You are on page 1of 19

Supporting Information

Nanoconfinement effect on surface tension: perspectives from molecular


potential theory

Dong Feng †, Keliu Wu*,†, Sahar Bakhshian*,‡, Seyyed Abolfazl Hosseini‡, Jing Li†, Xiangfang Li†
† State Key Laboratory of Petroleum Resources and Prospecting, China University of Petroleum
(Beijing), Beijing 102249, P.R. China.
‡ Bureau of Economic Geology, Jackson School of Geosciences, The University of Texas at Austin,
Austin, TX, USA
* Corresponding author: Keliu Wu & Sahar Bakhshian
E-mail address: kwu@cup.edu.cn & sahar.bakhshian@beg.utexas.edu
Postal address:18 Fuxue Road, Changping, Beijing China 102249 (Keliu Wu);Bureau of Economic
Geology University Station, Box X Austin, TX 78713-8924 (Sahar Bakhshian)

S1
Lists of supporting information:

Number of pages: 19

Number of figures: 7

Number of tables: 1

Table of contents

S1: Determination of the shift of critical temperature

S2: Determination of curvature-dependent coefficient

S3: Determination of ST from Peng-Robinson EOS

S2
In the Supporting Information we provide more details on the topics and methods discussed in the
paper. The Supporting Information is divided into appropriately entitled “S” for organizational
purposes; when applicable, these sections are referred to the paper by their respective numbers.

S1: Determination of the shift of critical temperature


In our work, the dimensionless shift of critical temperature is defined as
Tcb  Tc
T 
Tcb (S1)

where Tcb is the critical temperature of bulk fluid, Tc is the critical temperature of the nanoconfined
fluid, ΔT is the dimensionless critical temperature shift in a confined space.
In previous studies, the linear relationship in a log-log scale is found to be useful to describe the
relationship between the dimensionless shift of critical temperature ΔT and the normalized pore
dimension m, which is the ratio of pore radius to the LJ size parameter.S1 Meanwhile, MD simulations
have shown the difference between ΔT in cylindrical nanopores and that in slit nanopores with the
same pore dimension.S2 Therefore, the ΔT as a function of normalized pore dimension m in cylindrical
nanopores and slit nanopores can be separately expressed as

ln(T )  a1 ln(m)  b1 (S2)

ln(T )  a2 ln(m)  b2 (S3)

where a1 and b1 are the fitting parameters for cylindrical nanopores while a2 and b2 are the fitting
parameters for slit nanopores. In order to develop the correlation accurately, 29 cases for cylindrical
nanopores and 33 cases for slit nanopores, both collected from experimental studies and MD
simulations,S2-S11 are separately used to search optimal fitting parameters with a nonlinear regression
routine, and the model matches well with fitting parameters a1=-1.0355, b1=-0.2813 for cylindrical
nanopores while a2=-1.2804 and b2=-0.8358 for slit nanopores. The fitting results are shown in Fig. 3
in the manuscript, and a summary of the collected data used to reproduce Fig. 3 is listed in Table S1.
The last column of Table S1 represents the error of Tc, which can be calculated as

Tc  Tcc
Tc  (S4)
Tc

Our calculations show that the error of the majority of these data is less than 5%. Therefore, the
model is acceptable.

S3
Table S1. A summary of collected data used to reproduce Fig 3a in the manuscript and the error analysisS2-S11
Researchers Methods Fluids Pore shape m Tcb(K) Tc(K) ΔT ln (m) ln(ΔT) Calculated Tcc(K) ΔTc(%)
Burgess et al.,(1990) Exp C02 7.76 304.21 261.46 0.14 2.05 -1.96 276.68 -5.82
8.96 - - 0.10 2.19 -2.28 - -
Machin et al., (1994)(1) Exp Xe
6.21 - - 0.13 1.83 -2.03 - -
3.50 - - 0.35 1.25 -1.05 - -
Ar
9.91 - - 0.07 2.29 -2.62 - -
Morishige et al., 6.26 - - 0.15 1.83 -1.88 - -
Exp CO 2
(1997)(1) 2.91 - - 0.30 1.07 -1.20 - -
O2 3.40 - - 0.34 1.22 -1.08 - -
N2 3.22 - - 0.31 1.17 -1.16 - -
C1 5.45 190.6 171.2 0.10 1.70 -2.28 165.76 3.18
C2 4.65 305.4 272.1 0.11 1.54 -2.22 258.51 5.00
Jiang et al., (2004) MD
C3 4.08 369.8 322.6 0.13 1.41 -2.06 304.65 5.56
C4 3.79 425.2 374.2 0.12 1.33 -2.12 344.38 7.97
circle 25 1.22 1.19 0.03 3.22 -3.64 1.19 0.07
20 1.22 1.18 0.04 3.00 -3.35 1.18 -0.14
15 1.22 1.17 0.04 2.71 -3.17 1.16 0.41
12 1.22 1.16 0.05 2.48 -3.01 1.15 0.88
Singh et al., (2011) MD na(2) 8 1.22 1.13 0.08 2.08 -2.56 1.11 1.15
5 1.22 0.98 0.20 1.61 -1.62 1.05 -6.96
4 1.22 0.95 0.23 1.39 -1.49 1.00 -5.91
3 1.22 0.85 0.30 1.10 -1.20 0.92 -8.67
2 1.22 0.75 0.39 0.69 -0.95 0.77 -2.77
5.36 190.6 162.10 0.15 1.68 -1.90 165.32 -1.99
6.70 190.6 165.74 0.13 1.90 -2.04 170.54 -2.89
Jin et al., (2017) MD C1 8.04 190.6 170.20 0.11 2.08 -2.23 173.99 -2.23
9.38 190.6 174.68 0.08 2.24 -2.48 176.44 -1.01
10.72 190.6 177.40 0.07 2.37 -2.67 178.27 -0.49
S4
12.06 190.6 180.60 0.05 2.49 -2.95 179.68 0.51
13.40 190.6 182.70 0.04 2.60 -3.18 180.81 1.03
5.36 - - 0.04 1.68 -3.19 - -
4.02 - - 0.07 1.39 -2.71 - -
3.35 - - 0.09 1.21 -2.46 - -
C1
2.68 - - 0.12 0.99 -2.15 - -
2.01 - - 0.21 0.70 -1.58 - -
1.34 - - 0.36 0.29 -1.02 - -
3.72 - - 0.06 1.31 -2.81 - -
singh et al., (2009) MD 2.79 - - 0.09 1.03 -2.46 - -
C1 2.33 - - 0.11 0.84 -2.23 - -
1.86 - - 0.15 0.62 -1.89 - -
1.40 - - 0.24 0.33 -1.44 - -
2.86 - - 0.09 1.05 -2.45 - -
2.14 - - 0.12 0.76 -2.14 - -
C8
Slit 1.43 - - 0.19 0.36 -1.66 - -
1.07 - - 0.29 0.07 -1.24 - -
20.00 1.22 1.21 0.01 3.00 -4.71 1.18 2.51
15.00 1.22 1.20 0.01 2.71 -4.33 1.16 3.30
10.00 1.22 1.19 0.02 2.30 -3.81 1.14 4.85
8.00 1.22 1.18 0.03 2.08 -3.50 1.11 5.91
6.00 1.22 1.16 0.05 1.79 -2.98 1.08 7.08
Singh et al., (2011) MD na(2)
4.00 1.22 1.09 0.11 1.39 -2.25 1.00 8.26
3.00 1.22 1.02 0.16 1.10 -1.81 0.92 9.42
2.50 1.22 0.94 0.23 0.92 -1.48 0.86 8.38
2.00 1.22 0.84 0.31 0.69 -1.17 0.77 8.35
1.50 1.22 0.81 0.34 0.41 -1.09 0.61 24.08
2.96 369.85 321.80 0.13 1.09 -2.04 279.24 13.22
Li et al., (2013) MD C3
9.88 369.85 362.80 0.02 2.29 -3.96 343.81 5.24

S5
5.41 190.60 175.20 0.08 1.69 -2.52 165.53 5.52
C1 9.46 190.60 185.60 0.03 2.25 -3.64 176.56 4.87
13.51 190.60 188.20 0.01 2.60 -4.37 180.89 3.88
Jin et al., (2016) MD
4.47 305.33 279.70 0.08 1.50 -2.48 256.49 8.30
C2 8.01 305.33 295.60 0.03 2.08 -3.45 278.60 5.75
11.44 305.33 300.30 0.02 2.44 -4.11 286.86 4.48
(1) The data are collected from the work of Zarragoicoechea and Kuz, who analyzed the experimental results of Machin’s and Morishige’s.
(2) In the table, na represents the unknown simple fluids. In Singh’s work, no special fluids are mentioned and all the fluid–fluid interactions and wall–fluid
interactions are normalized with characteristic energy.

S6
S2: Determination of curvature-dependent coefficient
Due to the constraint of the pore wall and curvature, the liquid-vapor surface energy for a curved
interface in nanopores is dramatically different from the case of a planar interface. Similar to the
estimation method of planar case, the determination of surface energy for a curved interface in
nanopores is divided into three stages based on the magnitudes of pore radius R, minimum effective
interaction radius d0 and the distance between test molecule center to the center of curved interface l1.
(1) Stage І (R≤l1<R+d0)

R
R
R
l1 l1 r
r r
d0 ɵ ɵ
R

(a) d0≤r<R (b) r≥R


Fig. S1 Schematic diagram of the method for calculating the liquid-vapor ST in nanopores in Stage І. (a) d0≤r<R. Parts
of the nearest neighbors are located in the liquid phase (purple curves) while the others are located in the vapor phase (red
curves). (b) r≥ R. The dotted black curves represent that there are no fluid molecular interactions outside the pore due to
the constraint of the pore wall.

In this stage, as shown in Fig. S1(a), when other molecules are located at a distance in the range
of d0 to R from the center of test molecule, due to the constraint of curvature, the liquid-vapor molecule
potential energy and liquid-liquid molecule potential energy can be expressed as
R
U lvc (1 )   ui (r ) N vc  [2 r 2 (1  cos  )]dr (S5)
d0

R
U llc (1 )   ui (r ) N lc [4 r 2  2 r 2 (1  cos  )]dr (S6)
d0

r 2  l12  R 2
cos   (S7)
2rl1

where Ulvc and Ullc are the liquid-vapor molecule potential energy and liquid-liquid molecule potential
energy in nanopores, respectively. Nvc and Nlc are the number of vapor molecules and liquid molecules

S7
per unit volume in nanopores, respectively; and the value can be estimated by Eqs. (13-16) in the
manuscript with further considering the shift of critical temperature. 2πr2(1-cosɵ) is the surface area
represented by red curves in Fig. S1 (a).
As shown in Fig. S1(b), when other molecules are located at a distance larger than R from the
center of the test molecule, due to the coupling constraint of curvature and pore wall, the liquid-vapor
molecule potential energy and liquid-liquid molecule potential energy can be given as

U lvc ( 2 )   ui (r ) N vc  [2 r 2 (1  cos  )]dr (S8)
R


U llc ( 2 )   ui (r ) N lc [4 r 2  2 r 2 (1  cos  )  S1 ]dr (S9)
R

Where S1 is the surface area represented by the dotted black curves in Fig. S1(b), and the value can be
estimated as 4πr(r-R).
Therefore, the total potential energy of a molecule in stage І can be expressed as

U   U lvc (1 )  U lvc ( 2 )  U llc (1 )  U llc ( 2 ) (S10)

(2) Stage ІІ (R+d0≤l1<2R)

R R R

R R

l1 l1 l1
r
l 1 -R d0
ɵr ɵ r
R

(a) d0≤r<l1-R (b) l1-R≤r<R (c) r≥R


Fig. S2 Schematic diagram of the method for calculating the liquid-vapor ST in nanopores in Stage ІІ. (a) d0≤r<l1-R. All
molecules, which interact with the test molecule, are immersed in the liquid. (b) l1-R≤r<R. For the test molecule, parts of
the interacted molecules are located in the liquid phase (purple curves) while the others are located in the vapor phase (red
curves). (c) r≥ R. The dotted black curves represent that there are no fluid molecular interactions outside the pore due to
the constraint of the pore wall.

In this stage, as shown in Fig. S2(a), all molecules, located at a distance in the range of d0 to l1-R
from the center of the test molecule, are immersed in the liquid. The liquid-liquid molecule potential
energy can be expressed as
S8
l1  R

U llc (1 )  
d0
ui (r ) N lc 4 r 2 dr (S11)

As shown in Fig. S2(b), when other molecules are located at a distance in the range of l1-R to R
from the center of the test molecule, due to the constraint of curvature, the liquid-vapor molecule
potential energy and liquid-liquid molecule potential energy can be given as
R
U lvc (1 )  
l1  R
ui (r ) N vc  [2 r 2 (1  cos  )]dr (S12)

R
U llc ( 2 )  
l1  R
ui (r ) N lc [4 r 2  2 r 2 (1  cos  )]dr (S13)

As shown in Fig. S2(c), when other molecules are located at a distance larger than R from the
center of the test molecule, the determination of potential energy is influenced by the pore wall and
curvature. The liquid-vapor molecule potential energy and liquid-liquid molecule potential energy can
be given as

U lvc ( 2 )   ui (r ) N vc  [2 r 2 (1  cos  )]dr (S14)
R


U llc ( 3 )   ui (r ) N lc [4 r 2  2 r 2 (1  cos  )  S1 ]dr (S15)
R

Therefore, the total potential energy of a molecule in stage ІІ can be expressed as

U   U lvc (1 )  U lvc ( 2 )  U llc (1 )  U llc ( 2 )  U llc ( 3 ) (S16)

(3) Stage ІІІ (l1≥2R)

R R
R

l1
l1 l1
r
l 1 -R r
d0 r ɵ
R R

(a) d0≤r<R (b) R ≤r<l1-R (c) r≥l1-R


Fig. S3 Schematic diagram of the method for calculating the liquid-vapor ST in nanopores in Stage ІІІ. (a) d0≤r<R. All
molecules, which interact with the test molecule, are immersed in the liquid. (b) R ≤r<l1-R. Parts of the interacted
S9
molecules are located in the liquid phase while no liquid-vapor molecular interactions. (c) r≥ l1-R. The dotted black curves
represent that there are no fluid molecular interactions outside the pore.

In this stage, as shown in Fig. S3(a), all molecules, located at a distance in the range of d0 to R
from the center of test molecule, are immersed in the liquid. The liquid-liquid molecule potential
energy can be expressed as
R
U llc (1 )   ui (r ) N lc 4 r 2 dr (S17)
d0

When other molecules are located at a distance in the range of R to l1-R from the center of the test
molecule (Fig. S3-b), the determination of potential energy is influenced by the pore wall, and the
liquid-liquid molecule potential energy can be estimated as
l1  R

U llc ( 2 )  
R
ui (r ) N lc [4 r 2  S1 ]dr (S18)

When other molecules are located at a distance larger than l1-R from the center of test molecule
(Fig. S3-c), both the pore wall and curvature influence the determination of potential energy, and the
liquid-vapor molecule potential energy and liquid-liquid molecule potential energy can be given as

U lvc (1 )  
l1  R
ui (r ) N vc  [2 r 2 (1  cos  )]dr (S19)


U llc ( 3 )  
l1  R
ui (r ) N lc [4 r 2  2 r 2 (1  cos  )  S 1 ]dr (S20)

Therefore, the total potential energy of a molecule in stage ІІІ can be expressed as

U   U lvc (1 )  U llc (1 )  U llc ( 2 )  U llc ( 3 ) (S21)

Similar to the case of planar interface, the energy of surface layer with a radius of curvature R in
nanopores can be obtained with integrating from the interface to the infinite liquid depth

U c   (U   U   U   U ibc ) SN lc dl1 (S22)
R

In Eq. (S22), Uibc is the total potential energy for a molecule inside the nanoconfined liquid, and
can be estimated as
R 
U ibc   ui (r )N lc  4 r dr   ui (r )N lc  (4 r 2  S1 )dr
2
(S23)
d0 R

Based on Eq. (S10), Eq. (S16), Eq. (S21) and Eq. (S23), we have

S10

U   U ibc   ui (r )( N vc  N lc )  [2 r 2 (1  cos  )]dr (S24)
d0


U   U ibc  
l1  R
ui (r )( N vc  N lc )  [2 r 2 (1  cos  )]dr (S25)


U   U ibc  
l1  R
ui (r )( N vc  N lc )[2 r 2 (1  cos  )]dr (S26)

Combining Eqs. (S24-26) and Eq. (S22), we have


R  d0   
U c
   ui (r )( Nvc  Nlc )  [2 r (1  cos  )]drdl1    ui (r )( N vc  N lc )  [2 r 2 (1  cos  )]drdl1
2

SN lc R d0 R  d0 l1  R (S27)

Besides, the normalized pore dimension m and minimum effective interaction radius d0 are
expressed as

m  R /  i , d 0  2 i
6
(S28)
Taking Eq. (S28) into Eq. (S27) and rewriting the results as the form of ST for planar interface
(Eq. (9) in manuscript), the energy of the surface layer with a radius of curvature R in nanopores can
be expressed as

U c    4.3636 i ( N vc  N lc ) i 4 N lc S
(S29)
In Eq. (S29), parameter α is defined as the curvature-dependent coefficient, representing the
surface-layer energy ratio of the curved surface in nanopores to that of the planar surface.
The curvature-dependent coefficient is correlated with the normalized pore dimension m, which
is the ratio of pore radius to the LJ size parameter. For the convenience of calculation and application,
the numerical results are characterized by a simple analytical model. Both of the numerical solution
and fitting curves with non-linear least-square methods are shown in Fig. S4. The analytical
formulation for the curvature-dependent coefficient can be expressed as follows

m R / i
f ( m)  
m  1.812 R /  i  1.812 (S30)

S11
1

0.9

α (m), dimensionless
0.8

0.7

0.6
Numerical solution
0.5
fitting curve α(m)
0.4
0 10 20 30 40 50
m, dimensionless
Fig. S4 The numerical solution and analytical fitting curve of the curvature-dependent coefficient.

S3: Determination of ST from Peng-Robinson EOS


(1) Peng-Robinson EOS Model and calculation procedures
The classical Peng−Robinson (PR) EOR equation is given as S12-S14
RT a
P  (S31)
Vm  b Vm (Vm  b)  b(Vm  b)

R 2Tc 2 RT
a  0.45724  , b  0.07780 c  (S32)
Pc Pc

  [1   (1  Tr )]2 (S33)

  0.37464  1.54226  0.26992 2 (S34)

where P is the pressure; T is the temperature; Tc is the critical temperature; Tr is the reduced
temperature; Pc is the critical pressure; R is the gas constant; ν is the molar volume; ω is the acentric
factor; a and b are EOS constants.
In nanopores, the nanoconfinement effects lower the critical properties of pure components. The
changes in critical properties calculated by the Eq. (20) in the manuscript are adopted

Using compressibility factor Z  PVm RT , Eq. (S31) can be separately transformed to

Z L 3  (1  BL ) Z L 2  ( AL  3BL 2  2 BL ) Z L  ( AL BL  BL 2  BL 3 )  0 (S35)

ZV 3  (1  BV ) ZV 2  ( AV  3BV 2  2 BV ) ZV  ( AV BV  BV 2  BV 3 )  0 (S36)

with

S12
 aPL  aPV
 AL  R 2T 2  AV  R 2T 2
 
 B  bPL  B  bPV (S37)
 L RT  V RT

Where ZL and ZV are the liquid and vapor compressibility factors, respectively; PL and PV are the
liquid pressure and vapor pressure, respectively.
In our work, the nanopores are completely wetted by the liquid phase, the capillary pressure can
be given as

Pc  PV  PL
(S38)
The fugacity coefficient for pure components can be given as follows

AL Z  (1  2) BL
ln  L   ln( Z L  BL )  ( Z L  1)  ln L (S39)
2 2 BL Z L  (1  2) BL

AV Z  (1  2) BV
ln  V   ln( ZV  BV )  ( ZV  1)  ln V (S40)
2 2 BV ZV  (1  2) BV

Where  L and  V are the liquid and vapor fugacity coefficient.

The equilibrium of liquid and vapor phases in nanopores can be given as

 L PL   V PV
(S41)
Besides, the capillary pressure can be also expressed by the modified Young–Laplace equation
with considering the curvature-dependent coefficient in nano space

2 cos  2  cos 


Pc  
R R (S42)

Where γ∞ is the surface tension of the planar surface, which can be estimated with the parachor model

    pa  L  pa V 
n

(S43)
Where pa is the parachor of the component, which can be obtained from published literatures;S15
parameter n is the scaling exponent; ρL and ρv are the molar densities of liquid and vapor phase, which
can be calculated as

PL Z L PZ
L  V  V V
RT RT (S44)

S13
In the computational procedures, both the confinement effects of the critical shift of temperature and
capillary force with coupling the curvature-dependent effect are considered. A series of iterative
computation and Newton−Raphson method are applied to solve these nonlinear equations, and the
detailed flowchart can be seen in Fig.S5.

Fig S5. Flowchart for single component phase behavior and surface tension

(3) Results and discussions


Using the Peng-Robinson EOS model, we calculate the P-T phase diagram of several substances
(i.e., C1-C4, Ar, N2) in bulk and confined spaces, and compare the calculated ST with our model and
S14
published results. S16-S17 Fig. S6 shows the pressure−temperature phase diagram of several substances
(i.e., C1-C4, Ar, N2) in bulk and confined spaces. Taking C1 as a detailed example, compared with the
bulk phase, the model with the effects of both capillary pressure and confinement predicts the boiling
point (BP) temperature higher than that in the bulk phase but the critical point temperatures and
pressure lower than that in the bulk phase, narrowing the temperature range for the vapor-liquid phase
coexistence. However, these results are inconsistent with the engineering DFT predictions and
experimental phenomena. S18 Previous studies also suggested that there is a non-fugacity equality
region (corresponding to a limiting temperature) from the PR-EOS with the capillary-effect model,
which limits the applicability of PR-EOS model for single component in nanopores. S18-S20 Fig. S7
shows the comparisons of the ST with the EOS model, our model and published data. In our work, the
bulk data is firstly used to obtain the scaling exponent n in Eq. (S43);S21-S24 then, the nanoconfined ST
can be calculated with the PR-EOS model. Corresponding to the nanoconfined pressure−temperature
phase diagram in Fig. S6, the ST can be predicted with the PR-EOS model in the region that there is
vapor–liquid phase coexistence while our proposed model can a cover wider range.

5 5
C1(Bulk phase) Bulk CP C2(Bulk phase)
Nano CP
4 4 C2(H=3nm)
C1(H=3nm)
Pressure P, MPa
Pressure P, MPa

3 3

2 2

1 1

Bulk BP Nano BP
0 0
80 100 120 140 160 180 200 140 180 220 260 300
Temperature T, K Temperature T, K

(a) Methane (b) Ethane


5 4
C3(Bulk phase) C4(Bulk phase)
4 C3(H=3nm) C4(H=3nm)
3
Pressure P, MPa

Pressure P, MPa

3
2
2

1
1

0 0
140 180 220 260 300 340 380 200 240 280 320 360 400 440
Temperature T, K Temperature T, K

(c) Propane (d) Butane


S15
4 6
N2(Bulk phase) Ar(Bulk phase)
5
N2(R=2.2nm) Ar(R=2.2nm)
3

Pressure P, MPa
Pressure P, MPa

2 3

2
1
1

0 0
60 80 100 120 140 60 80 100 120 140 160
Temperature T, K Temperature T, K

(e) Nitrogen (f) Argon


Fig. S6 Pressure−temperature phase diagram of several substances in bulk and confined spaces. (a) methane; (b)ethane;
(c) propane; (d) butane; (e) nitrogen; (f) argon
20 20
C1(bulk, Garrido, 2016) C2(bulk,Hoang, 2017 )
C1(H=3nm,G) C2(H=3nm,G)
16 C1(H=3nm,M) 16 C2(H=3nm,M)
C1(EOS-bulk) C2(EOS-bulk)
C1(EOS-3nm) C2(EOS-3nm)
12 C1(model-bulk) 12 C2(model-bulk)
γ, mN/m

γ, mN/m

C1(model-3nm) C2(model-3nm)
8 8

4 4

0 0
100 120 140 160 180 200 140 190 240 290 340
Temperature T, K Temperature T, K

(a) Methane (b) Ethane


25 25
C3(bulk, Garrido, 2016) C4(bulk, Hoang, 2017)
C3(H=3nm,G) C4(H=3nm,G)
20 C3(H=3nm,M) 20 C4(H=3nm,M)
C3(EOS-bulk) C4(EOS-bulk)
C3(EOS-3nm) C4(EOS-3nm)
15 15
γ, mN/m
γ, mN/m

C3(model-bulk) C4(model-bulk)
C3(model-3nm) C4(model-3nm)
10 10

5 5

0 0
150 200 250 300 350 400 150 200 250 300 350 400 450
Temperature T, K Temperature T, K

(c) Propane (d) Butane

S16
20 20
N2(bulk, Stansfield, 1958) Ar(bulk, Sampayo, 2010)
N2(R=2.2nm) Ar(R=2.2nm)
16 N2(EOS-bulk) 16 Ar(EOS-bulk)
N2(EOS-2.2nm) Ar(EOS-2.2nm)
N2(model-bulk) Ar(model-bulk)
12 N2(model-2.2nm) 12
γ, mN/m

Ar(model-2.2nm)

γ, mN/m
8 8

4 4

0 0
60 80 100 120 140 60 80 100 120 140
Temperature T, K Temperature T, K

(e) Nitrogen (f) Argon


Fig. S7 Comparison of the ST with the EOS model, our model, and published data. (a)-(d): liquid-vapor ST of
hydrocarbons C1-C4 in slit nanopores of graphite (G) and mica (M) is obtained with MD simulations.S16 (e)-(f): liquid-
vapor ST of nonhydrocarbons N2 and Ar in cylindrical nanopores is obtained with perturbed-chain statistical associating
fluid theory. S17 The parachor for C1-C4, N2 and Ar are 74.05, 112.91, 154.03, 193.9, 61.62 and 54, respectively.S15 Scaling
exponent n in Eq. (S30) for C1-C4, N2 and Ar are 3.3, 3.6, 3.6, 3.6, 3.2 and 3.4, respectively.

References:
(S1) Yang, G.; Fan, Z.; Li, X. Determination of confined fluid phase behavior using extended Peng-
Robinson equation of state. Chem. Eng. J 2019, 378, 122032.
(S2) Singh, S.K.; Singh, J.K. Effect of pore morphology on vapor–liquid phase transition and
crossover behavior of critical properties from 3D to 2D, Fluid Phase Equilibr. 2011, 300, 182–
187.
(S3) Singh, S.K.; Sinha, A.; Deo, G.; Singh, J.K. Vapor− liquid phase coexistence, critical properties,
and surface tension of confined alkanes, J. Phys. Chem. C 2009, 113 (17), 7170-7180.
(S4) Sedghi, M.; Piri, M. Capillary condensation and capillary pressure of methane in carbon
nanopores: molecular dynamics simulations of nanoconfinement effects. Fluid Phase Equilib.
2018, 459, 196–207.
(S5) Burgess, C.G.; Everett, D.H.; Nuttall, S. Adsorption of carbon dioxide and xenon by porous glass
over a wide range of temperature and pressure-applicability of the Langmuir case VI equation,
Langmuir 1990, 6 (12), 1734–1738.
(S6) Machin, W.D. Temperature dependence of hysteresis and the pore size distributions of two
mesoporous adsorbents, Langmuir 1994, 1235–1240.
(S7) Morishige, K.; Fujii, H.; Uga, M.; Kinukawa, D. Capillary critical point of argon, nitrogen,
oxygen, ethylene, and carbon dioxide in MCM-41, Langmuir. 1997,13, 3494–3498.
S17
(S8) Jiang, J.; Sandler, S.I.; Smit, B. Capillary phase transitions of n-alkanes in a carbon nanotube,
Nano Lett. 2004, 4, 241–244.
(S9) Jin, B.; R, Bi.; Nasrabadi, H. Molecular simulation of the pore size distribution effect on phase
behavior of methane confined in nanopores. Fluid Phase Equilibr. 2017,452, 94–102.
(S10) Jin, B.; Nasrabadi, H. Phase behavior of multi-component hydrocarbon systems in nano-pores
using gauge-GCMC molecular simulation. Fluid Phase Equilib. 2016, 425, 324–334.
(S11) Li, Z.; Jin, Z.; Firoozabadi, A. Phase behavior and adsorption of pure substances and mixtures
and characterization in nanopore structures by density functional theory, SPE J. 2014,19, 91–109.
(S12) Song, Z.; Song,Y.; Guo, J.; Zhang, Z.; Hou, J. Adsorption induced critical shifts of confined
fluids in shale nanopores. Chem. Eng. J. 2020, 385, 123837.
(S13) Yang, G.; Fan, Z.; Li, X. Determination of confined fluid phase behavior using extended Peng-
Robinson equation of state, Chem. Eng. J. 2019, 378, 122032.
(S14) Zhang, K.; Jia, N.; Liu, L. Adsorption Thicknesses of Confined Pure and Mixing Fluids in
Nanopores. Langmuir 2018, 34, 12815−12826.
(S15) Schechter, D. S.; Guo, B. Parachors Based on Modern Physics and Their Uses in IFT Prediction
of Reservoir Fluids. SPE J. 1998, 1, 207−217.
(S16) Singh, S.K.; Sinha, A.; Deo, G.; Singh, J.K. Vapor−liquid phase coexistence, critical properties,
and surface tension of confined alkanes, J. Phys. Chem. C 2009, 113 (17), 7170-7180.
(S17) Tan, S.; Piri, M. Equation-of-state modeling of confined-fluid phase equilibria in nanopores.
Fluid Phase Equilib. 2015, 393, 48−63.
(S18) Liu, Y. L.; Jin, Z. H.; Li, H. Z. Comparison of Peng-Robinson Equation of State with Capillary
Pressure Model with Engineering Density-Functional Theory in Describing the Phase Behavior
of Confined Hydrocarbons. SPE J. 2018, 23 (5), 1784−1797.
(S19) Stimpson, B. C.; Barrufet, M. A. Thermodynamic Modeling of Pure Components Including the
Effects of Capillarity. J. Chem. Eng. Data 2016, 61 (8), 2844−2850.
(S20) Zuo, J. Y.; Guo, X.; Liu, Y.; Pan, S.; Canas, J.; Mullins, O. C. Impact of Capillary Pressure and
Nanopore Confinement on Phase Behaviors of Shale Gas and Oil. Energy Fuels 2018, 32,
4705−4714.
(S21) Garrido, J. M.; Mejía, A.; Piñeiro, M. M.; Blas, F. J.; Müller EA. Interfacial tensions of industrial
fluids from a molecular-based square gradient theory, AIChE J. 2016, 62, 1781–1794.
(S22) Hoang, H.; Delage-Santacreu, S.; Galliero, G. Simultaneous description of equilibrium,

S18
interfacial, and transport properties of fluids using a Mie chain coarse-grained force field, Ind.
Eng. Chem. Res. 2017, 56, 9213-9226.
(S23) Sampayo, J; Blas, F.; Miguel, E.; Müller, E.; Jackson, G. Monte Carlo simulations of the liquid-
vapor interface of Lennard-Jones diatomics for the direct determination of the interfacial tension
using the test-area method. J. Chem. Eng. Data 2010, 55, 4306−4314.
(S24) Stansfield, D. The surface tensions of liquid argon and nitrogen, Proc. Phys. Soc. 1958, 72, 854–
866

S19

You might also like