scispace - formally typeset
Open AccessJournal ArticleDOI

Molecular scale contact line hydrodynamics of immiscible flows.

TLDR
This work gives a continuum formulation of the immiscible flow hydrodynamics, comprising the generalized Navier boundary condition, the Navier-Stokes equation, and the Cahn-Hilliard interfacial free energy, and yields interfacial and velocity profiles matching those from the molecular dynamics simulations at the molecular-scale vicinity of the contact line.
Abstract
From extensive molecular dynamics simulations on immiscible two-phase flows, we find the relative slipping between the fluids and the solid wall everywhere to follow the generalized Navier boundary condition, in which the amount of slipping is proportional to the sum of tangential viscous stress and the uncompensated Young stress. The latter arises from the deviation of the fluid-fluid interface from its static configuration. We give a continuum formulation of the immiscible flow hydrodynamics, comprising the generalized Navier boundary condition, the Navier-Stokes equation, and the Cahn-Hilliard interfacial free energy. Our hydrodynamic model yields interfacial and velocity profiles matching those from the molecular dynamics simulations at the molecular-scale vicinity of the contact line. In particular, the behavior at high capillary numbers, leading to the breakup of the fluid-fluid interface, is accurately predicted.

read more

Content maybe subject to copyright    Report

Molecular scale contact line hydrodynamics of immiscible flows
Tiezheng Qian
Department of Physics and Institute of Nano Science and Technology, The Hong Kong University of Science and Technology,
Clear Water Bay, Kowloon, Hong Kong, China
Xiao-Ping Wang
Department of Mathematics, The Hong Kong University of Science and Technology, Clear Water Bay, Kowloon, Hong Kong, China
Ping Sheng
*
Department of Physics and Institute of Nano Science and Technology, The Hong Kong University of Science and Technology,
Clear Water Bay, Kowloon, Hong Kong, China
Received 23 October 2002; revised manuscript received 2 May 2003; published 17 July 2003
From extensive molecular dynamics simulations on immiscible two-phase flows, we find the relative slip-
ping between the fluids and the solid wall everywhere to follow the generalized Navier boundary condition, in
which the amount of slipping is proportional to the sum of tangential viscous stress and the uncompensated
Young stress. The latter arises from the deviation of the fluid-fluid interface from its static configuration. We
give a continuum formulation of the immiscible flow hydrodynamics, comprising the generalized Navier
boundary condition, the Navier-Stokes equation, and the Cahn-Hilliard interfacial free energy. Our hydrody-
namic model yields interfacial and velocity profiles matching those from the molecular dynamics simulations
at the molecular-scale vicinity of the contact line. In particular, the behavior at high capillary numbers, leading
to the breakup of the fluid-fluid interface, is accurately predicted.
DOI: 10.1103/PhysRevE.68.016306 PACS numbers: 47.11.j, 68.08.p, 83.10.Mj, 83.10.Ff
I. INTRODUCTION
Immiscible two-phase flow in the vicinity of the contact
line CL, where the fluid-fluid interface intersects the solid
wall, is a classical problem that falls beyond the framework
of conventional hydrodynamics 1–12. In particular, mo-
lecular dynamics MD studies have shown relative slipping
between the fluids and the wall, in violation of the no-slip
boundary condition 6,7. There have been numerous ad-hoc
models 1,8,1012 to address this phenomenon, but none
was able to give a quantitative account of the MD slip ve-
locity profile in the molecular-scale vicinity of the CL. While
away from the moving CL the small amount of relative slip-
ping was found to follow the Navier boundary condition
NBC兲关13, i.e., relative slipping proportional to the tangen-
tial viscous stress, in the molecular-scale vicinity of the CL
the NBC failed totally to account for the near-complete slip.
This failure casts doubts on the general applicability of the
NBC to immiscible flows and hinders a continuum formula-
tion of the hydrodynamics in the CL region. In particular, a
possible breakdown in the hydrodynamic description for
the molecular-scale CL region has been suggested 7.In
another approach 14, it was shown that the MD results can
be reproduced by continuum finite element simulations, pro-
vided the slip profile extracted from MD is used as input.
This work demonstrated the feasibility of the hybrid algo-
rithm, but left unresolved the problem concerning the bound-
ary condition governing the CL motion. Without a continuum
hydrodynamic formulation, it becomes difficult or impos-
sible to have realistic simulations of micro- or nanofluidics,
or of immiscible flows in porous media where the relative
wetting characteristics, the moving CL dissipation, and be-
havior over undulating solid surfaces may have macroscopic
implications.
From MD simulations on immiscible two-phase flows, we
report the finding that the generalized Navier boundary con-
dition GNBC applies for all boundary regions, whereby the
relative slipping is proportional to the sum of tangential vis-
cous stress and the uncompensated Young stress. The latter
arises from the deviation of the fluid-fluid interface from its
static configuration 10. By combining GNBC with the
Cahn-Hilliard CH hydrodynamic formulation of two-phase
flow 11,12, we obtained a consistent, continuum descrip-
tion of immiscible flow with material parameters such as
viscosity, interfacial tension, etc directly obtainable from
MD simulations. The convective-diffusive dynamics in the
vicinity of the interface and the moving CL also means the
introduction of two phenomenological dynamic parameters
whose values can be fixed by comparison with one MD flow
profile. Once the parameter values are determined from MD
simulations, our continuum hydrodynamics can yield predic-
tions matching those from MD simulations for different
Couette and Poiseuille flows. Our findings suggest the no-
slip boundary condition to be an approximation to the
GNBC, accurate for most macroscopic flows but failing in
immiscible flows. These results open the door to efficient
simulations of nano- or microfluidics involving immiscible
components, as well as to macroscopic immiscible flow cal-
culations, e.g., in porous media, that are physically meaning-
ful at the molecular level 15. The latter is possible, for
example, by employing the adaptive method based on the
iterative grid redistribution introduced in Ref. 15. This
*
Author to whom correspondence should be addressed. Email ad-
dress: sheng@ust.hk
PHYSICAL REVIEW E 68, 016306 2003
1063-651X/2003/681/01630615/$20.00 ©2003 The American Physical Society68 016306-1

method has demonstrated the capability of resolving, at the
same time, both the global behavior of a partial differential
equation solution with coarse mesh and a strong singularity
in a localized region with a refined local mesh of over 10
5
ratio to the coarse mesh.
II. MOLECULAR DYNAMICS SIMULATIONS
The MD simulations were performed for both static and
dynamic configurations in the Couette and Poiseuille flows.
The two immiscible fluids were confined between two paral-
lel walls separated along the z direction, with the fluid-solid
boundaries defined by z 0,H see Fig. 1 for the Couette
geometry. Interaction between the fluid molecules was mod-
eled by a modified Lennard-Jones LJ potential U
ff
4
(
/r)
12
ff
(
/r)
6
, where r is the distance between
the molecules,
and
are the energy scale and the range of
interaction, respectively, and
ff
1 for like molecules and
ff
⫽⫺1 for molecules of different species. Each of the two
walls was constructed by two or more兲关001 planes of an
fcc lattice see Appendix A, with each wall molecule at-
tached to the lattice site by a harmonic spring. The mean-
squared displacement of wall molecules was controlled to
obey the Lindemann criterion. The wall-fluid interaction was
also modeled by a LJ potential U
wf
, with energy and range
parameters
wf
1.16
and
wf
1.04
, and a
wf
for speci-
fying the wetting property of the fluid. Both U
ff
and U
wf
were cut off at r
c
2.5
. The mass of the wall molecule was
set equal to that of the fluid molecule m, and the average
number densities for the fluids and wall were set at
0.81/
3
and
w
1.86/
3
, respectively. The temperature
was controlled at 2.8
/k
B
, where k
B
is Boltzmann’s con-
stant. Moving the top and bottom walls at a constant speed V
in the x directions, respectively, induced the Couette flow
7. Applying a body force mg
ext
to each fluid molecule in
the x direction induced the Poiseuille flow 6. Periodic
boundary conditions were imposed on the x and y boundaries
of the sample. Most of our MD simulations were carried out
on samples consisting 6144 atoms for each fluid and 2880
atoms for each wall. The sample is 163.5
by 6.8
along the
x and y, respectively, and H 13.6
. Our MD results repre-
sent time averages over 2040 million time steps. For tech-
nical details of our MD simulations, we followed those de-
scribed in Ref. 16.
Two different cases were considered in our simulations.
The symmetric case refers to identical wall-fluid interactions
for the two fluids both
wf
1), which leads to a flat static
interface in the yz plane with a 90° contact angle. The asym-
metric case refers to different wall-fluid interactions, with
wf
1 for one and
wf
0.7 for the other. The resulting
static interface is a circular arc with a 64° contact angle. We
measured six quantities in the Couette-flow steady states of
V 0.25(
/m)
1/2
, H 13.6
for the symmetric case and V
0.2(
/m)
1/2
, H13.6
for the asymmetric case:
v
x
slip
, the
slip velocity relative to the moving wall; G
x
w
, the tangential
force per unit area exerted by the wall; the
xx
,
nx
compo-
nents of the fluid stress tensor (n denotes the outward surface
normal, and
v
x
,
v
z
.
We denote the region within 0.85
z
0
of the wall the
boundary layer BL. It must be thin enough to render suffi-
cient precision for measuring
v
x
slip
, while thick enough to
fully account for the tangential wall-fluid interaction force,
due to the finite range of the LJ interaction. Thus, it is not
possible to do MD measurements strictly at the fluid-solid
boundary, not only because of poor statistics, but also be-
cause of this intrinsic limitation. The wall force can be iden-
tified by separating the force on each fluid molecule into
wall-fluid and fluid-fluid components. For 0 zz
0
the fluid
molecules can detect the atomic structure of the wall. When
coupled with kinetic collisions with the wall molecules, there
arises a nonzero tangential wall force that varies along the z
direction and saturates at zz
0
. G
x
w
is the saturated total
tangential wall force per unit wall area Fig. 2. In Appendix
A we give account of our MD results on both the tangential
and normal components of the wall force, plus the effects
of increasing the wall thickness in our simulations from two
layers of wall molecules to four layers and to infinite layers
by using the continuum approximation beyond the four lay-
ers.
Spatial resolution along the x and z directions was
achieved by evenly dividing the sampling region into bins,
each x 0.425
by z 0.85
in size.
v
x
slip
was obtained
as the time average of fluid molecules’ velocities inside the
BL, measured with respect to the moving wall; G
x
w
was ob-
tained from the time average of the total tangential wall force
experienced by the fluid molecules in the BL, divided by the
bin area in the xy plane;
xx(nx)
was obtained from the time
averages of the kinetic momentum transfer plus the fluid-
fluid interaction forces across the constant-x(z) bin surfaces,
FIG. 1. Color Segments of the MD simulation sample for the
immiscible Couette flows. The colored dots indicate the instanta-
neous molecular positions of the two fluids projected onto the xz
plane. The black gray circles denote the wall molecules. The up-
per panel illustrates the symmetric case; the lower panel illustrates
the asymmetric case. The red circles and the blue squares represent
the time-averaged interface profiles, defined by
1
2
(
0), for
the two cases. The black solid lines are the interface profiles calcu-
lated from the continuum hydrodynamic model with the GNBC.
QIAN, WANG, AND SHENG PHYSICAL REVIEW E 68, 016306 2003
016306-2

and
v
x(z)
was measured as the time-averaged velocity com-
ponents within each bin. For the contribution of intermo-
lecular forces to the stress, we have directly measured the
fluid-fluid interaction forces across bin surfaces instead of
using the Irving-Kirkwood expression 17, whose validity
was noted to be not justified at the fluid-fluid or the wall-
fluid interface see the paragraph following Eq. 5.15 in the
above reference. In Appendix B we give some details on our
MD stress measurements. As reference quantities, we also
measured G
x
w0
,
xx
0
,
nx
0
in the static (V 0) configuration.
In addition, we measured in both the static and dynamic
configurations the average molecular densities
1
and
2
for
the two fluid species in each bin to determine the interface
profile. The shear viscosity
1.95
m/
2
and the interfa-
cial tension
5.5
/
2
were also determined.
We have also measured the interface and velocity profiles
for the Poiseuille flow in the asymmetric case, as well as for
the Couette flows with different V and H in the symmetric
case.
III. GENERALIZED NAVIER BOUNDARY CONDITION
In the presence of a fluid-fluid interface, the static fluid
stress tensor
0
reflects the static Young stress surface ten-
sion as well as those stresses arising from wall-fluid inter-
action. This is the case in spite of the fact that in all the MD
fluid stress measurements only the fluid-fluid interaction was
counted see Appendix B 2. The reason is that because the
MD measurements were carried out either in the static equi-
librium state or in the dynamic steady state, local force bal-
ance necessarily requires the fluid stress components to fully
reflect the influence of the wall-fluid interaction. For the con-
sideration of moving CL, we will be concerned with the part
of the fluid stress tensor, which is purely dynamic in origin,
i.e., arising purely from the hydrodynamic motion of the
fluid and the CL. In the notations below, the over tilde
denotes the difference between that quantity and its static
part. Thus, if
is the total stress, we will be concerned only
with the hydrodynamic part, denoted by
˜
0
. We note
that in the absence of body forces, the momentum equation
in bulk fluid is given by
m
v/
t (v•“)v
“•
˜
. In the
BL, the wall-fluid interaction means the existence of a dy-
namic, tangential wall force density g
˜
x
w
such that the force
balance equation is given by (“•
˜
)x
ˆ
g
˜
x
w
0 inside the
BL. The tangential wall force density g
˜
x
w
, shown explicitly
in the inset to Fig. 2, is a function sharply peaked at z
z
0
/2. Here we note that the boundary layer thickness is
extremely small (z
0
0.85
), hence the inertial effect may
be neglected (m
Vz
0
/
0.1). MD evidence for an inte-
grated form of the steady-state force balance is shown in Fig.
3. The total tangential force exerted by the wall on the fluid
is given by G
˜
x
w
0
z
0
dzg
˜
x
w
per unit wall area. In steady state,
this wall force is necessarily balanced by the tangential fluid
force G
˜
x
f
0
z
0
dz(
x
˜
xx
z
˜
zx
) inset to Fig. 3兲关18. Here
x,z,n
means taking partial derivative with respect to x, z,or
surface normal.
We now present evidences to show that everywhere on the
boundaries, relative slipping is proportional to G
˜
x
f
the
GNBC, see also Eq. 3 below:
G
˜
x
f
v
x
slip
, 1
where
is the slip coefficient and G
˜
x
f
can be written as
G
˜
x
f
x
0
z
0
dz
˜
xx
z
˜
nx
z
0
, 2
where we have used the fact that
˜
zx
(0) 0. More strictly,
˜
zx
(0
) 0 because there is no fluid below z 0 and hence
no momentum transport across z 0] Here the z coordinate
is for the lower fluid-solid boundary same below, with the
understanding that the same physics holds at the upper
boundary, and
n
⫽⫺
z
for the lower boundary.
Force balance means that at steady state, the frictional
force exerted by the solid wall on the moving slipping fluid
is fully accounted for in G
˜
x
f
. Thus, the GNBC or NBC can
be expressed in either G
˜
x
f
or G
˜
x
w
, but not both. In Fig. 3 we
show the measured MD data for the symmetric and asym-
metric cases in the Couette geometry. The symbols represent
the values of G
˜
x
f
measured in the BL. The solid lines repre-
sent the values of G
˜
x
f
calculated from
v
x
slip
by using
1
2
1.2
m/
3
for the symmetric case and
1
1.2
m/
3
,
2
0.532
m/
3
for the asymmetric case
away from the CL region straight line segments in Fig. 3,
and
(
1
1
2
2
)/(
1
2
) in the CL region 19, with
v
x
slip
and
1,2
obtained from MD simulations. It is seen that
FIG. 2. By subdividing the boundary layer into thin sections, we
plot the accumulated wall force per unit wall area as a function of
distance z away from the boundary. Here G
˜
x
w
(z) is defined by
G
˜
x
w
(z)
0
z
dz
g
˜
x
w
(z
), where g
˜
x
w
is the density of tangential wall
force. For different x positions, the absolute value of the saturating
total wall force is different. However, when normalized by the cor-
responding saturated total wall force per unit area at each x, all
points fall on a universal curve, nearly independent of x. It is seen
that at z z
0
the wall force has reached its saturation value. Inset:
Tangential wall force density plotted as a function of distance z
away from the boundary. The solid lines are averaged g
˜
x
w
in thin
sections at different x, normalized by the corresponding saturated
total wall force per unit area. The dashed line is a smooth Gaussian
fit. It is seen that g
˜
x
w
is a function sharply peaked at zz
0
/2. In the
sharp boundary limit this peaked wall force density is approximated
by G
˜
x
w
(z).
MOLECULAR SCALE CONTACT LINE HYDRODYNAMICS... PHYSICAL REVIEW E 68, 016306 2003
016306-3

for the lower boundary upper right panel, the MD data
agree well with the predictions of Eq. 1. For the upper
boundary lower left panel the straight line segments also
agree well with Eq. 1. However, there is some discrepancy
in the interfacial region of the upper boundary that seems to
arise from a ‘shear thinning’ effect of decreasing
at very
large tangential stresses 13.
The fact that the wall force density is distributed inside a
thin BL and vanishes beyond the BL necessitates the form of
G
˜
x
f
as defined by Eq. 2. However, it is intuitively obvious
that the fluids would experience almost the identical physical
effects from a wall force density G
˜
x
w
(z), concentrated
strictly at the fluid-solid boundary with the same total wall
force per unit area. In the inset to Fig. 2, it is shown that the
MD-measured wall force density is a sharply peaked func-
tion. The sharp boundary limit involves the approximation of
replacing this peaked function by
(z). The replacement of a
diffuse boundary by a sharp boundary can considerably sim-
plify the form of the GNBC, because local force balance
along x then requires
x
˜
xx
z
˜
zx
0 away from the
boundary z 0. Integration of this relation from 0
to z
0
yields
x
0
z
0
dz
˜
xx
z
˜
zx
z
0
˜
zx
0
0
FIG. 3. Color
1
V/G
˜
x
f
plotted as a function of V/
v
x
slip
. Sym-
bols are MD data measured in the BL at different x locations, where
the red circles denote the symmetric case and the blue squares de-
note the asymmetric case. The solid lines were calculated from Eq.
1 with values of
1,2
and the expression of
given in the text.
The statistical errors of the MD data are about the size of the sym-
bols. The upper-right data segment corresponds to the lower bound-
ary, whereas the lower-left segment corresponds to the upper
boundary. The slopes of the two dashed lines are given by
1,2
1
.
Inset: G
˜
x
w
plotted as a function of G
˜
x
f
, measured in the two BLs at
different values of x. The symbols have the same correspondence as
in the main figure. The data are seen to lie on a straight line with a
slope of 1, indicating G
˜
x
w
G
˜
x
f
0.
FIG. 4. Color Two components of the dynamic tangential
stress at z z
0
, plotted as a function of x. The dashed lines denote
˜
zx
Y
; solid lines represent the viscous component. Here red indicates
the symmetric case and blue indicates the asymmetric case. In the
CL region the nonviscous component is one order of magnitude
larger than the viscous component. The difference between the two
components, however, diminishes towards the boundary, z0, due
to the large interfacial pressure drop implying a large curvature in
the BL, thereby pulling
d
closer to
s
. Inset:
d,s
plotted as a
function of
cos
d,s
at different values of z. Here
d
⫽⫺
dx(
nx
nx
v
),
s
⫽⫺
dx
nx
0
, and
d,s
was measured from
the time-averaged interfacial profiles Fig. 1. The red circles denote
the symmetric case, the blue squares denote the asymmetric case,
the solid blue squares denote the asymmetric static case, and the
single solid red circle at the origin denotes the symmetric static
case. The data are seen to follow a straight dashed line with slope
1, indicating
d,s
cos
d,s
.
FIG. 5. Color S
0
z
0
˜
xx
(z)dz
0
z
0
xx
(z)
xx
0
(z)
dz plot-
ted as a function of x. Here red circles denote the symmetric case
and blue squares denote the asymmetric case. For clarity,
xx
0
was
vertically displaced such that
xx
0
0 far from the interface in the
symmetric case, and for the asymmetric case,
xx
0
0 at the center
of the interface.
QIAN, WANG, AND SHENG PHYSICAL REVIEW E 68, 016306 2003
016306-4

and as a consequence by comparing with Eq. 2兲兴 G
˜
x
f
⫽⫺
˜
nx
(0
). Therefore,
˜
zx
changes from
˜
zx
(0
) 0to
˜
zx
(0
) G
˜
x
f
at z 0, leading to
“•
˜
x
ˆ
G
˜
x
f
z
.
Comparing with the diffuse boundary, where (“•
˜
)x
ˆ
g
˜
x
w
0, we see that the form of the equation remains the same,
but the BL is now from 0
to 0
, instead of from 0
to z
0
as in the diffuse case. Thus, GNBC 1 becomes
˜
nx
0
v
x
slip
in the sharp boundary limit.
The tangential stress
˜
nx
can be decomposed into a vis-
cous component and a non-viscous component:
˜
nx
z
nx
v
z
˜
nx
Y
z
.
In Fig. 4 we show that away from the interfacial region the
tangential viscous stress
nx
v
(z)
(
n
v
x
x
v
n
)(z) is the
only nonzero component, but in the interfacial region
˜
nx
Y
nx
nx
v
nx
0
nx
Y
nx
0
is dominant, thereby account-
ing for the failure of NBC to describe the CL motion. There-
fore, away from the CL region the NBC is valid, but in the
interfacial region the NBC clearly fails to describe the CL
motion. We wish to clarify the origin of
nx
Y
and
nx
0
as the
dynamic and static Young stresses, respectively, so that
˜
nx
Y
nx
Y
nx
0
is the uncompensated Young stress. As shown in
the inset to Fig. 4, the integrals across the interface of
nx
Y
(
nx
nx
v
, calculated by subtracting the viscous compo-
nent
(
n
v
x
x
v
n
) from the total tangential stress
nx
) and
nx
0
are equal to
cos
d
and
cos
s
, respectively, at differ-
ent values of z, i.e.,
int
dx
nx
Y
z
cos
d
z
and
int
dx
nx
0
z
cos
s
z
,
where
d
(z) and
s
(z) are, respectively, the dynamic and the
static interfacial angles at z 20. Here
int
dx denotes the
integration across the fluid-fluid interface along x. These re-
sults clearly show the origin of the extra tangential stress in
the interfacial region to be the interfacial uncompensated
Young stress. Thus, the GNBC is given by
v
x
slip
⫽⫺
˜
nx
0
⫽⫺
n
v
x
兴共
0
˜
nx
Y
0
. 3
Here only one component of the viscous stress is nonzero,
due to
v
n
0 at the boundary, and
˜
nx
Y
(0) is the uncom-
pensated Young stress, satisfying
int
˜
nx
Y
0
dx
cos
d
surf
cos
s
surf
,
with
d(s)
surf
being a microscopic dynamic static contact angle
at the fluid-solid boundary. The fact that
˜
nx
Y
(0)0 away
from the CL shows that the GNBC implies NBC for single
phase flows.
Due to the diffuse nature of the BL in the MD simula-
tions, the contact angle
d(s)
surf
cannot be directly measured.
Nevertheless, they are obtainable through extrapolation by
using the integrated interfacial curvature within the BL. That
is, in the sharp boundary limit the force balance in the fluids
is expressed by
x
˜
xx
n
˜
nx
0. Integration in z across the
BL gives
x
0
z
0
dz
˜
xx
z
nx
v
z
0
nx
v
0
˜
nx
Y
z
0
˜
nx
Y
0
0.
4
Integration of Eq. 4 along x] across the fluid-fluid inter-
face then yields
0
z
0
dz
˜
xx
z
int
dx
nx
v
z
0
int
dx
nx
v
0
K
d
K
s
0, 5
where
0
z
0
dz
˜
xx
(z)
is the change of the z-integrated
˜
xx
across the interface, K
d
and K
s
denote the dynamic and the
static z-integrated interfacial curvatures:
K
d
cos
d
z
0
cos
d
surf
,
and
K
s
cos
s
z
0
cos
s
surf
.
Here
0
z
0
dz
˜
xx
(z)
,
nx
v
(z
0
),
d
(z
0
), and
s
(z
0
) are ob-
tainable from MD simulations, K
s
2z
0
cos
s
surf
/H for the
circular static interfaces, while
nx
v
(0)
n
v
x
(0) may be
obtained by extrapolating from the values of tangential vis-
cous stress at z z
0
,2z
0
, and 3z
0
. Therefore, the micro-
scopic dynamic contact angle
d
surf
can be obtained from Eq.
5. In Appendix B 3 we give a more detailed account of the
relationship between the MD measured stresses and the
stress components in the continuum hydrodynamics. The
above extrapolation is based on this correspondence.
We have measured the z-integrated
˜
xx
xx
xx
0
in the
BL. The dominant behavior is a sharp drop across the inter-
face, as shown in Fig. 5 for both the symmetric and asym-
metric cases. The value of
d
surf
obtained is 88° 0.5° for the
symmetric case and 63° 0.5° for the asymmetric case at the
lower boundary, and 64.5° 0.5° at the upper boundary.
These values are noted to be very close to
s
surf
. Yet the
small difference between the dynamic and static micro-
scopic contact angles is essential in accounting for the near-
complete slip in the CL region.
In essence, our results show that in the vicinity of the CL,
the tangential viscous stress
nx
v
as postulated by the NBC
cannot give rise to the near-complete CL slip without taking
into account the tangential Young stress
nx
Y
in combina-
MOLECULAR SCALE CONTACT LINE HYDRODYNAMICS... PHYSICAL REVIEW E 68, 016306 2003
016306-5

Figures
Citations
More filters
Journal ArticleDOI

Wetting and Spreading

TL;DR: In this article, the surface forces that lead to wetting are considered, and the equilibrium surface coverage of a substrate in contact with a drop of liquid is examined, while the hydrodynamics of both wetting and dewetting is influenced by the presence of the three-phase contact line separating "wet" regions from those that are either dry or covered by a microscopic film.
Journal Article

Heterogeneous multiscale methods: A review

TL;DR: This paper gives a systematic introduction to HMM, the heterogeneous multiscale methods, including the fundamental design principles behind the HMM philosophy and the main obstacles that have to be overcome.
Journal ArticleDOI

The heterogeneous multiscale method

TL;DR: The heterogeneous multiscales method (HMM), a general framework for designing multiscale algorithms, is reviewed and emphasis is given to the error analysis that comes naturally with the framework.
Journal ArticleDOI

A phase field model for the mixture of two incompressible fluids and its approximation by a Fourier-spectral method

TL;DR: A phase field model for the mixture of two incompressible fluids is presented in this article, which is based on an energetic variational formulation, consisting of a Navier-Stokes system coupled with a Cahn-Hilliard equation through an extra stress term and the transport term.
Journal ArticleDOI

A multi-phase SPH method for macroscopic and mesoscopic flows

TL;DR: A multi-phase smoothed particle hydrodynamics (SPH) method for both macroscopic and mesoscopic flows is proposed, and a new simple algorithm capable for three or more immiscible phases is developed.
References
More filters
Book

Computer Simulation of Liquids

TL;DR: In this paper, the gear predictor -corrector is used to calculate forces and torques in a non-equilibrium molecular dynamics simulation using Monte Carlo methods. But it is not suitable for the gear prediction problem.
Related Papers (5)
Frequently Asked Questions (12)
Q1. What are the contributions mentioned in the paper "Molecular scale contact line hydrodynamics of immiscible flows" ?

From extensive molecular dynamics simulations on immiscible two-phase flows, the authors find the relative slipping between the fluids and the solid wall everywhere to follow the generalized Navier boundary condition, in which the amount of slipping is proportional to the sum of tangential viscous stress and the uncompensated Young stress. 

In the BL, the wall-fluid interaction means the existence of a dynamic, tangential wall force density g̃x w such that the force balance equation is given by (“•s̃)• x̂1 g̃x w50 inside the BL. 

of increasing the wall thickness in their simulations from two layers of wall molecules to four layers and to infinite layers ~by using the continuum approximation beyond the four layers!. 

In the steady state, the external work done to the system is fully dissipated in the system through convection-diffusion of the composition, slipping at surface, and shear viscosity in the bulk. 

Gxw was obtained from the time average of the total tangential wall force experienced by the fluid molecules in the BL, divided by the bin area in the xy plane; sxx(nx) was obtained from the time averages of the kinetic momentum transfer plus the fluidfluid interaction forces across the constant-x(z) bin surfaces,6-2and vx(z) was measured as the time-averaged velocity component~s! 

It is only for a moving CL that there is a component of the Young stress, which is no longer balanced by the normal stress gradient, and this uncompensated Young stress is precisely the additional component captured by the GNBC but missed by the NBC. 

The rate of that work is positive, given by the integrated local force times the wall velocity, i.e., *dxuG̃x wuV5*dxbuvx slipuV per unit length. 

For the symmetric case, the resulting heat generation rate due to the CL is thus bV2WsL ~for one wall!, where L is the length of the CL and Ws defines the width of the CL region:Ws5 1 VE ~ uvxslipu2v0slip!dx . 

The wall-fluid interaction was also modeled by a LJ potential Uw f , with energy and range parameters ew f51.16e and sw f51.04s , and a dw f for specifying the wetting property of the fluid. 

To further verify that the boundary conditions and the parameter values are local properties and hence applicable to flows with different macroscopic conditions, the authors have varied the wall speed V, the system size H, and the flow geometry to check that the same set of parameters plus the GNBC are valid for reproducing ~a! 

As a first-order approximation, the authors formulate a hydrodynamic model based on the GNBC and the CH free energy functional @21# that has been successful in the calculations of fluid-fluid interfacial phenomena:F@f#5E drF12 K~¹f!21 f ~f!G , ~6! where f5(r22r1)/(r21r1), f (f)52 1 2 rf 21 14 uf 4, and K, r, u are the parameters that can be directly obtained from MD simulations through the interface profile thickness j 5AK/r @22#, the interfacial tension g52A2r2j/3u , and the two homogeneous equilibrium phases given by the condition of ] f /]f50, yielding f656Ar/u (561 in their case!. 

potential U f f 54e@(s/r)122d f f(s/r)6# , where r is the distance between the molecules, e and s are the energy scale and the range of interaction, respectively, and d f f51 for like molecules and d f f521 for molecules of different species.