulation
,”
J. Phys.: Condens. Matter
26
,
244108
(
2014
).
25.
P.
Li
,
Y.
Jiao
,
J.
Huang
, and
S.
Chen
, “
Electric double layer effects in electrocatalysis: Insights from ab initio simulation and hierarchical continuum modeling
,”
JACS Au
3
,
2640
2659
(
2023
).
26.
J.-B.
Le
,
Q.-Y.
Fan
,
J.-Q.
Li
, and
J.
Cheng
, “
Molecular origin of negative component of Helmholtz capacitance at electrified Pt(111)/water interface
,”
Sci. Adv.
6
,
eabb1219
(
2020
).
27.
J.-B.
Le
,
X.-H.
Yang
,
Y.-B.
Zhuang
,
M.
Jia
, and
J.
Cheng
, “
Recent progress toward ab initio modeling of electrocatalysis
,”
J. Phys. Chem. Lett.
12
,
8924
8931
(
2021
).
28.
V.
Quaranta
,
J.
Behler
, and
M.
Hellström
, “
Structure and dynamics of the liquid–water/zinc-oxide interface from machine learning potential simulations
,”
J. Phys. Chem. C
123
,
1293
1304
(
2018
).
29.
M. F.
Calegari Andrade
,
H.-Y.
Ko
,
L.
Zhang
,
R.
Car
, and
A.
Selloni
, “
Free energy of proton transfer at the water–TiO2 interface from ab initio deep potential molecular dynamics
,”
Chem. Sci.
11
,
2335
2341
(
2020
).
30.
T.
Dufils
,
L.
Knijff
,
Y.
Shao
, and
C.
Zhang
, “
PiNNwall: Heterogeneous electrode models from integrating machine learning and atomistic simulation
,”
J. Chem. Theory Comput.
19
,
5199
5209
(
2023
).
31.
L.
Scalfi
,
M.
Salanne
, and
B.
Rotenberg
, “
Molecular simulation of electrode-solution interfaces
,”
Annu. Rev. Phys. Chem.
72
,
189
212
(
2021
).
32.
G.
Jeanmairet
,
B.
Rotenberg
, and
M.
Salanne
, “
Microscopic simulations of electrochemical double-layer capacitors
,”
Chem. Rev.
122
,
10860
10898
(
2022
).
33.
R.
Sundararaman
,
D.
Vigil-Fowler
, and
K.
Schwarz
, “
Improving the accuracy of atomistic simulations of the electrochemical interface
,”
Chem. Rev.
122
,
10651
10674
(
2022
).
34.
M.
Stengel
,
N. A.
Spaldin
, and
D.
Vanderbilt
, “
Electric displacement as the fundamental variable in electronic-structure calculations
,”
Nat. Phys.
5
,
304
308
(
2009
).
35.
T.
Sayer
,
M.
Sprik
, and
C.
Zhang
, “
Finite electric displacement simulations of polar ionic solid-electrolyte interfaces: Application to NaCl(111)/aqueous NaCl solution
,”
J. Chem. Phys.
150
,
041716
(
2019
).
36.
C.
Zhang
,
J.
Hutter
, and
M.
Sprik
, “
Coupling of surface chemistry and electric double layer at TiO2 electrochemical interfaces
,”
J. Phys. Chem. Lett.
10
,
3871
3876
(
2019
).
37.
M.
Jia
,
C.
Zhang
, and
J.
Cheng
, “
Origin of asymmetric electric double layers at electrified oxide/electrolyte interfaces
,”
J. Phys. Chem. Lett.
12
,
4616
4622
(
2021
).
38.
C.
Zhang
,
T.
Sayer
,
J.
Hutter
, and
M.
Sprik
, “
Modelling electrochemical systems with finite field molecular dynamics
,”
J. Phys. Energy
2
,
032005
(
2020
).
39.
C.
Zhang
and
M.
Sprik
, “
Finite field methods for the supercell modeling of charged insulator/electrolyte interfaces
,”
Phys. Rev. B
94
,
245309
(
2016
).
40.
T. W.
Kim
and
K.-S.
Choi
, “
Nanoporous BiVO4 photoanodes with dual-layer oxygen evolution catalysts for solar water splitting
,”
Science
343
,
990
994
(
2014
).
41.
J.
Chi
,
Z.
Jiang
,
J.
Yan
,
A.
Larimi
,
Z.
Wang
,
L.
Wang
, and
W.
Shangguan
, “
Recent advancements in bismuth vanadate photoanodes for photoelectrochemical water splitting
,”
Mater. Today Chem.
26
,
101060
(
2022
).
42.
R.
Crespo-Otero
and
A.
Walsh
, “
Variation in surface ionization potentials of pristine and hydrated BiVO4
,”
J. Phys. Chem. Lett.
6
,
2379
2383
(
2015
).
43.
J.
Cheng
and
M.
Sprik
, “
Acidity of the aqueous rutile TiO2(110) surface from density functional theory based molecular dynamics
,”
J. Chem. Theory Comput.
6
,
880
(
2010
).
44.
J.
Hutter
,
M.
Iannuzzi
,
F.
Schiffmann
, and
J.
VandeVondele
, “
CP2K: Atomistic simulations of condensed matter systems
,”
WIREs Computat. Mol. Sci.
4
,
15
25
(
2014
).
45.
J.
VandeVondele
,
M.
Krack
,
F.
Mohamed
,
M.
Parrinello
,
T.
Chassaing
, and
J.
Hutter
, “
Quickstep: Fast and accurate density functional calculations using a mixed Gaussian and plane waves approach
,”
Comput. Phys. Commun.
167
,
103
128
(
2005
).
46.
J. P.
Perdew
,
K.
Burke
, and
M.
Ernzerhof
, “
Generalized gradient approximation made simple
,”
Phys. Rev. Lett.
77
,
3865
3868
(
1996
).
47.
C.
Hartwigsen
,
S.
Goedecker
, and
J.
Hutter
, “
Relativistic separable dual-space Gaussian pseudopotentials from H to Rn
,”
Phys. Rev. B
58
,
3641
3662
(
1998
).
48.
S.
Goedecker
,
M.
Teter
, and
J.
Hutter
, “
Separable dual-space Gaussian pseudopotentials
,”
Phys. Rev. B
54
,
1703
1710
(
1996
).
49.
J.
VandeVondele
and
J.
Hutter
, “
Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phases
,”
J. Chem. Phys.
127
,
114105
(
2007
).
50.
A. P.
Thompson
,
H. M.
Aktulga
,
R.
Berger
,
D. S.
Bolintineanu
,
W. M.
Brown
,
P. S.
Crozier
,
P. J.
In’t Veld
,
A.
Kohlmeyer
,
S. G.
Moore
,
T. D.
Nguyen
et al, “
LAMMPS - A flexible simulation tool for particle-based materials modeling at the atomic, meso, and continuum scales
,”
Comput. Phys. Commun.
271
,
108171
(
2022
).
51.
N. C.
Castillo
,
A.
Heel
,
T.
Graule
, and
C.
Pulgarin
, “
Flame-assisted synthesis of nanoscale, amorphous and crystalline, spherical BiVO4 with visible-light photocatalytic activity
,”
Appl. Catal., B
95
,
335
347
(
2010
).
52.
S.
Obregón
and
G.
Colón
, “
On the different photocatalytic performance of BiVO4 catalysts for methylene blue and rhodamine b degradation
,”
J. Mol. Catal. A: Chem.
376
,
40
47
(
2013
).
53.
S.
Su
,
I.
Siretanu
,
D.
van den Ende
,
B.
Mei
,
G.
Mul
, and
F.
Mugele
, “
Nanometer-resolved operando photo-response of faceted BiVO4 semiconductor nanoparticles
,”
J. Am. Chem. Soc.
146
,
2248
2256
(
2024
).
54.
J.
Cheng
,
X.
Liu
,
J.
VandeVondele
, and
M.
Sprik
, “
Reductive hydrogenation of the aqueous rutile TiO2 (110) surface
,”
Electrochim. Acta
179
,
658
667
(
2015
).
55.
M.
Kosmulski
, “
The significance of the difference in the point of zero charge between rutile and anatase
,”
Adv. Colloid Interface Sci.
99
,
255
264
(
2002
).
56.
J. P.
Fitts
,
M. L.
Machesky
,
D. J.
Wesolowski
,
X.
Shang
,
J. D.
Kubicki
,
G. W.
Flynn
,
T. F.
Heinz
, and
K. B.
Eisenthal
, “
Second-harmonic generation and theoretical studies of protonation at the water/α-TiO2 (110) interface
,”
Chem. Phys. Lett.
411
,
399
403
(
2005
).
57.
J. W.
Bullard
and
M. J.
Cima
, “
Orientation dependence of the isoelectric point of TiO2 (rutile) surfaces
,”
Langmuir
22
,
10264
10271
(
2006
).
58.
M.
Jia
,
C.
Zhang
,
S. J.
Cox
,
M.
Sprik
, and
J.
Cheng
, “
Computing surface acidity constants of proton hopping groups from density functional theory-based molecular dynamics: Application to the SnO2(110)/H2O interface
,”
J. Chem. Theory Comput.
16
,
6520
6527
(
2020
).
59.
S. M.
Ahmed
and
D.
Maksimov
, “
Studies of the double layer on cassiterite and rutile
,”
J. Colloid Interface Sci.
29
,
97
104
(
1969
).
60.
J.
Rosenqvist
,
M. L.
Machesky
,
L.
Vlcek
,
P. T.
Cummings
, and
D. J.
Wesolowski
, “
Charging properties of cassiterite (α-SnO2) surfaces in NaCl and RbCl ionic media
,”
Langmuir
25
,
10852
10862
(
2009
).
61.
N.
Bogdanova
,
A.
Klebanov
,
L.
Ermakova
,
M.
Sidorova
, and
D.
Aleksandrov
, “
Adsorption of ions on the surface of tin dioxide and its electrokinetic characteristics in 1:1 electrolyte solutions
,”
Colloid J.
66
,
409
417
(
2004
).
62.
M.
Kosmulski
, “
The ph dependent surface charging and points of zero charge. viii. update
,”
Adv. Colloid Interface Sci.
275
,
102064
(
2020
).
63.
K.
Bourikas
,
J.
Vakros
,
C.
Kordulis
, and
A.
Lycourghiotis
, “
Potentiometric mass titrations: Experimental and theoretical establishment of a new technique for determining the point of zero charge (PZC) of metal (Hydr)Oxides
,”
J. Phys. Chem. B
107
,
9441
9451
(
2003
).
64.
M.
Sprik
, “
Computation of the pk of liquid water using coordination constraints
,”
Chem. Phys.
258
,
139
150
(
2000
).
65.
M.
Sulpizi
and
M.
Sprik
, “
Acidity constants from vertical energy gaps: Density functional theory based molecular dynamics implementation
,”
Phys. Chem. Chem. Phys.
10
,
5238
5249
(
2008
).
66.
F.
Ambrosio
,
J.
Wiktor
, and
A.
Pasquarello
, “
ph-dependent surface chemistry from first principles: Application to the BiVO4(010)–water interface
,”
ACS Appl. Mater. Interfaces
10
,
10011
10021
(
2018
).
67.
J.
Cheng
,
M.
Sulpizi
, and
M.
Sprik
, “
Redox potentials and pKa for benzoquinone from density functional theory based molecular dynamics
,”
J. Chem. Phys.
131
,
154504
(
2009
).
68.
J.
Cheng
,
X.
Liu
,
J.
VandeVondele
,
M.
Sulpizi
, and
M.
Sprik
, “
Redox potentials and acidity constants from density functional theory based molecular dynamics
,”
Acc. Chem. Res.
47
,
3522
3529
(
2014
).
69.
I. D.
Brown
, “
Bond valence theory
,” in
Bond Valences
, edited by
I. D.
Brown
and
K. R.
Poeppelmeier
(
Springer Berlin Heidelberg
,
Berlin, Heidelberg
,
2014
), pp.
11
58
.
70.
F.
Maleki
,
G.
Di Liberto
, and
G.
Pacchioni
, “
pH- and facet-dependent surface chemistry of TiO2 in aqueous environment from first principles
,”
ACS Appl. Mater. Interfaces
15
,
11216
11224
(
2023
).
71.
W.
Matsumoto
,
T.
Fukushima
,
S.
Heguri
,
S.
Fujii
, and
S.
Higashimoto
, “
Boosting charge transport in the BiVO4 photoanode interface modified with an aluminum hydroxide layer for solar water oxidation
,”
Sustainable Energy Fuels
8
,
1626
1635
(
2024
).
72.
F.
Ambrosio
,
J.
Wiktor
, and
A.
Pasquarello
, “
pH-dependent catalytic reaction pathway for water splitting at the BiVO4–water interface from the band alignment
,”
ACS Energy Lett.
3
,
829
834
(
2018
).
73.
S.
Lyu
,
J.
Wiktor
, and
A.
Pasquarello
, “
Oxygen evolution at the BiVO4–water interface: Mechanism of the water dehydrogenation reaction
,”
ACS Catal.
12
,
11734
11742
(
2022
).
74.
M. T. M.
Koper
, “
Theory of multiple proton–electron transfer reactions and its implications for electrocatalysis
,”
Chem. Sci.
4
,
2710
2723
(
2013
).
75.
M.
Shen
,
A. J.
Kaufman
,
J.
Huang
,
C.
Price
, and
S. W.
Boettcher
, “
Nanoscale measurements of charge transfer at cocatalyst/semiconductor interfaces in BiVO4 particle photocatalysts
,”
Nano Lett.
22
,
9493
9499
(
2022
).
76.
E. H. G.
Backus
,
S.
Hosseinpour
,
C.
Ramanan
,
S.
Sun
,
S. J.
Schlegel
,
M.
Zelenka
,
X. i.
Jia
,
M.
Gebhard
,
A.
Devi
,
H. I.
Wang
, and
M.
Bonn
, “
Ultrafast surface-specific spectroscopy of water at a photoexcited TiO2 model water-splitting photocatalyst
,”
Angew. Chem., Int. Ed.
63
,
e202312123
(
2024
).
77.
N.
Österbacka
,
H.
Ouhbi
,
F.
Ambrosio
, and
J.
Wiktor
, “
Spontaneous oxygen vacancy ionization enhances water oxidation on BiVO4
,”
ACS Energy Lett.
9
,
153
158
(
2023
).
78.
Q.
Zhang
,
M.
Liu
,
W.
Zhou
,
Y.
Zhang
,
W.
Hao
,
Y.
Kuang
,
H.
Liu
,
D.
Wang
,
L.
Liu
, and
J.
Ye
, “
A novel Cl modification approach to develop highly efficient photocatalytic oxygen evolution over BiVO4 with AQE of 34.6%
,”
Nano Energy
81
,
105651
(
2021
).
79.
Z.
Li
,
Q.
Zhang
,
X.
Chen
,
F.
Yang
,
D.
Wang
,
L.
Liu
, and
J.
Ye
, “
Cl modification for effective promotion of photoelectrochemical water oxidation over BiVO4
,”
Chem. Commun.
56
,
13153
13156
(
2020
).
80.
X.-Y.
Li
,
X.-F.
Jin
,
X.-H.
Yang
,
X.
Wang
,
J.-B.
Le
, and
J.
Cheng
, “
Molecular understanding of the Helmholtz capacitance difference between Cu(100) and graphene electrodes
,”
J. Chem. Phys.
158
,
084701
(
2023
).
81.
J. G.
Brandenburg
,
A.
Zen
,
M.
Fitzner
,
B.
Ramberger
,
G.
Kresse
,
T.
Tsatsoulĩs
,
A.
Grüneis
,
A.
Michaelides
, and
D.
Alfè
, “
Physisorption of water on graphene: Subchemical accuracy from many-body electronic structure methods
,”
J. Phys. Chem. Lett.
10
,
358
368
(
2019
).
82.
Z.
Wei
,
J. D.
Elliott
,
A. A.
Papaderakis
,
R. A.
Dryfe
, and
P.
Carbone
, “
Relation between double layer structure, capacitance, and surface tension in electrowetting of graphene and aqueous electrolytes
,”
J. Am. Chem. Soc.
146
,
760
772
(
2024
).