arXiv:1703.02759v2 [cond-mat.mes-hall] 10 Apr 2017

10
Osmotic and diffusio-osmotic flow generation at high solute concentration. II. Molecular dynamics simulations Hiroaki Yoshida, 1,2, a) Sophie Marbach, 1, b) and Lyd´ eric Bocquet 1, c) 1) LPS, UMR CNRS 8550, Ecole Normale Sup´ erieure, 24 rue Lhomond, 75005 Paris, France 2) Toyota Central R&D Labs., Inc., Nagakute, Aichi 480-1192, Japan (Dated: 11 April 2017) In this paper, we explore osmotic transport by means of molecular dynamics (MD) simulations. We first con- sider osmosis through a membrane, and investigate the reflection coefficient of an imperfectly semi-permeable membrane, in the dilute and high concentration regimes. We then explore the diffusio-osmotic flow of a solute-solvent fluid adjacent to a solid surface, driven by a chemical potential gradient parallel to the surface. We propose a novel non-equilibrium MD (NEMD) methodology to simulate diffusio-osmosis, by imposing an external force on every particle, which properly mimics the chemical potential gradient on the solute in spite of the periodic boundary conditions. This NEMD method is validated theoretically on the basis of linear-response theory by matching the mobility with their Green–Kubo expressions. Finally, we apply the framework to more realistic systems, namely a water-ethanol mixture in contact with a silica or a graphene surface. I. INTRODUCTION Transport phenomena involving solute concentration difference or gradient of a solute-solvent fluid emerge in many scientific and industrial fields, from the chemical physics of biological membranes to the development of desalination processes. 1,2 Furthermore, there is a grow- ing interest in applications harnessing concentration gra- dients to drive flows, 3–9 in particular for energy conver- sion 10 or storage using nano-scale membranes. 11 There is accordingly a need for a better fundamental understand- ing of such transport phenomena. Osmosis across a membrane is a transport phenomenon driven by a solute concentration difference. Let us con- sider a situation where two fluid reservoirs with so- lute concentration difference c are separated by a mem- brane. If the membrane is completely semi-permeable, i.e., only the solvent particles are allowed to pass through the membrane, an osmotic pressure builds up, and is well described by the classical van ’t Hoff type equa- tion: Π= k B Tc, with k B the Boltzmann constant, T the temperature. In contrast, if the membrane is par- tially semi-permeable, i.e., solute particles are not com- pletely rejected, then also solute flux occurs across the membrane. Transport through the membrane in the lat- ter situation is described by the Kedem–Kachalsky equa- tions, 12–14 which include the reflection coefficient σ as a phenomenological correction to the van ’t Hoff equation. Relevant definitions of σ for the low concentration regime were given, e.g., by Manning, 15 and extended to arbi- trary concentrations in the first paper of this series. We also provide there a comprehensive theory to understand a) Electronic mail: [email protected] b) Electronic mail: [email protected] c) Electronic mail: [email protected] the origin of the reflection coefficient σ at a microscopic level. 16 Diffusio-osmotic flow is a more subtle phenomenon which occurs under solute gradients in the presence of a fluid-solid interface. In a bulk fluid, a concentration gradient of a solute will lead to a diffusive flux of both components, but there is no total fluid flow because the forces acting on the solvent and solute particles are bal- anced. However, in the presence of an interface, the so- lute concentration in a thin layer near the surface differs from that in the bulk, because of either an adsorption or a repulsion of solute particles. Consequently the force balance is broken in this thin layer and the driving force results in the fluid diffusio-osmotic flow. Such interfa- cially driven flow is especially relevant to small-scale sys- tems, typically in microfluidic devices with narrow chan- nels and through nanoporous membranes, because of the large surface-to-volume ratio. 17 Anderson and co-workers provided a theoretical framework of the diffusio-osmotic flow for the case of low concentration of solute, 18,19 and in the accompanying paper we extended the theory to the high-concentration regime of the solute. 16 In the present paper, we numerically study the micro- scopic aspects of these two problems, i.e., the osmosis and the diffusio-osmotic flow, using molecular dynamics (MD) simulations. Our first goal here is to validate the theoretical predictions developed in the accompanying paper, by means of direct measurements of the osmotic pressure and the diffusio-osmotic flow at a microscopic scale. However, in order to achieve this objective, one encounters a methodological difficulty in simulating the diffusio-osmotic flow directly; there is no existing method to implement directly a chemical potential gradient com- patible with periodic boundary conditions. In this study, we accordingly introduce a novel non-equilibrium MD (NEMD) technique, which circumvents this difficulty and allows to impose a proper external forcing representing a gradient in the chemical potential of the solute. We find arXiv:1703.02759v2 [cond-mat.mes-hall] 10 Apr 2017

Transcript of arXiv:1703.02759v2 [cond-mat.mes-hall] 10 Apr 2017

Osmotic and diffusio-osmotic flow generation at high solute concentration.II. Molecular dynamics simulations

Hiroaki Yoshida,1, 2, a) Sophie Marbach,1, b) and Lyderic Bocquet1, c)1)LPS, UMR CNRS 8550, Ecole Normale Superieure, 24 rue Lhomond, 75005 Paris,France2)Toyota Central R&D Labs., Inc., Nagakute, Aichi 480-1192, Japan

(Dated: 11 April 2017)

In this paper, we explore osmotic transport by means of molecular dynamics (MD) simulations. We first con-sider osmosis through a membrane, and investigate the reflection coefficient of an imperfectly semi-permeablemembrane, in the dilute and high concentration regimes. We then explore the diffusio-osmotic flow of asolute-solvent fluid adjacent to a solid surface, driven by a chemical potential gradient parallel to the surface.We propose a novel non-equilibrium MD (NEMD) methodology to simulate diffusio-osmosis, by imposingan external force on every particle, which properly mimics the chemical potential gradient on the solute inspite of the periodic boundary conditions. This NEMD method is validated theoretically on the basis oflinear-response theory by matching the mobility with their Green–Kubo expressions. Finally, we apply theframework to more realistic systems, namely a water-ethanol mixture in contact with a silica or a graphenesurface.

I. INTRODUCTION

Transport phenomena involving solute concentrationdifference or gradient of a solute-solvent fluid emerge inmany scientific and industrial fields, from the chemicalphysics of biological membranes to the development ofdesalination processes.1,2 Furthermore, there is a grow-ing interest in applications harnessing concentration gra-dients to drive flows,3–9 in particular for energy conver-sion10 or storage using nano-scale membranes.11 There isaccordingly a need for a better fundamental understand-ing of such transport phenomena.

Osmosis across a membrane is a transport phenomenondriven by a solute concentration difference. Let us con-sider a situation where two fluid reservoirs with so-lute concentration difference c are separated by a mem-brane. If the membrane is completely semi-permeable,i.e., only the solvent particles are allowed to pass throughthe membrane, an osmotic pressure builds up, and iswell described by the classical van ’t Hoff type equa-tion: Π = kBTc, with kB the Boltzmann constant, Tthe temperature. In contrast, if the membrane is par-tially semi-permeable, i.e., solute particles are not com-pletely rejected, then also solute flux occurs across themembrane. Transport through the membrane in the lat-ter situation is described by the Kedem–Kachalsky equa-tions,12–14 which include the reflection coefficient σ as aphenomenological correction to the van ’t Hoff equation.Relevant definitions of σ for the low concentration regimewere given, e.g., by Manning,15 and extended to arbi-trary concentrations in the first paper of this series. Wealso provide there a comprehensive theory to understand

a)Electronic mail: [email protected])Electronic mail: [email protected])Electronic mail: [email protected]

the origin of the reflection coefficient σ at a microscopiclevel.16

Diffusio-osmotic flow is a more subtle phenomenonwhich occurs under solute gradients in the presence ofa fluid-solid interface. In a bulk fluid, a concentrationgradient of a solute will lead to a diffusive flux of bothcomponents, but there is no total fluid flow because theforces acting on the solvent and solute particles are bal-anced. However, in the presence of an interface, the so-lute concentration in a thin layer near the surface differsfrom that in the bulk, because of either an adsorptionor a repulsion of solute particles. Consequently the forcebalance is broken in this thin layer and the driving forceresults in the fluid diffusio-osmotic flow. Such interfa-cially driven flow is especially relevant to small-scale sys-tems, typically in microfluidic devices with narrow chan-nels and through nanoporous membranes, because of thelarge surface-to-volume ratio.17 Anderson and co-workersprovided a theoretical framework of the diffusio-osmoticflow for the case of low concentration of solute,18,19 andin the accompanying paper we extended the theory tothe high-concentration regime of the solute.16

In the present paper, we numerically study the micro-scopic aspects of these two problems, i.e., the osmosisand the diffusio-osmotic flow, using molecular dynamics(MD) simulations. Our first goal here is to validate thetheoretical predictions developed in the accompanyingpaper, by means of direct measurements of the osmoticpressure and the diffusio-osmotic flow at a microscopicscale. However, in order to achieve this objective, oneencounters a methodological difficulty in simulating thediffusio-osmotic flow directly; there is no existing methodto implement directly a chemical potential gradient com-patible with periodic boundary conditions. In this study,we accordingly introduce a novel non-equilibrium MD(NEMD) technique, which circumvents this difficulty andallows to impose a proper external forcing representing agradient in the chemical potential of the solute. We find

arX

iv:1

703.

0275

9v2

[co

nd-m

at.m

es-h

all]

10

Apr

201

7

2

an excellent agreement between our method and the re-sults of a Green–Kubo approach based on linear-responsetheory, and furthermore validate Onsager’s reciprocal re-lation. The NEMD method is then used to validate thetheory, and applied to a more realistic system of a water-ethanol mixture in contact with a silica or a graphenesurface.

In Sec. II, we examine the osmotic pressure across amembrane, focusing on the evaluation of the reflectioncoefficient of incomplete semi-permeable membranes atlow and high concentrations. We next consider diffusio-osmosis in Sec. III, including the introduction and vali-dation of the new methodology mentioned above. Thena brief summary given in Sec. IV concludes the paper.

II. REFLECTION COEFFICIENT OF PARTIALLYSEMI-PERMEABLE MODEL MEMBRANES

In this section, we first consider the osmotic pressureacross a model membrane, which allows to gain muchinsight into the osmotic transport, and we introduce aversatile method to measure the osmotic pressure.

A. Theory

For the transport of a solute-solvent fluid across a fil-tration membrane with pressure and concentration dif-ferences between the two sides, the Kedem–Kachalskymodel is widely used to describe the volume flux (perunit area) of the solution Q and the particle flux of thesolute Js:

Q = −Lhyd (∆p− σkBT∆c) , (1)

Js = −LDω∆c+ c(1− σ)Q, (2)

where Lhyd is the permeability coefficient, c is the con-centration of the solute, LD = D/L is the solute perme-ability with D its diffusion coefficient and L the thicknessof the membrane, ω is the factor for the effective mobilityvalue in the membrane, and σ is the reflection coefficientthat is a measure of the semi-permeability of the mem-brane.15,20,21 The non-dimensional coefficients ω and σare expected to be related by a linear relationship,12 as1− σ ∝ ω.

Most of the approaches so far treat the reflection coef-ficient as a phenomenological parameter, and discussionon a direct connection with parameters characterizingthe membrane is rare. Following our theoretical discus-sion in the companion paper, we consider here a modelmembrane, taking the form of an energy barrier felt bythe solute particles.16 This simplified situation allows toobtain an expression for the reflection coefficient which

takes the form:

σ = 1−∫ L/2−L/2

dx′

λ[c(x′)]∫ L/2−L/2

dx′

λ[c(x′)]c0c(x′)

, (3)

where λ is the mobility of the solute particles, and c0is the average solute concentration far from the mem-branes; c(x) is the stationary concentration distribution,see Ref. 16. Since no assumption is made on the mag-nitude of c0, this expression is valid beyond the dilutesolute limit. In the dilute solute limit, the formula givenin Eq. (3) reduces to the one derived by Manning15

σ = 1− L∫ L/2−L/2 dx

′ exp[+βU(x′)], (4)

where U denotes the energy barrier representing themembrane.

B. MD simulations

In the present study, we validate the theoretical predic-tions for the reflection coefficient by means of MD sim-ulations. We use a system of identical Lennard-Jonesparticles for the fluid mixture with a potential barriermodel for the membrane, similar to that considered inRef. 22. Whereas they use a cubic box made of thesemi-permeable membrane, here we consider a more di-rect geometrical setup as shown in Fig. 1. Two reservoirsare separated by a membrane as shown in Fig. 1(a); themembrane is not visible in the figure. The left reservoircontains a pure liquid solvent, while the right reservoiris filled with a liquid solution containing solute parti-cles. The membrane is modeled by an energy barrier U ,which acts only on the solute particles, as illustrated in

x

(a)

(b)

U(x)

fixed

L

FIG. 1. (a) Simulation setup of two fluid reservoirs separatedby a membrane. (b) Illustration of the energy barrier U(x)felt only by the solute particles (red).

3

Fig. 1(b). The ends of the reservoirs are closed by rigidwalls consisting of an FCC lattice made of the same par-ticles as the solvent. The left wall serves as a piston,maintaining the normal pressure in the left reservoir atPL = P0 (see below for P0). On the other hand, the rightwall is fixed, and we measure the pressure in the rightreservoir PR from the total force acting on this wall. Ifthe membrane is perfectly semi-permeable, i.e., there isno flux of solute particles across the membrane, then thesystem reaches the steady state. In the present study,the osmotic pressure is measured as the pressure differ-ence Π = PR − PL; it could be measured alternativelywith summing all the forces exerted on each particle bythe membrane, which yields identical values. The case ofincomplete semi-permeability is less straightforward anddescribed below. Since the right wall is fixed in our setup,there is no net flux of solution; the case of finite flux ofmixture across the membrane could also be simulated bycontrolling the permeability coefficient Lhyd, e.g. by in-troducing a drag force acting on the solvent particles inthe membrane (see Ref. 16.) For simplicity we do notconsider any drag force here.

For the inter-particle interactions we assume aLennard-Jones (LJ) potential among the solvent andsolute particles: Uij(r) = 4εLJij [(σLJ

ij /r)12 − (σLJ

ij /r)6].

The parameters for the solute-solute, solute-solvent, andsolvent-solvent interactions are commonly set as εLJij =

εLJ0 and σLJij = σLJ

0 . The mass of all the particles ism0. Therefore, the solute and solvent particles are me-chanically identical, except that the solute particles feelthe energy barrier representing the membrane. The wallparticles are also described by the same interaction pa-rameter set. In presenting the simulation results usingthe LJ potential, we use the units normalized in terms ofthe LJ parameters εLJ0 and σLJ

0 , i.e., the reference length`0 = σLJ

0 , the energy ε0 = εLJ0 , the force f0 = ε0/`0, thepressure P0 = f0/`

20, and the time τ0 = `20m0/ε

LJ0 . The

energy barrier U(z) takes the one-dimensional Gaussianform:

U(z) = U0 exp(−a(z − z0)2), (5)

where z0 is the position of the membrane, and U0 con-trols the height of the energy barrier. The thickness ofthe membrane is ∼ √a, where a is fixed at 10/`20. Thispotential is cut off at a distance `cut, with `cut = 4`0. Thesize of the simulation box in y and z is 22.7×22.7 `20, andtypical number of particles in one reservoir is 8380. Thetemperature is kept constant at kBT/ε0 = 1 using theNose–Hoover thermostat in all directions, and then thedensity at pressure P0 is 0.75 `−30 . The time integrationis carried out with the time step 0.005τ0. For the actualMD implementation, the open-source code LAMMPS isused throughout the paper.23

Figure 2(a) shows the MD results of the osmotic pres-sure as a function of the solute concentration c in theright reservoir. In the case of U0 = 30 ε0, no solute par-ticles cross the membrane during the simulation up to

(a)

(b)

0

0.1

0.2

0.3

0.4

0.5

0.6

0 0.1 0.2 0.3 0.4c

× −30

Π

×P0

U0

×ε0

σ

U0 = 30 ε0

U0 = 3ε0

::

::

van’t Hoff (linearized)van’t Hoff

x

c/c

res

× 0

(c)

0

0.2

0.4

0.6

0.8

1

1.2

0 5 10 15

0

0.2

0.4

0.6

0.8

1

1.2

-4 -3 -2 -1 0 1 2 3 4

cres = 0.036 30:

cres = 0.36 30:

: Boltzman

distribution

c = 0 .036 30:

c = 0.36 30:

: linear theory

: non-linear theory

FIG. 2. (a) Osmotic pressure Π versus solute concentrationc in the right reservoir. The symbols indicate the MD results,and the solid line indicates the van ’t Hoff type formula givenin Eq. (6). The linearized van ’t Hoff law is shown by thedashed line. (b) Reflection coefficient σ versus height of theenergy barrier U0. The MD results are shown by the symbols.While the solid line is the theory for the low concentrationregime given in Eq. (4), the dashed line and dash-dotted lineare the results of the generalized theory (Eq. (3)). (c) Sta-tionary concentration profiles across the membrane, for thecase of U0 = 5ε0; here, the concentrations in both reservoirsare identical. The symbols indicate the MD results, which areused to calculate σ using Eq. (3) (and plotted as crosses andlines in panel (b)). The solid line is the Boltzmann distribu-tion.

5× 106 time steps, i.e., the membrane exhibits completesemi-permeability. In this regime, the reflection coeffi-cient is unity, σ = 1. The osmotic pressure then con-verges to the standard van ’t Hoff law, Π = kBTc, for di-

4

lute solutions. For the larger concentrations, Π departsfrom the linear line, but is still captured by the van ’tHoff law before linearization:

Π = −ρvkBT ln(1− χ), (6)

where χ is the molar fraction of the solute, χ = ρ−1v (1/c−1/ρu + 1/ρv)

−1, and ρu and ρv are the density of the so-lute and solvent, respectively. On the other hand, whenthe energy barrier is small (U0 = 3ε0), some solute par-ticles permeate through the membrane during the sim-ulations. The membrane is imperfectly semi-permeableand one expects σ < 1. We observe indeed that the os-motic pressure drops, coherently with σ < 1. In thissituation, the pressure in the right reservoir evolves withtime. We accordingly compute the osmotic pressure inthe following manner: after equilibration of the systemat U0 = 30ε0 for at least 105 time steps, we set the energybarrier at U0 < 30ε0. Then we average the results over5× 105 time steps to evaluate the osmotic pressure.

More quantitative data of the reflection coefficient σfor the incomplete semi-permeable membrane are givenin Fig. 2(b). Here, the MD values are obtained withσ = Π/Πcom, where Πcom denotes the value of thecomplete semi-permeable case, i.e., the data shown inFig. 2(a) for U0 = 30ε0. The MD results are plottedfor two values of initial concentration in the right reser-voir, c = 0.036/`30 and c = 0.36/`30. For comparison, thetheoretical predictions quoted in Sec. II A are also plot-ted in the figure. The prediction of Eq. (4), which isvalid in the dilute limit, agrees well with the MD dataat c = 0.036/`30. For the high concentration c = 0.36/`30,however, the MD data departs from the prediction ofEq. (4) and takes smaller values.

The prediction of Eq. (3) is evaluated by numeri-cally integrating the concentration profiles of the MDresults. Since the solute and solvent particles consid-ered here are mechanically identical, the mobility is in-dependent of the concentration, and thus λ cancels outin Eq. (3). The stationary concentration profiles areobtained at the equilibrium state, with the same con-centration in the two reservoirs. Typical concentrationprofiles are shown in Fig. 2(c) for U0 = 5ε0, togetherwith the Boltzmann distribution valid in the dilute limit,c = c0 exp(−U/kBT ). The deviation from the Boltzmanndistribution at c = 0.36/`30 is a non-linear effect due tohigh concentration. The interaction between solute par-ticles becomes significant, and it causes the oscillationsvisible in the concentration profile, similar to fluid den-sity oscillations that generally occur near solid surfaces;24

the solute-solvent interaction causes a similar oscillationin the solvent density profile (not shown.) Taking intoaccount this effect, Eq. (3) accurately predicts the reflec-tion coefficient as shown in Fig. 2(b). This demonstratesthe usefulness of the theoretical prediction for wide con-centration ranges beyond the dilute regime, once the con-centration profiles are measured or estimated.

III. NON-EQUILIBRIUM SIMULATIONS OFDIFFUSIO-OSMOTIC FLOW

In this section we now consider diffusio-osmosis, i.e.the flow induced by solute gradients, but now tangentialto a solid surface. We perform molecular dynamics ofdiffusio-osmosis for dense solute concentrations. To thisend we introduce a new methodology allowing to simulatethe effect of chemical potential gradients numerically.

A. Theory: a reminder

The geometry under consideration is shown inFig. 3(a), with a chemical potential gradient applied par-allel to the surface. In the bulk region, the total forceis zero yet solute and solvent fluxes are observed. Thesolute concentration in a layer adjacent to the surface de-viates from that in the bulk, because of either a prefer-ential adsorption or a depletion of the solute. The forceunbalance in the thin layer is the driving force of thediffusio-osmotic flow. As shown in the first paper of theseries,16 the fluid velocity is linearly proportional to thechemical potential gradient, according to

vx(z) =1

η

∫ z

0

dz′∫ ∞z′

dz′′ (c∞ − c(z′′))∇xµ, (7)

where η is the fluid viscosity, c∞ is the concentration inthe bulk region sufficiently far from the surface, and ∇xµis the chemical potential gradient. This formula correctlyrecovers the classical result for the dilute solution.18,19

The diffusio-osmotic mobility KDO, relating the velocityv∞ far from the surface to the gradient of the chemicalpotential as v∞ = KDOc∞∇xµ, is then given by

KDO = −1

η

∫ ∞0

dz′ z′(c(z′)

c∞− 1

). (8)

The mobility KDO is negative for an excess surface con-centration at the interface, i.e., the flow of solvent goestowards the low chemical potential area. Respectively,it is positive if there is a surface depletion, and the flowreverses. Note that in the case of slip at the interfacewith typical slip length b, a slip velocity adds to Eq. (7),such that KDO is enhanced by a factor (1 + b/Ls), withLs the thickness of the diffusion layer.25

B. Principles for an NEMD diffusio-osmosis

In this section, our goal is to develop a method tosimulate directly diffusio-osmosis on the basis of non-equilibrium MD (NEMD) simulations. This implies togenerate the diffusio-osmotic flow by applying an exter-nal field that is consistent with the application of a chem-ical potential gradient −∇xµ. A key characteristic ofdiffusio-osmosis, like all interfacially driven flows – in-

5

ρ`

03

0

0.5

1

1.5

2

0 10 20 30z/`0

z

x

Bulk region ΩB

Considered region Ω Slip wall(a)

(b)

z

x

(c)

(d)

Pressure gradient

Chemical potential gradient

−∇xP

−∇xµ

Surface

H

FIG. 3. (a) Computational geometry for the MD simulation for an LJ mixture (solute: red, solvent: blue) in contact witha solid wall (gray). (b) Typical concentration profile along the z direction. (c) Schematic illustration of the NEMD methodmodeling the pressure gradient. It is modeled by a force acting on each particle. (d) NEMD method for simulating the chemicalpotential gradient. It is modeled as a forward force per solute particle (red) and counter force per solvent particle (blue) suchthat the total force in the bulk is zero.

cluding electro- or thermo-osmosis flows26,27 –, is thatit is a force-free transport phenomenon. This can bedemonstrated from the fact that the hydrodynamic ve-locity profile is flat far away from the surface, so thatall forces acting on the surface – direct interactions andhydrodynamics – do vanish in the bulk. Accordingly, ifany force is applied to the system (solute+solvent), itshould be balanced so that the total force acting on thefluid should vanish in the bulk region. This is obviouslyin contrast with the pressure driven flow. In the MDsimulations, for the latter case, an external force Fp isapplied commonly to each particle in the fluid, as shownin Fig. 3(c).28–30 The applied pressure gradient, i.e., theforce per volume, is then identified as −∇xP = FpN/V ,where N is the number of particles, and V is the volumeof the whole system Ω.

To simulate diffusio-osmosis, we accordingly proposethe following scheme, where we apply a differential forceon the solute and on the solvent, see Fig. 3(d):

• an external force Fµ is applied to each solute par-ticle in the whole system Ω.

• a counter force −[NBs /(N

B −NBs )]Fµ is applied to

each solvent particle in Ω, where NBs and NB are

respectively the number of solute particles and thetotal number of particles in the bulk region.

Note that the bulk region, denoted by ΩB , is defined asa volume far from the wall such that the density (andconcentration) profile is flat, as depicted in Fig. 3(b).

The counter force therefore ensures the force balance inthe bulk volume ΩB . The external force strength is thenrelated to the chemical potential gradient as

−∇xµ = FµNB/(NB −NB

s ), (9)

as is confirmed below via the Green–Kubo approach.

C. Validation of the NEMD scheme: Green–Kuborelationships

We now validate this methodology on the basis of thelinear response theory. Indeed due to the Onsager sym-metry, one expects that the same diffusio-osmotic mobil-ity will relate two symmetric situations: on the one hand,the solvent flow under a solute chemical gradient, and onthe other hand, the (excess) solute flux under a pressuredrop.25 This is expressed in the transport matrix as:[

QJs − c∗∞Q

]=

[MQQ MQJ

MJQ MJJ

] [−∇xP−∇xµ

], (10)

where Q and Js−c∗∞Q are the total (volume) flux and theexcess solute flux, respectively, as described below. Theoff-diagonal coefficients in Eq. (10) are expected to beidentical MQJ = MJQ due to the Onsager time-reversal,symmetry relationship. Let us therefore demonstratethat our NEMD methodology complies to this symmetryrelationship. We will calculate the Green–Kubo expres-

6

sion for both MQJ and MJQ cross coefficients.

We first remind quickly the general statements of lin-ear response theory, i.e., on the response of an observedvariable B to an external potential field A(xi)F0, whereF0 is a constant microscopic force and A(xi) is a func-tion of the positions of the particles xi. The observedvariable is expressed as 〈B〉 = MBAF0, where 〈·〉 is theensemble average, and

MBA =1

kBT

∫ ∞0

〈B(t)A(0)〉dt. (11)

In the NEMD approach, the external field is A × Fµwith

A =∑

i∈solute

xi −NBs

NB −NBs

∑i∈solvent

xi, (12)

where xi is the coordinate along the x axis of particlenumber i. The observed variable is the total flux, i.e.,B = Q(t) = (1/N)

∑i∈all xi. Injecting these into the

definition ofMBA gives the Green–Kubo formula for thediffusio-osmotic flow 〈Q〉 (abbreviated Q) generated bythe NEMD scheme:

Q = MQJ

(NB

NB −NBs

)Fµ, (13)

with MQJ =V

kBT

∫ ∞0

〈Q(t)(Js − c∗∞Q)(0)〉dt. (14)

Here V is the volume of Ω, and c∗∞ = φBρav, with φB =NBs /N

B being the molar fraction of solute in the bulkregion ΩB , and ρav the density averaged over Ω. Thesolute flux Js is calculated in terms of the particle velocityas Js = (1/V )

∑i∈solute xi.

Similarly, in the reciprocal situation, we measure thesolute flux B = Js(t)− c∗∞Q(t) under a pressure gradientrepresented by A =

∑i∈all xi. We deduce the symmetric

formula:

Js − c∗∞Q = MJQ

(N

V

)Fp, (15)

with MJQ =V

kBT

∫ ∞0

〈(Js − c∗∞Q)(t)Q(0)〉dt. (16)

Comparing Eqs. (14) and (16), we find that MQJ = MJQ

so that the proposed scheme complies to Onsager’s recip-rocal relation. Regarding MQJ and MJQ as the diffusio-osmotic transport coefficients relating the fluxes with theexternal fields, one may interpret that the microscopicforces Fµ and Fp in terms of the thermodynamic forces:

−∇xµ = FµNB

NB −NBs

, (17)

and similarly −∇xP = FpN/V .

Equation (17) indicates that, given the chemical po-tential gradient, the force acting on the solute particles

is Fµ = −(1− φB)∇xµ and that on the solvent particlesis −[NB

s /(NB − NB

s )]Fµ = φB∇xµ. Physically, whilethe force directly originating in −∇xµ acts only on thesolute particles, the counteracting force φB∇xµ appliesto all the particles to ensure a vanishing net force.

D. Numerical validation of the NEMD methodology

We now apply this methodology in NEMD simulations.Our goals are first to highlight the implementation of theNEMD and second to validate the NEMD mobility bycomparing it to the equilibrium Green–Kubo estimates.

1. Numerical details

In this section, solvent, solute, and wall particles inter-act via the LJ potential. While the parameters for thesolute-solute, solute-solvent, solvent-solvent, and solvent-wall interactions are commonly set as εLJij = εLJ0 = ε0 and

σLJij = σLJ

0 = `0, the parameters for the solute-wall in-

teraction εLJsolute,wall and σLJsolute,wall are varied to control

the surface excess of solute particles. The wall particlesare fixed at z = 0, as in Fig. 3(a), on an FCC lattice

with lattice constant√

2`0. In this setting, the hydrody-namic slip at the interface between the wall and the fluidis negligible.25 An artificial reflecting wall is placed totruncate the computational domain, sufficiently far fromthe wall. At this reflecting wall, the incoming atoms aresimply reflected with no tangential momentum transfer,i.e., the wall is a complete slip boundary. Since an ar-tificial oscillation of density occurs in the vicinity of thereflecting boundary, we need to exclude this part fromall measurements. We thus consider a specific region Ω(shown in Fig. 3(a)), that extends to typically a distance10`0 from the reflecting boundary. The particle density isdetermined such that the normal pressure on the surfaceis P0.26,31

The lateral dimension of the simulation box is 17`0 ×17`0, and the height of domain Ω is H = 25`0. The bulkregion ΩB is defined as z ∈ [15, 25]`0. The total num-ber of fluid particles is 7424, and the reflecting wall istypically placed at z = 35`0 (this position slightly de-pends on the interaction parameters). The LJ parame-ters are varied in the ranges εLJsolute,wall/ε

LJ0 ∈ [0.5, 1.5]

and σLJsolute,wall/σ

LJ0 ∈ [0.8, 1.5]. Two concentrations

c = 0.15/`30 and 0.04/`30 are considered, where c is thesolute concentration averaged over Ω. Other computa-tional conditions, as well as notations for the referenceparameters, are the same as those described in Sec. II B.

2. NEMD results: velocity profiles

We show in Fig. 4(a) the velocity profiles obtained us-ing the present NEMD method for different solute-wall

7

interaction parameters. As expected the velocity profileis plug-like at a large distance from the wall, while ex-hibiting some structuration close to the interface. Herewe introduce the solute adsorption Γ, defined as

Γ =

∫ ∞0

dz′(c(z′)

c∞− 1

), (18)

which is a measure of the surface excess of solute in thelayer: Γ is positive for an excess surface concentration,and negative for a depletion. In Fig. 4, the two cases ofΓ = 3.9`0 and −0.9`0 are shown. The corresponding LJparameters are (εLJsolute,wall/ε0, σ

LJsolute,wall/`0) = (1.5, 1.5)

and (0.5, 0.8), respectively. The reversal of the velocityprofiles is associated with a sign change of the adsorption:the flow is forward for Γ = 3.9`0 and backward for Γ =−0.9`0.

Figure 5 plots the diffusio-osmotic mobility calculatedfrom the relationship KDO = v∞/(c∞∇xµ) (here shownfor −∇µx = 0.025f0). The horizontal axis is the theo-retical expression for the mobility given in Eq. (8). Itdepends on the local concentration profile data which wemeasure in the simulation, see Fig. 4(b). Clearly all thenumerical values drop on the line of slope equal to unity,validating the theoretical prediction in a wide parameterrange. We note that we used the value of η calculatedfrom pressure driven flow simulations. One may ques-tion whether it is pertinent to use this value to modelthe flow in the vicinity of the surface, where structuringof the fluid occurs, see Fig. 4(b). However the simulationdata show that this provides a fairly accurate predictionfor the diffusio-osmotic mobility, using the concentrationprofile (measured in the equilibrium situation) as an in-put.

We finally note that the theoretical predictions alsoallow to calculate the local velocity profiles in terms of theconcentration profile, given in Eq. (7). Here, the integralin Eq. (7) is performed using the concentration profile asshown in Figs. 4(b); the integration range is truncated atz/`0 = 8`0, after the concentration converges to the bulkvalue. The comparison is shown in Fig. 4(a) as solid lines,showing again an excellent agreement with the simulationdata.

3. Comparison of mobilities with equilibrium Green–Kuboestimates

As a final check, one can compare the previous valuesfor the mobilities with those obtained from the Green–Kubo relationships in Eq. (14) and Eq. (16). One keydifference is that the latter are now evaluated in equilib-rium simulations.

The calculated correlation functions are displayed inFig. 6(a). The time integration appearing in Eqs. (14)and (16) suffers from significant noise, and we thereforetake an average over a very large time-series sample tocompute the time-correlation functions. We accordingly

(a)

v x

× 0 /τ 0

z/ 0

.063∇xµ = 0 .125 f0:−

:

.025:

(b)

c

× −30

z/ 0

0

0.2

0.4

0.6

0.8

1

1.2

1.4

0 5 10 15 20 25

:

Γ = −0.9 0

Γ = 3.9 0

:

-0.01

0

0.01

0.02

0 5 10 15 20 25

Γ = −0.9 0

Γ = 3.9 0

FIG. 4. Velocity profiles of the diffusio-osmotic flow for thecase of positive surface excess Γ = 3.9`0 and negative surfaceexcess Γ = −0.9`0. The average concentration is c = 0.15/`30.The symbols indicate the MD results, and the solid line inpanel (a) indicates the theoretical result given in Eq. (7) wherewe integrated the concentration profiles shown in panel (b).In panel (b), the concentration profile of ∇xµ = 0 is alsoplotted (black) in addition to the cases of −∇xµ = 0.025f0,0.063f0, and 0.125f0, though the difference is negligible.

adopt the same strategy as in Refs. 26,32, i.e., we per-form ten independent MD simulation runs with differentinitial configurations, and average the time-correlationfunctions over the different samples and time-series. Thecorrelations up to the time difference t = 1000τ0 aretaken, and 4.8 × 106 time-series samples are averagedfor each of ten runs.

Then the diffusio-osmotic mobility MQJ and the recip-rocal counterpart MJQ are obtained by using Eqs. (14)and (16). Here, we truncate the integration range att = 150 τ0 – after a sufficient decay of the correlationfunctions – to avoid unnecessary noise. For the exampleshown in Fig. 6(a), one can check that the two mobil-ities, calculated using the two correlation functions, domatch within the numerical error, i.e., MQJ = MJQ =0.12 ± 0.005 (`0/f0τ0) for the case of Γ = 3.9`0, and−0.035± 0.005 (`0/f0τ0) for the case of Γ = −0.9`0.

Finally, we show in Figs. 6(b) and (c) the comparisonof the NEMD results (symbols) with the results of theGreen–Kubo approach (lines). We apply various valuesof the chemical potential gradient −∇xµ (tuning Fµ),

8

× 0 0τ0

× 0

f0τ0

–K

DO

c∞

-0.1

0

0.1

0.2

-0.1 -0.05 0 0.05 0.1 0.15

c = 0 .15

c = 0 .04

:

:

0

3

03

¡KDO c∞theo

FIG. 5. Numerical values of the diffusio-osmotic mobility−KDOc∞ obtained using the NEMD method versus its the-oretical counterpart −Ktheo

DO c∞ from Eq. (8). At each pointat least four simulation runs have been performed and theaverage value is plotted, with the error bar indicating thestandard deviation. The line indicating a slope equal to unitycorresponds to the theoretical prediction.

and the measured flux Q is plotted in panel (b). A goodagreement is obtained, which validates the direct imple-mentation of the diffusio-osmotic flow using the presentNEMD method. In panel (c), we also compare the resultsto the symmetric estimate of the mobility in terms of theexcess solute flux under an imposed pressure gradient.The measured solute flux Js − c∗∞Q is plotted for vari-ous values of applied pressure drop −∇xP (tuning Fp).Again we find good agreement with the Green–Kubo re-sults.

E. Application to the water-ethanol mixture

We finally demonstrate the versatility of the NEMDmethod by applying it to more realistic systems. Herewe keep the same geometry as shown in Fig. 3(a), butreplace the fluid with an aqueous ethanol solution, andthe wall with a silica surface (Fig. 7(a)) or a graphenesheet (Fig. 7(b)). We use the TIP4P/2005 model forwater molecules,33 and the united atom model of the op-timized potentials for liquid simulations (OPLS)34,35 forthe ethanol molecules. The model detailed in Ref. 36 isemployed for the silica surface, and the interaction pa-rameters for the carbon atoms of the wall are extractedfrom the AMBER96 force field.37 The Lorentz–Berthelotmixing rules38 are used to determine the LJ parametersfor the cross-interactions. The temperature is kept at300 K, using the Nose–Hoover thermostat for all direc-tion, and the pressure is at 1 atm. The time step is set to2 fs. The external force is applied to each atom individ-ually, and the value of the force per atom is obtained bydividing the force per molecule by the number of atomswithin a molecule.

Here we restrict ourselves to the case of high concen-

(a)

(b)

(c)

J s−

c∗ ∞Q

−∇xP

×10−3 20τ0

−∇xµQ

× 0/τ 0

×f 0

×10 3

10−30τ

20

t/ τ0

f030

-0.01

0

0.01

0.02

0.03

0 0.05 0.1 0.15

-0.1

0

0.1

0.2

0 0.5 1 1.5

-1

0

1

2

3

4

0.01 0.1 1 10 100 1000

Γ = −0.9 0

Γ = 3.9 0

Q(t)(Js− c∗∞Q)(0):

(Js− c∗∞Q)(t)Q(0):

FIG. 6. (a) Time correlation functions appearing in Eqs. (14)and (16), obtained using equilibrium MD simulations, for thecase of c = 0.15/`30. The results of ten simulation runs withdifferent initial configurations are averaged, and the standarderror is shown with the error bar. (b) Total flux Q versus thechemical potential gradient −∇xµ. (c) Solute flux Js − c∗∞Qversus the pressure gradient −∇xP . In panels (b) and (c),the symbols indicate the results of NEMD simulations, andthe slopes of the lines indicate the coefficients obtained usingEqs. (14) and (16).

tration, i.e., 20 % ethanol molar fraction, correspondingto 40 wt% ethanol. The lateral dimension of the simula-tion box is 4×4.3 nm2, and the height of the domain Ω isH = 4.8 nm for the case of the silica surface, and 6.3 nmfor the case of the graphene surface. The thickness of thebulk region ΩB is z ∈ [H − 2 nm, H].

As in the insets of Figs. 7(a) and (b), the pressuredriven flow shows no velocity slip on the silica surface,and a large slip on the graphene surface.39 By fitting

9

(c)

(d)

solvent: water 80%

surface: graphene

solute: ethanol 20%

z

xy

(b)

0

10

20

30

40

0 1 2 3 4 5 6z (nm)

vx

(m/s

)

−∇xP = 18 (J/mol nm4)

solvent: water 80%

surface: silica

solute: ethanol 20%

z

xy

: unit

(a)

vx

(m/s

)

z (nm)

−∇xP = 8.9×102 (J/mol nm4)

0

5

10

0 1 2 3 4

): −∇xµ = 2.09 (J/mol nm

: 1.05

: 0.52

z (nm)

vx

(m/s

)

z (nm)

vx

(m/s

)

: −∇xµ = 2.09×103 (J/mol nm)

: 1.05×103

: 0.52×103

(e)

–∇xµ (J/mol nm)

vx

(m/s

)

: graphene (b = 285 nm)

: silica (b = 0 nm)

10-6

10-4

10-2

100

102

104

10-2

100

102

104

0

1

2

0 1 2 3 4

0

1

2

3

0 1 2 3 4 5 6

FIG. 7. Illustrations of systems of water-ethanol mixture in contact with (a) silica surface and (b) graphene surface. Thevelocity profiles under a pressure gradient are also shown in each panel (circles), together with the continuum model (solidline); the z coordinate is measured from the position of Si atoms for the silica surface, and from the C atoms for the graphenesurface. The velocity profiles of the diffusio-osmotic flow are shown for the case of silica surface in (c), for the case of graphenesurface in (d). The symbols indicate the MD results, and the solid lines indicate the theoretical results given in Eq. (7). Theslip length is assumed to be 0 in panel (c) and 285 nm in panel (d). (e) Comparison of the diffusio-osmotic velocity obtainedusing Eq. (8). The solid line indicates the case of the graphene surface, and the dash-dotted line indicates the case of the silicasurface. The dots indicate the points shown in panels (c) and (d).

the formula based on the classical continuum theory,vx = (−∇xP/2η)(2Hb + 2Hz − z2), the slip length forthe graphene surface is estimated as b = 285 nm (seealso Ref. 39). In Figs. 7(c) and (d), the diffusio-osmoticflow profiles obtained by the present NEMD are plot-ted. The flow velocity still shows some noise in spite ofthe relatively large averages at least over 100 ns (5× 107

time steps). Nevertheless, the diffusio-osmotic flows aredirectly observed. The theoretical predictions given inEq. (7) are also shown in the figure, which exhibit rea-sonable agreement with the NEMD data. The applicabil-ity of the present NEMD method to a realistic system isthus confirmed. We note that the inverse diffusio-osmoticflow, which has been reported recently for the system ofaqueous ethanol solution with a silica surface,9 was notobserved in the parameter range we considered here.

We finally emphasize that the large slip length for thecase of the graphene surface is accounted for by correctingEq. (7) as remarked in Sec. III A (see also Refs. 16,25.)The magnitude of the diffusio-osmotic flow is comparedin Fig. 7(e), in which v∞ is plotted as a function of−∇xµ,using Eq. (8); the results corresponding to Fig. 7(c) and(d) are indicated by the dots. The diffusio-osmotic flowon the graphene surface is larger than that on the silicasurface by about three orders of magnitude. This indi-cates that the hydrodynamic slip enormously enhancesthe diffusio-osmotic flow, as expected theoretically, seeRef. 25.

IV. SUMMARY

Transports of fluid mixtures under chemical potentialdifference have been investigated numerically by meansof MD simulations. We first considered osmosis acrossmembranes, and examined the reflection coefficient ofimperfectly semi-permeable membranes. The theoreticalexpression given in Eq. (3), which we derived for high so-lute concentrations, was numerically validated. Next weconsidered the diffusio-osmotic flow near a solid-liquidinterface. We introduced a novel NEMD method allow-ing to simulate a chemical potential gradient, involv-ing a mixed force balance acting on solute and solventmolecules, as illustrated in Fig. 3(d). This method al-lows us to simulate a diffusio-osmotic flow using periodicboundary conditions. We validated the methodology onthe basis of linear response theory and numerical calcu-lations of the corresponding Green–Kubo expressions ofthe transport coefficients. Using the proposed NEMDmethod, the plug-like velocity profile was directly ob-tained, as shown in Figs. 4 and 7, both for the LJ fluidsand water-ethanol solutions. These results showed verygood agreement with the analytical predictions for boththe local velocity profile and mobility.16

The proposed methodology can be extended to explorediffusio-phoretic transport involving complex molecules,like polymers, which has not been explored theoreticallyup to now. Further work in this direction is in progress.

10

ACKNOWLEDGMENTS

L.B. thanks fruitful discussions with B. Rotenberg,P. Warren, M. Cates and D. Frenkel on these topics.This work was granted access to the HPC resourcesof MesoPSL financed by the Region Ile de France andthe project Equip@Meso (reference ANR-10-EQPX-29-01) of the programme Investissements d’Avenir super-vised by the Agence Nationale de la Recherche (ANR).L.B. acknowledges support from the European Union’sFP7 Framework Programme/ERC Advanced Grant Mi-cromegas. S.M. acknowledges funding from a J.-P.Aguilar grant. We acknowledge funding from ANRproject BlueEnergy.

1O. Kedem and A. Katchalsky, “Thermodynamic analysis ofthe permeability of biological membranes to non-electrolytes,”Biochim. Biophys. Acta 27, 229–246 (1958).

2S. S. Sablani, M. F. A. Goosen, R. Al-Belushi, and M. Wilf,“Concentration polarization in ultrafiltration and reverse osmo-sis: a critical review,” Desalination 141, 269–289 (2001).

3B. Abecassis, C. Cottin-Bizonne, C. Ybert, A. Ajdari, andL. Bocquet, “Boosting migration of large particles by solute con-trasts,” Nature Mat. 7, 785–789 (2008).

4V. Yadav, H. Zhang, R. Pavlick, and A. Sen, “Triggered “on/off”micropumps and colloidal photodiode,” J. Am. Chem. Soc. 134,15688–15691 (2012).

5C. Lee, C. Cottin-Bizonne, A.-L. Biance, P. Joseph, L. Bocquet,and C. Ybert, “Osmotic flow through fully permeable nanochan-nels,” Phys. Rev. Lett. 112, 244501 (2014).

6Y.-X. Shen, P. O. Saboe, I. T. Sines, M. Erbakan, and M. Kumar,“Biomimetic membranes: a review,” J. Membrane Sci. 454, 359–381 (2014).

7S. Shin, E. Um, B. Sabass, J. T. Ault, M. Rahimi, P. B. Warren,and H. A. Stone, “Size-dependent control of colloid transport viasolute gradients in dead-end channels,” Proc. Natl. Acad. Sci.113, 257–261 (2016).

8S. Marbach and L. Bocquet, “Active osmotic exchanger for ef-ficient nanofiltration inspired by the kidney,” Phys. Rev. X 6,031008 (2016).

9C. Lee, C. Cottin-Bizonne, R. Fulcrand, L. Joly, and C. Ybert,“Nanoscale dynamics versus surface interactions: What dictatesosmotic transport?” J. Phys. Chem. Lett. 8, 478–483 (2017).

10A. Siria, P. Poncharal, A.-L. Biance, R. Fulcrand, X. Blase, S. T.Purcell, and L. Bocquet, “Giant osmotic energy conversion mea-sured in a single transmembrane boron nitride nanotube,” Nature494, 455–458 (2013).

11W. J. van Egmond, M. Saakes, S. Porada, T. Meuwissen, C. J. N.Buisman, and H. V. M. Hamelers, “The concentration gradientflow battery as electricity storage system: Technology potentialand energy dissipation,” J. Power Sources 325, 129–139 (2016).

12O. Kedem and A. Katchalsky, “A physical interpretation ofthe phenomenological coefficients of membrane permeability,” J.Gen. Physiol. 45, 143–179 (1961).

13A. Katchalsky and O. Kedem, “Thermodynamics of flow pro-cesses in biological systems,” Biophys. J. 2, 53–78 (1962).

14K. S. Spiegler and O. Kedem, “Thermodynamics of hyperfiltra-tion (reverse osmosis): criteria for efficient membranes,” Desali-nation 1, 311–326 (1966).

15G. S. Manning, “Binary diffusion and bulk flow through apotential-energy profile: A kinetic basis for the thermodynamicequations of flow through membranes,” J. Chem. Phys. 49, 2668–2675 (1968).

16S. Marbach, H. Yoshida, and L. Bocquet, “Osmotic and diffusio-osmotic flow generation at high solute concentration. I. Mechan-ical approaches,” (2017), submitted to J. Chem. Phys.

17L. Bocquet and E. Charlaix, “Nanofluidics, from bulk to inter-faces,” Chem. Soc. Rev. 39, 1073–1095 (2010).

18J. L. Anderson, M. E. Lowell, and D. C. Prieve, “Motion of a par-ticle generated by chemical gradients Part 1. Non-electrolytes,”J. Fluid Mech. 117, 107–121 (1982).

19J. L. Anderson, “Colloid transport by interfacial forces,” Ann.Rev. Fluid Mech. 21, 61–99 (1989).

20J. L. Talen and A. J. Staverman, “Osmometry with membranespermeable to solvent and solute,” Trans. Faraday Soc. 61, 2794–2799 (1965); “Negative reflection coefficients,” 61, 2800–2804(1965).

21J. L. Anderson and D. M. Malone, “Mechanism of osmotic flowin porous membranes,” Biophys. J. 14, 957 (1974).

22T. W. Lion and R. J. Allen, “Osmosis in a minimal model sys-tem,” J. Chem. Phys. 137, 244911 (2012).

23See http://lammps.sandia.gov for the code.24J. N. Israelachvili, Intermolecular and surface forces 3rd Edition

(Academic press, 2011).25A. Ajdari and L. Bocquet, “Giant amplification of interfacially

driven transport by hydrodynamic slip: Diffusio-osmosis and be-yond,” Phys. Rev. Lett. 96, 186102 (2006).

26H. Yoshida, H. Mizuno, T. Kinjo, H. Washizu, and J.-L. Bar-rat, “Molecular dynamics simulation of electrokinetic flow of anaqueous electrolyte solution in nanochannels,” J. Chem. Phys.140, 214701 (2014).

27R. Ganti, Y. Liu, and D. Frenkel, “Molecular simulation ofthermo-osmotic slip,” arXiv preprint arXiv:1702.02499 (2017).

28D. K. Bhattacharya and G. C. Lie, “Molecular-dynamics simu-lations of nonequilibrium heat and momentum transport in verydilute gases,” Phys. Rev. Lett. 62, 897 (1989).

29B. D. Todd and D. J. Evans, “Temperature profile for Poiseuilleflow,” Phys. Rev. E 55, 2800 (1997).

30J.-L. Barrat and L. Bocquet, “Large slip effect at a nonwettingfluid-solid interface,” Phys. Rev. Lett. 82, 4671 (1999).

31H. Yoshida and L. Bocquet, “Labyrinthine water flow across mul-tilayer graphene-based membranes: molecular dynamics versuscontinuum predictions,” J. Chem. Phys. 144, 234701 (2016).

32H. Yoshida, H. Mizuno, T. Kinjo, H. Washizu, and J.-L. Barrat,“Generic transport coefficients of a confined electrolyte solution,”Phys. Rev. E 90, 052113 (2014).

33J. L. F. Abascal and C. Vega, “A general purpose model for thecondensed phases of water: TIP4P/2005,” J. Chem. Phys. 123,234505 (2005).

34W. L. Jorgensen, J. D. Madura, and C. J. Swenson, “Optimizedintermolecular potential functions for liquid hydrocarbons,” J. A106, 6638–6646 (1984).

35W. L. Jorgensen, “Optimized intermolecular potential functionsfor liquid alcohols,” J. Phys. Chem. 90, 1276–1284 (1986).

36S. H. Lee and P. J. Rossky, “A comparison of the structure anddynamics of liquid water at hydrophobic and hydrophilic surfaces– a molecular dynamics simulation study,” J. Chem. Phys. 100,3334–3345 (1994).

37W. D. Cornell, P. Cieplak, C. I. Bayly, I. R. Gould, K. M. Merz,D. M. Ferguson, D. C. Spellmeyer, T. Fox, J. W. Caldwell, andP. A. Kollman, “A second generation force field for the simulationof proteins, nucleic acids, and organic molecules,” J. Am. Chem.Soc. 117, 5179–5197 (1995).

38M. P. Allen and D. J. Tildesley, Computer Simulation of Liquids(Oxford Univ. Press, Oxford, 1989).

39K. Falk, F. Sedlmeier, L. Joly, R. R. Netz, and L. Bocquet, “Ul-tralow liquid/solid friction in carbon nanotubes: comprehensivetheory for alcohols, alkanes, OMCTS, and water,” Langmuir 28,14261–14272 (2012).