Philipp
Schienbein
* and
Jochen
Blumberger
*
Department of Physics and Astronomy and Thomas Young Centre, University College London, London, WC1E 6BT, UK. E-mail: p.schienbein@ucl.ac.uk; j.blumberger@ucl.ac.uk
First published on 10th June 2022
Metal oxide/water interfaces play an important role in biology, catalysis, energy storage and photocatalytic water splitting. The atomistic structure at these interfaces is often difficult to characterize by experimental techniques, whilst results from ab initio molecular dynamics simulations tend to be uncertain due to the limited length and time scales accessible. In this work, we train a committee neural network potential to simulate the hematite/water interface at the hybrid DFT level of theory to reach the nanosecond timescale and systems containing more than 3000 atoms. The NNP enables us to converge dynamical properties, not possible with brute-force ab initio molecular dynamics. Our simulations uncover a rich solvation dynamics at the hematite/water interface spanning three different time scales: picosecond H-bond dynamics between surface hydroxyls and the first water layer, in-plane/out-of-plane tilt motion of surface hydroxyls on the 10 ps time scale, and diffusion of water molecules from the oxide surface characterized by a mean residence lifetime of about 60 ps. Calculation of vibrational spectra confirm that H-bonds between surface hydroxyls and first layer water molecules are stronger than H-bonds in bulk water. Our study showcases how state of the art machine learning approaches can routinely be utilized to explore the structural dynamics at transition metal oxide interfaces with complex electronic structure. It foreshadows that c-NNPs are a promising tool to tackle the sampling problem in ab initio electrochemistry with explicit solvent molecules.
Force field MD simulations have often been conducted on oxide/water interfaces, see e.g. ref. 17. However, metal oxide interfaces are generally reactive, be it O atom exchange16 or frequent proton transfer between the metal oxide and water,18 and some materials feature a complex electronic structure.19 Such effects are generally not captured by standard non-polarizable force fields. It is therefore desirable to employ ab initio MD (AIMD, also called “DFTMD”) simulations,20 where the full electronic structure is explicitly considered using DFT. Such simulations have already been done for numerous systems,21–33 where mostly hybrid DFT functionals have been used. A large disadvantage of hybrid functionals is, that they are generally much more expensive compared to GGA functionals. MD simulations based on hybrid DFT are therefore computationally demanding and also limited to short times (about 10 picoseconds) and relatively small system sizes. Sampling dynamical properties, e.g. rate constants or residence lifetimes, require that the event of interest happens frequently along the MD trajectory. However, if the desired process is slower than the trajectory length (so-called “rare events”), the corresponding time correlation functions cannot reliably be converged. Notably, dynamical properties can also suffer from finite-size effects due to the limited size of the simulation box, which are, e.g. well known for the self-diffusion coefficient in a periodic simulation box.34–36
In the last years, Machine Learning (ML) models have been introduced to run MD simulations with the aim to replace the expensive DFT calculations by a computationally less expensive model while keeping the error reasonably small. A widely applied ML approach are Behler–Parrinello Neural Network potentials (NNPs),37 where atomic energies are trained for each atom species. The total energy of a given configuration is then obtained as the sum of all atomic energies. Here, so-called symmetry functions are used to transform the Cartesian coordinates into a representation based on intermolecular distances and angles only. Such NNPs have successfully been applied to a number of systems, including bulk liquid water,38 metal/water interfaces, e.g. Cu,39 and also metal oxide/water interfaces, e.g. ZnO18 and LixMn2O4.40 Very recently, so-called committee Neural Network potentials (c-NNP) have been introduced41 which are based on Behler–Parrinello NNPs. In a nutshell, a committee (or ensemble) of NNPs yields an average energy and a corresponding variance. The variance can be interpreted as a measure how “good” the energy prediction of a given configuration is and thereby allows active learning and automated fitting approaches which dramatically reduce the number of expensive training data needed compared to standard NNP training approaches.
In this work, we have trained a c-NNP for the hematite/liquid water interface at the hybrid DFT level of theory to uncover the structural dynamics of water at the interface over multiple time scales, from pico to nanoseconds. Hematite (Fe2O3, “rust”) is one of the most abundant metal oxide compounds on earth.42 In the field of photocatalysis, the potential use of hematite for water splitting2,7 as a source of “green” hydrogen motivated numerous studies including nanostructuring43 and doping.44 Despite all the efforts, hematite still suffers from slow reaction kinetics and a large overpotential which hinders its usage in real photoelectrochemical cells.2,45 To address these practical problems, a deep understanding of the atomistic processes (structural and dynamical) at the hematite/liquid water interface is key. Hematite is known to feature a complex electronic structure due to the antiferromagnetic spin pattern19 and the interface is a highly reactive region,16 where charge transfer and polarization effects are expected to be of major importance. We opt to describe the hematite/water system with a modified hybrid HSE06 functional, where the fraction of exact exchange is reduced from 25 to 12%. This functional was shown to give excellent agreement with experimental measurements for a wide range of different properties including crystal structure, band gap, antiferromagnetic ordering and spin density distribution.26,46 The trained c-NNP model allows us to significantly extend the simulation time (up to nanoseconds) and the size of the simulation box (up to about 3000 atoms). We show that such time and length scales are required to converge important dynamical properties such as mean residence lifetimes and diffusion coefficients of surface water molecules. This is clearly out of scope for explicit AIMD simulations, especially when using hybrid DFT, without the aid of ML. Note that even longer simulation times and larger boxes are generally possible using ML, but do not add any significant additional value to the physical properties discussed in this study. Using the c-NNP simulations, the aim of this paper is to explore the dynamics of water molecules at the hematite(001) surface. In particular, we are interested in (i) rate constants for exchange of water molecules between the surface and the bulk liquid, (ii) the intramolecular vibrational signature of surface water molecules compared to bulk water, (iii) self-diffusion of water at the interface, and finally, (iv) which time and length scales are required to converge such properties.
Here, we train our c-NNP based on a two-step procedure: In our previous work,26 about 15 ps of AIMD simulation data have been generated in total. From these available previous data, we randomly selected 20 configurations and trained a c-NNP. Thereafter, this c-NNP is used to predict the energies of all configurations of the same 15 ps AIMD trajectory. The 20 “most unknown” configurations are automatically selected, added into the training set, and a new c-NNP is trained. This cycle was repeated 10 times, such that 200 configurations are contained in our training set in total. In the second step, we further use an active learning procedure41,51 to systematically explore the phase space further and iteratively improve the quality of the c-NNP. In that protocol, we run a MD simulation of the hematite/water interface using the above trained preliminary c-NNP. While doing so, it is possible that new areas in the phase space are visited, which have not been included in the training set of the c-NNP yet. This manifests in an increase of the variance because the committee members start to extrapolate beyond the known training set and thus disagree on the energy. Importantly, the c-NNP does not constrain the MD to sample only that portion of potential energy surface which is already known, but allows to explore unknown regions. This is also desired to iteratively include more and more training data to mimic the full accessible DFT potential energy surface at a given temperature as close as possible. In full analogy to step one, we now take the 10 most unknown configurations (the ones with the largest variance) from the MD simulation. The forces for these unknown configurations are then calculated by explicit DFT calculations using the hybrid HSE06 functional52 with 12% HFX46 supplemented by D3 dispersion corrections;53 the details of the electronic structure calculations are described in Section S1 in the ESI.† We then compute the RMSE between the calculated energies/forces from the c-NNP and the explicit DFT calculation. If the RMSE is small enough, the training of the c-NNP is finished. Otherwise, the 10 new DFT energies and forces are added to the training set, a new c-NNP is trained and the whole process is repeated until the desired accuracy is met. Note that by adding only the most unknown configurations to the training set it is also ensured that only such configurations enter which are sufficiently different from the existing training set, i.e. which have not been added before. This way, 200 configurations were added to the training set such that the final training set is composed of only 400 statistically independent configurations.
All c-NNP simulations have been conducted using CP2k54 and the training of the individual Behler–Parrinello NNPs forming the committee has been done using n2p2.55 The final c-NNP used for production was validated against the previous explicit AIMD simulations26 and unknown data which was generated by the c-NNP simulations. Thereby, we ensure that the trained c-NNP mimics the explicit AIMD correctly as elaborately shown in Section S2 in the ESI.† For the final c-NNP we obtain a total force RMSE with respect to the unknown validation set of 149.8 meV Å−1. This RMSE agrees well with previous trainings of NNPs, where similar force RMSEs have been obtained, e.g. for liquid water at a Cu slab (207.5 meV Å−1),39 LixMn2O440 (240.0 meV Å−1) or liquid water at MoS2 slabs51 (190.1 meV Å−1).
It is well known in the literature that PBE functionals tend to overbind water–water interactions56–59 resulting in too rigid structures and slower dynamics than experimentally measured. This also holds true for PBE derived functionals, such as the HSE06 and PBE0 hybrid functionals.60,61 One possible way to counteract this overbinding effect is to increase the simulation temperature until the correct experimental quantities are recovered. For PBE (and derived hybrid functionals) an effective simulation temperature of 400 K has been shown to reproduce the structural dynamics, in terms of radial distribution functions and the self-diffusion coefficient, of bulk liquid water at ambient conditions satisfactorily.56,60–62 Indeed, we elaborately show in Section S4 in the ESI† that HSE06-D3 water at 400 K slightly overestimates the dynamics, but still slightly overbinds the structure and thus offers a very good compromise to describe ambient liquid water at 300 K correctly in terms of both, the structure and the dynamics. For this reason, we opt to run all simulations in this paper at that effective simulation temperature of 400 K to correctly model the structural dynamics of real liquid ambient water at 300 K.
The trained c-NNP makes possible the use of larger system sizes. We employ four different system sizes to explore the water dynamics at the hematite/water interface: the first (and smallest) box is taken from our previous simulations26 containing 93 water molecules and a height (box length perpendicular to the hematite surface) of about 47 Å and a surface area of about 10 × 10 Å2. We then expanded the dimension perpendicular to the slab to 80 Å and increased the number of water molecules to 208 while keeping the water density constant at 0.996 kg L−1. Finally, we increased the surface area of the hematite surface by a factor of two (“(1 × 2)”, 416 water molecules, 1560 atoms) and by a factor of four (“(2 × 2)”, 832 water molecules, 3120 atoms). A more detailed description and illustration of the employed boxes is given in Section S3 and Fig. S2 in the ESI.† For each box, a c-NNP simulation in the canonical (NVT) ensemble is conducted for a total trajectory length of 1 ns. From this NVT simulation we took 10 equidistant configurations as starting points for subsequent simulations in the microcanonical (NVE) ensemble. The NVE simulations were run for 400 ps each using a timestep of 0.5 fs. All the presented results in the following are averaged over 10 NVE simulations of the largest box, (2 × 2), except explicitly stated otherwise. We opted to use only the NVE simulations to obtain the presented results herein, because there are some properties (vibrational density of states, diffusion coefficient) which are calculated from atomic velocities. Within a NVT simulation, these velocities might be affected by the thermostat introducing artificial effects in those properties. Arguably, the influence of most thermostats is generally small, especially if the simulations are well equilibrated. However, if additional NVE simulations are affordable, this issue is guaranteed to be circumvented. Moreover, from a practical point of view, many independent NVE simulations can trivially be parallelized and block averaged.
Noteworthy, the hematite slab is terminated with hydroxide groups, see Fig. S2 in the ESI.† Given the length of the generated trajectories, one might expect that some of the OH groups dissociate during the simulation, but we do not find any of such dissociations. One might wonder if this could be a problem of the employed c-NNP potential instead of conducting explicit AIMD simulations. Generally, NNPs have shown to be reactive, e.g. at the ZnO/water interface,18 where proton transfer from the surface to the adjacent water molecules could be sampled using NNPs. Moreover, our employed active learning scheme described above, allows the MD to freely explore the phase space and does not put any constraint on the potential energy. Therefore it is more likely that dissociation does not occur due to physical reasons. The pKa value describing the deprotonation of these hematite(001) OH groups has been estimated before29 to be 18.5 indicating that the deprotonation is highly improbable; in fact, a terminal deprotonated O− group would even be a quite strong base. This explains why the terminal OH groups do not dissociate along the trajectories, even when going to nano second time scales.
![]() | ||
Fig. 1 Side view of the water layering at the hematite slab (top), in the previously26 used 10 × 10 × 47 Å box, see Section S3 in the ESI† for details, where Fe, O, and H atoms are painted ochre, red, and white, respectively. The blue horizontal lines illustrate the periodic box. Atom densities (bottom) obtained from the c-NNP simulations trained to mimic the hybrid HSE06 functional in this work of H atoms (filled blue curve) and O atoms (filled purple curve). For reference, the previously published atom densities obtained from explicit AIMD simulations of the terminal hydroxyl H atoms (black solid line), of the water H atoms (red solid line) and of the water O atoms (orange solid line) are shown. All atom densities are normalized by their respective value in bulk ambient water. The average location of the interfacial oxide O layer is set to a distance of 0 Å by convention. |
In Fig. 1 we also compare the atom densities as obtained by the previously published explicit AIMD simulations (pico-second timescale) with the ones obtained by the new c-NNP simulations (nano-second timescale). First, we deduce that the trained c-NNP simulations qualitatively recover the very same layering as the explicit AIMD simulations as expected. Quantitatively, we observe that the immediate layering at the slab up to about 3 to 4 Å is described well with the explicit AIMD simulations. At distances further away, the structure becomes much less clear because the density simply is not fully converged statistically. The c-NNP simulations adding about three orders of magnitude more data are fully converged also at large distances. In terms of the structure, explicit AIMD simulations are therefore sufficient to properly describe the immediate surrounding of the slab (in terms of the statistical quality of the results), however, did not provide enough data to converge the structure at larger distances.
The kinetics of water exchange between the first and second or higher water layers can be obtained using the stable-state-picture (SSP) introduced by Laage and Hynes.63 Within this technique, a time correlation function
![]() | (1) |
In Fig. 2C(t) is shown as a function of time. We observe that the time correlation function slowly but systematically converges from a total trajectory length of 200 to 400 ps. Note that the shortest length considered here is already about ten times longer than the original AIMD simulations26 and is thus clearly out of scope for present-day AIMD simulations at hybrid DFT level. The inset shows the resulting residence lifetimes τ determined from the respective correlation functions as a function of the total trajectory length. At 200 ps we find a residence lifetime of about 32.5 ps, whereas at 400 ps we find one of about 49 ps. The residence lifetimes fit remarkably well using a standard exponential “limited growth” function (f(t) = S − (S − A)exp[−kt], where S, A, and k are fitting constants and S marks the upper limit) which yields a residence lifetime of about 58.3 ps for an infinitely long simulation. The error bars are obtained by block averaging over the N water molecules in eqn (1). Here, we have averaged over 485 water molecules being in the first solvation layer at t = 0; recall that the data is collected from 10 independent NVE simulations of 400 ps each (4 ns in total), where each trajectory contained 3120 atoms (largest box investigated). To decrease the statistical error further, even larger simulation boxes or even more trajectories would therefore be required. Given the size of the error bars for the obtained residence lifetime, we estimate the error of our extrapolated result to be about 10 ps.
![]() | ||
Fig. 2 Time correlation functions according to eqn (1) describing the residence lifetime of water molecules in the first solvation layer at the hematite surface as a function of total simulation time, where the brown, light brown, light green, and green curves correspond to total simulation times of 200, 300, 350, and 400 ps, respectively. The inset shows the corresponding residence lifetimes τ which have been determined by fitting an exponential function (see text) to the respective correlation function as a function of total simulation length, using the same color code. The black solid line shows an exponential growth function which fitted to all τ values to extrapolate towards an infinite simulation length. The black horizontal dotted line indicates the limit of that growth function of 59.3 ps corresponding to an infinite total simulation length. |
Having determined the residence lifetime, we now turn to H-bond dynamics at the interface. H-Bond lifetimes have been computed from MD simulations for a long time, however, they are no physical observables and depend on the H-bond definition used. Moreover, lifetimes itself can be defined in many different ways, e.g. “continuous” and “intermittent” lifetimes64 or within the more sophisticated reactive-flux methodology,65 and their interpretation is slightly different. Here, we employ the continuous H-bond lifetime definition64
![]() | (2) |
![]() | ||
Fig. 3 Continuous H-bond lifetime correlation functions64CHB(t), see eqn (2), for H-bonds between two water molecules in the bulk phase (solid blue), between water molecules where at least one is located in the first solvation layer (dashed red), and between water and the hematite surface (green). The inset shows the residence lifetime of a terminal hematite O–H group in the IP or OOP state, respectively, see Fig. 1 which was calculated using the steady-state picture63 and eqn (1). |
Previously we have reported that the exposed terminal OH groups at the hematite surface can interchange between an “in-plane” (IP) and “out-of-plane” (OOP) configuration,26 where the OH bond vector is parallel or perpendicular to the surface plane, respectively (see Fig. 1). As shown in Section S6 in the ESI,† these two states are clearly separated by a transition state. We can therefore again apply the SSP in eqn (1) as employed above for the residence time of water molecules in the first solvation layer. The corresponding time correlation function is shown in the inset of Fig. 3 and yields an average lifetime of 11.26 ± 3.28 ps. The latter quantity describes how long the OH group resides in the IP or OOP configuration until it interchanges and is settled in the respective other state. We find that the correlation function is converged after roughly 300 ps total simulation time. Therefore an extrapolation to an infinite simulation duration, as it was done above for the residence time in the first solvation layer, see Fig. 2, is not necessary. Consequently, the IP/OOP interchange lifetime is significantly shorter compared to the average residence time of water molecules in the first layer lifetime, but substantially longer than the interfacial H-bond lifetimes.
A common (indirect) measure of the H-bond strength in liquid water is the spectral signature of the intramolecular OH stretch vibration.70 If a strong H-bond is formed between two water molecules, the intramolecular OH bond is weakened, causing a red-shift of the frequency and vice versa. Note that the intermolecular H-bond stretch can also directly be measured in bulk water using IR spectroscopy in the THz frequency range, also at elevated temperatures and pressures.68,71 Still, the intramolecular OH stretch has been used more frequently in the past, because it is readily available from the vibrational density of states (VDOS) of the H atoms,24,68,72 while the intermolecular H-bond mode does not appear in the H or O VDOS at all since charge transfer and polarization effects play a crucial role.71 The H VDOS is computed as usual from the Fourier transform of the velocity time auto correlation function of H atoms
![]() | (3) |
In Fig. 4 we show ΛH(ω) considering only H atoms belonging to water molecules in the first layer compared to H atoms belonging to water molecules in the bulk liquid phase. We find that the OH stretch peak of bulk water molecules is centered around 3500 cm−1 (black line). In case of the first layer water molecules we further differentiate between OH vibrations which are perpendicular to the surface and vibrations which are parallel to the surface (blue line). This can trivially be achieved by computing the time correlation functions in eqn (3) for each Cartesian component individually; note that the hematite surface is constructed such that it is parallel to the xy plane. We find that water OH bonds which are oriented perpendicular to the hematite surface are significantly red-shifted compared to those which are oriented mostly parallel to the slab. This observation can be understood in terms of the structure (see Fig. 1), where we showed that water actively donates H-bonds to the hematite slab. In such a case, the water OH bond points towards the surface and the bond vector is consequently also oriented perpendicular to the surface plane. By dissecting the spectrum into parallel and perpendicular contributions, we can thus deduce that H-bonds donated to the hematite slab are stronger compared to H-bonds between water molecules. Recall that the intramolecular OH stretch vibration is an indirect measure of the H-bond strengths.70 A red-shift of that band indicates that the intramolecular vibration is weakened due to a stronger intermolecular H-bond. Remarkably, we cannot find any difference in H-bonding for water/water H-bonds, even if at least one of the two molecules is in the first solvation layer. This implies that the water/water H-bond network at the surface is not changed compared to bulk, but only the hematite/water H-bonds differ from H-bonds in bulk liquid water.
![]() | ||
Fig. 4 Vibrational density of states (VDOS) of H atoms (ΛH(ω), eqn (3)) showing the intramolecular OH stretch of water molecules located in the pure bulk phase (black) and in the first solvation layer at the hematite surface. In the latter case, the VDOS is split into its xy components (blue) and z component (red) which are parallel and perpendicular to the surface, respectively. |
The dynamics of water molecules in a simulation box are commonly quantified in terms of the self-diffusion coefficient. To compute the diffusion coefficient we use the Green–Kubo formulation74
![]() | (4) |
In Fig. 5 we show D‖PBC as a function of the simulation box size. Remarkably, it significantly depends on the size of the simulation box: upon increasing the dimension perpendicular to the surface from 47 to 80 Å the diffusion coefficient is greatly enhanced and after increasing the surface area from 100 to 400 Å2 the diffusion coefficient decreases again. Surprisingly, in case of the 10 × 10 × 80 box, we find a diffusion coefficient of the bulk molecules around 7 which is more than twice as large as the diffusion coefficient calculated for a cubic bulk water box using the very same c-NNP as for the hematite/water interface (dashed horizontal line, see ESI† for details). Simultaneously, the error bars behave the same way. We further observe, that the diffusion coefficient of water molecules being in the first layer is generally smaller compared to the one of water molecules being in the bulk phase, irrespective of the simulation box size.
![]() | ||
Fig. 5 Self-diffusion coefficient D‖PBC computed by eqn (4), but only considering the x and y Cartesian components which are parallel to the hematite surface as a function of system size. The simulation boxes and simulation sizes are given in Section S3 in the ESI.† The red and orange bars show D‖PBC of water molecules located in the first solvation layer at the surface and in the bulk liquid phase, respectively. The error bars are obtained by block averaging. The horizontal black dashed line represents the experimental self-diffusion coefficient of pure bulk liquid water (2.61 × 10−5 cm2 s−1) obtained by NMR measurements.73 |
The behavior of DPBCxy as a function of the simulation box size implies the presence of significant finite-size effects. It is well-known for cubic simulation boxes that the diffusion coefficient depends on the simulation box size and there are formulae available to quantify this finite-size effect as well as to extrapolate the value to its value at infinity.34,35 Finite-size effects for confined systems are less well-known, however, they have been reported.36 The authors find that for wide, thin pores, i.e. when the simulation box size in the confined direction is large, but the corresponding surface area is small, DPBCxy “can be several times larger than Dbulk”.36 The opposite effect, i.e. that DPBCxy converges to its bulk value, is observed when the surface area is large compared to the length of the confined dimension. Qualitatively, we exactly find this behavior and thus ascribe it to a finite-size effect arising due to the confinement.
Moreover, we applied the given finite-size correction equation36 (see eqn (S2) in the ESI†) to estimate the error of the self-diffusion coefficient as a function of system size. The correction term requires knowledge of the viscosity, η, which we estimated from pure bulk liquid water simulations, see Section S5 in the ESI† for details. Due to the strong interactions between water and the hematite slab, we expect that the viscosity of bulk water molecules and molecules in the first solvation layer are different. Therefore, we only applied the correction to bulk water molecules in the hematite/water box and not to those water molecules being in the first solvation layer. Due to these two approximations (η estimated for pure bulk water and the probable dependence of η on the distance from the slab), we interpret the finite-size correction only as an estimate of the error of D as a function of system size. Inspecting Fig. 5 again, we indeed find that the finite-size correction is largest for the 80 Å box, where the surface area (about 10 × 10 Å) is comparably small with respect to the 80 Å in the perpendicular dimension. In contrast, for the largest box (20 × 20 × 80 Å) the finite-size correction is quite small. Therefore, we conclude that the calculated self-diffusion coefficients with the latter box should be a reliable estimate. This conclusion is further supported by available experimental self-diffusion coefficient of pure bulk liquid water which closely agrees with the calculated one in the largest box. Importantly, when using explicit AIMD simulations, usually only a single simulation box is affordable. Fig. 5 shows that if a unsuitable simulation box is chosen accidentally, the calculated self-diffusion coefficients can potentially be far off due to substantial finite-size effects.
A similar red-shift of the intramolecular OH stretch of water was also measured previously at the Al2O3(110)/water interface by vibrational SFG spectroscopy12 and assigned to strongly H-bonded species. By computing the VDOS of interfacial water molecules at Al2O3(0001) and SiO2(0001) interfaces, it could also be shown that H-bonds donated to and accepted from the oxide material can show different spectral signals.24 Two different bands could indeed also be measured by vibrational SFG spectra at the Al2O3/water interface and their respective band strengths change as function of pH value.15
Besides the H-bond kinetics at the interface, also the in-plane/out-of-plane kinetics of the terminal OH groups play an important role. We find that the IP/OOP interchange occurs every 11 ps on average (see Fig. 3). Note that this value, obtained from extensive sampling, is about one order of magnitude slower than previously reported by short explicit AIMD simulations.26 In contrast, the average H-bond lifetime between a terminal OH group and a water molecule is 1.58 ps. The difference between these two values (about one order of magnitude) indicate that the IP/OOP interchange and H-bond dynamics between the hematite slab and adjacent water molecules are two different dynamical processes which reside on different time scales, but which must obviously be coupled. Previously, we could show26 that there is a concerted IP/OOP interchange involving many different surface hydroxyls. It is suggestive, that this requires the first layer water molecules to form some pre-organization H-bond network state, which is capable to accept (or even to facilitate) such a concerted IP/OOP interchange of multiple surface hydroxyl groups. This might explain why the H-bond dynamics at the interface is faster than the concerted interchange of multiple surface hydroxyls.
In our previous publication, H-bond lifetimes on the order of 150 fs have been reported. We also came to the conclusion, that H-bond lifetimes between the hematite and adjacent water molecules are shorter compared to water/water H-bonds.26 This seeming discrepancy between the previous and current H-bond lifetimes can be understood in terms of the definition of the employed H-bond criteria: previously,26 the renowned Luzar–Chandler criterion has been used which had also been employed before in conjunction with the reactive flux technique to determine H-bond lifetimes.65 However, this criterion does not cut through the saddle point between H-bonded and non H-bonded states, see Section S7 in the ESI.† Therefore, it is known to include transient ruptures of the H-bond due to fast librational motion of the water molecules.75 This means that as soon as a participating water molecule slightly rotates out of the H-bond criterion (which is guaranteed to happen on the librational motion time scale), the H-bond is counted as broken. Note that the reactive flux technique explicitly accounts for such short time bond ruptures65,75 such that the ultra-short discontinuities of the H-bond are not counted as its rupture. Using the continuous H-bond definition64 together with the Luzar–Chandler geometrical H-bond criterion therefore maps the librational motion which resides in between 300 and 1000 cm−1 in case of liquid ambient water,76 corresponding to vibration periods in between about 34 and 111 fs. Clearly, the H-bond lifetime on the order of 150 fs reported previously falls into that time regime. It is known that the libration frequency of a water molecule blue-shifts, i.e. the vibration becomes faster, if the H-bond strength increases.68 In this work, we clearly show that the hematite/water H-bond is stronger than a water/water H-bond and thus the librational motion of first layer water molecules becomes faster than in bulk water. Since the continuous H-bond definition employing the Luzar–Chandler criterion exactly recovers that vibrational period, it is clear that the previously calculated lifetime26 was found to be shorter than the water/water H-bond lifetime.
Finally, the self-diffusion coefficient of first layer water molecules is significantly smaller compared to water molecules in the bulk phase (see Fig. 5) implying that water molecules in the first layer are systematically hindered or slowed down. Clearly, in case of hematite, the slowdown is mainly caused because of the strong H-bonds across the hematite/liquid water interface, see above. From a macroscopic perspective, a slowdown of the water molecules at the interface comes at no surprise, because friction is generally expected,36 also for solids which do not form any H-bonds with water, such as graphene and boron nitride77 or metals, such as Cu.39 However, the quantitative strength of this slowdown clearly depends on the atomistic structure of the solid and the strength of the interfacial interactions. It can also depend on the exposed facet as shown in case of the Cu/water interface39 as well as on the exposed microstructure, e.g. if solid defects are present at the interface.78
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d2cp01708c |
This journal is © the Owner Societies 2022 |