International Journal of Heat and Mass Transfer: Hari Krishna Chilukoti, Gota Kikugawa, Taku Ohara

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

International Journal of Heat and Mass Transfer 79 (2014) 846–857

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

Structure and transport properties of liquid alkanes in the vicinity


of a-quartz surfaces
Hari Krishna Chilukoti a,⇑, Gota Kikugawa b, Taku Ohara b
a
Department of Nanomechanics, Graduate School of Engineering, Tohoku University, Sendai, Japan
b
Institute of Fluid Science, Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, Japan

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

Article history: Structure and mass transport properties of several liquid n-alkanes, methane, decane and tetracosane, in
Received 8 November 2013 the vicinity of a-quartz surfaces of three crystal planes have been investigated by using molecular
Received in revised form 27 May 2014 dynamics simulations. Solid a-quartz surfaces were terminated with –H and –OH groups to create hydro-
Accepted 31 August 2014
phobic and hydrophilic surfaces, respectively. It was observed that adsorption of the liquid alkane mol-
Available online 19 September 2014
ecules is more noticeable near the (0 0 1) crystal plane when compared with other two planes (0 1 1) and
(1 0 0). For a given alkane, the number of molecules adsorbed near a a-quartz wall marginally depends
Keywords:
upon the crystal plane and type of surface termination. Ordering parameter and radius of gyration for
Molecular dynamics
Solid–liquid interface
liquid molecules were examined in the interface region to gain the knowledge about molecular alignment
n-Alkane and chain configuration in the interface region. Liquid molecules in the first adsorption layer near the
a-Quartz hydrophilic and hydrophobic surfaces are more parallel to the interface in the following order (0 0 1),
Crystal plane (0 1 1) and (1 0 0). Portions of decane and tetracosane liquid molecules tend to enter into the H-termi-
Self-diffusion coefficient nated side with an orientation in the parallel direction to the interface and into the OH-terminated side
with an orientation in the perpendicular direction for (1 0 0) crystal plane. Molecules are flattened in the
direction perpendicular to the interface and are elongated in the other direction in the interface region for
decane and tetracosane. It is observed that the in-plane self-diffusion coefficient of liquids, the self-dif-
fusion coefficient according to the migration of molecules in the directions parallel to the interface, is
smaller in the vicinity of the solid surface and approaches to the bulk liquid value away from the inter-
face. For methane, decane and tetracosane, the in-plane diffusion in the first adsorption layer near the
smoother silica surface (0 0 1), which has molecularly smoother surface than the other two surfaces
(0 1 1) and (1 0 0), for both of the terminations is higher as compared with the other two surfaces
(0 1 1) and (1 0 0). It was found that the diffusion coefficients in the two directions of the 2-D in-plane
diffusion are different significantly and this anisotropic nature is resulted by the topology of the solid sur-
face. At the same reduced temperature, a tendency was observed that the in-plane diffusion near the
H-terminated and OH-terminated surfaces for three crystal planes increases with an increase in the chain
length of the liquid alkanes.
Ó 2014 Elsevier Ltd. All rights reserved.

1. Introduction structural adhesives used in the aerospace and automotive appli-


cation, paints and coatings and microelectronics [1]. It is observed
The interface between liquid and solid is important in many that liquids behave differently in the vicinity of solid surfaces
industrial applications like catalysis, adhesion, wetting, corrosion from that in the bulk liquid. The chemical and physical nature
and biocompatibility. Structure and dynamic properties at the of the surface, hydrogen bond network on the surface, atomic-
interfaces need to be carefully understood at the molecular level scale roughness, partial charges on the surface, its distribution
to design many industrial processes and the proper function of and concentration might influence the transport properties and
many devices where interface has significant importance like structure of liquid in the vicinity of the solid surface [2,3]. From
a scientific point of view, it would be interesting to know how
⇑ Corresponding author at: Molecular Heat Transfer Laboratory, Institute of Fluid
the interactions between two contacting phases influence the
Science, Tohoku University, 2-1-1 Katahira, Aoba-ku, Sendai 980-8577, Japan. Tel./
interfacial physical properties in the macroscopic scale. It is also
fax: +81 22 217 5872. important to know how far the influence caused by the solid wall
E-mail address: chilukoti@microheat.ifs.tohoku.ac.jp (H.K. Chilukoti). penetrates into the bulk liquid.

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2014.08.089
0017-9310/Ó 2014 Elsevier Ltd. All rights reserved.
H.K. Chilukoti et al. / International Journal of Heat and Mass Transfer 79 (2014) 846–857 847

Nomenclature

D self-diffusion coefficient (m2/s) Greek symbols


L length of the simulation system (Å) h angle between the line joining two united atoms and
N number of molecules interface normal
P orientation order parameter s time interval (s)
r distance between two united atoms (Å)
rc cut-off radius (Å) Subscripts
Rg radius of gyration (Å2) PSk particular slab
S survival probability x, y, z x, y and z directions
Tcr critical temperature (K)

Due to the hidden nature and comparatively small size (few layered structures [12–14]. It would be of practical interest to
molecular lengths), investigations of molecules in the interface know the orientation and shape of the molecules in the layers
regions by experiments is not straightforward [4]. For instance, it and their influence on the diffusion coefficient in these layers,
is difficult to study the liquid confined between solid walls which especially in cases of long-chain polymer liquid molecules. Exper-
are closer than 1 nm from experiments because of large surface imental measurement of diffusion coefficient is sometime difficult
forces [5]. It is also observed that some experimental techniques under a circumstance like low concentration limit and far from
like the surface force apparatus (SFA) measurements cannot be ambient. Under these conditions computational simulations have
used for investigation of an isolated liquid–solid surface [5]. Sum an advantage. MD simulations are widely used to evaluate the dif-
frequency generation (SFG) vibrational spectroscopy is a widely fusion of liquid in the bulk liquid region [15–17] and liquid in the
used analytical tool to examine the structure at liquid–solid inter- vicinity of solid substrates [18–20]. Recently, diffusion in the inter-
faces. Nevertheless, quantitative analysis of the obtained spectra is face region was studied by a few researchers using modified
not easy. So, computer simulations have been also performed to Einstein relation [21–23]. The authors have examined the diffusion
obtain the equilibrium and non-equilibrium quantities in the inter- in several alkane liquids at the liquid–vapor interfaces [24].
face region and to understand the experimental findings. Solid/ The transport properties of liquid alkanes in the vicinity of silica
liquid interfaces are extensively studied by the computational substrates are not well understood issues. Among the possible
methods like molecular dynamics simulations. Using this method, crystallographic surfaces of the a-quartz, only a few are more sta-
transport, physical and structural quantities are calculated and ble. For each crystallographic plane, surface characteristics are dif-
analyzed. ferent and the interaction with liquids in the vicinity is also
Silica is ubiquitously found in the earth crust in the form of different. Detailed understanding of influences of crystal plane on
a-quartz. Due to its importance in geology, nanotribology, micro- transport properties of liquids at the interfaces helps to control
fluidics, micro/nanofabrication and to many processes in the nat- the mobility of liquids and diffusion process in the vicinity of solid
ure, scientists have extensively studied about liquid interaction surfaces. To the best of our knowledge, self-diffusion of alkanes in
with silica surfaces. Liquid alkanes have many industrial applica- the vicinity of silica substrates has not been investigated by any
tions like lubrication and coating. It is also important because it researcher.
is a basic block of various organic chemicals. Interfaces involving In the present study, structure and diffusion properties of liquid
alkane and silica substrate have been attracting scientific and methane (CH4), decane (C10H22) and tetracosane (C24H50) in the
engineering interest. For example, studies of structural and trans- vicinity of a-quartz substrates were investigated by molecular
port phenomena of alkanes in the vicinity of silica surfaces have dynamics simulations. To examine the influence of a-quartz sur-
applications in the oil production industry and MEMS. face topology, obtained properties in the interface region were
The a-quartz/water interface have been extensively studied by compared among (0 0 1), (0 1 1) and (1 0 0) crystal planes. The sys-
many researchers [2,3,6–8] using MD simulations to understand tems utilized in the present MD simulations have two a-quartz sil-
the hydrophobic and hydrophilic effect and to examine the struc- ica walls contacting the alkane liquid between the walls. One of the
ture of water in the interface. It is well understood that inorganic a-quartz substrate surfaces was terminated with –H (silane)
oxides like a-quartz has affinity to hydroxyl groups on its surface groups and the other wall surface with –OH (silanol) groups to
[4]. It is believed that fully hydroxylated quartz surfaces are ener- generate hydrophobic and hydrophilic surfaces, respectively. The
getically favored and stable when compared with bare and par- properties of liquid alkanes near –H and –OH terminated surfaces
tially hydroxylated surfaces [9]. Wright et al. [10] suggested that were compared to examine the effect of surface termination type.
a charge free hydroxylated surface would be a reasonable approx- By comparing the properties of liquid alkanes at the same reduced
imation to a-quartz at the surface at pH7. Therefore, in the present temperature, influences of alkane chain length have been studied.
work one wall is terminated with OH-group (hydroxylated) and
the other wall terminated with H-group (hydrogenated). Both sur- 2. Potential functions and simulation methodology
faces are modeled as electrically neural surfaces.
Mass transport phenomena and the measurement of diffusion MD simulations have been performed in the present study to
coefficient are important as a fundamental research and also in investigate the structure and self-diffusion coefficient of different
applications in industry such as coating, surface processing and liquid alkanes in the vicinity of (0 0 1), (1 0 0) and (0 1 1) crystal
modification by using chemical solutions, and adsorption filtration planes of a-quartz substrate. Methane (CH4), decane (C10H22) and
of chemical species. Study of diffusion in liquid alkanes under con- tetracosane (C24H50) were selected for the n-alkanes. An atomistic
finement at nanoscale has application in molecular sieving in model of silica was applied and alkane liquid was modeled using
nanofilters and pore filling in solid-state dye-sensitized solar cells the united atom model. In the united atom representation of
[11]. It is well known that liquid in the vicinity of solid wall has alkane, CHn groups are treated as single interaction site and the
848 H.K. Chilukoti et al. / International Journal of Heat and Mass Transfer 79 (2014) 846–857

interaction site is placed at the carbon atoms. In the literature, idealized crystal structure and real structure of silica surface is
many united atom force fields have been developed to reproduce not completely known [30].
physical and dynamic properties of alkanes using MD simulations. Initially, density of the liquid alkane at a required temperature
Among those, NERD potential parameters were used for decane was calculated from liquid–vapor interface simulation. After that,
and tetracosane in the present study, and TraPPE parameters for alkane molecules were randomly placed between the two solid
methane. Stretching of C–C bonds, bond bending and torsional walls to see to it that after the system reaches equilibrium, liquid
motion are considered as intra molecular degrees of freedom. density in the bulk liquid region that is sufficiently distant from
Intermolecular interaction between atoms on different molecules the solid walls and free from the influence of intermolecular poten-
and intramolecular interaction between atoms which are more tial of the solid wall atoms was close to that of the bulk liquid
than 3 atoms apart within an alkane molecule are considered using alkane obtained at the required temperature. After that, the system
LJ potential. Detailed force field parameters for alkanes were found temperature was slowly raised to the required temperature, where
in [25,26]. The silica wall was modeled based on the all atom the temperature was controlled using the velocity scaling method.
model. Lopes et al. [27] proposed a force field for silica by optimiz- After the system reached the required temperature, the simulation
ing the CHARMM force field from ab initio calculations. Among the system was equilibrated for 5 ns and then a production run was
available force fields in the literature, this force field is in compat- performed for 3 ns in the NVT ensemble using Nosé–Hoover ther-
ible with NERD force filed [25] and TraPPE force field [26]. So, this mostat for analysis.
was used to model the a-quartz substrate. Interaction parameters
between the atoms in the silica wall were available in Lopes et al.
[27]. In the present simulation, n-alkane molecules and the 3. Results and discussion
a-quartz substrate were interacted through LJ potential. For the
interaction parameter between different species of interaction 3.1. Density profiles
sites, the Lorentz–Berthelot rule was used. Coulombic interactions
were calculated using SPME method [28] with a charge grid of Structure and mass transport properties were obtained after the
about 1 Å. Cut-off radius (rc) for the LJ interaction was 16 Å. system reached equilibrium. First, mass density distributions of
Atomic equations of motion were integrated with the reversible liquid alkane between the two solid walls were examined before
reference system propagator algorithm method [29] with multiple studying other quantities at the interfaces. These density profiles
time scales. 1 fs and 0.2 fs integration time steps were used for the will provide a basic insight into the structure in the interface
intermolecular and intramolecular motions. Fig. 1 shows the MD region. In the present study, the results for methane, decane and
simulation model of the a-quartz solid and alkane liquid system. tetracosane were compared at the same reduced temperature
The simulation system has two solid silica walls with surfaces par- (0.56Tcr). Predictions of the critical temperatures for decane and
allel to the xy plane and an alkane liquid is in between them. In tetracosane modeled by NERD force field and methane by TraPPE
accordance with the length of the alkane molecule, several sizes force field are taken from the literature [26,31] as 191 K, 617 K
of the systems were created in such a way that the lengths of the and 805 K, respectively, which are used to determine the reduced
solid substrate in the x, y and z directions were larger than double temperatures. Bulk liquid densities of the alkanes by the present
of the alkane molecular length. The solid substrate length in the z- simulations have been validated with literature [24]. The density
direction was also chosen in such a way that the interaction distributions were obtained by dividing the rectangular simulation
between OH-terminated surface and H-terminated surface was box into slabs with a thickness of 0.1 Å in the z direction, and mea-
negligible, so that each of the two solid surfaces is not affected suring the density in each slab. Density profiles for liquid methane,
by the other surface beyond the periodic boundary. The length of decane and tetracosane at 106.96 K, 345.6 K and 450 K near the
the simulation system in the z-direction was 150 Å in all the cases hydroxylated and hydrogenated surfaces are shown in Figs. 2–4.
considered in the present study. Periodic boundary conditions The origin for the distance in the density profiles is located at the
were applied in all three directions. The a-quartz substrate was equilibrium position of the first layer of silicon atoms on which
created in such a way that the crystal plane required to contact the surface terminations are attached. Since for the (0 1 1) crystal
the liquid is normal to the z-axis. The solid wall was constructed plane, the surface atoms are not in a single plane, the origin repre-
by considering the crystal parameters of silica as specified by Lopes sents average position of all the surface silicon atoms. As expected,
et al. [27]. As suggested by Lopes et al. [27], all the dangling bonds density profiles for all alkane liquids exhibit oscillatory nature in
presented on one of the two surfaces were manually saturated by the vicinity of the solid walls. The oscillations in the density profile
terminating with –H atoms to generate a hydrophobic surface and indicate layering of liquid molecules in the interface region. The
the other surface was terminated with –OH atoms to generate a adsorption layer is defined here as the first layer of liquid mole-
hydrophilic surface, although this kind of silica substrate is an cules near the solid wall which can be identified as the first peak

Fig. 1. The a-quartz (0 0 1) /methane simulation system.


H.K. Chilukoti et al. / International Journal of Heat and Mass Transfer 79 (2014) 846–857 849

(a) 2500 (a) 2500


Methane at 106.96 K Decane at 345.6 K
2000 2000
H-terminated side H-terminated side
Density (kg/m3)

Density (kg/m3)
1500 (001)
1500
(001) (011)
(011) (100)
1000 (100) 1000

500 500

0 0
-40 -30 -20 -10 0 -40 -30 -20 -10 0
Distance from surface Si atoms (Å) Distance from surface Si atoms (Å)

(b) 2500 (b) 2500


Methane at 106.96 K Decane at 345.6 K
2000 2000 OH-terminated side
OH-terminated side
Density (kg/m3)

Density (kg/m3)
1500 1500 (001)
(001) (011)
(011) (100)
1000 (100) 1000

500 500

0 0
0 10 20 30 40 0 10 20 30 40
Distance from surface Si atoms (Å) Distance from surface Si atoms (Å)

Fig. 2. Density distributions for methane at 106.96 K on (a) the H-terminated side Fig. 3. Density distributions for decane at 345.6 K on (a) the H-terminated side and
and (b) the OH-terminated side. (b) the OH-terminated side.

in the density profile. For all density profiles, oscillations disappear instead of the single peak only. When large number of molecules
after some distance away from the wall and show a flat region accumulated, split twin peak is transformed into two simple peaks
which indicates the bulk-like liquid region which is free from the (two layers of molecules) which can be observed in Fig. 2(a) for
influence of solid surfaces. It is observed from the figure that the (1 0 0) crystal plane. In this case, it is observed that the peak density
layered structure near the (1 0 0) crystal plane is relatively close in the second layer is larger than that in the first layer. For longer
to the surface layer of silicon atoms than those for the other two chain molecules, decane and tetracosane, the alignment of liquid
crystal planes. It suggests that some liquid molecules might enter molecules in the adsorption layer is not sensitive to the surface
into the cavities on the solid surface, which will be examined later roughness so that the twin peaks of the adsorption layer disap-
again. Snapshots of the methane/a-quartz interface for all cases of pears. Fig. 2 illustrates that the number of oscillations formed near
crystal planes for H- and OH-terminations are shown in Fig. 5. It is the hydrophilic and hydrophobic surfaces and their propagation
observed in the figure that surface topology is different among the into the bulk liquid region is strongly depends on the crystal plane
tested crystal planes, and that shape and orientation of the surface and the surface termination type for methane. On the other hand,
termination group, –H or –OH, are different for each crystal plane. for decane and tetracosane, layering caused by the wall disappears
The (0 0 1), (1 0 0) and (0 1 1) crystal planes have surface terminal approximately at the same distance away from the wall irrespec-
densities of 10.02OH (or) H/nm2, 7.68OH (or) H/nm2, 5.92OH (or) tive of the surface termination type and crystal plane. The layered
H/nm2, respectively. It can be seen in Fig. 5(a) and (c) that methane structure of liquid molecules is reduced with an increase in chain
molecules in the adsorption layers on the (0 0 1) H- and OH-termi- length for all crystal planes at the same reduced temperature. Den-
nated surfaces and (1 0 0) H-terminated surface are aligned in a sity of all three liquids in the adsorption layer near hydrophilic and
plane. Therefore, it exhibits a simple peak in the density profile hydrophobic surfaces for (0 0 1) crystal plane is higher than that of
in Fig. 2(a) and (b). In cases of (0 1 1) H- and OH-terminated sur- (0 1 1) and (1 0 0) planes. This can be attributed to the strong inter-
faces and (1 0 0) OH-terminated surface, which have larger lat- action between (0 0 1) crystal plane and the liquid molecules,
tice-scale roughness than (0 0 1), methane molecules in the first which has a higher surface density of terminal groups than other
adsorption layer are not aligned in a single plane in a simple fash- two planes. The maximum density in the adsorption layer is higher
ion. They tend to be located at two typical positions of z. This on the H-terminated side for methane for all crystal planes,
results in the split twin peaks in the density profiles in Fig. 2. In whereas the maximum density for decane and tetracosane in the
such cases, the first adsorption layer is defined as the twin peaks adsorption layer depends on the type of crystal plane.
850 H.K. Chilukoti et al. / International Journal of Heat and Mass Transfer 79 (2014) 846–857

(a) 2500 (a)


Tetracosane at 450 K
2000 H-terminated side
Density (kg/m3)

(001)
1500
(011)
(100)

1000

500

0 (001) OH-side (001) H-side


-40 -30 -20 -10 0
Distance from surface Si atoms (Å) (b)

(b) 2500
Tetracosane at 450 K
2000 OH-terminaetd side
Density (kg/m3)

1500 (001)
(011)
(100)
1000

500 (011) OH-side (011) H-side

(c)
0
0 10 20 30 40
Distance from surface Si atoms (Å)

Fig. 4. Density distributions for tetracosane at 450 K on (a) the H-terminated side
and (b) the OH-terminated side.

Table 1 shows the penetration depths of the liquid methane,


decane and tetracosane into the solid wall for three crystal
planes. The penetration depth is defined here as the overlapping
length of the number density profiles of the liquid atoms and
the solid atoms as shown in Fig. 6. Zoomed region of the number
density profiles of the methane and the silica near the OH-termi-
nated side is shown in the Fig. 6. It is to note that, the chances of (100) OH-side (100) H-side
liquid molecules entering into the cavities on the solid surface are
very rare, which can be seen in the Fig. 5(a) and (b). However, the Fig. 5. Snapshots of OH-terminated (left) and H-terminated (right) a-quartz/
time-averaged number density profiles of liquids atoms and solid methane interfaces for crystal plane of (a) (0 0 1), (b) (0 1 1), and (c) (1 0 0).
atoms are overlapping as shown in Fig. 6. The penetration depth
suggests the tendency of the liquid molecules to enter into the molecules. This tendency might have been influenced by large
cavities on the solid wall. From Table 1, it is observed that at lattice vibrations of termination groups at high temperature. It
the same reduced temperature, liquid tetracosane molecules should be noted that the liquids were simulated at the same
show the tendency to enter deeply into the solid wall near both reduced temperature, which resulted in higher absolute tempera-
types of surface terminations than methane and decane liquid tures for longer alkane molecules. For all liquids, the tendency of

Table 1
Comparison of penetration depth of liquid molecules into the solid wall among methane, decane and tetracosane for three crystal planes.

(0 0 1) Crystal plane (0 1 1) Crystal plane (1 0 0) Crystal plane


H-terminated sideÅ OH-terminated side Å H-terminated side Å OH-terminated side Å H-terminated side Å OH-terminated side Å
Methane (CH4) – – 0.1 0.7 0.4 0.7
Decane (C10H22) – 0.1 0.5 1.0 0.9 2.1
Tetracosane (C24H50) 0.2 0.6 0.7 1.3 1.2 2.2
H.K. Chilukoti et al. / International Journal of Heat and Mass Transfer 79 (2014) 846–857 851

×1028 ×1028
15.0 (a) 0.5 10.0
Number density Decane at 345.6 K

Orientation order parameter


12.5 of silica atoms
H-terminated side 8.0
Methane at 106.96 K

Number density (m-3)


0.25
Number density (m-3)

(001)
10.0 (011) crystal plane
(011) 6.0
Number density (100)
7.5 of methane atoms
0 4.0
Penetration
5.0 depth
2.0
2.5 -0.25 Number density
of H-termination
0.0
0.0
-0.5 -2.0
-2.5 -40 -30 -20 -10 0
2 2.5 3 3.5 4 Distance from surface Si atoms (Å)
Distance from surface Si atoms (Å)
28
Fig. 6. Number density profiles of silica atoms and methane atoms for (0 1 1) crystal ×10
plane near OH-termination side. (b) 0.5 10.0
Decane at 345.6 K

Orientation order parameter


molecules penetration is deeper into the OH-terminated surface OH-terminated side 8.0

Number density (m-3)


than the H-terminated surface. 0.25
(001)
Table 2 shows the widths of the first adsorption layer in cases 6.0
(011)
of methane, decane and tetracosane in the vicinity of –H and –OH (100)
terminated three a-quartz crystal planes where the width of the
0 4.0
first adsorption layer is defined as the distance from the valley
between 1st and 2nd adsorption layer to the point where the
traces of liquid molecules are present close to the solid wall for 2.0
both type of terminations. The value in the bracket in the table -0.25 Number density
indicates average number of molecules contained in the first of OH-termination
0.0
adsorption layer per cross sectional area of the interface. Average
number of molecules was calculated by dividing the total number
-0.5 -2.0
of united atoms in the first adsorption layer with the number of 0 10 20 30 40
united atoms in a molecule. It is observed that the adsorption Distance from surface Si atoms (Å)
width near OH-terminated surface is larger than near H-termi-
nated surface for (0 1 1) and (1 0 0) crystal planes. On the other Fig. 7. Orientation order parameter distributions for decane at 345.6 K on (a) the H-
terminated side, and (b) the OH-terminated side. Density distribution is also plotted
hand, for (0 0 1) crystal plane the adsorption width on H-termi-
in the figure by the dashed lines and dash dotted lines.
nated surface is larger than the OH-terminated surface. It is
observed that the average number of molecules present in the
first adsorption layer per cross-sectional area of the interface on plane on the ordering of molecules in the layers of liquid molecules
the –H terminated side is slightly larger than the –OH terminated close to the solid wall. This represents the average alignment of
side for all three crystal planes. For a particular alkane, the num- molecules in the liquid and interface region. The orientation order
ber of molecules in the first adsorbed layer per cross sectional parameter is defined as follows [32]
area of the interface on both sides of terminations slightly
depends upon the crystal plane. 1 
PðzÞ ¼ 3 cos2 h  1 ; ð1Þ
2

3.2. Orientation order parameter and radius of gyration where h is defined as the angle between connecting line between
two united atoms which are separated by two covalent bonds in
The orientation order parameter was examined in order to an alkane molecule and a normal vector to the interface plane (z-
know how the molecules are aligned and the influence of crystal direction). The angular bracket represents that averages are taken

Table 2
Comparison of the first adsorption layer width for methane, decane and tetracosane for three crystal planes. Numbers in the bracket indicate the average number of molecules in
the first adsorption layer per cross sectional area of the interface.

(0 0 1) Crystal plane (0 1 1) Crystal plane (1 0 0) Crystal plane


H-terminated side Å OH-terminated side Å H-terminated side Å OH-terminated side Å H-terminated side Å OH-terminated side Å
(Å2) (Å2) (Å2) (Å2) (Å2) (Å2)
Methane (CH4) 3.00 (0.06131) 2.60 (0.06029) 2.30 (0.05924) 2.70 (0.05923) 4.00 (0.07684) 4.15 (0.06720)
Decane (C10H22) 4.40 (0.01434) 4.25 (0.01426) 4.00 (0.01455) 4.20 (0.01393) 5.00 (0.01457) 5.90 (0.01375)
Tetracosane 4.75 (0.00631) 4.70 (0.00624) 4.30 (0.00652) 4.40 (0.00595) 5.40 (0.00639) 6.20 (0.00619)
(C24H50)
852 H.K. Chilukoti et al. / International Journal of Heat and Mass Transfer 79 (2014) 846–857

28 −19
×10 ×10
(a) 0.5 10.0 (a) 7.0 2500
Tetracosane at 450 K Tetracosane at 450 K
Orientation order parameter

H-terminated side 8.0 H-terminated side


6.0 2000

Number density (m-3)


0.25
(001) (001)

Density (kg/m3)
(011) 6.0

Total R2g (m2)


(011)
(100) 5.0 (100) 1500
0 4.0

4.0
Radius of gyration 1000
2.0
-0.25 Number density
of H-termination Density
0.0 3.0 500

-0.5 -2.0 2.0 0


-40 -30 -20 -10 0 -40 -30 -20 -10 0
Distance from surface Si atoms (Å) Distance from surface Si atoms (Å)

×10
28 ×10−19

(b) 0.5 10.0 (b) 7.0 2500


Tetracosane at 450 K Tetracosane at 450 K
Orientation order parameter

OH-terminated side 8.0 OH-terminated side


6.0 2000
Number density (m-3)

0.25

Density (kg/m3)
(001) (001)
6.0
Total Rg2 (m2)
(011) (011)
5.0 1500
(100) (100)
0 4.0
4.0 Radius of gyration 1000
2.0
-0.25 Number density Density
of OH-termination 3.0 500
0.0

-0.5 -2.0 2.0 0


0 10 20 30 40 0 10 20 30 40
Distance from surface Si atoms (Å) Distance from surface Si atoms (Å)

Fig. 8. Orientation order parameter distributions for tetracosane at 345.6 K on (a) Fig. 9. Total radius of gyration as a function of distance for tetracosane at T = 450 K
the H-terminated side and (b) the OH-terminated side. Density distribution is also on (a) the H-terminated side and (b) the OH-terminated side. Density distribution is
plotted in the figure by the dashed lines and dash dotted lines. also plotted in the figure.

over all the vectors present in a particular slab over the entire
production run. It has maximum value of 1 and minimum of and align preferentially parallel to the interface in the next layer.
0.5. Zero value of the ordering parameter indicates that alkyl chain This is an indication of interlocking of molecules in the monolayers
segments are oriented randomly. A positive value specifies that which is called as interlayer interdigitation [1]. The amplitude of
molecular orientation is favored parallel to the z-direction, i.e., per- oscillation of the orientation order parameter is the largest for
pendicular to the interface. A negative value indicates that favored the (0 0 1) surface and smallest for the (0 1 1) surface, although it
orientation is perpendicular to the z-axis and parallel to the inter- is a slight difference. The difference between H-side and OH-side
face. Variation of the ordering parameter along the z-axis was is also very small.
obtained by dividing the rectangular simulation box into slabs with A remarkable feature is observed in the inner side of the first
a thickness of 1 Å. A vector joining two united atoms was assigned adsorption layer for the OH-terminated side of (1 0 0) surface. The
to a slab where the midpoint was situated. orientation order parameter exhibits large positive value. This
Variation of the orientation order parameter along the z-direc- occurs in the region close to the solid surface where the number
tion for decane and tetracosane at the same reduced temperature density profiles of the liquid molecules and the OH-termination
were shown in Figs. 7 and 8. Density profiles of alkane liquids mol- groups overlap. This implies that a few portions of liquid molecules
ecules were shown in the figures as dashed lines. Density profiles tend to get into the solid surface structure with a perpendicular
of –H and –OH surface termination groups for (1 0 0) crystal plane orientation only in the case of (1 0 0) surface. On the other hand,
are also shown in the figure by the dash dotted line. As expected, the molecules near the H-terminated side might enter into the
molecules are randomly oriented in the bulk liquid region. Accord- solid wall in the parallel orientation. At the same reduced temper-
ing to the oscillations in the density profile, the orientation order ature, decane and tetracosane liquids have approximately the same
parameter exhibits oscillations between positive and negative val- distributions of the orientation order parameter near the hydro-
ues. The value is negative in the layers of molecules located at the phobic and hydrophilic surfaces for a particular crystal plane. Sim-
peaks of the density, while it is weakly positive between the layers. ilar tendency of orientation ordering is observed for other kinds of
This indicates that portions of molecules in a layer are bent in the liquids near a-quartz substrates by experiments [5] and MD simu-
perpendicular directions to the interface between the two layers lations [1,4].
H.K. Chilukoti et al. / International Journal of Heat and Mass Transfer 79 (2014) 846–857 853

×10
−19 3.3. Self-diffusion coefficient in the vicinity of interface
6.0 3000
(a) Tetracosane at 450 K Diffusion coefficient is commonly obtained by MD simulations
using the Einstein relation and Green–Kubo relation. Recently,
H-terminated side 2500
Liu et al. [21] modified the Einstein relation to determine the
Component of R2g (m2)

4.0 2 self-diffusion coefficient in the heterogeneous systems which is


Rg in x-direction

Density (kg/m )
2000

3
2 applicable to study the diffusion in the interface regions. The var-
Rg in y-direction
iation of the in-plane self-diffusion coefficient, a self-diffusion
2
Rg in z-direction coefficient related to migration of molecules in the direction paral-
2.0 1500
lel to the interface, in the layers of liquid molecules in the vicinity
of the solid wall were analyzed in the present study using this
1000 method. Although the technical details are addressed in Ref. [22],
0.0 the method is briefly described here. The simulation system was
(001) 500 divided into slabs in the z-direction. The survival probability of
(011) the liquid molecules to be present in a slab is obtained from the
(100) following formula:
-2.0 0
-40 -30 -20 -10 0 Nð0; sÞ
Distance from surface Si atoms (Å) SðsÞ ¼ ; ð2Þ
Nð0Þ

−19
where N(0, s) indicates the number of molecules that are continu-
×10 ously present in the slab (PSk) during the entire time period s. Num-
8.0 3000
(b) Tetracosane at 450 K
ber of molecules in the slab at the beginning is represented by N(0).
Molecule was allocated to a slab where the center of mass of the
OH-terminated side 2500 molecule was situated. Mean square displacement in the x direction
6.0
Component of R2g (m2)

of the molecules that are present in the slab is calculated as follows:


2
R in x-direction D E
Density (kg/m )

g
2000 1 X
3

R2g in y-direction DxðsÞ2 ¼ ðxi ðsÞ  xi ð0ÞÞ2 : ð3Þ


4.0 PSk Nð0Þ i2PS
k
2
Rg in z-direction 1500
Utilizing this mean square displacement, self-diffusion coefficient
2.0 in the parallel direction was calculated by the following formula:
1000
hDxðsÞ2 iPSk
Dxx ðPSk Þ ¼ lim : ð4Þ
0.0 (001) s!1 2sSðsÞ
500
(011)
Using this modified Einstein relation, we can determine the diffu-
(100)
sion coefficient for each slab in the interface region. To observe
-2.0 0
0 10 20 30 40 the variation of diffusion in the interface region, small thickness
Distance from surface Si atoms (Å) of slabs is required. However, the reliability of the results deterio-
rates when thin slabs are applied, because of the small number of
Fig. 10. Components of radius of gyration as a function of distance for tetracosane molecules in the slab. So, there is a compromise between reliability
at T = 450 K on (a) the H-terminated side and (b) the OH-terminated side. Density
distribution is also plotted in the figure.
of result and finite details of diffusion in the interface region.
In the authors’ previous study [24], the modified Einstein rela-
tion has been used to obtain the variation of self-diffusion in the
In order to obtain knowledge about shape of the alkane mole- direction parallel to the interface in the liquid–vapor interface
cules in the interface region, the radius of gyration was examined. region. In the present work, the modified Einstein relation was
The detailed calculation procedure applied by the authors has been used to determine the variation of in-plane self-diffusion coeffi-
described in [32,24]. To obtain the variation of the radius of gyra- cient in the solid–liquid interface region in a similar way to the
tion, the simulation system was divided into slabs with a width of previous study for liquid–vapor interfaces. To obtain the variation
1 Å. Radius of gyration value of a molecule was allocated to a slab as a function of distance from the interface, the simulation box was
where the center of mass of the molecule was situated. The divided into slabs and the self-diffusion coefficient in each slab was
mean-squared total radius of gyration for tetracosane near the obtained from the mean square displacement (MSD) versus time
hydrophobic and hydrophilic surfaces, averaged within the slabs slope. In the present work, a slab thickness of approximately 8 Å
are shown in Fig. 9 for three crystal planes, and the component of was applied for the bulk liquid region and smaller values for the
the radius of gyration decomposed into the x, y and z direction interface regions so that the valley to valley (monolayer) in the
are shown in Fig. 10. The density distributions are also plotted in density profile for all major peaks is contained in one slab for all
both figures. It is observed from the figures that the total radius liquids. Fig. 11(a) shows an example of the slabs considered on
of gyration is constant in the bulk liquid region and increases stee- the density profile for the case of (0 0 1) crystal plane with liquid
ply in the first adsorption layers. This suggests that shape of the decane. Fig. 11(b) shows the survival probabilities as a function
alkane molecules is remarkably changed and stretched in the first of time for each slab. It should be noted that the width of slabs
adsorption layer in all cases of the three crystal planes and the #1–#4 and #12–#15 is smaller than that of slabs #5–#11 and, sur-
two terminations. It is observed from Fig. 10 that the z-component, vival probabilities decrease in the slabs #3–#4 and #12–#13. It is
which is normal to the solid surface, decreases while the parallel observed from the figure that the survival probabilities are remark-
components increase in the interface region. Based on the observa- ably higher for the slabs which are next to the solid wall, #1 and
tions of the orientation order parameter and the radius of gyration, #15, than the slabs in the bulk liquid region. An example of MSD
it is evident that molecules are flatten in the z direction and are versus time plots for decane at 345.6 K was shown in Fig. 11(c).
extended in the interface planar direction. In the bulk liquid region, all lines are almost overlapping each
854 H.K. Chilukoti et al. / International Journal of Heat and Mass Transfer 79 (2014) 846–857

−9
OH-side H-side ×10
(a) 2250 (a) 5.0 2500

15
12
13
14
5 7 8 9

1
6 10 11

2
3
4

Self-diffusion coefficient (m /s)


Self-diffusion coefficient

2
1800
4.0 2000
Density (kg/m3)

Density (kg/m3)
Methane at 106.96 K
1350 H-terminated side
3.0 1500
(001)
900 (011)
2.0 (100) 1000
Dxx
450 Dyy
1.0 500
Density
0
25 50 75 100 125 0.0 0
-50 -40 -30 -20 -10 0
Position-z (Å) Distance from surface Si atoms (Å)
1
(b) Decane at 345.6 K ×10−9
(001) crystal plane (b) 5.0 2500
0.8

Self-diffusion coefficient (m2/s)


Self-diffusion coefficient
Survival probability

slab # 1
slab # 2 4.0 2000
0.6 slab # 3

Density (kg/m3)
slab # 4
Methane at 106.96 K
3.0 OH-terminated side 1500
0.4 (001)
(011)
slab # 5-11 2.0 1000
slab # 12 (100)
0.2
slab # 13 Dxx
slab # 14 Dyy
slab # 15 1.0 500
0 Density
0 5 10 15 20
Time (ps)
0.0 0
0 10 20 30 40 50
×10−20 Distance from surface Si atoms (Å)
(c) 10.0
Fig. 12. Self-diffusion coefficients Dxx and Dyy as a function of position for methane
MSD parallel to the interface (m2)

at 106.96 K on (a) the H-terminated side and (b) the OH-terminated side. The
in OH-side interface density distributions are shown in dash dotted lines and the self-diffusion
8.0 in bulk liquid coefficient profiles are shown in solid lines with rectangular squares.
in H-side interface

Decane at 345.6 K Figs. 12–14 shows the variation of self-diffusion coefficients Dxx
6.0
and Dyy corresponding to the molecular migration in the parallel
(001) crystal plane direction to the interface, i.e., the in-plane self-diffusion coefficient.
4.0 The results for methane, decane and tetracosane at the same
reduced temperature are plotted as a function of distance from
the solid surface. The origin of the solid surface is again the equi-
2.0 librium position of the surface layer of silicon atoms which are ter-
minated with –H or –OH groups. It is observed from the figure that
self-diffusion coefficient is reduced near the solid walls and
0.0 increases as the position is distant from the wall. It approaches
0 2 4 6 8 10
the self-diffusion coefficient value in the bulk liquid region. Meth-
Time (ps) ane, decane and tetracosane liquids have similar in-plane diffusion
Fig. 11. For decane at 345.6 K near (0 0 1) crystal plane, (a) division of slabs on
characteristics near the –H and –OH terminated surfaces for a
density profile, (b) survival probabilities with respect to time for different slabs, and given crystal plane. For methane near the H-terminated side of
(c) mean square displacement as a function of time for different slabs. the (0 1 1) and (1 0 0) crystal planes, diffusion coefficient in the
few layers from the wall is extremely low (almost zero), which
indicates that molecules are strongly bound to the solid wall. Mol-
other. The red lines correspond to the slabs on the H-terminated ecules near the H-terminated surface for the (0 0 1) plane exhibit a
side while the blue lines correspond to the slabs on the OH-termi- certain value. On the other hand, in the side of the OH-terminated
nated side. The lines for the slabs in the interface regions are dis- surface, diffusion coefficient in the first and second adsorption lay-
tinguished from that of lines for the bulk liquid region, which ers is almost zero only in the case of the (0 1 1) plane, and it has
indicates that self-diffusion in the interface region is different from some significant nonzero values in the adsorption layer in other
that in bulk liquid region. cases of the crystal plane. In contrast with the cases of methane,
H.K. Chilukoti et al. / International Journal of Heat and Mass Transfer 79 (2014) 846–857 855

−9 −9
×10 ×10
(a) 5.5 2500 (a) 3.5 2500
Self-diffusion coefficient Self-diffusion coefficent
Self-diffusion coefficient (m /s)

Self-diffusion coefficient (m /s)


2

2
4.4 2000 2.8 2000
Decane at 345.6 K
Tetracosane at 450 K

Density (kg/m3)

Density (kg/m3)
H-terminated side
3.3 1500 2.1
H-terminated side 1500
(001)
(001)
(011)
(011)
2.2 (100) 1000 1.4 1000
(100)

Density Density
1.1 Dxx 500 0.7 Dxx 500
Dyy Dyy
0 0 0 0
-50 -40 -30 -20 -10 0 -50 -40 -30 -20 -10 0
Distance from surface Si atoms (Å) Distance from surface Si atoms (Å)

×10−9 ×10
−9

(b) 5.5 2500 (b) 3.5 2500


Self-diffusion coefficient
Self-diffusion coefficient (m /s)

Self-diffusion coefficient

Self-diffusion coefficient (m /s)


2

2
4.4 2000 2.8 2000
Decane at 345.6 K Tetracosane at 450 K
Density (kg/m3)

Density (kg/m )
OH-terminated side

3
OH-terminated side
3.3 1500 2.1 1500
(001) (001)
(011) (011)
2.2 1000 1.4 (100) 1000
(100)

1.1
Density 500 Density
Dxx 0.7 Dxx 500
Dyy Dyy
0.0 0 0 0
0 10 20 30 40 50 0 20 40
Distance from surface Si atoms (Å) Distance from surface Si atoms (Å)
Fig. 13. Self-diffusion coefficients Dxx and Dyy as a function of position for decane at
Fig. 14. Self-diffusion coefficients Dxx and Dyy as a function of position for
345.6 K on (a) the H-terminated side and (b) the H-terminated side. The density
tetracosane at 450 K on (a) the H-terminated side and (b) the OH-terminated side.
distributions are shown in dash dotted lines and the self-diffusion coefficient
The density distributions are shown in dash dotted lines and the self-diffusion
profiles are shown in solid lines with rectangular squares.
coefficient profiles are shown in solid lines with rectangular squares.

decane and tetracosane liquids show a comparatively large value For decane and tetracosane, near the OH-terminated side, diffu-
in the first adsorption layer near the solid wall for a particular crys- sion coefficient in the first adsorption layer is higher in the follow-
tal plane. For all three tested liquids, in-plane self-diffusion coeffi- ing order of the crystal plane, (0 0 1), (1 0 0) and (0 1 1). On the
cient Dxx in the adsorption layer contacting the (0 0 1) plane has H-terminated side of the (0 1 1) and (1 0 0) crystal planes, the diffu-
higher values than those of the other crystal plane cases on both sion coefficient in the first adsorption layer is approximately equal
terminations. This indicates that molecules migrate relatively fas- and significantly lower than those for the (0 0 1) crystal plane. It
ter on a smoother surface than that on a rough surface. It is should be noted that the surface terminal density for (1 0 0) crystal
observed in the figure that the self-diffusion coefficient along the plane is larger than the (0 1 1) crystal plane. It can be concluded
x direction is not same as the diffusion coefficient along the y direc- that configuration of solid molecules at the surface has also a sig-
tion for (0 0 1) and (1 0 0) crystal planes. Self-diffusion coefficient nificant effect on diffusion of liquid molecules contacting the sur-
values are not significantly different between the x and y directions face. It is also observed that for a particular liquid alkane,
for (0 1 1) crystal plane. Fig. 15 shows the surface terminal silica, H- molecules in contact with (0 0 1) crystal plane on the H-terminated
and OH-terminated atoms of three crystal planes. This anisotropic side migrate faster than the molecules in contact with the OH-
character in the in-plane self-diffusion coefficient is due to the dif- terminated side. For (0 1 1) and (1 0 0) crystal planes, molecules
ference in the topology along the x and y directions for a crystal on the OH-terminated side migrate relatively faster than those
plane. Fig. 15(a) and (b) illustrates that the lattice-scale surface on the H-terminated side. This suggests that surface terminal
roughness is small in the x-direction than that in the y direction. group has also an influence on in-plane diffusion of alkane. How-
Because of anisotropy, self-diffusion coefficient in the x-direction ever its tendency is different among the three crystal planes.
is significantly higher than that in the y-direction for the (0 0 1) Although the influence of liquid species on the diffusion in the
crystal plane near both terminations, which is more obvious in interface region is hard to be observed because self-diffusion coef-
case of longer liquid molecules. ficient in the bulk liquid region is not equal even at the same
856 H.K. Chilukoti et al. / International Journal of Heat and Mass Transfer 79 (2014) 846–857

Fig. 15. Snapshots of H-terminated and OH-terminated surface atoms for crystal plane of (a) (0 0 1), (b) (0 1 1), and (c) (1 0 0).

reduced temperature, the tendency can be observed that diffusion significant. For all the tested alkanes, the tendency of the liquid
coefficient value in the adsorption layer contacting the solid wall is molecules to penetrate into the solid wall is deeper on the OH-ter-
higher with larger chain length of molecules. It is suggested from minated side compared with that on the H-terminated side. It was
the result that the roughness of the solid surface, including the observed that the number of molecules in the first adsorption layer
termination groups, has an influence on the diffusion of liquid mol- is slightly dependent upon the crystal plane and type of surface
ecules in the first adsorption layer. termination for a given alkane. The parts of alkane molecules that
enter into cavities on the OH-terminated side of (1 0 0) crystal
4. Conclusions plane are in the direction perpendicular to the interface whereas
the parts of the molecules that enter into the cavities on the H-ter-
Structure and self-diffusion coefficient of three liquid alkanes in minated side are in the parallel direction. At the same reduced
the vicinity of three crystal planes of a-quartz silica substrate ter- temperature, decane and tetracosane liquid molecules have similar
minated with –H and –OH groups have been investigated using MD tendencies of alignment in the interface region for a particular
simulations. Density profiles exhibit oscillations that are a charac- crystal plane with both types of terminations. The variation of
teristic of liquid interfacial region contacting solids. The layering radius of gyration across the interfaces suggests the molecular
structure of liquid molecules is different according to the crystal shape is remarkably flattened in the z-direction in the layers close
plane of solid surface and termination type. This is the most obvi- to the wall.
ous in the case of methane. For long chain alkanes, influence of The variation of in-plane self-diffusion coefficient of three liquid
crystal plane and termination type on layering of molecules is less alkanes in the vicinity of a-quartz substrate for different crystal
H.K. Chilukoti et al. / International Journal of Heat and Mass Transfer 79 (2014) 846–857 857

planes is studied for the first time using MD simulations. It is [9] R. Notman, T.R. Walsh, Molecular dynamics studies of the interactions of water
and amino acid analogues with quartz surfaces, Langmuir 25 (2009) 1638–
observed that self-diffusion is lower near the wall and reaches to
1644.
the bulk liquid value away from the walls. It is shown that mole- [10] L.B. Wright, T.R. Walsh, Facet selectivity of binding on quartz surfaces: free
cules in touch with the smoother surface (0 0 1) migrate faster than energy calculations of amino-acid analogue adsorption, J. Phys. Chem. C116
the molecules in touch with rough surfaces (0 1 1) and (1 0 0). It was (2011) 8700–8709.
[11] T.S. Kim, R.H. Dauskardt, Molecular mobility under nanometer scale
found that the in-plane diffusion characteristics in the adsorption confinement, Nano Lett. 10 (2010) 1955–1959.
layer are anisotropic in nature and this is the most noticeable for [12] H.K. Christenson, D.W.R. Gruen, R.G. Horn, J.N. Israelachvili, Structuring in
the (0 0 1) crystal plane. At the same reduced temperature, self-dif- liquid alkanes between solid surfaces: Force measurements and meanfield
theory, J. Chem. Phys. 87 (1987) 1834–1841.
fusion coefficient of the tested alkane liquids in the first adsorption [13] A.N. Vasilyuk, R.M. Lynden-bell, A simulation study of films of n-hexane and n-
layer in cases of both types of terminations for all the crystal planes perfluorohexane on a solid surface, Mol. Phys. 99 (2001) 1407–1411.
tends to be higher with an increase in chain length of liquid alkane [14] E.B. Webb, G.S. Grest, Interfaces between silicate surfaces and liquid
hexadecane: a molecular dynamics simulation, J. Chem. Phys. 116 (2002)
molecules. From the results, it can be concluded that the surface 6311–6321.
topology and surface termination have significant influences on [15] D.K. Dysthe, A.H. Fuchs, B. Rousseau, Fluid transport properties by equilibrium
the diffusion of alkane liquids in the vicinity of a-quartz substrate. molecular dynamics. III: Evaluation of united atom interaction potential
models for pure alkanes, J. Chem. Phys. 112 (2000) 7581–7590.
[16] G.H. Goo, G. Sung, S.H. Lee, T. Chang, Diffusion behavior of n-alkanes by
Conflict of interest molecular dynamics simulations, Bull. Korean Chem. Soc. 23 (2002) 1595–
1603.
[17] V.A. Harmandaris, D. Angelopoulou, V.G. Mavrantzas, D.N. Theodorou,
None declared. Atomistic molecular dynamics simulation of diffusion in binary liquid n-
alkane mixtures, J. Chem. Phys. 116 (2002) 7656–7665.
Acknowledgments [18] S. Xu, G.C. Simmons, T.S. Mahadevan, G.W. Scherer, S.H. Garofalini, C. Pacheco,
Transport of water in small pores, Langmuir 25 (9) (2009) 5084–5090.
[19] M. Hu, J.V. Goicochea, B. Michel, D. Poulikakos, Water nanoconfinement
The work reported in this paper was supported by the Grant- induced thermal enhancement at hydrophilic quartz interfaces, Nano Lett. 10
in-Aid for Scientific Research and the Global COE Program ‘‘World (2010) 279–285.
[20] Niharendu Choudhury, Dynamics of water at the nanoscale hydrophobic
Center of Education and Research for Trans-Disciplinary Flow confinement, J. Chem. Phys. 132 (2010). 064505(1–5).
Dynamics’’ by the Japan Society for the Promotion of Science [21] P. Liu, E. Harder, B.J. Berne, On the calculation of diffusion coefficients in
(JSPS).Numerical simulations were performed on the SGI Altix confined fluids and interfaces with an application to the liquid–vapor interface
of water, J. Phys. Chem. B 108 (2004) 6595–6602.
UV1000 at the Advanced Fluid Information Research Center,
[22] C.D. Wick, L.X. Dang, Diffusion at the liquid–vapor interface of an aqueous
Institute of Fluid Science, Tohoku University. ionic solution utilizing a dual simulation technique, J. Phys. Chem. B 109
(2005) 15574–15579.
[23] S.R.V. Castrillon, N. Giovambattista, I.A. Aksay, P.G. Debenedetti, Evolution
References
from surface-influenced to bulk-like dynamics in nanoscopically confined
water, J. Phys. Chem. B 113 (2009) 7973–7976.
[1] M. Tsige, T. Soddemann, S.B. Rempe, G.S. Grest, J.D. Kress, M.O. Robbins, S.W. [24] H.K. Chilukoti, G. Kikugawa, T. Ohara, A molecular dynamics study on
Sides, M.J. Stevens, E. Webb III, Interactions and structure of poly transport properties and structure at the liquid–vapor interfaces of alkanes,
(dimethylsiloxane) at silicon dioxide surfaces: electronic structure and Int. J. Heat Mass Transfer 59 (2013) 144–154.
molecular dynamics studies, J. Chem. Phys. 118 (2003) 5132–5142. [25] S.K. Nath, R. Khare, New forcefield parameters for branched alkanes, J. Chem.
[2] D. Argyris, N.R. Tummala, A. Striolo, D.R. Cole, Molecular structure and Phys. 115 (2001) 10837–10844.
dynamics in thin water films at the silica and graphite surfaces, J. Phys. Chem. [26] M.G. Martin, J.I. Siepmann, Transferable potentials for phase equilibria. 1:
C112 (2008) 13587–13599. United-atom description of n-alkanes, J. Phys. Chem. B 102 (1998) 2569–2577.
[3] M.P. Gaigeot, M. Sprik, M. Sulpizi, Oxide/water interfaces: how the surface [27] P.E.M. Lopes, V. Murashov, M. Tazi, E. Demchuk, A.D. Mackerell, Development
chemistry modifies interfacial water properties, J. Phys.: Condens. Matter 24 of an empirical force field for silica. Application to the quartz-water interface,
(2012). 124106(1–11). J. Phys. Chem. B 110 (2005) 2782–2792.
[4] M. Ledyasuti, Y. Liang, M. Kunieda, T. Matsuoka, Asymmetric orientation [28] U. Essmann, L. Perera, M.L. Berkowitz, T. Darden, H. Lee, L.G. Pedersen, A
of toluene molecules at oil-silica interfaces, J. Chem. Phys 137 (2012). smooth particle mesh Ewald method, J. Chem. Phys. 103 (1995) 8577–8593.
064703(1–8). [29] M. Tuckerman, B.J. Berne, G.J. Martyna, Reversible multiple time scale
[5] H. Mo, G. Evmenenko, P. Dutta, Ordering of liquid squalane near a solid surface, molecular dynamics, J. Chem. Phys. 97 (1992) 1990–2001.
Chem. Phys. Lett. 415 (2005) 106–109. [30] A.A. Skelton, D.J. Wesolowski, P.T. Cummings, Investigating the quartz (1 0 0)/
[6] M.R. Warne, N.L. Allan, T. Cosgrove, Computer simulation of water molecules water interface using classical and Ab initio molecular dynamics, Langmuir 27
at kaolinite and silica surfaces, Phys. Chem. Chem. Phys. 2 (2000) 3663–3668. (2009) 1638–1644.
[7] R.T. Cygan, J.J. Liang, A.G. Kalinichev, Molecular models of hydroxide, [31] S.K. Nath, F.A. Escobedo, J.J. de Pablo, On the simulation of vapor–liquid
oxyhydroxide, and clay phases and the development of a general force field, equilibria for alkanes, J. Chem. Phys. 108 (1998) 9905–9911.
J. Phys. Chem. B108 (2004) 1255–1266. [32] J.G. Harris, Liquid–vapor interfaces of alkane oligomers. Structure and
[8] A.A. Skelton, P. Fenter, J.D. Kubicki, D.J. Wesolowski, P.T. Cummings, thermodynamics from molecular dynamics simulations of chemically
Simulations of the quartz/water interface: a comparison of classical force realistic models, J. Phys. Chem. 96 (1992) 5077–5086.
fields, ab initio molecular dynamics, and x-ray reflectivity experiments, J. Phys.
Chem. B115 (2011) 2076–2088.

You might also like