Biphasic interfaces are complex but fascinating regimes that display a number of properties distinct from those of the bulk. The CO2–H2O interface, in particular, has been the subject of a number of studies on account of its importance for the carbon life cycle as well as carbon capture and sequestration schemes. Despite this attention, there remain a number of open questions on the nature of the CO2–H2O interface, particularly concerning the interfacial tension and phase behavior of CO2 at the interface. In this paper, we seek to address these ambiguities using ab initio-quality simulations. Harnessing the benefits of machine-learned potentials and enhanced statistical sampling methods, we present an ab initio-level description of the CO2–H2O interface. Interfacial tensions are predicted from 1 to 500 bars and found to be in close agreement with experiment at pressures for which experimental data are available. Structural analyses indicate the buildup of an adsorbed, saturated CO2 film forming at a low pressure (20 bars) with properties similar to those of the bulk liquid, but preferential perpendicular alignment with respect to the interface. The CO2 monolayer buildup coincides with a reduced structuring of water molecules close to the interface. This study highlights the predictive nature of machine-learned potentials for complex macroscopic properties of biphasic interfaces, and the mechanistic insight obtained into carbon dioxide aggregation at the water interface is of high relevance for geoscience, climate research, and materials science.

The boundary between two immiscible fluids represents a unique and fascinating regime. The properties of this biphasic interface, noticeably different from those of bulk fluid, arise from its inhomogeneous composition and the anisotropic distribution of its constituent particles.1 The resulting asymmetry in forces acting between these particles promotes a series of interesting phenomena, including molecular orientational alignment, density gradients and structuring, phase transfer effects, and variations in reactivity and dielectric properties.1–4 

Arguably, the most important biphasic interface is that of carbon dioxide and water, CO2–H2O. This interface exhibits a diverse range of physiochemical properties, which stem from the distinct phases that CO2 can form with subtle changes to temperature and pressure. At ambient temperatures, CO2–H2O exists as a gas–liquid interface at pressures p less than the critical pressure (pC = 73.8 bars; TC = 304.1 K).5 At higher pressures, CO2–H2O forms a liquid–liquid (or supercritical–liquid) system, suggesting that temperature or pressure can be used as a lever to modulate biphasic properties. CO2–H2O finds use in a variety of real-word applications and phenomena, from compressible solvent design6 to ocean acidification monitoring7 to enhanced oil and gas recovery schemes.8 They are also central to various carbon capture and storage (CCS) techniques, where anthropogenic CO2 is captured at source and injected under pressure into predominately aqueous subsurface storage sites, e.g., saline aquifers or disused coal seams.9 In these storage sites, retention of the anthropogenic carbon is mediated by the interactions between injected CO2 and in situ H2O, meaning that both storage capacity and the duration are functions of the interfacial behavior.

Realizing these applications for CO2–H2O requires a detailed knowledge of its biphasic interface. This includes understanding both its microscopic properties (in terms of the structuring of molecules, their orientations, the nature of their bonding, etc.) and its macroscopic properties, such as phase densities and interfacial tensions (IFTs), γ. The IFT, in particular, is critical to understanding biphasic interfaces. This property encapsulates much of the macroscopic nature of the interfacial region and is used to gauge their relative miscibilities. In terms of CCS and CO2–H2O, γ is directly proportional to the sealing capillary pressure, which quantifies the volume of CO2 that can be stored and thus is a good estimate of the efficacy of a chosen storage site.10,11

An extensive picture of both micro- and macroscopic properties of CO2–H2O has been built up over the past decades. In the experimental literature, researchers have focused mostly on the measurement of γ at differing temperatures and pressures.12–22 In general, γ is observed to decrease with either increasing pressure or increasing temperature, although the exact nature of this reduction is dependent upon which CO2 phase is present (see later, Fig. 2). Measurements from experiment, however, are subject to significant variations depending on experimental setup (e.g., thermocouple placement, equilibration times, and density treatment).10 Therefore, experimental work is often supplemented by theoretical and computational estimates for γ. Statistical perturbation models, such as statistical associating fluid theory (SAFT), do a good job in reproducing experimental values but can sometimes lack sufficient microscopic insight into the origins of these results.23–25 In contrast, force field molecular dynamics (FF-MD) simulations provide both estimates for γ and full atomistic details on the behavior of CO2–H2O. Although the exact description of the CO2–H2O interface is sensitive to the choice of the empirical parameters, FFs have provided several important trends and insights. These include the following: CO2–H2O interfaces are molecularly sharp; molecules form a layered structure within several angstrom of the interface; water’s dipole orients parallel to the interface; and there exists a strong, lateral network of hydrogen bonds within the first contact layer of water.10,26–30

Despite the valuable insight uncovered from these studies, a number of open questions remain surrounding CO2–H2O. For example, how exactly does γ behave in the vicinity of CO2’s critical point (pC, TC)? How do high pressures impact the value of γ? In addition, what is the nature of interfacial CO2 and H2O molecules and how exactly do they interact with one another?

Building on previous work in the field, this paper aims to provide a new perspective for characterizing the CO2–H2O interface. Our work focuses on generating machine-learned potentials (MLPs), which we use to perform MD simulations with an ab initio level of accuracy. Unlike previously applied methods, our work breaks away from the empirical fitting to experimental or literature values and instead provides estimates from the ground up. MLPs already have a proven history of success in simulating complex systems,31–37 and their use in simulating CO2–H2O would help address many of the open questions on this system. To date, however, an ab initio or MLP description of biphasic fluid interfaces has remained elusive. Biphasic systems, by definition, require larger simulation boxes than bulk for direct ab initio simulations and for generating training data for MLPs. At the same time, converging the properties of the bulk and interfacial regions requires long simulation time, beyond the reach of standard molecular dynamics.

In this work, we bring together ab initio accuracy and thorough sampling to accurately simulate large and complex mixtures. We utilize a number of distinct methodological features—specifically, MLPs for fast and accurate ab initio total energy predictions, an active learning strategy to build compact and representative training sets, and replica exchange simulations for enhanced sampling of the interface. Our approach allows accurate and converged analysis of both macro- and microscopic properties of the CO2–H2O interface. We obtain estimates of the interfacial tension at room temperature as a function of pressure. These are found to be in good agreement with the literature and help discern experimental reference data across the gas–liquid phase transition for CO2. Improved predictions with respect to previous classical MD results, particularly for low pressure regions, are attributed to the greater accuracy of the underlying theory. Having obtained a correct description of the interfacial tension, we study the structural properties of the CO2–H2O interface to obtain key microscopic insights. Notable observations include the buildup of a liquid-like CO2 layer at low pressures and a reduced structuring in the aqueous phase with increasing pressure. We anticipate the future application of our potential to studying other important state points for CO2–H2O as well as characterizing the dynamical system properties, e.g., in the form of diffusion coefficients and interfacial residence times. We believe that our approach will serve as a robust blueprint for investigating other complex biphasic interfaces.

Our approach to modeling the biphasic CO2–H2O interface comprises three established components: high-dimensional neural network potentials38–40 (HD-NNPs) to act as our first-principles force generator; Query by Committee (QbC), an active learning technique for generating and optimizing our structural dataset;40 and replica exchange molecular dynamics (REMD) to expedite the statistical convergence of interfacial properties over multiple thermodynamic state points. Combining these features allows us to apply accurate potential energy surfaces for simulating large, complex molecular systems over extended time periods. An overview of our approach is provided in Fig. 1.

FIG. 1.

Modeling biphasic fluid interfaces. (a) 2D projection of the structural data used to train our neural network potential. (b) Query-by-committee workflow for the creation of a compact, structurally diverse dataset for each constituent trajectory. (c) Application of the model using replica exchange molecular dynamics toward the determination of macroscopic and microscopic properties.

FIG. 1.

Modeling biphasic fluid interfaces. (a) 2D projection of the structural data used to train our neural network potential. (b) Query-by-committee workflow for the creation of a compact, structurally diverse dataset for each constituent trajectory. (c) Application of the model using replica exchange molecular dynamics toward the determination of macroscopic and microscopic properties.

Close modal

Our work follows the committee NNP development procedure outlined in Ref. 40. We employ Behler–Parrinello HD-NNPs trained on DFT-level ab initio total energies and forces estimated with CP2K.38,41 A set of hand-crafted radial and angular symmetry functions was used.40 We employed the BLYP functional42,43 augmented by Grimme’s D3 corrections.44 This setup has been shown to closely reproduce the condensed-phase and structural properties of both pure CO2 and pure H2O.36,45,46 In addition, BLYP-D3 replicates the critical and vapor–liquid coexistence properties for CO2, making it an appropriate selection for treating the various CO2 phases exhibited within our chosen pressure range.36,46 Goedecker–Teter–Hutter (GTH) pseudopotentials were used for the treatment of the core electrons, TZV2P-GTH basis sets were used for the valence electron density, and a plane wave cutoff of 1050 Ry was used.

Training of our NNPs was implemented using QbC over multiple improvement cycles [see Fig. 1(b) for a schematic description]. This was performed using the open-source AML package.40 A candidate dataset was provided in the form of chemical structures labeled with energies and atomistic forces. Structures were generated using a mixture of flexible force field potentials (SPC/Fw47 + Zhu,48 Lorentz–Berthelot combining rules) as well as preliminary NNPs across three active learning cycles. This gave a total of more than 100 000 structures. QbC was used to trim and optimize this dataset through the iterative selection of those structures displaying the largest uncertainties across each committee member. This was repeated for each trajectory, resulting in a refined dataset of some 4000 structures. Using this dataset, a final model was trained using the openly available N2P2 package.49 

This final, refined dataset is shown in Fig. 1(a). A breadth of structures is included here, including gas–liquid, liquid–liquid, crystalline, and pure-phase structures, in order to maximize exploration of the relevant configuration space. A committee of NNPs was trained using N2P2,49 and the committee member displaying the lowest errors was selected as the final model. This final NNP was validated against the predictions of ab initio MD in the form of energies and forces (test RMSEs: 3.31 meV/atom and 88.26 meV/Å, of similar performance to that reported in Ref. 50 for pure water) as well as structural predictions of pure-phase properties, which are detailed in Sec. I of the supplementary material. Long-range tail corrections for the chosen NNP cutoff are expected to be on the order of 0.5–1 mN/m based on explicit classical model analysis,51 which is smaller than our usual statistical error obtained from block averaging. Due to these reasons, we have not applied any explicit tail correction to our data.

Typical system setups for our production run are shown in Fig. 1(c). Each system consisted of 600 water molecules and 200 CO2 molecules. Lateral dimensions xy were fixed at 20 Å, and periodic boundary conditions were applied along all three axes. Atomic configurations were initialized such that the CO2–H2O interface resides parallel to the xy plane. System sizes and compositions were chosen to mitigate periodic error and finite-size effects, which have been shown in FF-MD to impact IFTs and other interfacial properties.52,53 Results from our system-size analysis in which we varied lateral (xy) dimensions and molecular compositions are shown in Sec. II of the supplementary material. This analysis was performed using SPC/E + EPM2 for lx = ly = 12–30 Å and total number of molecules n = 168–3200. Our choice of box size and composition reflects the smallest system for which γ converges to within 0.5 mN/m of the largest box limit.

Fifteen representative pressures were selected spanning the 1–500 bar range. REMD was performed across 15 different pressures, from which we extracted macroscopic measurements (i.e., IFTs) as well as microscopic properties in the form of density profiles, molecular orientations, and radial distribution functions (RDFs). NNP-REMD was performed using the open-source i-pi package54 with LAMMPS as the force generator and substituting hydrogen for deuterium atoms.55,56 For each state point below 40 bars, simulations were run for 3 ns of simulation time, while above 40 bars simulations were run for 10 ns of simulation time corresponding to 3 × 106 and 107 steps, respectively. These were performed under the NPzAT ensemble, with a fixed number of particles N, a fixed lateral (xy) area A, a fixed temperature (300 K), and Pz set to the target pressure. Collectively, these simulations compile 150 ns of ab initio-quality data, forming a robust basis for our analysis. We have checked the sensitivity of our results on the choice of DFT functional by also training an NNP model to revPBE-D3. Resulting IFTs prove to be very similar overall, and these are shown in Sec. III of the supplementary material.

The interfacial tension is a key property for determining the miscibility of two phases. A number of previous experimental and computational studies have sought to characterize the IFT of CO2–H2O as a function of pressure. In Fig. 2, we plot the results of these studies at roughly room temperature alongside our own estimates obtained using the developed NNP. We compute γ using the statistical mechanical route of Kirkwood and Buff (KB),57 
(1)
where Lz gives the z length of the simulation box and P and P represent the normal and lateral pressure tensor components, respectively.
FIG. 2.

Measuring the interfacial tension (γ) of CO2–H2O. Interfacial tension results for our neural network potential (blue, filled) measured at 300 K for pressures 1–500 bars. Error bars represent the 1σ uncertainty, estimated by integrating the autocorrelation function to a lag of 100 sampling intervals. The selected previous experimental results are shown as hollow markers,12–14,16,17,22 while the computational work of Shiga et al. (force field molecular dynamics) is encompassed within the shaded region.30 

FIG. 2.

Measuring the interfacial tension (γ) of CO2–H2O. Interfacial tension results for our neural network potential (blue, filled) measured at 300 K for pressures 1–500 bars. Error bars represent the 1σ uncertainty, estimated by integrating the autocorrelation function to a lag of 100 sampling intervals. The selected previous experimental results are shown as hollow markers,12–14,16,17,22 while the computational work of Shiga et al. (force field molecular dynamics) is encompassed within the shaded region.30 

Close modal

Inspection of Fig. 2 yields several observations. At low pressures (p < 60 bars), our simulations predict a steep descent in γ with increasing pressure. This phenomenon is associated with the gradual accumulation of CO2 at the water surface, a process which reduces the spatial anisotropy at the phase boundary and thus also the magnitude of γ (see later, Fig. 3). At intermediate pressures (60 p< 100 bars), our NNP results indicate an abrupt change in the behavior of γ as we cross the critical pressure boundary (73.8 bars), going from gas–liquid to liquid–liquid, and finally, at high pressures (p ≥ 100 bars), our model yields a slight reduction in γ with increasing pressure. The comparatively small gradient in γ for this regime is attributed to a saturation in the concentration of CO2 close to the interface such that changes to the pressure lead to minimal change in the near-surface environment. A minimum of 24 ± 2 mN/m is recorded at 500 bars.

FIG. 3.

Profiling the change in CO2–H2O from 1 to 500 bars. (a) Representative snapshots of the CO2 phase and instantaneous interface (yellow mesh) at 1, 20, and 200 bars. H2O has been omitted for visual clarity. (b) Density profiles are shown at different pressures for water (red) and CO2 (blue) with distance from the instantaneous interface that separates them.

FIG. 3.

Profiling the change in CO2–H2O from 1 to 500 bars. (a) Representative snapshots of the CO2 phase and instantaneous interface (yellow mesh) at 1, 20, and 200 bars. H2O has been omitted for visual clarity. (b) Density profiles are shown at different pressures for water (red) and CO2 (blue) with distance from the instantaneous interface that separates them.

Close modal

Across Fig. 2, we then observe close agreement between experimental and our simulation results, within the range of pressures for which experimental data are available. This is true at both low pressures (p < 60 bars), where γNNP replicates the steep descent in gas–liquid IFTs, and high pressures (p ≥ 100 bars), where γNNP follows the gradual decline in liquid–liquid IFTs as reported by Hebach et al.13 and Bachu and Bennion16 A notable aspect of this work is the extension of IFT measurements beyond the currently available experimental pressure range for ∼ 300 K. Moreover, γNNP values recorded at intermediate pressures (60 p< 100 bars) provide additional clarity on the nature of γ in the vicinity of CO2’s phase transition pressure pT (73.8 bars). This region has previously been a point of contention, with experimental results differing by up to 20 mN/m for pressures close to pT, as seen by the selection of results displayed in Fig. 2. Our results clearly support the presence of a break in the slope between gas–liquid and liquid–liquid regimes, as reported by a number of authors.13,16,17 There is no evidence of the “dip” or “cusp” in γ near pT as predicted by some experiments.12,14 Such observations are now thought to be experimental artifacts that stem either from differences in thermocouple placement or from the inherent uncertainty in near-phase-transition properties.10,13

Our results also qualitatively corroborate the results of previous FF simulations. These are encompassed within the shaded portion of Fig. 2, the limits of which represent the extent of γ predictions for varying combinations of FF models (e.g., SPC/E + EPM2, SPC/E + PPL, and TIP4P/2005 + PPL).30 While the exact value of γ is sensitive to the choice of models and parameterization, the overall range of FF predictions is in qualitative agreement with our NNP predictions, with a notable exception of the low pressure regime.30 At these pressures, classical FFs systematically underpredict γ by 15–30 mN/m. We attribute the improved performance of our NNPs to the more accurate level of underlying theory compared to FFs and to the fact that our NNP is trained explicitly to replicate the ab initio-level interactions between CO2 and H2O. In comparison, bicomponent FFs are constructed using geometric/mathematical mixing rules, the choice of which is often ad hoc and can lead to substantial variations in the resulting thermodynamic properties.58–60 This is exemplified by the large spread in γ values shown for FF-MD in Fig. 2.

To better understand the interfacial tension shown in Fig. 2, we plot a series of microscopic profiles detailing the variation in density ρ with distance d from the instantaneous interface. This interface is calculated at each time step using the Willard–Chandler formalism,61 allowing for a full resolution of the dynamic nature of the interface. Resulting profiles are plotted in Fig. 3 for both H2O (red) and CO2 (blue). An inspection of Fig. 3 reveals a stark contrast in the phase behavior of water and CO2. Focusing first on the water phase, we observe a layered structuring in ρ across all pressures, in close agreement with the behavior of the air–water interface.35 The shape of these profiles is relatively unvarying with changes to the pressure, although we do note an enhanced structuring of molecules at p ≤ 20 bars approaching atmospheric pressure. In contrast to this behavior, the CO2 profile exhibits large changes in both magnitude and shape with changes to the pressure. At p < 10 bars, density profiles register a sharp peak at roughly 1 Å with a density that tails off exponentially with distance from the interface. Physically, this represents a film of monolayer CO2 adsorbed at the water surface and a gaseous CO2 phase extending beyond to large distance. At 20 p< 60 bars, a second peak is observed at 4 Å, suggesting the formation of a bilayer of CO2 at the interface. Beyond 60 bars, the exponential decay in density is replaced by liquid-like layering that tends toward bulk density for large distances. For the sampled pressure range, the fraction of CO2 located within the aqueous component increases with increasing pressure. This is a product of both a greater CO2 accumulation at the aqueous surface and a reduction in the interfacial sharpness at higher pressures. The calculated CO2 solubilities are shown in Sec. I of the supplementary material and found to replicate the general trend in experimental results.

Relating these microscopic observations to the IFT profile of Fig. 2, it is clear that the variation in γ is a product of the considerable changes observed in CO2 with changing pressure. Unlike water, CO2 is a nonpolar molecule that exhibits relatively weak intermolecular interactions. Its critical pressure pC resides within our range of sampling, so we end up simulating two distinct phases of CO2 and a phase transition region. In comparison, at 300 K, water is firmly within the liquid phase, and its high natural surface tension—a product of its extensive hydrogen bonding network—ensures preservation of its structural integrity and prevents extensive mixing with the CO2 phase. The most significant change in the structuring of water occurs at low pressures (p ≤ 20 bars). It is interesting to note that this change coincides with the emergence of the monolayer in CO2. The combination of these observations possibly suggests that structuring of the aqueous phase is mediated by the extent of monolayer coverage of the adsorbed phase and that subsequent adsorbed layers have a minimal impact.

Our ab initio-level profiles share many similarities with previous force field modeling.26,27 This includes the formation of monolayer and bilayer CO2 and the recovery of liquid-like properties at high pressures. Interestingly, however, the pressure at which our model predicts the emergence of a CO2 monolayer (and, therefore, also the starting of a second CO2 layer) is much lower compared to conventional FF predictions (e.g., those of SPC/E + EPM2). In our results, a fully saturated monolayer forms at 20 bars; in FFMD studies, this occurs at higher pressures (around 60 bars for SPC/E + EPM2), thereby suggesting a lower wettability for CO2.27 These observations corroborate the ab initio work of Morishita and Shiga,62 who have suggested that differences in the interfacial behavior predicted by ab initio-level modeling and FF-MD stem from the weaker attractive CO2–H2O forces given by the latter. Similarly, we conclude that differences in the density profiles arise from the improved CO2–H2O descriptions provided by NNPs, which are trained to account for both polarization and charge transfer effects. On the other hand, the classical FFs used so far to study this interface are limited by the nature of their mixing rules and their rigid-body formulation, which we would not expect to fully represent the interfacial regime.

Results from the microscopic profiling of CO2–H2O show the formation of a saturated CO2 monolayer at low pressures. To better understand this phenomenon, we investigate the structural properties of CO2 within 2.5 Å of the instantaneous interface, i.e., within the first contact layer. Figure 4 plots both the 2D lateral distribution function (LDF) and angular distribution of these molecules. The LDF, g(r), is calculated such that it accounts for the quasi-2D nature of this monolayer,
(2)
where dnr is the number of CO2 molecules within a shell of thickness dr and ρ gives the local density. Inspection of Fig. 4(a) shows how the structuring of this layer quickly converges toward a bulk-like character. By 20 bars, we see a distribution that emulates that of bulk CO2 and suggests a liquid-like nature for the first contact layer. This is true for pressures 20–500 bars.
FIG. 4.

Characterizing CO2 buildup. (a) Lateral distribution functions for CO2 located within the first layer. (b) Angular distributions of CO2 located within the first layer. Angles are calculated between C–O vector, vCO, and the surface-molecule vector, vsurf. The latter is defined as being the vector connecting the carbon of a CO2 molecule and the point on the instantaneous surface closest to that molecule.

FIG. 4.

Characterizing CO2 buildup. (a) Lateral distribution functions for CO2 located within the first layer. (b) Angular distributions of CO2 located within the first layer. Angles are calculated between C–O vector, vCO, and the surface-molecule vector, vsurf. The latter is defined as being the vector connecting the carbon of a CO2 molecule and the point on the instantaneous surface closest to that molecule.

Close modal

In Fig. 4(b), we show the distribution of angles between the instantaneous interface and the CO2 orientation. At all pressures, our results suggest a clear preference for CO2 to lie flat at the water surface. This corroborates the previous computational work, which suggests that interactions between water and CO2 are maximized when the latter molecule lies parallel at the surface.26 We note that the distribution is more pronounced at 90° at p < 20 bars, where the adsorbed CO2 is more exposed and has fewer molecules to interact with. The addition of more adsorbed molecules at the aqueous surface reduces the strength of water–CO2 interactions, allowing for greater freedom in the latter molecule’s orientational alignment with increasing pressure.

Biphasic interfaces represent complex and dynamic regimes. In this paper, we have highlighted an approach for analyzing these regions in a way that combines ab initio accuracy with converged statistics and computational tractability. Our methodology allowed for nanosecond treatment of large CO2–H2O systems, which has yielded a converged IFT profile at an ab initio level of accuracy. Our results show good agreement with the experimental literature for low- and high-pressure regimes, and a microscopic insight into this behavior has been provided. The reproduction of results in the vicinity of a phase transition (gas to liquid) is notable, given the difficulty associated with treating this highly dynamic regime. We observe the formation of a saturated CO2 monolayer at low pressures (20 bars) with structural properties akin to those of bulk CO2. The emergence of this monolayer coincides with reduced structuring for near-interface water molecules, suggesting that interactions between the first contact layers of water and CO2 are critical for aqueous phase structuring. We envisage that such insights will be important for realizing CO2–H2O’s many applications as well as shedding light on other significant biphasic systems, for example, in biological membranes63 and for liquid–liquid interfaces facilitating nanoparticle assembly.64,65

In terms of carbon capture and sequestration, our NNP provides a robust tool for extending IFT measurements and providing benchmark figures for higher pressure regimes. In this way, coupled with additional geological measurements, one might use our model to provide estimates for a particular storage site of known temperature and pressure, thereby also providing an estimate of that site’s suitability. Compared with statistical perturbation models, our work has the advantage of providing an additional microscopic insight to supplement predictions of γ. Future work could look to exploit this microscopic insight toward estimating other important storage parameters, for example, the contact angle between CO2 and water. It would also be interesting to investigate ideas related to thermodiffusion in underground storage sites. Previous classical work has suggested that CO2 aggregates in areas of low temperature,66,67 i.e., at the top of storage sites, as a result of the Soret effect. Investigating how these predictions change with NNP treatment would allow further comparison between classical and ab initio-level predictions and shed more light on this important phenomenon. A further degree of realism can also be added by incorporating Na+ and Cl ions into our training data and simulations. In doing this, we would better replicate the saline conditions of underground storage reservoirs and thus enable more accurate predictions of a site’s storage capacity.

Our work also unearths several differences between classical and ab initio-level modeling of the CO2–H2O interface. Compared with previous classical results, we suggest that previous analyses of low-pressure CO2 coverage in aqueous systems may require reanalyzing. This could have profound implications for our understanding of processes such as CO2 adsorption and its role in ocean acidification. Higher levels of CO2 adsorption at the ocean surface imply a greater CO2 dissolution and, therefore, a higher overall pH. Extending our current model through the addition of other relevant species, e.g., carbonate, bicarbonate, and carbonic acid, it would be interesting to investigate this acidification process and understand how the adsorption, dissolution, and subsequent reaction of CO2 proceeds under differing conditions.

In conclusion, we provide new insight into the nature of biphasic interfaces and demonstrate the significance of interactions occurring between molecules adsorbed in the first contact layer. We anticipate application of the combined set of techniques used here to other important biphasic systems. The immediate findings of this study about preferred CO2 layering at H2O are expected to be of direct relevance in geoscience, climate research, and materials science.

See the supplementary material for NNP validation tests, system-size convergence tests, and a comparison of IFT predictions for NNP and classical MD.

SGHB was supported by the Syntech CDT and funded by EPSRC (Grant No. EP/S024220/1). C.S. acknowledges financial support from the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) Project No. 500244608. V.K. acknowledges support from the Ernest Oppenheimer Early Career Fellowship and the Sydney Harvey Junior Research Fellowship, Churchill College, University of Cambridge. A.M. acknowledges support from the European Union under the “n-AQUA” European Research Council Project (Grant No. 101071937). We are grateful for computational support and resources from the UK Materials and Molecular Modeling Hub, which is partially funded by EPSRC (Grant Nos. EP/P020194/1 and EP/T022213/1). We are also grateful for computational support and resources from the UK national high-performance computing service, Advanced Research Computing High End Resource (ARCHER2), and the Swiss National Supercomputing Centre under project s1209. Access to both the UK Materials and Molecular Modeling Hub and ARCHER2 was obtained via the UK Car-Parrinello consortium, funded by EPSRC Grant Reference EP/P022561/1. Access to CSD3 was obtained through a University of Cambridge EPSRC Core Equipment Award No. (EP/X034712/1).

The authors have no conflicts to disclose.

Samuel G. H. Brookes: Conceptualization (equal); Investigation (lead); Formal analysis (equal); Writing - original draft (Lead); Writing – review and editing (equal). Venkat Kapil: Conceptualization (equal); Methodology (equal); Formal analysis (equal); Writing – review and editing (equal); Supervision (equal). Christoph Schran: Conceptualization (equal); Methodology (equal); Formal analysis (equal); Writing – review and editing (equal); Supervision (equal). Angelos Michaelides: Conceptualization (equal); Formal analysis (equal); Writing – review and editing (equal); Supervision (equal).

All data required to reproduce the findings of this study are available at https://github.com/water-ice-group/co2-on-h2o.

1.
L.
Benjamin
, “
Mechanism and dynamics of ion transfer across a liquid-liquid interface
,”
Science
261
,
1558
1560
(
1993
).
2.
M. F.
Ruiz-Lopez
,
J. S.
Francisco
,
M. T. C.
Martins-Costa
, and
J. M.
Anglada
, “
Molecular reactions at aqueous interfaces
,”
Nat. Rev. Chem.
4
,
459
475
(
2020
).
3.
S. A.
Shah
and
S.
Baldelli
, “
Chemical imaging of surfaces with sum frequency generation vibrational spectroscopy
,”
Acc. Chem. Res.
53
,
1139
1150
(
2020
).
4.
G.
Gonella
,
E. H. G.
Backus
,
Y.
Nagata
,
D. J.
Bonthuis
,
P.
Loche
,
A.
Schlaich
,
R. R.
Netz
,
A.
Kühnle
,
I. T.
McCrum
,
M. T. M.
Koper
,
M.
Wolf
,
B.
Winter
,
G.
Meijer
,
R. K.
Campen
, and
M.
Bonn
, “
Water at charged interfaces
,”
Nat. Rev. Chem.
5
,
466
485
(
2021
).
5.
R.
Span
and
W.
Wagner
, “
A new equation of state for carbon dioxide covering the fluid region from the triple-point temperature to 1100 K at pressures up to 800 MPa
,”
J. Phys. Chem. Ref. Data
25
,
1509
1596
(
1996
).
6.
Y.
Medina-Gonzalez
,
S.
Camy
, and
J.-S.
Condoret
, “
ScCO2/Green solvents: Biphasic promising systems for cleaner chemicals manufacturing
,”
ACS Sustain. Chem. Eng.
2
,
2623
2636
(
2014
).
7.
B.
Hönisch
and
N. G.
Hemming
, “
Surface ocean pH response to variations in pCO2 through two full glacial cycles
,”
Earth Planet. Sci. Lett.
236
,
305
314
(
2005
).
8.
I.
Klewiah
,
D. S.
Berawala
,
H. C.
Alexander Walker
,
P. Ø.
Andersen
, and
P. H.
Nadeau
, “
Review of experimental sorption studies of CO2 and CH4 in shales
,”
J. Nat. Gas Sci. Eng.
73
,
103045
(
2020
).
9.
D. Y. C.
Leung
,
G.
Caramanna
, and
M. M.
Maroto-Valer
, “
An overview of current status of carbon dioxide capture and storage technologies
,”
Renewable Sustainable Energy Rev.
39
,
426
443
(
2014
).
10.
L. C.
Nielsen
,
I. C.
Bourg
, and
G.
Sposito
, “
Predicting CO2–water interfacial tension under pressure and temperature conditions of geologic CO2 storage
,”
Geochim. Cosmochim. Acta
81
,
28
38
(
2012
).
11.
Y.
Hosseinzadeh Dehaghani
,
M.
Assareh
, and
F.
Feyzi
, “
Simultaneous prediction of equilibrium, interfacial, and transport properties of CO2-brine systems using molecular dynamics simulation: Applications to CO2 storage
,”
Ind. Eng. Chem. Res.
61
,
15390
15406
(
2022
).
12.
B.-S.
Chun
and
G. T.
Wilkinson
, “
Interfacial tension in high-pressure carbon dioxide mixtures
,”
Ind. Eng. Chem. Res.
34
,
4371
4377
(
1995
).
13.
A.
Hebach
,
A.
Oberhof
,
N.
Dahmen
,
A.
Kögel
,
H.
Ederer
, and
E.
Dinjus
, “
Interfacial tension at elevated Pressures Measurements and correlations in the water + carbon dioxide system
,”
J. Chem. Eng. Data
47
,
1540
1546
(
2002
).
14.
F.
Tewes
and
F.
Boury
, “
Thermodynamic and dynamic interfacial properties of binary carbon dioxide–water systems
,”
J. Phys. Chem. B
108
,
2405
2412
(
2004
).
15.
P.
Chiquet
,
J.-L.
Daridon
,
D.
Broseta
, and
S.
Thibeau
, “
CO2/water interfacial tensions under pressure and temperature conditions of CO2 geological storage
,”
Energy Convers. Manage.
48
,
736
744
(
2007
).
16.
S.
Bachu
and
D. B.
Bennion
, “
Interfacial tension between CO2, freshwater, and brine in the range of pressure from (2 to 27) MPa, temperature from (20 to 125) °C, and water salinity from (0 to 334000) mg·L−1
,”
J. Chem. Eng. Data
54
,
765
775
(
2009
).
17.
A.
Georgiadis
,
G.
Maitland
,
J. P. M.
Trusler
, and
A.
Bismarck
, “
Interfacial tension measurements of the (H2O + CO2) system at elevated pressures and temperatures
,”
J. Chem. Eng. Data
55
,
4168
4175
(
2010
).
18.
C. A.
Aggelopoulos
,
M.
Robin
,
E.
Perfetti
, and
O.
Vizika
, “
CO2/CaCl2 solution interfacial tensions under CO2 geological storage conditions: Influence of cation valence on interfacial tension
,”
Adv. Water Resour.
33
,
691
697
(
2010
).
19.
C. A.
Aggelopoulos
,
M.
Robin
, and
O.
Vizika
, “
Interfacial tension between CO2 and brine (NaCl+CaCl2) at elevated pressures and temperatures: The additive effect of different salts
,”
Adv. Water Resour.
34
,
505
511
(
2011
).
20.
P. K.
Bikkina
,
O.
Shoham
, and
R.
Uppaluri
, “
Equilibrated interfacial tension data of the CO2–water system at high pressures and moderate temperatures
,”
J. Chem. Eng. Data
56
,
3725
3733
(
2011
).
21.
Y.
Liu
,
J.
Tang
,
M.
Wang
,
Q.
Wang
,
J.
Tong
,
J.
Zhao
, and
Y.
Song
, “
Measurement of interfacial tension of CO2 and NaCl aqueous solution over wide temperature, pressure, and salinity ranges
,”
J. Chem. Eng. Data
62
,
1036
1046
(
2017
).
22.
Z. R.
Hinton
and
N. J.
Alvarez
, “
Surface tensions at elevated pressure depend strongly on bulk phase saturation
,”
J. Colloid Interface Sci.
594
,
681
689
(
2021
).
23.
X.-S.
Li
,
J.-M.
Liu
, and
D.
Fu
, “
Investigation of interfacial tensions for carbon dioxide aqueous solutions by perturbed-chain statistical associating fluid theory combined with density-gradient theory
,”
Ind. Eng. Chem. Res.
47
,
8911
8917
(
2008
).
24.
G.
Niño-Amézquita
,
D.
van Putten
, and
S.
Enders
, “
Phase equilibrium and interfacial properties of water+CO2 mixtures
,”
Fluid Phase Equilib.
332
,
40
47
(
2012
).
25.
T.
Lafitte
,
B.
Mendiboure
,
M. M.
Piñeiro
,
D.
Bessières
, and
C.
Miqueu
, “
Interfacial properties of water/CO2: A comprehensive description through a gradient Theory–SAFT-VR Mie approach
,”
J. Phys. Chem. B
114
,
11110
11116
(
2010
).
26.
S. R. P.
da Rocha
,
K. P.
Johnston
,
R. E.
Westacott
, and
P. J.
Rossky
, “
Molecular structure of the water–supercritical CO2 Interface
,”
J. Phys. Chem. B
105
,
12092
12104
(
2001
).
27.
H.
Zhang
and
S. J.
Singer
, “
Analysis of the subcritical carbon dioxide–water interface
,”
J. Phys. Chem. A
115
,
6285
6296
(
2011
).
28.
L.
Zhao
,
L.
Tao
, and
S.
Lin
, “
Molecular dynamics characterizations of the supercritical CO2–mediated hexane–brine interface
,”
Ind. Eng. Chem. Res.
54
,
2489
2496
(
2015
).
29.
W.
Li
,
Y.
Nan
,
Z.
Zhang
,
Q.
You
, and
Z.
Jin
, “
Hydrophilicity/hydrophobicity driven CO2 solubility in kaolinite nanopores in relation to carbon sequestration
,”
Chem. Eng. J.
398
,
125449
(
2020
).
30.
M.
Shiga
,
T.
Morishita
, and
M.
Sorai
, “
Interfacial tension of carbon dioxide - water under conditions of CO2 geological storage and enhanced geothermal systems: A molecular dynamics study on the effect of temperature
,”
Fuel
337
,
127219
(
2023
).
31.
O. T.
Unke
and
M.
Meuwly
, “
PhysNet: A neural network for predicting energies, forces, dipole moments, and partial charges
,”
J. Chem. Theory Comput.
15
,
3678
3693
(
2019
).
32.
V. L.
Deringer
,
M. A.
Caro
, and
G.
Csányi
, “
Machine learning interatomic potentials as emerging tools for materials science
,”
Adv. Mater.
31
,
1902765
(
2019
).
33.
V.
Kapil
,
C.
Schran
,
A.
Zen
,
J.
Chen
,
C. J.
Pickard
, and
A.
Michaelides
, “
The first-principles phase diagram of monolayer nanoconfined water
,”
Nature
609
,
512
516
(
2022
).
34.
M.
Yang
,
L.
Bonati
,
D.
Polino
, and
M.
Parrinello
, “
Using metadynamics to build neural network potentials for reactive events: The case of urea decomposition in water
,”
Catal. Today
387
,
143
149
(
2022
).
35.
Y.
Litman
,
J.
Lan
,
Y.
Nagata
, and
D. M.
Wilkins
, “
Fully first-principles surface spectroscopy with machine learning
,”
J. Phys. Chem. Lett.
14
,
8175
8182
(
2023
).
36.
R.
Mathur
,
M. C.
Muniz
,
S.
Yue
,
R.
Car
, and
A. Z.
Panagiotopoulos
, “
First-principles-based machine learning models for phase behavior and transport properties of CO2
,”
J. Phys. Chem. B
127
,
4562
4569
(
2023
).
37.
A.
Omranpour
,
P.
Montero De Hijes
,
J.
Behler
,
C.
Dellago
, and
C.
Dellago
, “
Perspective: Atomistic simulations of water and aqueous systems with machine learning potentials
,” arXiv:2401.17875 (
2024
).
38.
J.
Behler
and
M.
Parrinello
, “
Generalized neural-network representation of high-dimensional potential-energy surfaces
,”
Phys. Rev. Lett.
98
,
146401
(
2007
).
39.
C.
Schran
,
K.
Brezina
, and
O.
Marsalek
, “
Committee neural network potentials control generalization errors and enable active learning
,”
J. Chem. Phys.
153
,
104105
(
2020
).
40.
C.
Schran
,
F. L.
Thiemann
,
P.
Rowe
,
E. A.
Müller
,
O.
Marsalek
, and
A.
Michaelides
, “
Machine learning potentials for complex aqueous systems made simple
,”
Proc. Natl. Acad. Sci. U. S. A.
118
,
e2110077118
(
2021
).
41.
T. D.
Kühne
,
M.
Iannuzzi
,
M.
Del Ben
,
V. V.
Rybkin
,
P.
Seewald
,
F.
Stein
,
T.
Laino
,
R. Z.
Khaliullin
,
O.
Schütt
,
F.
Schiffmann
,
D.
Golze
,
J.
Wilhelm
,
S.
Chulkov
,
M. H.
Bani-Hashemian
,
V.
Weber
,
U.
Borštnik
,
M.
Taillefumier
,
A. S.
Jakobovits
,
A.
Lazzaro
,
H.
Pabst
,
T.
Müller
,
R.
Schade
,
M.
Guidon
,
S.
Andermatt
,
N.
Holmberg
,
G. K.
Schenter
,
A.
Hehn
,
A.
Bussy
,
F.
Belleflamme
,
G.
Tabacchi
,
A.
Glöß
,
M.
Lass
,
I.
Bethune
,
C. J.
Mundy
,
C.
Plessl
,
M.
Watkins
,
J.
VandeVondele
,
M.
Krack
, and
J.
Hutter
, “
CP2K: An electronic structure and molecular dynamics software package - Quickstep: Efficient and accurate electronic structure calculations
,”
J. Chem. Phys.
152
,
194103
(
2020
).
42.
A. D.
Becke
, “
Density-functional exchange-energy approximation with correct asymptotic behavior
,”
Phys. Rev. A
38
,
3098
3100
(
1988
).
43.
C.
Lee
,
W.
Yang
, and
R. G.
Parr
, “
Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density
,”
Phys. Rev. B
37
,
785
789
(
1988
).
44.
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
,
154104
(
2010
).
45.
M. J.
Gillan
,
D.
Alfè
, and
A.
Michaelides
, “
Perspective: How good is DFT for water?
,”
J. Chem. Phys.
144
,
130901
(
2016
).
46.
H.
Goel
,
Z. W.
Windom
,
A. A.
Jackson
, and
N.
Rai
, “
Performance of density functionals for modeling vapor liquid equilibria of CO2 and SO2
,”
J. Comput. Chem.
39
,
397
406
(
2018
).
47.
Y.
Wu
,
H. L.
Tepper
, and
G. A.
Voth
, “
Flexible simple point-charge water model with improved liquid-state properties
,”
J. Chem. Phys.
124
,
24503
(
2006
).
48.
A.
Zhu
,
X.
Zhang
,
Q.
Liu
, and
Q.
Zhang
, “
A fully flexible potential model for carbon dioxide
,”
Chin. J. Chem. Eng.
17
,
268
272
(
2009
).
49.
A.
Singraber
,
T.
Morawietz
,
J.
Behler
, and
C.
Dellago
, “
Parallel multistream training of high-dimensional neural network potentials
,”
J. Chem. Theory Comput.
15
,
3075
3092
(
2019
).
50.
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
,
8368
8373
(
2016
).
51.
J. M.
Míguez
,
M. M.
Piñeiro
, and
F. J.
Blas
, “
Influence of the long-range corrections on the interfacial properties of molecular models using Monte Carlo simulation
,”
J. Chem. Phys.
138
,
34707
(
2013
).
52.
J.
Janeček
, “
Effect of the interfacial area on the equilibrium properties of Lennard-Jones fluid
,”
J. Chem. Phys.
131
,
124513
(
2009
).
53.
F. G. J.
Longford
,
J. W.
Essex
,
C.-K.
Skylaris
, and
J. G.
Frey
, “
Unexpected finite size effects in interfacial systems: Why bigger is not always better—Increase in uncertainty of surface tension with bulk phase width
,”
J. Chem. Phys.
148
,
214704
(
2018
).
54.
V.
Kapil
,
M.
Rossi
,
O.
Marsalek
,
R.
Petraglia
,
Y.
Litman
,
T.
Spura
,
B.
Cheng
,
A.
Cuzzocrea
,
R. H.
Meißner
,
D. M.
Wilkins
,
B. A.
Helfrecht
,
P.
Juda
,
S. P.
Bienvenue
,
W.
Fang
,
J.
Kessler
,
I.
Poltavsky
,
S.
Vandenbrande
,
J.
Wieme
,
C.
Corminboeuf
,
T. D.
Kühne
,
D. E.
Manolopoulos
,
T. E.
Markland
,
J. O.
Richardson
,
A.
Tkatchenko
,
G. A.
Tribello
,
V.
Van Speybroeck
, and
M.
Ceriotti
, “
i-PI 2.0: A universal force engine for advanced molecular simulations
,”
Comput. Phys. Commun.
236
,
214
223
(
2019
).
55.
S.
Plimpton
, “
Fast parallel algorithms for short-range molecular dynamics
,”
J. Comput. Phys.
117
,
1
19
(
1995
).
56.
A.
Singraber
,
J.
Behler
, and
C.
Dellago
, “
Library-based LAMMPS implementation of high-dimensional neural network potentials
,”
J. Chem. Theory Comput.
15
,
1827
1840
(
2019
).
57.
J. G.
Kirkwood
and
F. P.
Buff
, “
The statistical mechanical theory of solutions. I
,”
J. Chem. Phys.
19
,
774
777
(
1951
).
58.
J.
Delhommelle
and
P.
Millié
, “
Inadequacy of the Lorentz-Berthelot combining rules for accurate predictions of equilibrium properties by molecular simulation
,”
Mol. Phys.
99
,
619
625
(
2001
).
59.
D.
Boda
and
D.
Henderson
, “
The effects of deviations from Lorentz–Berthelot rules on the properties of a simple mixture
,”
Mol. Phys.
106
,
2367
2370
(
2008
).
60.
M.
Rouha
and
I.
Nezbeda
, “
Non-Lorentz–Berthelot Lennard-Jones mixtures: A systematic study
,”
Fluid Phase Equilib.
277
,
42
48
(
2009
).
61.
A. P.
Willard
and
D.
Chandler
, “
Instantaneous liquid interfaces
,”
J. Phys. Chem. B
114
,
1954
1958
(
2010
).
62.
T.
Morishita
and
M.
Shiga
, “
Ab initio characterization of the CO2–water interface using unsupervised machine learning for dimensionality reduction
,”
J. Phys. Chem. B
128
,
5781
5791
(
2024
).
63.
B. S.
Pattni
,
V. V.
Chupin
, and
V. P.
Torchilin
, “
New developments in liposomal drug delivery
,”
Chem. Rev.
115
,
10938
10966
(
2015
).
64.
S.
Dasgupta
,
T.
Auth
, and
G.
Gompper
, “
Nano- and microparticles at fluid and biological interfaces
,”
J. Phys. Condens. Matter
29
,
373003
(
2017
).
65.
S.
Sokołowski
and
O.
Pizio
, “
Density functional theory for the microscopic structure of nanoparticles at the liquid–liquid interface
,”
Phys. Chem. Chem. Phys.
21
,
3073
3082
(
2019
).
66.
F. M.
Coelho
,
L. F. M.
Franco
, and
A.
Firoozabadi
, “
Thermodiffusion of CO2 in water by nonequilibrium molecular dynamics simulations
,”
J. Phys. Chem. B
127
,
2749
2760
(
2023
).
67.
F. M.
Coelho
,
L. F. M.
Franco
, and
A.
Firoozabadi
, “
Effect of salinity on CO2 thermodiffusion in aqueous mixtures by molecular dynamics simulations
,”
ACS Sustain. Chem. Eng.
11
,
17086
17097
(
2023
).
Published open access through an agreement with JISC Collections