<<

Bridging the gap between mesoscopic and molecular models of / interfaces out-of-equilibrium Aaron R. Finney1 and Matteo Salvalaglio∗1 Thomas Young Centre and Department of Chemical Engineering - University College London - London WC1E 7JE UK. (Dated: 7 September 2021) Solid/liquid interfaces control various processes of technological relevance in the process industry and many fundamental physicochemical phenomena. This work examines the link between the atomistic description of transfer at solid/liquid interfaces out of equilibrium and the constitutive mass transfer equations typically used to model these processes at the mesoscale. In our analysis we discuss the microscopic inconsistencies apparent in simplified models of mass transfer whenever non-idealities dominate the liquid phase in the proximity of solid/liquid interfaces. Using CµMD - a molecular simulation technique to investigate out-of- equilibrium, -driven processes with pseudo open-boundary conditions - we outline a strategy to capture and quantify non-idealities induced by specific interactions between the solid surface and in the fluid phase. We demonstrate our approach by studying electrolyte in technologically important, multi-component systems in which introducing a solid/liquid interface induces substantial deviations in the composition and electrochemical properties compared to the fluid phase bulk. To this aim, we analyse NaCl(aq) solutions in contact with i) the graphite basal surface and ii) {100}) NaCl(s) crystalline surfaces. We uncover the tendency of the sodium cation to preferentially adsorb at the surfaces considered, which induces local violations of electroneutrality that leads to the emergence of an electric potential in the fluid phase at the solid/liquid interface. Keywords: Molecular Simulations, Interfaces, Crystal Growth, Ion Adsorption, Electrolytes

I. INTRODUCTION time-scales for molecular processes which are often diffi- cult to probe in experiments. In particular, MD simula- tions are ideally suited to describe equilibrium conditions Solid/liquid interfaces play a central role in a range in dense systems such as single and multi-component of processes including catalysis,1 crystal growth2 and .19 Technological applications, however, often ex- filtration3. In order to control these processes for e.g., in- ploit out-of-equilibrium, concentration-driven processes novative materials design, it is often necessary to obtain to achieve their function. Crystal precipitation repre- a description of the interface at the molecular scale. For sents once again a particularly relevant example in this example, the selectivity and efficiency of advanced mem- context. Crystallization processes in fact operate out-of- brane materials is determined by factors that are directly equilibrium, and are governed by the kinetics of funda- linked to the molecular structure of the interface between mental steps such as nucleation and growth.13,21,22 Ide- the stationary and mobile phases.4,5 Similarly, the opera- ally, MD simulations aimed at understanding the con- tion of heterogeneous catalysts is determined by the envi- trol of solid/liquid interfaces on these steps need to be ronment of the active site at the atomistic level.6 Even in able to capture the evolution of an out-of-equilibrium sys- the ideal case of a perfectly planar surface in contact with tem under well-defined conditions of temperature, pres- a liquid, intermolecular interactions lead to highly non- sure and supersaturation. When dealing with out-of- ideal regions of the fluid phase.7 Measuring the proper- equilibrium, concentration-driven molecular simulations ties and structure of this interfacial region is challenging of systems containing more than one phase, however, the and often requires specialist experimental techniques8–12 accessible system sizes associated with MD (due to fi- This is compounded in the case of changing or fluctuat- nite computational resources) can limit the scope of such ing surfaces. Crystal growth is a paradigmatic example of techniques to tackle the problems of interest, and often arXiv:2109.00568v1 [cond-mat.soft] 1 Sep 2021 such a macroscopic process emerging from the collective requires the development of dedicated theoretical descrip- behavior of a large number of molecules and the evolu- tions and ad-hoc corrections.23–29 tion of a solid/liquid interface. Crucially, crystallization is affected by mechanistic, thermodynamic and kinetic The solid/liquid interface in macroscopic systems is in details regarding the fundamental physicochemical steps 13–16 contact with a bulk liquid phase that changes its prop- occurring at the scale of single growth units. erties on timescales that are significantly longer than Given the challenges alluded to above, it is common- those typically associated with molecular-scale phenom- place in many fields to employ molecular simulation tech- ena. For example, in the case of multi-component solu- niques, often based on Molecular Dynamics (MD), to gain tions, bulk solute can be considered un- atomistic-level insight into the structure and dynamics of changing on the timescales required for solute monomer collective phenomena at the solid/liquid interface.17–20 incorporation into a crystal growth site at the sur- MD simulations provide direct access to the length- and face. A faithful representation of these types of pro- 2 cesses in MD requires an ability to model systems out- of-equilibrium as a function of constant bulk properties. Achieving this is far from trivial, because it usually requires simulations with open boundaries to al- low exchange of species at the solid/liquid interface with the ‘surrounding bulk’ to maintain steady-state out-of- equilibrium conditions. Suitable techniques to perform these kinds of simulations include hybrid grand-canonical Monte Carlo/MD, in which there is a dynamic exchange of molecular species between the simulation cell and an external, virtual reservoir defined by a constant chem- ical potential.30–35 Despite their elegant formalism, hy- brid MC/MD approaches are extremely inefficient when dealing with dense fluids, as the probability of acceptance of exchange moves between a in the fluid and the external reservoir becomes very small. An alternative ap- proach, known as Constant Chemical Potential Molecular Dynamics (CµMD), was recently introduced to overcome these limitations that mimics open-boundary, constant composition, out-of-equilibrium conditions.36–42 CµMD, FIG. 1. Conceptual model of mass transfer in close proximity 43–46 to an interacting, non-exchanging surface. Time-dependent similarly to adaptive resolution simulation methods , solution of the microscopic material balance for a binary liq- introduces an internal reservoir of molecules in the fluid uid in contact with an interacting surface in a mono- phase, and applies ad-hoc forces to regulate the flux dimensional domain. Left. Molar fraction profile of the so- to/from the reservoir and the ‘bulk’ solution in contact lute (A) as a function of time. At a steady state, the so- with the interface. This technique has already proven to lute accumulates at the interface according to the potential be very effective to understand the molecular-scale pro- of mean force for adsorption at the interface. Right. The cesses associated with crystallization, surface adsorption driving force ∇µA becomes null at a steady state, where the and permeation, and its application is not limited by the µA profile is flat and the flux is null. These profiles were ob- of the fluid phase.7,16,36,42 tained by solving the time-dependent diffusion equation, with initial conditions xA(z, 0) = 0.15, and boundary conditions ∂µA xA(0.5, t) = 0.15 and ∂z |z=0,t= 0.

Adopting CµMD to gain atomistic-level insight into the solid/liquid interface allows for a direct comparison A. Microscopic inconsistencies in models of transport at of the structure and dynamics of interfacial species with interfaces: Film Theory the predictions of mean field models (based on constitu- tive equations) to describe mass transfer at mesoscopic To highlight the need for a reconciliation between and macroscopic length scales. In order to achieve this meso- and molecular-scale representations of transport goal, a fully consistent analysis and interpretation of processes at interfaces, we start by analysing Film The- atomistic-level information in the context of constitutive ory (FT). FT is a typical simplification of the mass trans- mass-transfer equations is necessary. fer constitutive equations applied to solid/liquid inter- faces that leads to an inconsistent description of the com- position field in the liquid phase in the presence of strong interactions between the solid and fluid phases. The mass The remainder of this paper is organised as follows: transfer flux at the solid/liquid interface in FT is ex- firstly, using conceptual examples we illustrate how ac- pressed as13 counting for microscopic interactions between the surface s b and the fluid phase is essential to capture the complex- Ji = −k ci − ci (1) ity of concentration fields at the nanometre scale with constitutive equations; secondly, we employ constitutive where Ji is the molar flux of species i, k is a mass transfer s b equations to interpret the concentration profiles obtained coefficient, and ci and ci are the concentrations of species studying the adsorption of NaCl(aq) at graphite and i at the solid surface and in the bulk of the fluid phase, NaCl(s), and extract the ion-specific excess free energy of respectively. The above equation is predicated on two interaction with the solid surface. The non-ideality of the key assumptions: i) the concentration gradient is linear solution structure stemming from asymmetric adsorption within a fluid film of thickness δ in contact with the sur- of species with different charge is responsible for local face; and, ii) the driving force for the diffusive mass trans- violations of electroneutrality in solution, and is key to fer of i is a gradient in concentration. FT is grounded develop consistent mesoscopic models of the solid/liquid in Fick’s diffusion equation, and the mass transfer co- interface. efficient is defined as the ratio between the Fick diffu- 3 sion coefficient, Di, and δ, which can be obtained from molar fraction, xA(r), and its activity coefficient, γA(r) adimensional number correlations.13 FT provides a sim- as: ple framework to express diffusion dominated flowrates 0 across interfaces as a function of the concentrations de- µA(r) = µA + RT ln xA(r) + RT ln γA(r) (3) scribed above where cs is typically considered equal to the i 0 equilibrium concentration. While this model is able to where µA is a reference chemical potential under stan- capture the qualitative behavior of solute fluxes in crys- dard conditions, and RT ln xA(r) is the free energy of tal growth processes, it demonstrates fundamental incon- an ideal mixture of composition xA (R and T being the sistencies when compared with results from simulations ideal constant and temperature, respectively). The E and experiments able to resolve molecular-level length term RT ln γA(r) defines gA (r), i.e. the partial molar ex- and time scales. Attractive and repulsive intermolecular cess free energy of A in B, at distance r from the 47 interactions between the surface and solute species can interacting surface . In contrast with the excess free en- lead to concentrations at the interface that deviate from ergy of component A in the bulk of a real fluid mixture, e ci q, thus introducing inconsistencies between the sign of which is a function of the composition only (in the ab- s b the term ci − ci and the direction of the net diffusive sence of an external field), the excess free energy term in flux at the interface. Indeed, the concentration profile the presence of an interface accounts for the asymmetric postulated by FT is verified only in the case of a weakly intermolecular forces between the species in solution and interacting surface, that does not induce significant so- the solid surface, and thus depends on the distance from lute adsorption at the solid/liquid interface. the interface. Equation 2 can be rewritten as:

∗ h xA E i JA = −cÐA,B ∇xA + ∇gA (r) (4) II. A CONSISTENT DESCRIPTION ACROSS SCALES: RT MAXWELL-STEFAN DIFFUSION. E ∇gA (r) can be interpreted as the mean thermodynamic force exerted by the solid surface on species A, which is A fully consistent description of diffusive mass transfer responsible for the emergence of features in the local com- at attractive/repulsive interfaces requires a framework in position field in the vicinity of a solid surface. In the case which the chemical potential gradient is explicitly con- of an inert surface, the excess chemical potential of A in E sidered as the driving force for the flux of species. Such a the fluid becomes independent from r, i.e. ∇gA (r) = 0 framework is provided by the generalised Maxwell-Stefan and there is no accumulation/depletion of solute in the (MS) diffusion equation. In this section, we progressively fluid phase near the interface. In such conditions, Fick   build the necessary complexity within the MS diffusion ∂ln γA diffusivity is recovered as DA,B = ÐA,B 1 + , equation to consistently interpret the effect of a strongly ∂ ln xA interacting surface on the multi-component, electrolyte and simplified approaches like FT will provide yield a solution modelled via fully-atomistic CµMD simulations. coherent representation of the concentration field. Fur- thermore, in the case of an ideal mixture (i.e. γA = 1), DA,B = ÐA,B. A. Interacting, non-exchanging surface. When the concentration field in the presence of a het- ∗ erogeneous interacting surface is stationary and JA = 0, ∇gE(r) We begin by considering the adsorption of an electrically the excess thermodynamic force A can be computed neutral solute A in solvent B interacting with a solid from the expression: surface unable to exchange mass with the fluid phase. RT ∇gE(r) = − ∇xss(r) (5) The MS equation for the diffusive flux of A in this system A xss(r) A is47–49 A xss ∗ ÐA,B where A is the steady state molar fraction field. Equa- J = −cA ∇µA (2) E ss A RT tion 5 is equivalent to gA (r) = −RT ln xA + C, where C is an arbitrary integration constant. The excess free ∗ ∗ ∗ E where JA = xA(vA − v ), with vA and v being the ve- energy term gA (r) can thus be computed directly from locity of component A and the molar weighted average molecular simulations by sampling the steady-state com- velocity, respectively. cA in the above equation is the position field of species A as a function of distance from of A, while ÐA,B is the MS diffusion the solid/liquid interface. As shown in Fig.1, solving the E coefficient, and ∇µA is the chemical potential gradient of microscopic material balance accounting for ∇gA (r) 6= 0 A. in a mono-dimensional domain leads to a microscopically The species-specific intermolecular forces established consistent concentration field. In the case of an attractive between the surface and each component of the fluid surface, from an initial homogeneous concentration of A, phase can be introduced into the MS diffusion equations accumulation of solute at the surface occurs at equilib- through ∇µA by explicitly considering the dependence rium. The steady state solution of this mass balance pro- of µA on the distance from an interacting surface. The vides the correct equilibrium conditions, i.e. ∇µA = 0, ss chemical potential of A can be written as a function of its while ∇xA (r) 6= 0. 4

B. Interacting, exchanging surface A. NaCl(aq)/graphite interfaces.

Let us now consider the example of a solid surface ex- The concentration field obtained in atomistic simu- changing molecules of solute (A) with a two-component lations of a graphite-electrolyte solution interface (see solution constituted by the electrically neutral solute A Fig.3 A) is a result of the combined effect of differing and solvent B. This example applies to crystal/solution affinities for ionic species to accumulate at the interface, interfaces out-of-equilibrium in which solute molecules and of the emergent electric fields in this region generated by local violations of electroneutrality associated with (A, in this example) are either released by, or incorpo- 7 rated into the solid surface. the asymmetric adsorption. The system considered in this section is represented in Fig.3 A. As such, additional In the current case, the diffusive flux of A in proxim- driving forces, such as an electric field (which is equal ity to a crystal surface is described by the MS diffusion to the gradient of the electric potential field, ∇φ), can equation reported in Equation 4 discussed in the context be introduced into the MS diffusion equation to describe of an interacting surface. However, in order to capture the global diffusive flux of cations and anions.50 The MS the flux of A across the solid/liquid phase boundary, the equations for Na+ and Cl− in solution, respectively read: microscopic mass balance should reflect the fact that, at constant temperature and pressure, the chemical poten- xtal NNaxCl − NClxNa NNaxW − NW xNa tial of a crystal comprised of A monomers, µA (T,P ), is + constant. One can therefore express the boundary condi- cÐNaCl cÐNaW (8) E tions in terms of chemical potentials of the solid and liq- ∇gNa ρNazNaF = −∇xNa − xNa − ∇φ uid species. Considering that the diffusion driving forces RT cRT MNa E include the contribution of gA (r), as introduced in Equa- tion 3, the microscopic material balance reads: and:

NClxNa − NNaxCl NClxW − NW xCl ∂xA 1  E  + = ∇ xAÐA,B∇ RT ln xA + g (r) (6) ∂t RT A cÐNaCl cÐClW (9) E ∇gCl ρClzClF = −∇xCl − xCl − ∇φ with the boundary condition at the solid/liquid interface: RT cRT MCl

where Ni, zi, ρi and Mi indicate the molar flux, density, xtal ∗ ∗ 50 µ|zint = µ (T,P ) = RT ln(xAγA) (7) valency and of species i, respectively. It should be noted that in the case of an uncharged surface, ∗ ∗ ∇φ where xA and γA are the molar fraction and activity co- the terms dependent upon on the right-hand side of efficient of component A at equilibrium in the bulk solu- Equations 8 and 9 emerge solely due to a non-zero net tion. As shown in Fig. 2, for a weakly interacting surface charge distribution in the fluid phase close to the surface; E φ where the contribution of ∇gA (r) to the mass transfer indeed is computed as: driving force is small compared to the contribution of 0 0 0 Z z "Z z Z z # ∇xss(r), the composition field remains qualitatively con- F c A φ(z) = − xNadz − xCldz dz (10) sistent with the predictions of simple models for mass 0 0r 0 0 transfer at interfaces. However, when the contribution of E ∇gA (r) becomes significant (see Fig. 2, strongly interact- where F is Faraday’s constant, c is the bulk concentra- ing surface) the composition profile shows the emergence tions of ions, and 0 and r are constants for the vacuum of a local solution structure with solute accumulation at and relative permittivity of the solution medium. the interface, incompatible with models such as FT. The source of the asymmetric adsorption of ions of op- posite electrical charge is the different potentials of mean force that characterise the surface-solution interactions, and which are captured by the excess free energy term E III. EXCESS SURFACE FREE ENERGY FROM CµMD gi (r) for ion i. We can exploit the steady state condi- SIMULATIONS tions obtained from open boundary MD simulations to extract the mean force terms Wi(r), which balances the combined effect of the electric field and that of the chem- Having introduced a framework to interpret complex ical potential of the , leading to a constant composition profiles, in this section we extend our de- electrochemical potential at equilibrium: scription to analyse results from CµMD simulations of interacting surfaces in contact with NaCl(aq) electrolyte E RT ρiziF ∇gi = − ∇xi − ∇φ (11) solutions with varying solute concentrations. We con- xi ci Mi sider first the case of a non-exchanging surface (graphite), and then a surface exchanging ions with the solution As shown in Fig. 3 C and D, the molar fraction profile of ({100} NaCl(s)). both cation and anion can be extracted from simulations, 5

FIG. 2. Conceptual model of mass transfer in the proximity of an out-of-equilibrium, interacting and exchanging surface provided by the time-dependent solution of the microscopic material balance for a binary liquid mixture in a mono-dimensional domain. Top: weakly interacting surface Molar fraction and chemical potential profiles of the solute (A) as a function of time. In this case the gradient of concentration and chemical potential have the same sign, and the use of the simplified description of mass transfer from Film Theory (FT) provides a reasonable description of the microscopic concentration profile. Bottom: strongly interacting surface In the case of a strongly interacting surface, the microscopic description of concentration and chemical potential in the interfacial region is inconsistent with the descriptions from FT. For example, the concentration difference between interface and bulk has the opposite sign of the chemical potential difference, and simplified models cannot capture a microscopically consistent composition field. and the electrostatic potential computed using Eq. 10, as second and third solution layers beyond the surface (see shown in Fig. 3 B. The profile of the ion-specific excess Fig. 3 E) deepen. E free energy of adsorption gi /RT = ln γi(r), reported in Fig. 3 E and F is thus computed using Eq. 11. The gE/RT i profiles overall capture the asymmetric driving B. NaCl(aq)/NaCl(s) interfaces. forces for ion adsorption at graphite surfaces responsi- ble for non-ideal character of the electrical double layer in these systems7. The excess free energy of adsorption Finally, we analyse out-of-equilibrium NaCl(s) {100} of the cation has a minimum in direct contact with the crystalline interfaces in contact with a bulk solution at graphite surface, while the anion excess free energy has constant composition. When the system evolves under a minimum that corresponds to an adjacent layer. diffusion-limited growth, the total flux of ions to and from the crystal is determined by the gradient in chem- We define the excess free energy of adsorption as, ical potential. In this case, ions at the interface are not E E E ∆gi = min gi (z) − gi (zb) for ionic species i, where zb restricted to the liquid phase and can be incorporated E E is a position in the bulk. ∆gNa and ∆gCl exhibit a dif- into the evolving crystalline surface. Their chemical po- ferent dependence on the value of bulk concentration as tential displays a discontinuity at the boundary between E shown in Fig. 3 G. While ∆gNa becomes less negative the solid and liquid phases. Together with the chemi- E with increasing bulk concentrations, ∆gCl remains fairly cal potential, the system’s composition, density and local independent from bulk concentration. The free energy structure also display discontinuities at the solid/liquid change for the adsorption of single cations at the car- boundary. bon surface is greater than for chloride ions;51 hence, the When the solution bulk is supersaturated, i.e., aq xtal result shown in Fig. 3 G for the lowest concentration. µNaCl > µNaCl (where superscript aq and xtl indi- At high bulk concentrations, however, crowding of solute cate the aqueaous and crystal chemical potential, respec- species at the interface leads to significant ordering of tively), the solid/liquid interface will act as a sink for the ions, particularly so for cations in the first solution the removal of ions from the liquid phase at a rate deter- layer. The relative accumulation of ions in this region is mined by the diffusion of ions. Conversely, in understau- aq xtal reduced relative to the bulk levels as concentration in- rated conditions, where µNaCl < µNaCl, the solid/liquid creases, and minima for the adsorption of cations in the interface supplies ions to the bulk solution. In order to 6

FIG. 3. CµMD Simulations of graphite interfaces in contact with NaCl aqueous solutions highlighting the asymmetric adsorption of ions (∼1-10M). A) A representation of the simulation box setup, displaying the ion reservoir in green, and the position at which CµMD forces are applied as a dashed line. B) Electric potential Φ(z) generated by the local ion charge fluctuations in solution. C-D) Na+ and Cl− molar fraction profiles. E-F) Excess chemical potential associated with the asymmetric interaction of Na+ (E) and Cl− (F) ions with the graphite surface.G) Excess adsorption free energy for Na+ (circles), and Cl− (squares) as a function of the concentration in the solution bulk cb. The color code corresponds to cb and is consistent with those in panels C-F. investigate the role of surface structure on the ion-specific to local violations of the solution electroneutrality per- excess free energy, we carry out simulations on both per- pendicular to the surface and a gradient in the electric fectly planar (smooth) and rough surfaces where at least potential (Fig. 4 B). three crystal terraces are exposed to solution initially (see At smooth NaCl(s) surfaces the cation excess is read- Methods). As in the case of graphite, CµMD simulations ily compensated by an anion excess in the immediately reveal a solution structure heavily affected by specific adjacent solution layer. In this limit, even if the accu- surface-solute interactions. For both smooth and rough mulation of ions from the solution is asymmetric, the surfaces, the solution composition profiles display a lay- strength of the adsorption interaction of both ions with ered solute structure at the interface, similar to the one the surface is comparable. This is quantitatively reflected observed in Fig. 3. Also in this case, the cations in solu- in the adsorption excess free energy (Fig. 4 E-G) that is tion tend to accumulate in the first adsorbed solute layer essentially the same for both ions and exhibits the same in contact with the surface (see Fig. 4 C and D), leading dependence on the bulk solute concentration, with in- 7

FIG. 4. CµMD Simulations of NaCl(s) {100} surfaces in contact with NaCl aqueous solutions ( 1-7M). A) A representation of the simulation box setup, displaying the ion reservoir in green, and the position at which CµMD forces are applied as a dashed line. B) Electric potential Φ(z) generated by the local charge fluctuations due to the asymmetric adsorption of ions at the interface. C-D) Na+ and Cl− molar fractions profiles, displaying an asymmetric behaviour in proximity to the system surface. E-F) Excess chemical potential associated with the asymmetric interaction of Na+ (E) and Cl− (F) ions with the graphite surface.G) Excess adsorption free energy for Na+ (circles), and Cl− (squares) as a function of the concentration in the solution bulk cb. The color code corresponds to cb and is consistent with those in panels C-F. Full symbols refer to smooth NaCl(s), while empty symbols correspond to values of the adsorption excess free energy for rough surfaces. creasing bulk ion concentrations leading to a less negative ditional surface area and exposed step edges lead to in- ∆gE. A more complex behaviour emerges in the case of creased ordering of ions in first few solution layers, cf. the roughened {100} NaCl crystal surfaces. At the rough sur- smooth NaCl surface, with fluctuations to the screening face the outermost crystal layer can exchange ions with of the charge associated with a preferentially adsorbed the liquid phase much more readily on the timescale of first sodium layer. This observation is consistent with the the simulations than occurs at perfectly planar crystal excess adsorption free energy trend as a function of bulk surfaces of NaCl. Moreover, under-coordinated surface concentration on rough surfaces (see Fig. 4 B). While ions tend to develop stronger interactions with molecules the excess adsorption free energy of the anion on rough in the fluid phase than ions in the ‘bulk’ of the crys- surfaces is only marginally lower than that of both the E tal. The combined effect is a significant reorganisation of anion and cation on smooth surfaces, ∆gNa is signifi- both the liquid and the solid at the interface. The ad- cantly lower, and does not show a clear dependence on 8 the bulk ionic concentration. This result suggests that complexity of the model. interactions with under-coordinated sites at the interface is an important contributing factor towards ion accumu- lation. IV. CONCLUSIONS

In this work we demonstrated how molecular-level in- C. Outlook: linking dynamics across the scales formation obtained from CµMD56 simulations of solid- liquid interfaces can be interpreted consistently and In the above discussion we have shown how Maxwell- quantitatively using the Maxwell-Stefan theory of diffu- Stefan diffusion theory can be applied to extract quan- sive mass transfer. This enables a characterisation of titative information from both equilibrium and out-of- non-idealities in two fundamental processes at the heart equilibrium molecular simulations of interfaces subject of a wide range of chemical technologies, namely: to a constant background concentration, at steady state. E • The equilibration of a fluid phase in contact with a gi (r) - the excess free energy of interaction between non-exchanging, interacting surface (Sec.III A). species i and an interface located at distance r - is a key component responsible for the emergence of local de- • The steady-state evolution of an out-of-eqilibrium, viations from ideal solution behavior. With knowledge interacting surface able to exchange molecules with E of gi (r), one can inform the development of fully consis- a fluid phase under the effect of a constant driving tent mean field models based on MS theory to capture force (Sec. III B) also the evolution of the liquid phase composition close to The former process drives adsorption-based technologies; an interface. The information needed to achieve this goal the latter is fundamental to understand and control crys- is the set of N(N − 1)/2 MS diffusion coefficients, where tallization. The fully constitutive approach is based on N is the number of species in the system. As discussed the calculation of an excess free energy profile as a func- 52 in detail in the comprehensive review of Liu et al. the tion of distance from a surface, which is otherwise ex- MS diffusion coefficients Ði,j can be computed using a tremely challenging to compute using alternative simula- variety of different simulation approaches. In particular tion approaches. This allows us to demonstrate how, in we highlight that the application of the multi-component the presence of strong interactions between solid- and 48,53–55 Darken equation allows for a straightforward cal- fluid-phase species, deviations in the spatially-varying culation of Ði,j from the self diffusion coefficient Di,self chemical potential and concentration fields differ from and from the component molar fraction xi: the predictions of simplified models of mass transport

N at interfaces. A consistent mean-field model, like the X xi one presented here, is key to efficiently transfer informa- Ði,j = Di,self Dj,self (12) D tion across the scales and investigate out-of-equilibrium, i=1 i,self concentration-driven processes such as crystallization in In the context of this study we have performed dedicated complex surface geomteries at the nanometer scale over bulk simulations to obtain the self-diffusion coefficients a wide range of bulk solute concentrations. of Na+, Cl−, and SPC/E water (see the Methods sec- Our analysis, applied to an extensive set of NaCl(aq) tion). Nevertheless, we highlight that calculation of the simulations in contact with graphite (Sec. III A) and self-diffusion coefficients can be performed on the same NaCl crystal surfaces (Sec. III B), reveals a ubiquitous trajectory obtained with CµMD, as done in Ref.7 to as- asymmetric adsorption behavior. In particular, we con- sess the ions’ mobility as a function of distance from the sistently observe the preferential adsorption of the cation, solid/liquid interface. Since in this work we are consid- driven by a more favourable excess adsorption free en- ering mass transfer in the vicinity of interacting surfaces ergy both on graphite as well as on NaCl(s) surfaces. that exert an asymmetric effect on different species, thus The asymmetric adsorption behaviour leads to local vio- leading to local violations of typical constraints such as lations of the solution electroneutrality and to the emer- electroneutrality (Fig. 3 and 4B), in the application of gence of an electric potential in the fluid phase in contact the multicomponent Darken equation we need to explic- with the crystal surface analogous to the double layer po- itly account for three independent molar fractions. The tential that NaCl(aq) develops in contact with graphite MS diffusivities Ði,j computed as function of three in- electrodes. dependent molar fractions subject to the stoichiometric constraint, are reported in the triangular plots repre- V. METHODS sented in Fig. 5. In particular, we note how local de- viations from the ideal solution behavior observed close to the surface may lead to fluctuations by up to a factor A. CµMD simulations. two in the MS diffusivities. Depending on the level of detail required for the solution of a mean field model in- The simulations herein were performed using the formed by MD simulations, this aspect may require con- GROMACS 2018.6 package57 and the PLUMED plu- sideration as it will likely impact on the computational gin (version 2.5)58. Unless otherwise stated, CµMD 9

FIG. 5. Triangle plots representing variations in the MS diffusion coefficients computed using the Darken equation (Eq. 12) for NaCl(aq). Since close to the S/L interface electroneutrality is locally violated, and molar fractions of sodium and chloride ions should be considered independent variables. simulations56 were carried out using a leapfrog time in- performed for at least 100 ns with target ion number den- tegration algorithm with a 2 fs timestep. Simulations sities in the range nt = 0.6022–6.022, corresponding to t t 24 were performed in the NVT ensemble at 298 K; the ion molar concentrations, c = n ×10 /NA (where NA is temperature was held constant within statistical fluctu- Avogadro’s constant), of 1–10 M (mol dm−3) in the case ations using the Bussi-Donadio-Parrinello thermostat.59 of simulations including graphite and ct = 1, 3, 5 and 7 M Periodic boundaries were imposed in three- to otherwise. the orthorhomibic simulation cells shown in e.g., FIG. 2 A. Electrostatic interactions were treated using Parti- 60 cle Mesh Ewald summation with a cut-off applied to B. Preparation of initial configurations. the calculation of the real-space contribution from within 0.9 nm. A truncation was also applied to the 62 calculation of pairwise Van der Waal interactions which A 5.4 × 5.5 × 2.7 nm (x × y × z) graphite supercell was was limited to those within 0.9 nm, and with a potential constructed; this comprised eight graphene layers and shift applied to minimise the fluctuations in the forces was positioned in the centre of an orthorhombic simu- around the cut-off distance. The LINCS algorithm61 was lation cell such that the c-axis was parallel to z. The adopted to constrain the structure of water molecules. graphite was placed in contact with an In the CµMD algorithm adopted in this work, a contin- containing 1,672 NaCl and 13,819 water molecules (for uous force, F µ, is applied to ions in the vicinity of a fixed a total NaCl of 6.7 mol/kg). With restraints distance from a reference point in the simulation cell, so applied to the position of atoms to avoid distor- as to maintain target ion number in control re- tion of the solid, a molecular dynamics (MD) simulation gions (CRs), nCR, beyond the solid/liquid interface: was performed for 0.2 ns at 298 K and 1 bar to relax the solution and equilibrate the simulation cell . µ CR t Fi (z) = ki(ni − ni)G(z) (13) The pressure of the system was held constant within sta- tistical fluctuations using the barostat of Berendsen et t 63 In the above equation, ki is a force constant and ni is the al. With the volume held constant, a series of simu- target number density of species i in the CRs. The force lations were subsequently performed to prepare the ion- is modelled according to a continuous function centered rich reservoir. Here, harmonic restraints were applied at zF : to the distances between ions in solution and a point at the centre of the simulation cell (where the solid slab   −1 1 z0 − zF was positioned). In a series of 0.2–0.5 ns simulations, the G(z) = 1 + cosh (14) 4ω ω minimum energy distance for the restraints was increased from 1 to 6 nm with a force constant of 3×105 kJ mol−1. The function can be made arbitrarily sharp by changing The protocol to prepare the initial configuration for ω. In the simulations in this work, ω was 0.001% of the simulations for NaCl(s) in contact with NaCl(aq) was total extent of z; zF was 5.9 nm and z0 was the z position the same as in the case of graphite; however, here a of the centre of the solid slab; the size of the CRs in z was 5.6 × 5.6 × 2.8 nm NaCl rock-salt supercell formed the 5 −1 2.2 nm; and, ki = 2 × 10 kJ mol . Simulations were solid substrate. To prepare ‘rough’ crystal surfaces, 79 10

NaCl ion pairs were removed from the uppermost layers E. Data availability of one side of the 100 NaCl(s) surface and the structure, in contact with water, was then optimised before per- GROMACS and Plumed input and example forming a 0.2 ns simulation to relax the surface further. output files, including the force field parame- NaCl were also restrained in the simulation cell by ters necessary to reproduce the simulation re- applying restraints to eight ions at the centre of the slab sults reported in this paper, are available on to their initial positions using harmonic potential biases −1 github (https://github.com/aaronrfinney/CmuMD- with k = 600 kJ mol . NaCl_at_graphite). The PLUMED input files are also accessible via PLUMED-NEST (https://www.plumed- nest.org72), the public repository for the PLUMED consortium, using the project IDs: plumID:21.035 C. Self-Diffusion calculations. (NaCl(s) surfaces) and plumID:21.011 (Graphite sur- faces). Details on how to use and implement the CµMD method within PLUMED is available on github (see The self-diffusion coefficients (Dself ) for ions and wa- ter were measured in simulations of bulk solutions which https://github.com/mme-ucl/CmuMD). approached infinite . Single ions were placed into bulk liquid water comprising 4,000 water molecules. F. Acknowledegments Simulations were performed at 298 K and 1 bar us- ing the thermostat described above and the Parinello- Rahman barostat,64 respectively. 50 ns simulations were The authors acknowledge funding from an EPSRC performed using a 1 fs timestep, from which the mean Programme Grant (Grant EP/R018820/1). The authors squared displacement of atoms was calculated according acknowledge the use of the UCL Myriad High Through- to put Computing Facility (Myriad@UCL), and associated support services, in the completion of this work. N 1 X MSD = h|r(t) − r(0)|2i (15) N i=1 REFERENCES

where the time origin was reset every 10 ps. A linear fit to 1 MSD(t) 1 − 10 Iwasawa, Y. Advances in Catalysis; Elsevier, 1987; Vol. 35; pp in the range ns allowed for the calculation 187–264. of the apparent self-diffusion coefficient (D) according to 2Pimpinelli, A.; Villain, J. Physics of crystal growth; Collection MSD(t) = 6Dt. A correction was applied to account for Aléa-Saclay 4; Cambridge University Press: Cambridge ; New 65 + − the finite system size, resulting in Dself for Na , Cl York, 1998. 3Burchell, T. D., Ed. Carbon materials for advanced technologies; and H2O of 1.223(0.005), 1.282(0.008) and 2.762(0.021)× −5 2 −1 Pergamon: Amsterdam ; New York, 1999; OCLC: ocm42274752. 10 cm s , respectively. 4Konatham, D.; Yu, J.; Ho, T. A.; Striolo, A. Simulation Insights for Graphene-Based Water Desalination Membranes. Langmuir 2013, 29, 11884–11897. 5Cohen-Tanugi, D.; Grossman, J. C. Water Desalination across Nanoporous Graphene. Nano Letters 2012, 12, 3602–3608. D. The classical force field. 6Hutchings, G. J. Spiers Memorial Lecture: Understanding reac- tion mechanisms in heterogeneously catalysed reactions. Faraday Discussions 2021, 229, 9–34. Pairwise intermolecular interactions between ions and ex- 7Finney, A. R.; McPherson, I. J.; Unwin, P. R.; Salvalaglio, M. 66 tended simple pint charge (SPC/E) water molecules Electrochemistry, ion adsorption and dynamics in the double were modelled using the Joung and Cheatham force layer: a study of NaCl(aq) on graphite. Chemical Science 2021, field,67 The equilibrium solute molality evaluated for this 10.1039.D1SC02289J. 8García, R. Dynamic atomic force microscopy methods. Surface force field is 3.7 mol/kg.68 Graphite C–C atomic inter- 69 Science Reports 2002, 47, 197–301. actions were modelled using the OPLS/AA force field; 9Hayes, R.; Borisenko, N.; Tam, M. K.; Howlett, P. C.; Endres, F.; C–water intermolecular interactions were modelled using Atkin, R. Double Layer Structure of Ionic Liquids at the Au(111) the pair potential provided by Wu and Aluru,70, Electrode Interface: An Atomic Force Microscopy Investigation. The Journal of Physical Chemistry C 2011, 115, 6855–6863. which was based on fitting to water adsorption energies 10 71 Vlieg, E.; van der Gon, A. W. D.; van der Veen, J. F.; Macdon- from random phase approximation calculations; while ald, J. E.; Norris, C. Surface X-Ray Scattering during Crystal C–ion intermolecular interactions were modelled using Growth: Ge on Ge(111). Physical Review Letters 61 . the atom pair potentials from Williams et al.,51 based on 11Collins, R. W.; Kim, Y.-T. Ellipsometry for Thin-Film and Sur- density functional theory calculations of the adsoprtion face Analysis. Analytical Chemistry 1990, 62, 887A–900A. 12 of ions in a continuum polarisable cavity. Both the C– Watts, K. E.; Blackburn, T. J.; Pemberton, J. E. Optical Spec- troscopy of Surfaces, Interfaces, and Thin Films: A Status Re- water and C–ion potentials were fitted to the same three port. Analytical Chemistry 2019, 91, 4235–4265. point charge model for water, ensuring a self-consistency 13Mullin, J. W. Crystallization, 4th ed.; Butterworth-Heinemann: in the force field used here. Oxford ; Boston, 2001. 11

14Anwar, J.; Zahn, D. Uncovering molecular processes in crys- 37Ozcan, A.; Perego, C.; Salvalaglio, M.; Parrinello, M.; Yazay- tal nucleation and growth by using molecular simulation. Ange- din, O. Concentration gradient driven molecular dynamics: a wandte Chemie International Edition 2011, 50, 1996–2013. new method for simulations of membrane permeation and sepa- 15Salvalaglio, M.; Vetter, T.; Giberti, F.; Mazzotti, M.; Par- ration. Chemical science 2017, 8, 3858–3865. rinello, M. Uncovering molecular details of urea crystal growth 38Radu, M.; Kremer, K. Enhanced crystal growth in binary in the presence of additives. Journal of the American Chemical Lennard-Jones . Physical review letters 2017, 118, Society 2012, 134, 17221–17233. 055702. 16Han, D.; Karmakar, T.; Bjelobrk, Z.; Gong, J.; Parrinello, M. 39Karmakar, T.; Piaggi, P. M.; Perego, C.; Parrinello, M. A canni- Solvent-mediated morphology selection of the active pharmaceu- balistic approach to grand canonical crystal growth. Journal of tical ingredient isoniazid: Experimental and simulation studies. chemical theory and computation 2018, 14, 2678–2683. Chemical Engineering Science 2019, 204, 320–328. 40Loganathan, N.; Bowers, G. M.; Wakou, B. F. N.; 17Frenkel, D.; Smit, B. Understanding molecular simulation: from Kalinichev, A. G.; Kirkpatrick, R. J.; Yazaydin, A. O. Under- algorithms to applications; Elsevier, 2001; Vol. 1. standing methane/carbon dioxide partitioning in clay nano-and 18De Pablo, J. J.; Escobedo, F. A. Molecular simulations in chemi- meso-pores with constant reservoir composition molecular dy- cal engineering: Present and future. American Institute of Chem- namics modeling. Physical Chemistry Chemical Physics 2019, ical Engineers. AIChE Journal 2002, 48, 2716. 21, 6917–6924. 19Allen, M. P.; Tildesley, D. J. Computer Simulation of Liquids; 41Namsani, S.; Ozcan, A.; Yazaydın, A. Ö. Direct Simulation of Oxford University Press, 2017; Vol. 1. Ternary Mixture Separation in a ZIF-8 Membrane at Molecular 20Palmer, J. C.; Debenedetti, P. G. Recent advances in molecular Scale. Advanced Theory and Simulations 2019, 2, 1900120. simulation: A chemical engineering perspective. AIChE Journal 42Ozcan, A.; Semino, R.; Maurin, G.; Yazaydin, A. O. Modeling of 2015, 61, 370–383. gas transport through polymer/MOF interfaces: a microsecond- 21Kashchiev, D. Nucleation; Elsevier, 2000. scale concentration gradient-driven molecular dynamics study. 22Piana, S.; Reyhani, M.; Gale, J. D. Simulating micrometre-scale Chemistry of Materials 2020, 32, 1288–1296. crystal growth from solution. Nature 2005, 438, 70–73. 43Fritsch, S.; Poblete, S.; Junghans, C.; Ciccotti, G.; Delle Site, L.; 23Reguera, D.; Bowles, R.; Djikaev, Y.; Reiss, H. Phase transitions Kremer, K. Adaptive resolution molecular dynamics simulation in systems small enough to be clusters. The Journal of chemical through coupling to an internal particle reservoir. Physical review physics 2003, 118, 340–353. letters 2012, 108, 170602. 24Wedekind, J.; Reguera, D.; Strey, R. Finite-size effects in simula- 44Potestio, R.; Fritsch, S.; Espanol, P.; Delgado-Buscalioni, R.; tions of nucleation. The Journal of chemical physics 2006, 125, Kremer, K.; Everaers, R.; Donadio, D. Hamiltonian adaptive res- 214505. olution simulation for molecular liquids. Physical review letters 25Grossier, R.; Veesler, S. Reaching one single and stable critical 2013, 110, 108301. cluster through finite-sized systems. Crystal Growth and Design 45Wang, H.; Hartmann, C.; Schütte, C.; Delle Site, L. Grand- 2009, 9, 1917–1922. canonical-like molecular-dynamics simulations by using an 26Salvalaglio, M.; Vetter, T.; Mazzotti, M.; Parrinello, M. Control- adaptive-resolution technique. Physical Review X 2013, 3, ling and predicting crystal shapes: The case of urea. Angewandte 011018. Chemie 2013, 125, 13611–13614. 46Agarwal, A.; Zhu, J.; Hartmann, C.; Wang, H.; Delle Site, L. 27Agarwal, V.; Peters, B. Solute precipitate nucleation: A review of Molecular dynamics in a grand ensemble: Bergmann–Lebowitz theory and simulation advances. Advances in Chemical Physics: model and adaptive resolution simulation. New Journal of Volume 155 2014, 97–160. Physics 2015, 17, 083042. 28Salvalaglio, M.; Perego, C.; Giberti, F.; Mazzotti, M.; Par- 47Prausnitz, J. M.; Lichtenthaler, R. N.; De Azevedo, E. G. Molecu- rinello, M. Molecular-dynamics simulations of urea nucleation lar thermodynamics of fluid-phase equilibria; Pearson Education, from aqueous solution. Proceedings of the National Academy of 1998. Sciences 2015, 112, E6–E14. 48Liu, X.; Vlugt, T. J.; Bardow, A. Predictive Darken equation for 29Salvalaglio, M.; Tiwary, P.; Maggioni, G. M.; Mazzotti, M.; Par- Maxwell-Stefan diffusivities in multicomponent mixtures. Indus- rinello, M. Overcoming time scale and finite size limitations to trial & engineering chemistry research 2011, 50, 10350–10358. compute nucleation rates from small scale well tempered meta- 49Liu, X.; Schnell, S. K.; Simon, J.-M.; Krüger, P.; Bedeaux, D.; dynamics simulations. The Journal of chemical physics 2016, Kjelstrup, S.; Bardow, A.; Vlugt, T. J. Diffusion coefficients from 145, 211925. molecular dynamics simulations in binary and ternary mixtures. 30Çağin, T.; Pettitt, B. M. Molecular dynamics with a variable International journal of thermophysics 2013, 34, 1169–1196. number of molecules. Molecular physics 1991, 72, 169–175. 50Bird, R. B. Transport phenomena. Appl. Mech. Rev. 2002, 55, 31Papadopoulou, A.; Becker, E. D.; Lupkowski, M.; van Swol, F. R1–R4. Molecular dynamics and Monte Carlo simulations in the grand 51Williams, C. D.; Dix, J.; Troisi, A.; Carbone, P. Effective Polar- canonical ensemble: Local versus global control. The Journal of ization in Pairwise Potentials at the Graphene–Electrolyte Inter- chemical physics 1993, 98, 4897–4908. face. The Journal of Physical Chemistry Letters 2017, 8, 703– 32Yau, D. H.; Liem, S. Y.; Chan, K.-Y. A contact cavity-biased 708. method for grand canonical Monte Carlo simulations. The Jour- 52Liu, X.; Schnell, S. K.; Simon, J.-M.; Krüger, P.; Bedeaux, D.; nal of chemical physics 1994, 101, 7918–7924. Kjelstrup, S.; Bardow, A.; Vlugt, T. J. H. Diffusion Coefficients 33Lo, C.; Palmer, B. Alternative Hamiltonian for molecular dynam- from Molecular Dynamics Simulations in Binary and Ternary ics simulations in the grand canonical ensemble. The Journal of Mixtures. International Journal of Thermophysics 2013, 34, chemical physics 1995, 102, 925–931. 1169–1196. 34Lynch, G. C.; Pettitt, B. M. Grand canonical ensemble molec- 53Darken, L. M. Diffusion, Mobility and Their Interrelation ular dynamics simulations: Reformulation of extended system through Free Energy in Binary Metallic Systems. Transactions dynamics approaches. The Journal of chemical physics 1997, of the American Institute of Mining and Metallurgical Engineers 107, 8594–8610. 1948, 175, 184. 35Delgado-Buscalioni, R.; Coveney, P. USHER: an algorithm for 54Liu, X.; Schnell, S. K.; Simon, J.-M.; Bedeaux, D.; Kjelstrup, S.; particle insertion in dense fluids. The Journal of chemical physics Bardow, A.; Vlugt, T. J. Fick diffusion coefficients of liquid mix- 2003, 119, 978–987. tures directly obtained from equilibrium molecular dynamics. 36Perego, C.; Salvalaglio, M.; Parrinello, M. Molecular dynamics The Journal of Physical Chemistry B 2011, 115, 12921–12929. simulations of solutions at constant chemical potential. The Jour- 55Liu, X.; Martín-Calvo, A.; McGarrity, E.; Schnell, S. K.; nal of chemical physics 2015, 142, 144113. Calero, S.; Simon, J.-M.; Bedeaux, D.; Kjelstrup, S.; Bardow, A.; 12

Vlugt, T. J. Fick diffusion coefficients in ternary liquid systems 64Parrinello, M.; Rahman, A. Polymorphic transitions in single from equilibrium molecular dynamics simulations. Industrial & crystals: A new molecular dynamics method. Journal of Applied engineering chemistry research 2012, 51, 10247–10258. Physics 1981, 52, 7182–7190. 56Perego, C.; Salvalaglio, M.; Parrinello, M. Molecular dynamics 65Yeh, I.-C.; Hummer, G. System-Size Dependence of Diffusion simulations of solutions at constant chemical potential. The Jour- Coefficients and Viscosities from Molecular Dynamics Simula- nal of Chemical Physics 2015, 142, 144113. tions with Periodic Boundary Conditions. The Journal of Phys- 57Hess, B.; Kutzner, C.; van der Spoel, D.; Lindahl, E. GROMACS ical Chemistry B 2004, 108, 15873–15879. 4: Algorithms for Highly Efficient, Load-Balanced, and Scalable 66Berendsen, H. J. C.; Grigera, J. R.; Straatsma, T. P. The missing Molecular Simulation. Journal of Chemical Theory and Compu- term in effective pair potentials. The Journal of Physical Chem- tation 2008, 4, 435–447. istry 1987, 91, 6269–6271. 58Tribello, G. A.; Bonomi, M.; Branduardi, D.; Camilloni, C.; 67Joung, I. S.; Cheatham, T. E. Determination of Alkali and Bussi, G. PLUMED 2: New feathers for an old bird. Computer Halide Monovalent Ion Parameters for Use in Explicitly Solvated Physics Communications 2014, 185, 604–613. Biomolecular Simulations. The Journal of Physical Chemistry B 59Bussi, G.; Donadio, D.; Parrinello, M. Canonical sampling 2008, 112, 9020–9041. through velocity rescaling. The Journal of Chemical Physics 68Benavides, A. L.; Aragones, J. L.; Vega, C. Consensus on the 2007, 126, 014101. of NaCl in water from computer simulations using the 60Darden, T.; York, D.; Pedersen, L. Particle mesh Ewald: An chemical potential route. The Journal of Chemical Physics 2016, N · logN method for Ewald sums in large systems. The Journal 144, 124504. of Chemical Physics 1993, 98, 10089–10092. 69Jorgensen, W. L.; Maxwell, D. S.; Tirado-Rives, J. Development 61Hess, B.; Bekker, H.; Berendsen, H. J. C.; Fraaije, J. G. E. M. and Testing of the OPLS All-Atom Force Field on Conforma- LINCS: A linear constraint solver for molecular simulations. tional Energetics and Properties of Organic Liquids. Journal of Journal of Computational Chemistry 1997, 18, 1463–1472. the American Chemical Society 1996, 118, 11225–11236. 62Trucano, P.; Chen, R. Structure of graphite by neutron diffrac- 70Wu, Y.; Aluru, N. R. Graphitic Carbon–Water Nonbonded Inter- tion. Nature 1975, 258, 136. action Parameters. The Journal of Physical Chemistry B 2013, 63Berendsen, H. J. C.; Postma, J. P. M.; van Gunsteren, W. F.; 117, 8802–8813. DiNola, A.; Haak, J. R. Molecular dynamics with coupling to 71Ma, J.; Michaelides, A.; Alfe, D.; Schimka, L.; Kresse, G.; an external bath. The Journal of Chemical Physics 1984, 81, Wang, E. Adsorption and diffusion of water on graphene from 3684–3690. first principles. Physical Review B 2011, 84, 033402. 72Bonomi, M. Promoting transparency and reproducibility in en- hanced molecular simulations. Nature methods 2019, 16, 670– 673.