Molecular Simulations of Binary Gas Mixture Transport and Separation

Aug 14, 2018 - Computational Screening of Metal–Organic Frameworks for Membrane-Based CO2/N2/H2O Separations: Best Materials for Flue Gas ...
0 downloads 0 Views 2MB Size
Subscriber access provided by University of Missouri-Columbia

C: Surfaces, Interfaces, Porous Materials, and Catalysis

Molecular Simulations of Binary Gas Mixture Transport and Separation in Slit Nanopores Tianhao Wu, and Abbas Firoozabadi J. Phys. Chem. C, Just Accepted Manuscript • DOI: 10.1021/acs.jpcc.8b04976 • Publication Date (Web): 14 Aug 2018 Downloaded from http://pubs.acs.org on August 19, 2018

Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.

is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.

Page 1 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

1

Molecular Simulations of Binary Gas Mixture Transport

2

and Separation in Slit Nanopores

3

Tianhao Wu† and Abbas Firoozabadi*,†

4



Reservoir Engineering Research Institute, 595 Lytton Avenue Suite B, Palo Alto, CA 94301, USA.

5

*

Corresponding author; Tel: 650-326-9172; Fax: 650-472-9285; E-mail: [email protected]

6

7

8

1

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

9

Page 2 of 29

ABSTRACT

10

We propose a new method to simulate the gas mixture transport and separation in slit nanopores. The

11

method is based on the random removal of molecules from the permeate side in the dual control volume-

12

grand canonical molecular dynamics (DCV-GCMD) method. Each step in the ramom removal is

13

independent of the previoud steps.The conventional method, DCV-GCMD, simulates the gas mixture in

14

control volumes based on the chemical potential of each component, in which the gas compositions have

15

to be known in advance and kept constant. However, the transport process affects the composition of the

16

produced gas mixture due to selective adsorption in the nanopores. This process has been modeled

17

iteratively in the literature. We propose an alternative method to calculate the composition in the

18

permeate side directly; in our approach, the computational efficiency is improved by an order of

19

magnitude compared to the iterative method. Transport of gas mixtures of CH4/He and CO2/CH4 are

20

investigated in graphene, graphite, and Tetrahedral-Octahedral-Tetrahedral (TOT) structure slit pores.

21

Pore width is the dominant factor in the species separation. The solid substrates and surface roughness

22

have pronounced effect on gas separation. The average pressure may have pronounced effect when the

23

pore width is less than 1.00 nm.

24

2

ACS Paragon Plus Environment

Page 3 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

25

INTRODUCTION

26

Gas mixture transport through nanopores is a classic topic and of great interest. The porous materials,

27

with pore scales ranging from micropores to mesopores, are widely utilized in adsorption and gas

28

separation, including natural gas sweetening, flue gas purification, gas storage, carbon dioxide capture,

29

and shale gas development.1-5 In gas separation, various parameters affect the transport and separation of

30

species, including the molecular size, the pore geometry and connectivity, and the interactions between

31

the molecules and the pore surface.1-2,6 In shale gas formations, due to the nanometer size of pores, there

32

may be a separation of species at high pressure. The investigation of the mechanisms will shed light on

33

the gas transport and separation, and the engineering of the material design accordingly.

34

The grand canonical Monte Carlo (GCMC) simulation has been widely applied to investigate the

35

selectivity based on adsorption for various materials,7-12 including porous carbon materials,4,13

36

zeolites,14-16 and metal-organic frameworks (MOFs).3,17-22 Various methods have also been proposed to

37

simulate the gas transport and separation in nanopores,6,23 including the external force (EF) method, the

38

boundary driven (BD) method, and the dual control volume-grand canonical molecular dynamics (DCV-

39

GCMD) method.24 The EF method performs an external force field on each atom in the simulation box

40

with periodic boundary conditions to achieve an equivalent pressure gradient in an infinite pore.25-26 Ho

41

and Striolo27 have studied the transport of a mixture of methane and water through slit muscovite

42

nanopores. The pressure gradient in the BD method is generated by means of the specific actions on the

43

boundary, such as the moving “piston”,28 the external force within a specific region,29 and the virtual

44

semi-membrane.26,30 The DCV-GCMD method is a hybrid technique of grand canonical Monte Carlo

45

(GCMC) and molecular dynamics (MD) simulations.31-33 It contains two control volumes (CV) at the

46

upstream side (feed side) and the downstream side (permeate side) of the slit pore, with different

3

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 4 of 29

47

pressures or chemical potentials. The simulation is carried out with an alternation of the Monte Carlo

48

(MC) steps and the MD steps based on the MC/MD ratio. Due to the stability of simulation and the

49

similarity to physical processes, the DCV-GCMD is widely used in gas transport simulation.

50

Various studies of single-component flow with DCV-GCMD method have been carried out.29,33-36 The

51

flow rates in 2-nm slit pores have been measured to be 1 to 2 orders of magnitude higher than the

52

predictions by the Knudsen diffusion. Jin and Firoozabadi34 performed the flow of single-component

53

species in slit nanopore to investigate the large disparity between measurements and the prediction from

54

the Knudsen diffusion. They found that the flow of the high density adsorbed layer contributes to flow

55

enhancement. For the multicomponent gas transport, much effort has been devoted to developing

56

effective algorithms and investigating the separation factor of advanced materials. Xu et al.37 performed

57

the binary gas mixture transport through structureless carbon slit pore by the DCV-GCMD simulations.

58

To improve the efficiency for the simulation of alkanes, the configurational-bias MC was also

59

proposed.38-40 These methods have been widely utilized to simulate the gas mixture transport through

60

various materials, such as graphitic slit pores,1-2,23,41-46 carbon nanotubes,47 nanoporous carbons with

61

complex pore-networks,38,48-50 zeolites,51 and silicon-carbide membranes.52-53 In the conventional

62

method, the chemical potential or partial pressure of each component is maintained by the GCMC

63

method. The value of target chemical potential is obtained using equation of state (EOS) based on

64

pressure, gas species, and compositions. One has to assign the compositions of gas mixture in the system

65

before launching the simulation. The compositions are often set the same at both the feed side and the

66

permeate side, which is kept constant in each control volume despite the fact that selectivity of the

67

substrate may lead to different compositions. This assumption will lead to the inconsistency of the flux

68

ratio between the two components in the pore and produced composition ratio in the permeate side.

4

ACS Paragon Plus Environment

Page 5 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

69

Generally, the pressure at the permeate side should be set, but the composition is unknown. In GCMC

70

simulations, the composition in each control volume should be known before launching the simulation

71

to . calculate the chemical potentials. The conventional method, namely DCV-GCMD, cannot be used to

72

simulate the condition of constant pressure with unknown composition in the permeate side for gas

73

mixtures transport. Wang et al.6 have proposed an iteration workflow based on the conventional method.

74

The computational cost of each loop from the iteration method is similar to one complete run of the

75

conventional DCV-GCMD simulation. The computations are time-consuming, and the stability of

76

computations should be managed carefully.6 Phan et al.54 have simulated the H2S/CH4 mixture transport

77

through a membrane containing water. The molecules in the permeate side are removed in different time

78

intervals, which can model the condition close to vacuum. Cabrales-Navarro et al.55 applied a similar

79

method for vacuum condition as well as a dynamic process of gas mixture transport with canonical

80

ensemble. Because the parameters in the two control volumes are changing with time, one cannot

81

determine the boundary conditions for macroscopic analysis. An accurate and efficient method will

82

facilitate the molecular simulation of gas mixture transport and separation in nanopores.

83

In this work, we introduce a method to simulate gas mixture transport and separation in slit nanopores.

84

In Section 2, the molecular model and the methods of gas mixture transport are introduced, including

85

conventional method, iterative method, and our proposed approach. In Section 3, our approach is

86

validated by methane transport with marked components. The results from the conventional method, the

87

iterative method, and the proposed method are compared. In Section 4, we use our proposed method to

88

perform the simulations of CH4/He and CO2/CH4 mixtures transport through slit pores made of graphite,

89

graphene, and Tetrahedral-Octahedral-Tetrahedral (TOT) structures. The effects of various parameters

90

on separation are discussed. The conclusions are drawn in Section 5.

91

METHODOLOGY 5

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 92 4 5 6 7 93 8 9 94 10 11 12 95 13 14 96 15 16 97 17 18 19 98 20 21 99 22 23 100 24 25 101 26 27 28 29 30 31 32 33 34 35 36 102 37 103 38 39 104 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 105 56 57 58 59 60

Page 6 of 29

Model description

A schematic representation of the system of this study is shown in Fig. 1. The simulation box consists of three regions. The H-CV, L-CV and C-regions correspond to the high-pressure CV at the upstream side (feed side), the low-pressure CV at the downstream side (permeate side), and the slit nanopore in the middle, respectively. The periodic boundary condition is applied in the y- and z-directions, while the fixed boundary condition is applied in the x-direction by means of a virtual reflective wall. Flow is along the x-direction. We performed simulations with three different solid structures including graphene, graphite (three sheets), and Tetrahedral-Octahedral-Tetrahedral (TOT) structure from pyrophyllite (see Fig. 2). The Tetrahedral-Octahedral-Tetrahedral (TOT) structure exists in many clay minerals, which can represent the main properties of ideal Si-O surface in clays and construct the symmetric slit pore.

Fig. 1. Sketch of the model configuration used in gas mixture transport in slit nanopores with the dual control volume.

6

ACS Paragon Plus Environment

Page 7 of 29 1 2 3 106 4 107 5 6 108 7 8 109 9 10 11 110 12 13 111 14 15 112 16 17 18 113 19 20 114 21 22 115 23 24 25 26 116 27 28 29 117 30 31 32 33 34 118 35 36 37 119 38 39 120 40 41 42 121 43 44 122 45 46 123 47 124 48 49 125 50 51 126 52 53 127 54 55 128 56 57 58 59 60

The Journal of Physical Chemistry

Fig. 2. Molecular structure of the solid substrates of the slit pores. (a) Graphene; (b) Graphite; (c) Tetrahedral-OctahedralTetrahedral (TOT) structure of clay mineral.

The solid substrates are fixed in the C-region which are parallel to the xy-plane. The pore lengths are 10.00 nm, 30.12 nm, and 90.36 nm in different settings. The pore width ranges from 0.80 nm to 10.00 nm. The widths in the y-direction are 4.47 nm, 4.47 nm, and 5.44 nm for graphene, graphite, and TOT, respectively. For H- and the L-CVs, the domain size in the x- and y-directions are 30.00 nm×4.47 nm, while the domain size in the z-direction will change with the settings for different pore widths. The gas molecules are modeled as the Lennard–Jones (LJ) fluid. The interactions between the gas–gas and the gas-solid particles i and j are modeled with the cut and shifted (CS) LJ 12-6 potential: φij ( rij ) − φij ( rcut ) φijCS =  0

rij ≤ rcut rij > rcut

(1)

where

 σ φij ( rij ) = 4ε ij  ij  rij 

12 6   σ ij    −      rij  

(2)

The cutoff distance rcut is set as 1 nm in this study. The effective LJ size and the energy parameters, σ and ε, are listed in Table 1.44,56 For all the cross-term LJ parameters, the Lorentz–Berthelot mixing rules are employed.

7

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 129 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 130 22 23 24 131 25 26 132 27 28 133 29 30 31 134 32 33 135 34 35 136 36 37 38 137 39 40 41 138 42 43 44 45 139 46 47 140 48 49 50 141 51 52 53 54 142 55 56 57 58 59 60

Page 8 of 29

Table 1. Parameters for LJ (12-6) potential44,56

σ (nm)

ε / kB (K)

CH4

0.381

148.1

He

0.255

10.2

CO2

0.379

225.3

C (graphene/graphite)

0.340

28.0

O (TOT)

0.317

78.2

Si (TOT)

0.395

47.8

Al (TOT)

0.411

32.7

The simulations are performed based on the open-source package LAMMPS,57 with the additional function of the proposed method developed by the authors. The time step for MD simulation ranges from 1 fs to 5 fs depending on the case. We use 25 ns simulation time for the system to reach steadystate, and another 25 ns simulation time to calculate the density and pressure. The simulation time is extended to 50 ns for each process in the slit pore less than 1.00 nm width. The snapshots are taken every 0.5 ps for density and velocity calculation. At every 0.5 ns, the average properties are recorded for the calculation of mean value and standard deviation. The Nose–Hoover thermostat is performed for the entire domain at 298 K except the simulations for examining the effect of temperature.

Conventional and iterative methods

In the conventional DCV-GCMD, the pressure is maintained by performing a sufficient number of MC attempts, including insertions and deletions of molecules. For component i, the probability of acceptance for insertion is given by

 Z V  ∆U   pi+ = min  i exp  −  ,1  k BT    Ni + 1 8

ACS Paragon Plus Environment

(3)

Page 9 of 29 1 2 3 143 4 5 6 144 7 8 145 9 10 146 11 12 147 13 14 15 16 148 17 18 19 149 20 21 22 150 23 24 151 25 26 152 27 28 29 153 30 31 32 33 154 34 35 36 155 37 38 39 156 40 41 157 42 43 158 44 45 159 46 47 48 160 49 50 161 51 52 162 53 54 55 163 56 57 58 59 60

The Journal of Physical Chemistry

where Zi = exp(µi/kBT)/Λi3 is the absolute activity at temperature T, Λi is the de Broglie wavelength, kB is the Boltzmann’s constant, µi is the chemical potential, V is the volume of the CV, Ni is the number of molecules, and ∆U is the potential energy change resulting from inserting or deleting the molecule. When the insertion is accepted, the velocity of the molecule is assigned based on the MaxwellBoltzmann distribution. The probability of acceptance for deletion is given by

 N  ∆U   pi− = min  i exp  −  ,1  k BT    ZiV

(4)

The MC moves in each CV are followed by one MD step. The MC/MD ratio is set to 10 in this study. The iterative method is a modified approach based on the conventional DCV-GCMD.6 It resets the chemical potential in the L-CV based on the flux ratio of the two species in each loop. The iteration process will continue until the flux ratio in the nanopore and the mole ratio in the L-CV converge to the same value.

Proposed method

At many experimental conditions and in many industrial processes, the pressure at the permeate side is not zero. The target chemical potentials in the H-CV can be obtained based on the feed composition and pressure. Then, the chemical potentials in the H-CV can be maintained by the conventional GCMC method. In the H-CV, the molecule exchanges come from the insertion and deletion by MC steps and the net flux between H-CV and C-region. However, in the L-CV, because the composition is unknown, the chemical potential cannot be determined before launching the simulation. As a result, the insertion attempt cannot be performed in the conventional GCMC simulation. The molecule exchanges are from the molecule deletion in MC steps and the net flux between L-CV and C-region. The removal of the molecules in the MC step is the key step in our proposed method. 9

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 164 4 5 6 165 7 8 166 9 10 167 11 12 168 13 14 15 169 16 17 170 18 19 171 20 21 22 172 23 24 25 173 26 27 28 174 29 30 31 175 32 33 176 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 10 of 29

The workflow is shown in Fig. 3. We check for pressure in the L-CV during a given time interval. When the current pressure Pcurrent is higher than the target pressure Ptarget, we remove one molecule randomly at a time to reach the target pressure. In each step the random removal is independent of the previous steps. The target pressure is treated as a threshold value. Every molecule in the L-CV has the same probability to be picked up. However, because a reliable current pressure should be calculated with the information from many time steps, a time window should be assigned for pressure calculation. For the initial period, the time window is set as same as the current simulation time, which is shorter than the assigned value of time window (see Fig. S1). Pcurrent is usually zero in the early period because the initial condition is vacuum. In our approach, we define the separation factor as:

S AB =

cL , A / cL , B cH , A / cH , B

(5)

where c is the molar concentration, the subscript H and L denote the H- and the L-CVs, respectively, and A and B denote Species A and Species B, respectively. Note that in our method in each step the random removal is independent of the previous steps.

10

ACS Paragon Plus Environment

Page 11 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 177 24 25 178 26 27 28 179 29 30 31 32 180 33 34 35 181 36 37 38 182 39 40 183 41 42 184 43 44 45 185 46 47 186 48 49 187 50 51 52 188 53 54 189 55 56 190 57 58 59 60

The Journal of Physical Chemistry

Fig. 3. Pressure control in the permeate side.

METHOD VALIDATION

Simulation with marked component

To validate the proposed method, we performed the simulation without pressure difference between the H-CV and the L-CV first. The simulation was carried out in the graphene slit pore with the length of 30.12 nm and the pore width of 2.00 nm. The flow of single component gas was investigated, while the gas molecules were marked as two different species with the property of methane, namely Species A and Species B. The pressure in the H-CV and the L-CV were set as 40 bar. The feed composition in the H-CV was maintained by GCMC method with 70% Species A and 30% Species B in mole percentage. If the method is valid, the composition in each region should be uniform, and the number density in the H-CV and the L-CV should be the same. The result in Fig. S2(a) is the average composition in each region. The number density in Fig. S2(b) and the mole percentage in Fig. S2(c) are the average values in each cross-section, including the adsorbed layer. The time window for pressure monitoring was set as 5 11

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 191 4 5 6 192 7 8 193 9 10 194 11 12 195 13 14 15 196 16 17 18 19 197 20 21 22 198 23 24 25 199 26 27 200 28 29 201 30 31 32 202 33 34 203 35 36 204 37 38 205 39 40 41 206 42 43 207 44 45 208 46 47 48 209 49 50 210 51 52 211 53 54 55 212 56 57 58 59 60

Page 12 of 29

ps, which should be compatible with the MC/MD ratio. Otherwise, it may lead to a delay response of current pressure or have significant fluctuations. A simulation with pressure difference was also performed in the same configuration as above; the pressure in the L-CV is 30 bar instead of 40 bar. The velocity in the y- and z-directions was taken into account for temperature calculation. The composition in each region should be uniform as well. The results are as expected (see Fig. S3).

Comparison of results from the conventional method and proposed method

The transport of CH4/He mixture was investigated in the graphite slit pore with the length of 30.12 nm and the pore width of 1.00 nm. The pressure is set as 60 bar and 50 bar in the H-CV and the L-CV, respectively. The feed composition of each component was 0.5 in mole fraction for the two methods in the H-CV. In the L-CV, the composition of each component was set as 0.5 in mole fraction in the conventional method as well, while in our method the composition was determined from the simulation. The results are presented in Figs. 4 and 5. In the conventional method, the composition is kept constant in both the H-CV and the L-CV. In our method, the composition in the L-CV shows significant difference from the feed composition due to the separation in the slit nanopore. The solid substrate selectively adsorbs CH4, which has a higher number density and a higher mole percentage than He. In the conventional method, the mole percentage is uniform in each region. In our method, the mole percentage of CH4 in the pore exhibits slight increasing trend along flow direction. Helium has the opposite trend. The molecule number density profiles at the locations of 1/4, 1/2, and 3/4 of the pore length have minor differences as well. The velocity profiles reveal pronounced difference between the two methods. The velocity profiles of CH4 and He are very close to each other in the conventional method, whereas the velocity of He is much higher than CH4 in our method. The concentration 12

ACS Paragon Plus Environment

Page 13 of 29 1 2 3 213 4 5 6 214 7 8 215 9 10 216 11 12 217 13 14 15 218 16 17 219 18 19 220 20 21 22 221 23 24 222 25 26 223 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

difference of He between the feed side and permeate side is underestimated in the conventional method because the mole percentage of He is much lower in the permeate side due to the selectivity of the pore. Therefore, He should have larger driving force for the gas transport, just like in the proposed method (see Fig. 5(b)), which will lead to higher He velocity than CH4. Graphite has an extremely smooth surface, then each component has pronounced slip velocity which will lead to substantial mobility near the solid surface. As a result, the adsorption has significant contribution to the total flux. The separation factor is related to the coupled effect between the two factors: one is the enhanced flux for the preferred component due to adsorption; the other is the higher concentration difference for the less adsorbed component. We also performed the comparative simulations for the two methods in nanopores with the widths of 2.00 nm and 10.00 nm. The results reveal much lower selectivity of 2.00 nm pore and almost no selectivity of 10.00 nm pore (Figs. S4 to S7).

13

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 224 34 35 225 36 226 37 227 38 228 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 14 of 29

Fig. 4. Simulation of the CH4/He gas mixture transport by the conventional method: Graphite slit pores; Pore width = 1.00 nm; Pore length = 30.12 nm; Pressure gradient = 0.33 bar/nm; Pressure = 60 bar in the H-CV; Pressure = 50 bar in the L-CV; Temperature = 298 K. (a) Composition along the x-direction; (b) Velocity in the x-direction at the middle of the slit pore; Molecular number density of CH4 (c) and He (d) along the z-direction at three locations in the pore.

14

ACS Paragon Plus Environment

Page 15 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 229 34 35 230 36 231 37 232 38 233 39 40 41 234 42 43 44 235 45 46 47 236 48 49 237 50 51 238 52 53 54 239 55 56 240 57 58 59 60

The Journal of Physical Chemistry

Fig. 5. Simulation of the CH4/He gas mixture transport by the proposed method: Graphite slit pores; Pore width = 1.00 nm; Pore length = 30.12 nm; Pressure gradient = 0.33 bar/nm; Pressure = 60 bar in the H-CV; Pressure = 50 bar in the L-CV; Temperature = 298 K. (a) Composition along the x-direction; (b) Velocity in the x-direction at the middle of the slit pore; Molecular number density of CH4 (c) and He (d) along the z-direction at three locations in the pore.

Comparison of the iterative method and proposed method

We have compared the performance of the iterative method and our proposed method using the data for example in Fig. 5. The parameters in the iterative method were set the same as the conventional method except for the chemical potentials in the L-CV. In the iterative method, the composition in the permeate side is set the same as in the feed side in the first iteration. Based on the flux, in the subsequent iterations, the composition in the permeate side is updated, and the flux ratio is then computed for the next iteration. The process of iterative method is presented in Fig. 6. The mole ratio in the L-CV (same 15

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 241 4 5 6 242 7 8 243 9 10 244 11 12 245 13 14 15 246 16 17 247 18 19 248 20 21 22 249 23 24 250 25 26 251 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 252 44 45 253 46 254 47 255 48 256 49 50 257 51 52 53 54 55 56 57 58 59 60

Page 16 of 29

as the separation factor SAB) from the proposed method is presented using the dashed line as reference. The flux ratio and mole ratio converge to the same value after 11 iterations. The results from the iterative method agree with the results from our proposed method. We noticed convergence issues in the iterative method. In some iterations, there were negative flux ratios. We cut back the mole ratio in the LCV for the next iteration to reduce the convergence issue (see the second to fourth iterations in Fig. 6). The convergence problems in the iterative method have been discussed by Wang et al.6 We used the workstation with two Intel Xeon 2690 v4 CPUs and two Nvidia P100 GPUs to compare the computational efficiency. It took 71 hours 44 minutes 28 seconds for the entire simulation of the proposed method. One iteration of the iterative method took 78 hours 15 minutes 27 seconds. The number of iteration may vary in different systems. For this particular example, there were 11 iterations to achieve the convergence, then our approach is about 12 times faster than the iterative method.

Fig. 6. Mole ratio of CH4/He in the L-CV and flux ratio in the slit nanopore from the iterative method. The mole ratio in the L-CV (separation factor SAB) from the proposed method is presented using dashed line as reference. Graphite slit pores; Pore width = 1.00 nm; Pore length = 30.12 nm; Pressure gradient = 0.33 bar/nm; Pressure = 60 bar in the H-CV; Pressure = 50 bar in the L-CV; Temperature = 298 K.

APPLICATION AND DISCUSSION

16

ACS Paragon Plus Environment

Page 17 of 29 1 2 3 258 4 5 6 259 7 8 260 9 10 261 11 12 262 13 14 15 263 16 17 264 18 19 265 20 21 22 266 23 24 267 25 26 268 27 28 29 269 30 31 270 32 33 271 34 35 272 36 37 38 273 39 40 274 41 42 275 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Three solid structures, including graphene, graphite, and TOT, were selected for investigation of transport and separation. Graphene and graphite have the same surface structure but different strengths for adsorption. The three layers of graphitic sheets can provide stronger interaction with the gas molecules than in graphene. The TOT surface has a different molecular structure and pairwise interaction which has a rougher surface than graphene and graphite. First, we performed the simulations for the CH4/He mixture in the slit pores with different widths ranging from 0.80 nm to 10.00 nm. The pore length is 30.12 nm. The pressure was set as 60 bar and 50 bar in the H-CV and the L-CV, respectively. The rest of the model configurations were same as in section 3. The separation factor was calculated from Eq. (5). The results are presented in Fig. 7. The separation factor increases significantly with the decrease of pore width, especially below 1.00 nm. When the pore width is larger than 2.00 nm, the material can hardly provide selectivity. The trend of separation factor is mainly due to the coupled effect between the enhanced flux of CH4 due to adsorption and the higher concentration difference of He due to selectivity. The flux from adsorbed layer has much larger contribution to the total flux in the smaller pores than in the larger pores. The effect of higher concentration difference of He is relatively weak when the pore width is very small. The separation factors for the three substrates are close to one in pore widths greater than 5.00 nm, while it is enhanced in pore widths less than 1.00 nm. TOT shows less selectivity than the other two materials. These phenomena are mainly controlled by the effect of substrate surface roughness.

17

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 276 19 277 20 278 21 279 22 23 280 24 25 281 26 27 28 282 29 30 283 31 32 284 33 34 35 285 36 37 286 38 39 287 40 41 42 288 43 44 289 45 46 290 47 48 291 49 50 51 292 52 53 293 54 55 56 57 58 59 60

Page 18 of 29

Fig. 7. Separation factor of the CH4/He mixture vs. pore width in three slit pores: Pore length = 30.12 nm; Pressure gradient = 0.33 bar/nm; Pressure = 60 bar in the H-CV; Pressure = 50 bar in the L-CV; Temperature = 298 K. Simulations based on the proposed method.

The simulations were performed in the slit pores with a width of 1.00 nm and 2.00 nm at different average pressures at the same pressure gradient of 0.33 bar/nm. The results show that the separation factor decreases with the average pressure increase in 1.00 nm pore width (see Fig. 8(a)), due to less contribution at high pressure for flux from adsorbed layer than at low pressure. However, there is almost no change with average pressure in 2.00 nm pore (see Fig. 8(b)). Extensive simulations for 10.00 nm pore width in graphite are also made (see Fig. S8). There is no clear relationship between the separation factor and average pressure in the 10.00 nm pore width. There is almost no separation for the gas mixture. In the pore widths larger than 2.00 nm, the mobility of adsorbed layer cannot provide pronounced contribution to the total flux, so that the separation factors as well as their changes are negligible. The results indicate that the average pressure may have pronounced effect when the pore width is less than 1.00 nm. Graphene and graphite have different macroscopic interaction strengths with gas molecules, but the separation factors are very close. The separation factor of TOT is lower than the other two substrates. The surface roughness has a more pronounced effect than the interaction strength.

18

ACS Paragon Plus Environment

Page 19 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 294 19 20 295 21 296 22 297 23 298 24 299 25 26 27 300 28 29 301 30 31 302 32 33 34 303 35 36 304 37 38 305 39 40 41 306 42 43 307 44 45 308 46 47 309 48 49 50 310 51 52 311 53 54 312 55 56 57 58 59 60

The Journal of Physical Chemistry

Fig. 8. Separation factor of the CH4/He mixture vs. average pressure in three slit pores. (a) Pore width = 1.00 nm; (b) Pore width = 2.00 nm. Pore length = 30.12 nm; Pressure gradient = 0.33 bar/nm; Temperature = 298 K. The pressures in the H-CV and the L-CV are assigned based on the average pressure, pressure gradient, and pore length. Simulations based on the proposed method.

We also carried out the simulations for transport of the CO2/CH4 mixture in the three materials. The pressure gradient was 0.33 bar/nm. For the effect of pore width, the pressures were set as 40 bar and 30 bar in the H-CV and the L-CV, respectively. For the effect of average pressure, the pressures in the HCV and the L-CV were adjusted to maintain the pressure gradient. The rest of the model configurations were the same as the CH4/He gas mixture simulations. Due to the preferred adsorption of CO2 instead of CH4, the slit pores show selectivity for CO2. The separation factors with respect to CO2 in the mixture are presented in Figs. 9 and 10. The results of 1.00 nm and 2.00 nm pore widths on composition, velocity profile, and number density are shown in Figs. S9 and S10. Similar to the CH4/He mixture, the separation factor shows clear trend of the effect of pore width. The reason is related to the coupled effect of the enhanced flux of CO2 due to selective adsorption and the higher concentration difference of CH4 between the two sides due to the selectivity, which is similar to the CH4/He mixture. When the pore width is larger than 5.00 nm, the material cannot provide pronounced selectivity, especially in TOT. The separation factors for graphite and graphene are still very close in pore widths larger than 2.00 nm,

19

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 313 4 5 6 314 7 8 315 9 10 316 11 12 317 13 14 15 318 16 17 18 319 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 320 35 36 321 37 322 38 323 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Page 20 of 29

whereas they show differences in pore widths less than 1.00 nm. In the 1.00 nm pore width, the separation factor presents minor downward trend as average pressure increases. The effect of average pressure is negligible in the 2.00 nm pore width. TOT has weaker selectivity than the other two materials in the whole range. The effects of pore length, temperature, pressure gradient, and feed composition are also investigated. The details and results are provided in the Supporting Information (see Figs. S11 and S12).

Fig. 9. Separation factor of the CO2/CH4 mixture vs. pore width in three slit pores: Pore length = 30.12 nm; Pressure gradient = 0.33 bar/nm; Pressure = 40 bar in the H-CV; Pressure = 30 bar in the L-CV; Temperature = 298 K. Simulations based on the proposed method.

20

ACS Paragon Plus Environment

Page 21 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 324 19 20 325 21 326 22 327 23 328 24 25 329 26 27 28 29 330 30 31 32 331 33 34 332 35 36 333 37 38 334 39 40 41 335 42 43 336 44 45 337 46 47 48 338 49 50 339 51 52 340 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

Fig. 10. Separation factor of the CO2/CH4 mixture vs. average pressure in three slit pores. (a) Pore width = 1.00 nm; (b) Pore width = 2.00 nm. Pore length = 30.12 nm; Pressure gradient = 0.33 bar/nm; Temperature = 298 K. The pressures in the H-CV and the L-CV are assigned based on the average pressure, pressure gradient, and pore length. Simulations based on the proposed method.

CONCLUSIONS

In this work, we have simulated separation of binary gas mixtures in slit nanopores. We introduce a new method based on the dual control volume-grand canonical molecular dynamics (DCV-GCMD) simulations. In our approach we directly compute the gas composition at the permeate side. Our method is based on random removal of molecules from the permeate side to keep the pressure constant. The inconsistency of the species flux ratio in the pore and the mole ratio of the produced fluid in the permeate side is eliminated. We improve the computational efficiency significantly compared to the iterative method. We used the proposed method to investigate flow and separation of binary mixtures of CH4/He and CO2/CH4 in nanopores of various widths and length and three different slit pores. We compared the results from our method and the conventional method, and the iterative method. The results from the iterative methods is the same as our proposed method. Our method is one order of magnitude faster. The main conclusions drawn from this work are:

21

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 341 4 5 6 342 7 8 343 9 10 344 11 12 345 13 14 15 16 346 17 18 19 20 347 21 22 23 24 348 25 26 27 28 349 29 30 31 350 32 33 34 35 351 36 37 38 39 352 40 41 353 42 43 354 44 45 46 355 47 48 356 49 50 357 51 52 53 358 54 55 359 56 57 58 59 60

Page 22 of 29

(1) The composition in the permeate side is significantly different from the feed composition when the pore widths are less than 2.00 nm. (2) There is a significant difference between the computed velocity of species in the nanopores from the conventional DCV-GCMD simulation and our method. (3) The separation in molecularly non-smooth nanopores is less than molecularly smooth nanopores.

SUPPORTING INFORMATION

Supporting information is available online at https://pubs.acs.org/.

ACKNOWLEDGEMENTS

The work was supported by the member companies of the Reservoir Engineering Research Institute (RERI). Their support is appreciated.

REFERENCES

(1) Xu, L.; Tsotsis, T. T.; Sahimi, M. Nonequilibrium Molecular Dynamics Simulation of Transport and Separation of Gases in Carbon Nanopores. I. Basic Results. J. Chem. Phys. 1999, 111, 3252–3264. (2) Xu, L.; Sedigh, M. G.; Tsotsis, T. T.; Sahimi, M. Nonequilibrium Molecular Dynamics Simulation of Transport and Separation of Gases in Carbon Nanopores. II. Binary and Ternary Mixtures and Comparison with the Experimental Data. J. Chem. Phys. 2000, 112, 910–922. (3) Yazaydın, A. O.; Benin, A. I.; Faheem, S. A.; Jakubczak, P.; Low, J. J.; Willis, R. R.; Snurr, R. Q. Enhanced CO2 Adsorption in Metal-Organic Frameworks Via Occupation of Open-Metal Sites by Coordinated Water Molecules. Chem. Mater. 2009, 21, 1425–1430. 22

ACS Paragon Plus Environment

Page 23 of 29 1 2 3 360 4 5 6 361 7 8 362 9 10 363 11 12 364 13 14 15 365 16 17 366 18 19 367 20 21 22 368 23 24 369 25 26 370 27 28 29 371 30 31 372 32 33 373 34 35 374 36 37 38 375 39 40 376 41 42 377 43 44 45 378 46 47 379 48 49 380 50 51 52 381 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(4) Liu, Y.; Wilcox, J. Molecular Simulation Studies of CO2 Adsorption by Carbon Model Compounds for Carbon Capture and Sequestration Applications. Environmental Science & Technology 2012, 47, 95– 101. (5) Venna, S. R.; Carreon, M. A. Highly Permeable Zeolite Imidazolate Framework-8 Membranes for CO2/CH4 Separation. J. Am. Chem. Soc. 2009, 132, 76–78. (6) Wang, S.-M.; Yu, Y.-X.; Gao, G.-H. Grand Canonical Monte Carlo and Non-Equilibrium Molecular Dynamics Simulation Study on the Selective Adsorption and Fluxes of Oxygen/Nitrogen Gas Mixtures through Carbon Membranes. Journal of Membrane Science 2006, 271, 140–150. (7) Thornton, A. W.; Dubbeldam, D.; Liu, M. S.; Ladewig, B. P.; Hill, A. J.; Hill, M. R. Feasibility of Zeolitic Imidazolate Framework Membranes for Clean Energy Applications. Energy & Environmental Science 2012, 5, 7637–7646. (8) Keskin, S.; Sholl, D. S. Selecting Metal Organic Frameworks as Enabling Materials in Mixed Matrix Membranes for High Efficiency Natural Gas Purification. Energy & Environmental Science 2010, 3, 343–351. (9) Watanabe, T.; Keskin, S.; Nair, S.; Sholl, D. S. Computational Identification of a Metal Organic Framework for High Selectivity Membrane-Based CO2/CH4 Separations: Cu (Hfipbb)(H2Hfipbb) 0.5. PCCP 2009, 11, 11389–11394. (10) Kim, N.; Harale, A.; Tsotsis, T. T.; Sahimi, M. Atomistic Simulation of Nanoporous Layered Double Hydroxide Materials and Their Properties. II. Adsorption and Diffusion. J. Chem. Phys. 2007, 127, 224701. (11) Lim, S. Y.; Tsotsis, T. T.; Sahimi, M. Molecular Simulation of Diffusion and Sorption of Gases in an Amorphous Polymer. J. Chem. Phys. 2003, 119, 496–504.

23

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 382 4 5 6 383 7 8 384 9 10 385 11 12 386 13 14 15 387 16 17 388 18 19 389 20 21 22 390 23 24 391 25 26 392 27 28 29 393 30 31 394 32 33 395 34 35 396 36 37 38 397 39 40 398 41 42 399 43 44 45 400 46 47 401 48 49 402 50 51 52 53 54 55 56 57 58 59 60

Page 24 of 29

(12) Le, T.; Striolo, A.; Cole, D. R. CO2–C4H10 Mixtures Simulated in Silica Slit Pores: Relation between Structure and Dynamics. The Journal of Physical Chemistry C 2015, 119, 15274–15284. (13) Cao, D.; Wu, J. Modeling the Selectivity of Activated Carbons for Efficient Separation of Hydrogen and Carbon Dioxide. Carbon 2005, 43, 1364–1370. (14) Liu, B.; Smit, B. Comparative Molecular Simulation Study of CO2/N2 and CH4/N2 Separation in Zeolites and Metal− Organic Frameworks. Langmuir 2009, 25, 5918–5926. (15) Krishna, R.; Van Baten, J. Using Molecular Simulations for Screening of Zeolites for Separation of CO2/CH4 Mixtures. Chem. Eng. J. 2007, 133, 121–131. (16) Dehghani, M.; Asghari, M.; Mohammadi, A. H.; Mokhtari, M. Molecular Simulation and Monte Carlo Study of Structural-Transport-Properties of Peba-Mfi Zeolite Mixed Matrix Membranes for CO2, CH4 and N2 Separation. Computers & Chemical Engineering 2017, 103, 12–22. (17) Yang, Q.; Zhong, C. Molecular Simulation of Carbon Dioxide/Methane/Hydrogen Mixture Adsorption in Metal− Organic Frameworks. The Journal of Physical Chemistry B 2006, 110, 17776– 17783. (18) Yang, Q.; Xue, C.; Zhong, C.; Chen, J. F. Molecular Simulation of Separation of CO2 from Flue Gases in Cu‐Btc Metal‐Organic Framework. AlChE J. 2007, 53, 2832–2840. (19) Liu, B.; Smit, B. Molecular Simulation Studies of Separation of CO2/N2, CO2/CH4, and CH4/N2 by Zifs. The Journal of Physical Chemistry C 2010, 114, 8515–8522. (20) Bae, Y.-S.; Mulfort, K. L.; Frost, H.; Ryan, P.; Punnathanam, S.; Broadbelt, L. J.; Hupp, J. T.; Snurr, R. Q. Separation of CO2 from CH4 Using Mixed-Ligand Metal− Organic Frameworks. Langmuir

2008, 24, 8592–8598.

24

ACS Paragon Plus Environment

Page 25 of 29 1 2 3 403 4 5 6 404 7 8 405 9 10 406 11 12 407 13 14 15 408 16 17 409 18 19 410 20 21 22 411 23 24 412 25 26 413 27 28 29 414 30 31 415 32 33 416 34 35 417 36 37 38 418 39 40 419 41 42 420 43 44 45 421 46 47 422 48 49 423 50 51 52 424 53 54 425 55 56 57 58 59 60

The Journal of Physical Chemistry

(21) Babarao, R.; Hu, Z.; Jiang, J.; Chempath, S.; Sandler, S. I. Storage and Separation of CO2 and CH4 in Silicalite, C168 Schwarzite, and Irmof-1: A Comparative Study from Monte Carlo Simulation. Langmuir 2007, 23, 659–666. (22) Babarao, R.; Jiang, J. Unprecedentedly High Selective Adsorption of Gas Mixtures in Rho Zeolite-Like Metal− Organic Framework: A Molecular Simulation Study. J. Am. Chem. Soc. 2009, 131, 11417–11425. (23) Firouzi, M.; Tsotsis, T. T.; Sahimi, M. Nonequilibrium Molecular Dynamics Simulations of Transport and Separation of Supercritical Fluid Mixtures in Nanoporous Membranes. I. Results for a Single Carbon Nanopore. J. Chem. Phys. 2003, 119, 6810–6822. (24) Arya, G.; Chang, H.-C.; Maginn, E. J. A Critical Comparison of Equilibrium, Non-Equilibrium and Boundary-Driven Molecular Dynamics Techniques for Studying Transport in Microporous Materials. J. Chem. Phys. 2001, 115, 8112–8124. (25) Falk, K.; Coasne, B.; Pellenq, R.; Ulm, F.-J.; Bocquet, L. Subcontinuum Mass Transport of Condensed Hydrocarbons in Nanoporous Media. Nat. Commun. 2015, 6, 6949. (26) Wu, T.; Zhang, D. Impact of Adsorption on Gas Transport in Nanopores. Scientific Reports 2016, 6, 23629. (27) Ho, T. A.; Striolo, A. Water and Methane in Shale Rocks: Flow Pattern Effects on Fluid Transport and Pore Structure. AlChE J. 2015, 61, 2993–2999. (28) Riewchotisakul, S.; Akkutlu, I. Y. Adsorption-Enhanced Transport of Hydrocarbons in Organic Nanopores. SPE J. 2016, 21, 1960–1969. (29) Collell, J.; Galliero, G.; Vermorel, R.; Ungerer, P.; Yiannourakou, M.; Montel, F.; Pujol, M. Transport of Multicomponent Hydrocarbon Mixtures in Shale Organic Matter by Molecular Simulations. The Journal of Physical Chemistry C 2015, 119, 22587–22595.

25

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 426 4 5 6 427 7 8 428 9 10 429 11 12 430 13 14 15 431 16 17 432 18 19 433 20 21 22 434 23 24 435 25 26 436 27 28 29 437 30 31 438 32 33 439 34 35 440 36 37 38 441 39 40 442 41 42 443 43 44 45 444 46 47 445 48 49 446 50 51 52 447 53 54 55 56 57 58 59 60

Page 26 of 29

(30) Li, J.; Liao, D.; Yip, S. Coupling Continuum to Molecular-Dynamics Simulation: Reflecting Particle Method and the Field Estimator. Phys. Rev. E 1998, 57, 7259. (31) Allen, M. P.; Tildesley, D. J. Computer Simulation of Liquids, 1st ed.; Oxford University Press: New York, 1989. (32) Frenkel, D.; Smit, B. Understanding Molecular Simulation: From Algorithms to Applications; Academic Press, 2001; Vol. 1. (33) Cracknell, R. F.; Nicholson, D.; Quirke, N. Direct Molecular Dynamics Simulation of Flow Down a Chemical Potential Gradient in a Slit-Shaped Micropore. Phys. Rev. Lett. 1995, 74, 2463. (34) Jin, Z.; Firoozabadi, A. Flow of Methane in Shale Nanopores at Low and High Pressure by Molecular Dynamics Simulations. J. Chem. Phys. 2015, 143, 104315. (35) Jin, Z.; Firoozabadi, A. Phase Behavior and Flow in Shale Nanopores from Molecular Simulations. Fluid Phase Equilib. 2016, 430, 156–168. (36) Pohl, P. I.; Heffelfinger, G. S. Massively Parallel Molecular Dynamics Simulation of Gas Permeation across Porous Silica Membranes. Journal of Membrane Science 1999, 155, 1–7. (37) Xu, L.; Sedigh, M. G.; Sahimi, M.; Tsotsis, T. T. Nonequilibrium Molecular Dynamics Simulation of Transport of Gas Mixtures in Nanopores. Phys. Rev. Lett. 1998, 80, 3511–3514. (38) Firouzi, M.; Sahimi, M. Molecular Dynamics Simulation of Transport and Separation of Carbon Dioxide–Alkane Mixtures in a Nanoporous Membrane under Sub-and Supercritical Conditions. Transport in Porous Media 2016, 115, 495–518. (39) Firouzi, M.; Tsotsis, T. T.; Sahimi, M. Molecular Dynamics Simulations of Transport and Separation of Supercritical Carbon Dioxide-Alkane Mixtures in Supported Membranes. Chem. Eng. Sci.

2007, 62, 2777–2789.

26

ACS Paragon Plus Environment

Page 27 of 29 1 2 3 448 4 5 6 449 7 8 450 9 10 451 11 12 452 13 14 15 453 16 17 454 18 19 455 20 21 22 456 23 24 457 25 26 458 27 28 29 459 30 31 460 32 33 461 34 35 462 36 37 38 463 39 40 464 41 42 465 43 44 45 466 46 47 467 48 49 468 50 51 52 469 53 54 55 56 57 58 59 60

The Journal of Physical Chemistry

(40) Firouzi, M.; Nezhad, K. M.; Tsotsis, T. T.; Sahimi, M. Molecular Dynamics Simulations of Transport and Separation of Carbon Dioxide–Alkane Mixtures in Carbon Nanopores. J. Chem. Phys.

2004, 120, 8172–8185. (41) Sedigh, M. G.; Onstot, W. J.; Xu, L.; Peng, W. L.; Tsotsis, T. T.; Sahimi, M. Experiments and Simulation of Transport and Separation of Gas Mixtures in Carbon Molecular Sieve Membranes. The Journal of Physical Chemistry A 1998, 102, 8580–8589. (42) Seo, Y. G.; Kum, G. H.; Seaton, N. A. Monte Carlo Simulation of Transport Diffusion in Nanoporous Carbon Membranes. Journal of Membrane Science 2002, 195, 65–73. (43) Wu, Z.; Liu, Z.; Wang, W.; Fan, Y.; Xu, N. Non-Equilibrium Molecular Dynamics Simulation on Permeation and Separation of H2/CO in Nanoporous Carbon Membranes. Sep. Purif. Technol. 2008, 64, 71–77. (44) Firouzi, M.; Wilcox, J. Molecular Modeling of Carbon Dioxide Transport and Storage in Porous Carbon-Based Materials. Microporous Mesoporous Mater. 2012, 158, 195–203. (45) Firouzi, M.; Wilcox, J. Slippage and Viscosity Predictions in Carbon Micropores and Their Influence on CO2 and CH4 Transport. J. Chem. Phys. 2013, 138, 064705. (46) Travis, K. P.; Gubbins, K. E. Transport Diffusion of Oxygen−Nitrogen Mixtures in Graphite Pores: A Nonequilibrium Molecular Dynamics (NEMD) Study. Langmuir 1999, 15, 6050−6059. (47) Düren, T.; Keil, F. J.; Seaton, N. A. Composition Dependent Transport Diffusion Coefficients of CH4/CF4 Mixtures in Carbon Nanotubes by Non-Equilibrium Molecular Dynamics Simulations. Chem. Eng. Sci. 2002, 57, 1343–1354. (48) Xu, L.; Sahimi, M.; Tsotsis, T. T. Nonequilibrium Molecular Dynamics Simulations of Transport and Separation of Gas Mixtures in Nanoporous Materials. Phys. Rev. E 2000, 62, 6942–6948.

27

ACS Paragon Plus Environment

The Journal of Physical Chemistry 1 2 3 470 4 5 6 471 7 8 472 9 10 473 11 12 474 13 14 15 475 16 17 476 18 19 477 20 21 22 478 23 24 479 25 26 480 27 28 29 481 30 31 482 32 33 483 34 35 484 36 37 38 485 39 40 486 41 42 487 43 44 45 488 46 47 489 48 49 490 50 51 52 491 53 54 492 55 56 57 58 59 60

Page 28 of 29

(49) Düren, T.; Jakobtorweihen, S.; Keil, F. J.; Seaton, N. A. Grand Canonical Molecular Dynamics Simulations of Transport Diffusion in Geometrically Heterogeneous Pores. PCCP 2003, 5, 369–375. (50) Kaganov, I.; Sheintuch, M. Nonequilibrium Molecular Dynamics Simulation of Gas-Mixtures Transport in Carbon-Nanopore Membranes. Phys. Rev. E 2003, 68, 046701. (51) Newsome, D. A.; Sholl, D. S. Predictive Assessment of Surface Resistances in Zeolite Membranes Using Atomically Detailed Models. The Journal of Physical Chemistry B 2005, 109, 7237–7244. (52) Rajabbeigi, N.; Tsotsis, T. T.; Sahimi, M. Molecular Pore-Network Model for Nanoporous Materials. II: Application to Transport and Separation of Gaseous Mixtures in Silicon-Carbide Membranes. Journal of Membrane Science 2009, 345, 323–330. (53) Naserifar, S.; Tsotsis, T. T.; Goddard III, W. A.; Sahimi, M. Toward a Process-Based Molecular Model of Sic Membranes: III. Prediction of Transport and Separation of Binary Gaseous Mixtures Based on the Atomistic Reactive Force Field. Journal of Membrane Science 2015, 473, 85–93. (54) Phan, A.; Cole, D. R.; Weiß, R. G.; Dzubiella, J.; Striolo, A. Confined Water Determines Transport Properties of Guest Molecules in Narrow Pores. ACS nano 2016, 10, 7646–7656. (55) Cabrales-Navarro, F. A.; Gómez-Ballesteros, J. L.; Balbuena, P. B. Molecular Dynamics Simulations of Metal-Organic Frameworks as Membranes for Gas Mixtures Separation. Journal of Membrane Science 2013, 428, 241–250. (56) Bhatia, S. K.; Bonilla, M. R.; Nicholson, D. Molecular Transport in Nanopores: A Theoretical Perspective. PCCP 2011, 13, 15350–15383. (57) Plimpton, S. Fast Parallel Algorithms for Short-Range Molecular Dynamics. J. Comput. Phys.

1995, 117, 1–19.

28

ACS Paragon Plus Environment

Page 29 of 29 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

493

The Journal of Physical Chemistry

TABLE OF CONTENTS IMAGE

494 495

29

ACS Paragon Plus Environment