Sie sind auf Seite 1von 17

International Journal of Heat and Mass Transfer 99 (2016) 805–821

Contents lists available at ScienceDirect

International Journal of Heat and Mass Transfer


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

A numerical study on electrowetting-induced jumping and transport


of droplet
K. Ashoke Raman ⇑, Rajeev K. Jaiman ⇑, Thong-See Lee, Hong-Tong Low
Department of Mechanical Engineering, National University Singapore, 10 Kent Ridge Crescent, Singapore 117576, Singapore

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

Article history: The evolution dynamics of wettability-induced jumping droplet is studied numerically by means of a
Received 19 February 2016 high-density ratio based lattice Boltzmann method. The three phase moving contact line is modelled
Received in revised form 13 April 2016 by a geometry based wetting formulation. We analyse the influence of pulse characteristics with respect
Accepted 14 April 2016
to its nature, pulse width and double pulse actuation on the jumping dynamics. It is shown that these
Available online 28 April 2016
factors strongly affect the droplet lift-off mechanism. The influence of surrounding gas density and vis-
cosity on the droplet behaviour is investigated systematically. It is found that the role of the ambient fluid
Keywords:
becomes prominent during the recoiling and the lift-off phase of the jumping droplet. Finally, we demon-
Droplet jumping
Lattice Boltzmann method
strate the applicability of wettability-induced droplet jumping in microfluidic devices where efficient
Gas density and viscosity transport of droplet is essential.
Droplet transport Ó 2016 Elsevier Ltd. All rights reserved.

1. Introduction drop is attached to the solid surface. Different methods can be


employed to actuate such forces leading to droplet detachment.
Detachment and removal of droplets from a solid substrate is an Self-propelled jumping of droplets on superhydrophobic surfaces
essential component in several industrial and biological applica- is observed in many natural water-repellent surfaces and engineer-
tions. It is an important procedure in self-cleaning [1], anti-dew ing applications. The droplets on these surfaces merge and coalesce
[2,3] and anti corrosion [4] applications, where removal of droplets with the adjoining droplets. This results in the release of excess
from the surface is required. The droplet jumping motion is essen- surface energy which primarily converts into the kinetic energy
tial for heat transfer systems to enhance dropwise condensation of the jumping droplet system. The first experimental study on
[5–7] and to promote thermal rectification [8,9]. The condensate self-propelled jumping of coalescence induced droplets was
drops which form on the solid surface due to condensation, reduce reported by Boreyko and Chen [2]. They performed experiments
the heat transfer between the surface and the vapour. Hence, on superhydrophobic surfaces with droplet sizes ranging from
removal of these condensate drops is vital to enhance heat transfer. 10 lm to 150 lm. The results indicated an increase in the vertical
In addition, detachment and removal of droplets is one of the sali- velocity of the droplet system with increase in droplet size. How-
ent operations in digital microfluidics applications [10,11]. These ever, after reaching a maximum vertical velocity, a decrease in
applications involve manipulation and removal of droplets to per- velocity is observed on further increasing the drop size. To estab-
form mixing, transport and chemical reactions. Apart from its rel- lish a relationship for coalescence induced velocity based on the
evance to industrial processes, the basic science of the fluid energy conservation of the coalescing droplets, theoretical analysis
dynamical aspects of an detaching droplet has attracted several based on energy analysis was conducted by Wang et al. [12]. The
researchers to explore this phenomenon. Therefore, understanding relationship for the coalescence induced velocity was based on
the mechanism of droplet jumping process from a solid surface is the balance between surface energy, kinetic energy and viscous
of great interest with a practical significance in optimising these dissipation. Farokhirad et al. [13] performed lattice Boltzmann
applications. simulations to investigate the effects of air inertia and viscosity
To detach a droplet adhered to a solid substrate, the applied on the jumping height of the droplet. The results showed that an
force should be greater than the threshold force with which the increase in air viscosity inhibited circulation within the droplet
and self-propelled jumping is not noticed. Increase in air inertia
⇑ Corresponding authors. resulted in higher jumping heights of the droplet system. Liu
E-mail addresses: ashokeraman@u.nus.edu (K. Ashoke Raman), mperkj@nus.edu.sg et al. [14] performed two dimensional lattice Boltzmann simula-
(R.K. Jaiman). tions for self-propelled jumping of droplets on textured surfaces.

http://dx.doi.org/10.1016/j.ijheatmasstransfer.2016.04.038
0017-9310/Ó 2016 Elsevier Ltd. All rights reserved.
806 K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821

Shi et al. [15] investigated three dimensional coalescence-induced elucidate the mechanism of droplet jumping. In order to optimise
droplet jumping behaviour on superhydrophobic complex textured the design and efficiency of several microfluidic applications which
surface by altering the wettability of the textured surface, changing involve droplet detachment and transport, it is imperative to have
the roughness of the textured surface, controlling the number and a basic understanding of the involved flow physics. In electrowet-
size of the droplets by the MRT pseudopotential lattice Boltzmann ting actuation, the dynamics of the droplet motion is predomi-
model. nantly governed by the evolution of the moving contact line. The
Droplet detachment is also accomplished from vibrational kinetics of the moving contact line in turn depends on the fluctua-
motion of the solid substrate [16,17]. Kim [16] performed exper- tions of the voltage which govern the surface contact angle. In this
iments to show that a liquid drop pendant from a solid surface work, we examine the influence of the nature voltage fluctuations
can be effectively detached by vibrating the solid at specific fre- (pulse characteristics) and resulting kinematic mechanisms of the
quency ranges with weak external vibrations. At these frequen- moving contact line leading to different dynamical droplet beha-
cies, the drop oscillation was found to exhibit resonant viour. In this context, we evaluate the effect of various parameters
behaviour with the solid motion. A novel way to create a spray like the nature of the actuating pulse and electrowetting number
was outlined by James et al. [17]. They placed a liquid drop on a on the jumping dynamics of the droplet. Secondly, in several cases,
vertically vibrating solid surface which lead to the formation of it is required to transport droplet when the surrounding medium is
waves along the droplet free surface. While secondary droplets not air. While the role of the ambient fluid density may not be
were ejected above a certain critical forcing amplitude, for small prominent prior to jumping, it has a significant effect on droplet
forcing frequency, a single drop was ejected from the tip of the dynamics once it lifts-off the surface [25]. Hence, it is crucial to
primary drop. Movement of mechanical parts can impose several understand the role of the properties of the ambient environment
restriction such as difficulties in integration and noise generation on the dynamic evolution of the jumping droplet. The current work
in many applications. Some researchers have employed electro- illustrates the effect of gas density and viscosity on the jumping
static forces to disengage droplets from solid surfaces [18,19]. process. Finally, we demonstrate the applicability of electrowetting
Roux et al. [18] observed continuous bouncing of the conductive induced droplet jumping method as an effective tool for droplet
droplet between the electrode plates when a uniform electric transport relevant in microfluidic applications. In this paper, we
field is exerted. This uncontrolled back and forth motion of the apply high density ratio based lattice Botlzmann method (LBM)
droplet is controlled by coating the upper electrode with a [26] which is based on the Cahn–Hilliard diffuse interface theory
dielectric layer and by adjusting the electric field direction. The for binary fluids. One of the distinct feature of LBM is its easily par-
experimental investigations of Koji et al. [19] revealed that a allelizable nature which allows to model complex three dimen-
vertical electric field can induce a water droplet to jump up from sional simulations by employing advanced high performance
a superhydrophobic surface. However, the authors insisted on computing techniques [27]. This method [26] is based on the
the need of precise control of Coulomb forces to prevent droplet model of He et al. [28] which has been further developed to
splitting. account high density ratios through stable discretisation of the
Electrowetting has been established as one of the most widely forcing term. The model has been applied for various multiphase
used method for manipulation of liquid droplets [20]. In contrast flows problems such as droplet impact on wet walls [29,30], dry
to chemical and topographical modulations of the solid surface, walls [31], buoyancy induced interfacial mixing [32,33], droplet
electrowetting offers better contact angle variations and switching slipping under shear flows [34] and bubble rise under buoyancy
speeds. The droplets can be propagated freely on surfaces along [35].
programmable paths where they can be split, merged and mixed The remainder of the paper is organised as follows: Section 2
with higher degree of flexibility. Huang et al. [21] performed lattice provides the details of the multiphase LBM and the geometry based
Boltzmann simulations to study diverse droplet behaviours result- contact line model. Section 3 briefly outlines the theory of elec-
ing due to spatio-temporal variations in surface wettability. They trowetting while Section 4 describes the problem specification
found that under certain conditions, the droplet displays rapid con- and involved characteristic parameters. The discussion of the
tinuous unidirectional movement. Abdelgawad et al. [11] numerical results is provided in Section 5. Section 6 describes the
employed electrowetting technique to manipulate droplet on regime classification of EW actuated droplet jumping. In Section 7
inclined, twisted and upside-down geometries. More recently, we demonstrate the applicability on EW-induced droplet jumping
detachment of droplets from solid surfaces has been established for droplet transport and mixing. Finally, we provide concluding
through electrowetting [22–24]. Lee et al. [22] introduced droplet remarks in Section 8.
jumping by electrowetting and demonstrated the transport of ses-
sile droplets to upper surfaces under diverse electrode configura-
tions. Lapierre et al. [23] presented a technique to dewet droplets 2. Numerical method
on superhydrophobic surfaces through periodic vibrations induced
due to electrowetting actuation. This approach allowed them to In the present study, wettability induced droplet jumping
investigate a large of superhydrophobic surfaces for electrowetting dynamics have been simulated by employing a high-density ratio
without droplet impalement. Employing square pulse signals in multiphase lattice Boltzmann method based on the formulation
electrowetting actuations resulted in droplet detachment from proposed by Lee and Lin [26]. Building upon our previous work
hydrophobic surfaces [24]. When the pulse width is synchronous [30], we have extended the solver to three dimensions [36,31] in
with spreading time, it is found that the droplet is detached more order to investigate the dynamics of droplet jumping. The method
easily. employs two particle distribution functions to recover incompress-
The contributions of this study is threefold. Firstly, we attempt ible Navier–Stokes equation (g a ) and a macro interface capturing
to analyse the governing fluid dynamical aspects involved during equation (f a ). The model adopts stress and potential forms of inter-
wettability-induced droplet jumping. In addition to the experi- molecular forcing terms in the momentum equation and the equa-
mental studies mentioned above, the present numerical simula- tion for order parameter, respectively. Stable discretization
tions provide additional details of the dynamic flow field and schemes were proposed by Lee and Lin [26] to discretize the
various transient energy conversions which would help to forcing terms in each collision step which helped in improving
K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821 807

rffiffiffiffiffiffi
numerical stability for high density ratio cases. The governing 4 j
D¼ ð12Þ
distribution functions were solved in a three step process: ðqsat
l  qv Þ
sat 2b
(i) Pre-streaming step
 where j is a constant related to the magnitude of surface tension.
a 
g a  g eq The surface tension force r is represented as
ga ðx; tÞ ¼ g a ðx; tÞ 
2s ðx;tÞ
sat 3 pffiffiffiffiffiffiffiffiffi
ðqsat
l  qv Þ
dt ðeai  ui Þ@ i ðqc2s Þ r¼ 2jb: ð13Þ
þ ðCa ðuÞ  Ca ð0ÞÞjðx;tÞ 6
2 c2s
dt ðeai  ui Þ½j@ i ð@ k q@ k qÞ  j@ j ð@ j q@ i qÞ The density of the fluid q, hydrodynamics pressure p and the
þ Ca ðuÞjðx;tÞ ð1Þ velocity u are calculated by taking the moments of the correspond-
2 c2s
 ing distribution function:
eq

f a ðx; tÞ ¼ f ðx; tÞ  f a  f a  X
f a
a
2s  ðx;tÞ
q¼ ð14Þ
a
dt ðeai  ui Þ½@ i qc2s  q@ i ð/  j@ 2j qÞ X dt @ qc2s
þ Ca ðuÞjðx;tÞ ð2Þ p¼ ga c2s þ ui ð15Þ
2 c2s 2 @xi
a
X     
(ii) Streaming dt @ @q @q @ @q @q
ui ¼ g a ea þ j  ð16Þ
2 @xi @xk @xk @xj @xi @xj
gðx þ ea dt; t þ dtÞ ¼ ga ðx; tÞ ð3Þ a
f ðx þ ea dt; t þ dtÞ ¼ f a ðx; tÞ ð4Þ The relaxation parameter s is related to the kinematic viscosity
m ¼ sc2s dt, which can be calculated by a linear interpolation
(iii) Post-streaming step
s ¼ C sl  ð1  CÞsv ð17Þ
g a ðx þ ea dt; t þ dtÞ ¼ ga ðx þ ea dt; t þ dtÞ
where sl and sv are the relaxation times for liquid and vapour,
1
 ðga  geq
a Þjðxþea dt;tþdtÞ
respectively and C is the composition approximated by
2s þ 1
ðq  qsat
v Þ
2s dt ðeai  ui Þ@ i ðqc2s Þ C¼ ð18Þ
þ ðCa ðuÞ  Ca ð0ÞÞjðxþea dt;tþdtÞ ðqsat  q sat Þ
v
2s þ 1 2 c2s l

2s dt ðeai  ui Þ½j@ i ð@ k q@ k qÞ  j@ j ð@ j q@ i qÞ Along with the consistent discretizations of the intermolecular
þ Ca ðuÞjxþea dt;tþdt forcing terms, the aforementioned two-phase LBE formulation pro-
2s þ 1 2 c2s
vides necessary stabilization at high density and viscosity ratios.
ð5Þ While the mixed difference scheme is used in the pre-streaming
f a ðx þ ea dt; t þ dtÞ ¼ f a ðx þ ea dt; t þ dtÞ collision step, the second order central difference scheme is
1 considered for the forcing terms in the post-streaming collision
ðf a  f a Þjðxþea dt;tþdtÞ
eq
 step. Further details on the discretization schemes can be
2s þ 1
2 found in [26].
2s dt ðeai  ui Þ½@ i qc2s  q@ i ð/  j@ j qÞ
þ Ca ðuÞjðxþea dt;tþdtÞ ð6Þ For the wetting boundary condition, we employ the geometric
2s þ 1 2 cs 2
formulation proposed by Ding and Spelt [37]. This geometric
The equilibrium distribution functions are given by scheme has been also employed in the framework of LBM to inves-
" # tigate the dynamics of impacting droplet onto surface with varying
eq ea :u ðea :uÞ2 ðu:uÞ wettabilities [36]. The procedure essentially involves updating the
f a ¼ wa q 1 þ 2 þ  ð7Þ
cs 2c4s 2c2s values of the order parameter (density in the present model)
" # enforcing the following equation:
p qea :u qðea :uÞ2 qðu:uÞ p
g a ¼ wa 2 þ
eq
þ  ð8Þ
cs c2s 2c4s 2c2s n  $q ¼  tan  h j$q  ðn:$qÞnj ð19Þ
2
pffiffiffi
where cs ¼ 1= 3 and wa is the corresponding integral weights for a The developed 3D solver based on the present model has been
D3Q19 lattice velocity model: calibrated to evaluate fluid–fluid and fluid–solid interactions for
1 both static and dynamic scenarios. Various benchmark cases such
wa ¼ ; a¼0 as verification of Laplace law for a stationary bubble, evaluation
3
1 of equilibrium contact angle for a droplet resting on a solid surface
wa ¼ ; a 2 ½1; 6 and experimental comparison of the dynamic case of a droplet
18
1 impacting on a solid surface have been outlined in our previous
wa ¼ ; a 2 ½7; 18 ð9Þ paper [36]. In the present study, droplet actuation by electrowet-
36
ting is realised by varying the equilibrium contact angle. The next
and section outlines the details of the influence of applied voltage on
" # contact angle which is observed in electrowetting.
ea :u ðea :uÞ2 ðu:uÞ
Ca ðuÞ ¼ wa 1 þ þ  ð10Þ
c2s 2c4s 2c2s
3. Theory of electrowetting
The chemical potential / is given by
When a drop is placed in contact with a solid surface, it drives
1 sat
v Þðq  ql Þðq  2 ðqv þ ql ÞÞ
/  4bðq  qsat sat sat
ð11Þ towards its equilibrium configuration by capillary forces. The three
phase contact line is balanced by the surface tension forces
where b is a constant, and qsat sat
v and ql are the saturation densities between pairs of phases forming an equilibrium shape charac-
of the vapour and liquid phases, respectively. The interface thick- terised by the contact angle. This is determined by the Young’s
ness, denoted as D, is given by law given as:
808 K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821

rsg  rsl smaller than the capillary length. For the parameters correspond-
cos hY ¼ ð20Þ
r ing to water droplets, the Bond number is of the order 104 , hence
where r; rsg and rsl are the liquid–gas, solid–gas and solid–liquid the influence of gravity can be assumed to be small. During the
surface tensions, respectively. Lippmann’s original experiment dealt wettability-induced jumping process, the surface energy of the
with direct metal and electrolyte interface [38]. However, in such drop is converted into kinetic energy and a part of the energy is
situations, electrowetting based droplet actuation was applicable dissipated due to viscous effect. In order to investigate the tran-
for very low voltages. Increase in voltage would lead to onset of sient energy variations involved during the jumping process, we
the electrolytic process. In current systems employing electrowet- monitor them in the current study. The surface energy is obtained
ting for droplet control, the solid substrate is provided with a layer from the free energy density model in our simulations:
of dielectric material over which the droplet is placed. This is more Z 
commonly known as electrowetting on dielectric (EWOD) [20]. The ES ¼ Eo þ jj$qj2 dV ð24Þ
V
presence of an insulating dielectric layer greatly reduces the capac-
2 2
itance of the system. When the electric potential is applied, forma- where Eo ¼ bðq  qsat
v Þ ðq  ql Þ . The free energy density has two
sat

tion of a layer of charged ions on the dielectric surface and components. The first term Eo is the bulk energy related to the ther-
subsequently a layer of oppositely charged ions in the liquid close modynamics pressure by the equation of state. It models fluid com-
to the surface takes place. This fringing field at the corners of the ponent’s immiscibility and separates the fluid into two stable
electrolyte droplet tends to pull the droplet closer to the substrate phases. The second component in Eq. (24) is the gradient energy
and thereby reduces the contact angle. From a thermodynamic which accounts for the diffusive interfacial region, where j is a
point of view, the Gibbs free energy of a system is the minimum parameter associated with the surface tension along the interface.
energy requirement to create a certain area of that surface and it The dissipation function for a continuum fluid is defined as
consists of chemical and electrical components. The chemical com-  2  2  2
ponent accounts for the natural surface tension of the solid–liquid @ux @uy @uz
/ ¼ 2 þ2 þ2
interface without electric field. The relationship between the equi- @x @y @z
librium static contact angle in an electrowetting system and applied  2  2  2
@ux @uy @ux @uz @uy @uz
voltage V is expressed as þ þ þ þ þ þ ð25Þ
@y @x @z @x @z @y
r o V 2 The viscous dissipation is obtained as a time integration of the
cos hLY ¼ cos hY þ ð21Þ
2de r dissipation function
Z s
where hLY is the Lippmann–Young equilibrium contact angle, de is
the thickness of the dielectric layer, r is the relative dielectric con- Ed ¼ ql Oh/ dt ð26Þ
0
stant of the material and o is the dielectric constant of vacuum. The pffiffiffiffiffiffiffiffiffiffiffiffiffi
second term on the right-hand-side of the equation is known as the where Oh ¼ ll = ql rDo , is the Ohnesorge number defined as the
dimensionless electrowetting number (g), which is defined as the ratio of viscous forces to inertial and surface tension forces.
ratio of the strength of the electrostatic energy to the surface Fig. 1 shows the schematic representation of the problem setup.
energy. This number (g) is always positive due to the voltage- The wettability of the lower wall is controlled by electrowetting.
squared term, hence electrowetting always decreases the Lipp- When the voltage is applied, the contact angle decreases due to
mann–Young contact angle from the equilibrium contact angle. the reduction in interfacial surface tension. Periodic boundary con-
After providing a brief review on electrowetting theory, we next ditions are imposed on the sides of the domain, while the equilib-
proceed to outline the characteristic parameters governing the rium bounce back boundary conditions [39] are used on the top
droplet jumping process under electrowetting, which is the main and bottom boundaries to impose the no-slip boundary conditions.
subject of the present study. We quantify different characteristic It is to be noted that in the present work, we consider a constant
parameters of the electric pulse and the relevant scales for normal- contact angle model. A more realistic modelling approach would
isation followed by the details of the computational setup. be to employ dynamic contact angle models [40,41], which how-
ever is not the focus of the current study. While analysing the
4. Characteristic parameters and computational setup dynamics of droplet jumping through electrowetting, we examine
the following two key derived parameters:
A spherical droplet with diameter Do is initialized on a solid sur-
face such that the three phase contact line makes an angle hY = 140° (i) The contact diameter (Dc) which is defined as the distance
with the surface. Unless otherwise mentioned, the liquid–gas between the two three phase contact points on the bottom
density ratio (qr ) and the viscosity ratio (lr ) are set to be 1000 wall extracted from the mid y-plane.
and 50, respectively for subsequent calculations. However, the (ii) Average drop velocity (U Z ) along the z-direction, which is
effect of viscosity ratio on the impact dynamics is considered later given by:
in the subsection where influence of drop viscosity is investigated.
All the length scales are non-dimensionalised by the droplet diam-
eter (Do). The time and droplet velocity are non-dimensionalized
with capillary inertial time (sci ) and velocity uci as defined below:
sffiffiffiffiffiffiffiffiffiffi
ql D3o
sci ¼ ð22Þ
8r
sffiffiffiffiffiffiffiffiffiffi
2r
uci ¼ ð23Þ
ql Do
The size of the droplet of interest considered here, relevant to
microfluidic applications, is of the order of micrometers which is Fig. 1. Schematic representation for electrowetting induced droplet transport.
K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821 809

Table 1 considered for all subsequent sections, unless otherwise mentioned.


Numerical convergence for various droplet radii with Oh = 0.0126, qr = 1000, lr = 50 Table 1 shows U zmax for three different grid resolutions. A variation
and hLY = 60°.
of less that 1% in U zmax is observed for Do = 75. To achieve a confi-
Do (in lattice units) 50 75 100 dence in the grid resolution employed in this study, Fig. 2 presents
U Zmax 0.565 0.579 0.583 droplets shapes for three different grid sizes corresponding to
Relative error 3.08% 0.68% – Do = 50, 75 and 100. We notice that the interface profiles for
Do = 75 and 100 nearly approaches each other, except small devia-
tions at the bottom part of the droplet. It is worth mentioning that
given the diffusive nature of the interface in phase field methods,
resolution of the interface with enough mesh points plays an impor-
tant role to capture the droplet dynamics accurately. This would in
turn lead to a higher computational cost. However, advances in par-
allel adaptive mesh refinement techniques, which have been
applied to resolve complex 3D multiphase flows [42], can reduce
the excessive computational cost associated with the uniformly
refined mesh throughout the domain. Based on the grid conver-
gence test, the droplet diameter was set to be 75 lattice units for
the simulations preformed in this study hereafter.

5. Results and discussion

We next proceed to evaluate the effect of various parameters


like pulse characteristics, electrowetting number, drop viscosity
and properties of the ambient fluid on the dynamics of the jumping
droplet. Finally, we will demonstrate the application of electrowet-
ting based droplet actuation employed in microfluidic applications
involving droplet transport and mixing.

5.1. Influence of pulse characteristics


Fig. 2. Effect of grid resolution on the droplet interface profiles at a representative
non-dimensional time T⁄ = 4.5.
Unlike previous studies using sinusoidal or trapezoidal wave-
forms [43] for electrowetting actuation, we consider square wave-
R
ux dX forms in the present study. Square pulse actuation has been found
UZ ¼ R
Vdrop
ð27Þ to be more effective and controllable for droplet jumping [22].
Vdrop
dX
Fig. 3 shows the imposed square waveforms for (a) single pulse
and (b) double pulse actuations, respectively. It is assumed that
where X encloses the region where the density (q) is greater than the voltage is applied at T⁄ = 0. As the voltage is applied, the contact
q which corresponds to the mean density. Along with the above angle of the lower wall is set to be hLY . The duration during which
mentioned parameters, we also monitor the transient energy varia- the voltage is applied is known as pulse width (TP). After the dura-
tions as defined previously. We perform grid independence test by tion of TP, the voltage is turned off and the surface contact angle is
comparing the maximum vertical velocity of the droplet (U zmax ) for changed back to its initial value of hY . In the case of DC EW
Oh = 0.0179. A computational domain of 2Do  2Do  1:8Do was actuation, there is a constant application of voltage to the solid

160 160

θY θY

120 120

80 80

θLY θLY

40 40 TP TP
TP

0 0
0 1 2 3 4 5 6 7 8 0 1 2 3 4 5 6 7 8

T T∗
(a) (b)
Fig. 3. Evolution of imposed contact angle with square waveforms for (a) single pulse and (b) double pulse actuations.
810 K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821

substrate. During this phase, the droplet spreads and recoils to droplet undergoing the AC EW actuation is stretched to its
attain a configuration corresponding to the Lippmann–Young con- maximum spread, resulting in higher storage of surface energy
tact angle for the given voltage. However, in the case of AC EW inside the droplet system. As the contact angle changes at T⁄ = TP,
actuation, the voltage is applied for a given pulse width and then the droplet system begins to minimise its it free energy. The stored
turned off. During this period, the droplet is still in its spreading surface energy is converted into the kinetic energy resulting in
or receding phase and has not attained its equilibrium configura- increase in UZ. The droplet undergoing the AC EW actuation has
tion. Double pulse actuation implies actuation of two successive higher maximum UZ value due to increased conversion of the
AC pulses within a gap of the corresponding pulse width between stored surface energy into kinetic energy. The observations
them. indicate that under AC EW actuation, the droplet acquires greater
jumping velocities as compared to the DC EW actuation. Hence,
it would be viable to employ the AC EW based jumping for droplet
5.1.1. AC versus DC electrowetting actuation transport in microfluidic applications.
The spreading and jumping behaviour of a droplet subjected to
electrowetting (EW) depends on the nature of EW actuation. We 5.1.2. Effect of pulse width
consider droplet detachment by both DC and AC EW cases. Fig. 4 The duration of the applied voltage is determined by the pulse
illustrates the temporal evolution of the droplet detachment pro- width (TP), which in turn governs the spreading behaviour of the
cess for (a) DC and (b) AC EW actuation scenarios. For the AC
EW, a square pulse with TP = 2.22, corresponding to the time taken
by the droplet to spread to its maximum extent is considered. The
spreading droplet is allowed to attain its equilibrium shape corre-
sponding to hLY = 60° until T⁄ = 8.27, after which the contact angle
is changed to hY = 140° in the case of DC EW. At T⁄ = 7.44, the inter-
face profile nearly attains its equilibrium shape as shown in Fig. 4
(a). As the surface wettability changes to h = 140°, the droplet base
begins to decrease sharply adjusting to the corresponding local
curvature. As the base radius continues to decrease, an inversion
in droplet morphology is observed from a broad base-narrow top
configuration (T⁄ = 8.44) to a narrow base-broad top system
(T⁄ = 10.59) owing to mass conservation. The base radius continues
to decrease leading to droplet detachment from the substrate
(T⁄ = 12.58). The sequence of events leading to droplet detachment
from the surface for AC EW actuation are similar to that of
observed for DC EW actuation. However, the notable difference
between the two situations is the interface profiles of the droplet
at the instant when the surface contact angle is changed to hY .
Unlike the case for the DC EW actuation, the droplet is in its max-
imum spread resembling a thin pancake like shape as shown at
T⁄ = 2.31. The droplet shape undergoes greater vertical elongation
in comparison to the corresponding droplet shapes for DC EW
actuation. The last snapshots in Fig. 4(a) and (b) depicts the droplet
shapes at DT  = 4.47 and 4.24 (DT  ¼ T  – TP), respectively. We
observe that the droplet in Fig. 4(b) lifts off to a greater height than
that in Fig. 4(a). This illustrates the presence of higher kinetic
energy for droplets detaching due to the AC EW actuation.
The kinetics of the above depicted droplet detachment beha-
viour is illustrated by the temporal evolution of the base diameter
(DC), as shown in Fig. 5. Fig. 5(a) and (b) outline the evolution of DC
during spreading and recoiling phases, respectively. For the DC EW
actuation, the base diameter DC increases initially to reach its max-
imum spread and undergoes recoiling during which it decreases.
Eventually, it attains a steady value corresponding to hLY . The evo-
lution of DC during the spreading phase shows a steady increase in
its value. When the surface contact angle is changed to hY at T⁄ = TP,
the droplet begins to recede leading to decreasing DC . We observe
earlier droplet detachment for the AC EW actuation. This difference
in droplet jumping behaviour can be explained by the surface
energy stored inside the droplet system at T⁄ = TP. When the dro-
plet behaviour is observed under the DC EW actuation, after attain-
ing its maximum spread it undergoes recoiling, during which some
amount of the surface energy is converted into the kinetic energy
of the recoiling droplet. This can be noticed from the evolution of
UZ, as shown in Fig. 6.
After DT⁄ = 6.0, we observe a gradual decrease in UZ before it
eventually becomes zero as the droplet attains its equilibrium con-
figuration at T⁄ = TP. The surface energy stored inside the droplet Fig. 4. Temporal evolution of the droplet detachment process for (a) DC EW
corresponds to its equilibrium shape. However, at T⁄ = TP the actuation and (b) AC EW actuation with hLY = 60°.
K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821 811

droplet. Hence, it is important to investigate the role of TP on the T⁄ = 4.44, the wettability of the surface changes to hLY and the sec-
dynamics of droplet detachment. To investigate the influence of tion of the droplet in contact with the surface begins to spread.
TP, we set hLY = 60° and define a parameter TPr = (TP/T P60 ), where However, as a result of the actuation of the previous pulse, the bulk
T P60 is the time taken for the droplet to spread to its maximum of the droplet recoils in the upward direction owing to its inertia.
extent for hLY = 60°. Fig. 7(a) illustrates the temporal evolution of This results in the formation of a neck region between the uprising
Dc for three different TPr. droplet bulk and the spreading contact region. This neck region is
It shows that Dc increases with increase in TPr. For TPr = 0.5, the clearly observed for all the cases, as shown in Fig. 9. However,
surface contact angle is changed before the droplet attains its max- the length of the neck is found to increase with decreasing hLY .
imum spread and subsequently its lifts off the substrate earlier As the contact region continues to spread and the bulk droplet con-
compared to the case with TPr = 1.5. The droplet attains its maximum tinues to rise, the neck region elongates and eventually breaks and
velocity in Z-direction for TPr = 1.0 as observed from Fig. 7(b). Since forms a satellite droplet resting on the surface. This volume of the
for TPr = 1.0, the contact angle is changed to hY when the droplet satellite droplet increases with decrease in hLY as shown in the bot-
is in its maximum spread, the surface energy converted into the tom row of Fig. 9. This can be attributed to the increase in contact
kinetic energy of the recoiling droplet is also higher. However, diameter at lower hLY which lead to greater adhesion forces
for TPr = 0.5 and 1.5, the contact angle is changed during the capil- between the droplet and the solid substrate. The current results
lary spreading and receding phases, respectively. Fig. 8 illustrates show qualitative similarities with the experimental investigation
the velocity field inside the droplet and the surrounding fluid near
the interface for (a) TPr = 0.5 and (b) TPr = 1.5. As the droplet
interface curvature at the three phase contact line changes due
to change in contact angle, we observe flow circulation near this 1
region for both the cases as shown as in Fig. 8. However, the
notable differences in the velocity field are observed in the upper
0.8
region inside the droplet. For TPr = 0.5, we observe the velocity DC EW actuation
vectors pointing radially downwards due to the inertia attained
by the droplet liquid during capillary spreading until T⁄ = TP. The 0.6 AC EW actuation
drag induced on the surrounding fluid by this downward moving
droplet interface sets up a vortex near the upper region. Due to
0.4
the opposing sense of movement of the interface near the contact Spreading
UZ

line and the upper bulk region, the corresponding vortex patterns
move in anti-clockwise and clockwise directions, respectively. 0.2
The velocity vectors in the upper bulk region move in radially
upward direction and have the same sense of motion with the flow
0
field near the contact region as observed in Fig. 8(b). However, the Receding
velocity magnitude near the contact region is higher compared to
the velocity magnitude in the bulk region. -0.2

5.1.3. Double pulse actuation -0.4


We next investigate the influence of double pulse actuation on -8 -6 -4 -2 0 2 4
the dynamics of droplet detachment. Two square pulses with a ΔT∗
pulse width TP = 2.22 were triggered at equal intervals. Fig. 9 illus-
Fig. 6. Time evolution of Z-component of drop velocity for DC and AC electro-
trates the droplet detachment process at two different time
wetting actuation: hLY = 60°.
instants for (a) hLY ¼ 40 (b) hLY ¼ 60 and (c) hLY ¼ 80 . After

4 4

3.5
3.5

3 DC EW actuation
3 AC EW actuation
2.5
Dc
Dc

2.5 2

1.5
2

DC EW actuation 1
AC EW actuation
1.5
0.5

1 0
-8 -6 -4 -2 0 0 0.5 1 1.5 2 2.5 3

ΔT T∗ - TP
(a) (b)
Fig. 5. Time evolution of droplet contact diameter for DC and AC electro-wetting actuation: hLY = 60°.
812 K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821

conducted by Lee et al. [24]. They observed the formation of satel- corresponding oscillations in ES as time proceeds. These oscilla-
lite droplet when the droplet is actuated using the double square tions in surface energy are greater for lower hLY . Time variations
pulse. of viscous dissipation shows an increase in ED as hLY decreases. This
observation can be attributed to the fact that with increase in dro-
plet deformation, the resistance offered by the viscous stresses to
5.2. Effect of Lippmann–Young contact angle the deforming fluid increases. For hLY = 40°, the droplet undergoes
sharp spreading and intense recoiling. Hence, a sharp rise in Ed is
Apart from the nature of actuating pulse, the behaviour of the noticed before it eventually lifts off the surface.
jumping droplet also depends on the applied voltage. The extent
to which the droplet spreads depends on the applied voltage,
which in turns governs the Lippmann–Young contact angle (hLY ). 5.3. Effect of drop viscosity
We next investigate the influence of hLY on the droplet jumping
behaviour. Fig. 10(a) presents the time evolution of the contact The mobility of the droplet during its spreading and receding
diameter for different hLY . The pulse width TP is fixed to be 2.22 phases is hindered by viscous forces. Hence, it will be interesting
for all the simulations considered in this subsection. We observe to investigate the effect of viscosity on the jumping behaviour of
that the rate of evolution of Dc increases as hLY decreases. After the droplet subjected to electrowetting. With increase in Ohne-
the contact angle changes to hY at T⁄ = TP, a sharp decline in Dc is sorge number, viscous effects begin to dominate the dynamics of
noticed which eventually converges to zero as the droplet lifts droplet jumping. We perform numerical simulations for Oh num-
off the surface. The present numerical results do not indicate a gen- bers ranging from 0.0179 to 0.179. Fig. 12 illustrates the time
eral trend in the lift-off time with hLY . While the cases with resolved images of the droplet behaviour which corresponds to
hLY = 40° and 100° lift off nearly at the same time, the cases consid- those shown in Fig. 4(b), except that the Oh number considered
ered with intermediate hLY detach earlier. The lift-off time for the in this case is larger (Oh = 0.179). The figure clearly indicates the
droplet for different hLY depends on the interplay between the sur- influence of increased viscous effects leading to suppression of dro-
face energy stored in the stretched droplet and the magnitude of Dc plet jumping in contrast to the inertia dominated process shown in
at T⁄ = TP. Since the droplet is stretched to a greater extent, it has Fig. 4(b). At T⁄ = 0.99, the droplet has a cusp-like structure on the
higher surface energy stored inside the droplet system for top as the base continues to expand radially for Oh = 0.0179. This
hLY = 40° when compared to the droplet with hLY = 100°. Hence, results due to the upward propagation of the capillary wave as
the droplet retracts faster due to its increased kinetic energy. How- the droplet spreads across the surface. However, increase in dro-
ever, as it covers a larger contact area unlike the slowly retracting plet viscosity leads to higher viscous damping resulting in a
droplet with hLY = 100°, the lift-off time for both the cases is nearly smooth droplet profile for Oh = 0.179. The droplet undergoes
same. This behaviour is indicated from the temporal evolution of higher radial expansion before it begins to recede after T⁄ = TP for
UZ shown in Fig. 10(b). We observe increase in UZ as hLY decreases Oh = 0.0179 as observed from the corresponding images at
due to a larger conversion of surface energy into the kinetic energy T⁄ = 2.31. While the droplet detaches and lifts off the surface in
of the droplet system. Fig. 11 shows the transient variation of (a) the last two snapshots shown in Fig. 4(b), the droplet recedes
surface energy (ES ) and (b) viscous dissipation (Ed ) for different and undergoes secondary spreading for the corresponding time
hLY . As discussed above, for hLY = 40° the droplet undergoes large instants observed in Fig. 12.
deformation from its initial configuration. This excess deformation To further illustrate the effects of viscosity the temporal
results in sharp rise in ES compared to the other two cases consid- evolution of instantaneous contact diameter and drop velocity
ered. The transient variation of ES also reveal the oscillations which (UZ) are shown in Fig. 13. Before T⁄ = TP, we observe that both Dc
the drop undergoes after it lifts-off the surface. As the droplet and the maximum contact diameter increases with decrease in
morphology changes from prolate to oblate shapes, we observe Oh. However, a reversal in this trend is observed as the droplet

3.5 0.8
TP r = 0.5
TP r = 1.0
3
TP r = 1.5

2.5
0.4
Dc

2
UZ

1.5
0

1 TP r = 0.5
TP r = 1.0
0.5 TP r = 1.5
-0.4
0
0 1 2 3 4 5 0 1 2 3 4 5
T∗ T∗
(a) (b)
Fig. 7. Time evolution of droplet (a) contact diameter and (b) Z-component of drop velocity for different pulse width (TPr): hLY = 60°.
K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821 813

impedes droplet deformation. In contrast to the other two


Z
Velocity Magnitude
cases, the droplet with Oh = 0.0179 continues to oscillate after it
0.03
0.025 lifts off the surface. Fig. 14(b) further asserts the role of drop
0.02
0.015 X
Y viscosity resulting in increase in viscous dissipation with
0.01 increasing Oh.
0.005

5.4. Effect of the properties of ambient fluid

The role of the ambient fluid on the behaviour of the jumping


drop becomes important after the droplet detaches from the sur-
face. Hence, it is important to understand the influence of the prop-
erties of the surrounding fluid on the droplet jumping dynamics.
Mukherjee and Abraham [29] investigated the influence of sur-
rounding gas density and viscosity on the crown dynamics of the
splash formed when single droplet impacts a thin film. They
observed a delay in crown breakup and decrease in rate of increase
of crown radius and height with increase in surrounding gas den-
sity. They attributed this to the effect of increased drag imposed by
the surrounding medium on the crown. We first investigate the
influence of gas density (qg ) by varying the density ratio,

(a) TP r = 0.5 qr ¼ ðql =qg Þ. The other parameters are kept fixed as those consid-
ered for the case shown in Fig. 4(b). The kinematic viscosity ratio is
kept constant at 0.05. It is expected that at high gas densities, the
Z
Velocity Magnitude influence of gas on the jumping behaviour will be greater due to
0.014 the increased effect of drag on the droplet by the surrounding med-
0.012
0.01 X Y ium. This is illustrated from the time evolution sequence of the
0.008
0.006 droplet shapes shown in Fig. 15 for qr = 20 (top row), qr = 100
0.004
0.002
(middle row) and qr = 1000 (bottom row). As the droplet begins
to recede and attain a spherical shape due to surface tension forces
after T⁄ = TP, the drag induced on the droplet increases with
increase in qg . At T⁄ = 4.96 we observe that while the droplet is
detached from the surface for qr = 100, the droplet is still in contact
with the surface for qr = 20. The droplet undergoes greater shape
deformation for qr = 100 compared to the nearly spherical droplet
shapes for qr = 20 at T⁄ = 5.46 and 6.12. It is also noticed that at the
same time instant the droplet jumps to a greater height when the
surrounding fluid is less dense due to the lower resistance to the
movement of the droplet as it penetrates a lesser dense ambient
fluid. Next, let us consider the evolution of Dc and UZ. In Fig. 16
(a), the evolution of Dc is compared for different density ratios. It
is observed that during the initial capillary driven spreading phase
the evolution of Dc is fairly independent of the gas density qg . As
time proceeds, with decrease in qr , i.e increase in the gas density,
(b) TP r = 1.5
the rate of increase in Dc decreases. The lift-off time is found to
Fig. 8. Velocity field inside the droplet for (a) TPr = 0.5 at T⁄ = 1.32 and (b) TPr = 1.5 at increase as qr decreases. We notice that for qr = 10 the droplet does
T⁄ = 3.44. not detach from the surface and undergoes secondary spreading.
The influence of qg on Uz is shown in Fig. 16(b). The maximum
Uz attained by the droplet increases with increase in qr which reit-
erates the influence of the drag induced by the denser surrounding
continues to recoil and Dc decreases. This can be explained by the fluid on the jumping droplet. To investigate the influence of ambi-
increased viscous resistance to the movement of droplet liquid ent fluid viscosity (lg ), we present simulation results for three dif-
during the spreading and receding stages with increase in Oh num- ferent viscosity ratio (lr ): 15, 30 and 50. It is to be noted that
ber. For the cases considered, apart from the case with Oh = 0.179, changing lr corresponds to changing sg while other simulation
we observe the outcome of droplet detachment. The lift-off time is parameters are kept constant. Fig. 17 illustrates the influence of
found to increase with increase in Oh. The delay in lift-off time and viscosity ratio on the temporal evolution droplet profiles. We
suppression of droplet jumping can be attributed to decrease in observe that the droplet height for lr = 15 is nearly constant in
surface energy stored inside the droplet at T⁄ = TP as Oh is the shown snapshots and the variation is observed only in the bot-
increased. Similarly, the amount of surface energy, which is con- tom region of the droplet due to the receding motion of the contact
verted into kinetic energy, lost due to viscous dissipation increases line. This elucidates the resistance offered by the viscous forces of
with increase in Oh. This results in decrease in the evolution of UZ the ambient fluid on the recoiling droplet surface. However, for
as observed from Fig. 13(b). lr = 30, we observe that inertial forces attained by the recoiling
Fig. 14(a) illustrates the transient evolution of surface energy droplet as a result of the conversion of surface energy into kinetic
for different Oh numbers. With increase in Oh, the temporal evolu- energy is large enough to overcome the resistance offered by the
tion of ES decreases. Increase in droplet viscosity results in higher ambient fluid. Subsequently, this results in droplet detachment
viscous resistance to the moving fluid element which in turn and jumping. With increase in lr , the rate of evolution of Dc and
814 K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821

UZ decreases and increases, respectively as shown in Fig. 18. liquid mass inside the droplet moves downwards under the action
Increased ambient fluid viscosity results in increased viscous drag of capillary forces leading to radial expansion of the droplet. A cir-
on the receding droplet. culatory flow pattern sets up around the droplet due to the drag
induced by the downward moving droplet interface on the sur-
rounding fluid as shown at T⁄ = 0.99 in Fig. 19(a). The next stage
6. Kinematic stages of EW-actuated jumping droplet (II) is the curvature switching stage, during which the local interface
curvature near the contact line changes due to switching of the
To further classify regimes of the jumping process shown in contact angle from hLY to hY . The sudden increase in local curvature
Fig. 4(b), temporal evolution of droplet velocity field is shown in leads to greater accumulation of droplet liquid inside the periph-
Fig. 19. We partition the entire jumping process into four stages: eral rims. This creates large surface tension forces resulting in
The first stage is (I) the capillary spreading stage during which the accelerated radially inward flow. The sharp change in direction in

a-1 b-1 c-1

a-2 b-2 c-2


Fig. 9. Time evolution images of EW process with double pulse actuation for (a) hLY = 40 (b) hLY = 60 and (c) hLY = 80 .

5 1.5

4.5 θLY = 40◦


θLY = 40◦
θLY = 60◦
4 θLY = 60◦
θLY = 80◦ 1
θLY = 80◦
3.5 θLY = 100◦
θLY = 100◦
3
UZ
Dc

2.5 0.5

1.5
0
1

0.5

0 -0.5
0 1 2 3 4 5 0 1 2 3 4 5

T T∗
(a) (b)
Fig. 10. Time evolution of droplet (a) contact diameter and (b) Z-component of drop velocity for different hLY .
K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821 815

the movement of the contact line leads to the formation of a vortex liquid velocity inside the droplet is higher in the upper region of
pattern near the contact edge as observed from Fig. 19(b). The liq- the droplet compared to the base region as observed in Fig. 19
uid inside the droplet continues to move in radially upward direc- (c). This results in vertical elongation of the droplet and the contact
tion under the restraining influence of surface tension forces. The diameter continues to decrease resulting in the detachment stage

0.003

16

0.002
θLY = 40◦
θLY = 80◦
E∗S

E∗d
12
θLY = 100◦

0.001
θLY = 40◦
θLY = 80◦
8
θLY = 100◦

0
0 2 4 6 8 0 2 4 6 8
∗ ∗
T T
(a) (b)
Fig. 11. Time evolution of droplet (a) surface energy (ES ) and (b) viscous dissipation (Ed ) for different hLY .

Fig. 12. Time evolution images of EW process identical to those of Fig. 4(b) except for a larger Oh = 0.179.

4 1

Oh = 0.0179
3 Oh = 0.0895
Oh = 0.1342
Oh = 0.179
UZ
Dc

1 Oh = 0.0179
Oh = 0.0895
Oh = 0.1342
Oh = 0.179
0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
T∗ T∗

(a) (b)
Fig. 13. Time evolution of droplet (a) contact diameter and (b) Z-component of drop velocity for different Ohnesorge number (Oh): hLY = 60°.
816 K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821

0.004

14

0.003

12 Oh = 0.0179
Oh = 0.1342

Ed∗
E∗S

0.002
Oh = 0.179

Oh = 0.0179
10
Oh = 0.1342 0.001
Oh = 0.179

8 0
0 2 4 6 8 0 2 4 6 8
T∗ T∗
(a) (b)
Fig. 14. Time evolution of droplet (a) surface energy (ES ) and (b) viscous dissipation (Ed ) for different Ohnesorge numbers.

T∗ = 4.96 T∗ = 5.46 T∗ = 6.12 T∗ = 7.61


Fig. 15. Time evolution images of EW process with different gas densities for qr = 20 (top row), qr = 100 (middle row) and qr = 1000 (bottom row).

(III) of the jumping process (Fig. 19(d)). The high magnitude of the 7. Application to EW-induced droplet jumping for transport and
velocity vectors in the lower elongated region, which forms a cusp- mixing
like shape, shown in Fig. 19(d) indicate the fast retraction of the
lower region due to interfacial tension. In the final stage (IV), the Manipulation and transport of droplet is an essential con-
droplet velocity continues to decrease due to drag induced by stituent in digital microfluidic applications. In this subsection, we
the ambient fluid and it attains a spherical shape after some investigate the transport of a droplet subjected to electrowetting,
oscillations. from the bottom surface to the top surface. Fig. 20 illustrates
K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821 817

4 1

ρr = 10
3 ρr = 20
ρr = 40
0.5
ρr = 100
ρr = 1000
Dc

UZ
2

0 ρr = 10
1
ρr = 20
ρr = 40
ρr = 100
ρr = 1000
0 -0.5
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
T∗ T∗
(a) (b)
Fig. 16. Time evolution of droplet (a) contact diameter and (b) Z-component of drop velocity for different gas densities at hLY = 60°.

T∗ = 3.64 T∗ = 4.63 T∗ = 5.62 T∗ = 7.11


Fig. 17. Time evolution images of EW process for two representative viscosity ratios lr = 15 (top row) and lr = 30 (bottom row).

4 1

μr = 15
3 μr = 30
μr = 50
UZ
Dc

1
μr = 15
μr = 30
μr = 50

0
0 1 2 3 4 5 6 7 0 1 2 3 4 5 6 7
T∗ T∗
(a) (b)
Fig. 18. Time evolution of droplet (a) contact diameter and (b) Z-component of drop velocity for different gas viscosities with qr = 1000: hLY = 60°.
818 K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821

Fig. 19. Instantaneous velocity field along the symmetry Y-plane at different time instants: hLY = 60°.

droplet transport from the bottom surface to the top plate with Owing to the low confinement ratio, the upper region of the recoil-
confinement ratios (CR) 1.0 (top row) and 2.0 (bottom row). The ing droplet attaches to the upper surface leading to the formation
confinement ratio is defined as the ratio of the distance between of a liquid bridge. As time proceeds, the droplet contact area on the
the two plates and the droplet diameter. The wettability of the upper surface increases while the one on the bottom surfaces
upper surface is fixed at h = 90°. For CR = 1.0, we observe that the decreases. This is attributed to the higher wettability of the upper
droplet is transferred to the upper plate by forming a liquid bridge. plate compared to the bottom surface. This is observed from the
K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821 819

reversal in broad base and narrow top layout of the liquid bridge at Certain digital microfluidic applications require the transport
T⁄ = 3.33 to the narrow base and broad top layout at T⁄ = 4.46. Once and merging of droplets in order to perform a chemical
transported to the upper surface, the droplet attains its equilibrium reaction. In these applications, each droplet which contains a
configuration corresponding to the wettability of the upper wall. chemical reagent, interacts and coalesces to produce precipi-
This way of droplet transport through formation of a liquid bridge tates a the result of the chemical reaction. Droplet manipulation
is termed as the droplet transport in mode A. With increase in con- through electrowetting can be employed to actuate such
finement ratio, the droplet transport occurs distinctly different ter- operations leading to chemical analysis and synthesis. We next
med as mode B, where we observe that the droplet detaches demonstrate various steps involved during this process in
completely from the substrate and is airborne (T⁄ = 5.79). Subse- Fig. 21. To visualise the internal flow evolution during droplet
quently, it is transported to the upper surface as observed in the impact and coalescence, different coloured passive tracer
snapshots shown in the bottom row of Fig. 20 for CR = 2.0. particles are seeded inside the two droplets. The evolution of

T∗ = 3.33 T∗ = 3.8 T∗ = 4.46 T∗ = 5.95

T∗ = 5.79 T∗ = 6.78 T∗ = 7.44 T∗ = 9.26


Fig. 20. Temporal sequence of vertical droplet transport on upper plate for mode A with confinement ratios of 1.0 (top row) and mode B for CR = 2.0 (bottom row). The upper
plate has a fixed contact angle of h = 90°.

T∗ = 0.66 T∗ = 3.80 T∗ = 5.95 T∗ = 8.27

T∗ = 8.93 T∗ = 9.76 T∗ = 10.59 T∗ = 12.41


Fig. 21. Temporal sequence of droplet transport and merging with a sessile drop situated on the upper plate with confinement ratio of 2.5. The upper plate has a fixed contact
angle h = 90°.
820 K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821

these particles were followed using a trilinear interpolation of The results and findings indicate that electrowetting based dro-
the velocity field from the surrounding lattice nodes and their plet lift-off method can be used as a useful tool in microfluidic
locations were updated after every time step using the fourth- applications. The effect of gravity has been neglected in this study
order Runge–Kutta method. Particles going out of the droplet which will induce additional complex dynamics into large jumping
due to numerical errors were omitted for the visualisation droplets. This remains to be explored in future study.
purpose. Consider two droplets required to undergo chemical
reaction situated on the upper and bottom plates as shown in Acknowledgement
Fig. 21. The droplet on the bottom plate begins to spread
(T⁄ = 0.66) due to EW actuation. With the contact angle of the One of the authors was supported by the Ministry of Education
bottom plate changing from h = 60° to 140°, the droplet recoils sponsored research Scholarship of the National University of
(T⁄ = 3.80) and lifts-off from the bottom plate (T⁄ = 5.95). As Singapore.
soon as the jumping drop interacts with the sessile droplet
resting on the upper plate (T⁄ = 8.27), capillary forces come into References
play and initiate coalescence between the two drops. As the
drops continue to coalesce, the liquid bridge formed due to [1] K.M. Wisdom, J.A. Watson, X. Qu, F. Liu, G.S. Watson, C.H. Chen, Self-cleaning of
the contact between the drops begins to expand, resulting in super-hydrophobic surfaces by self-propelled jumping condensate, Proc. Natl.
Acad. Sci. U.S.A. 110 (2013) 7992–7997.
the formation of a column-like structure as shown at [2] J.B. Boreyko, C.H. Chen, Self-propelled dropwise condensate on super-
T⁄ = 8.93. The droplet ensemble spreads (T⁄ = 10.59) and recoils hydrophobic surfaces, Phys. Rev. Lett. 103 (184501) (2009).
(T⁄ = 12.41) to attains its equilibrium shape on the top plate. [3] K. Rykaczewski, A.T. Paxon, S. Anand, X. Chen, Z. Wang, K.K. Varanasi,
Multimode multidrop serial coalescence effects during condensation on
Particle based visualisation reveals absence of mixing between hierarchical super-hydrophobic surfaces, Langmuir 29 (2013) 881–891.
the two merging drops. This can be attributed to the lack of [4] R.D. Narhe, W. González-Vinas, D.A. Beysens, Water condensation on zinc
sufficient stretching and folding of the liquid interface required surfaces treated by chemical bath deposition, Appl. Surf. Sci. 256 (2010) 4930–
4933.
for internal mixing. Once the chemical reaction is done, the [5] C. Dietz, K. Rykaczewski, A.G. Fedorov, Y. Joshi, Visualization of droplet
droplet system can be deposited back on the bottom plate departure on a superhydrophobic surface and implications to heat transfer
through EW actuation. enhancement during dropwise condensation, Appl. Phys. Lett. 97 (033104)
(2010).
[6] J. Cheng, A. Vandadi, C.L. Chen, Condensation heat transfer on two-tier
superhydrophobic surfaces, Appl. Phys. Lett. 101 (131909) (2012).
8. Concluding remarks [7] N. Miljkovic, R. Enright, Y. Nam, K. Lopez, N. Dou, J. Sack, E.N. Wang, Jumping-
droplet-enhanced condensation on scalable superhydrophobic nanostructured
surfaces, Nano Lett. 13 (2013) 179–187.
In this work we employed a high-density ratio based lattice [8] J.B. Boreyko, Y. Zhao, C.H. Chen, Planar jumping-drop thermal diodes, Appl.
Boltzmann method and performed three dimensional simulations Phys. Lett. 99 (234105) (2011).
to investigate the evolution dynamics of electrowetting-induced [9] J.B. Boreyko, C.H. Chen, Vapor chambers with jumping-drop liquid return from
superhydrophobic condensers, Int. J. Heat Mass Transfer 61 (2013) 409–418.
droplet jumping. A geometry based contact angle formulation [10] R.B. Fair, Digital microfluidics: is a true lab-on-a-chip possible? Microfluid.
has been used to model the three-phase contact line. The effect Nanofluid. 3 (2007) 245–281.
of voltage pulse characteristics, droplet viscosity and fluid [11] M. Abdelgawad, S.L. Freire, H. Yang, A.R. Wheeler, All-terrain droplet actuation,
Lab Chip 8 (2008) 672–677.
properties of the surrounding medium on the droplet jumping [12] F.C. Wang, F. Yang, Y.P. Zhao, Size effect on the coalescence-induced self-
was investigated. Details of transient energy variations among propelled droplet, Appl. Phys. Lett. 98 (053112) (2011).
surface energy and viscous dissipation were provided. The main [13] S. Farokhirad, J.F. Morris, T. Lee, Coalescence-induced jumping of droplet:
inertia and viscosity effects, Phys. Fluids 27 (102102) (2015).
conclusions can be summarised as follows: [14] X. Liu, P. Cheng, X. Quan, Lattice Boltzmann simulations for self-propelled
jumping of droplets after coalescence on a superhydrophobic surface, Int. J.
(i) The jumping behaviour has been found to be strongly depen- Heat Mass Transfer 73 (2014) 195–200.
[15] Y. Shi, G.H. Tang, H.H. Xia, Investigation of coalescence-induced droplet
dent on the nature of electrowetting actuation. A pulse width jumping on superhydrophobic surfaces and liquid condensate adhesion on slit
corresponding to the time taken for maximum spread would and plain fins, Int. J. Heat Mass Transfer 88 (2015) 445–455.
lead to higher maximum jumping velocities when compared [16] Ho-Young Kim, Drop fall-off from the vibrating ceiling, Phys. Fluids 16 (2)
(2006).
to other pulse widths. Formation of satellite droplets were
[17] A.J. James, B. Vukasinovic, M.K. Smith, A. Glezer, Vibration-induced drop
observed for the cases involving double pulse actuation. atomization and the numerical simulation of low-frequency single-droplet
(ii) Increase in applied voltage provided greater initial droplet ejection, J. Fluid Mech. 476 (2003) 29–62.
spreading. This resulted in higher transient variations of [18] J.M. Roux, Y. Fouillet, J.L. Achard, 3D droplet displacement in microfluidic
systems by electrostatic actuation, Sensors Actuators A 134 (2007) 486–493.
the surface energy and the viscous dissipation. Subse- [19] K. Takeda, A. Nakajima, K. Hashimoto, T. Watanabe, Jump of water droplet
quently, higher lift-off velocities were observed for the cases from a super-hydrophobic film by vertical electric field, Surf. Sci. Lett. 519
with lower hLY . (2002) 589–592.
[20] F. Mugele, J.C. Baret, Electrowetting: from basics to applications, J. Phys.:
(iii) Depending upon the kinetic evolution of the three-phase Condens. Matter 17 (2005) 705–774.
contact line, the entire droplet jumping process is classified [21] J.J. Huang, C. Shu, Y.T. Chew, Numerical investigation of transporting droplets
into four regimes, namely: capillary spreading, curvature by spatiotemporally controlling substrate wettability, J. Colloid Interface Sci.
328 (2008) 124–133.
switching, detachment, and spherical equilibrium. [22] S.J. Lee, S. Lee, K.H. Kang, Droplet jumping by electrowetting and its
(iv) The density and viscosity of the surrounding medium was application to the three-dimensional digital microfluidics, Appl. Phys. Lett.
found to cause a significant influence in determining the 100 (081604) (2012).
[23] F. Lapierre, Y. Coffinier, R. Boukherroub, V. Thommy, Electro-(de)wetting on
final outcome droplet jumping. For higher gas densities, an superhydrophobic surfaces, Langmuir 29 (2013) 13346–13351.
increase in droplet lift-off time was observed and in some [24] J. Hong, Y.K. Kim, D.J. Won, J. Kim, S.J. Lee, Electrowetting-induced droplet
case droplet jumping was suppressed. detachment from hydrophobic surfaces, Langmuir 29 (2013) 13346–13351.
[25] F. Liu, G. Ghigliotti, J.J. Feng, C.H. Chen, Numerical simulations of self-propelled
(v) For electrowetting driven droplet transport, two different
jumping upon drop coalescence on non-wetting surfaces, J. Fluid Mech. 752
modes of transport were identified based on the confine- (2014) 39–65.
ment ratio. Mode A corresponds to droplet transport through [26] T. Lee, C.L. Lin, A stable discretization of the lattice Boltzmann equation for
the formation of a liquid bridge. In mode B, the droplet is simulation of incompressible two-phase flows at high density ratio, J. Comput.
Phys. 206 (2005) 16–47.
completely detached from the bottom surface and is air- [27] S.P. Vanka, A.F. Shinn, K.C. Sahu, Computational fluid dynamics using graphics
borne before it is transported to the upper plate. processing units: challenges and opportunities, in: ASME 2011 International
K. Ashoke Raman et al. / International Journal of Heat and Mass Transfer 99 (2016) 805–821 821

Mechanical Engineering Congress and Exposition, Denver, Colorado, USA, [36] K.A. Raman, R.K. Jaiman, T.S. Lee, H.T. Low, Lattice Boltzmann simulations of
November 11–17, 2011, ASME. droplet impact onto surfaces with varying wettabilities, Int. J. Heat Mass
[28] X. He, X. Shan, R. Zhang, A lattice Boltzmann scheme for incompressible Transfer 95 (2016) 336–354.
multiphase flow and its application in simulation of Rayleigh–Taylor [37] H. Ding, P.D.M. Spelt, Wetting condition in diffuse interface simulations of
instability, J. Comput. Phys. 152 (1999) 642–663. contact line motion, Phys. Rev. E 75 (046708) (2007).
[29] S. Mukherjee, J. Abraham, Crown behavior in drop impact on wet walls, Phys. [38] M.G. Lippmann, Relation entre les phénomènes électriques et capillaires, Ann.
Fluids 19 (2007). Chim. Phys. 495 (1875).
[30] K.A. Raman, R.K. Jaiman, T.S. Lee, H.T. Low, On the dynamics of crown structure [39] T. Lee, L. Liu, Lattice Boltzmann simulations of micron-scale drop impact on
in simultaneous two droplets impact onto stationary and moving liquid film, dry surfaces, J. Comput. Phys. 229 (2010) 8045–8063.
Comput. Fluids 107 (2015) 285–300. [40] E.B. Dussan V, On the spreading of liquids on solid surfaces: static and dynamic
[31] K.A. Raman, R.K. Jaiman, T.S. Lee, H.T. Low, Lattice Boltzmann study on the contact lines, Annu. Rev. Fluid Mech. 11 (1979) 371–400.
dynamics of successive droplets impact on a solid surface, Chem. Eng. Sci. 145 [41] Y. Sui, H. Ding, P.D.M. Spelt, Numerical simulations of flows with moving
(2016) 181–195. contact lines, Annu. Rev. Fluid Mech. 46 (2014) 97–119.
[32] K.C. Sahu, S.P. Vanka, A multiphase lattice Boltzmann study of buoyancy- [42] M.K. Tripathi, K.C. Sahu, R. Govindrajan, Dynamics of an initially spherical
induced mixing in a tilted channel, Comput. Fluids 50 (2011) 199–215. bubble rising in quiescent liquid, Nat. Commun. 6 (6268) (2015).
[33] P.R. Redapangu, S.P. Vanka, K.C. Sahu, Multiphase lattice Boltzmann [43] J.J. Huang, C. Shu, J.J. Feng, Y.T. Chew, A phase-field based hybrid lattice-
simulations of buoyancy-induced flow of two immiscible fluids with Boltzmann finite-volume method and its application to simulate droplet
different viscosities, Eur. J. Mech. B/Fluids 34 (2012) 105–114. motion under electrowetting control, J. Adhes. Sci. Technol. 26 (2012) 1825–
[34] L. Wang, H. Huang, X.Y. Lu, Scheme for contact angle and its hysteresis in a 1851.
multiphase lattice Boltzmann method, Phys. Rev. E 87 (013301) (2013).
[35] H. Huang, J.J. Huang, X.Y. Lu, A mass-conserving axisymmetric multiphase
lattice Boltzmann method and its application in simulation of bubble rising, J.
Comput. Phys. 269 (2014) 386–402.

Das könnte Ihnen auch gefallen