You are on page 1of 12

Downloaded from https://iranpaper.

ir

Molecular Simulation

ISSN: (Print) (Online) Journal homepage: https://www.tandfonline.com/loi/gmos20

The effects of potential model of CO2 on its bulk


phase properties and adsorption on surfaces and
in pores

Xiu Liu, Allan Hua Heng Sim & Chunyan Fan

To cite this article: Xiu Liu, Allan Hua Heng Sim & Chunyan Fan (2022): The effects of potential
model of CO2 on its bulk phase properties and adsorption on surfaces and in pores, Molecular
Simulation, DOI: 10.1080/08927022.2022.2086276

To link to this article: https://doi.org/10.1080/08927022.2022.2086276

Published online: 15 Jun 2022.

Submit your article to this journal

Article views: 2

View related articles

View Crossmark data

Full Terms & Conditions of access and use can be found at


https://www.tandfonline.com/action/journalInformation?journalCode=gmos20
Downloaded from https://iranpaper.ir

MOLECULAR SIMULATION
https://doi.org/10.1080/08927022.2022.2086276

The effects of potential model of CO2 on its bulk phase properties and adsorption on
surfaces and in pores
Xiu Liu, Allan Hua Heng Sim and Chunyan Fan
Discipline of Chemical Engineering, WA School of Mines: Minerals, Energy and Chemical Engineering, Curtin University, Bentley, Australia

ABSTRACT ARTICLE HISTORY


Adsorption of CO2 using carbonaceous materials is one of the common approaches for carbon capture Received 11 April 2022
and storage. To improve this technology, it necessitates the understanding of the underlying Accepted 28 May 2022
mechanisms, for which molecular simulation has been recognised as a useful tool and compensation
KEYWORDS
of experimental study. However, to ensure the reliability of the simulation results, the selection of the Carbon dioxide; adsorption;
reliable potential model is vital. A systematic Monte Carlo simulation was conducted to evaluate the Monte Carlo simulation;
performance of two commonly used potential models of CO2, i.e. the simple 1C-LJ and the TraPPE 3C- potential model
LJ + 3Q models, in different scenarios including bulk phase, adsorption on graphite surface and in
carbonaceous pores. In addition, a special investigation on the occurrence of step-wised hysteresis in
wedge-shaped pore that has been observed with other simple gases. It reveals the consideration of
shape and quadrupole of CO2 play an important role in the description on the phase equilibria and
adsorption behaviour, especially at temperatures below the triple point.

1. Introduction
The main objective of this work is to evaluate systematically
There is rising global concern on climate change due to the the effects of potential models on their performances within
emission of greenhouse gases, particularly carbon dioxide different scenarios, including the vapour liquid/solid equili-
(CO2) [1–3]. Recent years have seen much work devoted to brium (VLE/VSE) of bulk phase, and adsorption in carbon-
the adsorption of CO2 using a carbonaceous material such as aceous materials. Two potential models of CO2, i.e. 1C-LJ
activated carbon (AC) [4–6], which is emerging as an impor- and TraPPE are selected. The 1C-LJ model has the simplest
tant research field. The AC possesses complex physical struc- CO2 representation as it simulates the molecule in a spherical
ture and chemical properties, including wide pore size shape with single LJ site and no charge; therefore, the quadru-
distributions and surface heterogeneity [5]. Thus, it is difficult pole moment and elongated shape of CO2 molecule are not
to understand the underlying mechanisms by experimental considered. The great advantage of applying 1C-LJ model is
study solely. Molecular simulation such as Monte Carlo its computational efficiency. The TraPPE model (i.e. 3C-LJ +
(MC) simulation [7,8] is an effective tool to investigate the 3Q) is triatomic CO2 model comprising three dispersive inter-
adsorption behaviours from a microscopic perspective and action sites, located on each atoms, and three discrete charges
therefore can facilitate the development of adsorbent with cus- are assigned at the same locations to consider the high quad-
tomised properties for more effective carbon capture rupole moment of CO2, and it has an elongated shape as com-
performance. pared to the 1C-LJ model. One of the major differences
To ensure the reliability of molecular simulation, it is vital between these two models for their adsorption in confined
to select the proper potential models, and verify the simulated spaces have been observed is the conformation of the adsorbed
results against experimental data. The Lennard–Jones (LJ) 12- phase. The 1C-LJ CO2 molecules tend to form hexagonal pack-
6 is one of the widely applied potential models for MC simu- ing in monolayer, while the 3C-LJ CO2 molecules show richer
lation, and the parameters for many commonly used species features, including the tilted configuration [13], pinwheel
are readily available in the literature. A number of intermole- structures [14–16] and herringbone structures [9,17,18]. The
cular potential models for CO2 have been proposed in the lit- type of conformation is decided by the principle that the
erature such as HMT [9], Five charge [10], TraPPE [11] and configuration energy of the system should be minimised
1C-LJ [12]. Do and Do [13] compared three multi-sites poten- [18,19,20].
tial models (i.e. TraPPE, HMT and Five charge) and the simple In this work, the bulk phase behaviour with the selected
1C-LJ model on their adsorption on graphitised carbon black, CO2 models was studied with the Bin-canonical Monte
and found the TraPPE model is superior in describing the Carlo [21], which has been approved to be an efficient method
experimental isotherms at the specified conditions. In the in calculating the thermodynamic properties of the bulk fluids.
same work, the effects of quadrupole interaction and molecu- The adsorption of CO2 in carbonaceous materials, which are
lar shape were also investigated briefly for CO2 adsorption modelled as the graphitic surface and simplified homogeneous
mainly in micro-slit pores. pores, is investigated by applying the Grand Canonical Monte

CONTACT Chunyan Fan Chunyan.fan@curtin.edu.au


© 2022 Informa UK Limited, trading as Taylor & Francis Group
Downloaded from https://iranpaper.ir

2 X. LIU ET AL.

Carlo [22]. This approach can mimic the process of the volu- 2.2. Simulation systems
metric adsorption method, which is commonly used in
Four simulation systems used in this work are bulk phase equi-
practical.
libria (VLE and VSE), graphitised carbon black, uniform slit
Although, it has become a common practice to simulate
pore and wedge pore. The corresponding schematic diagrams
activated carbons as a combination of uniform slit pores
are presented in Figure 1.
with different sizes [23–25]. It has been well recognised that
the geometric structures of AC are much more complex.
According to the X-ray analysis [26], the AC is composed of 2.2.1. Phase equilibria system
crystallites that are stacked randomly and the confined spaces As shown in Figure 1(a), the dense phase is placed at the
between the carbon stacks tend to be wedge shaped. Previous middle of a rectangular simulation box and two bulk gaseous
studies [4–6] also show that a wedge pore is considered as a phases are connected at the two ends, forming two interfaces.
more realistic model for characterisation of the AC, moreover, Periodic boundary conditions are applied in all directions. The
in which an interesting step-wise adsorption isotherm has initial configuration of the dense phase is either solid slab or
been reported for noble gases. For the first time, the perform- lattice structured liquid, depends on the temperature of the
ance of the two selected CO2 models in wedge pore is evaluated system is below or above the triple point.
in this work.
2.2.2. Graphitised carbon black
The graphitised carbon black is simulated as an infinite plane
surface positioned at the bottom of the simulation box, as
2. Methodology shown in Figure 1(b). Periodic boundary conditions are
applied in both x- and y-directions. The length of each side
2.1. Potential models L is set as 4 nm and the box height is 10 nm, which is large
The corresponding molecular parameters for the two selected enough for the local density in the z-direction to closely
potential models of CO2, i.e. the 1C-LJ and TraPPE (3C-LJ) are approach the bulk gas density. The interaction between CO2
listed in Table 1. The interaction energy between CO2 mol- molecule and surface is calculated by the Steele 10-4-3
ecules i and j is given by the sum of the Lennard–Jones (LJ) equation [29], with the detailed parameters for a carbon
and Coulomb interactions: atom in a graphene layer are listed in Table 1.

 12  6 

A 
B sab
ij sab
ij
2.2.3. Slit pore
wi,j = 41ab
ij − Semi-finite slit pore is composed of two parallel graphite walls.
a=1 b=1 rijab rijab
Each wall contains three graphene layers as shown in Figure 1

A 
B qai qbj (c). The dashed line of the box represents the boundaries of the
+ (1) simulation box in the y- and z-directions. Bojan–Steele poten-
a=1 b=1 4p10 rijab
tial is applied for the interaction between the CO2 and solids.
Periodic boundary condition is only applied in the x-direction.
where A and B are the numbers of LJ sites and the numbers of
The two open ends are connected to bulk gas reservoirs, each
partial charges on the molecules i and j, respectively. The rijab is
has a length of 3 nm along the y-axis, and the dimension in the
the separation distance between the LJ site a on molecule i and
z-direction is decided by the pore width. Thus, a mechanical
the LJ site b on molecule j, and the separation distance between
equilibrium can be maintained between the pore and the sur-
charge a on molecule i and charge b on molecule j having
roundings through the interaction of molecules inside the pore
charges qai and qbj . The 1ab ab
ij and sij are the combined LJ with the bulk phase. The pore width H is referred to the dis-
well-depth and collision diameter for the two LJ sites, which
tance between the plane passing through the centres of carbon
are calculated by the Lorentz–Berthelot (LB) combining
atoms in the innermost layer of one wall and the correspond-
rules. The ε0 is the vacuum permittivity.
ing plane of the opposite wall.
The solid–fluid interaction energy is calculated by Steele 10-
4-3 equation for the infinite graphite surface and Bojan–Steele
equation for the semi-finite graphene layer [27,28]. The par- 2.2.4. Wedge pore
ameters of the solid (i.e. σss and εss/kB for a carbon atom in a Figure 1(d) depicts an open wedge pore and its structural par-
graphene layer) are also included in Table 1. The carbon den- ameters: pore sizes at small end (SH) and big end (BH), axial
sity of a graphene layer, ρs = 38.2 nm−2 and spacing between pore length (L) and half angle (α). The dashed line of box rep-
two adjacent graphene layers Δ = 0.3354 nm. resents the boundaries of the simulation box in the y- and z-
directions. Same as for the slit pore, the gas surrounding at
the open end has a length of 3 nm along the y-axis, and the
Table 1. The LJ parameters of two CO2 models and the carbon atom [11,12,20].
dimension in the z-direction is decided by the width at the
Atom σ (nm) ε/kB (K) q (e) Bond length (nm)
big end. The pore length along the axial direction (i.e. the y-
Graphene layer C 0.34 28
1C-LJ 0.3648 246.15
direction) is finite L, while, the dimension in the x-direction
3C-LJ C 0.28 27 +0.7 is assumed to be infinite with the periodic boundary condition
O 0.305 79 −0.35 applied. The configurations of the wedge pore applied in this
l(C–O) 0.116
work are L = 8.5 nm, SH = 2 nm, BH = 3.5 nm, and α = 5°.
Downloaded from https://iranpaper.ir

MOLECULAR SIMULATION 3

Figure 1. (Colour online) Schematic diagrams of four simulation systems applied in this work: (a) phase equilibria system, (b) graphitic surface, (c) slit pore and (d)
wedge pore. The dashed line in (b) and (d) represents the boundaries of the simulation box in the y- and z-directions.

2.3. Monte Carlo simulation of 20% [33]. The dimension of the simulation box in the x-
direction is set as 10 times the collision diameter of CO2.
2.3.1. Bin-canonical Monte Carlo
The lengths of the simulation box in the y- and z-directions
Bin-canonical Monte Carlo (Bin-CMC) method is used to per-
follow what have been described in previous sections for
form the bulk phase equilibria study for temperatures ranging
each solid model. The cut-off radius is five times the collision
from below the triple point to the critical point of CO2. For the
diameter of the carbon atom of CO2 (i.e. 1.4 nm for TraPPE)
vapour–solid equilibria (VSE) below the triple point, the sys-
and of the CO2 molecule (i.e. 1.8 nm for 1C-LJ).
tem is initialised by placing a solid slab of CO2 molecules
arranged in the fcc structure at the middle of an elongated
simulation box with vacuum spaces on both sides of the slab
(see Figure 1(a)). The simulation box is divided into 28 bins 3. Results and discussions
for the purpose of determining the density distribution and
3.1. VSE/VLE of CO2
the bin size is smaller in the interfacial and condensed phase
regions and larger in the gas-phase region. To ensure the accu- The VSE/VLE coexistence curves of CO2 obtained with two
racy of output data, at least 8 × 105 cycles were employed, with models are shown in Figure 2 and compared against the exper-
1000 displacement moves in each cycle and, in both the equi- imental data (cross symbols). The experimental data above tri-
libration and sampling stages. The vapour pressure in the ple point (216.6 K in experiment) were taken from the NIST
simulation is calculated through a virial route presented in [34]. The experimental vapour pressure below triple point in
reference [30]. The evaporation (sublimation) heat is deter- Figure 2(c,d) was taken from Terlain and Larher [17] and
mined by considering the change of the molar enthalpy Spencer et al. [35]. The experimental solid phase density and
between the two phases (gas and liquid phases) [21]. Surface sublimation heat below triple point in Figure 2(a,b,e) were
tension is calculated using the Kirkwood and Buff equation taken from Chen et al. [36]. The isosteric heat at zero loading,
[31]. Details of the Bin-CMC scheme and the calculation q0st is computed using Monte Carlo integration [37]. The simu-
methods of various thermodynamic properties can be found lated results of both models (red circle for 3C-LJ model and
elsewhere [21]. For temperatures above the triple point, the blue circle for 1C-LJ model) covered from below the triple
same simulation procedure described above was performed point up to near the critical point (304.2 K in experiment).
to determine the vapour–liquid equilibria (VLE). The only In general, compared with the 1C-LJ model, the simulation
difference is that the solid slab is replaced with a lattice having results of the 3C-LJ model agree better with those reported
an initial density equal to the bulk liquid density at the boiling experimentally. Furthermore, the following observations can
point. be made:

(1) At temperature below the triple point, compared with the


2.3.2. Grand canonical Monte Carlo 3C-LJ, the 1C-LJ model is inadequate in describing the
Grand canonical Monte Carlo (GCMC) simulation [22,32] was experimental data, including the densities of gas and
employed with at least 5 × 105 cycles in both equilibration and solid phases, the saturated vapour pressure and sublima-
sampling stages to obtain the adsorption isotherm and micro- tion heat. However, the deviation is negligible for temp-
scopic properties. Each cycle consists of 1000 displacement eratures greater than 230 K.
moves, insertion and deletion with equal probability. As a (2) Both models capture the discontinuity in the branch of the
result, this generated a total of 5 × 108 configurations. In the dense phase density (see Figure 2(a)) as well in the
equilibration stage, the maximum displacement length is enthalpy of phase change (see Figure 2(e)), which corre-
initially configured as half of the largest box dimension and sponds to the triple point. However, the prediction
adjusted at the end of each cycle to reach an acceptance ratio made by 3C-LJ model (216 K) is much closer to the
Downloaded from https://iranpaper.ir

4 X. LIU ET AL.

Figure 2. (Colour online) Thermodynamic properties of 1C-LJ and TraPPE models at various temperatures obtained with Bin-CMC as compared to the experimental
data, (a) liquid (solid)–vapour coexistence curve in linear scale, (b) liquid (solid)–vapour coexistence curve in semi-log scale, (c) vapour pressure in linear scale, (d)
vapour pressure in semi-log scale, (e) evaporation (sublimation) heat and (f) surface tension. Statistical uncertainties are smaller than the size of the symbols used
in the above figures.

experimental value (216.6 K), compared to that of the 1C- which, the two models have the same thermodynamic proper-
LJ model (172 K). ties. In this work, the equivalent temperatures of the two
(3) For surface tension (see Figure 2(f)), 1C-LJ model predicts models were defined as
the correct trend of surface tension, and performs better
than the 3C-LJ model. Tr
T1C Tr
− T1C = T3C − T3C (2)

i.e. with same intervals from their respective triple points,


Tr Tr
As aforementioned, the triple points predicted by the 1C-LJ where T1C = 172 K and T3C = 216 K. Hence,
and 3C-LJ models are not the same, which indicate, at the same
temperature, the two models describe the thermodynamic T3C = T1C + 44K (3)
properties of CO2 differently. To evaluate adsorption beha-
viours described by the two models, it is necessary to compare
them at a hereafter so called ‘equivalent temperature’, i.e. at The corresponding temperatures used for the 1C-LJ and 3C-LJ
models are shown in Table 2, where T1 and T2 denote an
equivalent temperature below and above the triple point,
Table 2. The corresponding temperatures used for 1C-LJ and 3C-LJ models. respectively. The saturation vapour pressure of CO2 at various
CO2 model 1C-LJ 3C-LJ temperatures was calculated using Bin-CMC simulations and
T1 (below triple point) 150 K 194 K presented in Table 3 for 1C-LJ and 3C-LJ models and exper-
T2 (above triple point) 229 K 273 K
imental data [35,38], respectively.
Downloaded from https://iranpaper.ir

MOLECULAR SIMULATION 5

Table 3. Saturation vapour pressures of CO2 (P0, unit = kPa) at various temperatures (unit = K) for 1C-LJ and 3C-LJ models, and experimental data [33,36].
T (K) 150 160 170 173 194 216 216.6 229 273 298
1C 17.62 37.55 85.69 269 681.5 1060 3468 5820
3C 0.896 4.848 12.304 99.7 455.607 3479 6422
Exp. 96.7 3480

3.2. Adsorption on graphite surface the results reported in Xu et al. [18]. However, with the 1C-
LJ model, at 150 K there are four distinct sharp steps recog-
The excess surface concentration [39] of adsorbate on the sur-
nised, indicating four layers are formed through the first-
face is defined as below:
order transition. The steps are no longer observed when temp-
N − rg Vacc eratures are greater than 150 K for the 1C-LJ model.
Gexc = (4) The snapshots of the monolayer for 1C-LJ model at 150 K
A
and 3C-LJ model at 194 K are presented in Figure 4(c,d),
where N is the ensemble average of the number of molecules in
different configurations are observed for the two models,
the system, ρg is the density of gas-phase calculation using the
which are the hexagonal packing by the 1C-LJ model and
equation of state, Vacc is accessible volume and A is the surface
the 3C-LJ particles are mainly lying flat on the surface.
area of the solid.
The accessible volume is determined for each potential
Above triple point. In Figure 4(b), a Type II adsorption iso-
model and applied in the calculation. It is shown in Figure 3,
therm is observed for both models at temperature above the
the potential profile of CO2 molecule as a function of distance
triple point, with continuous wetting observed, i.e. the loading
between its mass centre to the solid surface, the 3C-LJ model
approaches infinity asymptotically with pressure increased to
exhibits deeper well-depth, and closer zero-potential distance
the bulk vapour pressure. The isotherm obtained by the 1C-
to the surface, implying larger Vacc obtained with the 3C-LJ
LJ model at 273 K is almost identical with that at the equivalent
model. This is in good agreement with the result of Do and
temperature 229 K and traces the experimental data in general
Do [13].
well. In the sub-monolayer region, a good agreement is
observed with both the 1C-LJ and 3C-LJ models, the latter
3.2.1. Adsorption isotherms
diverts from the rest after the completion of the monolayer
Figure 4 shows the surface excess adsorption isotherms on
though, implying better performance of the 1C-LJ model at
graphite at several temperatures for the 1C-LJ and 3C-LJ
the high loading region. To shed further light on these obser-
models, and the comparison with the experimental data at
vations, Henry constant of the models on graphite surface is
194 K [35] and 273 K [38].
analysed.
Below triple point. At the temperature of 194 K, which is below
3.2.2. Henry constant
the bulk triple point, the 3C-LJ model describes the exper-
Henry’s law expresses the amount adsorbed on the surface of a
imental data better than the 1C-LJ model, produces the same
solid at low loading as a linear function of the bulk gas concen-
monolayer capacity as in experiment, while the 1C-LJ model
tration. Given that the surface excess concentration
over-predicts it. Both models capture the occurrence of the
incomplete wetting, i.e. with finite number of layers formed KP
C= (5)
at the saturated vapour pressure, which are consistent with Rg T
where P is the absolute pressure, Rg is the gas constant, T is the
temperature of the system and K is the Henry constant. The
Henry constant K is a measure of the interaction between a
single molecule and the solid. It can be a useful tool to evaluate
the interaction strength between a fluid molecular and adsor-
bent [40].
Moreover, the Henry constant K is dependent on the expo-
nential of the reciprocal temperature and is therefore lower at
higher T. To reveal the effect of adsorbate–adsorbent inter-
actions and the influence of the adsorbate–adsorbate inter-
action, the surface excess concentration (5) can be rewritten
as follows:
 
K′ P
C= (6)
Rg T P0
where K′ = KP0 is a modified Henry law parameter, K rep-
resents the Henry constant and implies for the adsorbate–
Figure 3. (Colour online) Profile of solid–fluid potential energy versus the dis-
tance between the mass centres of CO2 molecule to the solid surface, 3C-LJ adsorbent interaction, and the saturation vapour pressure P0
model is parallel to the surface to maximise the potential. denotes the parameter for the interaction between adsorbate
Downloaded from https://iranpaper.ir

6 X. LIU ET AL.

Figure 4. (Colour online) Comparison of adsorption isotherms on graphitic surface at temperatures of (a) 194 K and (b) 273 K for both models and monolayer snap-
shots of CO2 for (c) 1C-LJ model at 150 K and (d) 3C-LJ model at 194 K, both at P0.

molecules. As a result, the implication of K′ is the relative intermolecular interaction between CO2 molecules. However,
importance of these two interaction potentials. 1C-LJ model has decreased in the modified Henry constant
As shown in Figure 5, as temperature T increases, the (K′ ), implying that the adsorbate–adsorbent interaction is
measured parameter of the adsorbate–adsorbent interaction stronger than the intermolecular interaction between CO2
K decreases while the measured parameter of the adsor- molecules. These results also indicate that the 1C-LJ model
bate–adsorbate interactions P0 increases and the temperature tends to be hydrophilic on the graphitic surface while the
dependence of K′ therefore provides a mean of estimating the 3C-LJ model tends to be hydrophobic on the graphitic sur-
relative importance of these two interactions. If K′ decreases face. This explains (1) at the same/equivalent temperature,
with temperature T, then the adsorbate–adsorbent interaction the monolayer is built up with a sharper slope with the 1C-
tends to be dominant. On the other hand, the adsorbate– LJ than the 3C-LJ model (see Figure 4(a,b)); (2) the saturated
adsorbate interaction potentials would be great as K′ increases loading of 3C-LJ model is lower than the 1C-LJ model at low
with temperature T [41]. For 3C-LJ model, the modified temperature; (3) as the temperature is increased, i.e. higher
Henry constant (K′ ) increases with temperature T, indicating thermal mobility, adsorbate–adsorbate interaction become
that the adsorbate–adsorbent interaction is weaker than the more dominating, a thick adsorbed film is formed by the

Figure 5. (Colour online) Comparison of adsorption isotherms (on a logarithmic scale) on graphitic surface at various temperatures for (a) 1C-LJ and (b) 3C-LJ models.
Downloaded from https://iranpaper.ir

MOLECULAR SIMULATION 7

3C-LJ model at a coexistence pressure P0∗ that is smaller than selected, with the 3C-LJ model produces slightly higher satur-
the bulk saturated vapour pressure P0 . ation densities. This agrees with the Do et al. [13] that for slit
Furthermore, the comparison between the heat of evapor- pore with sizes larger than 1.5 nm, the packing effects caused
ation and the isosteric heat of adsorption at zero loading q0st is by the shape of the molecules are negligible. As the pore size
an indication of the occurrence of the wetting or non-wetting is increased to 3.5 nm, there is a reflection point observed for
behaviour for the adsorbate/adsorbent pair. The heat of evapor- both models, with the 3C-LJ model reaches the inflection
ation is a measure of the cohesion in the bulk, while the q0st is a point at lower reduced pressure, due to the stronger intermole-
measure of adhesion strength to the surface. As shown in Figure cular interaction by considering the quadrupole interaction.
2(e), for both models, without considering the adsorbate–
adsorbate interactions, the triple points demarcate the wetting 3.3.2. At equivalent temperature T1 – below triple point
and non-wetting temperature ranges, that is, the q0st is smaller Figure 7 shows the isotherms at T1 in two slit pores, the 1C-LJ
than sublimation heat when below the triple point, non-wetting and 3C-LJ models diverge from one another in both slit pores.
is expected. However, this demarcation can be affected when the The following observations can be made:
amount of adsorbate is increased [18]. This is illustrated in
Figure 4(a), incomplete wetting is observed at temperatures (1) Although incomplete wetting occurred for adsorption on
below the triple points for both models. graphite surface (see Figure 4(a)), due to the enhanced
force filed with the presence of the opposite wall, in 2 nm
slit pore, Type HI hysteresis according to the IUPAC
3.3. Adsorption in slit pores
classification are observed for both 1C-LJ and 3C-LJ
3.3.1. 298 K models. However, the 1C-LJ model predicts a larger hyster-
Figure 6 presents the adsorption isotherms at 298 K of the 1C-LJ esis and higher saturation loading, which is even higher
and 3C-LJ models in two slit pores, with sizes of 2 and 3.5 nm, than the bulk density despite that the two models predict
both of which fall in the mesopore range, and corresponds to the similar bulk density at T1 (see Figure 2(a)). Moreover,
SH and BH of the wedge pore defined in this work. Identical the condensation and evaporation occurred at higher
reversible type I (IUPAC) adsorption isotherms are observed reduced pressures with the 3C-LJ than the 1C-LJ model.
in the 2 nm slit pore, while divergence appeared when the (2) In the 3.5 nm slit pore, Type H1 hysteresis is again
pore size is enlarged to 3.5 nm. The two CO2 models predict observed with 1C-LJ model and condensation and evapor-
almost the same adsorption capacities for the two pore sizes ation occur at higher pressures compared to in the 2 nm

Figure 6. (Colour online) Comparison of adsorption isotherms at 298 K for 1C-LJ and 3C-LJ models in slit pores with the pore size of (a) 2 nm and (b) 3.5 nm.

Figure 7. (Colour online) Comparison of isotherms at temperature T1 for 1C-LJ and 3C-LJ models in slit pores with the pore size of (a) 2 nm and (b) 3.5 nm.
Downloaded from https://iranpaper.ir

8 X. LIU ET AL.

slit. However, the 3C-LJ CO2 failed to condense with results at T1, some features remain the same, including (1)
incomplete wetting occur on pore walls. 1C-LJ model predicts higher saturated density than 3C-LJ
(3) In the low-pressure region, the two pore walls act as two model; (2) the monolayers are built at lower relative pressure
independent graphite surfaces, therefore, identical fea- with 1C-LJ model, although the discrepancy between the two
tures observed earlier with adsorption on graphic surfaces models is reduced with increased temperature. Other major
are again captured here. That is the monolayer of the 1C- impacts caused by the temperature rise are:
LJ model is formed at lower reduced pressure and picked
up with a rather sharp manner than the 3C-LJ model, due
to the 1C-LJ model has stronger interaction with the solid (1) The hysteresis vanished in the 2 nm pore for both models,
substrate than the 3C-LJ model. instead a reflection point and vertical reversible transition
were observed for 1C-LJ model, and a more gradual build-
ing up process with 3C-LJ model.
3.3.3. At equivalent temperature T2 – above triple and (2) In the 3.5 nm pore, the hysteresis of 1C-LJ model becomes
below critical point much smaller, the 3C-LJ model fills the pore, instead of
Figure 8 illustrates the adsorption isotherms of both models at only partially wet the surfaces due to increased thermal
T2 in the slit pores of width 2 and 3.5 nm. Compare with mobility by increasing the temperature from T1 to T2.

Figure 8. (Colour online) Comparison of adsorption isotherms at temperature T2 for 1C-LJ and 3C-LJ models in slit pores with the pore size of (a) 2 nm and (b) 3.5 nm.

Figure 9. (Colour online) Comparison of adsorption isotherms for 1C-LJ and 3C-LJ models in the wedge pore at various temperatures: (a) 298 K, (b) T1, (c) T2. The pores
are 8.5 nm in axial length.
Downloaded from https://iranpaper.ir

MOLECULAR SIMULATION 9

3.4. Adsorption in wedge pores 3.4.1. Effects of temperature


The possible disputation about the way of defining the equiv-
The adsorption isotherm in the wedge pore at different temp-
alent temperature was aware of, although intention was to
eratures is presented in Figure 9. The wedge pore has an axial
evaluate the two models at the same thermodynamic status.
length of 8.5 nm, widths of 2 nm (small end) and 3.5 nm (large
To eliminate possible discrepancies caused by this definition,
end), and tilt angle α = 5°. Some of the differences between the
this section systematically illustrates the evolution of the per-
two models that observed in the slit pores are also captured
formances of the two models in the wedge pore as a function
with the wedge pore, i.e. (1) higher saturated pore density pre-
of temperature. Figure 11 shows the adsorption isotherms for
dicted by the 1C-LJ model at T1 and T2, (2) the monolayer for-
temperatures ranged between 150 K (below the triple point)
mation is facilitated with 1C-LJ model at the same equivalent
and 273 K (above the triple point). The following observations
temperature, due to its stronger interaction with solid and (3)
can be made.
the discrepancy between two models diminishes with tempera-
ture. Furthermore, the type H2b hysteresis (IUPAC classifi-
cation) and step-wise desorption that has been observed as (1) Figure 11(a) depicts Type H2(b) hysteresis (of IUPAC),
the major features for adsorption in wedge pores, was repro- which is then transformed into Type H1 for temperatures
duced with the 1C-LJ model at T1 but not the 3C-LJ model. increased from 150 to 173 K. The closure point of the hys-
The latter instead predicts a smaller H1 type of hysteresis teresis loop shifts to a higher relative pressure and smaller
(see Figure 9(b)). The reason is due to the packing effects of size until it is completely vanished at 216.6 K, instead a
CO2 molecules, as it has been known that the alternating com- sharp transition is observed. In the desorption boundary
mensurate and incommensurate are accounted for the step- of the hysteresis, the number of steps remains un-changed
wise desorption behaviour, which was observed with 1C-LJ between 150 and 160 K. Once the temperature reaches
model at T1 (see Figure 10(a)) in consistent with other noble 173 K, which is above the triple point, the last two steps
gases [9–11], but not the 3C-LJ model. This indicates the are gradually smoothed out. This implies that a critical
impacts of the quadrupole and molecular shape on the pack- temperature of this step-wise behaviour is between 172
ing. The alternative packing of the 1C-LJ model no longer and 174 K for the reference wedge pore.
exists with temperature increased. As seen in Figure 10(b), (2) Figure 11(b) clearly shows with the 3C-LJ model, when
the configuration clearly appears to be disordered. temperature is no higher than 170 K, the fluid only can
partially wet the surface. At 194 K, Type H1 in IUPAC

Figure 10. (Colour online) Snapshots of CO2 molecules in the reference wedge pore at saturation vapour pressures for (a) 1C-LJ model (at temperatures below triple
point) and (b) 1C-LJ (at temperatures above triple point).

Figure 11. (Colour online) Isotherms of CO2 adsorption corresponding to the reduced pressure for the reference wedge pore at various temperatures using (a) 1C-LJ
model (triple point = 172 K) and (b) 3C-LJ model (triple point = 216 K).
Downloaded from https://iranpaper.ir

10 X. LIU ET AL.

is observed, and 216 K is the hysteresis critical tempera- Funding


ture for the studied wedge pore, i.e. above which the iso- This work was supported by resources provided by the Pawsey Supercom-
therm becomes reversible. However, the pore critical puting Centre with funding from the Australian Government and the
temperature should be above 216 K, where exhibits no Government of Western Australia.
sharp jump in adsorption isotherm, which implies no
occurrence of condensation.
(3) In the monolayer region, the 1C-LJ model describes oppo- References
sitely as the 3C-LJ model for the trend of the uptake rate/ [1] Dantas S, Struckhoff KC, Thommes M, et al. Phase behavior and
pressure as a function of temperature. The uptake of 1C-LJ capillary condensation hysteresis of carbon dioxide in mesopores.
CO2 and completion of the monolayer shifts to higher Langmuir. 2019;35(35):11291–11298.
reduced pressure with temperature, and this is consistent [2] Mukherjee A, Okolie JA, Abdelrasoul A, et al. Review of post-com-
bustion carbon dioxide capture technologies using activated carbon.
with experimental data [1,17,42]. Same trend can be J Environ Sci. 2019;83:46–63.
observed on isotherm results using a uniform slit meso- [3] Siegelman RL, Kim EJ, Long JR. Porous materials for carbon diox-
pore and graphitic surface with 1C-LJ model, while 3C- ide separations. Nat Mater. 2021;20:1060–1072.
LJ model consistently describes in the opposite way. [4] Liu X, Fan C, Do D. Microscopic characterization of xenon adsorp-
tion in wedge pores. Adsorp. 2019;25:1043–1055.
[5] Loi Q, Prasetyo L, Tan J, et al. Wedge pore modelling of gas adsorp-
4. Conclusions tion in activated carbon: consistent pore size distributions. Carbon.
2020;166:414–426.
This work has presented a systematic simulation of CO2 bulk [6] Loi Q, Prasetyo L, Tan J, et al. Order-disorder transition of an argon
phase coexistence properties and its adsorption on graphitic adsorbate in graphitic wedge pores. Chem Eng J. 2020;384:123286.
[7] Norman G, Filinov V. Investigation of phase transitions by a Monte
surfaces, in slit and wedge pores by evaluating the performance Carlo method. High Temp (USSR). 1969: 216–222.
of two CO2 potential models (i.e. 1C-LJ and 3C-LJ). An equiv- [8] Frenkel D, Smit B. Understanding molecular simulation. 2nd ed.
alent temperature was introduced for comparing 1C-LJ and New York: Academic Press; 2002.
3C-LJ models at the same thermodynamic condition. Simu- [9] Hammonds K, McDonald I, Tildesley D. Computational studies of
lations with 3C-LJ model accounting for the presence of a the structure of carbon dioxide monolayers physisorbed on the
basal plane of graphite. Mol Phys. 1990;70(2):175–195.
quadrupole moment have better agreement for the vapour– [10] Murthy C, O’Shea S, McDonald I. Electrostatic interactions in mol-
liquid and vapour–solid equilibria, and adsorption, especially ecular crystals lattice dynamics of solid nitrogen and carbon diox-
at the temperature below triple point and at high-pressure ide. Mol Phys. 1983;50(3):531–541.
region. On the other hand, the 1C-LJ model has better agree- [11] Potoff J, Siepmann J. Vapor-liquid equilibria of mixtures containing
ment with the experimental data at low-pressure region. The alkanes, carbon dioxide, and nitrogen. AIChE J. 2001;47(7):1676–
1682.
difference between simulation results and experimental data [12] Vishnyakov A, Ravikovitch PI, Neimark AV. Molecular level
for both models is negligible at the temperature above triple models for CO2 sorption in nanopores. Langmuir. 1999;15
point. (25):8736–8742.
Simulations for 1C-LJ model at the temperature below the [13] Do D, Do H. Effects of potential models on the adsorption of carbon
triple point show occurrence of step-wised hysteresis in dioxide on graphitized thermal carbon black: GCMC computer
simulations. Colloids Surf A. 2006;277(1-3):239–248.
wedge-shaped pore. This occurrence is well-aligned with [14] Golebiowska M, Firlej L, Kuchta B, et al. Structural transformations
other simple gases. It is also evident that the consideration of of nitrogen adsorbed on graphite: Monte Carlo studies of spatial
shape and quadrupole of CO2 are significant in the correct heterogeneity in multilayer system. J Chem Phys. 2009;130
description of adsorption isotherm and packing density in (20):204703.
slit and wedge pores, especially at temperatures below the tri- [15] Kuchta B, Etters RD. Calculated properties of monolayer and multi-
layer N2 on graphite. Phys Rev B. 1987;36:3400–3406.
ple point. [16] Xu H, Phothong K, Do D, et al. Wetting/non-wetting behaviour of
In the monolayer region, the 3C-LJ model describes in the quadrupolar molecules (N2, C2H4, CO2) on planar substrates. Chem
opposite trend as compared to the 1C-LJ model. Same obser- Eng J. 2021;419:129502.
vations have been found on isotherms of the slit pore and gra- [17] Terlain A, Larher Y. Phase diagrams of films of linear molecules
phitic surface. As the temperature increases, the uptake of 1C- with large quadrupole moments (CO2, N2O, C2N2) adsorbed on
graphite. Surf Sci. 1983;125(1):304–311.
LJ model and completion of the monolayer shift to higher [18] Xu H, Zeng Y, Do DD, et al. On the nonwetting/wetting behavior of
reduced pressure. These observations are consistent with carbon dioxide on graphite. J Phys Chem C. 2019;123(14):9112–
experimental data [1,17,42]. 9120.
[19] Steele WA. Monolayers of linear molecules adsorbed on the graph-
ite basal plane: structures and intermolecular interactions.
Acknowledgement Langmuir. 1996;12(1):145–153.
[20] Do D, Junpirom S, Nicholson D, et al. Importance of molecular
Authors X. Liu and A.H.H Sim contributed equally to this work. This shape in the adsorption of nitrogen, carbon dioxide and methane
work was supported by resources provided by the Pawsey Supercomput- on surfaces and in confined spaces. Colloids Surf A. 2010;353
ing Centre with funding from the Australian Government and the Gov- (1):10–29.
ernment of Western Australia. [21] Fan C, Do D, Nicholson D. A new and effective Bin–Monte Carlo
scheme to study vapour–liquid equilibria and vapour–solid equili-
bria. Fluid Phase Equilib. 2012;325:53–65.
Disclosure statement [22] Do D, Do H. Pore characterization of carbonaceous materials by
DFT and GCMC simulations: A review. Adsorption Sci Technol.
No potential conflict of interest was reported by the author(s). 2003;21(5):389–424.
Downloaded from https://iranpaper.ir

MOLECULAR SIMULATION 11

[23] Ravikovitch PI, Vishnyakov A, Russo R, et al. Unified approach to [34] Lemmon E, Mclinden M, Friend D. Thermophysical properties of
pore size characterization of microporous carbonaceous materials fluid system in NIST chemistry webbook. Gaithersburg (MD):
from N2, Ar, and CO2 adsorption isotherms. Langmuir. 2000;16 NIST Standard Reference Database Number 69; 2018.
(5):2311–2320. [35] Spencer WB, Amberg CH, Beebe RA. Further studies of adsorption
[24] Velasco L, Guillet-Nicolas R, Dobos G, et al. Towards a better on graphitized carbon blacks. J Phys Chem. 1958;62(6):719–723.
understanding of water adsorption hysteresis in activated carbons [36] Chen B, Siepmann JI, Klein ML. Direct Gibbs ensemble Monte
by scanning isotherms. Carbon. 2016;96:753–758. Carlo simulations for solid-vapor phase equilibria: applications to
[25] Kohmuean P, Inthomya W, Wongkoblap A, et al. Monte Carlo Lennard-Jonesium and carbon dioxide. J Phys Chem B. 2001;105
simulation and experimental studies of CO2, CH4 and their mixture (40):9840–9848.
capture in porous carbons. Molecules. 2021;26(9):2413. [37] Do D, Nicholson D, Do H. On the henry constant and isosteric heat
[26] Rosalind EF. Crystallite growth in graphitizing and non-graphitiz- at zero loading in gas phase adsorption. J Colloid Interface Sci.
ing carbons. Proc R Soc London, Ser A. 1951;209:196–218. 2008;324(1):15–24.
[27] Bojan M, Steele W. Computer simulation of physisorption on a het- [38] Guillot A, Stoeckli F. Reference isotherm for high pressure adsorp-
erogeneous surface. Surf Sci. 1988;199(3):L395–L402. tion of CO2 by carbons at 273 K. Carbon. 2001;39(13):2059–2064.
[28] Bojan M, Steele W. Computer simulation of physical adsorption on [39] Do D, Do H. Appropriate volumes for adsorption isotherm studies:
stepped surfaces. Langmuir. 1993;9(10):2569–2575. The absolute void volume, accessible pore volume and enclosing
[29] Steele W. The physical interaction of gases with crystalline solid: particle volume. J Colloid Interf Sci. 2007;316(2):317–330.
I. gas-solid energies and properties of isolated adsorbed atoms. [40] Zeng Y, Do DD, Nicholson D. Existence of ultrafine crevices and
Surf Sci. 1973;36(1):317–352. functional groups along the edge surfaces of graphitized thermal
[30] Allen MP, Tildesley DJ. Computer simulation of liquids. Oxford: carbon black. Langmuir. 2015;31(14):4196–4204.
Oxford University Press; 1989. p. 385. [41] Nguyen VT, Horikawa T, Do D, et al. On the relative strength of
[31] Kirkwood JG, Buff FP. The statistical mechanical theory of surface adsorption of gases on carbon surfaces with functional groups:
tension. J Chem Phys. 1949;17:338–343. fluid–fluid, fluid–graphite and fluid–functional group interactions.
[32] Frenkel D, Smit B, Tobochnik J, et al. Understanding molecular Carbon. 2013;61:551–557.
simulation. Comput Phys. 1997;11(4):351–354. [42] Bottani E, Bakaev V, Steele W. A simulation/experimental study of
[33] Mountain R, Thirumalai D. Quantative measure of efficiency of the thermodynamic properties of carbon dioxide on graphite. Chem
Monte Carlo simulations. Phys A. 1994;210(3-4):453–460. Eng Sci. 1994;49(17):2931–2939.

You might also like