You are on page 1of 14

Home Search Collections Journals About Contact us My IOPscience

Stability and dynamics of droplets on patterned substrates: insights from experiments and

lattice Boltzmann simulations

This article has been downloaded from IOPscience. Please scroll down to see the full text article.

2011 J. Phys.: Condens. Matter 23 184112

(http://iopscience.iop.org/0953-8984/23/18/184112)

View the table of contents for this issue, or go to the journal homepage for more

Download details:
IP Address: 134.147.131.166
The article was downloaded on 21/04/2011 at 11:05

Please note that terms and conditions apply.


IOP PUBLISHING JOURNAL OF PHYSICS: CONDENSED MATTER
J. Phys.: Condens. Matter 23 (2011) 184112 (13pp) doi:10.1088/0953-8984/23/18/184112

Stability and dynamics of droplets on


patterned substrates: insights from
experiments and lattice Boltzmann
simulations
F Varnik1,2, M Gross1,2 , N Moradi1 , G Zikos1 , P Uhlmann3 ,
P Müller-Buschbaum4, D Magerl4 , D Raabe2 , I Steinbach1 and
M Stamm3
1
Interdisciplinary Center for Advanced Materials Simulation (ICAMS), Ruhr University
Bochum, Stiepeler Straße 129, 44780 Bochum, Germany
2
Max-Planck-Institut für Eisenforschung, Max-Planck Straße 1, 40237 Düsseldorf, Germany
3
Leibniz-Institut für Polymerforschung Dresden e.V., Hohe Straße 6, 01069 Dresden,
Germany
4
Physik-Department, Technische Universität München, LS E13, James-Franck-Straße 1,
85748 Garching, Germany

Received 3 June 2010


Published 20 April 2011
Online at stacks.iop.org/JPhysCM/23/184112

Abstract
The stability and dynamics of droplets on solid substrates are studied both theoretically and via
experiments. Focusing on our recent achievements within the DFG-priority program 1164
(Nano- and Microfluidics), we first consider the case of (large) droplets on the so-called gradient
substrates. Here the term gradient refers to both a change of wettability (chemical gradient) or
topography (roughness gradient). While the motion of a droplet on a perfectly flat substrate upon
the action of a chemical gradient appears to be a natural consequence of the considered situation,
we show that the behavior of a droplet on a gradient of topography is less obvious. Nevertheless,
if care is taken in the choice of the topographic patterns (in order to reduce hysteresis effects), a
motion may be observed. Interestingly, in this case, simple scaling arguments adequately account
for the dependence of the droplet velocity on the roughness gradient (Moradi et al 2010
Europhys. Lett. 89 26006). Another issue addressed in this paper is the behavior of droplets on
hydrophobic substrates with a periodic arrangement of square shaped pillars. Here, it is possible
to propose an analytically solvable model for the case where the droplet size becomes
comparable to the roughness scale (Gross et al 2009 Europhys. Lett. 88 26002). Two important
predictions of the model are highlighted here. (i) There exists a state with a finite penetration
depth, distinct from the full wetting (Wenzel) and suspended (Cassie–Baxter, CB) states.
(ii) Upon quasi-static evaporation, a droplet initially on the top of the pillars (CB state) undergoes
a transition to this new state with a finite penetration depth but then (upon further evaporation)
climbs up the pillars and goes back to the CB state again. These predictions are confirmed via
independent numerical simulations. Moreover, we also address the fundamental issue of the
internal droplet dynamics and the terminal center of mass velocity on a flat substrate.
(Some figures in this article are in colour only in the electronic version)

1. Introduction includes using droplets as a transport medium for medical


agents (e.g. drug delivery and lab-on-chip medical devices) as
The wetting behavior and dynamics of liquid droplets on solid well as industrial processes such as infiltration, coating and
substrates play a fundamental role for many applications. This printing. In inkjet printing, for example, tiny droplets (ink) are

0953-8984/11/184112+13$33.00 1 © 2011 IOP Publishing Ltd Printed in the UK & the USA
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

deposited onto a solid substrate at specified locations. Here, a introductions to the general foundations of the lattice
safe control of the droplet volume and the extent of the droplet Boltzmann method can be found in recent monographs [29–31]
spread (contact or base area) on the substrate are of fundamen- as well as in review articles [32, 33].
tal importance. On an ideally flat and chemically homogeneous Within a lattice Boltzmann method, one basically iterates
substrate, the contact area is larger on hydrophilic as compared two simple steps generally referred to as (1) relaxation
to hydrophobic substrates. Real surfaces, however, are neither (collision) and (2) free propagation (streaming),
perfectly flat nor chemically homogeneous. This makes an 1
f i (r , t) = f i (r , t) −
eq
experimental study of wetting phenomena a challenging task [ f i (r , t) − f i (r , t)], (1)
τ
(see e.g. [1] and references therein).
f i (r + dt ci , t + dt) = f i (r , t), (2)
In this context, computer simulations may provide a useful
and complementary tool. This is particularly interesting since where we introduced f i
in order to formally separate the
in a computer simulation one can easily focus on ideal cases, relaxation and streaming steps. In equation (1), the quantity
such as the behavior of a droplet on a perfectly flat substrate τ is the so-called relaxation time which determines the
with a well-defined chemical gradient. Conversely, one can fluid kinematic viscosity, ν = η/ρ (η = viscosity, ρ =
study the droplet dynamics on a chemically homogeneous fluid density). For the present three-dimensional model
substrate containing a regular topographic pattern. Additional with 15 non-zero velocities (D3Q15), one finds [30] ν =
features such as a spatially variable roughness (roughness (τ − 0.5) dx 2 /(3 dt), where dx is the distance between
gradient) or a combination of topographic and chemical two neighboring nodes connected along the main coordinate
directions and dt is the time step.
patterning may then be introduced into the model in order to
The physical properties of the system enter the LB
study more complex situations. eq
iteration scheme via the quantity f i equation (1). Obviously,
Furthermore, computer simulations often provide access eq
the system is ‘pushed’ towards f i with a rate 1/τ . The
to a number of mutually related quantities allowing us to shed eq
population density f i is, therefore, referred to as the
light onto the problem from different perspectives. An example
‘equilibrium distribution’. It is noteworthy that the term
is the possibility to obtain the center of mass velocity of the
‘equilibrium’ does not refer to a global thermal equilibrium,
droplet via two independent approaches, namely (i) from a
where no flow exists. Rather, it describes the local velocity
sequence of images of the droplet during its motion or (ii) by distribution in a portion of fluid moving at a velocity u(r ).
direct averaging over the velocity field within the droplet. eq
Within the present LB model, one expands fi in powers of
While the first approach is the usual way of determining the the fluid velocity u up to the second order [30]:
droplet velocity in real experiments (see e.g. figure 13), the  
three-dimensional velocity field inside the droplet is hard to eq 1 1
f i = ρwi 1 + 2 u · ci + 4 [(u · ci )2 − cs2 u · u] , (3)
access in an experiment. cs 2c s
In this paper, we provide some examples on the above where cs is the sound speed and wi is a set of weights
mentioned issues. This includes the behavior of liquids on normalized to unity. For the two-dimensional nine velocity
chemically patterned substrates, the motion of droplets driven model (D2Q9) used in our studies one finds w0 = 4/9, w1 =
by a spatial change of roughness density and the behavior w2 = w3 = w4 = 1/9 and w5 = w6 = w7 = w8 = 1/36.
of small droplets on a hydrophobic substrate with a regular Once the discrete populations, f i , are known, the fluid density,
topographic pattern. In the latter case, an analytic model is ρ(x, t), and velocity, u(x, t), at a given point and time are
proposed, capable of adequate description of the full stability obtained via
phase diagram. In the case of roughness gradient driven  
ρ= fi (x, t) and ρu = f i (x, t)ci . (4)
dynamics, on the other hand, simple scaling arguments are i i
found to provide a qualitative understanding of the relation
The approach described above must be supplemented by
between the average droplet velocity and roughness gradient.
additional information accounting for the non-ideal character
Furthermore, we also investigate the dynamics of a droplet on
of a liquid–vapor system. This can be done either by
a flat and chemically homogeneous substrate and compare our
(i) keeping the equilibrium distribution f eq unchanged but
results to recent molecular dynamics simulations [2] as well as
introducing density dependent interaction forces [16] or
experimental findings [3].
(ii) introducing a Cahn–Hilliard type free energy functional
and modifying the equilibrium distribution in such a way
2. The simulation method as to obtain both the thermodynamic and the irreversible
contributions to the pressure tensor correctly [34]. This
We use a simple but powerful computational tool, the latter approach has been further modified in order to ensure
lattice Boltzmann (LB) method [4–7], which has proved to Galilean invariance [35] as well as to include fluid–solid
be a versatile theoretical tool for the study of a variety interactions [36]. We emphasize here that, in this paper, we
of fluid dynamical phenomena such as the flow through are not presenting a new numerical method. Rather, we use
porous media [8], roughness effects on inertial flows in existing two-phase lattice Boltzmann models to study issues
narrow channels [9–11], flow of polymer solutions and relevant to the stability and dynamics of liquid drops. Unless
suspensions [12–15], multiphase flows [16–19], droplets on otherwise stated, we will use the approach proposed in [34]
topographically patterned substrates [20–23] and on gradients with appropriate modifications. Details of this method can be
of wettability [24–27] and topography [28]. Comprehensive found e.g. in [20, 27, 36] and references therein.

2
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

Figure 1. Creating wettability patterns via fluorosilane treatment. (a) Schematic representation of a substrate, partially covered with a
self-organized hydrophobic fluorosilane monolayer. The fluorosilane molecules reach the substrate from the vapor phase and chemically bind
to it (CVD technique). The uncovered area is initially coated with a photomask, which prevents fluorosilane from grafting to the substrate.
This part of the substrate thus remains hydrophilic. (b) Tiny water drops sprayed on a substrate, prepared as in (a). The contact angle of 106◦
observed here is in accord with table 1. It lies slightly below the average between the maximum values of θA and θR . (c) Another route to
chemical patterning: here, the role of the photomask is played by a polymer film, deposited via a commercially available overhead marker
(Staedtler™ Lumocolor permanent). Since the polymer film is not chemically bonded to the substrate (only physical adhesion), it can be
rinsed after the fluorosilane treatment (again via CVD technique), thus letting free the hydrophilic part of the substrate. (d) Water drops used
to visualize a chemical pattern produced via this procedure.

3. Droplets on chemically patterned substrates Table 1. Results of experimental measurements of the contact angle
of water on fluorosilanized glass substrates. θA (θR ) denotes the
In this section, we briefly address the behavior of droplets advancing (receding) contact angle. The number of investigated
samples is eleven.
on substrates containing spatially variable wettability. The
discussion is kept simple by focusing on an example of a Contact Average Standard Maximum Minimum
substrate with a chess board like chemical pattern. This angle value (deg) deviation (deg) (deg) (deg)
provides a nice example of how to confine a droplet to a well θA 113.2 5.0 122.6 104.4
specified region by the use of wetting gradients. θR 82.8 4.9 91.3 74.7

3.1. Preparation of chemically patterned surfaces


In order to provide a test case between experiments
There are various techniques to create hydrophobic coatings of and computer simulations, we used the photomask variant
solid surfaces. Here, we briefly sketch two variants of the so- of the silanization technique described above, and coated
called fluorosilane treatment (figure 1). In the first approach, glass substrates with a (tridecafluoro-1,1,2,2-tetrahydrooctyl)-
the parts of the substrate which shall remain hydrophilic are dimethylchlorosilane hydrophobic layer, thereby producing
covered by a photomask. The photomask is thus the negative different chemical patterns. One of these patterns is shown
image of the hydrophobic pattern which is to be created on in figure 2. The experiments of liquid behavior on these
a hydrophilic substrate. The sample is then exposed to the chemically patterned substrates were then paralleled by
vapor of fluorosilane—(tridecafluoro-1,1,2,2-tetrahydrooctyl)- computer simulation studies of the same situation.
dimethylchlorosilane in our experiments—which deposits on
the substrate and chemically grafts to it. Once this chemical 3.2. Computer simulation of droplets on a chemical pattern
vapor deposition (CVD) of fluorosilane has completed, the
photomask is removed thus uncovering the hydrophilic part of In the modeling part of our studies, the surface is assumed
the substrate. to be perfect (no contact angle hysteresis). A survey of
In the second approach, a commercially available table 1, however, reveals the presence of significant hysteresis
overhead marker is used in order to cover parts of the effects in our experiments. This is best seen by a comparison
substrate before silanization, which proceeds—similar to the of the advancing and receding contact angles. Thus, we
case of the photomask—via the CVD technique. Finally, the do not expect a full quantitative agreement between theory
polymer film is washed away with a solvent (often ethanol and experiments. Rather, we want to examine whether the
or acetone). While the use of a photomask is well suited to qualitative features observed in the experiments (e.g., the
create large structures, the marker-based method allows the localization of the droplet to the chemically patterned area)
design of finer patterns. Within this latter procedure, however, are also found in our computer simulations. As can be seen
nanoscale hydrophilic domains persist and the contact angles in figure 2, this is indeed the case. The observed deviations
thus reached are significantly lower than those obtained via the between simulation and experiment, on the other hand, are
use of a photomask [37]. rather expected. In the present simulation, the system is able to
The samples produced via the above described procedures avoid high curvatures at the contact zone of the two hydrophilic
are characterized by contact angle measurements and parts by slightly expanding towards the hydrophobic domains.
ellipsometry, the latter yielding the thickness of the In the corresponding experiment, however, such a process
hydrophobic coating layer. In the case of glass substrates, is very probably hindered by the contact angle hysteresis.
however, ellipsometry measurements were not possible and Regarding the dynamics, our studies of droplet spreading
the only route to characterize the samples was via contact on chemical gradients reveal that, despite complications due
angle measurements. The results of these measurements are to the geometry and spatial variation of wetting properties,
collected in table 1. the capillary time still remains one of the most important

3
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

Figure 2. (a) A glass substrate with a chess board like chemical pattern. (b) Liquid deposited on this substrate. (c) Lattice Boltzmann
computer simulation of a similar situation. (d) Footprint of the same droplet shown in (c). While the overall confinement of the droplet caused
by chemical patterning is well reproduced in the simulations, parts of the hydrophobic domain are also covered by the liquid. This is
necessary in order to avoid the high curvature of the liquid–vapor interface at the contact zone of the two hydrophilic domains. The presence
of such a high curvature in the experiment is very probably caused by the presence of hysteresis and the related pinning forces. The size of the
simulation box is 100 × 100 lattice units.

Figure 3. Quasi-static evaporation of a droplet on patterned (top) and perfectly smooth (bottom) hydrophobic substrates (Young contact angle
on the flat substrate = θY = 95◦ ). The comparison of the upper and lower images reveals that the rather large contact angle hysteresis
observed here (≈difference between the contact angles in (d) and (a)) largely arises from the presence of topography.

characteristic times of the separation process. Further studies Indeed, in these simulations, the contact angle continu-
of other related aspects and the governing scaling relations can ously decreases without ever reaching a steady value. This
be found in [27]. implies that, for the case considered here, the receding contact
angle is quite small (if not zero). The advancing contact angle,
on the other hand, is larger than 90◦ (see figure 3(a)).
3.3. Pinning at substrate roughness A few words are necessary regarding the definition and
We already remarked on the rather large difference between implementation of the topographic patterns in our simula-
tions. We introduce rectangular posts of linear dimension
the advancing and receding contact angles for our experimental
4 × 4 × 5 (length × width × height) lattice units. The distance
samples. A major reason for this large contact angle hysteresis
between two neighboring posts is set to eight lattice units along
is the surface roughness. This effect is nicely demonstrated
both the x and y directions. Thus, only 50% of the substrate
via our lattice Boltzmann simulations, shown in figure 3. The
is covered by the posts. At each solid node, two conditions are
aim of these simulations is to highlight the effect of surface
applied. The first condition is to ensure that the solid acts as an
topography on the receding contact angle. Instead of directly impenetrable barrier with no-slip boundary conditions.
subtracting the mass from the liquid, we allow here a low but The second condition regards the wetting behavior of the
constant outflow of vapor from the topmost layer (a horizontal substrate. A possible way to take account of this feature is
plane at L z − 1) of the simulation box. By doing so, we mimic a by introducing the contribution of the fluid–solid interaction
real situation corresponding to an open system with controlled to the total free energy of the system in terms of the fluid
outflow, without direct modification of the transport processes density. Linearizing this free energy [38], one obtains a
on the liquid–vapor interface. The term ‘low’ here means that simple condition for the gradient of the fluid density at
the resulting evaporation process is sufficiently slow so that, the solid surface [27, 36]. However, since this gradient
at each instant, the shape of the droplet corresponds to the must be computed in the direction normal to the surface,
local minimum of the surface free energy (evaporation time  its implementation in the case of an arbitrary orientation of
capillary time). In this sense, the studied evaporation process the solid surface becomes rather complicated. This explains
is quasi-static. why the topographic patterns used in our studies are made of

4
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

Figure 4. Left: top view of two step gradient substrates. In the upper panel (referred to as case A), the pillar density to the left side ( x < 50) is
φLeft = 0.187 (square posts of length a = b = 3) while it is set to φRight = 0.321 on the right side ( x > 50, rectangular posts of length a = 9
and width b = 3). The spacing distance of the pillars in the x -direction is dx = 5 and in the y -direction is d y = 3 overall on the substrate. The
height of the posts is c = 6. The lower panel (case B) is obtained from A by shifting the posts on each second row horizontally by an amount
(a + dx )/2 with dx = 5, a = 3 for x < 50 and a = 9 for x > 50. All lengths are given in LB units. Middle: the initial setup and final states of
a spherical droplet placed on the substrates of type A (top) and B (bottom). The droplet remains pinned to the roughness gradient zone in case
A while it passes over this zone in the case of substrate B. Right: images showing that similar final states are also observed for the case of
cylindrical droplets. Adapted from [28].

pieces of planar surfaces with normal vectors along the lattice roughness density, one may expect that a droplet placed at the
directions. Despite this simple geometry, careful book-keeping contact zone will leave this area towards the region of higher
is still necessary. It is particularly important here to realize and roughness density.
take account of the fact that corners, edges and faces provide However, the experimental observation of a roughness
different topologies for a computation of density gradients and gradient induced spontaneous motion is not an easy task [42].
thus require a separate treatment. A more detailed discussion In fact, even though equation (5) predicts a decrease in the
of this issue can be found in [20]. effective contact angle upon an increase in the roughness
density and hence a driving force along the gradient of φ ,
4. Spontaneous droplet motion induced by a gradient the contact angle hysteresis [46] may be strong enough in
of texture order to prevent a spontaneous droplet motion [42]. The
authors of [41, 42], for example, resort to vertically shaking the
While droplet behavior on homogeneous roughness has widely substrate in order to overcome the pinning forces. Furthermore,
been investigated in the literature, only a few works exist in these and similar experiments, it is often observed that the
dealing with the case of inhomogeneous topography [39–43]. behavior of droplets is not unique [39–42].
The starting point is the Cassie–Baxter (CB) equation for the In order to study this issue via computer simulations, we
effective contact angle of a droplet suspending on the top of adopt the above described case of a step wise change in the
roughness tips [44], roughness density and design a substrate divided into two
regions, each with a constant pillar density. Furthermore,
cos θCB = φ cos θY − (1 − φ) = φ(1 + cos θY ) − 1, (5) in order to underline the crucial role of pillar arrangement
on the behavior of droplets, we consider two different pillar
where the roughness density φ gives the fraction of the arrangements leading to the same pillar density gradient as
droplet’s base area, which is in contact with the solid. It is shown in the left panels of figure 4.
important to realize that equation (5) does not explicitly take The simulation results shown in figure 4 clearly
account of three-phase contact line structure. This shortcoming demonstrate that the behavior of a droplet on substrates
may, however, be neglected as long as the contact area reflects patterned by a pillar microstructure with the same pillar density
the structure and energetics of the three-phase contact line [45]. gradient but different pillar arrangement and geometries can be
Noting that 1 + cos θY > 0 also holds for hydrophobic qualitatively different. In particular, in the case of the substrate
substrates (θY > 90◦ ), it is easily seen from equation (5) of type A, the droplet motion is stopped on the gradient zone,
that cos θCB decreases upon a decrease of φ , resulting in a while in the case of substrate B it completely reaches the more
larger effective contact angle. Thus, the lower the pillar favorable region of higher φ .
density, the higher the effective hydrophobic character of the A detailed comparison of the effects of arrangements
patterned substrate. If one brings into contact two hydrophobic A and B on pinning forces seems quite a complicated and
substrates, each homogeneously decorated with a different demanding task. Instead, we investigate the dependence of the

5
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

in the viscous regime, we neglect inertial terms in the Navier–


Stokes equation and write for the steady state 0 = −∇ p+η u .
Here, u is the fluid velocity, p is the hydrostatic pressure and
η is the viscosity. The velocity u varies only over a distance
of the order of the droplet radius, hence u ∼ u/R 2 . On
the other hand, ∇ p ∼ −d pLaplace (θCB )/R = (σ/R 2 ) d R/R ,
assuming that the driving force originates from the Laplace
pressure variation (over a length of the order of R ) within the
droplet. For the case of a cylindrical droplet of unit length,
the condition of constant droplet volume,
= R 2 [θCB −
sin(2θCB )/2], equation
 (5) and some algebra lead  to d R/R ∼
(π − θCB ) dφ ∼ φ̄ dφ (the relation π − θCB ∼ φ̄ follows
Figure 5. The x -component of the center of mass position versus from equation (5) assuming θCB close to π [42] and φ̄ is defined
time for a cylindrical droplet using, φLeft = 0.187 and φRight = 0.321, as (φRight + φLeft )/2). Putting all this together, and after a
0.333, 0.35 and 0.375. The two groups of curves belong to two
different surface tensions of σ0 = 5.4 × 10−4 (LB units) (right) and
change of notation √ dφ ≡ φ = φRight − φLeft , we arrive at
4σ0 (left). Adapted from [28]. ηu/R 2 ∼ (σ/R 2 ) φ φ . Hence,

σ
u∼ φ̄ φ. (6)
η
average droplet velocity during its passage over the gradient
zone as a function of the pillar density gradient. For this It is noteworthy that both equations (6) and equation (5) in [42]
purpose, we create substrate patterns of type B with various predict a linear dependence of the droplet velocity on φ .
values of φ by keeping φLeft unchanged and varying φRight . In [42], the lateral velocity is estimated from the roughness
To simplify the analysis further, we place a cylindrical droplet gradient induced asymmetry of the dewetting of a droplet,
whose axis is along the y -direction. This allows us to use flattened due to impact. However, the situation we consider
the periodicity of the pillar arrangement along the y -direction is different. There is no impact and hence a related flattening
and thus also reduce the computation time. The results on is absent in the present case. Furthermore, the dynamics we
the dynamics of a cylindrical drop on such texture gradient study is in the viscous regime whereas the high impact velocity
substrates are shown in figure 5. in [42] supports the relevance of inertia. These differences
A survey of the center of mass position versus time in show up in different predictions regarding the dependence of
figure 5 reveals that the droplet motion is first linear in time the droplet velocity on surface tension, fluid viscosity and
until it reaches a constant value. The plateau corresponds to density. While equation (5) in [42] predicts a dependence on
the case where the droplet has completely left the region of the square root of σ , equation (6) suggests that, in our case, a
lower pillar density. This is in line with the fact that no driving linear dependence on σ should be expected.
force exists in this final state. Using the linear section of the We therefore examine equation (6) not only with regard to
data in figure 5, we define an average velocity for the motion the relation between the droplet velocity u and the difference
of the droplet’s center of mass. Importantly, figure 5 reveals in roughness density φ (left panel in figure 6), but we also
the strong effect of the surface tension on droplet dynamics. check how u changes upon a variation of the surface tension
Both the absolute values of the droplet velocity for a given φ σ for a fixed φ . The results of this latter test are depicted
as well as the slope of the data significantly depend upon σ . in the right panel of figure 6, confirming the expected linear
It is possible to rationalize these observations by simple dependence of u on σ . It is noteworthy that σ in the right panel
scaling arguments. First, for the sake of simplicity, we do not of figure 6 varies roughly by a factor of ten so that a square root
consider pinning effects here. Since the flow we consider is dependence can definitely be ruled out.

Figure 6. Left: the droplet’s center of mass velocity versus the difference in pillar density, φ = φRight − φLeft , extracted from the linear part
of the center of mass motion (see e.g. figure 5). Results for three different liquid–vapor surface tensions are depicted. From top to bottom:
4σ0 , 2σ0 and σ0 , where σ0 = 5.4 × 10−4 (LB units). In all cases, a linear variation is seen in accordance with the simple model, equation (6).
Right: a further test of equation (6), where the dependence of droplet velocity on surface tension is shown for a fixed φ .

6
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

Figure 7. A small droplet on a hydrophobic pillar array. Here, the term ‘small’ means that the linear dimension of the droplet is of the same
order of magnitude as the scale of the topographic pattern. (a) shows the impaled state as a typical situation in the present simulations. In the
analytical model (b), the droplet is assumed to be pinned at the edges of the pillars, i.e. it has a fixed base radius a . For a given droplet size,
the only way to minimize the overall free energy is thus through a change in the penetration depth p. This also fixes the apparent contact angle
θ . Adapted from [23].

5. Stability of small droplets The total volume of the droplet shall be fixed, Vtot =
const = Vsph (θ ) + Vcyl ( p), with Vsph = 13 πa 3 (2 − 3 cos θ +
The issues discussed above have one common point, namely cos3 θ )/ sin3 θ the volume of the cap and Vcyl = (πa 2 − b 2) p
that the size of the droplet is large compared to the typical the volume of the penetrating cylinder. In the following,
roughness scale of the substrate. This separation of length instead of the droplet volume Vtot , we will usually refer to the
scales often justifies the use of arguments based on average effective droplet radius Reff which corresponds to a spherical
properties (see equation (6)). Here, we would like to consider droplet of the same volume (4π/3 Reff 3
= Vtot ). We consider p
a situation where the droplet size is comparable to the length as the free variable and determine the dependence of θ on p
scale of the roughness. This situation is not only important for via the fixed volume condition.
a better understanding of the wetting properties of microscale Since the volume of the drop (and the temperature) is
systems [47], but also in many industrial applications, as, constant, only surface energy contributions play a role for a
for example, in the production of efficient self-cleaning change in the total free energy. The free energy f ( p) of the
surfaces [48], robust metal coatings, or in plasma spraying model droplet, neglecting gravity and terms associated with the
techniques. Moreover, small droplets naturally occur in many Wenzel transition, and normalizing such that f (0) = 0, is then
condensation or evaporation processes [49, 50]. given by

5.1. An analytically tractable model f ( p) = σLV [Ssph ( p)− Ssph (0)+ Scyl,LV ( p)− 8bp cos θY ]. (7)
The fact that we deal with droplets of a size comparable to the Here, Ssph = 2πa 2 (1 − cos θ )/ sin2 θ is the surface area of
roughness scale allows the use of a simple analytical model. the spherical cap Scyl,LV = (2πa − 4b) p is the lateral liquid–
The roughness of a surface is modeled by a regular array of vapor surface area of the cylinder and −σLV 8bp cos θY is the
cuboidal pillars with width b , height h and spacing d (figure 7). (positive, since θY > 90◦ ) free energy associated with the
The intrinsic hydrophobicity of the flat parts of the surface is wetting of the (eight) side walls of the pillars.
described by the Young contact angle θY . In this work, we
only focus on the case θY > 90◦ , which is also a necessary 5.2. Analytical results
condition for a (meta-) stable Cassie–Baxter state. Gravity will
be neglected throughout, as we only consider droplets that are Figures 8(a) and (b) show the dependence of the free energy on
smaller than the capillary length (∼2.7 mm for water). the penetration depth p for varying Young contact angles and
In the analytical model, we assume the part of the droplet droplet sizes. Several interesting observations can be made:
located above the pillars to be a spherical cap with base firstly, as also found in the case of droplets which are large
a = b/2 + d = R sin θ (with R being the radius of the compared to the roughness scale [52–56], the stability of the
cap and θ the apparent contact angle). The impaled part is CB state, determined by the slope of f at p = 0, depends not
approximated as a cylindrical liquid column with radius a and only on the contact angle but also on the size of the droplet.
height p (penetration depth), surrounding the central pillar. A novel feature is the appearance of a local minimum
The macroscopic contact line of the droplet is assumed to of the free energy at large penetration depths, existing in
remain pinned at the edges of the outer pillars. Note that addition to the possible minimum associated with the CB and
in this model the fully penetrated state (Wenzel [51]) would Wenzel states. From the condition d f /d p = 0, which is
correspond to p = h and the suspended (CB) state to p = 0. easily evaluated with the help of the fixed volume constraint,
Since the mechanism of the Wenzel transition has been a necessary condition for the existence of a minimum of the
discussed in detail by many authors in previous publications, free energy arises, namely, θY < arccos(−1/2 + b/(4a)),
we will ignore the Wenzel state completely, and, for the rest with a = b/2 + d being the base radius of the spherical cap.
of this work, assume the pillars to be so tall that no contact Interestingly, this condition does not depend on the droplet
between the liquid and the bottom of the grooves is possible. size.

7
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

Figure 8. Predictions of the analytical model. (a) Dependence of the free energy f /4πa 2 σLV on the penetration depth p and the Young
contact angle θY for a droplet of fixed size Reff = a . (The inset shows a magnification of the curve for θY = 108◦ .) (b) Dependence of the free
energy f /4πa 2 σLV on the penetration depth p and the droplet size Reff for a fixed contact angle of θY = 100◦ . (c) Contributions to the free
energy f /4πa 2 σLV for Reff = a and θY = 100◦ . f sph is the surface free energy of the liquid–vapor interface of the spherical cap, f cap is the
free energy due to the wetting of the ‘capillary’ constituted by the pillars and f = f sph + f cap is the total free energy. The insets sketch the
droplet configuration for different p according to the analytical model. All curves in (a)–(c) are given for b/d = 1 and plotted up to a value of
p where no further volume is left in the spherical cap. Adapted from [23].

Figure 9. Left: regions of stability for the CB state. We fix the pillar width b and plot the set of points (Reff , θY ) fulfilling
d f /d p(Reff , θY )| p=0 = 0, for varying values of the mutual pillar distance d (solid curves). The CB state is stable to the right hand side of each
solid curve. The dotted lines mark the limiting value of θY for the existence of a local minimum of the free energy. Right: reentrant transition.
Given a moderate contact angle, a large droplet is predicted by our analytical model to assume a metastable CB state (stage 1). On reduction
of its size the CB state becomes unstable and the droplet penetrates into the grooves (stage 2). If the height of the pillars prohibits a Wenzel
transition, the droplet eventually enters the CB state again upon further reduction of its size (stage 3). The boundary curve of the CB stability
region is shown for the case b = d . Adapted from [58].

The origin of this new state with a finite penetration depth surface geometries (e.g. omitting the central pillar) clearly
can be understood by imagining the pillars to represent a underline this assertion [57].
(partly open) hydrophobic capillary tube, wetted by a small These results also show under which conditions we are
droplet that is placed at its entrance. In this situation, the allowed not to consider the Wenzel state in the first place: a
equilibrium state of the droplet is a consequence of the balance transition to this state can be inhibited, if the pillar height h
between the Laplace pressure within the spherical cap (pushing is larger than the penetration depth p of a droplet at the local
the droplet into the capillary) and an opposing capillary force minimum, plus a small correction of the order of d 2 /Reff [1]
due to the hydrophobicity of the substrate. that accounts for the curvature of the lower droplet interface.
To illustrate this idea, we split the free energy As becomes clear from figure 9, the shape of the stability
(equation (7)) into the contributions of the spherical cap and region of the CB state is largely independent of the geometry,
the remaining ‘capillary’ part. As shown in (figure 8(c)), an and its characteristic shape offers the interesting possibility of
increase of the droplet penetration p leads to a linear increase a reentrant transition: for a given Young contact angle, imagine
of capillary free energy, while the free energy associated with a droplet that is very large and initially deposited at the top of
the spherical cap decreases in a nonlinear fashion. As a the pillar array. According to our analytical model, the droplet
result, the total free energy f may exhibit a local minimum. will adopt a CB state (stage 1 in figure 9). If the droplet is
This simple reasoning suggests that the intermediate minimum now reduced in size (e.g. through evaporation) and the Young
constitutes a generic equilibrium state of a droplet, occurring contact angle is not too large, the droplet will enter the region
in any situation of filling hydrophobic capillaries by a spherical of CB instability (stage 2). Depending on the tallness of the
liquid reservoir. Indeed, further simulations using various pillars, the droplet will now either go over into the Wenzel state

8
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

Figure 10. Reentrant transition through evaporation of a droplet. In the simulation, the droplet is initially (t = 0) prepared in a (meta-)stable
(partially impaled) Cassie state (a). Evaporation is then switched on and proceeds by reducing, at a sufficiently low rate satisfying quasi-static
equilibrium, the mass of the vapor phase across a x y -plane close to the top of the simulation box. In the course of the evaporation process, the
Cassie state becomes unstable and adopts an impaled state (b), from where it gradually climbs up the pillars (c) until it again reaches a Cassie

state (d). (e) shows the penetration depth of the lower droplet interface as it is observed in the simulation ( ) and predicted by the analytical
model (——). Simulation parameters: contact angle θY = 103◦ , initial droplet size Reff = 1.2a , b = d = 12 LB units, evaporation rate
5 × 10−7 /LB time steps. All lengths are given in units of the LB grid spacing. Adapted from [23].

or the impaled state. In the latter case, further reduction of the such situations adequately, the present model must be extended
droplet’s size brings it back again into the CB state (stage 3). appropriately. Such a study presents an interesting topic for
This remarkable behavior is due to the peculiar dependence of future work. Second, it is interesting to address the range of
the position of the intermediate minimum of the free energy on length scales for the validity of the present predictions. For this
droplet size (figure 10(b)). purpose, we recall that solid deformability, thermal fluctuations
Figure 10 provides evidence from simulations for the and gravity are neglected in the investigated analytic model. In
existence of the reentrant transition. This is achieved through other words, only surface free energies are considered in the
a quasi-static evaporation process of a relatively large droplet. model. This makes the model scale invariant since, in this case,
In the beginning (figure 10(a)), the droplet resides on the top f (αa, αb, αd, αp, α R) = α 2 f (a, b, d, p, R), where α is an
of the pillars with a tiny penetration depth. However, after arbitrary scale factor.
reaching a critical size, the droplet suddenly penetrates into As a consequence, all the predictions made within our
the pillar grooves and goes over to the intermediate minimum model are scale invariant as long as the above approximations
of the free energy (figure 10(b)). During the impaled phase hold. While the first two of the above mentioned
(figures 10(b) and (c)) the droplet gradually climbs up the approximations set a lower bound to the available length
pillars again, still residing in the local minimum. Note that scales (of the order of microns, the exact value depending
its penetration depth is in nice agreement with the analytical on the substrate elastic modulus and the surface tension),
predictions (figure 10(e)). In the end of the process, the droplet the third one implies that the droplet size should be less
again attains a stable Cassie–Baxter state (figure 10(d)), as than the capillary length (≈2.7 mm for water). Thus, when
expected from the phase diagram, figure 9. However, we
experimentally testing the present predictions, it is a priori
remark here that this transition actually happens for a slightly
not necessary to use micro-pillars with correspondingly small
larger radius than predicted analytically, since the pinning
micro-droplets (which are well known to be prone to strong
condition cannot be maintained if the penetration depth is
evaporation). Rather, experiments on sufficiently large droplets
below the thickness of the liquid–vapor interface, which is of
with diameters, d , below the capillary length (e.g., d
1 mm)
the order of 3–4 lattice units in the present case.
would also be adequate for such a purpose.
According to the common understanding of self-cleaning,
impalement is considered unfavorable and the droplet cleans
the surface by rolling over the top of the texture. Interestingly,
the existence of a reentrant transition suggests the possibility of
6. Droplet dynamics on a flat substrate
a qualitatively new self-cleaning mechanism, since the droplet
not only touches the top of the substrate, but also its inner Here, we employ a recent variant of the two-phase lattice
parts. The existence of a reentrant transition can also explain Boltzmann method in order to investigate the detailed
some recent experimental observations, which found that small dynamics of fluid drops on solid substrates. The method
evaporating droplets indeed tend to remain close to the top of combines a free energy approach to the liquid–vapor
the substrate [55, 59] and not get trapped inside of the texture. coexistence with a numerical scheme allowing us to literally
We close this section with two important remarks. First, eliminate the so-called parasitic currents [61]. It is important
in the present analytical model as well as in the accompanying to realize that the elimination of the spurious currents is an
lattice Boltzmann simulations, the pillars are assumed to be important step towards a reliable description of fluid dynamics
perfectly rigid. In the case of sufficiently tall micron-sized inside a droplet. Particular attention has also been paid to the
pillars, however, the elasticity of the pillars and the resulting correct implementation of the forcing term [62] in order to
deformation may become important [60]. In order to describe minimize error terms arising from the inhomogeneous nature

9
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

Figure 11. (a) Fluid momentum in a channel containing a cylindrical droplet which moves under the action of a force proportional to the fluid
density. Obviously, there is hardly momentum in the vapor phase. (b) Depiction of the same data viewed from a frame of reference comoving
with the drop’s center of mass. The dark (bright) color indicates large (small) density. All quantities are in LB units.

Figure 12. (a) Shear stress across the channel visualized via a color code (dark = high stress). The arrows show the velocity field in the lab
frame. The strongest shear stresses occur at the three-phase contact line and decay towards the middle of the droplet. Furthermore, the shear
stress is significantly higher close to the substrate than in the center of the drop. In particular, it is practically negligible at the liquid–vapor
interface. (b) Velocity profiles along vertical lines across the droplet (i.e. u x versus y ) at different lateral positions x as indicated in the panel
(a). (c) A plot similar to (b) but using the momentum data (ρu x (y)). (d) Terminal drop velocity versus drop volume for different choices of
kinematic viscosities νliquid and νvapor. Viscous liquid: νliquid > νvapor. Viscous vapor: νliquid < νvapor. Equal viscosities: νliquid = νvapor . All
quantities are in LB units.

of the mean field type interactions, the latter being essential for small. However, as the plot clearly shows, the momentum
the two-phase behavior of the system. contained in the vapor phase is negligible as compared to the
Figure 11 shows the results of LB simulations on the liquid momentum. This also reflects itself in the significant
dynamics of a two-dimensional droplet under the action of decrease of the shear stress as the vapor phase is approached
a gravity like force (equivalent to the case of a cylindrical (see figure 12).
droplet in the 3D studies of [2]). In the left panel, the Shear stress plays a central role in theoretical models
fluid momentum field is shown in the laboratory frame of aimed at a description of the droplet or liquid film dynamics. It
reference, while the right panel depicts the same data within is often assumed that shear stress vanishes at the liquid–vapor
a reference frame, which moves with the droplet’s center of interface (see e.g. [63]). It is, therefore, important to examine
mass. While the no-slip boundary condition is evident from how this quantity varies across the channel in the present case.
the left panel (fluid momentum is zero close to the substrate) an For this purpose we show in figure 12(a) the shear stress field in
observer moving with the droplet’s center of mass will confirm the entire channel using a color code. Obviously, the strongest
the presence of a well established rotational flow inside the shear stresses occur at the three-phase contact line and close to
droplet. Interestingly, similar rotational flows are also observed the liquid–solid contact. Away from the solid, the shear stress
in molecular dynamics simulations of polymeric liquids [2]. decays rapidly. In the center of the droplet, where rotational
It is noteworthy that the vapor velocity is not necessarily flow is observed in the comoving frame (figure 11(b)), the

10
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

Figure 13. (a) Selected optical microscopy snap-shots of a toluene droplet (volume 0.3 μl) moving down an incline (from top to bottom, left
to right) on a HDMS covered Si substrate. The time between individual pictures is 0.34 s and in total 5.1 s were measured. Each image shows
an area of 4.3 × 3.2 mm2 . (b) Terminal droplet velocity v plotted as a function of the contact area A for toluene droplets of different volumes
on basic cleaned Si. The solid line is a fit to the data with a model as explained in the text.

shear stress is quite small in accordance with the idea that the shape is no longer constant. On hexamethyldisiloxane
rotational motion minimizes the dissipation loss. Figure 12(b) (HDMS) covered Si substrates the contact angle is significantly
depicts the fluid velocity profiles at various vertical cuts across increased which allowed for experiments with moving toluene
the droplet, i.e. u x versus y for various lateral positions, x . droplets of constant shape (see figure 13(a)). The experiments
Analytical treatments of droplet motion often assume a simple were always prepared on freshly prepared HDMS films.
parabolic shape for the fluid velocity inside the droplet [64]. Evaporation of the toluene from the droplet was prevented by
Interestingly, a parabolic envelope for the velocity field has a surrounding toluene atmosphere.
recently been reported in molecular dynamics simulations [2]. Figure 13 shows the example of a toluene droplet moving
Within our lattice Boltzmann simulations and for the chosen down a HDMS covered Si substrate (inclination angle 60◦ ).
set of parameters, however, we see that the velocity profile Very clearly the droplet exhibits a spherical cap shape and its
changes with the lateral position x in a rather non-trivial way. front is well visible. During movement the droplet does not get
Parts of this profile can nevertheless be approximated by a expanded along the direction of the acting body force, which
parabolic profile. results in a constant shape. The movement of the droplet front
Another closely related issue addressed in our studies is was considered. For each image the position of the droplet
the scaling of terminal (steady state) droplet velocity with front was probed and a distance–time diagram was extracted
droplet volume. As shown in figure 12(d), the steady from the data. After an acceleration phase the slope of the
state drop velocity is a linear function of the drop volume. distance–time diagram is constant being the measured droplet
Interestingly, this behavior does not depend on the ratio of velocity. It has to be noted that unavoidable evaporation causes
kinematic viscosities of the liquid and vapor phases. A a deceleration of the droplet for long distances measured from
linear dependence between the terminal drop velocity and the point of droplet deposition.
drop volume has been suggested in [2] for the case of large In addition, it turned out to be more reliable to use the
droplets but the observation of this behavior was hampered contact area of the droplet with the underlying HDMS covered
by the rather small droplet sizes accessible to molecular Si substrate instead of the droplet volume. Depending on the
dynamics simulations [2]. However, as will be discussed actual volume, and in particular for very small sub-microliter
below, providing direct experimental evidence for this relation droplets, the real deposited volume and the nominal volume
still remains a challenge. showed non-negligible deviations. Thus figure 13(b) shows
the droplet velocities as a function of the contact area. The
7. Experimental studies of droplet motion on solid critical contact area, which droplets needed to have in order
substrates to exhibit motion on an incline with 60◦ inclination was
A = 2.0 mm2 . Smaller droplets did not move. The data
In the experimental part of the study, we use optical suggest a linear dependence of the terminal droplet velocity on
microscopy to detect the droplet motion on solid substrates. the contact area, if the latter exceeds the mentioned threshold
To simplify the experiments and allow for comparison with value.
theory, simple liquids were used. However, it turned out that However, due to a rather strong scatter of droplet volume
the movement of liquids such as toluene on simple substrates in these experiments, it is not possible to make a statement
such as chemically pre-treated silicon (Si) substrates cannot on the relation between the terminal drop velocity and its
be easily compared with theory due to the droplet shape. volume. Note that a linear dependence on the contact area
Caused by the small contact angle of toluene droplets on Si would suggest that the main dissipation mechanism arises from
substrates (e.g. 5◦ for toluene on basic cleaned Si), the droplets the friction between the substrate and the fluid. Following the
get elongated during movement [3]. As a consequence, analysis in [64], surface dissipation is expected to dominate

11
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

when the slip length becomes large compared to the vertical distribution of the shear stress. The latter is quite strong in the
dimension of the droplet, which is of macroscopic size (1 mm) proximity of the three-phase contact line, whereas it becomes
in our experiments. However, the sticky behavior observed in negligibly small at the liquid–vapor interface.
the present studies does not support such a large slip length. These results call for more elaborate studies of droplet
The present experiments thus must be complemented by new dynamics on solid substrates. A particularly important aspect
ones allowing safe control of the droplet volume so that reliable here is the effect of wettability and topographic roughness
data on the volume dependence of the terminal velocity can be as well as chemical patterning on the hysteresis and the
obtained. Another interesting challenge for new experiments resulting pinning. Hysteresis makes reproducible experiments
would be the optical visualization of the droplet velocity a challenging task. A better understanding of this issue could
in order to obtain direct information on the internal droplet help to eliminate or at least significantly reduce the hysteresis
dynamics. and related undesired effects. It is also highly desirable to
build upon the valuable knowledge gained via these studies in
8. Conclusion order to design new experiments as well as perform well tuned
computer simulations of more complex phenomena such as the
In this report, we present our recent achievements within the flow-mediated deposition of ordered structures on top of solid
DFG-priority program 1164, Nano- and Microfluidics. In surfaces.
particular, we discuss some selected issues relevant for the
behavior of liquid drops on chemically or topographically Acknowledgments
patterned substrates. An example is the effect of a chemical
step on droplet dynamics, the motion of droplets driven by a We gratefully acknowledge stimulating discussions with
spatial change of roughness density and the behavior of small
various colleagues within the DFG-priority program 1164
droplets on a hydrophobic substrate with a regular topographic
Nano- and Microfluidics during the past six years: M Müller,
pattern.
B Wu, P Truman, T Haraszti, A Wixforth, M F Schneider,
When placed on a perfectly flat substrate, the action of
O Vinogradova, J Harting, H Zabel, K Jacobs, M Rauscher.
a chemical gradient drives the fluid from a less towards a
Most of the simulations reported here (sections 3–5) were
more hydrophilic region. It is shown how this property can
performed using a variant of the D3Q15 non-ideal fluid LB
be used in order to separate an emulsion into its individual
code provided by A Dupuis. Financial support was provided
components. As to the behavior of a droplet on a gradient of
by the DFG in the framework of the priority program 1164
topography, it is found that contact angle hysteresis often plays
Nano- and Microfluidics (grats MU1487/2, STA324/27 and VA
a crucial role. Nevertheless, if care is taken in the choice of the
205/3).
topographic patterns (in order to reduce the hysteresis effects),
a motion may be observed. Interestingly, in this case, simple
scaling arguments adequately account for the dependence of References
the droplet velocity on the roughness gradient.
In the case of hydrophobic substrates with a periodic [1] Quéré D 2008 Annu. Rev. Mater. Res. 38 71
[2] Servantie J and Müller M 2008 J. Chem. Phys. 128 014709
arrangement of pillars, it is possible to propose an analytically [3] Müller-Buschbaum P et al 2011 J. Phys.: Condens. Matter
solvable model for the case where the droplet size becomes 23 184111
comparable to the roughness scale. Two important predictions [4] McNamara G and Zanetti G 1988 Phys. Rev. Lett. 61 2332
of the model are highlighted and confirmed via independent [5] Higuera F and Jimenez J 1989 Europhys. Lett. 9 663
numerical simulations. (i) There exists a state with a finite [6] Benzi R, Succi S and Vergassola M 1992 Phys. Rep. 3 145
[7] Qian Y, d’Humieres D and Lallemand P 1992 Europhys. Lett.
penetration depth, distinct from the full wetting (Wenzel) and 17 479
suspended (CB) states. (ii) Upon quasi-static evaporation, a [8] Gunstensen A and Rothman D 1993 J. Geophys. Res. 98 6431
droplet initially on the top of the pillars (CB state) undergoes [9] Varnik F and Raabe D 2006 Modelling Simul. Mater. Sci. Eng.
a transition to this new state with a finite penetration depth but 14 857
then climbs up the pillars and goes back to the Cassie–Baxter [10] Varnik F, Dorner D and Raabe D 2007 J. Fluid Mech. 573 191
[11] Varnik F and Raabe D 2007 Mol. Simul. 33 583
state again. [12] Ahlrichs P and Dünweg B 1999 J. Chem. Phys. 111 8225
Furthermore, we also discuss the dynamics of liquid drops [13] Ladd A and Verberg R 2001 J. Stat. Phys. 104 1191
on solid substrates and investigate the motion of drops under [14] Dupin M M et al 2007 Phys. Rev. E 75 066707
the action of a gravity like force and its dependence on [15] Krüger T, Varnik F and Raabe D 2011 Comput. Math. Appl.
parameters such as the drop volume and kinematic viscosity at press
[16] Shan X and Chen H 1993 Phys. Rev. E 47 1815
of the liquid and vapor phases. Experiments clearly show [17] Swift M, Orlandini E, Osborn W and Yeomans J 1996 Phys.
a strong effect of the droplet volume. This is best seen Rev. E 54 5041
by the observation that drops with a contact area smaller [18] Luo L S 1998 Phys. Rev. Lett. 81 1618
than ≈14 mm2 did not move at all. For larger drops it is [19] Benzi R, Chibbaro S and Succi S 2009 Phys. Rev. Lett.
found that the terminal drop velocity scaled linearly with the 102 026002
[20] Dupuis A and Yeomans J M 2005 Langmuir 21 2624
drop-substrate contact area. Our lattice Boltzmann computer [21] Dupuis A and Yeomans J M 2006 Europhys. Lett. 75 105
simulations shed light onto very important aspects of the [22] Sbragaglia M and Succi S 2006 Europhys. Lett. 73 370
problem such as the velocity field inside the droplet and the [23] Gross M, Varnik F and Raabe D 2009 Europhys. Lett. 88 26002

12
J. Phys.: Condens. Matter 23 (2011) 184112 F Varnik et al

[24] Leopoldes J, Dupuis A, Bucknall D G and Yeomans J M 2003 [42] Reyssat M, Pardo F and Quéré D 2009 Europhys. Lett.
Langmuir 19 9818 87 36003
[25] Dupuis A and Yeomans J M 2004 Future Gener. Comput. Syst. [43] Fang G, Li W, Wang X and Qiao G 2008 Langmuir 24 11651
20 993 [44] Cassie A B D and Baxter S 1944 Trans. Faraday Soc. 40 546
[26] Chang Q and Alexander J 2006 Microfluidics Nanofluidics [45] Gao L and McCarthy T 2007 Langmuir 23 3762
2 309 [46] Kusumaatmaja H and Yeomans J M 2007 Langmuir 23 6019
[27] Varnik F et al 2008 Phys. Fluids 20 072104 [47] Seemann R et al 2005 Proc. Natl Acad. Sci. USA 102 1848
[28] Moradi N, Varnik F and Steinbach I 2010 Europhys. Lett.
[48] Blossey R 2003 Nature Mater. 2 301
89 26006
[29] Rothman D H and Zaleski S 1997 Lattice-Gas Cellular [49] Narhe R D and Beysens D A 2004 Phys. Rev. Lett. 93 076103
Automata (Simple Models of Complex Hydrodynamics) [50] Dorrer C and Rühe J 2007 Langmuir 23 3820
(Cambridge: Cambridge University Press) [51] Wenzel R N 1936 Indust. Eng. Chem. 28 988
[30] Wolf-Gladrow D 2000 Lattice-Gas Cellular Automata and [52] Lafuma A and Quéré D 2003 Nature Mater. 2 457
Lattice Boltzmann Models (Berlin: Springer) [53] McHale G et al 2005 Langmuir 21 11053
[31] Succi S 2001 The Lattice Boltzmann Equation: For Fluid [54] Moulinet S and Bartolo D 2007 Eur. Phys. J. E 24 251
Dynamics and Beyond, Numerical Mathematics and [55] Reyssat M, Yeomans J M and Quéré D 2008 Europhys. Lett.
Scientific Computation (Oxford: Oxford University Press) 81 26006
[32] Chen S and Doolen G 1998 Annu. Rev. Fluid Mech. 30 329 [56] Kusumaatmaja H, Blow M L, Dupuis A and Yeomans J M 2007
[33] Raabe D 2004 Modelling Simul. Mater. Sci. Eng. 12 R13 Europhys. Lett. 81 36003
[34] Swift M, Osborn W and Yeomans J 1995 Phys. Rev. Lett. [57] Moradi N, Varnik F and Steinbach I 2010 in preparation
75 830 [58] Gross M, Varnik F, Raabe D and Steinbach I 2010 Phys. Rev. E
[35] Holdych J-G G D J, Rovas D and Buckius R O 1998 Int. J. 81 051606
Mod. Phys. C 9 1393 [59] Jung Y C and Bhushan B 2007 Scr. Mater. 57 1057
[36] Briant A J, Wagner A J and Yeomans J M 2004 Phys. Rev. E
[60] Brücker C 2011 J. Phys.: Condens. Matter 23 184120
69 031602
[61] Lee T and Fischer P E 2006 Phys. Rev. E 74 046709
[37] Trumann P 2007 PhD Thesis Technical University of Dresden
[38] de Gennes P G 1985 Rev. Mod. Phys. 57 827 [62] Guo Z, Zheng C and Shi B 2002 Phys. Rev. E 65 04630
[39] Zhu L, Feng Y, Ye X and Zhou Z 2006 Sensors Actuators A [63] Landau L D and Lifschitz E M 1991 Hydrodynamik vol 6
130/131 595 (Berlin: Akademie)
[40] Yang J, Chen J, Huang K and Yeh J 2006 J. Microelectromech. [64] de Gennes F B P G and Quéré D 2004 Capillarity and Wetting
Syst. 15 697 Phenomena: Drops, Bubbles, Pearls, Waves (New York:
[41] Shastry A, Case M and Böhringer K 2006 Langmuir 22 6161 Springer)

13

You might also like