8
Water Flow inside Polamide Reverse Osmosis Membranes: A Non- Equilibrium Molecular Dynamics Study Yang Song, Fang Xu, Mingjie Wei,* and Yong Wang* State Key Laboratory of Materials-Oriented Chemical Engineering, Jiangsu National Synergetic Innovation Center for Advanced Materials, and College of Chemical Engineering, Nanjing Tech University, Nanjing 210009, Jiangsu, P. R. China * S Supporting Information ABSTRACT: Water ow inside polyamide (PA) reverse osmosis (RO) membranes is studied by steady state nonequilibrium molecular dynamics (NEMD) simulations in this work. The PA RO membrane is constructed with the all-atom model, and the density and average pore size obtained thereby are consistent with the latest experimental results. To obtain the time-independent water ux, a steady state NEMD method is used under various pressure drops. The water ux in our simulations, which is calculated under higher pressure drops, is in a linear relation with the pressure drops. Hence, the water ux in lower pressure drops can be reliably estimated, which could be compared with the experimental results. The molecular details of water owing inside the membrane are considered. The radial distribution function and residence time of water around various groups of polyamide are introduced to analyze the water velocities around these groups, and we nd that water molecules ow faster around benzene rings than around carboxyl or amino groups in the membrane, which implies that the main resistance of mass transport of water molecules comes from the carboxyl or amino groups inside the membranes. This nding is in good consistency with experimental results and suggests that less free carboxyl or amino groups should be generated inside RO membranes to enhance water permeance. 1. INTRODUCTION The freshwater scarcity has become a worldwide problem due to the rapidly growing population, energy demand, and industrialization. 1 In many regions of the world, desalination is an increasingly common solution to supply potable water. Among all desalination technologies, reverse osmosis (RO) technology is the most internationally widespread technology, due to its less stucost, lower energy consumption, and smaller equipment size. 2 In this technology, the performance of membranes predominantly determines the eciency of this process. For most commercial RO membranes, which are mostly interfacially polymerized polyamide (PA) from m- phenylenediamine (MPD) and trimesoyl chloride (TMC), 3 a salt rejection as high as >99% is typically reached, which meets the demand for the seawater or wastewater desalination. However, the water ux remains insucient so that the net cost of RO water is not as low as expected. 4 High permeability for water and excellent exclusion of solutes are important for selective permeable membranes, which both evaluate the performance of RO membranes. Owning higher performance of membranes could not only reduce equipment sizes but also improve the eciency of the membrane separation process. Therefore, developing RO membranes with high water permeability without sacricing salt rejection is greatly demanded. 5 In the process of developing advanced RO membranes, the state-of-the-art PA RO membranes have achieved overwhelming success since the 1970s. 4 From that time, a signicant development has been witnessed in terms of materials synthesis as well as fabrication processes. 6 To further enhance the water ux of PA RO membranes, it is important to understand the molecular mechanism of water transport through the membrane. However, the conventional exper- imental methods could hardly detect behaviors of uid molecules at a microscopic scale, especially in the nanoscale connement. As a consequence of this, the theory of mass transport in RO membranes is far from elucidated. 7 Nowadays, with the improvement of the computational capabilities, computer simulations are widely used to observe these behaviors of nanoscale uids so as to study the membranesperformance. 8 As a classical method of computer simulations, the molecular dynamics (MD) simulation oers a useful way to investigate the behaviors of water molecules inside nanoconned channels. 9 Kotelyanskiia et al. investigated the statistical and dynamic properties of water molecules in an atomic PA membrane. 10 They found the rate of water diusion depended signicantly on waterwater hydrogen bonding interactions. Ding et al. were also focused on the molecular details of water molecules inside the PA RO membranes and found that the hydrogen bond number per molecule and mean square displacements of water molecules signicantly decreased inside the membrane with respect to the bulk phase due to the extreme connement. 3,11 These molecular understandings from Received: November 16, 2016 Revised: January 4, 2017 Published: January 30, 2017 Article pubs.acs.org/JPCB © 2017 American Chemical Society 1715 DOI: 10.1021/acs.jpcb.6b11536 J. Phys. Chem. B 2017, 121, 17151722

Water Flow inside Polamide Reverse Osmosis Membranes: A

  • Upload
    others

  • View
    6

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Water Flow inside Polamide Reverse Osmosis Membranes: A

Water Flow inside Polamide Reverse Osmosis Membranes: A Non-Equilibrium Molecular Dynamics StudyYang Song, Fang Xu, Mingjie Wei,* and Yong Wang*

State Key Laboratory of Materials-Oriented Chemical Engineering, Jiangsu National Synergetic Innovation Center for AdvancedMaterials, and College of Chemical Engineering, Nanjing Tech University, Nanjing 210009, Jiangsu, P. R. China

*S Supporting Information

ABSTRACT: Water flow inside polyamide (PA) reverse osmosis (RO) membranes isstudied by steady state nonequilibrium molecular dynamics (NEMD) simulations in thiswork. The PA RO membrane is constructed with the all-atom model, and the densityand average pore size obtained thereby are consistent with the latest experimental results.To obtain the time-independent water flux, a steady state NEMD method is used undervarious pressure drops. The water flux in our simulations, which is calculated underhigher pressure drops, is in a linear relation with the pressure drops. Hence, the waterflux in lower pressure drops can be reliably estimated, which could be compared with theexperimental results. The molecular details of water flowing inside the membrane areconsidered. The radial distribution function and residence time of water around variousgroups of polyamide are introduced to analyze the water velocities around these groups,and we find that water molecules flow faster around benzene rings than around carboxylor amino groups in the membrane, which implies that the main resistance of masstransport of water molecules comes from the carboxyl or amino groups inside themembranes. This finding is in good consistency with experimental results and suggests that less free carboxyl or amino groupsshould be generated inside RO membranes to enhance water permeance.

1. INTRODUCTION

The freshwater scarcity has become a worldwide problem dueto the rapidly growing population, energy demand, andindustrialization.1 In many regions of the world, desalinationis an increasingly common solution to supply potable water.Among all desalination technologies, reverse osmosis (RO)technology is the most internationally widespread technology,due to its less stuff cost, lower energy consumption, and smallerequipment size.2 In this technology, the performance ofmembranes predominantly determines the efficiency of thisprocess. For most commercial RO membranes, which aremostly interfacially polymerized polyamide (PA) from m-phenylenediamine (MPD) and trimesoyl chloride (TMC),3 asalt rejection as high as >99% is typically reached, which meetsthe demand for the seawater or wastewater desalination.However, the water flux remains insufficient so that the net costof RO water is not as low as expected.4 High permeability forwater and excellent exclusion of solutes are important forselective permeable membranes, which both evaluate theperformance of RO membranes. Owning higher performanceof membranes could not only reduce equipment sizes but alsoimprove the efficiency of the membrane separation process.Therefore, developing RO membranes with high waterpermeability without sacrificing salt rejection is greatlydemanded.5 In the process of developing advanced ROmembranes, the state-of-the-art PA RO membranes haveachieved overwhelming success since the 1970s.4 From thattime, a significant development has been witnessed in terms of

materials synthesis as well as fabrication processes.6 To furtherenhance the water flux of PA RO membranes, it is important tounderstand the molecular mechanism of water transportthrough the membrane. However, the conventional exper-imental methods could hardly detect behaviors of fluidmolecules at a microscopic scale, especially in the nanoscaleconfinement. As a consequence of this, the theory of masstransport in RO membranes is far from elucidated.7

Nowadays, with the improvement of the computationalcapabilities, computer simulations are widely used to observethese behaviors of nanoscale fluids so as to study themembranes’ performance.8 As a classical method of computersimulations, the molecular dynamics (MD) simulation offers auseful way to investigate the behaviors of water moleculesinside nanoconfined channels.9 Kotelyanskiia et al. investigatedthe statistical and dynamic properties of water molecules in anatomic PA membrane.10 They found the rate of water diffusiondepended significantly on water−water hydrogen bondinginteractions. Ding et al. were also focused on the moleculardetails of water molecules inside the PA RO membranes andfound that the hydrogen bond number per molecule and meansquare displacements of water molecules significantly decreasedinside the membrane with respect to the bulk phase due to theextreme confinement.3,11 These molecular understandings from

Received: November 16, 2016Revised: January 4, 2017Published: January 30, 2017

Article

pubs.acs.org/JPCB

© 2017 American Chemical Society 1715 DOI: 10.1021/acs.jpcb.6b11536J. Phys. Chem. B 2017, 121, 1715−1722

Page 2: Water Flow inside Polamide Reverse Osmosis Membranes: A

MD simulations help to reveal the behavior of water moleculesinside the PA membranes.12 However, in the RO process, themovement of water molecules is forced by pressure drops,13,14

so this is a typical nonequilibrium condition. Correspondingly,the statistical mechanical linear response theory used in anequilibrium condition may not faithfully reflect water flowingthrough membrane pores in the real RO process.15 Comparedto the equilibrium MD simulations, the NEMD simulations candirectly obtain the macroscopic water flux from the moleculartrajectories of the simulations.11,16,17 Thus, the NEMD couldtruly simulate the RO process. Ding et al.7 and Shen et al.13

used this method to study the water flux in RO membranes.Shen et al.13 applied forces on rigid sheets outside the waterreservoir to produce a pressure drop. However, this processcannot be maintained in a steady state due to the graduallydecreasing water molecules in the water reservoir during theprocess of the solution passing through PA membranes. Forthis reason, the water flux they calculated in small time scale(10−8 s) cannot be expanded to the experimental time scale(several hours). Hence, their nonsteady state simulations arehard to reliably predict the flux of fluids in the experimentaltime scale.14

In most simulation works on mass transport of ROmembranes, the system was usually built up by consideringboth the hydrated membrane and outside water or solutionreservoirs.3,7,13,18,19 Such a built system is advantageous wheninvestigating the gating effect and simulating the wholemembrane separation process. However, this system is notreliable to analyze the transport resistance. Generally, the masstransport resistance comes from two parts of the RO process,namely, the water entrance into and exit from the membrane(interfacial resistance) and the water transport inside themembrane (intramembrane resistance). In all-atom MDsimulations, it is practically impossible to build up themembrane system with the thickness at the experimentalscale (>100 nm20). Usually, the water flux is estimated byscaling the simulated flux with the membrane thickness.13 Inthis scaling estimation, the interfacial resistance is alsomultiplied for the systems with water reservoirs, which isapparently not reliable, as interfacial resistance is independentof the membrane thickness. In addition, the system consideringalso outside water reservoirs does not allow the effect ofpressure drops on the two specific transport resistances to bedistinguished.In this work, we are focused on understanding the main

resistance for the water transport inside RO membranes. Weapply a steady state nonequilibrium dynamics simulation (SS-NEMD) to study the behavior of water molecules flowinginside the RO membranes, by directly applying external forceon the water molecules, since there are no water reservoirsoutside the membranes. One of the advantages of this SS-NEMD lies in the fact that once the simulations reach thesteady state the statistic and macroscopic properties of waterremain unchanged regardless of the simulation time. During thesimulations, once the steady states are reached, we examinehow the pressure drop affects the macroscopic water flux.Furthermore, the effect of functional groups on water transportinside PA RO membranes is also revealed, which has not beenreported yet.

2. METHODS2.1. Force Field. We performed MD simulations with the

LAMMPS program.15 The atomic interaction was modeled

with the all-atom optimized potentials for liquid simulations(OPLS-AA) force field.21−25 The OPLS-AA potential iscomposed of nonbonded interactions and intramolecularinteractions. Intramolecular interactions are described bybond vibration, bond angle, and torsion angle potentials.

= + + +U U U U UOPLS nb bond ang tor (1)

Nonbonded interactions Unb consist of a sum of standard 12-6 Lennard-Jones (LJ) and electrostatic potentials21−25

ϵσ σ

= + = − +⎡

⎣⎢⎢⎛⎝⎜⎜

⎞⎠⎟⎟

⎛⎝⎜⎜

⎞⎠⎟⎟

⎦⎥⎥U r U U

r rk

q q

r( ) 4ij ij

ij

ij

ij

ij

i j

ijnb LJ coul

12 6

coul

(2)

where ϵij is the LJ energy, σij is the LJ diameter for atoms i andj, qi and qj are their partial charges, and kcoul represents theelectrostatic constant. Mixing rules are used: ϵij = (ϵiϵj)

1/2 andσij = (σiσj)

1/2, when atoms are different species. Nonbondedinteractions are calculated between all atom pairs of differentmolecules in addition to all pairs on the same moleculeseparated by three or more bonds, though the interaction isreduced by a factor of 1/2 for atoms separated by three bonds.The cutoff distance of all LJ interactions is 10 Å. Allelectrostatic interactions are calculated in real space, if atompairs are closer than 10 Å. However, outside this range, theinteractions are calculated in reciprocal space by using theEwald algorithm with a precision of 10−4. The potential forbond stretching is presented by eq 3

= −U r k r r( ) ( )ij r ijbond 02

(3)

where rij represents the distance between atoms i and j and r0 isthe equilibrium separation. The harmonic form of the anglestretching potential is the following:

θ = −θU k r r( ) ( )ijk ijang 02

(4)

The i, j, and k represent three atoms. θijk is the angle of thevectors rij and rjk, and θ0 is the equilibrium value. At last, thetorsional potential in OPLS is given by

∑ϕ ϕ= − −=

Uk

n( )2

[1 ( 1) cos( )]n

n ntor

1

4

(5)

in this equation, ϕ is the dihedral angle, and kn is the energy.26

2.2. Construction of the RO Membrane. In this work, weconstructed an all-atom model. The main advantage of theatomistic model is that it describes the complex macro-molecular system in a detailed all-atom representation andprovides a direct quantitative prediction of the polymerstructural and dynamical properties. The chemical structuresof monomers of PA considered in our simulations, TMC andMPD, as well as the repeat unit of PA are shown in Figure 1a.Each PA chain is set to have 50 repeat units, as shown in Figure1a, and the units are linked with each other by forming amidebonds through the reaction between COOH and NH2 groups.The constructed PA chains are therefore terminated with aCOOH group on one end and NH2 group on the other. Ninesuch PA chains are packed in the simulation box, and at thevery beginning, the PA chains are almost linear. To obtain therandom structure of PA and to minimize the energy of the PAmembrane, a simulation was performed with a thermostat at atemperature of 1000 K for 1 ns and then set to 300 K in a box

The Journal of Physical Chemistry B Article

DOI: 10.1021/acs.jpcb.6b11536J. Phys. Chem. B 2017, 121, 1715−1722

1716

Page 3: Water Flow inside Polamide Reverse Osmosis Membranes: A

with a three-dimensional periodic boundary, in which Lx = 8nm, Ly = 9 nm, and Lz = 15 nm.After that, we constructed the hydrated PA membranes. Ding

et al. constructed the hydrated PA membranes by an NEMDsimulation,19 by which the process simulated the experimentalmethod. Unlike their work, we constructed the hydrated PAmembranes by adding water molecules directly into thesimulation box, since we needed to construct only a hydratedmembrane without water reservoirs. As the water content in thehydrated membranes is experimentally measured to 23 wt %,27

we randomly inserted 1200 water molecules into the simulationbox. After the initial minimization, the simulations wereperformed 2 ns in a canonical ensemble (NVT), in which thetemperature was set to 300 K. In addition, the system pressurewas increased from 1 to 100 bar using an isothermal−isobaricensemble (NPT). The purpose of this step is to compress thesystem sufficiently to render it ready for the method ofshrinking boxes (T = 300 K and P = 100 bar). The finaldimensions of the orthorhombic cell of the final hydratedmembranes were Lx = 4 nm, Ly = 4.5 nm, and Lz = 7.5 nm, asshown in Figure 1b.

2.3. Details of SS-NEMD Simulations. To investigatewater molecules flowing inside the membrane, the periodicboundary conditions were employed in all three directions. Thepressure drops in the simulations were 2 orders of magnitudehigher than experimental values, since a small force led to a lowsignal-to-noise ratio, such that extremely long sampling wouldbe needed to measure the streaming motion with a smalluncertainty, which resulted in the large computational cost. Asfluid atoms could not distinguish between a situation where theflow is driven by a pressure or by an external field,28 the forceor gravity field could represent the pressure drop in the NEMDsimulations. The pressure drops inside the membrane can berelated to the constant applied force by ΔP = P1 − P2 = nFz/Aon each water molecule in the z-direction.29 After simulations,the water flux in the membrane was directly obtained bycounting numbers of water molecules though a particular planeacross the flowing direction.To simulate the driven flow, Frentrup et al.,30 Huang et al.,31

and Muscatello et al.32 achieved a pressure gradient by applyinga force to particles in a thin slab some distance away from themembrane. In this manner, the particles with dynamicsperturbed by the external force are not close to the system ofinterest, i.e., the region close to the membrane. However, as weare focused on water flow inside the membrane in this work, wehave to construct the hydrated PA membranes without theoutside water reservoirs; thus, there is only the membrane andits hydrations. Therefore, following a frequently usedmethod,29,33 we applied the external force directly to thewater molecules. A set of external forces (in the z direction ofthe simulation boxes) was added on water molecules so as tosimulate the pressure drops (180, 270, 360, 450, 540, and 630MPa) across the membranes, which could drive the waterflowing through the RO membranes. In the SS-NEMDsimulations, a microcanonical ensemble (NVE) was performedon water molecules with the NVT on polyamides, so that watermolecules could move freely from being interrupted bythermostats, which comes with the NVT ensemble, and thatthe extra heat generated by the external force can be removedcompletely by the polymer. The temperature of watermolecules remains around 300 K no matter what with thepressure drops ranging from 0 to 630 MPa (shown in FigureS2). In addition, to avoid the polyamide chains flowing withwater molecules in the z direction, we pinned 5% of themembrane to fixed points during the simulations. During thefirst 3 ns (or sooner) of each simulation, the system was goingto reach the steady state and the data collected in this periodwere not used for further analysis. The whole data acquisition

Figure 1. (a) The chemical equation of polymerization of MPD andTMC to produce polyamide. (b) Snapshot of the final configuration ofhydrated PA. The atoms of PA are marked with the blue, and watermolecules are represented by red (O) and white (H), respectively.

Figure 2. (a) Change of PWF with time through the RO membrane at ΔP = 270 MPa. (b) PWF through the RO membrane as a function ofpressure drops.

The Journal of Physical Chemistry B Article

DOI: 10.1021/acs.jpcb.6b11536J. Phys. Chem. B 2017, 121, 1715−1722

1717

Page 4: Water Flow inside Polamide Reverse Osmosis Membranes: A

stage lasted 10 ns, and atomic trajectories were recorded every10000 steps with a time step of 1 fs to examine the structuraland dynamic properties of water in the RO membrane.

3. RESULTS AND DISCUSSION3.1. Pure Water Flux. In this work, the pure water flux

(PWF) is calculated as water molecule flow rate per membranearea. Six pressure drops are applied on water molecules toinvestigate the effect of pressure drops on the transportproperties. The water flux at ΔP = 270 MPa is shown in Figure2a. In this case, the points of water flux recorded by thetrajectories are oscillated around the line y = 37.22, whichindicates a water flux of 37.22 molecules/ns/nm2. Since there isno obvious tendency of continuous increase or decrease in thePWF, we believe the simulation is in a steady state.The PWFs of the other five pressure drops are calculated the

same as in the case of ΔP = 270 MPa. Figure 2b shows thewater flux as a function of pressure drops. It is obvious that thewater flux increases almost linearly with pressure drops despitesmall fluctuations due to the random thermal motion. Thepressure drops in our simulations are much higher than those inexperimental RO processes. This is because it is practicallyimpossible to investigate the water flux at a relatively lowpressure drop within a few nanoseconds. However, it is worthnoting that the fitted line is zeroaxial which indicates that adoubled pressure drop will approximately lead to a doubledPWF. Hence, the water flux at low pressure drops can bereliably predicted. If ΔP = 5 MPa, the PWF will be calculated as0.687 molecules/ns/nm2, which is a number of magnitudeshigher than those obtained in experiments (0.001−0.004molecules/ns/nm2).13

There is another issue which will take effect on the PWF forthe RO process. Shen et al.13 pointed out that experimentallymeasured macroscale water flux is inversely proportional withthe thickness of the PA selective layer. Usually, the thickness ofthe PA selective layer is around 200 nm.34 Accordingly, theequivalent experimental water flux at 5 MPa with a membranethickness of 7.5 nm (Lz) should be in the range 0.026−0.107molecules/ns/nm2. From Figure 2b, the estimated water flux at5 MPa is about 0.687 molecules/ns/nm2. The value isobviously higher than the experimental result because weonly consider water flow inside the membrane. The transportresistance is actually composed of several processes, includingthe concentration gradient in the fluid bulk phase, the waterentering membranes, the one inside membrane pores, etc. Inthis work, we are focused on the transport behaviors of waterinside RO membranes and try to find out the main factor oftransport resistance for water molecules through the PA ROmembranes. The process of water molecules entering into ROmembranes and the concentration gradient are not consideredhere. Since the process of water molecules flowing into themembrane will undoubtedly consume a significant amount ofenergy due to the molecules colliding with the membrane, thisprocess would directly decline the velocity of water molecules.3.2. Physical Properties of the Membrane. Differing

from the carbon nanotubes which have well-defined hollowchannels with permanent shapes, pores in RO membranes aredynamic and ever changing while water molecules flowinginside the membrane, and few researchers use SS-NEMD toanalyze the behavior of water flowing through RO mem-branes.35 It is important to understand the physical propertiesof the RO membrane, as they determine the dynamicproperties of water molecules in it.

Figure 3 shows the average density profiles across thehydrated membrane. It is obvious that the density of hydrated

membrane fluctuates around 1.3 g/cm3, which is in goodagreement with the experimental results.36 The density of thehydrated membrane is almost constant in different positions inthe z direction of simulation boxes, and the oscillation within areasonable range is the influence of the presence of voids insidethe membrane. Figure 3 also depicts the average density profilesof water molecules inside the membrane. The total density ofthe water is around 0.38 g/cm3. By comparing the density ofwater and hydrated membrane (about 0.38 g/cm3 of water per1.3 g/cm3 hydrated polyamide),18 the pore volume fraction(Φ) can be calculated to be 0.3. In addition, the density of bulkwater ∼1 g/cm3 or unexpected high density cannot be observedin Figure 3. This predicts no macroscopic phase separationbetween polymer and water molecules.To further investigate the existence of macroscopic phase

separation and the free volume size of the membranes, theaverage radius of pores is calculated by the water−water radialdistribution function g(r) inside the membrane. Figure 4adescribes the radial distribution function g(r) among watermolecules, the tendency of which is consistent with the work ofDing et al.3 We continued our analysis of the membranes byinvestigating the pore size. The effective pore size can beestimated by eq 6 according to Debye37 with an ideal randomporous structure

ξ− ∝ −g r

rln[ ( ) 1] const

(6)

where ξ is the correlation length that determines thecharacteristic size of the minority phase, i.e., water-filled spacesin the membranes.18 Figure 4b represents the correlationfunction derived from Figure 4a as defined by eq 6.Then, the oscillated slope can be roughly fitted into a linear

slope after the first peak, from which ξ = 1.5 was calculated. Toconnect the correlation length ξ and the pore volume fractionΦ to the average pore area per unit volume (S/V), the Debyetheory follows.37

ξ= Φ − ΦS

V4 (1 )

(7)

From the previous calculations, S/V = 5.6 nm−1 is estimated. Ifthe pores are close to cylinders, spheres, and slit-like shapes, theaverage pore radii are 0.36 nm (2V/S), 0.54 nm (3V/S), and0.18 nm (V/S), respectively. Therefore, the pore diameter is inthe range 0.36−1.08 nm, which is in good consistency with the

Figure 3. Profile of the average mass density of water and hydratedmembrane along the z-direction of the RO membrane.

The Journal of Physical Chemistry B Article

DOI: 10.1021/acs.jpcb.6b11536J. Phys. Chem. B 2017, 121, 1715−1722

1718

Page 5: Water Flow inside Polamide Reverse Osmosis Membranes: A

experimental values.38 In all cases of water transport, the flux isalmost linear with the square of the pore radius.39 That is whythe pore size is an essential parameter we have to take intoaccount. In this work, the diameter of the pores is in the rangeof a couple of water molecules (about 0.32 nm for a singleone), so that most water molecules will collide with the porewall composed of various functional groups of PA membranes.Hence, the interaction between water molecules and the PAmembranes can be expected to dominate the water moleculesflowing through the pores due to the slippery theory of masstransport.40,41

3.3. Interactions between Water Molecules and theFunctional Groups in PA Membranes. To reveal theinteractions between water molecules and functional groups,the radial distribution function, g(r), was explored. The radialdistribution function between water molecules and functionalgroups of polyamide is shown in Figure 5a. Oxygen atoms ofcarboxyl groups, nitrogen atoms of amino groups, and carbonatoms of benzene rings were considered as characteristic atomswhich were presented as functional groups of polyamide. Theradial distribution function here could be defined as the localdensity of water molecules around the functional groupsdivided by the bulk density of water inside the membrane. InFigure 5a, r presents the distance between water molecules andfunctional groups and g(r) represents the degree of aggregation.For the pair of oxygen atoms in water molecules and oxygenatoms of carboxyl groups (OW−OPA), the curve begins to risewhen r = 0.23 nm, which means that the minimum distancebetween water and carboxyl groups in the membrane is 0.23nm. At r = 0.26 nm, there is a sharp peak. The intensity of thepeak is 2.7, implying that the water density here is roughly 1.7

times higher than the average density in the membrane. From r= 0.4 nm, g(r) begins to oscillate around 1, which presents thatthis functional group hardly has any affinity to water in thatrange. For the pair of oxygen atoms in water molecules andnitrogen atoms in amino groups (OW−NPA), the minimumdistance between the water and amino groups is 0.25 nm. Theintensity of a small peak is 1.3 at r = 0.28 nm, and the aminogroups have little affinity to water when r is beyond 0.5 nm. Incontrast, the curve of the OW−CPA (the pair of oxygen atoms inwater and carbon atoms of benzenes) is not like the two curvesdiscussed above, and one can hardly observe any small peak,which indicates that water molecules are randomly locatedaround benzene rings in a short distance.The shape of the hydration shell is obviously different from

one functional group to the other. As shown in Figure 5a, watermolecules have the highest affinity to carboxyl, followed byamino groups, and then benzene rings. It is mostly possible thatwater molecules are more likely to attach to the carboxyl oramino groups rather than benzene rings when water moleculesare moving inside the membrane. Consequently, watermolecules around the carboxyl or amino groups might havelower velocities than those around benzene rings.To further confirm this speculation, the residence time is

calculated from time correlation functions42 to investigatevelocities of water around functional groups. The dynamicalproperties of a fluid are conveniently described throughconsideration of a time correlation function R(t), which canbe cast in the form

Figure 4. (a) Radial distribution function calculated for water oxygenatoms within polyamide. (b) Correlation function derived from part aas defined by eq 6.

Figure 5. (a) Radial distribution function of oxygen atoms of water(OW) with oxygen atoms of carboxyl groups (OPA), nitrogen atoms ofamino groups (NPA), and carbon atoms of benzene rings (CPA). (b)Residence correlation functions for water molecules in the firsthydration shell of carboxyl groups, amino groups, and benzene rings(ΔP = 270 MPa).

The Journal of Physical Chemistry B Article

DOI: 10.1021/acs.jpcb.6b11536J. Phys. Chem. B 2017, 121, 1715−1722

1719

Page 6: Water Flow inside Polamide Reverse Osmosis Membranes: A

∑ ∑ θ θ=Ω

+= =

⎡⎣⎢⎢

⎤⎦⎥⎥R t

N Nt t t( )

1 1 1( ) ( )

t

t

i

N

i it 1 occu 1

0 0

0

0max

(8a)

where t0 is the initial time. t01 and t0

max are the first and lastimmediate times, respectively, and the moments are bothselected in 1 ns after the system being in steady state. Nt is thetotal number of time origins, and Noccu is the number of watermolecules in the confined space, which is averaged over all ofthe time origins. Ω in the above equation is the normalizationconstant given by

∑ ∑ θ θΩ == =N

t t1

( ) ( )t

t

i

N

i it 1 1

0 0

0

0max

(8b)

From the radial distribution function, we know that the firstcoordination shells of carboxyl groups, amino groups, andbenzene rings are 2.7, 2.8, and 3.7 Å, respectively. If watermolecules selected lie within the first coordination shell at bothtime steps, the θi(t) takes 1. Otherwise, it takes 0.Figure 5b shows the residence time distribution for water

molecules around carboxyl groups, amino groups, and benzenerings (OW−OPA, OW−NPA, OW−CPA, respectively). R(t)represents the percentage of initial selected water moleculesremaining inside the shell. As seen in Figure 5b, all three curvesdecay very rapidly at the beginning and then follow by a slowerdecay. Compared to the curves of OW−OPA and OW−NPA,OW−CPA drops faster. All sets of residence times of wateraround the three distinct groups of PA are listed in Table 1.

The water around benzene rings has a shorter residence timethan the other two, which means that benzene rings have theweakest affinity to water and water molecules around benzenerings flow with the fastest velocity in the three cases. Thesesimulation findings are nicely supported by experimentalworks.43,44 Wang et al. used 3,5-diamino-N-(4-aminophenyl)benzamide (DABA) to partially replace MPD to react withTMC in the interfacial polymerization of PA RO membranes.With the ratio of DABA to MPD increased from 0 to 12.5%,the water permeance of the produced RO membranes wasenhanced evidently from 37.5 to 55.4 L/m2/h without anysacrifice of salt rejection.43 Each DABA molecule carries threeprimary amino groups and two benzene rings, whereas eachMPD molecule carries two amino groups and one benzene ring.That is, DABA has a higher ratio of benzene to amino groupsthan MPD, and RO membranes prepared from larger dosagesof DABA have a higher concentration of benzene ringscompared to amino and carboxyl groups. Therefore, aspredicted by our simulation, PA RO membranes containingmore benzene rings would reduce the mass transport resistanceof water inside the membranes. Similarly, they also obtainedenhanced permeance by partially replacing TMC withisophthaloyl dichloride (IPC) which has a higher ratio ofbenzene to acyl chloride as a result of the increased

concentration of benzene in the membranes.44 Therefore, assupported by the experimental results, our simulation suggeststhat more benzene rings help to enhance the water transportinside the membrane.To obtain the residence time with different pressure drops,

we compute the integral for the corresponding residencecorrelation functions. Interestingly, with increasing pressuredrops, the residence time of water molecules near all threetypes of groups drops obviously. However, unlike the rule ofPWF with pressure drops, the tendency of the residence time isnot linear any more. This suggests that the effect of thehydrophilicity of groups on PWF is not proportional. After all,it is still evident that the residence time of water near carboxylor amino groups is always longer than those near benzene ringsindependent of the pressure drops. Therefore, the results ofresidence time confirm that the benzene rings of ROmembranes have lower transport resistance to water molecules.This is in good consistency with previous works reporting thatcarbon materials, such as carbon nanotubes45−48 andgraphene49−51 which have hydrophobic surfaces, exhibit fasterwater transport than hydrophilic surfaces.

4. CONCLUSIONSIn this work, the SS-NEMD method is conducted to investigatewater molecules flowing inside the RO membrane. By thismethod, a set of the steady state pure water flux is measured. Asexpected, the pure water flux increases linearly with thepressure drop. From our simulations of higher pressure drop(around 100 MPa), the values of fluid flux in the condition oflower pressure drop (less than 5 MPa) could be reliablydetermined, so as to compare the simulation results withexperimental data. In addition, the molecular details of hydratedmembranes are revealed. The physical density of the hydratedmembranes is measured to be 1.3 g/cm3, and the averaged poresize is estimated to be 0.6−1.22 nm. Both are consistent withexperimental data. We find that the residence time of watermolecules around benzene rings is obviously shorter than thosearound carboxyl or amino groups in the membrane. The shortresidence time indicates the faster flow of water molecules, sowater molecules move faster around benzene rings of themembranes, which is also confirmed by the experimental workon hydrophobic carbons. These findings are expected toincrease our understanding on the water transport behavior inRO membranes and also the rational design of advanced ROmembranes with upgraded permeance.

■ ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acs.jpcb.6b11536.

PWF of water molecules as a function of simulation time,which could confirm the achievement of steady state; thetemperature of the water molecules was also provided tomake sure the extra heat, which came from the externalforces, was completely removed by the PA atoms (PDF)

■ AUTHOR INFORMATIONCorresponding Authors*E-mail: [email protected].*E-mail: [email protected] Wei: 0000-0001-7601-4749

Table 1. Residence Time of Water near the Different Groupsof PA under Various Pressure Drops

residence time (ps)

functional groups ΔP = 180 270 360 450 540 630 MPa

carboxyl groups 66.1 51.9 39.2 36.0 33.2 31.1amino groups 68.9 48.5 41.4 34.4 33.6 32.8benzene rings 56.5 41.0 33.3 28.5 27.5 26.3

The Journal of Physical Chemistry B Article

DOI: 10.1021/acs.jpcb.6b11536J. Phys. Chem. B 2017, 121, 1715−1722

1720

Page 7: Water Flow inside Polamide Reverse Osmosis Membranes: A

Yong Wang: 0000-0002-8653-514XNotesThe authors declare no competing financial interest.

■ ACKNOWLEDGMENTSFinancial support from the National Basic Research Program ofChina (2015CB655301), the National Natural ScienceFoundation of China (21506091), the Jiangsu Natural ScienceFoundations (BK20150944, BK20150063), the Special Pro-gram for Applied Research on Super Computation of theNSFC-Guangdong Joint Fund (the second phase), and theProject of Priority Academic Program Development of JiangsuHigher Education Institutions (PAPD) is gratefully acknowl-edged.

■ REFERENCES(1) Shannon, M. A.; Bohn, P. W.; Elimelech, M.; Georgiadis, J. G.;Marinas, B. J.; Mayes, A. M. Science and Technology for WaterPurification in the Coming Decades. Nature 2008, 452, 301−310.(2) Drioli, E.; Stankiewicz, A. I.; Macedonio, F. MembraneEngineering in Process Intensificationan Overview. J. Membr. Sci.2011, 380, 1−8.(3) Ding, M.; Ghoufi, A.; Szymczyk, A. Molecular Simulations ofPolyamide Reverse Osmosis Membranes. Desalination 2014, 343, 48−53.(4) Lee, K. P.; Arnot, T. C.; Mattia, D. A Review of Reverse OsmosisMembrane Materials for DesalinationDevelopment to Date andFuture Potential. J. Membr. Sci. 2011, 370, 1−22.(5) Zhao, K.; Wu, H. Size Effects of Pore Density and Solute Size onWater Osmosis through Nanoporous Membrane. J. Phys. Chem. B2012, 116, 13459−13466.(6) Shenvi, S. S.; Isloor, A. M.; Ismail, A. F. A Review on RoMembrane Technology: Developments and Challenges. Desalination2015, 368, 10−26.(7) Ding, M.; Szymczyk, A.; Ghoufi, A. On the Structure andRejection of Ions by a Polyamide Membrane in Pressure-DrivenMolecular Dynamics Simulations. Desalination 2015, 368, 76−80.(8) Araki, T.; et al. Molecular Dynamics Study of CarbonNanotubes/Polyamide Reverse Osmosis Membranes: Polymerization,Structure, and Hydration. ACS Appl. Mater. Interfaces 2015, 7, 24566−24575.(9) Renou, R.; Szymczyk, A.; Ghoufi, A. Water Confinement inNanoporous Silica Materials. J. Chem. Phys. 2014, 140, 044704.(10) Kotelyanskii, M. J.; Wagner, N. J.; Paulaitis, M. E. MolecularDynamics Simulation Study of the Mechanisms of Water Diffusion in aHydrated, Amorphous Polyamide. Comput. Theor. Polym. Sci. 1999, 9,301−306.(11) Ding, M.; Szymczyk, A.; Goujon, F.; Soldera, A.; Ghoufi, A.Structure and Dynamics of Water Confined in a Polyamide Reverse-Osmosis Membrane: A Molecular-Simulation Study. J. Membr. Sci.2014, 458, 236−244.(12) Wei, T.; Zhang, L.; Zhao, H.; Ma, H.; Sajib, M. S. J.; Jiang, H.;Murad, S. Aromatic Polyamide Reverse-Osmosis Membrane: AnAtomistic Molecular Dynamics Simulation. J. Phys. Chem. B 2016, 120,10311−10318.(13) Shen, M.; Keten, S.; Lueptow, R. M. Dynamics of Water andSolute Transport in Polymeric Reverse Osmosis Membranes ViaMolecular Dynamics Simulations. J. Membr. Sci. 2016, 506, 95−108.(14) Arya, G.; Chang, H.-C.; Maginn, E. J. A Critical Comparison ofEquilibrium, Non-Equilibrium and Boundary-Driven MolecularDynamics Techniques for Studying Transport in MicroporousMaterials. J. Chem. Phys. 2001, 115, 8112−8124.(15) Plimpton, S. Fast Parallel Algorithms for Short-Range MolecularDynamics. J. Comput. Phys. 1995, 117, 1−19.(16) Zwanzig, R. Time-Correlation Functions and TransportCoefficients in Statistical Mechanics. Annu. Rev. Phys. Chem. 1965,16, 67−102.

(17) Luo, Y.; Harder, E.; Faibish, R. S.; Roux, B. ComputerSimulations of Water Flux and Salt Permeability of the ReverseOsmosis FT-30 Aromatic Polyamide Membrane. J. Membr. Sci. 2011,384, 1−9.(18) Kolev, V.; Freger, V. Hydration, Porosity and Water Dynamicsin the Polyamide Layer Of reverse Osmosis Membranes: A MolecularDynamics Study. Polymer 2014, 55, 1420−1426.(19) Ding, M.; Szymczyk, A.; Ghoufi, A. Hydration of a PolyamideReverse-Osmosis Membrane. J. Membr. Sci. 2016, 501, 248−253.(20) Li, D.; Yan, Y.; Wang, H. Recent Advances in Polymer andPolymer Composite Membranes for Reverse and Forward OsmosisProcesses. Prog. Polym. Sci. 2016, 61, 104−155.(21) Jorgensen, W. L.; Madura, J. D.; Swenson, C. J. OptimizedIntermolecular Potential Functions for Liquid Hydrocarbons. J. Am.Chem. Soc. 1984, 106, 6638−6646.(22) Jorgensen, W. L.; Maxwell, D. S.; TiradoRives, J. Developmentand Testing of the OPLS All-Atom Force Field on ConformationalEnergetics and Properties of Organic Liquids. J. Am. Chem. Soc. 1996,118, 11225−11236.(23) Damm, W.; Frontera, A.; TiradoRives, J.; Jorgensen, W. L.OPLS All-Atom Force Field for Carbohydrates. J. Comput. Chem.1997, 18, 1955−1970.(24) Rizzo, R. C.; Jorgensen, W. L. OPLS All-Atom Model forAmines: Resolution of the Amine Hydration Problem. J. Am. Chem.Soc. 1999, 121, 4827−4836.(25) Price, M. L. P.; Ostrovsky, D.; Jorgensen, W. L. Gas-Phase andLiquid-State Properties of Esters, Nitriles, and Nitro Compounds withthe OPLS-Aa Force Field. J. Comput. Chem. 2001, 22, 1340−1352.(26) Watkins, E. K.; Jorgensen, W. L. Perfluoroalkanes: Conforma-tional Analysis and Liquid-State Properties from Ab Initio and MonteCarlo Calculations. J. Phys. Chem. A 2001, 105, 4118−4125.(27) Kotelyanskii, M. J.; Wagner, N. J.; Paulaitis, M. E. AtomisticSimulation of Water and Salt Transport in the Reverse OsmosisMembrane FT-30. J. Membr. Sci. 1998, 139, 1−16.(28) Todd, B. D.; Evans, D. J.; Daivis, P. J. Pressure Tensor forInhomogeneous Fluids. Phys. Rev. E: Stat. Phys., Plasmas, Fluids, Relat.Interdiscip. Top. 1995, 52, 1627−1638.(29) Zhu, F.; Tajkhorshid, E.; Schulten, K. Pressure-Induced WaterTransport in Membrane Channels Studied by Molecular Dynamics.Biophys. J. 2002, 83, 154−160.(30) Frentrup, H.; Avendano, C.; Horsch, M.; Salih, A.; Muller, E. A.Transport Diffusivities of Fluids in Nanopores by Non-EquilibriumMolecular Dynamics Simulation. Mol. Simul. 2012, 38, 540−553.(31) Huang, C.; Choi, P. Y. K.; Kostiuk, L. W. A Method for Creatinga Non-Equilibrium Nt(P1 - P2) Ensemble in Molecular DynamicsSimulation. Phys. Chem. Chem. Phys. 2011, 13, 20750−20759.(32) Muscatello, J.; Jaeger, F.; Matar, O. K.; Muller, E. A. OptimizingWater Transport through Graphene-Based Membranes: Insights fromNonequilibrium Molecular Dynamics. ACS Appl. Mater. Interfaces2016, 8, 12330−12336.(33) Suk, M. E.; Aluru, N. R. Molecular and ContinuumHydrodynamics in Graphene Nanopores. RSC Adv. 2013, 3, 9365−9372.(34) Malaeb, L.; Ayoub, G. M. Reverse Osmosis Technology forWater Treatment: State of the Art Review. Desalination 2011, 267, 1−8.(35) Wang, L.; Dumont, R. S.; Dickson, J. M. NonequilibriumMolecular Dynamics Simulation of Water Transport through CarbonNanotube Membranes at Low Pressurea. J. Chem. Phys. 2012, 137,044102.(36) Zhang, X.; Cahill, D. G.; Coronell, O.; Marinas, B. J. Absorptionof Water in the Active Layer of Reverse Osmosis Membranes. J.Membr. Sci. 2009, 331, 143−151.(37) Debye, P.; Anderson, H. R.; Brumberger, H. Scattering by anInhomogeneous Solid. Ii. The Correlation Function and ItsApplication. J. Appl. Phys. 1957, 28, 679−683.(38) Singh, P. S.; Ray, P.; Xie, Z.; Hoang, M. Synchrotron Saxs toProbe Cross-Linked Network of Polyamide ‘Reverse Osmosis’ and‘Nanofiltration’ Membranes. J. Membr. Sci. 2012, 421−422, 51−59.

The Journal of Physical Chemistry B Article

DOI: 10.1021/acs.jpcb.6b11536J. Phys. Chem. B 2017, 121, 1715−1722

1721

Page 8: Water Flow inside Polamide Reverse Osmosis Membranes: A

(39) Murad, S.; Nitsche, L. C. The Effect of Thickness, Pore Size andStructure of a Nanomembrane on the Flux and Selectivity in ReverseOsmosis Separations: A Molecular Dynamics Study. Chem. Phys. Lett.2004, 397, 211−215.(40) Day, M. A. The No-Slip Condition of Fluid Dynamics.Erkenntnis 1990, 33, 285−296.(41) Striolo, A. From Interfacial Water to Macroscopic Observables:A Review. Adsorpt. Sci. Technol. 2011, 29, 211−258.(42) Choudhury, N.; Pettitt, B. M. Dynamics of Water Trappedbetween Hydrophobic Solutes. J. Phys. Chem. B 2005, 109, 6422−6429.(43) Wang, H. F.; Li, L.; Zhang, X. S.; Zhang, S. B. Polyamide Thin-Film Composite Membranes Prepared from a Novel Triamine 3,5-Diamino-N-(4-Aminophenyl)-Benzamide Monomer and M-Phenyl-enediamine. J. Membr. Sci. 2010, 353, 78−84.(44) Zhao, Y.; Zhang, Z.; Dai, L.; Mao, H.; Zhang, S. Enhanced BothWater Flux and Salt Rejection of Reverse Osmosis Membrane throughCombining Isophthaloyl Dichloride with Biphenyl Tetraacyl Chlorideas Organic Phase Monomer for Seawater Desalination. J. Membr. Sci.2017, 522, 175−182.(45) Whitby, M.; Quirke, N. Fluid Flow in Carbon Nanotubes andNanopipes. Nat. Nanotechnol. 2007, 2, 87−94.(46) Joseph, S.; Aluru, N. R. Why Are Carbon Nanotubes FastTransporters of Water. Nano Lett. 2008, 8, 452−458.(47) Kannam, S. K.; Todd, B. D.; Hansen, J. S.; Daivis, P. J. How FastDoes Water Flow in Carbon Nanotubes. J. Chem. Phys. 2013, 138,094701.(48) Walther, J. H.; Ritos, K.; Cruz-Chu, E. R.; Megaridis, C. M.;Koumoutsakos, P. Barriers to Superfast Water Transport in CarbonNanotube Membranes. Nano Lett. 2013, 13, 1910−1914.(49) Huang, H.; Song, Z.; Wei, N.; Shi, L.; Mao, Y.; Ying, Y.; Sun, L.;Xu, Z.; Peng, X. Ultrafast Viscous Water Flow through Nanostrand-Channelled Graphene Oxide Membranes. Nat. Commun. 2013, 4,2979.(50) Wei, N.; Peng, X.; Xu, Z. Understanding Water Permeation inGraphene Oxide Membranes. ACS Appl. Mater. Interfaces 2014, 6,5877−5883.(51) Jiao, S.; Xu, Z. Selective Gas Diffusion in Graphene OxidesMembranes: A Molecular Dynamics Simulations Study. ACS Appl.Mater. Interfaces 2015, 7, 9052−9059.

The Journal of Physical Chemistry B Article

DOI: 10.1021/acs.jpcb.6b11536J. Phys. Chem. B 2017, 121, 1715−1722

1722