You are on page 1of 10

Fluid Phase Equilibria 531 (2021) 112909

Contents lists available at ScienceDirect

Fluid Phase Equilibria


journal homepage: www.elsevier.com/locate/fluid

Phase behaviour of confined associating fluid in a functionalized slit


pore: a Monte Carlo study
Sashanka Sekhar Mandal a, Sudhir Kumar Singh b, Sanchari Bhattacharjee a, Sandip Khan a,∗
a
Department of Chemical & Biochemical Engineering, Indian Institute of Technology Patna, Patna 801103, India
b
Department of Chemical Engineering, Thapar Institute of Engineering and Technology, Patiala 147004, India

a r t i c l e i n f o a b s t r a c t

Article history: Vapour-liquid phase equilibria of associating fluids under confinement are investigated using Grand-
Received 15 August 2020 Canonical Transition Matrix Monte Carlo (GC-TMMC) method. Various properties of coexistence phases
Revised 15 November 2020
and critical properties are estimated for multiple associating fluids at different slit widths. The structural
Accepted 18 November 2020
features of coexistence phases are examined through monomer fraction, density profile, and orientation
Available online 24 November 2020
profile. The role of surface sites on the phase behaviour under confinement is critically examined. It has a
Keywords: significant effect on the structural behaviour of coexistence phases, leading to a change in the properties
Phase transition of coexistence phases. In particular, the orientation of associating fluid molecules near the functionalized
Associating fluid surface is entirely different than in the case of a smooth surface. The orientation of two sites associating
Slit pore (when sites are located oppositely to each other) fluid molecules is normal to surface, forming bonds
Functionalized surface with surface sites in one direction and another bond with molecules in an adjacent layer.
© 2020 Published by Elsevier B.V.

1. Introduction level. The advantages of statistical mechanics and computer simu-


lation made it possible now to study these systems. Molecular sim-
The fluid phase behaviour in a nanopore is significantly dif- ulations have been used for many years to study the properties of
ferent than those in a bulk phase [1–6]. The relative strength molecular systems using mainly two approaches, namely, molecu-
between fluid-fluid and fluid-substrate interaction induces vari- lar dynamics (MD) and Monte Carlo(MC) [19,23].
ous phase transitions, including layering, prewetting, vapour-liquid, In this direction, many investigations are performed using
melting/freezing, capillary condensation, along with a shift in crit- molecular simulation for model fluids and molecular fluids con-
ical properties of the system [7–18]. A detailed understanding of fined at the nanoscale to understand the various phase transitions
such systems will provide significant insight into many biological, [6,16,17,24–32]. For example, square-well fluids are studied in a
natural, and engineering processes such as shale gas extraction, slit pore from quasi 3D to 2D system to understand the effect of
microfluidics devices, fabrication of nanomaterials, nanotribology, confinement on the vapour-liquid transition. They observed four to
etc. [19]. Over the last few decades, substantial technological ad- five distinct linear regimes based on the critical temperature shift
vancement in nano-fabrication technology made it possible to im- with slit width, clearly visible at higher surface strength. More-
print nano-scale surfaces with well-defined chemical, physical and over, the change in the critical pressure and critical density does
geometric characteristics [20]. Modern electronic devices are made not show any specific trend; however, they approached the 3D
possible by extensive use of micro-electromechanical systems bulk value with an increase in slit pore width beyond 40 molec-
(MEMS) and nanoelectromechanical systems (NEMS). Micro/nano- ular diameters [29]. In another work [33,34] triangle-well fluids
fluidic devices like Lab-on-a-chip (LOC) have the potential to with variable well widths confined in different slit pores are in-
change the way we diagnose complex samples in medical, envi- vestigated using Monte Carlo simulation. An interesting observa-
ronmental, and many other industries [21,22]. Although we have a tion of this study is that inside the narrow slit pore the density
relatively good understanding of phenomena like adsorption, cap- profile shows a higher value at the centre than to that near the
illary action, phase transition, etc. at a macroscopic scale, new in- slit surface. In another study, MC simulation is used to investigate
sight is needed as the dimensions reduce to the molecular/atomic the effect of pore shape on the vapour-liquid phase transition of
square-well fluid. In this work, cylindrical pores of varying pore
width and surface nature are investigated. This investigation in-

Corresponding author. dicates that the surface nature can significantly alter the vapour-
E-mail address: skhan@iitp.ac.in (S. Khan). liquid coexistence envelopes and the critical point [35]. Vapour-

https://doi.org/10.1016/j.fluid.2020.112909
0378-3812/© 2020 Published by Elsevier B.V.
S.S. Mandal, S.K. Singh, S. Bhattacharjee et al. Fluid Phase Equilibria 531 (2021) 112909

liquid phase equilibria of square-well fluids confined in patterned by MD study. The critical temperature is found to decrease with
slit pores are also investigated using Monte Carlo simulations [28]. an increase in the external electric field. However, critical density
In this work, phase coexistence envelopes, spreading pressure, crit- increases for graphite pore but decreases in mica pore with an in-
ical properties, and local density profiles of the confined square- crease in the electrical field.
well fluid are reported for chemically and physically patterned slit Although many studies on confined polar fluids such as water,
pores. Liu et al. [36] studied the effect of hard slit pore on the alkanes, etc. but very few studies explore the functionalized sur-
vapour-liquid coexistence and critical properties of the confined face effect on the phase transition of associating fluids [43]. In this
Lenard-Jones (LJ) fluid and ST2 model of water using GCMC sim- direction, we have extensively studied the phase behaviour of two-
ulations. In this study, they observe three distinct regions based site associating fluids confined in a functionalized slit pore to un-
on the shift in critical temperature, critical density, and saturated derstand the role of associating sites present on the surface. The
liquid density as a function of slit width while crossing from 2- presence of sites on the fluid molecules favor the bond formation
dimensional to a 3-dimensional system. and leads to a chain-like structure for two-site associating fluids.
Along with simple fluids, molecular fluids such as water, alkane, Therefore, it is quite expected that there would be an intense layer-
alkanes, etc. are also studied to understand the effect of structural ing of associating fluid inside the pores. On the other hand, surface
integrity and the polarity of a molecule on the various phase tran- sites can compete with the fluid-fluid association and may trigger
sitions under confinement. In this direction, Brovchenko and his fluid molecules different orientation and result in different struc-
co-workers [17,37–39] have extensively studied the phase transi- tural behaviour inside pores.
tions of water in nanopores using Gibbs Ensemble Monte Carlo
(GEMC) simulation and observed various kind of phase transi- 2. Computational model and simulation methods
tions including vapour-liquid, layering, prewetting based on the
strength of wall-fluid interactions. In particular, the vapour-liquid 2.1. Model detail
phase transition was observed in hydrophobic and moderately hy-
drophilic pores, whereas surface phase transitions such as prewet- In this work, a two-site associating fluid model is considered
ting, layering, etc. were observed for strong substrate interac- in which Lenard-Jones potential is used for isotropic interaction
tion. Singh et al. [31] uses configurational-bias grand-canonical between fluid molecules and a short-range directional square-well
transition-matrix Monte Carlo simulations to investigate various potential is used for site-site interaction [44]. The sites are located
thermophysical properties of alkanes (C1 - C8) in slit pore of oppositely on the surface of fluid molecules (off-cantered). Thus,
graphite and mica along with in bulk. The walls of slit pore geome- the complete potential for fluid-fluid interaction is as follows:
try are taken as a featureless smooth surface. Wall-fluid interaction  
u f f ri j , θi , θ j = uLJ−tr (ri j ) + ua f (ri j , θi , θ j ),
is depicted by the 9-3 Steele potential. The critical temperature is 
found to be sensitive to the nature of the surface, whereas critical   −εa f if σ <ri j < rc , θi < θc and θ j < θc ,
density decreases with the slit width until the system approaches
u a f r i j , θi , θ j =
0 Otherwise
   (1)
single layer 2D geometry. Bonnaud et al. [40]investigated the struc-
12  r 6 
4ε σi j
r
ture and dynamics of water in a hydroxylated silica nanopore us- − σi j if ri j ≤ rcut ,
uLJ−tr (ri j ) =
ing MD simulation and found multilayer adsorption prior to the
0 Otherwise
capillary condensation. The study also shows a strong layering of
water in the vicinity of the silica surfaces. They also report the where θ i , θ j are the angle between the centre-centre and centre-
MC study of confined water in cementitious material (calcium- site vectors for the molecule i and molecule j, respectively. The θ c =
silicate- hydrates) in Grand Canonical ensemble [30]. This study 27o and rc =1.0 are the angle, and the distance criteria, respectively,
shows how the relative humidity can affect the phase behaviour of to consider two molecules are in association with the well-depth
water confined in different pore sizes of calcium-silicate-hydrates of ε af . Here, rc is the distance between two fluid particles. There-
grain. The kinetics of capillary condensation and water evapora- fore, the association of particles can occur if the distance between
tion in a smooth, ideal hydrophilic mesopore were investigated us- two particles is less than or equal to 1. Many groups have also
ing combined GCMC and MD by Yamashita et al. [27]. The effect used these parameters and predicted various phase transitions us-
of the deviation in chemical potential between the gas phase and ing classical Density Functional Theory (DFT) and MC simulation
pore surface on the dynamics of water evaporation and capillary [45,46]. The size (σ ) and energy (ε ) parameters for LJ potential of
condensation was explored. GCMC method was employed to con- fluid-fluid interaction are taken as unity with a cut-off distance of
trol the chemical potential of water vapour, whereas MD depicts rcut =2.5. This short-ranged and orientationally dependent site-site
the dynamics of water adsorption/desorption on/from the pore sur- interaction can be compared with hydrogen bonding between two
face. Srivastava et al. [26] studied the phase transition of confined molecules. For example, two-site associating fluid can represent
water in graphite and mica slit pores of different sizes ranging the hydrogen fluoride (HF) molecules, which can form two hydro-
from 10Å to 60 Å. Their study found that the confinement de- gen bonds with neighbour molecules and can make a linear chain.
creases the critical temperature and introduces non-uniformity in However, the associating strength of the site interaction in the case
density profile. This heterogeneity leads to different vapour-liquid of HF is comparatively large in comparison to the value used in
coexistence densities at different layers. Hydrophilic surfaces like this work (ε af is varied from 4 to 10, which allows us to sample the
mica reduced the critical temperature of confined water more ef- phase space efficiently and determine the critical properties accu-
fectively than hydrophobic graphite surfaces. The water molecules rately). It should be noted that all the parameters are reduced with
at the surface layer of the pore behave more like a quasi-2D fluid. respect to fluid-fluid interaction parameters (ε , σ ), such that pore
The effects of the electric field on the vapour-liquid equilibria of width H∗ =H/σ , number density ρ ∗ = ρσ 3 , temperature T∗ =KB T/ε
methanol and ethanol, confined in a graphite slit pore, is reported (where KB is Boltzmann constant) etc.
by Bhandary et al. [41]. The critical temperature of ethanol and The fluid molecules are confined by the structure less surface
methanol decreases with the increasing external electrical field to with active sites at different slit pore width. The sites are dis-
a certain minimum value and then increases. The orientation or- tributed uniformly in a rectangular grid on the surface based on
der for both alcohols is found to increase with a stronger external the site density, ρ s , defined as the number of associating surface
electrical field. Srivastava et al. [42] report the effects of the ex- sites per unit area. The substrate-fluid interaction is represented by
ternal electrical field on confined water in graphite and mica pore steele 10-4-3 potential [47], which corresponds to argon-graphite

2
S.S. Mandal, S.K. Singh, S. Bhattacharjee et al. Fluid Phase Equilibria 531 (2021) 112909

the histogram reweighting technique is used to locate the coex-


istence chemical potential at which the probability of liquid and
vapour phase are the same. Details of GC-TMMC simulation tech-
niques are given elsewhere [50]. The coexistence probability distri-
bution is used to estimate the saturation pressure (P) using Eq. (3).

c (N )
β PV = l n − l n (2 ) (3)
c (0 )
where, β = 1/KB T
The density difference between the two coexistence phases de-
creases as the system approaches the critical point. Thus, it is chal-
lenging to estimate the critical properties from a direct simulation.
The following scaling laws are used to determine the critical prop-
erties from the subcritical parameters [54]:
Fig. 1. Schematic diagram of-two-site associating fluid confined in a functionalized
slit pore. ρl − ρv = A(T − Tc )B (4)

ρ + ρ
interaction [48] along with active site interaction as given below: v
l
= ρc + D(T − Tc ) (5)
2
uw f = u10−4−3 (z ) + uaw (ri j , θ
i, θ j )   where subscript “c” denotes the corresponding critical properties.
 σw f 10  σw f 4 σw4 f A, B and D are fitting parameters. ρl and ρv are liquid and vapor
u10−4−3 (z ) = 2π ρw εw f σ 2
 2
− −
 wf 5 z z 3(z+0.61)3 density at subcritical regime.
−εaw i f ri j < rc , θi < θc and θ j < θc
uaw (ri j , θi , θ j ) =
0 Otherwise
2.3. Simulation details
(2)
All simulations are carried out in a rectangle box with a length
where z is the distance between fluid and substrate molecules
of 9 in x and y direction, and the z dimension is equal to the width
along z-coordinate,  is the distance between two graphite lay-
of the slit pore, H. The top and bottom surfaces are placed in the
ers, and ρ w is the density of carbon atom of graphite layers. The
xy plane of the box at z = 0 and z = H. In this work, the pore
effective interaction parameters ε wf and σ wf between fluid and
width is varied from 4 to 40 molecular diameter of fluid parti-
substrate molecules are calculated through Lorentz-Berthelot rule
cles, and the associating strength is varied from 0 to 10. The differ-
[49]. However, the energy parameter for solid-fluid interaction is
ent MC moves are carried out with probability 0.1, 0.35, 0.35, 0.1
taken weaker (ε wf = 0.2) to explore the effect of surface sites
and 0.1 for translation, insertion, deletion, rotation and Unbonding-
on the phase diagram in confinement, whereas σ wf is equal to
bonding (UB) respectively. The UB technique [55] is used to bias
0.9912, which is correspond the argon-graphite system. The site-
the sampling towards the bonding regime of associating fluids,
site interaction between associating fluid and the functionalized
which is particularly important at lower temperature range and
surface is also represented by square-well potential with a cut off
higher associating strength. The attractive part of the GC-TMMC is
distance of rc =1.0, and with associating strength of ε aw =5, which
that the range of particle state for a specific simulation can be dis-
is an intermediate value for the range of site-site interaction of
cretized among different cores [56]. Hence, multiple cores (8-32)
fluid particles to understand the interplay between fluid-fluid and
are used based on the range of particle state that depends on the
fluid-solid interaction on the phase diagram.
system temperature and the associating strength. In this work, the
In this study, the surface is placed at z = 0 and z = H (pore
statistical errors, calculated from four independent simulations, are
width), and active sites are placed on the top of the virtual sur-
of the order of symbol size.
face atoms. The centre of virtual molecules with the same radius
of fluid molecules, is located at z=0 and z =H. The schematic dia-
gram of the system is shown in Fig. 1. 3. Result and discussion

2.2. Methodology 3.1. Phase transition inside smooth slit pore

Grand Canonical Transition Matrix Monte Carlo [50] (GC-TMMC) We start our discussion with the vapour-liquid phase transition
simulation, along with histogram reweighting [51] is employed to of confined associating fluids with various associating strength at
extract the properties of the coexistence phases under confine- a slit width, H∗ = 6, as shown in Fig. 2(a). In general, the vapour
ment. GC-TMMC is basically Grand Canonical Monte Carlo (GCMC) density increases and liquid density decreases with a temperature
simulation conducted at constant chemical potential μ, volume V, rise and become indistinguishable at a critical temperature. With
and temperature T in which transition probability among differ- an increase in associating strength, the critical temperature of the
ent microstates are also computed during different Monte Carlo coexistence phases increases, which is more sensitive at higher as-
(MC) moves and collected in a matrix form. There are four basic sociating strength ε af ∗ > 6. The increase of critical temperature
moves, namely displacement, insertion and deletion and, rotation is mainly due to the strong association of fluid molecules, which
[52], which are used to sample the phase space efficiently. The increases the free energy barrier between the vapour and liquid
macro-state probability (probability distribution of particle states, phase at a given temperature and raises the critical temperature.
N ) is then evaluated from the collection matrix. At regular inter- Therefore, at a given temperature, the liquid density increases, and
vals, the multi-canonical sampling technique [53] is used to sam- vapour density decreases with an increase of associating strength,
ple all the particle states uniformly using macro-state probability which results in the broadening of the phase diagram. Interest-
density obtained from the updated collection matrix. These simu- ingly, an insignificant change in the critical density is observed
lations are performed at a given chemical potential μ, and hence, with the change in the associating strength of the fluid, attributed

3
S.S. Mandal, S.K. Singh, S. Bhattacharjee et al. Fluid Phase Equilibria 531 (2021) 112909

to the symmetrical shift in the coexistence envelope with the as-


sociating strength. We have analysed the system size effect on the
critical properties of the system (ε af ∗ = 6 and H∗ = 6) with the
different surface area. We found the error associate with the es-
timation of critical points is within 1%, which is also reported by
other authors [29] for the simple fluid confined in an attractive slit
pore.
Fig. 2(b) presents the effect of confinement on the coexistence
envelopes of associating fluid for ε af ∗ = 6, at different pore widths
ranging from H∗ =3 to 60. Clearly, under confinement, the coex-
istence envelope shrinks, and the critical temperature decreases.
Moreover, with the increase in the degree of confinement or the
decrease of slit width, the coexistence envelope shrinks further,
and the critical temperature monotonically decreases. However,
critical density under confinement shows nonmonotonic behaviour
with respect to bulk critical density. This behaviour can be at-
tributed to a significant change in coexistence densities with the
change in the slit pore width. Similar nonmonotonic behaviour of
critical density with respect to bulk is observed with simple fluids
[29,31] confined in attractive slit pores. Further, to understand the
phase coexistence behaviour under confinement on the reduced
scale, we have plotted the corresponding state coexistence en-
velopes of associating fluid for ε af ∗ =6 at different pore widths and
shown in Fig. 2(c). The corresponding state behaviour has shown
an interesting trend of coexistence envelopes. The highest shrink-
age in the corresponding state coexistence envelope is observed
with the slit pore width, H∗ = 15, as shown in Fig. 2(c). Further,
with the decrease in the slit pore width, the corresponding state
coexistence envelope broadens, and finally, at H∗ = 4, it overlaps
with the bulk corresponding state behaviour. Moreover, at H∗ = 3,
the coexistence envelope further broadened up and exceeded the
bulk coexistence envelope. Similar behaviour is also observed with
associating fluid for ε af ∗ =4 at the studied pore widths (not shown).
Interestingly, with the largest studied slit pore width, H∗ = 40, the
corresponding state coexistence envelope again starts broadening
as compared to H∗ =15, as reflected in Fig. 2(c), which indicates
that at a much larger slit pore width, it would approach towards
bulk corresponding state coexistence envelope, as expected.

3.2. Monomer fraction

The structure of coexistence phases of associating fluid largely


depends on the association of molecules and can be examined
through monomer fraction present in the system. Monomer frac-
tion is defined as the fraction of molecules that are not bonded
with other molecules. Fig. 3(a) shows the values of the monomer
fraction of coexistence phases for different associating strength at
a slit width, H∗ = 6, as well as for bulk. The solid line with
an open symbol represents the monomer fraction for respective
bulk systems. At a given associating strength, for both the bulk
and confined systems, the values of monomer fraction signifi-
cantly increase for the liquid phase with a decrease in temper-
ature, whereas for the vapour phase, the change of monomer
fraction is very small or negligible, depending on the associat-
ing strength, which is in line with previous studies employed in
bulk [57,58]. Therefore, it is quite expected that there would be
a more compact arrangement in the liquid phase through associa-
Fig. 2. (a). Temperature-density vapour-liquid coexistence envelope of associating tion and result in an increase of liquid density with a decrease in
fluids in a slit pore of width H∗ =6. (b). Vapour-liquid coexistence envelope of an temperature.
associating fluids (ε af ∗ =6) in slit pore of width ranging from H∗ =3 to 60. The phase On the other hand, at a given temperature, the monomer frac-
coexistence envelope of the bulk fluid (black solid line with square symbol) is also tion value for the liquid phase increases and that for the vapour
included in the plot for comparison. All parameters are in reduced unit. (c). Corre-
sponding state of vapour-liquid coexistence envelope of an associating fluid (ε af ∗ =6)
phase decreases under confinement compared to the respective
in slit pore of width ranging from H∗ =3 to 40. The phase coexistence densities of bulk systems. For the liquid phase, the presence of the substrate
the bulk fluid (black solid line with square symbol) are included in the plot for hinders the bond formation of fluid molecules near the surface,
comparison. and hence the monomer fraction in the liquid phase increases.

4
S.S. Mandal, S.K. Singh, S. Bhattacharjee et al. Fluid Phase Equilibria 531 (2021) 112909

Fig. 4. Saturation pressure for different associating strength at H∗ = 6. Solid lines


with open symbols represent the saturation pressure at bulk for corresponding as-
sociating strengths.

a given temperature, which is due to the decrease in saturation


pressure with slit width, discussed in the next section.

3.3. Saturation pressure

The saturated vapour pressure of the system is estimated us-


ing the Clausius-Clapeyron equation and is shown in Fig. 4. As ex-
pected, the saturated vapour pressure decreases linearly with a de-
crease in temperature, which is also observed for various model
fluids [59–61]. The slope of the line decreases with an increase in
associating strength, which is also seen for other associating flu-
ids in bulk [57,58]. The increase in associating strength and the
degree of association in the liquid phase increases the latent heat
of vaporization and, hence, decreases the saturated vapour pres-
Fig. 3. (a). Monomer fraction of different associating fluid at slit width H∗ =6. Here
sure in the system. Therefore, the saturation pressure of associat-
open symbol with solid line represent the respective bulk systems, (b) monomer
fraction at different slit width for associating fluid ε af ∗ =6. ing fluids in confinement increases at a given temperature as the
degree of association decreases. The difference of saturation pres-
sure between bulk and confinement (H∗ =6) at a given temperature
decreases with an increase of associating strength. Therefore, the
effect of confinement on the saturation pressure is less for higher
associating strength fluids.
On the other hand, in the vapour phase, the numbers of fluid
molecules are significantly less, and hence, the association of the
fluid molecules is typically found to be less in bulk. However, the 3.4. Density and orientation profile
presence of attractive surface in the vapour phase, induces strong
layering near the surface (can be observed from the density pro- To understand the effect of association strength on the coex-
file discussed in the next section), and therefore, bond formation istence phases, we examine the number density profile along z-
within the layer (near the surface) increases and hence the de- direction normal to the surface. Fig. 5 represents the z-density pro-
crease in monomer fraction. file for different association strength at a pore width of H∗ =6. Due
The degree of association increases significantly with an in- to the substrate interaction, the fluid molecules are found in an
crease in associating strength, and therefore the phase envelops ordered structure near the surface. The density near the surface
shifts to the lower monomer fraction with an increase in associat- is more compared to that of the middle regime. With an increase
ing strengths. In particular, the degree of association in the vapour in association strength, the density of each layer increases due to
phase is more influenced compared to the liquid phase due to con- an increase in orientational ordering and, hence, compact packing
finement for higher associating fluids (ε af ∗ > 6), which may be due in the fluid layers. On the other hand, the layering in the vapour
to the layering formation of vapour phase near the surface can be phase decreases due to a decrease in vapour pressure. The number
examined through density profile (discussed in the next section). of layers in the vapour phase is less compared to the liquid phase,
Fig. 3(b) shows the effect of confinement on the monomer frac- and hence bulk regime is visible in the vapour phase, whereas for
tion of an associating fluid (ε af ∗ = 6) at coexistence. Interestingly, the liquid phase, we do not observe bulk regime at H∗ =6. How-
at a given temperature, the monomer fraction of the liquid and ever, the bulk regime for the liquid phase, can be observed at slit
vapour phase increases with the degree of confinement, which in- width H∗ > 10, as shown in Fig. 5(b). To understand the orienta-
dicates the association of molecules in the vapour phase as well tional ordering of fluid molecules confined in slit pore, we examine
as in the liquid phase increases with an increase in slit width at the orientational order parameter, as shown in Fig. 6(a). The aver-

5
S.S. Mandal, S.K. Singh, S. Bhattacharjee et al. Fluid Phase Equilibria 531 (2021) 112909

Fig. 6. Orientational order parameter for different associating fluids at H∗ =6 and at


T∗ =1.1. Solid line represents the liquid orientation profiles and dash-dot line repre-
sents the corresponding vapour orientation profiles.

Fig. 7. Temperature-density vapour-liquid coexistence envelope of an associating


fluid (ε af ∗ =6) in a slit width H∗ =6 for different surface site density.

3.5. Phase transition inside functionalized slit pore

The effect of surface sites on the phase transition of confined


fluid is shown in Fig. 7. At a given temperature, with an increase
in surface sites, the vapour density significantly increases. In con-
Fig. 5. (a). Density profile for different associating fluids at slit width H∗ =6 and at trast, the liquid density remains the same at lower surface site
T∗ =1.1. Solid line represents the liquid density and dash-dot line represents the density and slightly increases at higher surface site density. This
correspondingvapour density. (b). Density profile for a associating fluid (ε af ∗ =6) at
implies that the surface sites play an important role in triggering
different slit width H∗ =6,8 and 10 and, at T∗ =1.1. Solid line represents the liquid
density and dash-dot line represents the corresponding vapour density profiles.
the arrangement of the fluid molecules in coexistence phases. Con-
sequently, the critical density increases and critical temperature
decreases with an increase in surface site density. A similar ob-
age orientation order parameter is evaluated as follows: servation is also found for other model fluids with an increase in
 substrate strength [62]. However, in this study, the interaction be-
3cos2 (θ ) − 1
Sz = (6) tween associating fluid molecules is directional dependent; there-
2 fore, it is quite expected that the sites on the fluid particles would
where, θ is the angle between the z-coordinate normal to the sur- be oriented towards the surface, which is just opposite in the case
face and site vector of the fluid molecule. of a smooth surface. Hence, we examine the z-density profile and
We observe strong orientational ordering near the surface with the orientational profile to understand the molecular ordering in-
an increase in association strength. The centre to site vector of a side the system, which is discussed in the next section.
fluid molecule near the surface tends to lie parallel to the surface,
which is more pronounced at higher association strength (ε af ∗ =8). 3.6. Density and orientation profile
The orientational ordering of fluid molecules slowly diminishes in
the subsequent layers. However, the orientational ordering of fluid The density profile normal to the surface is analysed in Fig. 8
molecules in all layers is preferentially parallel to the surface. at different surface site density. It can be observed from Fig. 8 that

6
S.S. Mandal, S.K. Singh, S. Bhattacharjee et al. Fluid Phase Equilibria 531 (2021) 112909

the effect of surface sites on the ordering of the fluid molecules,


we analyse the orientation order parameter profile of the liquid
phase, as shown in Fig. 9.
Interestingly, we found that the orientation of fluid molecules
near the surface in the presence of surface site is perpendicular
to the surface, which is just opposite without surface sites. There-
fore, the orientation of the subsequent layers also gets affected and
preferentially oriented toward the surface. The orientation profile
of the middle layer (for ρ s =1) also suggests that the layer has
two kinds of molecules with different orientations. Like the liquid
phase, the surface site also influences the vapour phase and leads
to strong layering near the surface, as shown in Fig. 9(b). Interest-
ingly, the density of the layering near the surface is very close to
the liquid density and increases with an increase in surface site
density. The orientation of the fluid molecules in the vapour phase
is very similar to that of the liquid phase, i.e., the fluid molecules
are preferentially oriented toward the surface.

Fig. 8. Density profile of associating strength ε af ∗ =6 at slit width H∗ =6 and at T∗ 3.7. Monomer fraction profile
=1.00 for different surface site density. Solid line represents the liquid density and
dot line represents the corresponding vapour density profiles. To understand the association of fluid molecules in different
layers, we analyse the monomer fraction profile along the axis nor-
mal to the surface, as shown in Fig. 10. It should be noted that
the liquid density near the surface is quite high and is very sharp, the molecules bonded with only surface sites are also treated as
which means that the fluid molecules are strongly ordered near a monomer. It can be observed that the values of monomer frac-
the surface, shift the subsequent layers toward the surface, and tion in the layer near the surface are quite high (~0.7). Also the
result in a partial split in the middle peak at high surface site molecules in that layer are highly (orientational order parameter ~
density. Overall, the structural behaviour of the coexistence phases 0.7, i.e., almost normal to the surface) oriented towards the surface,
completely different in the presence of surface sites. To understand as seen from the orientation profile in the presence of surface sites

Fig. 9. (a). Orientational order parameter of liquid phase of an associating fluids (ε af ∗ =6) at H∗ =6 and at T∗ =1.0 for different surface site density. Corresponding snapshots
are shown in right panel of the figure. (b). Orientational order parameter of vapour phase of an associating fluids (ε af ∗ =6) at H∗ =6 and at T∗ =1.0 for different surface site
density. Corresponding snapshots are shown in right panel of the figure.

7
S.S. Mandal, S.K. Singh, S. Bhattacharjee et al. Fluid Phase Equilibria 531 (2021) 112909

Fig. 10. (a). Monomer fraction profile of liquid phase of an associating fluids
(ε af ∗ =6) at H∗ =6 and at T∗ =1.0 for different surface site density. (b). Monomer frac-
tion profile of vapour phase of an associating fluids (ε af ∗ =6) at H∗ =6 and at T∗ =1.0 Fig. 11. (a). Density profile of coexistence phases of an associating fluids (ε af ∗ =6) at
for different surface site density. H∗ =6 and at T∗ =1.0 for different surface site density. (b). Monomer fraction profile
of coexistence phases of an associating fluids (ε af ∗ =6) at H∗ =6 and at T∗ =1.0 for
different surface site density.
(for ρ s =1). Therefore, the probability of forming bonds between
molecules within the layers is very less. Hence, it suggests that
some fraction of fluid molecules in that layer form bonds with the behaviour of confined fluids, we also study the phase behaviour of
surface site in one direction, and in other direction, it forms bonds the two-site associating fluids in which the sites are located at an
with the molecules in the adjacent layers. Therefore, the bond for- angle of 90 degrees (termed as type B). Fig. 11(a) represents the
mation with the surface sites allows propagating the bond forma- density profile of coexistence phases for two types of associating
tion to the adjacent layers, which result in shifting of density peaks fluids at H∗ =6 and T∗ =1.0. In general, we do not observe much sig-
towards the surface, which is found from the density profile. The nificant difference in the density profile for type B with and with-
maximum density for second layers from the surface is found at out surface sites and found to be less ordered compared to that of
z~1.8 in the presence of surface sites where the orientation order type A. In particular, in the case of type B with surface sites, the
parameter is maximum, and the monomer fraction is found to be density of the layer close to the surface is quite sharp and higher
minimum, which implies that the molecules in the layers preferen- than that without surface sites. However, this change in 1st layer
tially bonded with adjacent layers. A similar observation is found does not affect much in the subsequent layers, and therefore the
in the vapour phase shown in Fig. 10(b). behaviour of density profile for type B is more or less similar with
and without surface sites except in the 1st layer. This behaviour is
3.8. Effect of site location mainly due to the site location in the fluid molecules. In the case
of type B, when fluid molecules form a bond with surface sites,
From the above discussion, it is understood that in the case then another site of the molecules is not available to form bonds
of two-site associating fluids (sites are on opposite sides termed with adjacent layers, and therefore surface sites only affect the 1st
as type A), the fluid molecules near the surface form bonds with layer. On the other hand, in the case of type A, the formation of
surface sites in one direction. In another direction, it forms bonds bond with surface sites also enhance the bond formation with ad-
with adjacent layers and, hence, affects the structural behaviour jacent layers, and hence the effect of surface sites can be visualized
of confined fluids. Therefore, two-site associating fluids, with op- to adjacent layers.
positely located sites, promote the interlayer bond formation and Monomer fraction profile, as shown in Fig. 11(b), also suggests
propagate the effect of the surface to the adjacent layers. To under- less association of type B compared to that of type A molecules.
stand the impact of site location of fluid molecules on the phase For type B, the monomer fraction profile for both the liquid and

8
S.S. Mandal, S.K. Singh, S. Bhattacharjee et al. Fluid Phase Equilibria 531 (2021) 112909

vapour phase is similar with and without surface sites except in ture. This behaviour is further supported by the local density, ori-
the 1st layer. On the other hand, for type A, the monomer fraction entation, and monomer fraction profiles. It can be observed that
profile is found to be more layering in the case with surface sites the molecules near the surface are highly oriented towards the sur-
for both the phases and results in a more ordered structure in the face in the presence of surface sites, which promote the interlayer
coexistence phases. bond formation to the adjacent layers. Thus, the presence of sur-
face site propagates the effect of a surface to the adjacent layers in
4. Conclusions the coexistence phases.

In this work, the vapour-liquid phase equilibria of associating Declaration of Competing Interest
fluids under confinement are investigated through grand-canonical
transition matrix Monte Carlo (GC-TMMC) simulations. Various The authors declare that they have no known competing finan-
properties of coexistence phases, and critical properties, are esti- cial interests or personal relationships that could have appeared to
mated for multiple associating fluids at different slit widths. For a influence the work reported in this paper.
given slit width, with an increase in associating strength, the crit-
ical temperature of the coexistence phases increases. Interestingly,
CRediT authorship contribution statement
an insignificant change in the critical density is observed with the
change in the associating strength of the fluid, which is attributed
Sashanka Sekhar Mandal: Methodology, Formal analysis, Writ-
to the symmetrical shift in the coexistence envelope with the as-
ing - original draft, Visualization, Investigation. Sudhir Kumar
sociating strength. The corresponding state coexistence behaviour
Singh: Writing - review & editing, Visualization, Investigation.
has also investigated for an associating strength, ε af ∗ = 4 & 6 at dif-
Sanchari Bhattacharjee: Writing - review & editing, Visualization.
ferent pore widths varied from H∗ = 4 to 40. Corresponding state
Sandip Khan: Conceptualization, Supervision, Writing - review &
behaviour has shown an interesting trend of coexistence envelopes.
editing, Project administration.
The highest shrinkage in the corresponding state coexistence enve-
lope is observed with the studied slit width, H∗ = 15. Further, with
the decrease in the slit width, the corresponding state coexistence Acknowledgement
envelope broadens and overlaps the bulk corresponding state co-
existence envelope at H∗ = 4 and even broader than that of the We would like to acknowledge the Department of Chemical and
bulk phase at H∗ = 3. Moreover, at larger slit pore widths, such as Biochemical at IIT Patna for providing computational resources for
H∗ = 40, the corresponding state coexistence envelope again starts this research.
broadening and indicating its approach towards bulk, as expected.
The structural properties of coexistence phases are exam- References
ined through monomer fraction, density, and orientation profiles.
[1] X. Liu, D. Zhang, A review of phase behaviour simulation of hydrocarbons in
Monomer fraction of different associating fluid at a particular slit confined space: implications for shale oil and shale gas, J. Nat. Gas Sci. Eng. 68
width, H∗ =6, and monomer fraction at different slit width for a (2019) 102901.
given associating fluid, ε af ∗ =6 are also investigated. At a given as- [2] C. Sun, R. Zhou, Z. Zhao, B. Bai, Nanoconfined fluids: what can we expect from
them? J. Phys. Chem. Lett. 11 (2020) 4678–4692.
sociating strength, with a decrease in temperature (for H∗ =6), the [3] L. Bocquet, E. Charlaix, Nanofluidics, from bulk to interfaces, Chem. Soc. Rev.
values of monomer fraction significantly decrease for liquid phase 39 (2010) 1073–1095.
while the change of monomer fraction in the vapour phase is very [4] R. Evans, Fluids adsorbed in narrow pores: phase equilibria and structure, J.
Phys. 2 (1990) 8989–9007.
small or negligible depending on the associating strength. On the [5] M. Foroutan, S.M. Fatemi, F. Esmaeilian, A review of the structure and dynam-
other hand, at a given temperature, the values of monomer frac- ics of nanoconfined water and ionic liquids via molecular dynamics simulation,
tion for the liquid phase increases and the vapour phase decreases Eur. Phys. J. E 40 (2017) 19.
[6] P. Huber, Soft matter in hard confinement: phase transition thermodynamics,
under confinement compared to the respective bulk systems. In- structure, texture, diffusion and flow in nanoporous media, J. Phys. 27 (2015)
terestingly, for an associating fluid at a given temperature, the 103102.
monomer fraction of the liquid and vapour phase increases with [7] N. Sheibani, M. Kamalvand, Liquid-liquid phase transition in simple
Lennard-Jones nano-confined fluids, Fluid Phase Equilib. 510 (2020) 112495.
the increase in the degree of confinement. The saturation pressure
[8] P.T. Cummings, H. Docherty, C.R. Iacovella, J.K. Singh, Phase transitions in
of associating fluids in confinement increases at a given tempera- nanoconfined fluids: the evidence from simulation and theory, AIChE J. 56
ture as the degree of association decreases. The difference of sat- (2010) 842–848.
uration pressure between bulk and confinement at a given tem- [9] H.K. Christenson, Confinement effects on freezing and melting, J. Phys. 13
(2001) R95–R133.
perature decreases with an increase of associating strength. This [10] K. Mochizuki, K. Koga, Solid-liquid critical behaviour of water in nanopores,
indicates that the effect of confinement on the saturation pressure Proc. Natl. Acad. Sci. U. S. A. 112 (2015) 8221–8226.
is less for higher associating strength fluids. To understand the im- [11] M. Foroutan, S.M. Fatemi, F. Shokouh, Graphene confinement effects on melt-
ing/freezing point and structure and dynamics behaviour of water, J. Mol.
pact of association strength, the local densities, and the orientation Graphics Model. 66 (2016) 85–90.
profile of the coexistence phases, are also examined. The density [12] C. Alba-Simionesco, B. Coasne, G. Dosseh, G. Dudziak, K.E. Gubbins, R. Rad-
near the surface is more compared to that of the middle regime. hakrishnan, M. Sliwinska-Bartkowiak, Effects of confinement on freezing and
melting, J. Phys. 18 (2006) R15–R68.
With an increase in association strength, the densities of each layer [13] K. Binder, J. Horbach, R. Vink, A. De Virgiliis, Confinement effects on phase
increases due to an increase in orientational ordering and hence behaviour of soft matter systems, Soft Matter 4 (2008) 1555–1568.
result in compact packing in the fluid layers. Strong orientational [14] M.T. Miyahara, R. Numaguchi, T. Hiratsuka, K. Nakai, H. Tanaka, Fluids in
nanospaces: molecular simulation studies to find out key mechanisms for en-
ordering is observed near the surface, with an increase in asso-
gineering, Adsorption 20 (2014) 213–223.
ciation strength. The orientational order of fluid molecules slowly [15] L.D. Gelb, K.E. Gubbins, R. Radhakrishnan, M. Sliwinska-Bartkowiak, Phase sep-
diminishes in the subsequent layers. However, the orientational or- aration in confined systems, Rep. Prog. Phys. 62 (1999) 1573–1659.
[16] Y. Liu, A.Z. Panagiotopoulos, P.G. Debenedetti, Finite-size scaling study of the
dering of fluid molecules in all layers are preferentially parallel to
vapour-liquid critical properties of confined fluids: crossover from three di-
the surface. mensions to two dimensions, J. Chem. Phys. 132 (2010) 144107.
Further, the role of surface sites on the phase behaviour un- [17] I. Brovchenko, A. Geiger, A. Oleinikova, Water in nanopores. I. Coexistence
der confinement is critically examined, and found that with an in- curves from Gibbs ensemble Monte Carlo simulations, J. Chem. Phys. 120
(2004) 1958–1972.
crease in surface sites, the vapour density significantly increase. In [18] W.E.I. Shi, X. Zhao, J.K. Johnson, Phase transitions of adsorbed fluids computed
contrast, the liquid density slightly increases at a given tempera- from multiple-histogram reweighting, Mol. Phys. 10 0 (20 02) 2139–2150.

9
S.S. Mandal, S.K. Singh, S. Bhattacharjee et al. Fluid Phase Equilibria 531 (2021) 112909

[19] K.E. Gubbins, J.D. Moore, Molecular modeling of matter: impact and prospects [40] P.A. Bonnaud, B. Coasne, R.J. Pellenq, Molecular simulation of water confined
in engineering, Ind. Eng. Chem. Res. 49 (2010) 3026–3046. in nanoporous silica, J. Phys. 22 (2010) 284110.
[20] J. Bae, J. Lee, Q. Zhou, T. Kim, Micro-/nanofluidics for liquid-mediated pattern- [41] D. Bhandary, K. Srivastava, R. Srivastava, J.K. Singh, Effects of electric field on
ing of hybrid-scale material structures, Adv. Mater. 31 (2019) 1804953. the vapour-liquid equilibria of nanoconfined methanol and ethanol, J. Chem.
[21] L. Bocquet, Nanofluidics coming of age, Nat. Mater. 19 (2020) 254–256. Eng. Data 59 (2014) 3090–3097.
[22] Y. Xu, Nanofluidics: a new arena for materials science, Adv. Mater. 30 (2018) [42] R. Srivastava, J.K. Singh, P.T. Cummings, Effect of electric field on water con-
1702419. fined ingraphite and mica pores, J. Phys. Chem. C 116 (2012) 17594–17603.
[23] K.E. Gubbins, Y.-C. Liu, J.D. Moore, J.C. Palmer, The role of molecular modeling [43] S. Khan, J.K. Singh, Prewetting transitions of one site associating fluids, J. Chem.
in confined systems: impact and prospects, Phys. Chem. Chem. Phys. 13 (2011) Phys. 132 (2010) 144501.
58–85. [44] W.G. Chapman, Prediction of the thermodynamic properties of associating
[24] M. Ramdin, S.H. Jamali, T.M. Becker, T.J.H. Vlugt, Gibbs ensemble Monte Carlo Lenard-Jones fluids: theory and simulation, J. Chem. Phys. 93 (1990) 4299.
simulations of multicomponent natural gas mixtures, Mol. Simul. 44 (2018) [45] J. Alejandre, Y. Duda, S. Sokołowski, Computer modeling of the liquid-vapour
377–383. interface of an associating Lenard-Jones fluid, J. Chem. Phys. 118 (2003)
[25] S.K. Singh, A. Mehta, Corresponding state behaviour of capillary condensation 329–336.
of confined alkanes, Mol. Simul. 45 (2019) 1014–1028. [46] S. Khan, J.K. Singh, Surface phase transition of associating fluids on functional-
[26] R. Srivastava, H. Docherty, J.K. Singh, P.T. Cummings, Phase transitions of ized surfaces, J. Phys. Chem. C 115 (2011) 17861–17869.
water ingraphite and mica pores, J. Phys. Chem. C 115 (2011) 12448– [47] W.A. Steele, The physical interaction of gases with crystalline solids: I. Gas–
12457. solid energies and properties of isolated adsorbed atoms, Surf. Sci. 36 (1973)
[27] K. Yamashita, K. Kashiwagi, A. Agrawal, H. Daiguji, Grand canonical monte 317–352.
carlo and molecular dynamics simulations of capillary condensation and evap- [48] W.B. Streett, L.A.K. Staveley, Calculation on a corresponding states basis of the
oration of water inhydrophilic mesopores, Mol. Phys. 115 (2017) 328–342. volume change on mixing simple liquids, J. Chem. Phys. 47 (1967) 2449–2454.
[28] S.K. Singh, S. Khan, S. Jana, J.K. Singh, Vapour-liquid phase equilibria of simple [49] , in: J.S. Rowlinson, F.L. Swinton (Eds.), Liquids and Liquid Mixtures, third ed.,
fluids confined in patterned slit pores, Mol. Simul. 37 (2011) 350–360. Butterworth-Heinemann, 1982, p. iii.
[29] S.K. Singh, A.K. Saha, J.K. Singh, Molecular simulation study of vapour-liquid [50] J.R. Errington, Evaluating surface tension using grand-canonical transition-ma-
critical properties of a simple fluid in attractive slit pores: crossover from 3D trix Monte Carlo simulation and finite-size scaling, Phys. Rev. E 67 (2003)
to 2D, J. Phys. Chem. B 114 (2010) 4283–4292. 012102.
[30] P. Bonnaud, Q. Ji, B. Coasne, R.-M. Pellenq, K. Van Vliet, Thermodynam- [51] A.M. Ferrenberg, R.H. Swendsen, New Monte Carlo technique for studying
ics of waterconfined in porous calcium-silicate-hydrates, Langmuir 28 (2012) phase transitions, Phys. Rev. Lett. 61 (1988) 2635.
11422–11432. [52] J.K. Singh, D.A. Kofke, Molecular simulation study of effect of molecular asso-
[31] S.K. Singh, A. Sinha, G. Deo, J.K. Singh, Vapour-liquid phase coexistence, critical ciation on vapour-liquid interfacial properties, J. Chem. Phys. 121 (2004) 9574.
properties, and surface tension of confined alkanes, J. Phys. Chem. C 113 (2009) [53] B.A. Berg, T. Neuhaus, Multicanonical ensemble: a new approach to simulate
7170–7180. first-order phase transitions, Phys. Rev. Lett. 68 (1992) 9.
[32] K. Mochizuki, K. Koga, Solid-liquid critical behaviour of water in nanopores, [54] J.K. Singh, J. Adhikari, S.K. Kwak, Vapour-liquid phase coexistence curves for
Proc. Natl. Acad. Sci. 112 (2015) 8221. Morse fluids, Fluid Phase Equilb. 248 (2006) 1–6.
[33] A. Sengupta, P. Behera, J. Adhikari, Molecular simulation study of triangle-well [55] S. Wierzchowski, D.A. Kofke, A general-purpose biasing scheme for Monte
fluids confined in slit pores, Mol. Phys. 112 (2014) 1969–1978. Carlo simulation of associating fluids, J. Chem. Phys. 114 (2001) 8752–8762.
[34] A. Sengupta, J. Adhikari, Fluid phase equilibria of triangle-well fluids confined [56] J.R. Errington, Prewetting transitions for a model argon on solid carbon dioxide
inside slit pores: a transition matrix Monte Carlo simulation study, J. Mol. Liq. system, Langmuir 20 (2004) 3798.
221 (2016) 1184–1196. [57] J.K. Singh, D.A. Kofke, Molecular simulation study of the vapour-liquid inter-
[35] S.K. Singh, J.K. Singh, Effect of pore morphology on vapour-liquid phase transi- facial behaviour of a dimer-forming associating fluid, Mol. Simul. 30 (2004)
tion and crossover behaviour of critical properties from 3D to 2D, Fluid Phase 343–351.
Equilib. 300 (2011) 182–187. [58] J.K. Singh, D.A. Kofke, Molecular simulation study of effect of molecular as-
[36] Y. Liu, A.Z. Panagiotopoulos, P.G. Debenedetti, Finite-size scaling study of the sociation on vapour-liquid interfacial properties, J. Chem. Phys. 121 (2004)
vapour-liquid critical properties of confined fluids: crossover from three di- 9574–9580.
mensions to two dimensions., J. Chem. Phys. 132 (2010) 144107–144116. [59] J.R. Errington, A.Z. Panagiotopoulos, A new intermolecular potential model for
[37] I. Brovchenko, A. Geiger, A. Oleinikova, Phase equilibria of water in cylindrical the n-alkane homologous series, J. Phys. Chem. B 103 (1999) 6314–6322.
nanopores, Phys. Chem. Chem. Phys. 3 (2001) 1567–1569. [60] J.R. Errington, Direct calculation of liquid-vapour phase equilibria from transi-
[38] I. Brovchenko, A. Oleinikova, Water in Nanopores: III. Surface phase transitions tion matrix Monte Carlo simulation, J. Chem. Phys. 118 (2003) 9915–9925.
of water on hydrophilic surfaces, J. Phys. Chem. C 111 (2007) 15716–15725. [61] S. Jana, J.K. Singh, S.K. Kwak, Vapour-liquid critical and interfacial properties of
[39] I. Brovchenko, A. Geiger, A. Oleinikova, Water in nanopores: II. The liq- square-well fluids in slit pores, J. Chem. Phys. 130 (2009) 214707.
uid-vapour phase transition near hydrophobic surfaces, J. Phys. 16 (2004) [62] J.K. Singh, S.K. Kwak, Surface tension and vapour-liquid phase coexistence of
S5345–S5370. confined square-well fluid, J. Chem. Phys. 126 (2007) 024702.

10

You might also like