We investigate the density isobar of water and the melting temperature of ice using six different density functionals. Machine-learning potentials are employed to ensure computational affordability. Our findings reveal significant discrepancies between various base functionals. Notably, even the choice of damping can result in substantial differences. Overall, the outcomes obtained through density functional theory are not entirely satisfactory across most utilized functionals. All functionals exhibit significant deviations either in the melting temperature or equilibrium volume, with most of them even predicting an incorrect volume difference between ice and water. Our heuristic analysis indicates that a hybrid functional with 25% exact exchange and van der Waals damping averaged between zero and Becke–Johnson dampings yields the closest agreement with experimental data. This study underscores the necessity for further enhancements in the treatment of van der Waals interactions and, more broadly, density functional theory to enable accurate quantitative predictions for molecular liquids.

Water is ubiquitous on our planet, and life on Earth is unthinkable without it. Therefore, it is not surprising that water has received a great deal of attention from the scientific community. It was also one of the first liquids to be intensively studied by computer simulations.1,2 At present, carefully constructed force fields play a key role in the study of water.3–33 The best force fields are constructed to give an excellent description of all relevant properties of water (density, viscosity, interfacial free energy, phase transitions, etc.) while also giving outstanding agreement with high-level quantum chemistry calculations for small clusters. More recently, there has been a gradual shift toward supplementing or even substituting force fields with first-principles electronic structure calculations. These calculations predominantly rely on density functional theory (DFT), with the initial studies emerging in the early 1990s.34–37 As computing power advanced, the calculations evolved to encompass more meaningful timescales and systems.38–51 The advent of machine learning potentials has led to a further leap in the number of properties that can be predicted and the timescales that can be simulated.52–65 In particular, notable is the capability to simulate density isobars and melting temperatures within a matter of days or even hours, rendering the comparison of these properties with experimental data straightforward.

Indeed, there is little doubt that the density isobars predicted by density functional theory are not yet on par with those obtained from carefully constructed semi-empirical models.66–68 So why bother with the costly density functional theory calculations? The explanation is straightforward: as our microscopic understanding of water has progressed over time, the scientific community has redirected its focus toward more intricate questions. These encompass the interactions between water and surfaces, the nucleation of water at nanoparticles, and processes such as the electrolysis-induced hydrogen production or the hydrogen-to-water conversion in fuel cells.65 Understanding the interface between water and surfaces, frequently metallic, is imperative in these cases. In particular, the simulations of electrolysis and fuel cells, where bonds are continuously broken and reformed, necessitate the application of first-principles calculations. However, conducting high-level quantum chemical calculations, for which some force fields have been calibrated, remains challenging for surfaces or practically impossible for metallic ones. Consequently, density functional theory is the go-to approach. Employing density functional theory potentially adversely impacts the properties of water, as already discussed above. So ideally, one seeks a density functional that accurately describes the liquid state of water, encompassing at least the density and melting temperature. It is, however, understood that this alone will not suffice to describe other properties such as the solvation energy of ions, transport properties of protons or ions, or interfacial energies, but simple thermodynamic properties are a first stepping stone. Overall, identifying a density functional that reasonably describes the density isobars and melting point of water is a pressing scientific issue that will strongly impact future water studies. This is the core goal of the present work, where we focus on testing commonly applied density functionals via machine learning potentials. Ionic quantum effects are neglected in the present study but are expected to be small (1% and 2% for the density and temperatures, respectively).60 

Our methodology closely follows our previous work,69 where we have constructed machine-learning potentials (MLPs) for water using the RPBE-D3/zd density functional. The adopted strategy uses first on-the-fly training along the isobar using NPT simulations. Simulations cover the melting of ice and water in a temperature range between the supercooled state and slightly below the evaporation point. Further structures are added from parallel-tempering simulations. The final dataset consists of about 1500 structures with typically 64 molecules and corresponds to dataset 1 in Ref. 69. This will be referred to as the “base” dataset in the following. To obtain a highly accurate database, we have opted to use the hardest and most accurate potentials available for VASP70,71 (O_h and H_h). Following Ref. 69, the Brillouin zone was sampled at the Γ-point only, and the final datasets correspond to a 2000 eV energy cutoff to avoid errors in the pressure. For the present work, we use the MLP implementation available in VASP. The on-the-fly training relies on Gaussian process regression using Bayesian regression.72–74 For the production runs, we rely on kernel-ridge regression to solve the least squares problem using a singular value decomposition.75,76

In a first attempt, we re-calculated the entire dataset 1 from Ref. 69 using all considered semi-local functionals at 2000 eV. We expect this strategy to work well for functionals that yield a volume within 5% of the base functional (RPBE-D3/zd), as the volume fluctuations during training are typically 5%. For SCAN (very different melting temperature and density) and optPBE (significantly larger density), we additionally performed a complete on-the-fly training using the same strategy as outlined in Ref. 69. In Table I, this strategy is named “on-the-fly.” For the hybrid functional revPBE0, we first upgraded a smaller on-the-fly RPBE-D3/zd database containing 662 structures (dataset 2 in Ref. 69) to revPBE0-D3/zd using Δ-learning, then continued training starting from this force field and refining it using on-the-fly training. This on-the-fly training was done using the hybrid functional, an energy cutoff of 1100 eV, and cooling liquid water twice from 350 to 320 K, respectively, to 250 K, as well as melting cubic ice. The final dataset comprises 1010 structures and covers a much larger set of volumes than the original database. We term the structures in this database “hybrid” in the following.

TABLE I.

Training strategy for the various functionals used in the present work. For the “base” strategy, the original RPBE-D3/zd dataset 1 of Ref. 69 was recalculated using the desired functional and then used to train an MLP, except for the revPBE0-D3/BJ base. For this case, the baseline RPBE-D3/zd database of 1500 structures was upgraded using Δ-learning as described in the main text. For “hybrid,” a smaller database of 1010 structures covering a wider volume range was upgraded using Δ-learning. For optPBE and revPBE0-D3/BJ, we obtained results for two different strategies. The “on-the-fly” strategy follows the same protocol as for dataset 2 of Ref. 69. References to functionals and the D3 method with the respective damping functions are also shown.

DFTvdWStrategy
RPBE77  D3/zd78  Base 
RPBE D3/BJ79  Base 
revPBE80  D3/zd Base 
revPBE D3/BJ Base 
optPBE81   Base and on-the-fly 
SCAN82   On-the-fly 
revPBE083  D3/zd Hybrid 
revPBE0 D3/BJ Base and hybrid 
PBE0 D3/BJ Hybrid 
PBE0 D3/zd Hybrid 
DFTvdWStrategy
RPBE77  D3/zd78  Base 
RPBE D3/BJ79  Base 
revPBE80  D3/zd Base 
revPBE D3/BJ Base 
optPBE81   Base and on-the-fly 
SCAN82   On-the-fly 
revPBE083  D3/zd Hybrid 
revPBE0 D3/BJ Base and hybrid 
PBE0 D3/BJ Hybrid 
PBE0 D3/zd Hybrid 

To obtain the final MLPs for hybrid functionals, baseline RPBE-D3/zd calculations were conducted at 2000 eV for all structures in the original dataset (either the base or the hybrid dataset). Subsequently, Δ-learning was employed to “upgrade” this dataset to different hybrid functionals. This was achieved by machine learning the energy, force, and stress differences between the desired hybrid functional and the RPBE-D3/zd functional at 1100 eV for 300–600 structures. When learning the difference between the two functionals, we found that using only 5 and 64 structures resulted in test set errors of 6 and 2.5 meV/Å, respectively. This does not add a significant error to the baseline error of about 26 meV/Å for semi-local functionals (error propagation follows the Gaussian law). The Δ-force field obtained in this manner is calculated for all structures in the original RPBE-D3/zd dataset, and the predicted energy, force, and stress differences are added to the baseline dataset. This upgraded dataset is then passed to the machine-learning algorithm and subsequently used in the production runs. Δ-learning was used to upgrade the “base” dataset for revPBE0-D3/BJ and the “hybrid” dataset for all hybrid functionals.

The Δ-learning approach taken here does not introduce sizable additional errors. Indeed, comparison with a test set calculated at 2000 eV using the hybrid functional gives a force error of 28 meV/Å, which is only slightly larger than the errors we found for semi-local density functionals (26 meV/Å). The main source for the slight error increase is that we added the difference between the hybrid functional and RPBE-D3/zd evaluated at 1100 eV instead of 2000 eV.

Nevertheless, it is important to be cautious when employing Δ-learning. The primary sources of errors stem not from capturing the difference between the two functionals but rather from potential gaps in the training dataset, which may fail to cover the necessary range of configurations. Consequently, predictions for the baseline semi-local functional can become relatively inaccurate, in the present case, if there is a significant difference between the volumes predicted by the hybrid and semi-local functional. The hybrid dataset remedies this issue since it spans a volume range covering ±10% around the experimental density.

Figure 1 shows the density–temperature isobars obtained using a Langevin thermostat and a Parrinello–Rahman barostat (friction coefficients in both cases being 2 ps−1). Parallel tempering is performed for 128 molecules (for further details, we refer to Ref. 69). As shown in Ref. 88, parallel tempering notably improves the sampling of liquid water densities. In most simulations, we see that images at low temperatures tend to stop exchanging with images at higher temperatures. Although the structure of ordered ice is not recovered, the density around 210 K is always very close to the density of hexagonal ice reported in Table II. This can also be seen in Fig. 2, where we show the difference in density between ice and the parallel tempering runs along the isobar. Therefore, these structures are best categorized as low-density amorphous ice.89–91 

FIG. 1.

Density as a function of temperature at a pressure of ∼1 bar (isobars). Results have been obtained using parallel-tempering simulations for 128 molecules, as detailed in Ref. 69. Results for optPBE and revPBE0-D3/BJ are shown using open symbols and obtained using the base dataset. The black solid line corresponds to the experimental measurements (exp.).84–87 The type of dispersion correction (D3) has been omitted in the legend, keeping only the damping.

FIG. 1.

Density as a function of temperature at a pressure of ∼1 bar (isobars). Results have been obtained using parallel-tempering simulations for 128 molecules, as detailed in Ref. 69. Results for optPBE and revPBE0-D3/BJ are shown using open symbols and obtained using the base dataset. The black solid line corresponds to the experimental measurements (exp.).84–87 The type of dispersion correction (D3) has been omitted in the legend, keeping only the damping.

Close modal
TABLE II.

Density of hexagonal ice ρice and water ρwater at the melting temperature Tm for the functionals considered in the present work in g/cm3. Values for the melting temperature Tm and the temperature at the maximum density Tmax are also shown in K.

DFTvdWρiceρwaterTmTmax
SCAN  0.966 1.025 299 321 
92   0.949 1.002 311 333 
optPBE  0.909 1.053 239 220 
PBE0 D3/zd 0.960 1.045 303 300 
PBE0 D3/BJ 0.951 0.987 330 327 
revPBE0 D3/zd 0.905 1.073 246 220 
60   0.881 0.942 280 277 
93   ∼0.9 1.000 291 292 
revPBE0 D3/mix 0.910 0.997 270 263 
revPBE0 D3/BJ 0.915 0.958 289 280 
revPBE D3/zd 0.875 0.921 283 284 
revPBE D3/BJ 0.895 0.890 314 287 
RPBE D3/zd 0.861 0.892 278 283 
RPBE fly D3/zd 0.861 0.892 278 283 
54   ⋯ 0.901 274 274 
RPBE D3/BJ 0.892 0.887 310 289 
Experimental  ∼0.92 ∼1.00 ∼273 ∼277 
DFTvdWρiceρwaterTmTmax
SCAN  0.966 1.025 299 321 
92   0.949 1.002 311 333 
optPBE  0.909 1.053 239 220 
PBE0 D3/zd 0.960 1.045 303 300 
PBE0 D3/BJ 0.951 0.987 330 327 
revPBE0 D3/zd 0.905 1.073 246 220 
60   0.881 0.942 280 277 
93   ∼0.9 1.000 291 292 
revPBE0 D3/mix 0.910 0.997 270 263 
revPBE0 D3/BJ 0.915 0.958 289 280 
revPBE D3/zd 0.875 0.921 283 284 
revPBE D3/BJ 0.895 0.890 314 287 
RPBE D3/zd 0.861 0.892 278 283 
RPBE fly D3/zd 0.861 0.892 278 283 
54   ⋯ 0.901 274 274 
RPBE D3/BJ 0.892 0.887 310 289 
Experimental  ∼0.92 ∼1.00 ∼273 ∼277 
FIG. 2.

Density difference between liquid water and ice as a function of temperature at a pressure of ∼1 bar (isobars). The solid black line shows the difference between the experimental value for liquid in Ref. 84 and for ice in Ref. 94. The type of the dispersion correction (D3) has been omitted in the legend, keeping only the damping.

FIG. 2.

Density difference between liquid water and ice as a function of temperature at a pressure of ∼1 bar (isobars). The solid black line shows the difference between the experimental value for liquid in Ref. 84 and for ice in Ref. 94. The type of the dispersion correction (D3) has been omitted in the legend, keeping only the damping.

Close modal

The first important observation is that none of the functionals gives a satisfactory density for liquid water in the full range. At standard conditions, it should be around 1 g/cm3, but the best-performing functionals are about 4% away from experiments, only SCAN being in good agreement. However, such an agreement only holds at that particular thermodynamic state where it accidentally crosses the experimental isobar. In addition, the experimental density difference between liquid water and ice is not well reproduced; experimentally, ice is about 90% less dense than water.

We start the discussion with the semi-local revised Perdew–Burke–Ernzerhof (RPBE77 and revPBE80) functionals, all of which consistently give too low densities (8% lower than experiments, in the best case). The results also depend very much on the choice of damping, i.e., how the attractive van der Waals (vdW) terms are “damped” to zero as the constituents approach each other. The damping is necessary because the approximate semi-local DFT functionals already account for some of the interactions associated with correlation effects. Observing such a strong dependence on the choice of damping is disturbing. In particular, only the zero damping (zd), i.e., the original functional form proposed by Grimme et al.,78 yields a higher density for water than for ice, while the Becke–Johnson damping (BJ)79 incorrectly predicts ice to be denser than water so that the slope of the melting line is likely to be inverted.

The baseline DFT functionals also have some influence on the results. The RPBE and revPBE functionals are closely related. The revPBE functional is a revised version of the original PBE functional. It includes modifications to the exchange part of the functional aimed at improving its accuracy in describing properties such as the reaction energies of molecules. RPBE was primarily aimed at improving adsorption energies at surfaces and, as opposed to revPBE, locally observes the Lieb–Oxford criterion.77 Overall, revPBE-D3/zd gives the best agreement with the experimental density, although it underestimates the density of ice and, even more so, water. This underestimation is also true for the difference in densities between water and ice, as illustrated in Fig. 2. The ice melting points are presented in Table II. They were determined by performing direct coexistence simulations in anisotropic NpT calculations on 864 molecules (half being ice and half being liquid, divided in two rectangular prisms separated by two planar interfaces). Pressure is kept at standard conditions, and various temperatures are covered using a 2 K grid. The temperatures reported are usually the average of the lowest temperature at which melting occurred and the highest temperature at which freezing occurred.

Considering the intrinsic error of direct coexistence simulations and our 2 K grid, we expect the errors to be of about ±2 K. We note that the agreement with our previously reported melting temperature for larger ensembles for RPBE-D3/zd69 as well as with previous work54 is excellent. Table II shows that most semi-local functionals overestimate the experimental value of 273.15 K. Among these, the -zd ones align reasonably well (5 and 10 K higher for RPBE-D3/zd and revPBE-D3/zd, respectively), whereas the -BJ versions notably deviate (37 K higher in the best case), even exceeding the temperature of maximum density.

Now we discuss the results for hybrid functionals as well as optPBE. For optPBE and revPBE0-D3/BJ (“base”), we illustrate results for the original database (open symbols), as well as on-the-fly results for optPBE and results using the hybrid database covering a larger density variation for revPBE0-D3/BJ (filled symbols). The predicted densities are about 1% smaller using the databases constructed from RPBE-D3/zd structures, and the error increases with the density difference to that density functional. The error becomes even larger for revPBE0-D3/zd (not shown). This is a sign of the somewhat limited extrapolation capabilities of the MLPs applied in the present work, but we note that despite a density difference of 10%–20%, density errors remain smaller than 1% for optPBE and revPBE0-D3/BJ.

Let us begin our more detailed discussion with optPBE, a functional that introduces vdW interactions through terms dependent on the density at two distinct points in space. This departs from conventional semi-local density functionals, which typically only consider semi-local information at a single point in space. As already noted above, the isobars for the two training sets are very similar. Both simulations lack a clear density maximum. Instead, the density increases almost linearly with decreasing temperature, with an abrupt jump to a low-density amorphous ice phase around 210 K. This result is reproducible, as confirmed by the two independent MLPs and, thus, independent of database details. A very similar unsatisfactory result is observed for revPBE0-D3/zd. Again, there is no discernible density maximum, and an abrupt transition to low-density amorphous ice structure occurs around 210 K. Both functionals, optPBE and revPBE0-D3/zd, overestimate the density difference between water and ice, which appears to be the primary cause of the lack of a discernible density maximum. In addition, their melting points occur at a very low temperature (around 240 K). On the other hand, the revPBE0-D3/BJ functional predicts densities that are about 5% lower than in experiments, somewhat similar but slightly better than for the semi-local functionals with vdW corrections. Similar to revPBE0-D3/zd, PBE0-D3/zd yields a larger density than in experiments, but now a density maximum is discernible in the isobar, whose shape agrees quite well with experiments although shifted about 30 K to higher temperatures. Its difference in density between water and ice gives the best agreement with experiments but shifted by about 20 K to higher temperatures. For PBE0-D3/BJ, the density isobar is very similar in shape to revPBE0-D3/BJ but shifted to higher densities and temperatures. In fact, the melting points for all the functionals under consideration are, in general, significantly too high (by ∼30 K), except for optPBE and revPBE0-D3/zd, which are too low (by ∼30 K).

It is worth noting that there are certain similarities between semi-local and hybrid functionals regarding the choice of the damping term. In particular, the use of Becke–Johnson (BJ) damping tends to increase the density of ice compared to water. While this adversely affects the performance of semi-local functionals, it has a somewhat positive effect on hybrid functionals, restoring the density maximum. However, now the density difference between ice and water is again too small, so the agreement with the experiment is still far from good.

Empirically mixing the machine-learning potentials for revPBE0-D3 with zero damping and Becke–Johnson damping yields the most favorable outcome (revPBE0-D3/mix). To obtain the revPBE0-mix results, we averaged the energies, forces, and stresses of the revPBE0-D3/BJ and revPBE0-D3/zd datasets with equal weighting and retrained a force-field on that dataset. This is equivalent to averaging the damping factor before the D3 dispersion corrections are multiplied by the damping function and added to the DFT energies and forces. The approach demonstrates excellent agreement in predicting the density of water, as well as its difference with ice, and yields a well-defined density maximum, although at temperatures that are ∼10 K lower than the experimental ones (with the maximum in density occurring slightly below the melting point).

Finally, we comment on the results using the SCAN functional. We observe a difference in the water and ice density, as well as a water density isobar that closely matches the experimental one but shifted to higher temperatures by about 50 K. Overall, the density of water is about 5% too large, but that seems acceptable considering the larger errors of other functionals. The density maximum is also at a temperature (323 K) that is significantly too high, whereas the melting temperature is too low compared to the temperature of the density maximum but still higher than the experimental value (by ∼ 30). Hence, if one desires to simulate liquid water using SCAN, the temperature should be set well above 330 K.

The density isobars and melting temperatures have been calculated before for several functionals, including RPBE-D3/zd, SCAN, and revPBE0-D3/zd. For RPBE-D3/zd, we have already documented good agreement (density and melting temperatures within 1% and 2%, respectively) with FHI-AIMS54 in our previous work.69 We also summarize these results in Table II and show the isobars in the supplementary material.

For SCAN, about 10 K higher melting temperatures and temperatures of the density maxima92 and 1.5% smaller densities have been reported,95 overall in good agreement with the present work. The discrepancies might arise from a different plane wave energy cutoff (1500 eV in Refs. 92 and 95 and 2000 eV here), which we found quite relevant for water simulations.69 

For revPBE0-D3/zd, the disagreement with previous work60,93 is more troublesome, but previous publications also differ significantly from each other. Most of the discrepancies seem to be a consequence of the higher water density in the present work.

To confirm the water densities predicted via the MLPs, we estimated the DFT volume using histogram reweighting for RPBE-D3/zd, SCAN, and revPBE0-D3/zd. We performed 1.5 ns simulations (1 mio time steps, H mass set to 8 a.u.) for 64 molecules at 273 K (RPBE-D3/zd and revPBE0-D3/zd) and 325 K (SCAN), and selected 1000 equally spaced configurations for the three cases. These were re-weighted using the energy difference between the desired functional and the MLP. For revPBE0-D3/zd, we chose an energy cutoff of 1100 eV; for RPBE-D3/zd and SCAN, 1100 and 2000 eV were used (both cutoffs yield identical results for the corrected volume). In all cases, the reweighted average DFT density is slightly smaller than the MLP density (0.1%–0.2%); however, the statistical uncertainties are about 0.5%. In particular, we note that the volume and the energy difference between DFT and the MLPs are not statistically significantly correlated. This evaluation confirms that the MLP-predicted volume is consistent with the VASP revPBE0-D3/zd ground truth.

We speculate that the shortcoming of previous calculations is related to specifics of the codes (see the supplementary material) or holes in the MLP database: when we set out from the initial RPBE-D3/zd database, the density isobars were similar to previous calculations, but as we added more high-density data from the on-the-fly learning, the isobars approached the present results. Furthermore, we show in the supplementary material that a cutoff of 6 Å on the exact exchange—as used in previous calculations—favors lower densities.

It is not easy to predict or disentangle the trends associated with different density functionals, but some relationships can be identified. It appears that the equilibrium density of water and ice is related to the strength of the two-body interactions predicted by the functional. In particular, it has been found that SCAN predicts a significantly too large two-body interaction,96–98 while RPBE-D3 predicts too weak two-body interactions.98 In line with these considerations, the BLYP-D3 functional predicts quite accurate two-body interactions and very good densities.53,98

However, the shape of the density isobars and melting curves are difficult to relate to simple properties. The likely explanation is that the maximum in the density isobar relates to a subtle balance between three-body and two-body interactions, for which the present density functionals show a relatively large scatter. The only consistent observation that emerges from our study is that an accurate reproduction of the density difference between ice and water seems to be crucial for a reasonable isobar shape. Functionals that overestimate the density difference fail to produce a density maximum similar to the experiment. Instead, they predict an abrupt transition to a low-density amorphous ice phase (optPBE and revPBE0-D3/zd). Conversely, functionals with too small a density difference, or one in the right range, tend to produce more reasonable density isobar results, although potentially shifted in temperature and density. Notably, the SCAN functional falls into the latter category, with a density that is too high (5%) and a temperature deviation of almost 20%.

Regarding the damping of the vdW interactions, we can rationalize our observations as follows: VdW interactions only affect the two-body terms (3-body vdW terms were not considered in the present work). For water, the vdW corrections are typically 25% larger for zero damping than for Becke–Johnson damping for all functionals considered. This implies that the density is consistently greater for zero damping (compare Fig. 1). Furthermore, since ice is less dense and the inter-molecular distances are larger, vdW corrections tend to be smaller for ice-like structures. This means that ice is less affected by vdW corrections than water, especially for Becke–Johnson damping. In other words, increasing the strength of the vdW corrections (BJ → zd) also increases the density difference between water and ice, which can be advantageous or disadvantageous depending on the baseline functional.

In summary, we have assessed the density isobar of water and the melting temperature of ice for semi-local and hybrid functionals with vdW corrections across a range of commonly used density functionals. The outcomes, unfortunately, are not particularly pleasing, as none of the investigated functionals show good agreement with experimental density isobars. Although the shape of the isobars may resemble the experimental one for some functionals, a quantitative agreement would require a shift in density and temperature. It is challenging to determine whether the inadequacy lies in the vdW correction or the specifics of the density functionals, but both factors likely contribute. In addition, it is important to recognize that correlation effects are inherently non-separable, as they operate at all distances and wavelengths, and electronic excitations at one length-scale mix with those at other length scales. This makes it challenging to isolate vdW interactions from other contributions. Employing a damping function to truncate vdW interactions at a certain range is necessarily heuristic, and even parameter fitting to an extensive dataset does not ensure accurate predictions beyond the test set or for specific problems such as water and ice.

Among the semi-local functionals, we recommend either revPBE-D3/zd, as it provides the best overall density and accurate temperatures for the density maximum and melting, or RPBE-D3/zd, which is only slightly inferior for the density, although better in the melting point, both giving similar differences in the density between water and ice. However, a common issue with all semi-local functionals based on revised versions of PBE is an underestimation of the density of ice and water. This inadequacy is likely related to too weak two-body interactions.98 

In the case of hybrid functionals, there are often only small improvements in terms of densities and melting points, although previous publications report notably better performance for other crucial properties such as proton transport and reactivity.99,100 Zero damping results are generally unsatisfactory, as the density difference between ice and water is too large, precluding the observation of an isobar density maximum resembling experimental results. Although BJ damping yields more reasonable results, they are not markedly superior to those using the semi-local functionals with zero damping for the considered properties. We only achieve excellent agreement with the experiment when combining an MLP with zero damping and BJ damping. We conclude that the state of density functionals for water is unsatisfactory, and more development effort will be required to resolve this issue. A simple fix suggested by the present work is to adjust the damping parameters in the vdW functionals specifically to water. However, this will adversely affect the performance of the vdW corrections for other systems. For revPBE0, we have empirically observed that averaging the standard (zero-damping) and Becke–Johnson damping functions yields a very reasonable density isobar.

In the supplementary material, the density isobars of this work are compared to those of previous studies, and we attempt to clarify discrepancies.

This research was funded in part by the Austrian Science Fund (FWF) through the SFB TACO No. 10.55776/F81 and in part by 10.55776/I6467. For open access purposes, the author has applied a CC BY public copyright license to any author-accepted manuscript version arising from this submission. Computer resources and technical assistance were provided by the Vienna Scientific Cluster (VSC).

The authors have no conflicts to disclose.

Pablo Montero de Hijes: Conceptualization (equal); Data curation (equal); Investigation (equal); Methodology (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal). Christoph Dellago: Conceptualization (equal); Funding acquisition (equal); Investigation (equal); Writing – review & editing (equal). Ryosuke Jinnouchi: Methodology (equal); Software (equal); Writing – review & editing (equal). Georg Kresse: Conceptualization (equal); Data curation (equal); Funding acquisition (equal); Investigation (equal); Methodology (equal); Resources (equal); Software (equal); Supervision (equal); Validation (equal); Visualization (equal); Writing – original draft (equal); Writing – review & editing (equal).

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

1.
J.
Barker
and
R.
Watts
, “
Structure of water; a Monte Carlo calculation
,”
Chem. Phys. Lett.
3
(
3
),
144
145
(
1969
).
2.
A.
Rahman
and
F. H.
Stillinger
, “
Molecular dynamics study of liquid water
,”
J. Chem. Phys.
55
(
7
),
3336
3359
(
1971
).
3.
V.
Molinero
and
E. B.
Moore
, “
Water modeled as an intermediate element between carbon and silicon
,”
J. Phys. Chem. B
113
(
13
),
4008
4016
(
2009
).
4.
K. M.
Dyer
,
J. S.
Perkyns
,
G.
Stell
, and
B.
Montgomery Pettitt
, “
Site-renormalised molecular fluid theory: On the utility of a two-site model of water
,”
Mol. Phys.
107
(
4-6
),
423
431
(
2009
).
5.
W. L.
Jorgensen
, “
Quantum and statistical mechanical studies of liquids. 10. Transferable intermolecular potential functions for water, alcohols, and ethers. application to liquid water
,”
J. Am. Chem. Soc.
103
(
2
),
335
340
(
1981
).
6.
H. J.
Berendsen
,
J. P.
Postma
,
W. F.
van Gunsteren
, and
J.
Hermans
, “
Interaction models for water in relation to protein hydration
,” in
Intermolecular Forces: Proceedings of the fourteenth Jerusalem Symposium on Quantum Chemistry and Biochemistry Held in Jerusalem, Israel, April 13–16, 1981
(
Springer
,
1981
), pp.
331
342
.
7.
W. L.
Jorgensen
,
J.
Chandrasekhar
,
J. D.
Madura
,
R. W.
Impey
, and
M. L.
Klein
, “
Comparison of simple potential functions for simulating liquid water
,”
J. Chem. Phys.
79
(
2
),
926
935
(
1983
).
8.
H. J.
Berendsen
,
J. R.
Grigera
, and
T. P.
Straatsma
, “
The missing term in effective pair potentials
,”
J. Phys. Chem.
91
(
24
),
6269
6271
(
1987
).
9.
H. W.
Horn
,
W. C.
Swope
,
J. W.
Pitera
,
J. D.
Madura
,
T. J.
Dick
,
G. L.
Hura
, and
T.
Head-Gordon
, “
Development of an improved four-site water model for biomolecular simulations: TIP4P-Ew
,”
J. Chem. Phys.
120
(
20
),
9665
9678
(
2004
).
10.
J.
Abascal
,
E.
Sanz
,
R.
García Fernández
, and
C.
Vega
, “
A potential model for the study of ices and amorphous water: TIP4P/ice
,”
J. Chem. Phys.
122
(
23
),
234511
(
2005
).
11.
J. L.
Abascal
and
C.
Vega
, “
A general purpose model for the condensed phases of water: TIP4P/2005
,”
J. Chem. Phys.
123
(
23
),
234505
(
2005
).
12.
S.
Izadi
,
R.
Anandakrishnan
, and
A. V.
Onufriev
, “
Building water models: A different approach
,”
J. Phys. Chem. Lett.
5
(
21
),
3863
3871
(
2014
).
13.
S.
Piana
,
A. G.
Donchev
,
P.
Robustelli
, and
D. E.
Shaw
, “
Water dispersion interactions strongly influence simulated structural properties of disordered protein states
,”
J. Phys. Chem. B
119
(
16
),
5113
5123
(
2015
).
14.
M. A.
González
and
J. L.
Abascal
, “
A flexible model for water based on TIP4P/2005
,”
J. Chem. Phys.
135
(
22
),
224516
(
2011
).
15.
S.
Habershon
,
T. E.
Markland
, and
D. E.
Manolopoulos
, “
Competing quantum effects in the dynamics of a flexible water model
,”
J. Chem. Phys.
131
(
2
),
024501
(
2009
).
16.
F. H.
Stillinger
and
A.
Rahman
, “
Improved simulation of liquid water by molecular dynamics
,”
J. Chem. Phys.
60
(
4
),
1545
1557
(
1974
).
17.
M. W.
Mahoney
and
W. L.
Jorgensen
, “
A five-site model for liquid water and the reproduction of the density anomaly by rigid, nonpolarizable potential functions
,”
J. Chem. Phys.
112
(
20
),
8910
8922
(
2000
).
18.
S. W.
Rick
, “
A reoptimization of the five-site water potential (TIP5P) for use with Ewald sums
,”
J. Chem. Phys.
120
(
13
),
6085
6093
(
2004
).
19.
H.
Nada
, “
Anisotropy in geometrically rough structure of ice prismatic plane interface during growth: Development of a modified six-site model of H2O and a molecular dynamics simulation
,”
J. Chem. Phys.
145
(
24
),
244706
(
2016
).
20.
S.
Izadi
and
A. V.
Onufriev
, “
Accuracy limit of rigid 3-point water models
,”
J. Chem. Phys.
145
(
7
),
074501
(
2016
).
21.
L.-P.
Wang
,
T. J.
Martinez
, and
V. S.
Pande
, “
Building force fields: An automatic, systematic, and reproducible approach
,”
J. Phys. Chem. Lett.
5
(
11
),
1885
1891
(
2014
).
22.
R.
Fuentes-Azcatl
and
J.
Alejandre
, “
Non-polarizable force field of water based on the dielectric constant: TIP4P/ɛ
,”
J. Phys. Chem. B
118
(
5
),
1263
1272
(
2014
).
23.
P. T.
Kiss
and
A.
Baranyai
, “
A systematic development of a polarizable potential of water
,”
J. Chem. Phys.
138
(
20
),
204507
(
2013
).
24.
J.
Reimers
,
R.
Watts
, and
M.
Klein
, “
Intermolecular potential functions and the properties of water
,”
Chem. Phys.
64
(
1
),
95
114
(
1982
).
25.
C.
Tainter
,
P. A.
Pieniazek
,
Y.-S.
Lin
, and
J. L.
Skinner
, “
Robust three-body water simulation model
,”
J. Chem. Phys.
134
(
18
),
184501
(
2011
).
26.
H.
Yu
,
T.
Hansson
, and
W. F.
van Gunsteren
, “
Development of a simple, self-consistent polarizable model for liquid water
,”
J. Chem. Phys.
118
(
1
),
221
234
(
2003
).
27.
G. S.
Fanourgakis
and
S. S.
Xantheas
, “
The flexible, polarizable, thole-type interaction potential for water (TTM2-F) revisited
,”
J. Phys. Chem. A
110
(
11
),
4100
4106
(
2006
).
28.
H.
Jiang
,
O. A.
Moultos
,
I. G.
Economou
, and
A. Z.
Panagiotopoulos
, “
Hydrogen-bonding polarizable intermolecular potential model for water
,”
J. Phys. Chem. B
120
(
48
),
12358
12370
(
2016
).
29.
E. R.
Pinnick
,
S.
Erramilli
, and
F.
Wang
, “
Predicting the melting temperature of ice-Ih with only electronic structure information as input
,”
J. Chem. Phys.
137
(
1
),
014510
(
2012
).
30.
X.
Zhu
,
M.
Riera
,
E. F.
Bull-Vulpe
, and
F.
Paesani
, “
MB-pol (2023): Sub-chemical accuracy for water simulations from the gas to the liquid phase
,”
J. Chem. Theory Comput.
19
,
3551
3566
(
2023
).
31.
V.
Babin
,
C.
Leforestier
, and
F.
Paesani
, “
Development of a “first principles” water potential with flexible monomers: Dimer potential energy surface, VRT spectrum, and second virial coefficient
,”
J. Chem. Theory Comput.
9
(
12
),
5395
5403
(
2013
).
32.
V.
Babin
,
G. R.
Medders
, and
F.
Paesani
, “
Development of a “first principles” water potential with flexible monomers. II: Trimer potential energy surface, third virial coefficient, and small clusters
,”
J. Chem. Theory Comput.
10
(
4
),
1599
1607
(
2014
).
33.
G. R.
Medders
,
V.
Babin
, and
F.
Paesani
, “
Development of a “first-principles” water potential with flexible monomers. III. Liquid phase properties
,”
J. Chem. Theory Comput.
10
(
8
),
2906
2910
(
2014
).
34.
K.
Laasonen
,
F.
Csajka
, and
M.
Parrinello
, “
Water dimer properties in the gradient-corrected density functional theory
,”
Chem. Phys. Lett.
194
(
3
),
172
174
(
1992
).
35.
K.
Laasonen
,
M.
Sprik
,
M.
Parrinello
, and
R.
Car
, “
Ab initio” liquid water
,”
J. Chem. Phys.
99
(
11
),
9080
9089
(
1993
).
36.
M.
Sprik
,
J.
Hutter
, and
M.
Parrinello
, “
Ab initio molecular dynamics simulation of liquid water: Comparison of three gradient-corrected density functionals
,”
J. Chem. Phys.
105
(
3
),
1142
1152
(
1996
).
37.
S. S.
Xantheas
, “
Ab initio studies of cyclic water clusters (H2O)n, n=1–6. III. Comparison of density functional with MP2 results
,”
J. Chem. Phys.
102
(
11
),
4505
4517
(
1995
).
38.
J. C.
Grossman
,
E.
Schwegler
,
E. W.
Draeger
,
F.
Gygi
, and
G.
Galli
, “
Towards an assessment of the accuracy of density functional theory for first principles simulations of water
,”
J. Chem. Phys.
120
(
1
),
300
311
(
2004
).
39.
M.
Chen
,
H.-Y.
Ko
,
R. C.
Remsing
,
M. F.
Calegari Andrade
,
B.
Santra
,
Z.
Sun
,
A.
Selloni
,
R.
Car
,
M. L.
Klein
,
J. P.
Perdew
, and
X.
Wu
, “
Ab initio theory and modeling of water
,”
Proc. Natl. Acad. Sci. U. S. A.
114
(
41
),
10846
10851
(
2017
).
40.
L.
Ruiz Pestana
,
O.
Marsalek
,
T. E.
Markland
, and
T.
Head-Gordon
, “
The quest for accurate liquid water properties from first principles
,”
J. Phys. Chem. Lett.
9
(
17
),
5009
5016
(
2018
).
41.
M. J.
Gillan
,
D.
Alfe
, and
A.
Michaelides
, “
Perspective: How good is DFT for water?
,”
J. Chem. Phys.
144
(
13
),
130901
(
2016
).
42.
A. P.
Gaiduk
,
F.
Gygi
, and
G.
Galli
, “
Density and compressibility of liquid water and ice from first-principles simulations with hybrid functionals
,”
J. Phys. Chem. Lett.
6
(
15
),
2902
2908
(
2015
).
43.
J.
Schmidt
,
J.
VandeVondele
,
I.-F. W.
Kuo
,
D.
Sebastiani
,
J. I.
Siepmann
,
J.
Hutter
, and
C. J.
Mundy
, “
Isobaric–isothermal molecular dynamics simulations utilizing density functional theory: An assessment of the structure and density of water at near-ambient conditions
,”
J. Phys. Chem. B
113
(
35
),
11959
11964
(
2009
).
44.
J.
Wang
,
G.
Román-Pérez
,
J. M.
Soler
,
E.
Artacho
, and
M.-V.
Fernández-Serra
, “
Density, structure, and dynamics of water: The effect of van der Waals interactions
,”
J. Chem. Phys.
134
(
2
),
130901
(
2011
).
45.
G.
Miceli
,
S.
de Gironcoli
, and
A.
Pasquarello
, “
Isobaric first-principles molecular dynamics of liquid water with nonlocal van der Waals interactions
,”
J. Chem. Phys.
142
(
3
),
034501
(
2015
).
46.
M.
Ceriotti
,
J.
Cuny
,
M.
Parrinello
, and
D. E.
Manolopoulos
, “
Nuclear quantum effects and hydrogen bond fluctuations in water
,”
Proc. Natl. Acad. Sci. U. S. A.
110
(
39
),
15591
15596
(
2013
).
47.
M. D.
Baer
,
C. J.
Mundy
,
M. J.
McGrath
,
I.-F. W.
Kuo
,
J. I.
Siepmann
, and
D. J.
Tobias
, “
Re-examining the properties of the aqueous vapor–liquid interface using dispersion corrected density functional theory
,”
J. Chem. Phys.
135
(
12
),
124712
(
2011
).
48.
I.-C.
Lin
,
A. P.
Seitsonen
,
I.
Tavernelli
, and
U.
Rothlisberger
, “
Structure and dynamics of liquid water from ab initio molecular dynamics: Comparison of BLYP, PBE, and revPBE density functionals with and without van der Waals corrections
,”
J. Chem. Theory Comput.
8
(
10
),
3902
3910
(
2012
).
49.
J.
VandeVondele
,
F.
Mohamed
,
M.
Krack
,
J.
Hutter
,
M.
Sprik
, and
M.
Parrinello
, “
The influence of temperature and density functional models in ab initio molecular dynamics simulation of liquid water
,”
J. Chem. Phys.
122
(
1
),
014515
(
2005
).
50.
L.
Zheng
,
M.
Chen
,
Z.
Sun
,
H.-Y.
Ko
,
B.
Santra
,
P.
Dhuvad
, and
X.
Wu
, “
Structural, electronic, and dynamical properties of liquid water by ab initio molecular dynamics based on scan functional within the canonical ensemble
,”
J. Chem. Phys.
148
(
16
),
164505
(
2018
).
51.
A.
Zen
,
Y.
Luo
,
G.
Mazzola
,
L.
Guidoni
, and
S.
Sorella
, “
Ab initio molecular dynamics simulation of liquid water by quantum Monte Carlo
,”
J. Chem. Phys.
142
(
14
),
144111
(
2015
).
52.
J.
Behler
and
M.
Parrinello
, “
Generalized neural-network representation of high-dimensional potential-energy surfaces
,”
Phys. Rev. Lett.
98
(
14
),
146401
(
2007
).
53.
T.
Morawietz
,
V.
Sharma
, and
J.
Behler
, “
A neural network potential-energy surface for the water dimer based on environment-dependent atomic energies and charges
,”
J. Chem. Phys.
136
(
6
),
064103
(
2012
).
54.
T.
Morawietz
,
A.
Singraber
,
C.
Dellago
, and
J.
Behler
, “
How van der Waals interactions determine the unique properties of water
,”
Proc. Natl. Acad. Sci. U. S. A.
113
(
30
),
8368
8373
(
2016
).
55.
A.
Singraber
,
T.
Morawietz
,
J.
Behler
, and
C.
Dellago
, “
Density anomaly of water at negative pressures from first principles
,”
J. Phys.: Condens. Matter
30
(
25
),
254005
(
2018
).
56.
P.
Montero de Hijes
,
S.
Romano
,
A.
Gorfer
, and
C.
Dellago
, “
The kinetics of the ice–water interface from ab initio machine learning simulations
,”
J. Chem. Phys.
158
,
204706
(
20
) (
2023
).
57.
B.
Cheng
,
J.
Behler
, and
M.
Ceriotti
, “
Nuclear quantum effects in water at the triple point: Using theory as a link between experiments
,”
J. Phys. Chem. Lett.
7
(
12
),
2210
2215
(
2016
).
58.
V.
Kapil
,
J.
Behler
, and
M.
Ceriotti
, “
High order path integrals made easy
,”
J. Chem. Phys.
145
(
23
),
234103
(
2016
).
59.
V.
Kapil
,
D. M.
Wilkins
,
J.
Lan
, and
M.
Ceriotti
, “
Inexpensive modeling of quantum dynamics using path integral generalized Langevin equation thermostats
,”
J. Chem. Phys.
152
(
12
),
124104
(
2020
).
60.
B.
Cheng
,
E. A.
Engel
,
J.
Behler
,
C.
Dellago
, and
M.
Ceriotti
, “
Ab initio thermodynamics of liquid and solid water
,”
Proc. Natl. Acad. Sci. U. S. A.
116
(
4
),
1110
1115
(
2019
).
61.
T.
Morawietz
,
O.
Marsalek
,
S. R.
Pattenaude
,
L. M.
Streacker
,
D.
Ben-Amotz
, and
T. E.
Markland
, “
The interplay of structure and dynamics in the Raman spectrum of liquid water over the full frequency and temperature range
,”
J. Phys. Chem. Lett.
9
(
4
),
851
857
(
2018
).
62.
A.
Reinhardt
and
B.
Cheng
, “
Quantum-mechanical exploration of the phase diagram of water
,”
Nat. Commun.
12
(
1
),
588
(
2021
).
63.
B.
Cheng
,
M.
Bethkenhagen
,
C. J.
Pickard
, and
S.
Hamel
, “
Phase behaviours of superionic water at planetary conditions
,”
Nat. Phys.
17
(
11
),
1228
1232
(
2021
).
64.
O.
Wohlfahrt
,
C.
Dellago
, and
M.
Sega
, “
Ab initio structure and thermodynamics of the RPBE-D3 water/vapor interface by neural-network molecular dynamics
,”
J. Chem. Phys.
153
(
14
),
144710
(
2020
).
65.
A.
Omranpour
,
P.
Montero De Hijes
,
J.
Behler
, and
C.
Dellago
, “
Perspective: Atomistic simulations of water and aqueous systems with machine learning potentials
,”
J. Chem. Phys.
160
(
17
),
170901
(
2024
).
66.
J. R.
Espinosa
,
J. L. F.
Abascal
,
L. F.
Sedano
,
E.
Sanz
, and
C.
Vega
, “
On the possible locus of the liquid–liquid critical point in real water from studies of supercooled water using the TIP4P/ice model
,”
J. Chem. Phys.
158
(
20
),
204505
(
2023
).
67.
M. A.
González
,
C.
Valeriani
,
F.
Caupin
, and
J. L.
Abascal
, “
A comprehensive scenario of the thermodynamic anomalies of water using the TIP4P/2005 model
,”
J. Chem. Phys.
145
(
5
),
054505
(
2016
).
68.
S. K.
Reddy
,
S. C.
Straight
,
P.
Bajaj
,
C.
Huy Pham
,
M.
Riera
,
D. R.
Moberg
,
M. A.
Morales
,
C.
Knight
,
A. W.
Götz
, and
F.
Paesani
, “
On the accuracy of the MB-pol many-body potential for water: Interaction energies, vibrational frequencies, and classical thermodynamic and dynamical properties from clusters to liquid water and ice
,”
J. Chem. Phys.
145
(
19
),
194504
(
2016
).
69.
P.
Montero de Hijes
,
C.
Dellago
,
R.
Jinnouchi
,
B.
Schmiedmayer
, and
G.
Kresse
, “
Comparing machine learning potentials for water: Kernel-based regression and Behler–Parrinello neural networks
,”
J. Chem. Phys.
160
(
11
),
114107
(
2024
).
70.
G.
Kresse
and
J.
Furthmüller
, “
Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set
,”
Phys. Rev. B
54
(
16
),
11169
(
1996
).
71.
G.
Kresse
and
J.
Furthmüller
, “
Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set
,”
Comput. Mater. Sci.
6
(
1
),
15
50
(
1996
).
72.
R.
Jinnouchi
,
J.
Lahnsteiner
,
F.
Karsai
,
G.
Kresse
, and
M.
Bokdam
, “
Phase transitions of hybrid perovskites simulated by machine-learning force fields trained on the fly with Bayesian inference
,”
Phys. Rev. Lett.
122
(
22
),
225701
(
2019
).
73.
R.
Jinnouchi
,
F.
Karsai
, and
G.
Kresse
, “
On-the-fly machine learning force field generation: Application to melting points
,”
Phys. Rev. B
100
(
1
),
014105
(
2019
).
74.
R.
Jinnouchi
,
F.
Karsai
,
C.
Verdi
,
R.
Asahi
, and
G.
Kresse
, “
Descriptors representing two- and three-body atomic distributions and their effects on the accuracy of machine-learned inter-atomic potentials
,”
J. Chem. Phys.
152
(
23
),
234102
(
2020
).
75.
C.
Verdi
,
F.
Karsai
,
P.
Liu
,
R.
Jinnouchi
, and
G.
Kresse
, “
Thermal transport and phase transitions of zirconia by on-the-fly machine-learned interatomic potentials
,”
npj Comput. Mater.
7
(
1
),
156
(
2021
).
76.
P.
Liu
,
C.
Verdi
,
F.
Karsai
, and
G.
Kresse
, “
α-β phase transition of zirconium predicted by on-the-fly machine-learned force field
,”
Phys. Rev. Mater.
5
(
5
),
053804
(
2021
).
77.
B.
Hammer
,
L. B.
Hansen
, and
J. K.
Nørskov
, “
Improved adsorption energetics within density-functional theory using revised Perdew-Burke-Ernzerhof functionals
,”
Phys. Rev. B
59
(
11
),
7413
(
1999
).
78.
S.
Grimme
,
J.
Antony
,
S.
Ehrlich
, and
H.
Krieg
, “
A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu
,”
J. Chem. Phys.
132
(
15
),
154104
(
2010
).
79.
S.
Grimme
,
S.
Ehrlich
, and
L.
Goerigk
, “
Effect of the damping function in dispersion corrected density functional theory
,”
J. Comput. Chem.
32
(
7
),
1456
1465
(
2011
).
80.
Y.
Zhang
and
W.
Yang
, “
Comment on “Generalized gradient approximation made simple”
,”
Phys. Rev. Lett.
80
(
4
),
890
(
1998
).
81.
J.
Klimeš
,
D. R.
Bowler
, and
A.
Michaelides
, “
Chemical accuracy for the van der Waals density functional
,”
J. Phys.: Condens. Matter
22
(
2
),
022201
(
2009
).
82.
J.
Sun
,
A.
Ruzsinszky
, and
J. P.
Perdew
, “
Strongly constrained and appropriately normed semilocal density functional
,”
Phys. Rev. Lett.
115
(
3
),
036402
(
2015
).
83.
L.
Goerigk
and
S.
Grimme
, “
A thorough benchmark of density functional methods for general main group thermochemistry, kinetics, and noncovalent interactions
,”
Phys. Chem. Chem. Phys.
13
(
14
),
6670
6688
(
2011
).
84.
V.
Holten
,
C.
Bertrand
,
M.
Anisimov
, and
J.
Sengers
, “
Thermodynamics of supercooled water
,”
J. Chem. Phys.
136
(
9
),
094507
(
2012
).
85.
D.
Hare
and
C.
Sorensen
, “
The density of supercooled water. II. Bulk samples cooled to the homogeneous nucleation limit
,”
J. Chem. Phys.
87
(
8
),
4840
4845
(
1987
).
86.
D. R.
Lide
,
CRC Handbook of Chemistry and Physics
(
CRC Press
,
2004
), Vol.
85
.
87.
T.
Sotani
,
J.
Arabas
,
H.
Kubota
,
M.
Kijima
, and
S.
Asada
, “
Volumetric behaviour of water under high pressure at subzero temperature
,”
High Temp.—High Pressures
32
(
4
),
433
440
(
2000
).
88.
V.
Kapil
,
M.
Rossi
,
O.
Marsalek
,
R.
Petraglia
,
Y.
Litman
,
T.
Spura
,
B.
Cheng
,
A.
Cuzzocrea
,
R. H.
Meißner
,
D. M.
Wilkins
et al, “
i-PI 2.0: A universal force engine for advanced molecular simulations
,”
Comput. Phys. Commun.
236
,
214
223
(
2019
).
89.
A.
Eltareb
,
G. E.
Lopez
, and
N.
Giovambattista
, “
A continuum of amorphous ices between low-density and high-density amorphous ice
,”
Commun. Chem.
7
(
1
),
36
(
2024
).
90.
J.
Wong
,
D. A.
Jahn
, and
N.
Giovambattista
, “
Pressure-induced transformations in glassy water: A computer simulation study using the TIP4P/2005 model
,”
J. Chem. Phys.
143
(
7
),
074501
(
2015
).
91.
R.
Khusnutdinoff
, “
Structural and dynamic features of water and amorphous ice
,”
Colloid J.
79
,
152
159
(
2017
).
92.
P. M.
Piaggi
,
A. Z.
Panagiotopoulos
,
P. G.
Debenedetti
, and
R.
Car
, “
Phase equilibrium of water with hexagonal and cubic ice using the scan functional
,”
J. Chem. Theory Comput.
17
(
5
),
3065
3077
(
2021
).
93.
Z.
Chen
,
M. L.
Berrens
,
K.-T.
Chan
,
Z.
Fan
, and
D.
Donadio
, “
Thermodynamics of water and ice from a fast and scalable first-principles neuroevolution potential
,”
J. Chem. Eng. Data
69
(
1
),
128
140
(
2023
).
94.
A. D.
Fortes
, “
Accurate and precise lattice parameters of H2O and D2O ice Ih between 1.6 and 270 K from high-resolution time-of-flight neutron powder diffraction data
,”
Acta Crystallogr., Sect. B: Struct. Sci., Cryst. Eng. Mater.
74
(
2
),
196
216
(
2018
).
95.
T. E.
Gartner
III
,
L.
Zhang
,
P. M.
Piaggi
,
R.
Car
,
A. Z.
Panagiotopoulos
, and
P. G.
Debenedetti
, “
Signatures of a liquid–liquid transition in an ab initio deep neural network model for water
,”
Proc. Natl. Acad. Sci. U. S. A.
117
(
42
),
26040
26046
(
2020
).
96.
K.
Sharkas
,
K.
Wagle
,
B.
Santra
,
S.
Akter
,
R. R.
Zope
,
T.
Baruah
,
K. A.
Jackson
,
J. P.
Perdew
, and
J. E.
Peralta
, “
Self-interaction error overbinds water clusters but cancels in structural energy differences
,”
Proc. Natl. Acad. Sci. U. S. A.
117
(
21
),
11283
11288
(
2020
).
97.
S.
Dasgupta
,
E.
Lambros
,
J. P.
Perdew
, and
F.
Paesani
, “
Elevating density functional theory to chemical accuracy for water simulations through a density-corrected many-body formalism
,”
Nat. Commun.
12
(
1
),
6359
(
2021
).
98.
E.
Palos
,
E.
Lambros
,
S.
Swee
,
J.
Hu
,
S.
Dasgupta
, and
F.
Paesani
, “
Assessing the interplay between functional-driven and density-driven errors in DFT models of water
,”
J. Chem. Theory Comput.
18
(
6
),
3410
3426
(
2022
).
99.
M.
Chen
,
L.
Zheng
,
B.
Santra
,
H.-Y.
Ko
,
R. A.
DiStasio
, Jr.
,
M. L.
Klein
,
R.
Car
, and
X.
Wu
, “
Hydroxide diffuses slower than hydronium in water because its solvated structure inhibits correlated proton transfer
,”
Nat. Chem.
10
(
4
),
413
419
(
2018
).
100.
V.
Kapil
,
D. P.
Kovács
,
G.
Csányi
, and
A.
Michaelides
, “
First-principles spectroscopy of aqueous interfaces using machine-learned electronic and quantum nuclear effects
,”
Faraday Discuss.
249
,
50
68
(
2024
).

Supplementary Material