Fei-Teng
Wang‡
a,
Jia-Xin
Zhu‡
a,
Chang
Liu
a,
Ke
Xiong
a,
Xiandong
Liu
*de and
Jun
Cheng
*abc
aState Key Laboratory of Physical Chemistry of Solid Surfaces, iChEM, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361005, China. E-mail: chengjun@xmu.edu.cn
bLaboratory of AI for Electrochemistry (AI4EC), IKKEM, Xiamen 361005, China
cInstitute of Artificial Intelligence, Xiamen University, Xiamen 361005, China
dState Key Laboratory for Mineral Deposits Research, School of Earth Sciences and Engineering, Nanjing University, Nanjing, Jiangsu 210023, P. R. China. E-mail: xiandongliu@nju.edu.cn
eFrontiers Science Center for Critical Earth Material Cycling, Nanjing University, Nanjing, Jiangsu 210023, P. R. China
First published on 23rd December 2024
The altered solvation structures and dynamical properties of water molecules at the metal/water interfaces will affect the elementary step of an electrochemical process. Simulating the interfacial structure and dynamics with a realistic representation will provide us with a solid foundation to make a connection with experimental studies. To surmount the accuracy-efficiency tradeoff and provide dynamical insights, we use state-of-the-art machine learning molecular dynamics (MLMD) to study the water exchange dynamics, which are fundamental to adsorption/desorption and electrochemical reaction steps. We reproduce interfacial structures consistent with ab initio molecular dynamics (AIMD) results and obtain diffusion and reorientation dynamics in agreement with the experiment. We show that the hydrogen bonds at the interface become stronger than those in bulk water, which makes the diffusion, reorientation, and hydrogen-bond dynamics slower. Our findings reveal that the spatial correlation of desorption events, driven by the breaking and making of hydrogen bonds, accelerates water exchange dynamics. These dynamics occur on timescales of several hundred picoseconds at 337 K and 302 K. We take a solid step forward toward studying the in situ interface water dynamics and attribute the fast water exchange dynamics to the spatial correlation of the desorption events.
Of the many simulations of the interfacial dynamics (i.e. electron transfer,14 proton-coupled electron transfer15,16), water exchange dynamics are of fundamental interest as this process constitutes the thermodynamic and kinetic basis of many elementary steps. Limmer and Willard et al.17–19 studied the water exchange dynamics at the Pt/water interfaces using classical molecular dynamics and suggested the timescale to be over several nanoseconds at 298 K. Nevertheless, the high water coverage (approximately 0.85 monolayer) and oxygen density profiles at the Pt/water interfaces are distinct from the results of ab initio molecular dynamics (AIMD)20 simulations. For AIMD simulations, observations of water exchange at the interfaces are reported21,22 within 40 picoseconds. Nevertheless, the samplings are deemed insufficient for capturing the full picture of the dynamics of water exchange. Aside from the efficiency issue, the choice of functionals23–25 and thermostats26–28 is also reported to affect the structural dynamics, making the comparison of different theoretical efforts and the connection to experimental studies less clear.
To address the sampling issues we conducted machine learning molecular dynamics (MLMD) simulations and compared the oxygen density distributions from MLMD with those from previous AIMD simulations.29 We also made a benchmark comparison of the pure bulk water dynamics between MLMD and experimental observations, which reveals that by increasing the MD temperature by approximately 50 ± 10 K, the PBE-D3 predictions regarding the water dynamics closely align with the experimental results of the diffusion dynamics30 and reorientation dynamics.31,32 Based on the above theoretical calibration of the simulation temperature, we determine the residence timescale to be a few hundred picoseconds for water at the Pt/water interfaces at 337 and 302 K, which is much faster than classical MD results.19 This heightened frequency of water exchange underscores the critical importance of explicitly considering interface dynamics, especially in scenarios where water functions as a reacting species. For example, the probability of the first peak of adsorbed water dissociation may be suppressed if the water molecules are exchanged to the bulk region, which is closely related to the interfacial neutral pH. We reveal that the acceleration of water exchange is driven by the correlation between initial and subsequent desorption events, and this correlation is key to the observed acceleration. This correlation primarily operates within the first solvation shell of water molecules and during the first few picoseconds following the initial desorption event. This research significantly advances our comprehension of the dynamics of water exchange at metal/water interfaces.
All the density functional theory (DFT) calculations are performed in CP2K/Quickstep.33 The Goedecker–Teter–Hutter (GTH) pseudopotentials34,35 are used to describe the atoms: H is described by GTH-PBE-q1, with all electrons in the valence, the O atom is described by GTH-PBE-q6 with 2s and 2p electrons in the valence, and Pt is described by Pt GTH-PBE-q10 with 5d and 6s electrons in the valence. The DZVP-MOLOPT-SR-GTH Gaussian basis36 set is used for all atom types except for Pt, which uses a developed basis set for the GTH-PBE-q10 of Pt by Le et al.20 The cutoff of the plane wave energy is set to 1000 Ry. The Perdew–Burke–Ernzerhof (PBE) functional is utilized to describe the exchange-correlation effect.37 Grimme D3 correction is included to consider the dispersion interaction.38
The molecular dynamics (MD) simulations are performed with both the Large-scale Atomic/Molecular Massively Parallel Simulator (LAMMPS) package41 and CP2K package.33 For the exploration in the DPGEN, we use LAMMPS to conduct MD. The simulation is conducted in the NVT ensemble and we set the temperature to 330/430/530 K to include the configurations distributed around these temperatures. The time step is set to 0.5 fs and the Nose–Hoover thermostat is used to control temperature with the temperature damping parameter setting to 100 fs.42,43 For the equilibration and production run, we used the CP2K package. For the convenience of comparison to previous ab initio molecular dynamics (AIMD) simulations,29,44 we performed second-generation Car–Parrinello MD in a canonical ensemble (NVT) using a time step of 0.5 fs. The Langevin friction coefficient γD is set to 0.001 fs−1. The intrinsic friction coefficients γL are set to 2.2 × 10−4 fs−1 for H2O and 5 × 10−5 fs−1 for Pt based on preliminary tests.44 The validation runs involve performing a 1 ns simulation using the MLP with K point correction included. Then, we extracted 200 structures and calculated their energy and force using the first-principles calculations (2 × 2 × 1 k point setup). We compared the RMSE (root mean square error) of the energy and force to those determined by the MLP. The results are shown in ESI Table S1† (columns 5 and 7). To determine whether the thermostat affects the intrinsic dynamics, we conducted a subsequent 10 ns MD simulation in the NVE ensemble after 10 ns simulations in the NVT ensemble.
![]() | (1) |
The three-dimensional diffusion coefficient is obtained based on the mean square displacement (MSD):
![]() | (2) |
We used the three-dimensional formula to determine the diffusion coefficients for pure bulk water molecules. The finite size correction to the diffusion coefficients was estimated using the expression proposed by Dünweg and Kremer:45,46. Here kB is the Boltzmann constant, T is the temperature, ζ equals 2.83729, and η is the viscosity of liquid water whose value was taken from experimental data. L is the size of the cubic simulation cell.
The hydrogen-bond dynamic correlation function is defined as follows:
![]() | (3) |
Crot(t) = 〈Pl(μ(t)a(0)μ(0))〉 | (4) |
![]() | (5) |
![]() | (6) |
The vertical distances of the peaks and valleys of the oxygen density profile are used to distinguish water molecules in different regions. To study the residence dynamics and avoid the inclusion of water molecules that are at the boundary of different adsorption peaks, we use the middle values between the peaks and valleys. For example, the first chemisorbed peak is at around 2.3 Å and the first valley is at around 2.65 Å. So, we use 2.45 Å as the plane to choose those water molecules in the chemisorbed state in the A region. Based on these distinct regions, we conduct the residence correlation function analysis.
![]() | ||
Fig. 2 (A) Benchmark comparison of water bisector reorientation dynamics and diffusion dynamics from simulations and experimental observations. (B) The oxygen diffusion along the surface from its original position within 20 ps for the Pt(100)/water interface. (C) Mean squared displacements (MSD) of the oxygen atom of water L parallel to the Pt surfaces. For the NVE ensemble, the subscript 1/2/3 stands for the three temperatures. (D) The diffusion coefficients for water B, C, and L of Pt(100) and Pt(111). The data of expt.diff is reproduced with permission from ref. 30. The data of expt.Rs is reproduced with permission from ref. 47. |
For the interface models, we use the MSD method to calculate the diffusion coefficients and track the displacement every 10 ps over a 5 ns trajectory (Fig. 2C). Due to the isotropic diffusion along the surfaces on Pt(111)/Pt(100) surfaces that are different from the stepped Pt/water interface as we recently observed,52 we choose to show the one-dimensional diffusion. Here, we did not include the finite size correction as the correction is relatively small and the comparison between different types of water molecules makes the correction less important. The MSDs of water molecules L obtained from simulations in NVE ensembles for Pt(111) and Pt(100) are presented in Fig. 2C (see Fig. S17† for NVT results). We obtain the diffusion coefficients from the slope of the MSD over time. Here, we did not include the finite size correction as the target is to compare the relative trend of the diffusion dynamics. Fig. 2D shows the diffusion coefficients of water molecules in B, C, and L regions. Compared to the diffusion coefficients of pure bulk water molecules (Fig. 2A) at the respective temperature, the 1D diffusion coefficients of water L along the xy plane parallel to the Pt surface are faster by about 0.2/0.6/0.9 × 10−5 cm2 s−1 at 270/302/337 K, which is also observed in a previous study53 and we attribute this difference to the confinement of the interfacial water molecules. Moving from region L to C and B (Fig. 2D), we observe a decrease in diffusion coefficients, with water B exhibiting nearly half the diffusion coefficient of water L. Taking the deviation into consideration, the diffusion coefficient difference between the Pt(111) and Pt(100) interfaces is mild. This trend aligns with observations on copper single crystal surfaces.53 The distributions of diffusion coefficients are further detailed in Fig. S9.†
![]() | ||
Fig. 3 (A) Hydrogen bond correlation function. CHB(t) represents the hydrogen bond correlation function for water molecules in specific regions. The inset figure defines hydrogen bonds using the distance between two oxygen atoms and the angle between HO and OO′. The distance is set to 3.2 Å and the angle to 30°. (B–D) The fitted lifetime of the slow hydrogen bond dynamics of water L, AB, and C. See NVT results in Fig. S18.† (E) The joint probability distribution of finding an O–H–O′ geometry with a given ν–θ configuration for water molecules. Here, ν is the difference between the distance of OH and HO′. θ is defined as the angle between OH and OO′. Arrows are used to point to the stable/metastable geometry for a, b, and c. For the scale bar, P denotes the frequency count, and Pmax represents the maximum frequency count of a given O–H–O′ geometry. P/Pmax denotes the normalized probability distribution. (F) The difference between the normalized joint probability of water AB and L at 387 K. Arrows are used to indicate the main geometry difference at the interface. (G) The radial distribution function (RDF) for O–O of interfacial water molecules at 337 K and water L at 270/302/337 K. The experimental data is reproduced with permission from ref. 58. |
To further investigate the trend of longer lifetimes observed for interfacial water molecules, we analyzed the joint probability distribution of the O–H–O′ geometry. As defined in the caption of Fig. 3, a more positive value of ν in the O–H–O′ geometry suggests a stronger hydrogen bond. Fig. 3E illustrates three geometry domains (a), (b), and (c), indicated by arrows. In domain (a), cosθ is nearly 1.0 and ν is close to −1.0 Å (Fig. S12A†), suggesting a strong hydrogen bond. In domain (b), cos
θ is close to 0.60 and ν is close to −3.60 Å. This geometry does not correspond to a stable hydrogen bond between the observed donor and acceptor water molecules (Fig. S12B†). However, these donor water molecules may form hydrogen bonds with other surrounding water molecules. For geometry domain (c), the values of cos
θ and ν are approximately −0.40 and −2.20 Å, respectively.56 This hydrogen bond geometry involves donor water molecules (O) that accept hydrogen from defined acceptor water molecules (O′), but do not donate hydrogen to them (Fig. S12C†). An analysis of the disparity in joint probability between water molecules in the bulk and interface regions (Fig. 3F) reveals that hydrogen bonds in the interface (green) exhibit a more positive ν compared to those in the bulk region (purple). This observation is supported by the radial distribution function (RDF) of O–O, where the hydrogen bonds in the AB region are approximately 0.1 Å shorter than those in the L region (Fig. 3G). These findings highlight the distinct characteristics of interfacial water molecules: stronger hydrogen bonds with longer lifetimes, and lower diffusion coefficients. It should be further noted that PBE-D3 calculations predict a shorter O–O RDF peak compared to experimental observations.57 By raising the simulated temperature by 50 K, a closer alignment of the peak position and height in the simulation results with the experimental data is achieved.58
The reorientation of water molecules is intricately linked to the dynamic rearrangement and restructuring of the hydrogen bond (HB) network.59 Studies have demonstrated that rotational anisotropy is directly related to the second-order rotational autocorrelation function of the transition dipole moment of excited molecules.31 To assess the reorientation dynamics at the interface, we analyzed the second Legendre polynomials of the correlation function associated with the water bisector, which closely aligns with the direction of the dipole moment.32 The same trajectory employed for the analysis of hydrogen bond dynamics was utilized to study the reorientation dynamics. Subsequently, a two-exponential form was employed to fit the correlation functions for water A, B, C, and L for Pt(100)/water and Pt(111)/water interfaces. Fig. 4A illustrates the slow lifetime of pure bulk water (blue) and water L for the reorientation dynamics as determined by MLMD (see Fig. S13† for a fast lifetime). Compared to pure bulk water (Fig. 4A, green, also shown in Fig. 2A, orange), the reorientation dynamics of water L exhibit a slower rate at 270 K. The lifetime of water L is approximately 10–15 ps, about 5 ps slower than that of bulk water at the Pt(111)/water and Pt(100)/water interfaces. The deviation between NVE and NVT results is approximately 5 ps (see Fig. S19† for NVT results). As with hydrogen bond dynamics, this difference is likely attributable to water exchange dynamics and the time interval used for reorientation correlation function analysis. At the other two temperatures, the lifetime difference between water L and pure bulk water is minor, a trend also observed for hydrogen bond dynamics and diffusion dynamics. This minor difference may be attributed to the time interval used for correlation function analysis (Table S3,† 60 ps at 302 K and 20 ps at 337 K), which is short enough to be mildly affected by water exchange dynamics. Similar to hydrogen bond lifetime, the reorientation lifetime progressively increases from water L to B and A (Fig. 4B–D). Conversely, water C exhibits a slightly faster reorientation rate compared to other regions. The results presented herein contrast with those reported by Limmer et al.,17 who suggest that the reorientation dynamics timescale at the interface is significantly slower than the characteristic relaxation time for bulk water.
![]() | ||
Fig. 4 The fitted lifetime for water bisector reorientation dynamics. Here only the slow lifetimes are shown. See Fig. S13† for the fast lifetimes. (A) Water L at Pt(100)/water and Pt(111)/water interfaces and pure bulk water molecules (green). (B) Water A at the interfaces. (C) Water B at the interfaces. (D) Water C at the interfaces. |
To study the residence dynamics and avoid the inclusion of water molecules that are at the boundary of different adsorption peaks, we use the middle vertical distance between the peaks and valleys of the oxygen density profiles. Specifically, we designate the region 1.8 to 2.45 Å from the Pt surface as the stable adsorbed region for water A, and 2.9 to 3.75 Å for water B. The collective set of adsorbed water molecules is denoted as AB. Additionally, we define the non-adsorbed state as the region spanning 5.2 to 12.75 Å away from the Pt surface, which allows us to examine water exchange dynamics with the adsorbed state. This strategy is also used by Natarajan and Behler.53 The aforementioned classification permits the investigation of both the residence time of all adsorbed water molecules and the residence time of individual water molecules within the adsorbed layer (Fig. 5A). This distribution encompasses all residence times of individual water molecules. Here, we denote the residence time of all adsorbed water molecules as R and the residence time of a single individual water molecule as RS (Table 1).
Ensemble | T (K) | τ R (ps) | τ RS (ps) | ||
---|---|---|---|---|---|
AB | A | AB | |||
Pt(100) | NVT | 302 | 208(19) | 50(17) | 74 |
337 | 90(8) | 18(8) | 42 | ||
NVE | 302 | 224(16) | 44(12) | 79 | |
337 | 99(7) | 20(4) | 45 | ||
Pt(111) | NVT | 302 | 240(20) | 33(7) | 112 |
337 | 107(10) | 13(3) | 55 | ||
NVE | 302 | 251(21) | 34(7) | 128 | |
337 | 106(13) | 13(3) | 60 |
As shown in Table 1, the residence time for water A (τR) at 302/337 K is only on the order of tens of picoseconds for both interfaces, indicating that water molecules in the A region exhibit a high degree of flexibility and readily depart from the adsorbed layer. The timescales for water AB at 302/337 K are significantly faster than the residence times of adsorbed water molecules at Pt(100)/water and Pt(111)/water interfaces obtained from classical MD simulations60 at 298 K (ranging from several to tens of nanoseconds)17,19 with high water coverage (larger than 85%). To facilitate comparison with the previous MLMD study, we also fitted the lifetime of water AB using two exponential forms. At the Pt(111)/water interface, the slow residence time at 302 K (352 K without adjustment) in this work is approximately 20% (100 ps) faster than that obtained from previous MLMD simulations at 350 K for the same interface,61 which may be related to the model difference (the inclusion/exclusion of the liquid/vacuum interface, the cell size). It is difficult to definitively exclude the possibility of both timescales fitted using different exponential forms based solely on residence dynamics analysis. However, further analysis of the water flux correlation function, presented in the next section, indicates that exchange processes associated with the slower timescale extracted from two exponential forms cease much earlier. Therefore, we favor the timescale obtained from τR. If individual water molecules resided in the adsorbed layer independently, the overall residence time would be significantly longer than the residence time of individual water molecules. However, the observed ratio between the residence time of individual water molecules and the residence time (τR) of water AB is only 2–3. This suggests that when an individual water molecule departs from the adsorbed layer, its departure influences the residence times of other water molecules, effectively reducing the overall residence time of the adsorbed water molecules.
To investigate the spatial correlation of exchange events, we examined the RDF of adsorbed water molecules. The reference water molecule is the first to desorb, while the observed water molecules are those desorbing shortly after the initial event. We scrutinized the final snapshot just before the first desorption transpires, testing periods of 5/10 ps to ensure consistency (Fig. S14†). The results remained unaffected by the choice of period. Fig. 5B demonstrates that the RDF exhibits a peak at a pair distance of approximately 2.7 Å, corresponding predominantly to the first solvation shell of water molecules within the adsorbed layer. When an adsorbed water molecule breaks free from the solvation shell at the adsorbed layer, the original hydrogen bond network is broken. Consequently, the surrounding adsorbed water molecules are afforded a heightened likelihood of exchanging with non-adsorbed water. This spatial correlation gradually diminishes beyond 2.7 Å and becomes less pronounced at distances exceeding 10 Å.
To investigate the temporal correlation during the exchange process, we employed the flux correlation function as outlined in eqn (6) in the method section. The rapid exchange between water A and B (Fig. S15†) aligns with their respective residence times. The correlation results for the exchange between water AB and non-adsorbed water molecules are illustrated in Fig. 5C. Fig. 5D displays the slope of this flux correlation function over time, reflecting the rate of water exchange. When the exchange rate decays to zero, all adsorbed water molecules no longer reside in the adsorbed layer. Notably, the exchange rate demonstrates an initial increase within the first 5–10 ps, suggesting the efficacy of correlation at this timescale.
To provide a concise overview of the water exchange process, we propose a three-stage model based on the desorption of all initially adsorbed water molecules as shown in Fig. 5E.
Pre-desorption stage: prior to the initiation of the exchange process (Fig. S16B†), the hydrogen bond network evolves, and local adsorbed water molecules lose one or two hydrogen bonds (Fig. S16C†). This loss of hydrogen bonds results in faster instantaneous water diffusion (Fig. S16D†).
Acceleration stage: following the initiation of the first desorption event, the local hydrogen-bond environment is distorted, significantly increasing the likelihood of desorption for adjacent adsorbed water molecules. This distortion triggers a chain reaction of successive desorption events, driven by both spatial and temporal correlations. It is important to note that the relatively low coverage of A (0.15 ML) and AB (0.58/0.71 ML for Pt(111) and Pt(100)) means that space for incoming water molecules to adsorb is not a limiting factor, in contrast to classical MD simulations.19
Post-desorption stage: approximately 10 ps into this stage, a significant fraction of the initially adsorbed water molecules have transitioned to the bulk region. The remaining adsorbed water molecules are isolated within their new local hydrogen bond environments. This isolation results in a diminishing of spatial and temporal correlations, leading to a gradual decline in the exchange rate until it eventually halts.
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4sc06967f |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |