You are on page 1of 7

Supplementary Material for

Pressure-dependent transport characteristic of methane gas in slit

nanopores

Hao Yua, JingCun Fana, Jie Chena,b, YinBo Zhua, HengAn Wua,*

aCAS Key Laboratory of Mechanical Behavior and Design of Materials, Department of Modern

Mechanics, University of Science and Technology of China, Hefei, Anhui 230027, China
bZhongtian
Technology Submarine Cables Co., Ltd, Nantong Jiangsu 226010, China

*E-mail: wuha@ustc.edu.cn

This document contains following sections:

Fig. S1. Histogram for mineralogical compositions of several shale samples.

Fig. S2. Normalized mass transport velocity profiles of methane molecules in slit nanopores (H = 2

nm) under different system pressures: (a) 50 MPa, (b) 10 MPa, (c) 2 MPa, (d) 0.5 MPa, respectively.

Fig. S3. Normalized mass transport velocity profiles of methane molecules in slit nanopores (H =

10 nm) under different system pressures: (a) 50 MPa, (b) 10 MPa, (c) 2 MPa, (d) 0.5 MPa,

respectively.

Table S1. The mass flux contribution of three transport mechanisms as functions of system pressure

in slit nanopores with different pore widths (2 nm, 6 nm, and 10 nm).

Text S1. The determination of parameters in apparent permeability model based on MD simulations.

Fig. S4. Normalized density profiles of methane molecules in slit nanopores with different pore

widths (2nm, 6nm, and 10nm) under different system pressures: (a) 50 MPa, (b) 10 MPa, (c) 2 MPa,

(d) 0.5 MPa, respectively.

Fig. S5. Configuration for the simulation system of methane molecules in the box.
Fig. S1. Histogram for mineralogical compositions of several shale samples. Here, the statistical

data are based on the ref. [S1].

Fig. S2. Normalized mass transport velocity profiles of methane molecules in slit nanopores (H = 2
nm) under different system pressures: (a) 50 MPa, (b) 10 MPa, (c) 2 MPa, (d) 0.5 MPa, respectively.
Fig. S3. Normalized mass transport velocity profiles of methane molecules in slit nanopores (H =
10 nm) under different system pressures: (a) 50 MPa, (b) 10 MPa, (c) 2 MPa, (d) 0.5 MPa,
respectively.

Table S1. The mass flux contribution of three transport mechanisms as functions of system pressure
in slit nanopores with different pore widths (2nm, 6nm, and 10nm).
Channel width(nm) Pressure (MPa) Qsurface QKundsen Qviscous
0.1 0.4125 0.5462 0.0413
0.3 0.3916 0.5230 0.0854
0.5 0.3786 0.5155 0.1059
1 0.3622 0.4863 0.1515
3 0.2771 0.4223 0.3006
2 5 0.2574 0.3857 0.3569
10 0.2301 0.3197 0.4502
20 0.1773 0.2459 0.5768
30 0.1187 0.2124 0.6689
40 0.1079 0.2010 0.6911
50 0.1062 0.1840 0.7098
0.1 0.1773 0.7762 0.0465
0.3 0.1429 0.7060 0.1511
0.5 0.1339 0.6578 0.2083
1 0.1126 0.6048 0.2826
3 0.0735 0.4480 0.4785
6 5 0.0653 0.3914 0.5433
10 0.0547 0.3038 0.6415
20 0.0511 0.2562 0.6927
30 0.0337 0.1969 0.7694
40 0.0421 0.1589 0.7990
50 0.0108 0.1143 0.8749
0.1 0.0485 0.8645 0.0870
0.3 0.0421 0.7643 0.1936
0.5 0.0399 0.7046 0.2555
1 0.0358 0.6002 0.3640
3 0.0273 0.4393 0.5334
10 5 0.0221 0.3000 0.6779
10 0.0154 0.2454 0.7392
20 0.0103 0.1429 0.8468
30 0.0069 0.0939 0.8992
40 0.0034 0.0466 0.9500
50 0.0020 0.0273 0.9707

Text S1. The determination of parameters in apparent permeability model based on MD


simulations:
Adsorption parameters: Langmuir adsorption theory is widely used to describe single
layer adsorption behavior of shale gas:

C  Cmax P / ( PL  P) , (1)

where Cmax and PL donate the adsorption capacities at infinite pressure and Langmuir
pressure, which can be obtained by fitting isothermal adsorption capacity under a wide
range pressure. To discuss pore width dependence of the isothermal adsorption capacity,
we compared adsorption layers in slit nanopores with different pore widths (Fig. S4). It
can be found that density distribution in adsorption layer remains unchanged with pore
width changing, demonstrating the same isothermal adsorption capacity in slit
nanopores with different pore widths (H > 2 nm). Previous literatures also reported
similar MD simulation results [S2, S3]. Hence the Cmax and PL are constant for slit
nanopores with different pore widths (H > 2 nm) in apparent permeability model [S4-
S6]. It is noted that apparent permeability model proposed here will be invalid when
pore width is less than 2 nm, where adsorption configuration changes significantly [S7,
S8].

Self-diffusion coefficient: as an essential property of gas, self-diffusion coefficient


changes with system pressure. Therefore, we built a box (40 Å × 200 Å × 80 Å)
containing pure methane molecules with periodic boundary conditions in all three
directions (Fig. S5). After conducting GCMC and EMD simulations, the mean square
displacements of methane molecules (Fig. S5) were obtained under different pressures.
Then the self-diffusion coefficient can be calculated depending on:

y2
D , (2)
6t
where〈y2〉is the MSD of the methane molecules in box.

Viscosity: the Green-Kubo relation and Einstein equation are two common approaches
to calculate gas viscosity [S9]. Here, we employed Green-Kubo formulation to
determine gas viscosity of methane molecules (Fig. S5) under different system
pressures:

V 
η
10kBT  P
0
αβ
αβ (0) Pαβ (t ) dt , (3)

where Pαβ is the symmetrized traceless portion of the stress tensor, σαβ, defined as:
1 1
Pαβ  (σ αβ  σ βα )  δαβ ( σ yy ) . (4)
2 3 y

Our calculation results are comparable to experimental data reported by the National
Institute of Standards and Technology (NIST) [S10]. For example, the viscosity
calculated by MD simulations (13.052 μPa·s) is close to that by experimental method
(13.857 μPa·s) at 10 MPa and 298K.
Fig. S4. Normalized density profiles of methane molecules in slit nanopores with different pore
widths (2nm, 6nm, and 10nm) under different system pressures: (a) 50 MPa, (b) 10 MPa, (c) 2 MPa,
(d) 0.5 MPa, respectively.

Fig. S5. Configuration for the simulation system of methane molecules in the box.
References
[1] C.R. Clarkson, N. Solano, R.M. Bustin, A.M. Bustin, G.R. Chalmers, et al., Pore structure
characterization of North American shale gas reservoirs using USANS/SANS, gas adsorption,
and mercury intrusion, Fuel 103 (2013) 606-616.
[2] K. Mosher, J.J. He, Y.Y. Liu, E. Rupp, J. Wilcox, Molecular simulation of methane adsorption
in micro-and mesoporous carbons with applications to coal and gas shale systems, Int. J. Coal.
Geol. 109 (2013) 36-44.
[3] Z.H. Jin, Effect of nano-confinement on high pressure methane flow characteristics, J. Nat. Gas
Sci. Eng. 45 (2017) 575-583.
[4] X. Xiong, D. Devegowda, G.G. Michel, R.F. Sigal, F. Civan, A fully-coupled free and adsorptive
phase transport model for shale gas reservoirs including non-Darcy flow effects, in: SPE
annual technical conference and exhibition, Society of Petroleum Engineers, 2012.
[5] W.H. Song, J. Yao, Y. Li, H. Sun, L. Zhang, et al., Apparent gas permeability in an organic-rich
shale reservoir, Fuel 181 (2016) 973-984.
[6] H. Sun, J. Yao, D.Y. Fan, C.C. Wang, Z.X. Sun, Gas transport mode criteria in ultra-tight porous
media, Int. J. Heat Mass Transf. 83 (2015) 192-199.
[7] H.A. Wu, J. Chen, H. Liu, Molecular dynamics simulations about adsorption and displacement
of methane in carbon nanochannels, J. Phys. Chem. C 119 (24) (2015) 13652-13657.
[8] S. Wang, Q.H. Feng, F. Javadpour, Y.B. Yang, Breakdown of fast mass transport of methane
through calcite nanopores, J. Phys. Chem. C 120 (26) (2016) 14260-14269.
[9] T. Chen, B. Smit, A. Bell, Are pressure fluctuation-based equilibrium methods really worse than
nonequilibrium methods for calculating viscosities?, J. Chem. Phys. 131 (24) (2009) 246101.
[10] National Institute of Standards and Technology. Thermophysical properties of fluid systems;
2011. http://webbook.nist.gov/chemistry/fluid/.

You might also like