You are on page 1of 10

Microfluid Nanofluid

DOI 10.1007/s10404-014-1498-4

RESEARCH PAPER

Investigation ofpressuredriven gas flows innanoscale channels


using molecular dynamics simulation
FubingBao YuanlinHuang
YonghaoZhang JianzhongLin

Received: 3 June 2014 / Accepted: 29 September 2014


Springer-Verlag Berlin Heidelberg 2014

Abstract Pressure-driven gas flows in nanoscale channels are investigated in the present study using a molecular dynamics method. A reflecting particle membrane is
used to produce the pressure difference between the inlet
and outlet. The ratio of pressure difference to the average
pressure along the channel is studied in detail, which is
independent of the Knudsen number and temperature, but
increases with the membranes reflecting probability and
channel aspect ratio. A scaling law for the pressure ratio
is obtained from the simulation results, to determine the
membranes reflecting probability in molecular dynamics
simulation, so the required pressure difference over a given
channel can be produced. A case study of backward-facing
step flow is then conducted, which demonstrates the necessity of using the pressure-driven method in non-uniform
cross-sectional flow.
Keywords Pressure-driven Nanoscale gas flow
Molecular dynamics Poiseuille flow
Backward-facing step flow

F.Bao Y.Huang J.Lin(*)


Institute ofFluid Measurement andSimulation,
China Jiliang University, Hangzhou310018, China
e-mail: jzlin@sfp.zju.edu.cn
F.Bao Y.Zhang
James Weir Fluids Laboratory, Department ofMechanical
andAerospace Engineering, University ofStrathclyde,
GlasgowG1 1XJ, UK
J.Lin
Department ofMechanics, State Key Laboratory ofFluid
Power Transmission andControl, Zhejiang University,
Hangzhou310027, China

1Introduction
Nanoscale confined gas flows are usually encountered in
many micro- and nano-electromechanical systems (MEMS
and NEMS). For example, in magnetic disk drive units,
the distance between the head and media is on the order of
10nm (Juang and Bogy 2007). This characteristic length
scale is smaller than the mean free path of air at standard
ambient temperature and pressure, which is about 65nm.
It is usually considered that the continuum assumption is
no longer valid and the conventional NavierStokes equations cannot describe the flow behaviors at this scale (Colin
2013; Reese etal. 2003).
The extended hydrodynamic equations, such as the Burnett equations (Agarwal etal. 2001; Bao and Lin 2008) and
the high-order moment methods (Gu and Emerson 2009;
Taheri etal. 2009), may be able to describe flow behaviors
up to the early transition regime. But these models have
many problems, e.g., the boundary conditions (Barber and
Emerson 2006; Cao etal. 2009) and computational stability (Bao etal. 2012; Bobylev 2006). The direct simulation
Monte Carlo (DSMC) method (Bird 1994), which is especially developed for the rarefied gas flows, seems to be suitable to model the rarefied gas flows at micro-/nanoscale.
But in nanoscale confined gas flows, the gas-wall molecular
interactions dominate the flow characteristics (Barisik and
Beskok 2012). The DSMC method is not a good choice for
modeling the interactions between gas and wall molecules
(Markvoort etal. 2005).
The molecular dynamics (MD) method, which studies
the physical movements of atoms and molecules in the
context of N-body simulation, has been adopted by many
researchers in the past decade to investigate the gaseous
flow and heat transfer characteristics in nanoscale confined
channels (Babac and Reese 2014; Barisik and Beskok

13

2011, 2012; Cao etal. 2006; Xie and Liu 2012). Barisk
and Beskok (2011, 2012) investigated shear-driven gas
flows in nanoscale channels to reveal the gaswall interaction effects for flows in the transition and free molecular regimes. They found that density and stress profiles
exhibit a universal behavior inside the wall force penetration region at different flow conditions. Cao etal. (2006)
focused on the effects of the surface roughness on slip
flow of gaseous argon in submicron platinum channels.
They found that the friction coefficient increases with the
increase in surface roughness, and they owned this to the
distortion of the streamlines and the enhancement of the
penetrability near the rough surface. Xie and Liu (2012)
studied the gas flows in nano-channels with a Janus interface and found that the temperature has a significant influence on the particle number near the hydrophilic surface.
Recently, Babac and Reese (2014) investigated classical
thermosize effects by applying a temperature gradient
within the different-sized domains. These results provide
useful new insights into diffusive transport in nonequilibrium gas flows.
But in majority of the above studies, an artificial gravitation force is imposed on each molecule to drive the flow
accompanied by a periodic boundary condition in the flow
direction, recognized as gravitation-driven method (Koplik
etal. 1988). In order to induce appreciable flow, the gravitation force would need to be as high as 1012 times of the
earths gravitation. As a consequence, significant kinetic
energy rescaling is required to remove the generated heat.
This rescaling process will also disturb the particle dynamics (Li etal. 1998). In gravitation-driven method, the pressure and density keep uniform along the channel. This is
different from most nanoscale confined gas flows in MEMS
and NEMS, where compressibility is important and must
be taken into consideration (Karniadakis etal. 2006). Furthermore, when the geometry is non-uniform in cross
section, e.g., in backward-facing step flow, there may be
reverse pressure gradient in the channel; as such, a uniform
body forcing is no longer hydrodynamically equivalent to
the flow generated by an imposed pressure difference over
the same geometry.
Pressure-driven liquid flows using the MD method have
been investigated. Lupkowski and van Swol (1990) placed
two walls at the inlet and outlet and applied external forces
on them to control the pressure in simulations of liquid
liquid interfacial systems. Similarly, Huang etal. (2006)
applied external forces on self-adjusting plates located at the
outer boundaries of the system to achieve a specific pressure in each reservoir. These methods work well for dense
systems. However, since the number of molecules in the
simulation is fixed, all molecules will eventually be forced
to move out of the upstream reservoir, effectively ending the
simulation. In order to control the inlet and outlet pressures,

13

Microfluid Nanofluid

Sun and Ebner (1992) introduced an upstream source, which


is maintained at constant density and temperature, and a
downstream reservoir kept at vacuum. Density variation
along the channel can be achieved in 2D finite-length channel without periodic boundary conditions. But molecule
insertion and deletion are required and such a process would
disturb the dynamics of the system. Li etal. (1998) proposed a fictitious membrane to act as a filter. This membrane
allows the molecules to pass from one direction freely and
forces the other molecules to be elastically reflected with
a given probability. This method works well in liquid flow
simulation, and it does not require a thermostat. However,
no quantitative relationship between membranes reflecting
probability and pressure difference is known, which needs
to be investigated. A channel-moving pressure-driven
model was proposed by Zhang etal. (2009). The upper and
bottom wall moves and the molecules in the channels are
induced by the shear viscous force to move against a fixed
wall, hence resulting in a constant pressure gradient along
the channel. But this flow is not hydrodynamically equivalent to the pressure-driven flow.
In order to study the difference between osmotic permeability and diffusion permeability in water channels, Zhu
etal. (2004) applied a uniform gravitation force in a certain
region to induce a hydrostatic pressure difference across
the membrane. This method was later adopted by Nicholos etal. (2012) to study the water flow through nanotubes
and was further developed by Borg etal. (2013), where
they used a Gaussian distributed forcing to maintain a spatially smooth imposition of momentum to the MD fluid.
Recently, Docherty etal. (2014) compared the effects of
periodic region-based gravitation-driven Gaussian forcing
(Borg etal. 2013), non-periodic with reservoir pressure
control (Nicholls etal. 2012), and periodic gravitationdriven uniform forcing on generation of pressure-driven
water flows through carbon nanotubes.
But up to now, little has been done on pressure-driven
gas flows using the MD method. According to the pressure
tensor decomposition proposed by Irving and Kirkwood
(2004),

p = pK + pV ,

(1)

the kinetic term pK depends on gas density and molecular


square velocity, while the virial term pV depends on pairwise interactions between molecules. The hydrodynamic
pressure can be obtained using.

p=

1
tr(p).
3

(2)

In gaseous flow, the kinetic term in Eq. (1) dominates the


pressure because of large intermolecular distance. To etal.
(2012) modified the periodic boundary condition to account
for the pressure difference by maintaining the difference in

Microfluid Nanofluid

squared velocity at two opposite faces x=0 and x=L. But


the difference in squared velocity between the inlet and
outlet results in the difference of temperature.
According to the ideal gas law, the kinetic pressure is
proportional to the particle number density when temperature is constant. Hence, the pressure difference between the
inlet and outlet can be achieved by maintaining the density difference. The reflecting particle membrane (RPM)
method proposed by Li etal. (1998) is adopted and further
developed in the present study to precisely control the pressure difference between the inlet and outlet. This work is
organized as follows: In Sect.2, the implementations of
RPM method and MD method are presented; in Sect.3,
flow characteristics of gas confined between two walls are
investigated; a case study of non-uniform cross-sectional
backward-facing step flow is discussed in Sect.4; and
finally, Sect.5 concludes the present study.

2Setup ofpressuredriven gas flow


A fictitious membrane is placed at x =0. This membrane acts as a filter to allow the molecules crossing from
one direction (from left to right in Fig.1) to pass through
without hindrance, while the molecules moving from the
other direction (from right to left in Fig.1) are specularly
reflected with a certain probability, P. The membrane has
no other effect on the molecules. Molecules interact with
each other across the membrane in the same manner as
anywhere else in the fluid. Since no new particle or energy
is injected at any time, the particle number and total energy
of the system are conserved, thus allowing steady state flow
to be possible and no extra temperature rescaling is needed
(Li etal. 1998).
According to the ideal gas law,
(3)

p = nkB T ,

free pass

free pass, 1-P

Table1Parameters used in the LJ potential


ArAr
(m)
(J)

ArPt

3.4051010

3.0851010

21

0.8941021

1.65410

the pressure is proportional to the particle number density


when the temperature is constant. The gas number density
on the right of the membrane is larger than that on the left,
when membranes reflecting probability is larger than zero.
Thus, the pressure on the right of the membrane is larger
than the pressure on the left. The pressure difference across
the membrane increases when increasing the reflecting
probability.
Gas argon confined between two platinum walls that are
a distance H apart is investigated first. Periodic boundary
conditions are applied in all three directions. The interactions between gas molecules and gaswall molecules are
both described by the truncated and shifted Lennard-Jones
(LJ) 126 potential, given as:
12 6 12 6
4 r
rij rc
+ rc
rij rc
ij
V rij =
, (4)

0
rij > rc

where rij is the intermolecular distance between atoms i and


j, is the depth of the potential well, is the molecular
diameter and rc is the cutoff radius. In the present study, rc
is chosen to be 0.851nm, which is approximately equal to
2.5 Ar.
The parameters used in the LJ potential are listed
in Table1. These parameters have been used by many
researchers (Cao etal. 2006; Sun and Li 2008). The
mass of argon and platinum atoms is 6.641026 and
3.241025kg, respectively.
In our simulations, the neighbor-list method is used to
calculate the force between atoms, while the velocity-Verlet
algorithm is adopted to integrate the equations of motion
(Rapaport 2004). The time step in the simulation is set to
be 10.8fs. The first 1 million steps are used to equilibrate
the system, and another 5 million steps are used to accumulate properties in 200100 meshed bins along xy. The
Langevin thermostat method (Schneider and Stoll 1978)
is employed to control the gas temperature before equilibrium. Only thermal velocities are used to compute the temperature and pressure.

reflected, P

3Flow characteristics ofpressuredriven gaseous flow


membrane

Fig.1Schematic of reflecting particle membrane

Three-dimensional gas argon flow in a simulation box


of 340.535.036.81nm in the xyz directions is first
studied. Two face-centered-cubic (110) platinum walls

13

Microfluid Nanofluid

are placed along the xz plane, and the lattice constant is


0.392nm. The reflecting membrane is placed at x =0,
and the reflecting probability is chosen to be 0.8. A total
of 23,760 gas argon molecules and 121,450 solid platinum
atoms are used in the simulation. Wall atoms are frozen in
their initial positions, and interactions between them are
excluded when building the neighbor list to reduce computation. This treatment is acceptable as we do not focus on
the heat transfer characteristics near the wall. The thickness
of the wall is about 2.88 Ar, which is larger than rc and
prevents gas molecules interacting with each other crossing the wall. The initial temperature of gas is 300K and
the density is 20.0kg/m3, so the mean free path is about
6.44nm according to the hard sphere model,

1
=
,
2nd 2

(5)

where n and d are the gas number density and the diameter of the fluid molecules, respectively. For LJ fluids, the
molecular diameter d is approximately taken equal to the
parameter appearing in the LJ potential formula in Eq.
(4). The dimension of simulation box in lateral direction is
larger than one mean free path in the present study and is
sufficient to obtain MD solutions for gas flow independent
of the periodicity effects (Barisik and Beskok 2011). The
average Kn,

Kn =


,
H

(6)

is about 0.19. So the flow is in the early transition regime.


The location of the wall surface is defined at the center
of first row of the wall atoms facing the fluid. A number
of gas molecules are trapped by the wall, because of the
wall potential field. As a result, a density buildup near the
wall can be found, as illustrated in Fig.2a. The location of
(a)

density peak is about 1.1 ArPt away from the wall surface,
where the potential energy is lowest. When the distance of
a molecule to the wall is smaller than this value, the molecule experiences a repulsive force and moves away from
the wall; while the distance is larger than this value, the
molecule is subject to an attractive force and moves toward
the wall. The density profile varies with gaswall interaction strength and gas density (Barisik and Beskok 2012).
The streamwise velocity profile across the channel is
shown in Fig.2b, where we observe that in the bulk region
of the channel, the velocity profile agrees well with the
parabolic fitting. The NavierStokes equations can still
characterize the flow characters in this region. While in the
near-wall region, the velocity deviates from the parabolic
profile and its gradient is much larger than that in the bulk
region. The thickness of near-wall region is about 6 Ar
from the wall. Significant velocity slip can be found on the
wall. The velocity slip has been studied by many researchers, and it varies with gaswall interaction strength (Barisik
and Beskok 2012; Sun and Li 2008), wall roughness (Cao
etal. 2006), temperature (Cao etal. 2005) and Kn (Prabha
and Sathian 2012). Since we only consider pressure-driven
phenomenon in the present study, the correlation between
the velocity slip and these parameters is not discussed here.
The distributions of average pressure, density, temperature and streamwise velocity in the flow direction are
shown in Fig.3. Although the periodic boundary condition
is applied in the flow direction, macro-quantities are discontinuous across the membrane because of the existence
of RPM at x=0. As shown in Fig.3a, there is a large pressure difference between the inlet and outlet. The pressure
decreases gradually along the channel, so the negative pressure gradient drives the gas to flow. Because of the compressibility effect, the pressure distribution is nonlinear
along the channel. This is different from gravitation-driven
(b)

40

40

vx / ms-1

/ kgm

-3

30

20

10

50

30

MD simulation
parabolic fitting

20

0.2

0.4

y/H

0.6

0.8

10

0.2

0.4

y/H

0.6

0.8

Fig.2Profiles of gas density and streamwise velocity across the channel: a density, b streamwise velocity. The x-coordinate is normalized with
the channel height, H

13

Microfluid Nanofluid

(a) 1600

(b)

26
24

1400

/ kg m-3

p / kPa

22
1200

20
18

1000

800

pressure in pressure-driven flow


RT in pressure-driven flow
pressure in gravitation-driven flow

16
14

10

density in pressure-driven flow


density in gravitation-driven flow

(c)

10

x/H

x/H

(d)

400

60
55

350

vx / ms-1

T/K

50
300

45
40

250
35

temperature in pressure-driven flow


temperature in gravitation-driven flow

200

10

velocity in pressure-driven flow


velocity in gravitation-driven flow

30

10

x/H

x/H

Fig.3Distributions of flow quantities along the channel: a pressure, b density, c temperature and d streamwise velocity. The x-coordinate is
normalized with H

flow, where the pressure is constant along the channel.


The product of RT is also illustrated in Fig.3a, compared
with hydrodynamic pressure calculated by Eq. (2). The differences between them are about 0.39% at the inlet and
0.23% at the outlet. RT agrees very well with p, indicating that the virial term of pressure tensor in Eq. (1) is
small compared with the kinetic term because of the large
intermolecular distance (rave=4.64 Ar in the present study
when the density is 20.0kg/m3). So it can be neglected.
The bulk nanoscale gas can still be considered as ideal gas.
The pressure gradient in pressure-driven flow is mainly
caused by the density gradient, see Fig.3b, while the temperature almost keeps the same along the channel, see
Fig.3c. With the decrease in density along the channel, the
streamwise velocity increases gradually because of mass
conservation. The distributions of density and streamwise
velocity along the channel in the corresponding gravitationdriven flow are also illustrated in Fig.3, where the constant
density and streamwise velocity distributions along the
channel can be found. So, if we want to study flow characteristics in finite-length nanoscale channels, the gravitationdriven method is not appropriate for pressure-driven flow.

By adjusting membranes reflecting probability, different


pressure ratios, Rp, across the membrane can be achieved.
The pressure ratio is defined as:

Rp =

pin pout
,
pave

(7)

where pin and pout are obtained at a distance of 20 Ar away


from the membrane. The average pressure, pave, is defined
as the average pressure of the inlet and outlet pressures,

pave =

pin + pout
.
2

(8)

This average pressure may be a little smaller than that calculated from the pressure distribution along the channel,

pave =

L20
 Ar

pdx

x=20Ar

L 40Ar

(9)

However, Eq. (8) is much more convenient than Eq. (9).


From Eq. (6) and (8), we can get the relationship between
the pressure ratio Rp and the inlet/outlet pressure ratio,

13

Microfluid Nanofluid

2 + Rp
pin
=
.
pout
2 Rp

(10)

Thus, we can obtain the pressure ratio when the inlet and
outlet pressures are specified.
Variations of pressure ratio with membranes reflecting
probability are then investigated, by taking into account the
effects of gas density, temperature and channel aspect ratio,
L/H. The effect of gas density on pressure ratio is first studied
at several channel aspect ratios (L/H=5, 10, 15, 20, 25 and
40). The variation tendencies of pressure ratio with reflecting
probability at different channel aspect ratios are the same. So
only the results of L/H=10 are shown in Fig.4, as a representative example. The channel heights are 34.05nm, and
the temperatures are 300K. Five gas densities are studied,
from 1.68 to 33.6kg/m3, corresponding to Kn from 2.25 to
0.045, respectively. The gas flows are in the slip or transition
regimes. A second-order polynomial fitting,

Rp = C1 P2 + C2 P,

(11)

is also shown in Fig.4, where C1 and C2 are coefficients,


which are 0.63 and 0.78, respectively, when L/H =10.
When P =0, no molecule is reflected by the membrane
and the pressure ratio is zero. With the increase in membranes reflecting probability, more molecules are reflected
back from the membrane in the inlet region, which results
in larger pressure difference across the membrane. For
a certain P, the pressure ratios at all these five densities
are almost the same, indicating that the pressure ratio is
independent of gas density. This is because the average

intermolecular distances are all larger than rc in these five


cases, e.g., the average intermolecular distance is 3.68 Ar
when =33.6kg/m3. With the increase in density, the
intermolecular distance decreases and the gas can no longer
be considered as ideal gas. Then, the variation of pressure
ratio with P deviates from the second-order polynomial fitting of Eq. (11). In the present study, we only focus on gas
flows with densities smaller than 33.6kg/m3.
The effect of gas temperature on pressure ratio is also
studied. Five temperatures (T =200, 250, 300, 350 and
400K) are studied, and also we find that the pressure ratio
is independent of gas temperature.
The variations of pressure ratio with membranes reflecting probability at different channel aspect ratios are then
studied, at several gas densities (=1.68, 3.36, 8.4, 16.8
and 33.6kg/m3). Only the results of =16.8kg/m3 are
shown in Fig.5. Six different aspect ratios are investigated,
from L/H =5 to L/H =40. The pressure ratio increases
with membranes reflecting probability. This may be due
to the inlet and outlet effects. The variation of pressure
ratio with membranes reflecting probability can also be
described by the second-order polynomial in Eq. (11). A
polynomial fitting in the form of Eq. (11) can be obtained
for each channel aspect ratio with different gas densities
considered. Interestingly, the coefficients of linear term, C2,
in Eq. (11) are all close to 0.78 at different channel aspect
ratios, while the coefficient of quadratic term, C1, increases
with channel aspect ratio and gradually toward a constant
value. The variations of the coefficient C1 with channel
aspect ratio are shown in Fig.6. An exponent fitting

1.5

L /H = 5
L /H = 10
L /H = 15
L /H = 20
L /H = 25
L/H = 40

= 1.68 kg/m
3
= 3.36 kg/m
3
= 8.40 kg/m
3
= 16.8 kg/m
3
= 33.6 kg/m
Eq. (11)

1.2

1.6

1.2

Rp

Rp

0.9

0.6

0.8

0.3

0.4

0.2

0.4

0.6

0.8

Fig.4Pressure ratios as a function of membranes reflecting probability at different gas densities, where H=34.05nm, L/H=10 and
T=300K

13

0.2

0.4

0.6

0.8

Fig.5Pressure ratios as a functions of membranes reflecting probability at different channel aspect ratios, where H =34.05nm,
=16.8kg/m3 and T=300K

Microfluid Nanofluid

0.8

0.6

0.4

0.2
coefficient C1 in Eq. (11)
Eq.(12)

10

20

30

40

50

L/H

Fig.6Variations of the coefficient C1 in Eq. (11) with aspect ratios

C1 = 0.93 0.91e0.11L/H

(12)

can be obtained from the simulation results and is also


shown in the figure.
Finally, a quantitative relationship of pressure ratio with
membranes reflecting probability and channel aspect ratio
can be summarized as:


Rp = 0.93 0.91e0.11L/H P2 + 0.78P.
(13)
In the MD simulation of confined gas flows, for a given
geometry, when the inlet and outlet pressures are specified,
we can first use Eq. (8) to calculate the average pressure
and Eq. (7) to obtain the pressure ratio. The reflecting probability of membrane can then be obtained by solving Eq.
(13). The initial gas density in the MD simulation can be
calculated from the average pressure and temperature using
the ideal gas law.
From Eq. (13), we can find that the largest pressure ratio
across the membrane can be achieved is 1.71, because the
largest reflecting probability is 1. Thus, the largest inlet/
outlet pressure ratio is approximately 12.8, according to
Eq. (10). So the present RPM method is suitable when the
inlet/outlet pressure ratio is smaller than 12.8.
Pressure distributions in micro-channel have been
studied experimentally in a number of investigations
(Pong etal. 1994; Turner etal. 2004; Zohar etal. 2002).
Due to technical limitations, the ratio of channel length to
height is often on the order of 1,000. For instance, Zohar
etal. (2002) investigated gas argon flow in a straight and
uniform micro-channel utilizing standard micromachining techniques. The dimensions of the channel studied

are 4,0000.5340m in the xyz directions. The


pressure distribution along this channel was measured by
integral pressure micro-sensors. In their experiment, the
inlet and outlet pressures are 400 and 100kPa, respectively. These design parameters were selected based on
the limitations of the fabrication technology and the constraints of the connection to the external fluid handling
system. However, this geometry is still too large for a
MD simulation. If the same geometry and density are
used in the MD simulation, approximately 5 trillion gas
molecules should be used. This is beyond the computational capability of our small computer cluster. Thus,
a comparison with DSMC simulation (Liou and Fang
2006) is conducted here. The channel length is 2,000nm,
and the height is 400nm. Thus, the channel aspect ratio
is 5. The inlet and outlet pressures are 250 and 100kPa,
respectively. So the pressure ratio is 0.857, according to
Eq. (7). The reflecting probability of membrane used in
the MD simulation is solved to be 0.7816, according Eq.
(13). When the channel geometry, gas average density
and membranes reflecting probability are setup in the
MD simulation, 1.57 million argon molecules are used in
the simulation.
The simulation is carried out on a computer cluster
with 16 CPU for approximately 192h. The pressure distribution along the channel is shown in Fig.7, compared
with that of the DSMC simulation. The inlet and outlet
pressures sampled from the MD simulation are 253.4
and 101.4kPa, respectively. The deviations from objective pressures are 1.36 and 1.4%, indicating that RPM
method can be used to model the nanoscale pressuredriven flow.

250
DSMC
MD

200

p / kPa

150

100

x/H

Fig.7Pressure distributions along the channel, compared with


DSMC results

13

Microfluid Nanofluid

A pressure-driven backward-facing step flow is investigated


here as a case study, using the RPM method. The simulation box is 204.335.0317.03nm. The height of the
step is Hs=0.3 H and the length is Ls=1.5 H, where H is
the height of the channel after the step, which is 34.05nm
here. The membrane is placed at x =0 and its reflecting
probability is 0.8. The initial density of gas is 16.8kg/m3.
A total of 27,852 gas argon molecules and 185,745 wall
platinum atoms are used in the present simulation. The
results are sampled in 300100 bins over 5 million steps,
after 2 million steps for equilibrium.
The contour of pressure near the step is shown in Fig.8a,
and the pressure distributions along the flow direction at
y/H=0.15 (bottom-center) and y/H=0.65 (top-center) are
shown in Fig.8b. The top-center pressure decreases gradually along the channel. This is different from the gravitation-driven flow, where the pressures at the inlet and outlet
are the same. The pressure gradient before the step is larger
than that after the step, because of smaller cross-sectional
area. There is a reverse pressure gradient at the bottomcenter line after the step. Hence, the streamwise velocity
can be stagnant.
The contour of streamwise velocity near the wall is shown
in Fig.9a. This velocity first increases along the channel
before the step and then decreases because of flow expansion.

p / kPa

(a)

1400
1350
1300
1250
1200
1150
1100
1050

top-center

bottom-center

(b)

1600
top-center
bottom-center

p / kPa

1400

1000

800

x/H

Fig.8Pressure characteristics of nanoscale backward-facing step


flow: a the pressure contour near the step, b the pressure distributions
at the top-center and bottom-center lines

13

x/H=1.57

vx / ms-1
60
50
40
30
20
10
0
-10

75

(b)

60

45

30

15

0
0

0.2

0.4

0.6

0.8

y/H
Fig.9Streamwise velocity characteristics of nanoscale backwardfacing step flow: a the streamwise velocity contour, b the velocity
profile at x/H=1.57

The velocity profile after the step at x/H=1.57 is shown in


Fig.9b. The streamwise velocity profile at upper half of the
channel is parabolic, but at the lower half of the channel, it is
stagnant by the reverse pressure gradient. Reverse streamwise
velocities can be found after the step, indicating that there
exists vortex. The streamlines after the step are also shown in
Fig.9a. The reattachment length is about 0.5Hs in the present
case. Nanoscale backward-facing step flow has similar flow
characteristics as the macroscale equivalent.

5Conclusions

1200

600

(a)

vx / ms-1

4Nanoscale backwardfacing step flowa case study

The pressure-driven gas flows in the nanoscale channels are


investigated using the MD method. A RPM is used to produce the pressure difference between the inlet and outlet.
The molecules crossing the membrane from one direction
can pass through without hindrance, while the other molecules are specularly reflected with a certain probability.
The effects of membranes reflecting probability, Knudsen number, channel aspect ratio and temperature on pressure ratio are investigated. The pressure ratio is independent of Knudsen number and temperature, but increases with
membranes reflecting probability and channel aspect ratio.
A scaling law is proposed based on the simulation results,

Microfluid Nanofluid

which can be used to choose appropriate reflecting probability of the membrane to simulate pressure-driven flows
using the RPM method. A case study of nanoscale pressure-driven backward-facing step flow is then studied. A
reverse pressure gradient can be found after the step. The
streamwise velocity is stagnant and a vortex exists after the
step. The reattachment length is found to be about 0.5 Hs.
The nanoscale backward-facing step flow shares similar
flow characteristics with the corresponding macroscale one.
According to the scaling law, the largest pressure ratio
across the membrane can be achieved is 1.71. Thus, the
largest inlet/outlet pressure ratio is approximately 12.8. The
present RPM method is suitable when the inlet/outlet pressure ratio is smaller than 12.8. Further work is required to
produce large pressure ratio between the inlet and outlet of
micro-/nanoscale gaseous flows.
Acknowledgments This work was supported by the National Natural Science Foundation of China (Grant No. 11372298), the Major
State Basic Research Development Program of China (973 Program)
(GrantNo. 2011CB706501) and the Major Scientific and Technological Project of Zhejiang Province [Grant No. 2012C11015-3]. FB Bao
would also like to thank China Scholarship Council.

References
Agarwal RK, Yun KY, Balakrishnan R (2001) Beyond NavierStokes:
Burnett equations for flows in the continuum-transition regime.
Phys Fluids 13(10):30613085
Babac G, Reese JM (2014) Molecular dynamics simulation of classical thermosize effects. Anosc Microsc Therm 18(1):3953
Bao FB, Lin JZ (2008) Burnett simulation of gas flow and heat
transfer in micro Poiseuille flow. Int J Heat Mass Tranf
51(15):41394144
Bao FB, Zhu ZH, Lin JZ (2012) Linearized stability analysis of twodimension Burnett equations. Appl Math Model 36(5):19021909
Barber RW, Emerson DR (2006) Challenges in modeling gas-phase
flow in microchannels: from slip to transition. Heat Transf Eng
27(4):312
Barisik M, Beskok A (2011) Molecular dynamics simulations of
shear-driven gas flows in nano-channels. Microfluid Nanofluid
11(5):611622
Barisik M, Beskok A (2012) Surface-gas interaction effects on
nanoscale gas flows. Microfluid Nanofluid 13(5):789798
Bird GA (1994) Molecular gas dynamics and the direct simulation of
gas flows. Oxford University Press, Oxford
Bobylev AV (2006) Instabilities in the Chapman-Enskog expansion
and hyperbolic Burnett equations. J Stat Phys 124(24):371399
Borg MK, Lockerby DA, Reese JM (2013) A hybrid molecularcontinuum simulation method for incompressible flows in
micro/nanofluidic networks. Microfluid Nanofluid 15(4):541557
Cao BY, Chen M, Guo ZY (2005) Temperature dependence of the
tangential momentum accommodation coefficient for gases. Appl
Phys Lett 86(9):091905
Cao BY, Chen M, Guo ZY (2006) Effect of surface roughness on gas
flow in microchannels by molecular dynamics simulation. Int J
Eng Sci 44(13):927937
Cao BY, Sun J, Chen M, Guo ZY (2009) Molecular momentum transport at fluid-solid interfaces in MEMS/NEMS: a review. Int J Mol
Sci 10(11):46384706

Colin S (2013) Microfluidics. ISTE and Wiley, London


Docherty SY, Nicholls WD, Borg MK, Lockerby DA, Reese JM
(2014) Boundary conditions for molecular dynamics simulations
of water transport through nanotubes. Proc Inst Mech Eng Part C
J Mech Eng Sci 228(1):186195
Gu XJ, Emerson DR (2009) A high-order moment approach for capturing non-equilibrium phenomena in the transition regime. J
Fluid Mech 636:177216
Huang C, Nandakumar K, Choi PY, Kostiuk LW (2006) Molecular
dynamics simulation of a pressure-driven liquid transport process
in a cylindrical nanopore using two self-adjusting plates. J Chem
Phys 124(23):234701
Irving J, Kirkwood JG (2004) The statistical mechanical theory of
transport processes. IV. The equations of hydrodynamics. J Chem
Phys 18(6):817829
Juang JY, Bogy DB (2007) Air-bearing effects on actuated thermal pole-tip protrusion for hard disk drives. J Tribol 129(3):
570578
Karniadakis G, Beskok A, Aluru NR (2006) Microflows and nanoflows: fundamentals and simulation. Springer, New York
Koplik J, Banavar JR, Willemsen JF (1988) Molecular dynamics of Poiseuille flow and moving contact lines. Phys Rev Lett
60(13):1282
Li J, Liao DY, Yip S (1998) Coupling continuum to molecular-dynamics simulation: reflecting particle method and the field estimator.
Phys Rev E 57(6):7259
Liou WW, Fang Y (2006) Microfluid mechanics: principles and modeling. McGraw-Hill, New York
Lupkowski M, van Swol F (1990) Computer simulation of fluids
interacting with fluctuating walls. J Chem Phys 93:737
Markvoort AJ, Hilbers PAJ, Nedea SV (2005) Molecular dynamics
study of the influence of wall-gas interactions on heat flow in
nanochannels. Phys Rev E 71(6):066702
Nicholls WD, Borg MK, Lockerby DA, Reese JM (2012) Water transport through (7, 7) carbon nanotubes of different lengths using
molecular dynamics. Microfluid Nanofluid 12(14):257264
Pong K-C, Ho C-M, Liu J, Tai Y-C (1994) Non-linear pressure distribution in uniform microchannels. ASME-FED 197:51
Prabha SK, Sathian SP (2012) Molecular-dynamics study of poiseuille
flow in a nanochannel and calculation of energy and momentum
accommodation coefficients. Phys Rev E 85(4):041201
Rapaport DC (2004) The Art of Molecular Dynamics Simulation.
Cambridge University Press, New York
Reese JM, Gallis MA, Lockerby DA (2003) New directions in fluid
dynamics: non-equilibrium aerodynamic and microsystem flows.
Philos Trans R Soc Lond A 361(1813):29672988
Schneider T, Stoll E (1978) Molecular-dynamics study of a threedimensional one-component model for distortive phase transitions. Phys Rev B 17(3):1302
Sun M, Ebner C (1992) Molecular-dynamics simulation of compressible fluid flow in two-dimensional channels. Phys Rev A
46(8):4813
Sun J, Li ZX (2008) Effect of gas adsorption on momentum accommodation coefficients in microgas flows using molecular dynamic
simulations. Mol Phys 106(19):23252332
Taheri P, Torrilhon M, Struchtrup H (2009) Couette and Poiseuille
microflows: analytical solutions for regularized 13-moment equations. Phys Fluids 21(1):017102
To QD, Tung PT, Guy L, Cline L (2012) Molecular dynamics simulations of pressure-driven flows and comparison with acceleration-driven flows. Adv Mech Eng 2012:580763
Turner SE, Lam LC, Faghri M, Gregory OJ (2004) Experimental investigation of gas flow in microchannels. J Heat Transf
126(5):753763
Xie H, Liu C (2012) Molecular dynamics simulations of gas flow in
nanochannel with a Janus interface. AIP Adv 2(4):042126

13


Zhang ZQ, Zhang HW, Ye HF (2009) Pressure-driven flow in parallelplate nanochannels. Appl Phys Lett 95(15):154101
Zhu FQ, Tajkhorshid E, Schulten K (2004) Theory and simulation of
water permeation in aquaporin-1. Biophys J 86(1):5057

13

Microfluid Nanofluid
Zohar Y, Lee SYK, Lee WY, Jiang L, Tong P (2002) Subsonic gas
flow in a straight and uniform microchannel. J Fluid Mech
472:125151

You might also like