Sunteți pe pagina 1din 10

PHYSICAL REVIEW E 68, 061603 共2003兲

Spreading dynamics of polymer nanodroplets


David R. Heine, Gary S. Grest, and Edmund B. Webb III
Sandia National Laboratories, Albuquerque, New Mexico 87185, USA
共Received 6 August 2003; published 30 December 2003兲
The spreading of polymer droplets is studied using molecular dynamics simulations. To study the dynamics
of both the precursor foot and the bulk droplet, large hemispherical drops of 200 000 monomers are simulated
using a bead-spring model for polymers of chain length 10, 20, and 40 monomers per chain. We compare
spreading on flat and atomistic surfaces, chain length effects, and different applications of the Langevin and
dissipative particle dynamics thermostats. We find diffusive behavior for the precursor foot and good agree-
ment with the molecular kinetic model of droplet spreading using both flat and atomistic surfaces. Despite the
large system size and long simulation time relative to previous simulations, we find that even larger systems are
required to observe hydrodynamic behavior in the hemispherical spreading droplet.

DOI: 10.1103/PhysRevE.68.061603 PACS number共s兲: 68.47.Pe

I. INTRODUCTION vaporization and condensation, subsequent simulations used


short bead-spring chain molecules since they have a very low
The spreading of liquid droplets on a surface is an impor- vapor pressure. In most cases, the simulations reproduced the
tant issue for several industries including adhesion, lubrica- experimentally observed R⬃t 1/2 scaling of the contact radius
tion, coating, and printing. Emerging nanotechnology in ar- of the precursor foot on both atomistic 关11,33–35兴 and flat
eas such as lithography and microfluidics has made the issue 关36,37兴 surfaces, though logarithmic scaling has also been
of droplet spreading on small length scales even more rel- observed 关35兴. It is believed that this difference is due to the
evant. Experiments on droplet spreading have revealed sev- corrugation of the substrate, producing t 1/2 scaling for a suf-
eral phenomena involved in the spreading process, some of ficiently small lattice dimension and a logarithmic scaling for
which occur on the atomic level and others that become rel- large, i.e., strong corrugation 关38兴. Milchev and Binder 关39兴
evant at mesoscopic length scales 关1–15兴. These include the have studied wetting using Monte Carlo simulations on a flat
spreading of a precursor foot ahead of the droplet 关3兴, ter- substrate which suggest Tanner’s spreading law for the
raced spreading of monomolecular layers 关4,16,17兴, and vis- growth dynamics of the droplet holds on the nanoscopic
cous losses due to rolling motion 关1,18兴. scale. Other comparisons to theoretical models have strongly
Several models have been proposed to describe the spon- supported the molecular kinetic theory of wetting 关40– 44兴,
taneous spreading of liquid droplets on a surface. These probably due to the relatively small droplet sizes and short
models can be classified as molecular kinetic models, con- simulation times employed.
tinuum hydrodynamic models, or combined models. The mo- In this paper, we present results from extensive molecular
lecular kinetic theory of Eyring 关19兴 has been applied to the dynamics simulations of coarse-grained models of polymer
kinetics of wetting by Blake and Haynes 关20,21兴 as well as droplets wetting a surface. Although most recent simulations
Cherry and Holmes 关22兴. This theory treats the surface ad- of droplet spreading use droplets containing 20 000 to 32 000
sorption of liquid molecules as the dominant factor in the monomers 关11,33–35,40,42兴, we consider drops composed
spreading of a droplet. The hydrodynamic theory 关1,2,23,24兴 of up to 200 000 monomers to simultaneously study the pre-
focuses on the energy dissipation due to viscous flow in the cursor foot and bulk regions for long times. We compare
droplet. It has been claimed that hydrodynamic dissipation is simulations performed using both a flat surface and an ato-
dominant for small contact angles and nonhydrodynamic dis- mistic substrate to determine if the computationally expen-
sipation is dominant for relatively large contact angles 关25兴. sive atomistic substrate is required to obtain correct spread-
Since both mechanisms are present in spreading droplets, ing dynamics. We also evaluate different implementations of
several groups have proposed combined theories 关3,8,23,26 – the Langevin and dissipative particle dynamics 共DPD兲 ther-
28兴. Experimental results for the spreading of poly mostats for efficiency and realism in preserving hydrody-
dimethylsiloxane 共PDMS兲 drops on bare silicon wafers have namic effects. Also, the difference in using a spherical drop-
shown good agreement with one combined model 关10兴. let as the starting configuration as opposed to a
The study of droplet spreading using molecular dynamics hemispherical droplet is discussed. We find that the method
simulation has been hindered due to computational limita- which captures all of the physics of the spreading drop in the
tions restricting simulations to small droplet sizes and short most computationally efficient manner is to simulate large
times. Molecular dynamics simulations were first used to drops on flat substrates with a coupling to the thermostat
study the spreading of monomer and dimer liquids 关29–32兴. which falls off exponentially with distance from the substrate
However, the spreading of monomer and dimer droplets are 关45兴. For atomistic substrates, we find coupling only the sub-
clearly influenced by the volatility of the small molecules, strate monomers to the Langevin thermostat significantly
allowing them to vaporize and condense independent of the more efficient than coupling the DPD thermostat to all
dynamics of the droplet. To separate the spreading from the monomers.

1063-651X/2003/68共6兲/061603共10兲/$20.00 68 061603-1 ©2003 The American Physical Society


HEINE, GREST, AND WEBB III PHYSICAL REVIEW E 68, 061603 共2003兲

The paper is organized as follows. Section II describes the reaches the edge of the substrate and interacts with the peri-
details of the molecular dynamics simulations and the appli- odic image of the precursor foot. Although this can be related
cation of the thermostats. Section III presents the results for to the spreading of an array of nanodroplets, such as in mi-
the time dependence of the contact radius. The contact angle crocontact printing, we do not include any data for the pre-
data are fit to models of droplet spreading in Sec. IV and cursor foot once it reaches the periodic image. The droplets
conclusions are presented in Sec. V. consist of 10 000–200 000 monomers for nonwetting drop-
lets and ⬃200 000 monomers for wetting droplets. All simu-
II. SIMULATION DETAILS lations are run at a temperature of T⫽1.0␧/k B .
For the flat surface, the interaction between the monomers
A. System in the droplet and the surface is modeled by an integrated LJ
We perform molecular dynamics 共MD兲 simulations using potential

再 冋 冉 冊 冉 冊册
a coarse-grained model for the polymer chains in which the
polymer is represented by spherical beads of mass m at- 2␲␧w 2 ␴ 9
␴ 3
⫺ z⭐z c
tached by springs. We use a cutoff Lennard-Jones 共LJ兲 po- U LJ 共 z 兲 ⫽
wall 3 15 z z 共3兲
tential to describe the interaction between all monomers. The 0 z⬎z c ,
LJ potential is given by

再 冋冉 冊 冉 冊 册
␴ ␣␤ 12
␴ ␣␤ 6 with z c ⫽2.2␴ .
␣␤
4␧ ␣␤ ⫺ r⭐r c The equations of motion are integrated using a velocity-
U LJ 共 R 兲⫽ r r 共1兲 Verlet algorithm. We use a time step of ⌬t⫽0.009␶ , where
0 r⬎r c , ␶ ⫽ ␴ (m/␧) 1/2. The simulations are performed using the
LAMMPS code 关47兴 on 36 to 100 Dec Alpha processors of
where ␧ ␣␤ and ␴ ␣␤ are the LJ units of energy and length and Sandia’s CPlant cluster. Simulating 1⫻106 steps for a wet-
the cutoff is set to r c ⫽2.5␴ ␣␤ . We denote the polymer ting drop of 200 000 monomers on the medium atomistic
monomers as type 1 and substrate monomers as type 2. The substrate takes between 90 and 250 h on 64 processors, de-
monomer-monomer interaction, ␧ 11⫽␧, is used as the refer- pending on the thermostat.
ence and all monomers have the same diameter ␴ ␣␤ ⫽ ␴ . For
bonded monomers, we apply an additional potential where B. Thermostats
each bond is described by the finite extensible nonlinear elas-
tic 共FENE兲 potential 关46兴 The choice of thermostat employed can greatly affect the


droplet spreading dynamics, so we compare simulations that

U FENE 共 r 兲 ⫽
⫺k 2
2
R 0 ln 1⫺
R
r
0
冋 冉 冊册 2
r⭐R 0
共2兲
use the Langevin 关48兴 and DPD 关49,50兴 thermostats. The
purpose is to find an approach that is both computationally
efficient and provides a realistic representation of the transfer
⬁ r⬎R 0 , of energy in the spreading droplet.
The Langevin thermostat simulates a heat bath by adding
with k⫽30␧ and R 0 ⫽1.5␴ . Gaussian white noise and friction terms to the equation of
Droplets consisting of chains of length N⫽10, 20, or 40 motion,
monomers per chain are created by first equilibrating a melt
of the polymer and then removing molecules whose centers m i r̈i ⫽⫺⌬U i ⫺m i ␥ L ˙
ri ⫹Wi 共 t 兲 , 共4兲
are outside of a hemisphere of a given radius. The droplet is
then placed on either an atomistic substrate or a flat sub- where ␥ L is the friction parameter for the Langevin thermo-
strate. stat, ⫺⌬U i is the force acting on monomer i due to the
The atomistic substrate is composed of LJ particles form- potentials defined above, and Wi (t) is a Gaussian white
ing four layers of the (111) surface of an fcc lattice where noise term such that
the bottom layer is frozen and the top three layers maintain
their structure through a strong LJ interaction, ␧ 22⫽5␧. The 具 Wi 共 t 兲 •W j 共 t ⬘ 兲 典 ⫽6k B Tm i ␥ L ␦ i j ␦ 共 t⫺t ⬘ 兲 . 共5兲
masses of the substrate monomers are set to m 2 ⫽2m 1
⫽2m. For nonwetting droplets, each layer of the substrate The Langevin thermostat can either be coupled to all mono-
contains 12 000 monomers and the dimensions of the sub- mers in the system or just to those in the substrate. The
strate are 110.0␴ ⫻115.4␴ . For the wetting droplets, we advantage of the latter is that the long-range hydrodynamic
study two substrates, containing either 49 200 or 99 960 interactions are preserved in the droplet, whereas coupling
monomers per layer. The dimensions of the substrates are all monomers to the Langevin thermostat screens the hydro-
231.2␴ ⫻231.0␴ and 330.8␴ ⫻331.4␴ , respectively. We re- dynamic interactions. Both approaches are applied in the
fer to these as the small, medium, and large substrates. The simulations to test the various models for droplet spreading
large substrates are necessary because the finite size of the discussed below in Sec. IV. The damping constant is chosen
atomistic substrates require the use of periodic boundary to be ␥ L ⫽0.1␶ ⫺1 in most cases, which is much smaller than
conditions at their edges whereas the flat surface can extend that arising from collisions between monomers.
indefinitely in the x and y directions. For the atomistic sub- Our next approach is to apply the thermostat from the
strate, during the course of the simulation, the precursor foot DPD simulation method. The DPD technique includes a dis-

061603-2
SPREADING DYNAMICS OF POLYMER NANODROPLETS PHYSICAL REVIEW E 68, 061603 共2003兲

sipative force term in the equations of motion along with


random forces. The equation of motion for the DPD thermo-
stat is

m i r̈i ⫽ 兺
j⫽i
共 ⫺⌬U i j ⫹FD
i j ⫹Fi j 兲 .
R
共6兲

In Eq. 共6兲, FD R
i j and Fi j are the dissipative and random terms
given by

i j ⫽⫺m i ␥ D PD w 共 r i j 兲关 r̂i j • 共 ṙi ⫺ṙ j 兲兴 ,r̂i j


FD 共7兲
2

FRi j ⫽m i ␴ D PD w 共 r i j 兲 ␨ i j r̂i j , 共8兲


FIG. 1. Profile of the N⫽10 polymer droplet spreading on the
where ␥ D PD is the DPD friction parameter, ␴ D 2
PD
atomistic substrate at three different times using the Langevin ther-
⫽2k B T ␥ D PD , ␨ i j is a Gaussian noise term with mostat applied only to the substrate monomers with ␥ L ⫽0.1␶ ⫺1
and ␧ 12⫽1.5␧.
具 ␨ i j (t) ␨ kl (t ⬘ ) 典 ⫽( ␦ ik ␦ jl ⫹ ␦ il ␦ jk ) ␦ (t⫺t ⬘ ), ri j ⫽ri ⫺r j , r i j
⫽ 兩 ri j 兩 , and r̂i j ⫽ri j /r i j . The weight function w(r i j ) is de-
fined as III. SIMULATION RESULTS


A droplet containing about 200 000 monomers for a wet-
共 1⫺r i j /r c⬘ 兲 r i j ⬍r c⬘
w共 ri j 兲⫽ 共9兲 ting droplet is large enough to allow us to simultaneously
0 r i j ⭓r c⬘ . study the bulk and precursor foot regions. This can be seen in
the profile views for chain length N⫽10 in Figs. 1 and 2,
We take r c⬘ ⫽r c ⫽2.5␴ . The advantage of this thermostating which show the foot extending beyond the bulk region for
technique is that the momentum is conserved locally and wetting droplets on an atomistic substrate and a flat surface,
long-range hydrodynamic interactions are preserved even in respectively. The lower half of the exposed surface of the
the case where all monomers are coupled to the thermostat. droplet acts as the source of monomers for the foot. The rest
All simulations with the DPD thermostat use ␥ D PD of the droplet tends to spread such that the final position of a
⫽0.1␶ ⫺1 , so the dissipation from the thermostat is much less monomer has very little dependence on its initial height from
than from monomer collisions as seen in Sec. IV. Simula- the surface. Note that the third frame of Fig. 1 shows the
tions that use DPD couple the thermostat to all atoms in the thickness of the foot increasing after it reaches the periodic
system. image. The same behavior is seen when periodic boundaries
In the case of the flat substrate, we study several methods are applied to the flat surface, so this is not an effect of the
to thermostat the system. In the first, we simply couple the corrugation of the substrate. The local ordering in the z di-
Langevin or DPD thermostat to all monomers. However, this rection that is typically seen in the first two or three atomic
is somewhat unphysical since monomers near the substrate layers above the substrate is present, but not visible on the
are expected to have a stronger damping than those in the scale of these figures.
bulk of the droplet. In the case of the Langevin thermostat, To characterize the spreading dynamics of these droplets,
this coupling of all monomers also means that the hydrody- we extract the instantaneous contact radius and contact angle
namic interactions are screened. In addition, chains which every 10 000⌬t –40 000⌬t. The contact radius is calculated
separate from the droplet move across the substrate very rap- by defining a two-dimensional radial distribution function,
idly, particularly for the DPD thermostat. For this reason, we
did not further pursue the DPD thermostat on the flat sub-
strate. To overcome these difficulties, we follow the ap-
proach of Braun and Peyrard 关45兴 and add an external Lange-
vin coupling with a damping rate that decreases
exponentially away from the substrate. We choose the form

␥ L 共 z 兲 ⫽ ␥ Ls exp共 ␴ ⫺z 兲 , 共10兲

where ␥ Ls is the surface Langevin coupling and z is the dis-


tance from the substrate. We choose values of ␥ Ls ⫽1.0␶ ⫺1 ,
3.0␶ ⫺1 , and 10.0␶ ⫺1 . There is no obvious a priori way to
define the appropriate value of ␥ Ls . However, one way is to
choose ␥ Ls so that the diffusion constant of the precursor foot
is comparable for the flat and atomic substrates for compa- FIG. 2. Profile of the N⫽10 polymer droplet spreading on the
rable departures from the wetting/nonwetting transition 共see flat surface at three different times using the surface Langevin ther-
Fig. 4 below兲. mostat. ␥ Ls ⫽10.0␶ ⫺1 , ␧ w ⫽2.0␧.

061603-3
HEINE, GREST, AND WEBB III PHYSICAL REVIEW E 68, 061603 共2003兲

FIG. 3. Contact angle of nonwetting droplets of N⫽10 poly-


mers on an atomistic substrate starting from a hemispherical droplet
with ␥ D PD ⫽0.1␶ ⫺1 . The droplet contains 100 000 monomers.
FIG. 4. Equilibrium contact angle as a function of polymer-
g(r)⫽ ␳ (r)/ ␳ , based on every monomer within 1.5␴ of the surface interaction strength showing the transition from nonwetting
surface. The local density at a distance r from the center of to wetting. 共a兲 Droplets consisting of 10 000 (䊊), 22 000 (䊐),
mass of the droplet is 49 000 (〫), and 100 000 monomers (䉭) on a flat surface. The line
is a guide for the eye. 共b兲 N⫽10 polymer droplets containing
N共 r 兲 100 000 monomers on an atomistic substrate (䊊) and a flat surface
␳共 r 兲⫽ , 共11兲 (䊐), and N⫽40 polymer droplets containing 200 000 monomers
2 ␲ r⌬r on a flat surface (〫).

where N(r) is the number of monomers at a distance be-


tween r and r⫹⌬r from the center of mass and ␳ is the The time dependence of the contact radius of the precur-
integral of ␳ (r) over the entire surface. The contact radius is sor foot and bulk region is shown in Fig. 5 for wetting drop-
defined as the distance r at which g(r)⫽0.98. This approach lets on an atomistic substrate for three chain lengths. The t 1/2
provides a robust measure of the radius at any point during behavior is evident for the precursor foot at all chain sizes,
the spreading simulation. The same calculation is used to while the kinetics of the main droplet is clearly significantly
obtain the droplet radius for ten slices of the droplet at in- slower. The N⫽10 data shown in Fig. 5 is taken from simu-
cremental heights every 1.5␴ from the surface. A line is fit to lations on both the large and medium substrates whereas the
the resulting points and the instantaneous contact angle is N⫽20 and N⫽40 simulations are on the medium substrate.
determined from the slope of the line. For simulations that
exhibit a precursor foot, the monomers within 4.5␴ of the
surface are ignored in the contact angle calculation.
The nonwetting droplets reach their equilibrium configu-
rations fairly rapidly, as shown by the contact angle data in
Fig. 3. The equilibrium contact angles measured as a func-
tion of polymer-surface interaction strength are shown in
Fig. 4. The variation in equilibrium contact angle for differ-
ent droplet sizes is shown in Fig. 4共a兲. Finite size effects are
evident for drops containing 10 000 and 22 000 monomers.
They are almost negligible for drops of 49 000 monomers as
compared to droplets of 100 000 monomers. Away from the
transition region, the smaller droplets have a smaller equilib-
rium contact angle ␪ 0 due mostly to the decrease in the
liquid/vapor surface tension ␥ with size 关31,51–53兴. The
equilibrium contact angles for different surfaces and chain
lengths are shown in Fig. 4共b兲. The transition from nonwet- FIG. 5. Time dependence of the contact radius of the precursor
foot and bulk droplet for wetting droplets on an atomistic substrate
ting to wetting occurs near ␧ 12
c
⯝1.05␧ for N⫽10 droplets on
at three different chain lengths starting from a hemisphere with a
the atomistic substrate and ␧ wc ⯝1.75␧ for N⫽10 droplets on contact angle of ⬇90°. The Langevin thermostat, ␥ L ⫽0.1␶ ⫺1 , is
a flat surface. The contact angles for N⫽40 are included to applied only to the substrate monomers and ␧ 12⫽1.5␧. Results for
show that the transition is shifted to higher ␧ w s for larger N⫽10 are for both the medium and large atomistic substrates, while
chain lengths due to the increase in the liquid vapor surface those for N⫽20 and 40 are for the medium atomistic substrate. The
tension. For most of the wetting simulations, we use ␧ 12 inset shows the contact radius for N⫽10 starting with a spherical
⫽1.5␧ for the atomistic substrate and ␧ w ⫽2.0␧ for the flat droplet 共solid line兲 compared to a hemispherical droplet 共dotted兲.
substrate, both within the wetting regime for the range of Results for the hemisphere in the inset have been shifted downward
chain lengths studied. to easily compare the late time behavior.

061603-4
SPREADING DYNAMICS OF POLYMER NANODROPLETS PHYSICAL REVIEW E 68, 061603 共2003兲

The contact radius of the bulk droplet increases steadily for


all three chain lengths on the medium substrate. However,
the run on the large substrate shows a slowing down and
eventual contraction of the bulk contact radius as the foot
continues outward, depleting the supply of material in the
bulk faster than the drop can transfer material downward.
This suggests that for our largest substrate, the drop size
must be even larger to be able to study both the precursor
foot and bulk droplet in the same simulation.
The inset in Fig. 5 shows the spreading of a spherical N
⫽10 droplet compared to an initial hemisphere. The sphere
is placed just above the substrate with zero initial velocity to
avoid any effect due to impact velocity. The difficulty in
measuring the spreading rate for this case is evident as it
takes roughly 1200␶ for the sphere to adopt a hemispherical
shape, 1600␶ for the spreading rate of the foot to match that FIG. 6. Effect of thermostat on contact radius of precursor foot
and bulk region for wetting droplets of N⫽10 polymers on an ato-
of the hemisphere, and 5000␶ for the spreading rate of the
mistic substrate for ␧ 12⫽1.5␧. The thermostats applied are DPD
bulk to match that of the hemisphere. 共The hemisphere data
共solid line兲, Langevin on all monomers 共dotted兲, and Langevin on
for the foot and bulk regions are shifted downward to easily
only substrate monomers 共dashed兲. ␥ D PD ⫽ ␥ L ⫽0.1␶ ⫺1 .
compare the spreading rates.兲
Voué et al. 关6,11兴 found both experimentally for PDMS
Langevin thermostat is applied either to all monomers
droplets and in numerical computer simulations that the dif-
fusion constant of the precursor foot varies nonmonotoni- 共curves labeled with ␥ L ) or with the surface Langevin cou-
cally with increasing coupling to the substrate. At first, in- pling 共curves labeled with ␥ Ls ). The value of ␥ L clearly has a
creasing the coupling to the substrate increases the driving strong influence on the spreading rate. ␥ Ls ⫽3.0␶ ⫺1 gives a
force and the fluid spreads on the substrate more rapidly. diffusion constant comparable to the atomistic substrate with
However, further increase in the strength of the fluid sub- ␥ L ⫽0.1␶ ⫺1 . The chain length dependence of the contact ra-
strate coupling, while increasing the driving force, also in- dius is shown in Fig. 8. Again, the t 1/2 behavior is evident in
crease the friction of the fluid monomers with the substrate, the foot region but not the bulk region. The chain length
resulting in a decrease in the diffusion constant. From the dependence on the flat surface is similar to the atomistic
time dependence of R(t), the diffusion constant D f for the substrate, showing a moderate decrease in spreading rate for
foot can be determined from longer polymers.

具 关 R 共 t 兲 ⫺R 共 0 兲兴 2 典 ⫽4D f t. 共12兲

The resulting diffusion constants are D f ⫽0.34␴ 2 / ␶ ,


0.30␴ 2 / ␶ , and 0.23␴ 2 / ␶ for N⫽10, 20, and 40, respectively
for ␧ 12⫽1.5␧, indicating a very weak dependence on chain
length, at least for these unentangled chains. This weak de-
pendence on N is probably partially due to the fact that ␧ 12
⫽1.5␧ is closer to the nonwetting transition for longer chain
lengths than for shorter chain lengths. Increasing ␧ 12 to 2.0
and 5.0␧ for N⫽10, we find D f ⫽0.16 and 0.029␴ 2 / ␶ , re-
spectively, thus the droplets are in the high friction regime
for these values of fluid substrate coupling.
Figure 6 shows the time dependence of the contact radius
for wetting droplets on an atomistic substrate using different
thermostating techniques. These results show that there is
essentially no difference in the spreading rate between the
DPD thermostat applied to all monomers and the Langevin
thermostat applied only to the substrate. We can see that
applying the Langevin thermostat to all monomers slightly
decreases the spreading rate as the viscous heating is re-
moved from the system, though the resulting loss of hydro-
dynamic flow, at least for the droplet size studied here, has
no significant impact. FIG. 7. Effect of thermostat on contact radius of 共a兲 precursor
For wetting droplets on a flat surface, the thermostat de- foot and 共b兲 bulk region for wetting droplets of N⫽10 polymers on
pendence of the contact radius is shown in Fig. 7. Here, the a flat surface with ␧ w ⫽2.0␧.

061603-5
HEINE, GREST, AND WEBB III PHYSICAL REVIEW E 68, 061603 共2003兲

dR
dt
⫽⫺
3V
␲ 冉 冊 1/3
共 1⫺cos ␪ 兲 2
共 2⫺3 cos ␪ ⫹cos3 ␪ 兲 4/3
d␪
dt
. 共15兲

Combining Eqs. 共14兲 and 共15兲 gives an expression for the


time dependence of the contact angle,

d␪
dt
⫽⫺

3V 冉 冊 1/3
⍀共 ␪ 兲

␨0
共 cos ␪ 0 ⫺cos ␪ 兲 , 共16兲

where

共 2⫺3 cos ␪ ⫹cos3 ␪ 兲 4/3


FIG. 8. Chain length dependence of the contact radius of the ⍀共 ␪ 兲⫽ 共17兲
precursor foot and bulk droplet for wetting droplets on a flat surface 共 1⫺cos ␪ 兲 2
with ␧ w ⫽2.0␧. The surface Langevin thermostat is applied with
␥ Ls ⫽10.0␶ ⫺1 . and ␨ 0 is the friction coefficient defined as ␨ 0
⫽⌬nk B T/K␭, which has units of viscosity.
IV. MODELS OF DROPLET SPREADING DYNAMICS The hydrodynamic model 关24兴 describes the flow pattern
that forms in the bulk of the droplet as material is transferred
A. Overview of models
to the advancing contact line. This model can be obtained by
The dynamics of droplet spreading are controlled by the solving the equations of motion and continuity for the drop-
driving force 共the difference in surface tension ␥ at each let described as a cylindrical disk 关54兴 instead of a spherical
interface兲 and by the energy dissipation. The total energy cap. Neglecting the flow perpendicular to the surface and
dissipation can be represented by a sum of three different balancing the radial shear stress at the top of the cylinder
components, T(⌺̇ w ⫹⌺̇ f ⫹⌺̇ l ) 关3兴. The first term T⌺̇ w repre- with the effective radial surface tension, the velocity of the
sents energy dissipation due to the hydrodynamic flow in the contact line is written as
bulk of the droplet as more material is transferred to the
surface. T⌺̇ f relates to the viscous dissipation in the precur- dR 4␥V3 ␥␤ V
⫽ 3 9⫺ , 共18兲
sor foot present in cases of complete wetting. The third term dt ␲ ␩ R 2␲␩R3
T⌺̇ l refers to the dissipation in the vicinity of the contact line
due to the adsorption and desorption of liquid molecules to where V is the droplet volume, ␩ is the viscosity of the
the solid surface. Here, we compare models that incorporate liquid, and ␤ ⫽1⫺cos ␪0. Equation 共18兲 is in agreement with
one or more of these dissipation mechanisms to our simula- Tanner’s spreading law 关2兴 for completely wetting systems
tion results. ( ␪ 0 ⫽0) and for nonwetting systems with small equilibrium
The molecular kinetic theory of liquids developed by Ey- contact angles, giving R⬃t 1/10 at long times. Instead of di-
ring and co-workers 关19兴 has been applied to droplet spread- rectly combining Eqs. 共15兲 and 共18兲, we apply the approach
ing by Blake and Haynes 关21兴. It focuses on the adsorption of de Ruijter et al. 关8,10兴 in order to make a direct compari-
of liquid molecules to the surface as the dominant factor in son with the combined model presented below. Using the
energy dissipation. In this theory, the liquid molecules jump same cylindrical disk model, they neglect the flow perpen-
between surface sites separated by a distance ␭ with a fre- dicular to the surface and specify that the velocity at the
quency K. The velocity of the contact line is related to the upper edge of the cylinder is the actual droplet spreading rate
contact angle ␪ by dR/dt. With this approach, they find that the hydrodynamic

冋冉 冊 册
dissipation term can be written as
dR ␥
dt
⫽2K␭ sinh
2⌬nk B T
共 cos ␪ 0 ⫺cos ␪ 兲 , 共13兲
T 兺w ⫽6 ␲ R 共 t 兲 ␩ ␾ 关 ␪ 共 t 兲兴 冉 冊
dR
dt
2
ln关 R 共 t 兲 /a 兴 , 共19兲
where ␥ is the surface tension of the liquid/vapor interface,
⌬n is the density of sites on the solid surface, and ␪ 0 is the where ␾ ( ␪ ) is a geometric factor defined as
equilibrium contact angle. For sufficiently low velocities, the
equation can be written in its linearized form sin3 ␪
␾共 ␪ 兲⫽ 共20兲
2⫺3 cos ␪ ⫹cos3 ␪
dR K␭ ␥
⫽ 共 cos ␪ 0 ⫺cos ␪ 兲 . 共14兲
dt ⌬nk B T and a is an adjustable parameter that represents the radius of
the core region of the droplet, where the radial flow is neg-
Assuming the droplet maintains constant volume and the ligible. For the hydrodynamic model, they obtain

冉 冊
shape of a spherical cap, the velocity of the contact line can
be expressed in terms of the time dependence of the contact d␪ ␲ 1/3
␥ 共 cos ␪ 0 ⫺cos ␪ 兲
⫽⫺ ⍀共 ␪ 兲 . 共21兲
angle purely from geometric arguments giving dt 3V 6 ␩ ␾ 共 ␪ 兲 ln关 R/a 兴

061603-6
SPREADING DYNAMICS OF POLYMER NANODROPLETS PHYSICAL REVIEW E 68, 061603 共2003兲

TABLE I. Bulk properties of bead-spring chains obtained from MD simulation for T⫽␧/k B , P⯝0.

N ␳ ( ␴ ⫺3 ) ␥ (␧/ ␴ 2 ) ␩ (m/ ␶ ␴ ) 103 D( ␴ 2 / ␶ ) ␨ R ( ␶ ⫺1 )

10 0.8691 0.85⫾0.02 11.1⫾0.4 6.17⫾0.06 16.2


20 0.8803 0.92⫾0.02 17.4⫾0.7 3.04⫾0.03 16.4
40 0.8856 0.95⫾0.02 41.7⫾1.4 1.23⫾0.01 20.4

Both types of dissipation are present in the spreading functions are averaged to improve statistical uncertainty.
droplet. The hydrodynamic mechanism is expected to domi- From this, the viscosity can be calculated using 关57兴
nate at low velocities and small contact angles while the
kinetic mechanism is expected to dominate at high velocities
and large contact angles 关25兴. We include in our comparison ␩⫽
V
k BT
冕0

dt 具 ␴ ␣␤ 共 t 兲 ␴ ␣␤ 共 0 兲 典 . 共25兲
a model developed by de Ruijter et al. 关8,10兴 containing both
kinetic and hydrodynamic terms. In this model, the velocity The results for ␩ are summarized in Table I.
of the contact line is written as Estimates of the friction coefficients ␨ R obtained from the
dR ␥ 关 cos ␪ 0 ⫺cos ␪ 兴 melt simulations, are included in Table I. The diffusion con-
⫽ . 共22兲 stant D is determined from the mean square displacement of
dt ␨ 0 ⫹6 ␩ ␾ 共 ␪ 兲 ln关 R/a 兴 the middle monomers of each chain and using the Rouse
model one can extract ␨ R from D⫽k B T/mN ␨ R 关59兴.
Combining this with Eq. 共15兲 gives
With the above values for the surface tension and viscos-
d␪
dt
⫽⫺

3V冉 冊 1/3
⍀共 ␪ 兲
␥ 共 cos ␪ 0 ⫺cos ␪ 兲
␨ 0 ⫹6 ␩ ␾ 共 ␪ 兲 ln关 R/a 兴
. 共23兲
ity, the simulation data are fit to each of the models described
above. The fit is performed by taking initial guess values for
the independent parameters and integrating the expression
for d ␪ /dt defined in one of the equations 共16兲, 共21兲, or 共23兲.
B. Analysis of models The integration uses the fourth-order Runge-Kutta method to
Fitting simulation data to the models described above re- generate a set of data, ␪ calc (t). The parameters are varied
quire both the liquid/vapor surface tension and the bulk vis- using the downhill simplex method 关58兴 until the difference
cosity of the polymer. The surface tension ␥ is obtained by between the model and simulation data, 兩 ␪ calc (t)
first constructing a slab of the polymer melt containing ⫺ ␪ (t) 兩 / ␪ (t), is minimized.
10 000 chains of N⫽10, 5000 chains of N⫽20, or 5000 The kinetic, hydrodynamic, and combined models are fit
chains of N⫽40 centered in the simulation box such that to the contact angles of droplets spreading on a flat surface in
there are two surfaces perpendicular to the z direction. The Fig. 9. The Langevin thermostat is applied either to all
simulations are run at temperature T⫽␧/k B . The simulations monomers or only to those near the surface. We find that
are run until the two liquid/vapor interfaces are equilibrated, both the kinetic and combined models fit the data well de-
as determined by the density profiles across the interfaces. spite the fact that they predict that the friction coefficient ␨ 0
From the equilibrium values of the pressure, parallel and is larger in the combined model than in the kinetic model.
perpendicular to the interfaces, ␥ can easily be determined The hydrodynamic model produces a very poor fit to each
from Ref. 关55兴 data set as shown in Fig. 9共b兲. The best fit parameters for
these models applied to data for wetting droplets on a flat

␥⫽
1
2
冕0
Lz
关 p⬜ 共 z 兲 ⫺p 储 共 z 兲兴 dz. 共24兲
surface are shown in Table II. We also fit the hydrodynamic
model excluding the first 10 000 ␶ with the understanding
that hydrodynamic behavior should not be present for these
The values for the surface tension are summarized in Table I. short time scales. We found that the model still underesti-
These values can be compared to ␥ ⫽0.08␧/ ␴ 2 for a system mated the angle at early times and overestimated the angle at
of monomers 关56兴. late times. The error reported for each model is calculated as
The viscosity is computed from the equilibrium fluctua- N
1 兩 ␪ calc 共 t 兲 ⫺ ␪ 共 t 兲 兩 2
tions of the off-diagonal components of the pressure tensor
关57兴. The pressure tensors are recorded from simulations of
␹ ⫽
2
N 兺
i⫽1 ␪共 t 兲
, 共26兲
systems containing melts of 500 chains of the N⫽10 poly-
mer, 250 chains of the N⫽20 polymer, and 500 chains of the where N is the number of data points in each set of data.
N⫽40 polymer at T⫽␧/k B with the bulk pressure P⯝0 Figure 10 shows the kinetic, hydrodynamic, and com-
without tail corrections. We leave out tail corrections to the bined model fits to the contact angle data for wetting droplets
pressure in order to match the system of the spreading drop- on the medium substrate. Again, the hydrodynamic model
let. These simulations are run at a time step ⌬t⫽0.006 for up gives a significantly worse fit to the data. The best fit param-
to 25 000␶ . The autocorrelation function of each off-diagonal eters for these models for wetting droplets on an atomistic
component of the stress tensor is calculated using the nu- substrate are shown in Table III. The kinetic and combined
merical recipes routine CORREL 关58兴. The autocorrelation models give friction coefficients that are generally larger

061603-7
HEINE, GREST, AND WEBB III PHYSICAL REVIEW E 68, 061603 共2003兲

FIG. 10. Fits to contact angle data 共symbols兲 of 共a兲 kinetic, 共b兲
hydrodynamic, and 共c兲 combined models for wetting droplets on the
medium atomistic substrate with ␧ 12⫽1.5␧. The Langevin thermo-
FIG. 9. Fits to contact angle data 共symbols兲 of 共a兲 kinetic, 共b兲 stat is applied to all monomers or to just substrate monomers. ␥ L
hydrodynamic and 共c兲 combined models for wetting droplets on a ⫽0.1␶ ⫺1 for all cases except DPD where ␥ D PD ⫽0.1␶ ⫺1 . The data
flat surface with ␧ w ⫽2.0␧. The chain length is N⫽10 unless oth- sets are shifted by 10° increments 共except for N⫽10 DPD兲 for
erwise specified. The Langevin thermostat is applied to all mono- clarity.
mers ( ␥ L ) or just monomers near the surface ( ␥ Ls ). The data sets are
shifted by 10o increments 共except for ␥ Ls ⫽1.0␶ ⫺1 ) for clarity. greater in size than those previously studied. We find this to
be necessary to adequately model the behavior of the precur-
than the bulk viscosity for the range of coupling parameters
sor foot and the bulk material simultaneously. Starting from a
used here. We find that, in contrast with previous work by de
hemispherical droplet, we find that the precursor foot forms
Ruijter et al. 关8,10兴, the combined model predicts a larger
immediately and spreads diffusively for each system where
friction coefficient than the kinetic model. Also, the hydro-
the surface interaction strength is above the wetting/
dynamic and combined models give a value of a that is of
nonwetting transition. The bulk region of the droplet spreads
the order of the radius of the droplet, indicating that hydro-
at a significantly slower rate, but the data are too imprecise
dynamic flow is not a dominant feature of the spreading of
to distinguish between, for example, a t 1/7 and a t 1/10 scaling.
these droplets, at least for the time scales accessible to simu-
We perform spreading simulations on both an atomisti-
lation.
cally realistic substrate and a perfectly flat surface. The simu-
V. CONCLUSIONS lations using a flat surface exhibit the same behavior as the
realistic substrate and greatly improve the computational ef-
In this study, we perform molecular dynamics simulations ficiency since the number of monomers on the realistic sub-
of polymer droplets that are roughly an order of magnitude strate is typically several times greater than the number of
TABLE II. Model parameters and error estimates resulting from fits to contact angle data from simulations of wetting droplets on a flat
surface. Values for ␥ L and ␥ Ls are listed in the first column. ␧ w ⫽2.0␧.

Kinetic Combined

Thermostat N 冉
␨0 units of
m
␶␴ 冊 Hydrodynamic
a(units of ␴ ) 冉
␨0 units of
m
␶␴ 冊 a(units of ␴ ) ␹ kin
2
␹ hydro
2
␹ comb
2

␥ L ⫽1.0␶ ⫺1 10 9.55 44.40 35.41 71.23 0.0022 0.047 0.0039


␥ Ls ⫽3.0␶ ⫺1 10 25.97 42.29 57.27 71.55 0.00028 0.025 0.0011
␥ Ls ⫽10.0␶ ⫺1 10 56.30 38.14 89.98 83.80 0.00024 0.018 0.00028
␥ Ls ⫽10.0␶ ⫺1 20 81.37 38.83 137.99 84.77 0.00015 0.015 0.00023
␥ Ls ⫽10.0␶ ⫺1 40 101.29 38.63 200.45 86.49 0.00022 0.024 0.00036

061603-8
SPREADING DYNAMICS OF POLYMER NANODROPLETS PHYSICAL REVIEW E 68, 061603 共2003兲

TABLE III. Model parameters and error estimates resulting from fits to contact angle data from simulations of wetting droplets on an
atomistic substrate. ␧ 12⫽1.5␧, ␥ D PD ⫽ ␥ L ⫽0.1␶ ⫺1 . The Langevin thermostat is applied to either all atoms 共Lang on All兲 or to substrate
atoms 共Lang on Sub兲.

Kinetic Combined

Thermostat N 冉
␨0 units of
m
␶␴ 冊 Hydrodynamic
a(units of ␴ ) 冉
␨0 units of
m
␶␴ 冊 a(units of ␴ ) ␹ kin
2
␹ hydro
2
␹ comb
2

DPD 10 36.7 42.1 53.4 65.5 0.001 0.015 0.001


Lang on All 10 50.8 38.1 81.8 70.0 0.001 0.015 0.001
Lang on Sub 10 38.0 41.8 64.9 69.6 0.001 0.020 0.001
Lang on Sub 20 54.4 43.2 91.9 68.3 0.001 0.020 0.001
Lang on Sub 40 65.5 42.9 126 64.0 0.002 0.019 0.002

monomers in the droplet. However, to do so, it is critical to tion. Evidence for hydrodynamic effects on spreading has
apply a thermostat that couples only to monomers near the been observed experimentally for macroscopic drops
surface. On an atomistic substrate, the most efficient method 关2,12,60兴. The length scale where hydrodynamic effects be-
is to couple only the substrate particles to the thermostat. come important remains an open question.
This is computationally faster than coupling all monomers to Future work will include studying the spreading behavior
the DPD thermostat and leads to the same results. of binary droplets and developing more realistic surface in-
Several droplet spreading models have been developed to teractions.
fit contact angle data. A simple kinetic mechanism for energy
dissipation fits the data well and provides reasonable values
for the friction coefficients, which we verified through sepa- ACKNOWLEDGMENTS
rate polymer melt simulations. Using a combined model that
adds a hydrodynamic energy dissipation mechanism slightly We thank M. O. Robbins for helpful discussions. Sandia
improves the fit, but resulted in less accurate estimates of the is a multiprogram laboratory operated by Sandia Corpora-
friction coefficients. The fact that we do not observe evi- tion, a Lockheed Martin Company, for the United States De-
dence of hydrodynamic flow behavior may be due to the partment of Energy’s National Nuclear Security Administra-
small droplet sizes accessible to molecular dynamics simula- tion under Contract No. DE-AC04-94AL85000.

关1兴 C. Huh and L.E. Scriven, J. Colloid Interface Sci. 35, 85 关14兴 M. Cachile, O. Bénichou, and A.M. Cazabat, Langmuir 18,
共1971兲. 7985 共2002兲.
关2兴 L.H. Tanner, J. Phys. D 12, 1473 共1979兲. 关15兴 M. Cachile, O. Bénichou, C. Poulard, and A.M. Cazabat,
关3兴 P.G. de Gennes, Rev. Mod. Phys. 57, 827 共1985兲. Langmuir 18, 8070 共2002兲.
关4兴 F. Heslot, N. Fraysse, and A.M. Cazabat, Nature 共London兲 338, 关16兴 F. Heslot, A.M. Cazabat, P. Levinson, and N. Fraysse, Phys.
640 共1989兲. Rev. Lett. 65, 599 共1990兲.
关5兴 M.J. de Ruijter, J. De Coninck, T.D. Blake, A. Clarke, and A. 关17兴 N. Fraysse, M.P. Valignat, A.M. Cazabat, F. Heslot, and P.
Rankin, Langmuir 13, 7293 共1997兲. Levinson, J. Colloid Interface Sci. 158, 27 共1993兲.
关6兴 M. Voué, M.P. Valignat, G. Oshanin, A.M. Cazabat, and J. De 关18兴 V.E. Dussan and S. Davis, J. Fluid Mech. 65, 71 共1974兲.
Coninck, Langmuir 14, 5951 共1998兲. 关19兴 S. Gladstone, K.J. Laidler, and H.J. Eyring, The Theory of Rate
关7兴 S. Gerdes, A.M. Cazabat, G. Ström, and F. Tiberg, Langmuir Processes 共McGraw-Hill, New York, 1941兲.
14, 7052 共1998兲. 关20兴 T.D. Blake, Ph.D. thesis, University of Bristol, 1968 共unpub-
关8兴 M.J. de Ruijter, J. De Coninck, and G. Oshanin, Langmuir 15, lished兲.
2209 共1999兲. 关21兴 T.D. Blake and J.M. Haynes, J. Colloid Interface Sci. 30, 421
关9兴 S. Semal, T.D. Blake, V. Geskin, M.J. de Ruijter, G. Castelein, 共1969兲.
and J. De Coninck, Langmuir 15, 8765 共1999兲. 关22兴 B.W. Cherry and C.M. Holmes, J. Colloid Interface Sci. 29,
关10兴 M.J. de Ruijter, M. Charlot, M. Voué, and J. De Coninck, 174 共1969兲.
Langmuir 16, 2363 共2000兲. 关23兴 O.V. Voinov, Fluid Dyn. 11, 714 共1976兲.
关11兴 M. Voué and J. De Coninck, Adv. Mater. 共Weinheim, Ger.兲 48, 关24兴 R.G. Cox, J. Fluid Mech. 168, 169 共1986兲.
4405 共2000兲. 关25兴 F. Brochard-Wyart and P.G. de Gennes, Adv. Colloid Interface
关12兴 E. Pérez, E. Schäffer, and U. Steiner, J. Colloid Interface Sci. Sci. 39, 1 共1992兲.
234, 178 共2001兲. 关26兴 P.G. de Gennes, Liquids at Interfaces 共North-Holland, New
关13兴 T.D. Blake and Y.D. Shikhmurzaev, J. Colloid Interface Sci. York, 1990兲, p. 371.
253, 196 共2002兲. 关27兴 P.G. Petrov and J.G. Petrov, Langmuir 8, 1762 共1992兲.

061603-9
HEINE, GREST, AND WEBB III PHYSICAL REVIEW E 68, 061603 共2003兲

关28兴 T.D. Blake, Wettability 共Marcel Dekker, New York, 1993兲. 关44兴 F. Heslot, A.M. Cazabat, and P. Levinson, Phys. Rev. Lett. 62,
关29兴 J.X. Yang, J. Koplik, and J.R. Banavar, Phys. Rev. Lett. 67, 1286 共1989兲.
3539 共1991兲. 关45兴 O.M. Braun and M. Peyrard, Phys. Rev. E 63, 046110 共2001兲.
关30兴 J.X. Yang, J. Koplik, and J.R. Banavar, Phys. Rev. A 46, 7738 关46兴 K. Kremer and G.S. Grest, J. Chem. Phys. 92, 5057 共1990兲.
共1992兲. 关47兴 S. Plimpton, J. Comput. Phys. 117, 1 共1995兲.
关31兴 S.M. Thompson, K.E. Gubbins, J.P.R.B. Walton, R.A.R. 关48兴 G.S. Grest and K. Kremer, Phys. Rev. A 33, 3628 共1986兲.
Chantry, and J.S. Rowlinson, J. Chem. Phys. 81, 1 共1984兲. 关49兴 P.J. Hoogerbrugge and J.M.V.A. Koelman, Europhys. Lett. 19,
关32兴 P. van Remoortere, J.E. Metrz, L.E. Scriven, and H.T. David, J. 155 共1992兲.
Chem. Phys. 110, 2621 共1999兲. 关50兴 P. Español and P. Warren, Europhys. Lett. 30, 191 共1995兲.
关33兴 J. De Coninck, U. D’Ortona, J. Koplik, and J.R. Banavar, 关51兴 R.C. Tolman, J. Chem. Phys. 17, 333 共1949兲.
Phys. Rev. Lett. 74, 928 共1995兲.
关52兴 R. McGraw and A. Laaksonen, J. Chem. Phys. 106, 5284
关34兴 U. D’Ortona, J. De Coninck, J. Koplik, and J.R. Banavar,
共1997兲.
Phys. Rev. E 53, 562 共1996兲.
关53兴 P.R. ten Wolde and D. Frenkel, J. Chem. Phys. 109, 9001
关35兴 X. Wu, N. Phan-Thien, X. Fan, and T.Y. Ng, Phys. Fluids 15,
共1998兲.
1357 共2003兲.
关54兴 A.E. Seaver and J.G. Berg, J. Appl. Polym. Sci. 52, 431
关36兴 J.A. Nieminen, D.B. Abraham, M. Karttunen, and K. Kaski,
Phys. Rev. Lett. 69, 124 共1992兲. 共1994兲.
关37兴 J.A. Nieminen and T. Ala-Nissila, Phys. Rev. E 49, 4228 关55兴 M.J.P. Nijmeijer, A.F. Bakker, C. Bruin, and J.H. Sikkenk, J.
共1994兲. Chem. Phys. 89, 3789 共1988兲.
关38兴 S. Bekink, S. Karaborni, G. Verbist, and K. Esselink, Phys. 关56兴 S.W. Sides, G.S. Grest, and Martin-D. Lacasse, Phys. Rev. E
Rev. Lett. 76, 3766 共1996兲. 60, 6708 共1999兲.
关39兴 A. Milchev and K. Binder, J. Chem. Phys. 116, 7691 共2002兲. 关57兴 M. Allen and D. Tildesley, Computer Simulation of Liquids
关40兴 T.D. Blake, A. Clarke, J. De Coninck, and M.J. de Ruijter, 共Clarendon Press, Oxford, 1987兲.
Langmuir 13, 2164 共1997兲. 关58兴 W.H. Press, S.A. Teukolsky, W.T. Vetterling, and B.P. Flan-
关41兴 M.J. de Ruijter, T.D. Blake, and J. De Coninck, Langmuir 15, nery, Numerical Recipes in C 共Cambridge University Press,
7836 共1999兲. Cambridge, 1992兲.
关42兴 T.D. Blake, A. Clarke, J. De Coninck, M. de Ruijter, and M. 关59兴 M. Doi and S.F. Edwards, The Theory of Polymer Dynamics
Voué, Colloids Surf., A 149, 123 共1999兲. 共Clarendon Press, Oxford, 1986兲.
关43兴 M. de Ruijter, T.D. Blake, A. Clarke, and J. De Coninck, J. Pet. 关60兴 G. Garnier, M. Bertin, and M. Smrckova, Langmuir 15, 7863
Sci. Eng. 24, 189 共1999兲. 共1999兲.

061603-10

S-ar putea să vă placă și