You are on page 1of 11

Advances in Colloid and Interface Science 210 (2014) 2–12

Contents lists available at ScienceDirect

Advances in Colloid and Interface Science


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

Electrowetting — From statics to dynamics


Longquan Chen, Elmar Bonaccurso ⁎
Experimental Interface Physics, Center of Smart Interfaces, Technische Universität Darmstadt, Alarich-Weiss-Str. 10, 64287 Darmstadt, Germany

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

Available online 10 October 2013 More than one century ago, Lippmann found that capillary forces can be effectively controlled by external
electrostatic forces. As a simple example, by applying a voltage between a conducting liquid droplet and the
Keywords: surface it is sitting on we are able to adjust the wetting angle of the drop. Since Lippmann's findings,
Wetting electrocapillary phenomena – or electrowetting – have developed into a series of tools for manipulating
Electrowetting microdroplets on solid surfaces, or small amounts of liquids in capillaries for microfluidic applications. In this
Static & dynamic capillary phenomena
article, we briefly review some recent progress of fundamental understanding of electrowetting and address
Electrocapillarity
some still unsolved issues. Specifically, we focus on static and dynamic electrowetting. In static electrowetting,
we discuss some basic phenomena found in DC and AC electrowetting, and some theories about the origin of
contact angle saturation. In dynamic electrowetting, we introduce some studies about this rather recent area.
At last, we address some other capillary phenomena governed by electrostatics and we give an outlook that
might stimulate further investigations on electrowetting.
© 2013 Elsevier B.V. All rights reserved.

Contents

1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2. Wetting fundamentals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1. Contact angle and contact angle hysteresis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1.1. Contact angle . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.1.2. Contact angle hysteresis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3
2.2. Wetting dynamics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2.1. Fast wetting stage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
2.2.2. Slow wetting stage . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3. Static electrowetting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.1. The Young–Lippmann equation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4
3.2. DC electrowetting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.2.1. Contact angle hysteresis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.2.2. Polarity effects . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5
3.3. AC electrowetting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.3.1. Why AC? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.3.2. Frequency dependence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.3.3. Hydrodynamic flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6
3.4. Contact angle saturation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

Abbreviations: AC, alternating current; DC, direct current; EWOD, electrowetting on dielectric; HD, hydrodynamic (model); MK, molecular kinetic (theory); A, area (m2); C, capacitance
per unit area (F/m2); E, electric field strength (V/m); F, force (N); H, drop height (m); H⁎, characteristic drop height (m); L, characteristic length (m); LC, capillary length (m); R, drop
wetting radius (m); R⁎, characteristic drop radius (m); R0, initial drop radius; T, absolute temperature (K); U, contact line velocity (m/s); U⁎, characteristic contact line velocity (m/s);
V, voltage, applied potential (V); VS, saturation voltage (V); Vth, threshold voltage (V); Veff, effective voltage (V); c, coefficient; d, thickness (m); f, frequency (Hz); f0, molecular jump
frequency (Hz); fC, critical frequency (Hz); g, gravitational acceleration (m/s2); kB, Boltzmann constant; l, slip length (m); t, time (s); α, wetting exponent; γ, surface tension (N/m);
γLS, liquid–solid interfacial tension (N/m); γSV, solid–vapor interfacial tension (N/m); ε, relative permittivity; ε0, free space permittivity; λ, molecular displacement (m); μ, viscosity
(Pa s); θeq, equilibrium contact angle; θ, contact angle; θA, advancing contact angle; θR, receding contact angle; Δθ, contact angle hysteresis; ρ, density (kg/m3); σ, conductivity (S).
⁎ Corresponding author at: Experimental Interface Physics, Center of Smart Interfaces, Technische Universität Darmstadt, Alarich-Weiss-Str. 10, 64287 Darmstadt, Germany. Tel.: +49
615116 2282; fax: +49 615116 2048.
E-mail address: bonaccurso@csi.tu-darmstadt.de (E. Bonaccurso).

0001-8686/$ – see front matter © 2013 Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.cis.2013.09.007
L. Chen, E. Bonaccurso / Advances in Colloid and Interface Science 210 (2014) 2–12 3

4. Dynamic electrowetting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.1. “Fast” dynamics of electrowetting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8
4.2. “Slow” dynamics of electrowetting . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
5. Other capillary phenomena governed by electrostatics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
5.1. Electro-coalescence . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
5.2. Electrically assisted capillary wrapping . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 9
5.3. Electrostatic control of interfacial flow . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
5.4. Electric enhancement of heat and mass transfer . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
6. Conclusion and outlook . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11

1. Introduction 2. Wetting fundamentals

Wettability is a key parameter to describe the chemical–physical 2.1. Contact angle and contact angle hysteresis
properties of a surface and is usually characterized by a very simple
method: measuring the angle of contact – or wetting angle – of a 2.1.1. Contact angle
droplet of a test liquid with the surface. Triggered by many industrial When a liquid drop is brought into contact with a solid surface, the
applications, such as coating, printing, cleaning, or friction and wear drop spreads on the surface to minimize the free energy of the system.
control, many chemical or physical methods have been developed to Eventually, the drop comes to rest on the surface in a minimum energy
control the wettability of surfaces [1–3]. Eventually, lyophilic, state.q the drop size, R0, is smaller than the capillary length, LC, e.g. R0 ≪
If ffiffiffiffi
γ
superlyophilic, lyophobic and superlyophobic surfaces can be LC ¼ ρg , with γ and ρ respectively the surface tension and the density
fabricated in laboratory by modifying the surface chemistry and of the liquid and g the acceleration due to gravity, the gravity does not
introducing multiscale physical roughness [1–3]. Such surfaces distort the spherical drop shape and can thus be neglected [25]. This
maintain their wetting properties over some time, but do not allow condition is usually satisfied in all published electrowetting studies.
for an active control of their wettability after being manufactured. For chemically and physically homogenous surfaces, the drop in
In practical applications, however, active control of the wettability equilibrium adopts a spherical cap shape, as shown in Fig. 1. The
is more attractive. Indeed, smart approaches such as thermal tuning equilibrium contact angle, θeq, near the contact line is determined by
[4,5], optical switching [6], as well as electrostatic controlling of the interfacial tensions, γLV, γLS, and γSV at the liquid–solid–vapor
contact angles [7,8] have been developed. Among them, the interfaces.
electrostatic method is the most popular one due to its real-time
actuation, fast response, long term reliability, and good stability of γ SV −γLS
cosθeq ¼ ð1Þ
the actuation. γ
During his works, Lippmann found that applying a voltage
between mercury and aqueous electrolytes allowed for controlling Usually, γLV is denoted as γ for brevity. The above equation is
the position of the mercury meniscus in a capillary. In 1875, he was called Young's equation, in honor of Thomas Young who expressed
probably the first to report this electrocapillary phenomenon, which it with words in his work published in 1805 [26]. Young's equation
is at the foundations of electrowetting [9]. He further proposed a can be derived either from a mechanical perspective [1,3] or from
physical model and developed a number of applications. Later, a thermodynamic perspective [27]. θeq is a useful parameter to
Möller [10], Frumkin et al. [11], Gorodetskaya and Kabanov [12], characterize the wettability of surfaces, as one can easily relate the
Smolders [13], and Nakamura et al. [14] conducted contact angle contact radius R to θeq with
measurements at the mercury/metal-electrolyte interfaces. They 2 31=3
found that the contact angle decreased with applied potential, and
6 4 7
argued that the decrease of contact angle was due to the change of R ¼ R0 sinθeq 4 2  5 : ð2Þ
interfacial energy [10–15]. However, Lippmann's discovery and the 1− cosθeq 2 þ cosθeq
other works did not attract much attention until the 1980s, when
the term electrowetting was coined and proposed for designing Taking the limit as θeq → 0°, we find R → ∞ which means the liquid
display devices [16,17]. Since then, electrowetting started to develop tends to wet the surface completely. While if θeq → 180°, we obtain
rapidly and nowadays it has been successfully applied in areas like R → 0, which reflects that the surface repels the liquid extremely.
lab-on-chip systems [18–20], adaptive optical lenses [21], electronic
display technology [17,22], or mixing in microfluidic channels 2.1.2. Contact angle hysteresis
[23,24]. In principle, electrowetting can be applied to drops sitting Natural surfaces are decorated with physical roughness or chemical
on a bare electrode, or on thin dielectric layer on top of an electrode. moieties. The physical or chemical heterogeneity of surfaces leads to
However, most of the recent electrowetting studies and applications deviations of the contact angle from the one predicted by Young's
are carried out on dielectric, giving rise to the definition of equation. Pinning of the contact line of a wetting/dewetting drop due
electrowetting-on-dielectric (EWOD).
In this review, we focus on the latest progress on some
fundamental aspects of electrowetting. In Section 2, we give a short
description of the basics of static and dynamic wetting. Section 3 is
devoted to static electrowetting. We will discuss the basic
phenomena in DC and AC electrowetting, as well as the contact
angle saturation (CAS) phenomenon. In Section 4, recent results
about fast and low speed electrowetting will be presented. Finally,
we briefly address some other interesting capillary phenomena
governed by electrostatics. Fig. 1. Sketch of a drop sitting on a solid substrate in equilibrium.
4 L. Chen, E. Bonaccurso / Advances in Colloid and Interface Science 210 (2014) 2–12

to surface heterogeneity causes an advancing contact angle θA/a L is a macroscopic characteristic length and is typically of the order of
receding angle θR. The difference between θA and θR is defined as the the drop size. l denotes a microscopic slip length of the order of a
contact angle hysteresis Δθ [1,3,28,29] molecular size. If the surface is very lyophilic, e.g. θeq ≪ 1, one obtains
the growth of the spreading radius as a function of time [28,29,39].
Δθ ¼ θA −θR : ð3Þ
1=10
Ret ð6Þ
The contact angle hysteresis characterizes the homogeneity of
surfaces. Larger values of Δθ correspond to more pronounced surface The above equation is sometimes referred to as Tanner's law [40].
heterogeneities. The spreading could also be dominated by microscopic friction near
the contact line. In the molecular kinetic (MK) theory, the drop rim
2.2. Wetting dynamics moves via individual molecular jumps activated by thermal energy,
with an equilibrium frequency f0 and a displacement distance λ. The
When a drop touches a surface, the initial contact angle θ is largest relation between the spreading velocity and the dynamic contact angle
and close to 180°. Contact leads to a net horizontal capillary force of is given by [27,41]
γ(cos θeq − cos θ). This force drives drop wetting until equilibrium is
2 2 3
attained. Three sources resist drop wetting: (i) the kinetic energy of γλ cosθeq − cosθ
the spreading drop, (ii) viscous dissipation within the liquid, and U ¼ 2 f 0 λ sinh 4 5: ð7Þ
2kB T
(iii) contact line dissipation at the drop rim. In recent years, benefiting
from the development of high speed imaging techniques, dynamic
wetting has received considerable attention. Many studies showed kB is the Boltzmann constant and T is the absolute temperature. If the
that the wetting proceeds in two stages, a “fast” early stage followed term of sinh in Eq. (7) is small and the surface is very lyophilic, the
by a “slow” later stage. spreading radius can be simplified to [42]

1=7
2.2.1. Fast wetting stage Ret : ð8Þ
For low viscosity liquids, the spreading velocity is very high just
after a drop touches a solid surface. The Reynolds number in Blake et al. also combined both the effects of hydrodynamic stress
 
characteristic (*) experiments, Re ¼ 2 ρUμ R , compares the balance and molecular friction in the modified molecular kinetic theory [43].
of inertial and viscous forces and is larger than unity.  is the ratio Details about these wetting models can be found in a recent review by
of the characteristic drop height H⁎ and the characteristic drop Ralston et al. [44].
radius R⁎, and is approximately unity. U⁎ is the characteristic velocity
of the moving contact line and is of order 1m/s [30–34]. μ is the liquid 3. Static electrowetting
viscosity. Thus, capillarity drives and inertia resists the fast wetting
stage. The change (release) of surface energy is transformed into 3.1. The Young–Lippmann equation
kinetic energy of the moving drop. Based on this energy balance,
Bird et al. found that the spreading radius R grows with time t Fig. 2a shows the sketch of a typical electrowetting setup in which a
according to a power law [31,35]. thin dielectric layer with thickness d is deposited between a conductive
drop and a flat electrode. The insulating layer is few micrometers thick
α
R ¼ ct ð4Þ and is mainly used to prevent electrolytic processes at the interface
between drop and electrode. When a direct current (DC) voltage is
c is a coefficient, while the exponent α is only related to the applied between the drop and the electrode, the contact angle θ decreases
wettability of the surface. It was experimentally found that α takes with voltage V according to the so-called Young–Lippmann equation
values between 0.5 and 0.25 when θeq changes from ~0° to ~110°
[31–33,35,36]. Biance et al. carried out a scaling analysis based on C 2
cosθ ¼ cosθeq þ V ð9Þ
drop coalescence theory. They obtained a power law dynamics with 2γ
α = 0.5 for complete wetting [30], which was also observed by Winkels
et al. in molecular dynamics (MD) simulations [34]. Shanahan et al. θeq is the equilibrium contact angle with zero applied voltage and is
alternatively derived an inertial drop spreading model and found that determined by Young's equation. C is the capacitance per unit area. If
α takes values between 0.5 and 0.2 [33] depending on θeq. However, we treat the capacitor between the drop and the electrode as a plate
all these models are at most semi-quantitative and do not capture all capacitor, C = ε0εd/d. εd and ε0 are the permittivities of the dielectric
details of the wetting dynamics. layer and of vacuum, respectively. For alternating current (AC) voltage,
the root mean square (rms) value of the voltage – or effective voltage –
2.2.2. Slow wetting stage V2eff, is used instead of V2. The Young–Lippmann equation can be derived
The inertial wetting stage lasts several milliseconds only, depending by several approaches, as summarized, e.g., by Mugele and Baret [7]. A
on drop size. After this fast stage wetting crosses into a slow stage in rigorous thermodynamic derivation of Eq. (9) was also presented by
which drop spreading is resisted by viscous friction within the liquid Bormashenko recently [45]. These theories can be understood from two
and by molecular friction near the contact line. Two types of models perspectives:
were proposed, accounting for different mechanisms. The hydrodynamic
(HD) model considers viscous dissipation as the main resisting force.
Based on the introduction of a cutoff or slip length near the contact
line by Huh and Scriven [37] to get rid of the shear stress singularity,
Voinov derived a relationship between the dynamic contact angle θ
and the spreading velocity U [38].
 
3 3 μ L
θ ¼ θeq þ 9 U ln ð5Þ
γ l Fig. 2. (a) Standard electrowetting setup. (b) Schematic plot of electrowetting curve.
L. Chen, E. Bonaccurso / Advances in Colloid and Interface Science 210 (2014) 2–12 5

• Thermodynamic or electrochemical perspective. The classic model for Similarly, the advancing/receding contact angle as the drop
electrowetting was proposed by Lippmann [9]. In his experiments, a spreads/recedes with increasing/decreasing applied voltage can be
voltage was directly applied between the electrode (mercury) and the modified by considering the effects of the contact line pinning force
aqueous electrolyte. An electric double layer builds up spontaneously [50,58]
at the liquid–solid interface, with charges of opposite polarity being
repelled from the interface. Thus, the liquid–solid interfacial energy,
2
γLS, decreases along with the new equilibrium contact angle, i.e. the V¼0 CV
cosθA or R ¼ cosθA or R þ ð11Þ
surface becomes more hydrophilic [46]. In EWOD, the thickness of the 2γ
dielectric layer is significantly larger than that of the electric double
layer [16,47], and hence electrostatic energy stored in it is the main
source to decrease γLS. where θVA = 0
or R is the advancing/receding contact angle measured with

• Electromechanical perspective. Electrowetting could also be understood zero voltage applied. As the contact angle hysteresis is independent of
considering the forces exerted on the liquid near the contact line, as the applied voltage [50–52], reversible electrowetting could be achieved
proposed by Jones et al. [48,49]. When a voltage is applied, the electric with low contact angle hysteresis systems such as in the liquid/liquid/
field near the contact line attracts the free charges as well as the solid system [57,61] or on superhydrophobic/superlyophobic surfaces
polarized dipoles, which gives rise to a Maxwell stress on the liquid– [62].
air interface. In order to balance this stress, the liquid–air interface
must decrease its curvature to reduce the Laplace pressure. Eventually, 3.2.2. Polarity effects
a smaller apparent contact angle depending on the magnitude of the According to the Young–Lippmann equation in Section 3.1, the
applied voltage is attained. contact angle decreases with the applied voltage following the same
parabolic curve regardless of the polarity of the voltage. However,
A large amount of experimental studies showed that the Young–
early studies showed different polarity effects of electrowetting on
Lippmann equation can well predict the contact angle when the voltage
different materials. On one side, when Parylene was used as an
is smaller than a critical value VS [7,8,23,47]. When V ≥ VS, the
insulating layer, the change of contact angle was independent of
electrowetting contact angle does not increase with applied voltage
polarity [63–65]. On the other side, when Teflon was used several
(Fig. 2b), which is the effect known as contact angle saturation (CAS).
groups reported a strong deviation from the theory with a positive
The physics underlying CAS is not fully clear yet. A detailed description
applied voltage [66–68]. Since the liquids (water or aqueous electrolyte
of the recent progress on theories about CAS is presented in Section 3.4.
solutions) used in experiments were similar, authors attributed the
polarity effects to material properties [66–68]. Werner and co-workers
3.2. DC electrowetting
investigated the interfacial charge on Teflon in aqueous electrolyte
solutions [69]. They found that hydroxyl ions (OH−) have stronger
3.2.1. Contact angle hysteresis
interaction to the Teflon surface than the hydronium ions (H3O+) due
The phenomenon of contact angle hysteresis also plays a role in
to the presence of oxygen atoms in Teflon [66,70]. As a result, the
electrowetting [50–52]. Typically, the contact angle hysteresis has two
capacitance of the liquid–solid interface is better described by both a
main effects: first, electrowetting can start only when the applied
dielectric layer and an electric double layer, rather than by only a
voltage is larger than a threshold value Vth [53–56]; second, the contact
dielectric layer, as done in the Young–Lippmann model [68,71], and it
angle hysteresis leads to different contact angles with the same applied
can explain the polarity effects of electrowetting on Teflon surfaces.
voltage, depending if the voltage is increasing or decreasing. This is
In standard electrowetting experiments or applications, water and
referred as irreversibility or hysteresis of electrowetting [50,57].
aqueous electrolyte solutions were normally used [18,19,50–68,70,71].
The existence of the threshold voltage Vth can be explained by
Although water has useful solvent properties, its poor thermal stability,
accounting for contact line pinning. Due to the natural heterogeneities
it tendency to evaporate, and its corrosion ability lead to significant
of real surfaces, the equilibrium contact angle has a value between
limitations in microfluidic applications. Ionic liquids, as first pointed
the receding and the advancing contact angle, e.g. θR b θeq b θA.
out by Ralston et al., are more promising for robust applications due to
When electrowetting is initiated, the electrostatic force wants to
their unique properties: high thermal stability, non-flammability, and
reduce the contact angle. However, the contact line will move only
no significant vapor pressure [72]. In the same paper, Ralston et al.
if the electrostatic force exceeds the pinning force [50,51,58–60].
showed polarity effects and related them to the different properties/
Balancing these two forces, Fréchette et al. predicted the threshold
sizes of cations and anions [72]. A systematic study was carried out
voltage by [50]
later by Armstrong and coworkers. They found that anions or cations
with smaller size have stronger influence on the asymmetry of the
rffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
2γ   electrowetting curve since they interact more strongly with the Teflon
V th ¼ cosθeq − cosθA : ð10Þ surface than larger anions or cations do [73].
C
As discussed above, polarity effects in electrowetting are recognized
to be most probably due to molecular processes near the liquid–solid
Thus, the threshold voltage can be decreased by using, e.g., low interface, which cannot be captured by simple optical instruments.
contact angle hysteresis surfaces. Until then, the lowest threshold Thus, MD simulations could help in understanding the physical
voltages in electrowetting were reported in the range of 15–20 V [19]. mechanisms. Luzar and co-workers studied the electrowetting of
In their paper, Fréchette et al. showed that nearly no threshold voltage water nanodrops on apolar hydrophobic surfaces, such as graphene
existed on an oil-impregnated polydimethylsiloxane (PDMS) surface or paraffin [74–77]. They observed that the electrowetting contact
with only 1° contact angle hysteresis [50]. Eq. (10) also indicates that angle is sensitive to the polarity of an electric field directed
Vth can be reduced by decreasing the interfacial tension γ. This was perpendicular to the surface. The explanation was that water is
achieved by changing the fluid around the drop from air to a liquid, as easier to be polarized along the outgoing, rather than along the
liquid–liquid interfacial tensions are usually smaller than liquid–air incoming, electric field to the surface [74–76]. However, in the
interfacial tensions [55,57,61]. Moreover, the contact angle hysteresis presence of any salt ions, the strong affinities of them towards the
in liquid/liquid/solid systems is also smaller than in liquid/solid/air surface increase the screening of the electric field, and eventually
systems, as liquid–liquid friction is much smaller than liquid–solid suppress the polarity effects caused by different electric responses
friction [61]. of hydrogen bonding [78].
6 L. Chen, E. Bonaccurso / Advances in Colloid and Interface Science 210 (2014) 2–12

3.3. AC electrowetting σd, σl and εl are the conductivities of the dielectric layer and liquid
and the permittivity of the liquid, respectively. k ~ d/R is the size ratio
3.3.1. Why AC? of the dielectric layer and the drop. For deionized water, fC is typically
In most microfluidic electrowetting systems [19], alternating current few kHz, which matches with experimental observations [48,49,58].
(AC) voltage is applied because of following benefits: The above discussion qualitatively introduces the frequency
dependence of AC electrowetting, while a quantitative understanding is
• Reduction of contact angle hysteresis. As discussed above, the contact still missing. Hong et al. carried out a numerical investigation of AC
angle hysteresis exists in DC electrowetting due to pinning effects electrowetting [85]. They found that the electric field strength was
[50–52]. In contrast, using AC voltage electrowetting continuously remarkably reduced in the dielectric layer near the contact line, which
perturbs the force balance at the contact line and essentially leads to may lead to a decrease of the electrostatic contribution to the wetting
the depinning of the contact line from the surface [51,58]. Eventually, tension. Their numerical results were consistent with experiments. Most
the contact angle hysteresis is smaller in AC than in DC electrowetting recently, Klarman et al. proposed a model based on a similar configuration
[51,58,79]. as shown in Fig. 3 [88]. They also accounted for the effects of the electric
• Delay of contact angle saturation. Many experimental results showed double layer at the immersed counter-electrode and discussed this effect
that the contact angle saturation occurs at a smaller contact angle on AC and DC electrowetting. Among others, they concluded that the
and at a higher effective voltage for AC rather than DC electrowetting, effect of the polarization of the counter-electrode could be neglected in
in both liquid/solid/air [80,81] and liquid/liquid/solid systems [82–84]. most practical cases, since the area of the counter-electrode is usually
Hong et al. provided the explanation that the effective electric field much smaller than the contact area between the drop and the solid.
strength is reduced by increasing the frequency, and that this can
lead to a delay of the dielectric breakdown [85]. The underlying 3.3.3. Hydrodynamic flow
physics of delayed saturation in AC electrowetting is still not fully In recent years, researchers have reported hydrodynamic flows
understood. inside drops in AC electrowetting [24,89–93]. The flow patterns are
• Reduction of ion adsorption. Another possible reason leading to dependent on the frequency of the applied voltage as well as on the
irreversibility or a large contact angle hysteresis in DC electrowetting conductivity of the liquid (Fig. 4a). In the low frequency range (typically
is ion adsorption at the liquid–solid interface. It was reported that the up to 1 kHz), the flow forms two axisymmetric toroidal vortexes whose
ion adsorption at the liquid–solid interface was reduced by applying center is not visible [90,93] (Fig. 4b). The low frequency flow is rather
AC instead of DC voltage [86,87]. unstable and is very sensitive to the counter-electrode position. In
contrast, in the high frequency range (larger than few tens of kHz),
3.3.2. Frequency dependence the flow is highly stable and the vortex center is always located around
The Young–Lippmann equation is only valid for AC electrowetting the counter-electrode [90]. Also, the vortex direction of the low
when the liquid can be treated as a perfect conductor. This holds if the frequency and high frequency flows are different (Fig. 4b & d), which
AC frequency is larger than the resonance frequency of sessile drops indicates that the flow generation mechanisms are different. For
(typically of few hundreds hertz) and is far smaller than a critical intermediate frequencies, no flow is observed in the drop (Fig. 4c).
frequency fC (typically of 10–100 kHz) [7]. The electrowetting efficiency The low frequency flow is mainly driven by the oscillations of the
diminishes with frequency increasing beyond the upper limit, while it is drop [90]. The oscillation of the contact line caused by AC potential can
dependent on the frequency between the two limits [48,49,58,85]. The result in a hydrodynamic viscous streaming inside the drop. However,
general AC electrowetting behavior can be captured by the analogy the low frequency flow observed is not as steady as the viscous flow.
with an equivalent electric circuit, as shown by Mugele and Baret [7] Moreover, the direction of the low frequency flow in air [90] is opposite
(Fig. 3). The drop is represented as a resistor and a capacitor, and the to the one observed in mixing experiments of water/glycerol drops in
dielectric layer is treated as another capacitor. The electric double layers silicone oil ambient [24,93]. Both phenomena point out that some
formed at the liquid-dielectric layer or at the liquid-counter electrode additional mechanisms may influence the hydrodynamics. Based on
interfaces are not considered. At low frequency (≪fC), the drop behaves these experiments in oil, Mugele et al. suggested that the capillary
more like a conductor rather than an insulator, which satisfies the Stokes drift induced by the capillary wave drives the internal flow in
assumption for Young–Lippmann's equation. However, at very high the drop [93,94]. These studies pointed to more open questions, like
frequency (≫fC), the drop behaves more like an insulator, which how do the viscosity or the density of the atmosphere surrounding the
leads to the breakdown of electrowetting. The critical frequency is drop influence flow pattern and flow stability inside the drop.
related to the physical properties of the drop as well as to its geometry The high frequency flow is induced by the so-called electrothermal
[7]. effects [89,91]. With high frequency AC voltage applied, the liquid
behaves like an insulator as there is not sufficient time to completely
σ d þ σ lk charge the capacitor of the dielectric layer [7,88]. As a result, the electric
fCe ð12Þ
ε0 ðεd þ εl kÞ field inside the drop leads to Joule heating of the liquid, giving rise to
a gradient in conductivity and permittivity [89,91]. Eventually, the
voltage applied on the electrically inhomogeneous drop generates an
electric body force resulting in an internal flow.

3.4. Contact angle saturation

In the past years, many models or hypotheses have been proposed to


explain the contact angle saturation, as reviewed by Mugele [7,23], and
recently by Chevalliot et al. [80], Koopal [95] and Sedev [96]. Here, we
briefly recap the latest progress:

(1) Zero interfacial tension. According to the classical thermodynamic


model of electrowetting, the change of the contact angle is
entirely ascribed to the reduction of the liquid–solid interfacial
Fig. 3. The equivalent circuit for AC electrowetting. tension (γLS). Ralston and coworkers suggested that γLS should
L. Chen, E. Bonaccurso / Advances in Colloid and Interface Science 210 (2014) 2–12 7

Fig. 4. (a) The influence of electrolyte concentration on the frequency range of hydrodynamic flows with V = 80 V and the three typical flow patterns with frequency of 1 kHz (b), 18 kHz
(c) and 128 kHz (d). The electrolyte concentration in (b)–(d) is 10−3 M.
Reproduced with permission from Ref. [90].

not be less than zero to keep thermodynamic stability [47,97]. the electric field strength, the interfacial curvature near the
Thus, the thermodynamic limit with γLS = 0 causes the contact contact line, or the type of dielectric used [80].
angle saturation. They showed a consistence between their (3) Contact line instability. In early AC electrowetting studies, Vallet
model and experimental results [47,97]. Berry and coworkers et al. observed an instability of the contact line close to the
confirmed this model by low voltage electrowetting in both critical (saturation) voltage for low conductivity liquids such as
liquid/liquid/solid systems and liquid/solid/air systems [98,99]. water [110]. They found that microdrops spontaneously ejected
One big challenge to verify this model is that it is hardly possible from the contact line as the voltage was larger than a critical
to directly measure liquid–solid interfacial tensions. However, value. This phenomenon was reproduced later by Mugele and
recent measurements of thin films [100] and adhesion force Herminghaus with various liquids [111] and was attributed to
[101] showed no major change of the liquid–solid interfacial the diverging of charge density around the contact line. While
tension upon applying an external potential. Other groups also the applied voltage is larger than a critical value, the Maxwell
noticed that the zero interfacial tension theory does not predict force exceeds the capillary force near the contact line, which
reliably the saturation contact angle, e.g. with cos θS = γSV/γ leads to emission of (charged) satellite droplets. This instability
[80,82,83]. Moreover, Mugele and co-workers numerically at the contact line was identified as a possible cause for contact
found that the electrostatic force does not affect the interfacial angle saturation [110,111]. Park et al. also suggested that the
tensions force balance at the contact line [102], and they further instability with droplet ejection is preceded by a contact angle
experimentally observed that the microscopic contact angle reduction and the extrusion of a thin layer from the edge of the
always equals Young's angle [103]. This was later confirmed droplet [112,113]. However, these models cannot explain the
by MD simulation by Liu et al. [104]. Till now, no studies suppression of the instability by increasing the conductivity of
considered the possibility of a significant change of the the liquids, e.g. with salts [110,111]. As contact line instability
liquid–vapor or solid–vapor interfacial tensions upon does not arise in DC electrowetting, it hence may not capture
applying an electric potential. Recent work, however, showed the actual physics of contact angle saturation.
that the surface tension of water decreased remarkably by (4) Gas ionization or insulating fluid charging. Vallet et al. also reported
approx. 50% upon electrically charging the water surface air ionization around the contact line for voltages larger than the
[105] – even if the external potential was in the kV range; or saturation voltage and suggested that it could be responsible for
that the solid–vapor and solid–liquid interfacial energy can contact angle saturation [110]. Indeed, it was experimentally
be reduced upon weak electron irradiation [106] – causing a found that the contact angle of a drop can be controlled by
reversed electrowetting effect, i.e. increasing the contact charging its surface via a corona discharge [114,115] or by
angle upon irradiation. These results suggest that electric electrons [106]. As corona discharge or air ionization is dependent
charges can affect either of the interfacial tensions playing a on relative humidity [115], one possible way to validate this
role in electrowetting or in contact angle saturation. hypothesis would be to check whether the relative humidity of
(2) Dielectric breakdown. Near the contact line, the electric field the surrounding air influences contact angle saturation. On the
diverges as the interface curvature is extremely small. The electric other hand, corona discharge is not applicable to liquid–liquid–
field may locally become so strong that it exceeds the dielectric solid electrowetting systems. In such systems, Chevalliot et al.
breakdown strength of the insulating layer [64]. As a result, the suggested that it could be possible that the insulating fluid is
dielectric layer locally breaks down [107,108]. The charges charged due to ejection of charges or satellite drops from the
transferred to the dielectric layer lead to screen the electric field, conducting drop [80]. However, this hypothesis calls for further
which eventually reduces the electrostatic contribution and experimental evidence.
causes electrowetting saturation. This hypothesis was proposed (5) Minimization of the electrostatic energy. Lin et al. proposed a
by Drygiannakis et al. who showed a good match between their theoretical model based on an energy balance [116] instead of a
experimental results and theory [107]. Similar conclusions were force balance, as is used in the derivation of the Young–Lippmann
also drawn by Berry et al. based on their experiments on thin equation. In AC electrowetting, especially at higher voltages and in
amorphous fluoropolymer films [98,109]. However, Chevalliot a well-defined frequency range, a strong hydrodynamic flow
et al. recently reported experimental results on DC electrowetting develops inside the drop [24,89–93]. The flow causes additional
showing that the saturation contact angle is invariant with energy dissipation and the dissipated energy increases with the
respect to a number experimental parameters, among which are applied potential. Eventually, this leads to a minimum electrostatic
8 L. Chen, E. Bonaccurso / Advances in Colloid and Interface Science 210 (2014) 2–12

contribution to the total energy of the system when V ≥ VS [116]. 4.1. “Fast” dynamics of electrowetting
However, this model fails to explain the contact angle saturation
of DC electrowetting, as no internal flow develops there, and The early wetting stage after a drop touches a surface runs very fast,
does not reproduce the correct dependence cos θ ∝ V2 for V b VS. as discussed in Section 2.2. This type of electrowetting is ubiquitous as
Klarman and Andelman also developed a model based on energy most surfaces are naturally charged, which is the reason – except for
dissipation [88], and accounted for the electrostatic energy of the well controlled conditions – that there is always difference of potential
electric double layer around the counter-electrode. This is a between drops and surfaces before they wet each other. These naturally
reasonable approach, as the electric field is always stronger near arising excess surface charges need to be redistributed between drop
the electrode than that in the bulk drop [89,91]. They found a and surface upon meeting, and thus play a role in most spontaneous
minimum of the electrostatic energy contribution to the saturation wetting processes like rain drops hitting the soil or ink drops wetting
voltage for both DC and AC electrowetting systems. This was based paper. Stone and co-workers were the first to investigate the fast
on the assumption that the capacitance C is not constant but dynamic wetting of sessile drops just after contact with a surface
function of θ and has a minimum at VS. However, this assumption under an applied electric potential [35]. They found that the drop
is not based on first principles and would require a rigorous proof. spreading radius grows with a power law with a larger exponent,
(6) Tailor cone. Most recently, Chevalliot et al. carried out a systematic α′ ≈ 2/3. Without applied potential α = 1/2 on completely wetting
study of all physical and chemical parameters which may lead to surfaces. The effect of the applied potential was to deform the spherical
contact angle saturation according to the above models [80]. drop into a conical shape close to the point of contact. The authors set up
Based on their experimental results, they state that there are an energy balance accounting for this new contact geometry and
parallels between the phenomena of Tailor cone formation and derived a power law with an exponent of 2/3. Exponents larger than
contact angle saturation. This, if confirmed, could provide a new 1/2 were also observed by Chen et al. [117] in experiments. However,
direction for a universal theory of the saturation problem in exponents were not constant, but a function of the applied potential,
electrowetting. of surface wettability, and of electrolyte concentration. Similar to static
electrowetting, where the contact angle saturates when the applied
voltage is larger than VS, they found a saturation of the wetting
4. Dynamic electrowetting exponent α′ for aqueous electrolyte drops with low salt concentration.
For high electrolyte concentrations, no saturation was observed at the
In many applications, such as lab-on-chip, electrowetting dynamics is voltages applied (up to V ≈ 1000 V). Chen et al. modified Stone's early
of great interest as it controls the response time of the devices [19]. During model for partial wetting [31] by adding the contribution of the
the wetting process, the dynamics is determined by the energy balance electrostatic energy stored in the electrostatic double layer near the
between the driving and the resisting forces [28,29,39]. The driving liquid–solid interface as a driving force. The proposed model can explain
force of the spreading drop is always the change of surface energy, the experimental observations. MD simulations confirmed that the
which is related to the difference between the initial and equilibrium electric double layer is formed in a time scale of few hundred picoseconds
contact angle. The resisting force could be the kinetic energy of the [117], which satisfies the assumption that the electrostatic double layer
moving drop [30–33], the viscous dissipation in the vicinity of the moving exists during the fast wetting process (typically, ~10 ms). However, the
contact line [28], or the microscopic dissipation of the molecular jumps saturation of the wetting exponent in dependence of salt concentration
[67]. In electrowetting systems it is thus expected that the addition of and applied potential is still an open question.
electrostatic energy to the energy balance will modify the dynamics of The inertial wetting stage was also observed in the MD simulations
spreading. In literature, many experimental configurations have been of the early wetting of nanodrops [117]. However, the electrostatic
applied to study the dynamics of electrowetting, as shown in Figs. 2a & 5. energy did not enhance the wetting exponent here. One plausible

Fig. 5. Typical experimental configurations for dynamic electrowetting setups. (a) Dynamic wetting experiments with a spreading sessile drop. (b) Wilhelmy plate method and (c)
capillary rise experiments.
L. Chen, E. Bonaccurso / Advances in Colloid and Interface Science 210 (2014) 2–12 9

explanation is that the electrostatic double layer is not yet fully formed fit their data. They showed that the applied potential controlled the
in this short period. In fact, the ion migration time needed to form the wettability of the surface and that average molecular displacement, λ,
electrostatic double layer near the liquid–solid interface by applying a decreased with the applied potential [124]. Puah et al. applied a similar
potential was ~500 ps, which is two times longer than the inertial method to investigate the influence of the electric surface charge on
wetting time [117]. Thus electrostatics does not influence the early wetting dynamics [125]. Also there, MK theory was applied. This time
spreading of nanodrops, but it does for their later (viscous) wetting it was found that the equilibrium displacement frequency, k0, depended
and for their equilibrium contact angle. on the applied potential and was determined by the charges at the
liquid–solid interface, while λ did not show a significant relation to
4.2. “Slow” dynamics of electrowetting surface charge density. The values of λ were determined mainly by the
spacing of the adsorption sites on the solid surface. This was consistent
The electrostatic effects on the viscous stage of dynamic wetting are with previous work of the group using a standard electrowetting
difficult to investigate experimentally with a configuration as in Fig. 5a, configuration [82,83,120]. Blake et al. modified the MK theory by
since the drop detaches from the needle during this stage. Losing contact introducing the actual capillary force in Eq. (14) [126]. However, they
with the needle (which also acts as counter-electrode) leads to an open found that both k0 and λ were independent of the applied voltage.
circuit configuration. Alternatively, MD simulations can be applied to This brief presentation of experimental results is just to show that also
investigate the slow electrowetting. It was found that wetting follows a experiments performed with different configurations and materials
power law R ~ t0.1 without applied electric field [117] — which is are not always in agreement, and that the cause of disagreement is not
consistent with the HD model [28,29,39]. With an applied electric field yet clear. On the other hand, it becomes increasingly clear that MD
of 0.1 V/nm the wetting exponent was larger and the wetting law was simulations are now at a level that they can well capture the processes
R ~ t0.15 [117]. Similar wetting exponents from MD simulations were at the scale of the contact line [127]. They represent thus a promising
also obtained by Yuan and Zhao [118]. Despite these preliminary findings, tool for exploring the influence of external potentials on molecular
there is still a lack of knowledge on the effect of an applied electric displacements of the contact line in particular and of the electrowetting
potential on slow electrowetting dynamics. Another unsolved issue is process in general.
the influence of the electrolyte concentration on slow electrowetting In recent years, electrostatic assisted flow in capillaries or
dynamics. microchannels attracted quite some attention because of envisioned
Most studies of dynamic electrowetting in literature were carried out applications in microfluidics and lab-on-chip applications (Fig. 5c).
with the standard configuration shown in Fig. 2a, with the needle Wang and Jones proposed a hydrodynamic model to describe the
(counter-electrode) permanently inserted in the drop for applying the electrowetting-induced capillary rise in a millimeter-sized plane-
electric potential. The driving force due to the electrostatic contribution is parallel channel [128]. According to this model, there is a critical
potential below which contact line friction dominates over viscous
 
E
F ¼ γ cosθeq − cosθeq : ð14Þ dissipation, and above which these effects are reversed. However,
when the channel size shrinks to several tens of micrometers,
Herminghaus and co-workers showed that viscous dissipation always
θEeq is the new equilibrium electrowetting contact angle that is a dominates the electrowetting dynamics [129–131].
function of the applied voltage. This force is compensated by viscous
dissipation or contact line friction during the spreading. Decamps and 5. Other capillary phenomena governed by electrostatics
De Coninck modified the MK theory with this electrocapillary force and
applied it to spontaneous spreading experiments of glycerol drops on 5.1. Electro-coalescence
Teflon [119]. They found that the contact line friction was constant over
a large range of applied potentials. Ralston and co-workers studied the When two drops of the same liquid come into contact, they merge
dynamic electrowetting of various ionic liquids in liquid/liquid/solid with each other to minimize the surface energy. Drop surface charges
systems [82,83] and liquid/solid/air systems [120]. They used both HD or externally applied potentials can favor – or in some cases delay –
model and MK theory to interpret their data. They found that viscous coalescence [132,133]. If the drops are oppositely charged, non-
dissipation dominates the dynamics when the contact angle and contact coalescence was observed when the electric field was larger than a
line velocity are small, while the MK theory was more suitable to explain critical value EC (Fig. 6). The failing of coalescence can be explained by
the dynamics at high contact angles and with high contact line velocity considering the local pressure at the drop apex during coalescence
[83,120]. They also showed that the equilibrium frequency of molecular [132,134]. Exposed to an electric field, the drop deformed into a conical
jumps decreased with liquid viscosity, which indicates that it is more shape upon approaching the oppositely charged interface. The aperture
difficult for liquid molecules with higher viscosities to move from one angle of the cone depends on the field strength. If the cone angle is
adsorption site to another since the cohesion force between the molecules larger than 31°, the local curvature in the liquid thread results in such
is much stronger. Very recently, Hong et al. reported that liquid viscosity a high pressure that prevents coalescence. A similar phenomenon was
has a weak effect on electrowetting velocity over a broad range of also observed in the coalescence of a drop with a flat solid surface
viscosities (from approx. 1 to 200 mPa s) [121]. McHale and co-workers [117]. When the electric field was below the critical value EC, the
investigated the dynamic dielectrowetting of dielectric liquids impinging drop partially coalesced with the interface [135–138]. The
[122,123]. In their paper, they modified the HD model and found that coalescence was only partial, since a daughter drop was ejected and
when the applied voltage was larger than a threshold value the wetting moved away from the interface, as it acquired an opposite charge during
exponent increased with the voltage [123]. Yuan and Zhao studied the contact. The size of the daughter drop is determined by electrostatic and
dynamic electrowetting of cylindrical water nanodrops with MD capillary forces [136].
simulations [118]. They observed a power-law growth of the
spreading radius with time and an increase of the wetting exponent 5.2. Electrically assisted capillary wrapping
with the electric field. When a critical value of the field strength was
reached, the exponent saturated similarly to the saturation found in It was recently demonstrated by several authors that capillary forces
static electrowetting [7,23]. can be used to fold thin films and to wrap them around droplets
Dynamic electrowetting studies were also carried out with the [139–141]. In order to achieve a reversible wrapping, an additional force
Wilhelmy plate configuration (Fig. 5b). Schneemilch et al. studied the is needed to oppose the capillary folding force. One way is using the
electrowetting of a Teflon-covered rod and applied the MK theory to electrostatic force. Pineirua et al. controlled the folding and unfolding of
10 L. Chen, E. Bonaccurso / Advances in Colloid and Interface Science 210 (2014) 2–12

Fig. 6. Influence of an electric field on drop coalescence: below a critical field strength EC the drop coalesces with the interface, and above EC the impinging drop rebounds from the
interface. The scale bar is 0.5 mm.
Reproduced with permission from Ref. [132].

capillary origami by varying the unfolding torques with an external DC temperatures exceeding the Leidenfrost temperature [149–151]. They
voltage [142]. The authors showed that this technique can be used to found that the evaporation was enhanced under an applied field, as the
assemble small scale three dimensional soft structures. Zhao and co- field induced interfacial instability leading to direct contact between the
workers reported a similar effect using low frequency AC voltages [143]. levitating drop and the heated substrate [150]. They concluded that the
enhancement was more remarkable for polar liquids than for non-polar
5.3. Electrostatic control of interfacial flow liquids, as charge relaxation time on them was much shorter [150,152].
Recently, Butt and co-workers calculated that the presence of an electric
The electrostatic force can also be used to control interfacial flows, as field can reduce the saturation vapor pressure and lead to a field-
it deforms the shape of a drop or a rivulet. Lee et al. transferred the induced condensation of, e.g., water [153]. This explains the anodic
electrostatic energy into the surface energy by applying a voltage to oxidation at the nanoscale in dependence of relative humidity [154,155].
spherical drops sitting on superhydrophobic surfaces [144]. If the
increased surface energy overcomes a critical barrier, the drop can be 6. Conclusion and outlook
induced to jump off the surface. The voltage for making a 5 μl water
drop jump off is approx. 100V. The authors also showed that it is possible In the last 10years the topic of electrowetting spawned an increasing
to transfer drops between two parallel plates, which is envisioned to find number of theoretical, numerical, and experimental works. A good deal
applications in three-dimensional digital microfluidics. Bormashenko of fundamental and especially technological issues has been addressed
et al. investigated the deformation of liquid marbles under electric field and solved. The Young–Lippmann equation, despite its simplicity, is
and found that liquid properties influence the deformation [145]. able to quantitatively describe the physical processes and to predict
When a composite liquid marble was exposed to a sufficiently high the wetting behavior of most systems most of the time. The unbound
electric field, the highly polarizable drop climbed on top of the low creativity of researchers, however, continuously finds new ways of
polarizable drop to minimize the total energy of the system. Noblin combining capillary and electric forces with new material properties
and Celestini proposed a new method for controlling water jet dynamics. and system geometries. The Young–Lippmann equation can thus not
Applying an AC voltage between the metallic nozzle generating the jet account for all new observed phenomena. Some issues remain to be
and a metallic electrode below a dielectric polymer layer, they showed solved, most important being the understanding of the saturation
that they could control the reflection angle of the jet, or even to potentials in static and in dynamic electrowetting, of the molecular
completely prevent bouncing. For water, the authors determined a transport processes in the bulk and at the interfaces associated with
maximum threshold frequency of approx. 1.5 kHz above which the applying electric potentials to various liquids, and of the ionic and
effects disappear, since the liquid starts behaving like a dielectric [146]. electrochemical processes taking place at the surface of the electrodes.
Very recently, Yun et al. presented an experimental and numerical The natural contact angle hysteresis inherent to real surfaces leads
study about the prevention of drop rebound from solid surfaces by to poor controllability and sometimes to the irreversibility of the
electrostatic charging of the drops. Electric charges were transferred to electrowetting process. Controllability could be improved either using
water droplets while they fell through variously shaped ring electrodes. liquid/liquid/solid systems, by applying AC instead of DC potentials, or
Such an electrostatic charging affected drop oscillations and thus by using superhydrophobic/superlyophobic surfaces. Polarity effects in
induced a kinetic energy dissipation, which in turn reduced – or even DC electrowetting, generating a non-symmetric wetting behavior for
suppressed – drop rebounding [147]. Most recently, the electrostatic positive and negative voltages, have been found to be most probably
forces were also applied to suppress the Leidenfrost effect [148]. By caused by molecular processes in the liquids. So the use of special ad-
applying an electric field between a Leidenfrost droplet and the heated hoc liquids is recommended in certain applications. AC voltage is applied
substrate on which it was levitating, the vapor layer thickness decreased. more and more in electrowetting, as it can overcome many disadvantages
Eventually, a millimeter-sized drop contacted the heated substrate with of DC electrowetting. However, additional phenomena arise and set
an applied voltage of approx. 40 V. limitations for the applicability, such as the frequency dependence of
the conductivity of the liquid, or the generation of hydrodynamic flows
5.4. Electric enhancement of heat and mass transfer inside the liquid. Such side effects need also to be considered in designing
lab-on-chip devices. Saturation, referring to the existence of a critical
Many authors also reported that the heat and mass transfer can be threshold voltage above which the electrowetting effectiveness does not
controlled by applying an external electric field. Takano et al. further increase, remains a challenge for all electrowetting theories.
experimentally studied the evaporation of various drops of liquids at Since saturation is most probably dominated by microscopic interfacial
L. Chen, E. Bonaccurso / Advances in Colloid and Interface Science 210 (2014) 2–12 11

processes, molecular dynamics simulations could represent a useful tool [41] Blake TD, Haynes JM. Kinetics of liquid/liquid displacement. J Colloid Interface Sci
1969;30:421.
for triggering novel investigations. First papers in this areas appeared, [42] Blake TD. The physics of moving wetting lines. J Colloid Interface Sci 2006;299:1–13.
but more are required. [43] Blake TD, Clarke A, Ruschak KJ. Hydrodynamic assist of dynamic wetting. AlChE J
1994;40:229–42.
[44] Ralston J, Popescu M, Sedev R. Dynamics of wetting from an experimental point of
References view. Annu Rev Mater Res 2008;38:23–43.
[45] Bormashenko E. Contact angles of sessile droplets deposited on rough and flat
[1] Bico J, Tordeux C, Quere D. Rough wetting. Europhys Lett 2001;55:214–20. surfaces in the presence of external fields. Math Model Nat Phenom 2012;7:1–5.
[2] Bonn D, Eggers J, Indekeu J, Meunier J, Rolley E. Wetting and spreading. Rev Mod [46] Das S, Mitra SK, Chakraborty S. Wenzel and Cassie–Baxter states of an electrolytic
Phys 2009;81:739–805. drop on charged surfaces. Phys Rev E 2012;86.
[3] Quere D. Wetting and roughness. Annual review of materials research, vol. 38. Palo [47] Peykov V, Quinn A, Ralston J. Electrowetting: a model for contact-angle saturation.
Alto: Annual Reviews; 2008. p. 71–99. Colloid Polym Sci 2000;278:789–93.
[4] Darhuber AA, Troian SM. Principles of microfluidic actuation by modulation of [48] Jones TB, Fowler JD, Chang YS, Kim CJ. Frequency-based relationship of electrowetting
surface stresses. Annual review of fluid mechanics, vol. 37. Palo Alto: Annual and dielectrophoretic liquid microactuation. Langmuir 2003;19:7646–51.
Reviews; 2005. p. 425–55. [49] Jones TB, Wang KL, Yao DJ. Frequency-dependent electromechanics of aqueous
[5] Sefiane K, Ward CA. Recent advances on thermocapillary flows and interfacial liquids: electrowetting and dielectrophoresis. Langmuir 2004;20:2813–8.
conditions during the evaporation of liquids. Adv Colloid Interface Sci [50] Gupta R, Sheth DM, Boone TK, Sevilla AB, Frechette J. Impact of pinning of the triple
2007;134–35:201–23. contact line on electrowetting performance. Langmuir 2011;27:14923–9.
[6] Verplanck N, Coffinier Y, Thomy V, Boukherroub R. Wettability switching [51] Li F, Mugele F. How to make sticky surfaces slippery: contact angle hysteresis in
techniques on superhydrophobic surfaces. Nanoscale Res Lett 2007;2:577–96. electrowetting with alternating voltage. Appl Phys Lett 2008;92.
[7] Mugele F, Baret JC. Electrowetting: from basics to applications. J Phys Condens [52] Nelson WC, Sen P, Kim CJ. Dynamic contact angles and hysteresis under
Matter 2005;17:R705–74. electrowetting-on-dielectric. Langmuir 2011;27:10319–26.
[8] Quilliet C, Berge B. Electrowetting: a recent outbreak. Curr Opin Colloid Interface [53] Chang JH, Choi DY, Han S, Pak JJ. Driving characteristics of the electrowetting-on-
Sci 2001;6:34–9. dielectric device using atomic-layer-deposited aluminum oxide as the dielectric.
[9] Lippmann G. Relations entre les phenomenes electriques et capillaires. Ann Chim Microfluid Nanofluid 2010;8:269–73.
Phys 1875;5:494. [54] Pollack MG, Fair RB, Shenderov AD. Electrowetting-based actuation of liquid
[10] Möller HG. Electrolytic phenomena at the surfaces of electrodes. Z Phys Chem droplets for microfluidic applications. Appl Phys Lett 2000;77:1725–6.
1908;65:226–54. [55] Pollack MG, Shenderov AD, Fair RB. Electrowetting-based actuation of droplets for
[11] Frumkin A, Gorodetskaya A, Kabanov B, Nekrasov N. Electrocapillary phenomena and integrated microfluidics. Lab Chip 2002;2:96–101.
the wetting of metals by electrolytic solutions, I. Phys Z Sowjetunion 1932;1:255–84. [56] Song JH, Evans R, Lin YY, Hsu BN, Fair RB. A scaling model for electrowetting-on-
[12] Gorodetskaya A, Kabanov B. Electrocapillary phenomena and the wetting of metals dielectric microfluidic actuators. Microfluid Nanofluid 2009;7:75–89.
by electrolytic solutions, II. Phys Z Sowjetunion 1934;5:418–31. [57] Kornyshev AA, Kucernak AR, Marinescu M, Monroe CW, Sleightholme AES, Urbakh
[13] Smolders CA. Contact angles-wetting and dewetting of mercury, Part III. Rec Trav M. Ultra-low-voltage electrowetting. J Phys Chem C 2010;114:14885–90.
Chim 1961;80:699–720. [58] Bhushan B, Pan YL. Role of electric field on surface wetting of polystyrene surface.
[14] Nakamura Y, Kamada K, Katoh Y, Watanabe A. Studies on secondary Langmuir 2011;27:9425–9.
electrocapillary effects. I. The confirmation of Young–Duprè equation. J Colloid [59] Luo MX, Gupta R, Frechette J. Modulating contact angle hysteresis to direct fluid
Interface Sci 1973;44:517–24. droplets along a homogenous surface. ACS Appl Mater Interfaces 2012;4:890–6.
[15] de Bruyn PL, Agar GE. Surface chemistry of flotation. The American Institute of [60] Walker SW, Shapiro B, Nochetto RH. Electrowetting with contact line pinning:
Mining Metallurgical, and Petroleum Engineers, Inc.; 1962 computational modeling and comparisons with experiments. Phys Fluids 2009;21.
[16] Berge B. Electrocapillarite et mouillage de films isolants par l'eau. C R Acad Sci II [61] Maillard M, Legrand J, Berge B. Two liquids wetting and low hysteresis
1993;317:157. electrowetting on dielectric applications. Langmuir 2009;25:6162–7.
[17] Beni G, Hackwood S. Electro-wetting displays. Appl Phys Lett 1981;38:207–9. [62] Heikenfeld J, Dhindsa M. Electrowetting on superhydrophobic surfaces: present
[18] Cho SK, Moon HJ, Kim CJ. Creating, transporting, cutting, and merging liquid status and prospects. J Adhes Sci Technol 2008;22:319–34.
droplets by electrowetting-based actuation for digital microfluidic circuits. J [63] Prins MWJ, Welters WJJ, Weekamp JW. Fluid control in multichannel structures by
Microelectromech Syst 2003;12:70–80. electrocapillary pressure. Science 2001;291:277–80.
[19] Fair RB. Digital microfluidics: is a true lab-on-a-chip possible? Microfluid Nanofluid [64] Verheijen HJJ, Prins MWJ. Reversible electrowetting and trapping of charge: model
2007;3:245–81. and experiments. Langmuir 1999;15:6616–20.
[20] Velev OD, Prevo BG, Bhatt KH. On-chip manipulation of free droplets. Nature [65] Welters WJJ, Fokkink LGJ. Fast electrically switchable capillary effects. Langmuir
2003;426:515–6. 1998;14:1535–8.
[21] Kuiper S, Hendriks BHW. Variable-focus liquid lens for miniature cameras. Appl [66] Fan SK, Yang HP, Wang TT, Hsu W. Asymmetric electrowetting — moving droplets
Phys Lett 2004;85:1128–30. by a square wave. Lab Chip 2007;7:1330–5.
[22] Hayes RA, Feenstra BJ. Video-speed electronic paper based on electrowetting. [67] Koo B, Kim C-J. Evaluation of repeated electrowetting on three different
Nature 2003;425:383–5. fluoropolymer top coatings. J Micromech Microeng 2013;23:067002.
[23] Mugele F. Fundamental challenges in electrowetting: from equilibrium shapes to [68] Moon H, Cho SK, Garrell RL, Kim CJ. Low voltage electrowetting-on-dielectric. J
contact angle saturation and drop dynamics. Soft Matter 2009;5:3377–84. Appl Phys 2002;92:4080–7.
[24] Mugele F, Baret JC, Steinhauser D. Microfluidic mixing through electrowetting- [69] Zimmermann R, Dukhin S, Werner C. Electrokinetic measurements reveal
induced droplet oscillations. Appl Phys Lett 2006;88. interfacial charge at polymer films caused by simple electrolyte ions. J Phys Chem
[25] McHale G, Rowan SM, Newton MI, Banerjee MK. Evaporation and the wetting of a B 2001;105:8544–9.
low-energy solid surface. J Phys Chem B 1998;102:1964–7. [70] Seyrat E, Hayes RA. Amorphous fluoropolymers as insulators for reversible low-
[26] Young T. An essay on the cohesion of fluids. Philos Trans R Soc A 1805;95:65–87. voltage electrowetting. J Appl Phys 2001;90:1383–6.
[27] Berg JC. Wettability. New York: Marcel Dekker, Inc.; 1993. [71] Quinn A, Sedev R, Ralston J. Influence of the electrical double layer in
[28] De Gennes PG. Wetting: statics and dynamics. Rev Mod Phys 1985;57:827–63. electrowetting. J Phys Chem B 2003;107:1163–9.
[29] De Gennes PG, Brochard-Wyart F, Quere D. Capillarity and wetting phenomena. [72] Millefiorini S, Tkaczyk AH, Sedev R, Efthimiadis J, Ralston J. Electrowetting of ionic
Springer; 2004. liquids. J Am Chem Soc 2006;128:3098–101.
[30] Biance AL, Clanet C, Quere D. First steps in the spreading of a liquid droplet. Phys [73] Nanayakkara YS, Moon H, Payagala T, Wijeratne AB, Crank JA, Sharma PS, et al. A
Rev E 2004;69. fundamental study on electrowetting by traditional and multifunctional ionic
[31] Bird JC, Mandre S, Stone HA. Short-time dynamics of partial wetting. Phys Rev Lett liquids: possible use in electrowetting on dielectric-based microfluidic applications.
2008;100. Anal Chem 2008;80:7690–8.
[32] Chen LQ, Auernhammer GK, Bonaccurso E. Short time wetting dynamics on soft [74] Bratko D, Daub CD, Leung K, Luzar A. Effect of field direction on electrowetting in a
surfaces. Soft Matter 2011;7:9084–9. nanopore. J Am Chem Soc 2007;129:2504–10.
[33] Chen LQ, Bonaccurso E, Shanahan MER. Inertial to viscoelastic transition in early [75] Bratko D, Daub CD, Luzar A. Water-mediated ordering of nanoparticles in an
drop spreading on soft surfaces. Langmuir 2013;29:1893–8. electric field. Faraday Discuss 2009;141:55–66.
[34] Winkels KG, Weijs JH, Eddi A, Snoeijer JH. Initial spreading of low-viscosity drops [76] Daub CD, Bratko D, Leung K, Luzar A. Electrowetting at the nanoscale. J Phys Chem
on partially wetting surfaces. Phys Rev E 2012;85. C 2007;111:505–9.
[35] Courbin L, Bird JC, Reyssat M, Stone HA. Dynamics of wetting: from inertial [77] von Domaros M, Bratko D, Kirchner B, Luzar A. Dynamics at a Janus interface. J Phys
spreading to viscous imbibition. J Phys Condens Matter 2009;21. Chem C 2013;117:4561–7.
[36] Muralidhar P, Bonaccurso E, Auernhammer GK, Butt HJ. Fast dynamic wetting of [78] Daub CD, Bratko D, Luzar A. Electric control of wetting by salty nanodrops:
polymer surfaces by miscible and immiscible liquids. Colloid Polym Sci molecular dynamics simulations. J Phys Chem C 2011;115:22393–9.
2011;289:1609–15. [79] 't Mannetje DJCM, Murade CU, van den Ende D, Mugele F. Electrically assisted drop
[37] Huh C, Scriven LE. Hydrodynamic model of steady movement of a solid/liquid/fluid sliding on inclined planes. Appl Phys Lett 2011;98.
contact line. J Colloid Interface Sci 1971;37:85–101. [80] Chevalliot S, Kuiper S, Heikenfeld J. Experimental validation of the invariance
[38] Voinov OV. Hydrodynamics of wetting. Fluid Dyn 1976;11:714. of electrowetting contact angle saturation. J Adhes Sci Technol 2012;26:1909–30.
[39] Cazabat AM. How does a droplet spread. Contemp Phys 1987;28:347–64. [81] Nanayakkara YS, Perera S, Bindiganavale S, Wanigasekara E, Moon H, Armstrong
[40] Tanner LH. The spreading of silicone oil drops on horizontal surfaces. J Phys D Appl DW. The effect of AC frequency on the electrowetting behavior of ionic liquids.
Phys 1979;40:1473–84. Anal Chem 2010;82:3146–54.
12 L. Chen, E. Bonaccurso / Advances in Colloid and Interface Science 210 (2014) 2–12

[82] Paneru M, Priest C, Sedev R, Ralston J. Electrowetting of aqueous solutions of ionic [120] Li H, Paneru M, Sedev R, Ralston J. Dynamic electrowetting and dewetting of ionic
liquid in solid–liquid–liquid systems. J Phys Chem C 2010;114:8383–8. liquids at a hydrophobic solid–liquid interface. Langmuir 2013;29:2631–9.
[83] Paneru M, Priest C, Sedev R, Ralston J. Static and dynamic electrowetting of an ionic [121] Hong J, Kim YK, Kang KH, Oh JM, Kang IS. Effects of drop size and viscosity on
liquid in a solid/liquid/liquid system. J Am Chem Soc 2010;132:8301–8. spreading dynamics in DC electrowetting. Langmuir 2013;29.
[84] Zhang SG, Hu XD, Qu C, Zhang QH, Ma XY, Lu LJ, et al. Enhanced and reversible [122] McHale G, Brown CV, Newton MI, Wells GG, Sampara N. Dielectrowetting driven
contact angle modulation of ionic liquids in oil and under AC electric field. spreading of droplets. Phys Rev Lett 2011;107.
Chemphyschem 2010;11:2327–31. [123] McHale G, Brown CV, Sampara N. Voltage-induced spreading and superspreading
[85] Hong JS, Ko SH, Kang KH, Kang IS. A numerical investigation on AC electrowetting of liquids. Nat Commun 2013;4.
of a droplet. Microfluid Nanofluid 2008;5:263–71. [124] Schneemilch M, Welters WJJ, Hayes RA, Ralston J. Electrically induced changes in
[86] Yoon JY, Garrell RL. Preventing biomolecular adsorption in electrowetting-based dynamic wettability. Langmuir 2000;16:2924–7.
biofluidic chips. Anal Chem 2003;75:5097–102. [125] Puah LS, Sedev R, Fornasiero D, Ralston J. Influence of surface charge on wetting
[87] Kumar A, Pluntke M, Cross B, Baret JC, Mugele F. Finite conductivity effects and kinetics. Langmuir 2010;26:17218–24.
apparent contact angle saturation in AC electrowetting. Mater Res Soc Symp Proc [126] Blake TD, Clarke A, Stattersfield EH. An investigation of electrostatic assist in
2006;899:69–76. dynamic wetting. Langmuir 2000;16:2928–35.
[88] Klarman D, Andelman D, Urbakh M. A model of electrowetting, reversed [127] Zhu XY, Yuan QZ, Zhao YP. Capillary wave propagation during the delamination of
electrowetting, and contact angle saturation. Langmuir 2011;27:6031–41. graphene by the precursor films in electro-elasto-capillarity. Sci Rep 2012;2.
[89] Garcia-Sanchez P, Ramos A, Mugele F. Electrothermally driven flows in AC [128] Wang KL, Jones TB. Electrowetting dynamics of microfluidic actuation. Langmuir
electrowetting. Phys Rev E 2010;81. 2005;21:4211–7.
[90] Ko SH, Lee H, Kang KH. Hydrodynamic flows in electrowetting. Langmuir [129] Baret JC, Decre MMJ, Herminghaus S, Seemann R. Transport dynamics in open
2008;24:1094–101. microfluidic grooves. Langmuir 2007;23:5200–4.
[91] Lee H, Yun S, Ko SH, Kang KH. An electrohydrodynamic flow in AC electrowetting. [130] Khare K, Herminghaus S, Baret JC, Law BM, Brinkmann M, Seemann R. Switching
Biomicrofluidics 2009;3. liquid morphologies on linear grooves. Langmuir 2007;23:12997–3006.
[92] Mampallil D, van den Ende D, Mugele F. Controlling flow patterns in oscillating [131] Seemann R, Brinkmann M, Herminghaus S, Khare K, Law BM, McBride S, et al.
sessile drops by breaking azimuthal symmetry. Appl Phys Lett 2011;99. Wetting morphologies and their transitions in grooved substrates. J Phys Condens
[93] Mugele F, Staicu A, Bakker R, van den Ende D. Capillary stokes drift: a new driving Matter 2011;23.
mechanism for mixing in AC-electrowetting. Lab Chip 2011;11:2011–6. [132] Ristenpart WD, Bird JC, Belmonte A, Dollar F, Stone HA. Non-coalescence of
[94] Oh JM, Legendre D, Mugele F. Shaken not stirred — on internal flow patterns in oppositely charged drops. Nature 2009;461:377–80.
oscillating sessile drops. EPL 2012;98. [133] Yokota M, Okumura K. Dimensional crossover in the coalescence dynamics of
[95] Koopal LK. Wetting of solid surfaces: fundamentals and charge effects. Adv Colloid viscous drops confined in between two plates. Proc Natl Acad Sci U S A
Interface Sci 2012;179:29–42. 2011;108:6395–8.
[96] Sedev R. Electrowetting: electrocapillarity, saturation, and dynamics. Eur Phys J [134] Bird JC, Ristenpart WD, Belmonte A, Stone HA. Critical angle for electrically driven
Spec Top 2011;197:307–19. coalescence of two conical droplets. Phys Rev Lett 2009;103.
[97] Quinn A, Sedev R, Ralston J. Contact angle saturation in electrowetting. J Phys Chem [135] Aryafar H, Kavehpour HP. Electrocoalescence: effects of DC electric fields on
B 2005;109:6268–75. coalescence of drops at planar interfaces. Langmuir 2009;25:12460–5.
[98] Berry S, Kedzierski J, Abedian B. Low voltage electrowetting using thin fluoroploymer [136] Hamlin BS, Creasey JC, Ristenpart WD. Electrically tunable partial coalescence of
films. J Colloid Interface Sci 2006;303:517–24. oppositely charged drops. Phys Rev Lett 2012;109.
[99] Kedzierski J, Berry S. Engineering the electrocapillary behavior of electrolyte [137] Mousavichoubeh M, Ghadiri M, Shariaty-Niassar M. Electro-coalescence of an
droplets on thin fluoropolymer films. Langmuir 2006;22:5690–6. aqueous droplet at an oil–water interface. Chem Eng Process 2011;50:338–44.
[100] Gupta R, Olivier GK, Frechette J. Invariance of the solid–liquid interfacial energy in [138] Mousavichoubeh M, Shariaty-Niassar M, Ghadiri M. The effect of interfacial tension
electrowetting probed via capillary condensation. Langmuir 2010;26:11946–50. on secondary drop formation in electro-coalescence of water droplets in oil. Chem
[101] Revilla RI, Guan L, Zhu XY, Quan BG, Yang YL, Wang C. Electrowetting phenomenon Eng Sci 2011;66:5330–7.
on nanostructured surfaces studied by using atomic force microscopy. J Phys Chem [139] Chen LQ, Wang X, Wen WJ, Li ZG. Critical droplet volume for spontaneous capillary
C 2012;116:14311–7. wrapping. Appl Phys Lett 2010;97.
[102] Buehrle J, Herminghaus S, Mugele F. Interface profiles near three-phase contact [140] Guo XY, Li H, Ahn BY, Duoss EB, Hsia KJ, Lewis JA, et al. Two- and three-dimensional
lines in electric fields. Phys Rev Lett 2003;91. folding of thin film single-crystalline silicon for photovoltaic power applications.
[103] Mugele F, Buehrle J. Equilibrium drop surface profiles in electric fields. J Phys Proc Natl Acad Sci U S A 2009;106:20149–54.
Condens Matter 2007;19. [141] Py C, Reverdy P, Doppler L, Bico J, Roman B, Baroud CN. Capillary origami:
[104] Liu J, Wang M, Chen S, Robbins MO. Uncovering molecular mechanisms of spontaneous wrapping of a droplet with an elastic sheet. Phys Rev Lett 2007;98.
electrowetting and saturation with simulations. Phys Rev Lett 2012;108:216101. [142] Pineirua M, Bico J, Roman B. Capillary origami controlled by an electric field. Soft
[105] Santos LP, Ducati TRD, Balestrin LBS, Galembeck F. Water with excess electric Matter 2010;6:4491–6.
charge. J Phys Chem C 2011;115:11226–32. [143] Wang ZQ, Wang FC, Zhao YP. Tap dance of a water droplet. Proc R Soc Ser A Math
[106] Aronov D, Molotskii M, Rosenman G. Electron-induced wettability modification. Phys Eng Sci 2012;468:2485–95.
Phys Rev B 2007;76. [144] Lee SJ, Lee S, Kang KH. Droplet jumping by electrowetting and its application to the
[107] Drygiannakis AI, Papathanasiou AG, Boudouvis AG. On the connection between three-dimensional digital microfluidics. Appl Phys Lett 2012;100.
dielectric breakdown strength, trapping of charge, and contact angle saturation in [145] Bormashenko E, Pogreb R, Balter R, Gendelman O, Aurbach D. Composite non-stick
electrowetting. Langmuir 2009;25:147–52. droplets and their actuation with electric field. Appl Phys Lett 2012;100.
[108] Papathanasiou AG, Papaioannou AT, Boudouvis AG. Illuminating the connection [146] Noblin X, Celestini F. Electrowetting control of bouncing jets. Appl Phys Lett 2012;101.
between contact angle saturation and dielectric breakdown in electrowetting [147] Sungchan Y, Jiwoo H, Kwan Hyoung K. Suppressing drop rebound by electrically
through leakage current measurements. J Appl Phys 2008;103. driven shape distortion. Phys Rev E Stat Nonlin Soft Matter Phys 2013;87
[109] Berry S, Kedzierski J, Abedian B. Irreversible electrowetting on thin fluoropolymer [033010 (5 pp.)-(5pp.)].
films. Langmuir 2007;23:12429–35. [148] Celestini F, Kirstetter G. Effect of an electric field on a Leidenfrost droplet. Soft
[110] Vallet M, Vallade M, Berge B. Limiting phenomena for the spreading of water on Matter 2012;8:5992–5.
polymer films by electrowetting. Eur Phys J B 1999;11:583–91. [149] Takano K, Tanasawa I, Nishio S. Active enhancement of evaporation of a liquid drop
[111] Mugele F, Herminghaus S. Electrostatic stabilization of fluid microstructures. Appl on a hot solid surface using a static electric field. Int J Heat Mass Transfer
Phys Lett 2002;81:2303–5. 1994;37:65–71.
[112] Bhaumik SK, Chakraborty M, Ghosh S, Chakraborty S, DasGupta S. Electric [150] Takano K, Tanasawa I, Nishio S. Enhancement of evaporation of a liquid droplet
field enhanced spreading of partially wetting than liquid films. Langmuir using EHD effect: criteria for instability of gas–liquid interface under electric field.
2011;27:12951–9. J Enhanc Heat Transf 1996;3:73–81.
[113] Park J, Feng XQ, Lu W. Instability of electrowetting on a dielectric substrate. J Appl [151] Takano K, Tanasawa I, Nishio S. Enhancement of evaporation of a droplet using
Phys 2011;109. EHD effect — (measurement of steady-state heat flux during evaporation of a single
[114] Castaner L, Di Virgilio V, Bermejo S. Charge-coupled transient model for droplet). JSME Int J Ser B 1996;39:583–9.
electrowetting. Langmuir 2010;26:16178–85. [152] Vancauwenberghe V, Di Marco P, Brutin D. Wetting and evaporation of a sessile drop
[115] Di Virgilio V, Bermejo S, Castaner L. Wettability increase by “Corona” ionization. under an external electrical field: a review. Colloids Surf A Physicochem Eng Asp
Langmuir 2011;27:9614–20. 2013;432.
[116] Lin JL, Lee GB, Chang YH, Lien KY. Model description of contact angles in [153] Butt HJ, Untch MB, Golriz A, Pihan SA, Berger R. Electric-field-induced condensation:
electrowetting on dielectric layers. Langmuir 2006;22:484–9. an extension of the Kelvin equation. Phys Rev E 2011;83.
[117] Chen LQ, Li CL, van der Vegt NFA, Auernhammer GK, Bonaccurso E. Initial [154] Sugimura H, Uchida T, Kitamura N, Masuhara H. Tip‐induced anodization of
electrospreading of aqueous electrolyte drops. Phys Rev Lett 2013;110. titanium surfaces by scanning tunneling microscopy: a humidity effect on
[118] Yuan QZ, Zhao YP. Precursor film in dynamic wetting, electrowetting, and electro- nanolithography. Appl Phys Lett 1993;63:1288–90.
elasto-capillarity. Phys Rev Lett 2010;104. [155] Garcia R, Calleja M, Perez-Murano F. Local oxidation of silicon surfaces by dynamic
[119] Decamps C, De Coninck J. Dynamics of spontaneous spreading under electrowetting force microscopy: nanofabrication and water bridge formation. Appl Phys Lett
conditions. Langmuir 2000;16:10150–3. 1998;72:2295–7.

You might also like