Download as pdf or txt
Download as pdf or txt
You are on page 1of 10

Journal of Molecular Liquids 333 (2021) 116039

Contents lists available at ScienceDirect

Journal of Molecular Liquids


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

Dynamic spreading of a water nanodroplet on a nanostructured surface


in the presence of an electric field
Ben-Xi Zhang, Shuo-Lin Wang, Xin He, Yan-Ru Yang, Xiao-Dong Wang ⇑
State Key Laboratory of Alternate Electrical Power System with Renewable Energy Sources, North China Electric Power University, Beijing 102206, China
Research Center of Engineering Thermophysics, North China Electric Power University, Beijing 102206, China

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

Article history: Molecular dynamics simulations are implemented to investigate the statics and dynamics of wetting for a
Received 19 January 2021 water nanodroplet on a nanostructured surface in the presence of a vertical electric field. The results
Revised 25 March 2021 show that the electric field induces electro-stretching, electro-wettability, modified solid-liquid interfa-
Accepted 30 March 2021
cial tension, and pinning at the triple line, and they jointly affect the spreading exponent and static con-
Available online 6 April 2021
tact angle of the nanodroplet. Under an upward electric field, the spreading is always hindered, and the
static contact angle monotonously increases with the field strength. Interestingly, under a downward
Keywords:
electric field, the spreading is first hindered and then is promoted as increasing the field strength, leading
Polar water molecules
Electro-wettability
to a first increased and then decreased static contact angle. The increased solid-liquid interfacial tension
Electro-stretching and the enhanced pinning are found to be two main mechanisms for the slowing down of spreading and
Pinning the increase in static contact angles for both electric fields at low field strengths. The same trends are
Hydrogen bonds observed at high field strengths under the upward electric field; however, exactly the reverse trends
occur under the downward electric field, leading to the acceleration of spreading and the decrease in sta-
tic contact angle at high field strengths. Moreover, it is found that the enhancement in intrinsic wettabil-
ity of the surface can suppress the breaking of hydrogen bonds, thereby reducing the solid-liquid
interfacial tension and accelerating the spreading. The enhancement in intrinsic wettability can also
reduce the pinning at the triple line as well as the energy barrier of the Cassie-Wenzel transition, which
contributes to the reduced static contact angle.
Ó 2021 Elsevier B.V. All rights reserved.

1. Introduction adsorption and desorption events of liquid molecules on solid sur-


faces, and hence, follows a different scaling law of r ~ t1/7 [7–9].
Dynamic wetting occurs when a droplet spreads over a solid With the development of surface fabrication technologies [10],
surface, which is of great significance in many practical applica- various textured surfaces with nano/microstructures are widely
tions, such as inkjet printing, surface coating, boiling, and conden- employed to manipulate dynamic wetting behaviors of droplets.
sation [1]. Two kinds of classical models, hydrodynamics (HD) On such surfaces, droplets can suspend on or collapse into nano/
models [2] and molecular kinetic theory (MKT) models [3], have microstructures, forming to the Cassie or Wenzel state [11,12].
been proposed to interpreted the spreading kinetics of droplets Because the size of macrodroplets is serval orders of magnitude
on solid surfaces. Within the framework of HD models, the larger than the characteristic size of nano/microstructures, nano/
dynamic wetting is driven by capillary force or gravitational force, microstructures only modify the surface wettability. However, for
and is resisted by viscous force. Depending on the driving force, the nanodroplets, the droplet size is comparable to the characteristic
spreading obeys a scaling law of r ~ t1/10 in the capillary regime and size of nanostructures, so that nanostructures not only change
r ~ t1/8 in the gravitational regime [4–6]. However, within the the surface wettability but also may cause a significantly enhanced
framework of MKT models, the dynamic wetting is attributed to pinning effect. This enhanced effect is recently confirmed by an
analysis of the local atomic density in the vicinity the triple line
[13]. Thus, an extra energy dissipation is generated by the local
pinning effect, leading to the distinctly different spreading kinetics
⇑ Corresponding author at: State Key Laboratory of Alternate Electrical Power of nanodroplets on nanostructured surfaces. Zong et al. [14] inves-
System with Renewable Energy Sources, North China Electric Power University, tigated the spreading of water nanodroplets on hydrophilic nanos-
Beijing 102206, China.
tructured surfaces. They found that because of the strong pinning
E-mail address: wangxd99@gmail.com (X.-D. Wang).

https://doi.org/10.1016/j.molliq.2021.116039
0167-7322/Ó 2021 Elsevier B.V. All rights reserved.
Ben-Xi Zhang, Shuo-Lin Wang, X. He et al. Journal of Molecular Liquids 333 (2021) 116039

effect, the movement of water molecules is much more directional the Wenzel state, the electro-wettability and pinning effects
on nanostructured surfaces than that on smooth surfaces. The dominate over the electro-stretching effect, leading to a decreased
directional movement causes an extra dissipation in the bulk liq- static contact angle.
uid, hindering spreading. Yuan et al [15] demonstrated that when There are electro-wettability, electro-stretching, and pinning
a nanodroplet spreads over a hydrophilic textured surface with effects on nanostructured surfaces, and they jointly affect the stat-
flexible nanopillars, the liquid is accelerated by nanopillars when ics and dynamics of wetting of nanodroplets. However, the under-
approaching nanopillars but is pinned when passing nanopillars. lying mechanism remains poorly understood. In this work, MD
Owing to these two effects, the spreading obeys a scaling law of simulation are implemented to study the statics and dynamics of
r ~ t1/3 in the early stage, whereas the spreading exponent wetting of nanodroplets on a nanostructured surface in the pres-
decreases in the late stage because the pinning force gradually ence of a vertical electric field. The field strength, field direction,
becomes dominant as the nanodroplet are impaled on more and intrinsic wettability of the surface are altered in simulations
nanopillars. These two effects are also found on hydrophilic nanos- to reveal the effects of electro-wettability, electro-stretching, and
tructured surfaces with rigid nanopillars, and hence, the scaling pinning.
law of r ~ t1/3 occurs again on such surfaces [16]. The nanogrooves
between nanopillars provide an additional driving force to acceler- 2. Simulation method
ate the infiltration of suspended liquid. On the other hand, nanopil-
lars also bring extra resistance to the movement of collapsed 2.1. Interaction potentials
liquid. The competition between the two opposite effects affects
the advancing and receding of the triple line as well as the static The spreading of a water nanodroplet on a gold nanostructured
apparent contact angle of nanodroplets. As a result, a reasonable surface is studied. The simple- point charge/extension (SPC/E)
design of nanostructures is very important for the manipulation model is employed to characterize the properties of water mole-
of dynamic wetting of nanodroplets. cules, which is kept rigid by the SHAKE algorithm [27]. This model
In parallel with nanostructures, applying an external electric was also employed in previous studies of nanodroplets in the pres-
field to nanodroplets is also a feasible approach to manipulate nan- ence of an external electric field [28–30]. The interactions between
odroplets because the electric field can modify the static contact water molecules consist of the Lennard-Jones (L-J) 12-6 and
angle of nanodroplets. For macrodroplet, the modified contact Coulombic potentials, expressed as,
angle is described by the Lippmann-Young equation, " 12  6 #
coshe = coshY+(ere0V2)/2dclv, where he is the static contact angle rij rij qi qj
when voltage V is applied, hY is the Young contact angle, clv is U ij ¼ 4eij  þ ð1Þ
r ij r ij r ij
the liquid-vapor interfacial tension, d is the electric double layer
thickness, er is the dielectric constant of the insulator layer, and where Uij is the pair potential between the ith and jth atoms, rij is
e0 is the permittivity of vacuum [17]. However, this equation can- the distance between the ith and jth atoms, eij and rij are the energy
not describe the static contact angle of nanodroplets in the pres- and distance parameters, and qi and qj are the charges of the ith and
ence of electric fields because there is a scale effect for jth atoms.
electrowetting of nanodroplets. The electric double layer that can A gold (100) nanostructured substrate is constructed by face-
shield the electric field hardly forms for nanodroplets, and hence, centered cubic (FCC) crystals, in which the gold atoms in the bot-
nanodroplets are more sensitive to the electric field than macro- tom three layers are fixed to prevent the substrate deformation.
droplets [18–22]. Furthermore, the enhanced electric field effect The interactions between gold atoms are described by the embed-
also causes a direction-dependent static contact angle of nan- ded atom method potential, expressed as,
odroplets. Zong et al. [23] presented that for a water nanodroplet
X 1XX  
spreading over a smooth surface subjected to a vertical electric Ei ¼ F i ðqi Þ þ / r ij ð2Þ
field, the solid-liquid interfacial tension decreases in the presence i
2 i j–i ij
of an upward electric field, whereas it increases when a downward
electric field is applied. As a result, a larger static contact angle is where Ei is the energy of ith atom, Fi is the embedding energy that is
observed for the water nanodroplet under the downward electric a function of the atomic electron density qi, and /ij is a pair poten-
field when the field strength remains unchanged. Song et al. [24] tial. There are three main force field parameters, a, b, and t, for the
investigated the spreading of a nanodroplet on a smooth surface Au-Au interactions, which are taken from Ref. [31] and listed in
subjected to a horizontal electric field. They found asymmetric sta- Table 1. The interactions between water molecules and gold atoms
tic contact angles on two sides of the nanodroplet. The asymmetry are modeled by the L-J 12-6 potential. The intrinsic wettability is
first increases and then decreases with the field strength, and associated with the interactions between liquid and solid, the
finally disappears when the field strength exceeds a critical value. strong solid-liquid interactions can enhance the intrinsic wettabil-
This interesting phenomenon is explained by a combined ity. On the basis of such a fact, the intrinsic wettability is adjusted
interaction of electro-wettability, electro-stretching, and inter- by altering the energy parameter eO-Au, as shown in Table 1.
molecular forces.
Recently, several studies also paid their attention to the spread- 2.2. Simulated system
ing of a nanodroplet over a nanostructured surface in the presence
of an electric field. Zhao et al. [25] demonstrated that as the field Fig. 1 shows the schematic of the simulated system, which is a
strength increases, a suspended nanodroplet eventually collapses cuboid box with dimensions of 57 nm (x)  1.6 nm (y)  80 nm (z).
into a superhydrophobic nanostructured surface, corresponding The boundary condition applied to the box is periodic in the x- and
to a wetting transition from the Cassie state to the Wenzel state. y- directions, whereas it is fixed in the z-direction, with a reflect
Yen et al. [26] reported that for a nanodroplet in the Cassie state wall being assigned to the top of the box. When atoms collide with
subjected to a vertical electric field, the small solid-liquid contact the reflect wall, they will bounce back to the box without the loss
area weakens the electro-wettability effect, so that the electro- of kinetic energy. In the present simulations, a two-dimensional
stretching effect becomes dominant, which lifts up the droplet cylindrical droplet is adopted to save the computational cost.
and increases the static contact angle as compared with that in Although the cylindrical droplet is nonexistent in the nature, the
the absence of the electric field; however, for a nanodroplet in study of its spreading is still of great importance to understand
2
Ben-Xi Zhang, Shuo-Lin Wang, X. He et al. Journal of Molecular Liquids 333 (2021) 116039

Table 1
Values of the potential parameters employed in the present simulations.

Particles i j ri,j (Å) eij (eV) qe (eV) h (°) a b t


O-O 3.1660 0.0068 0.8476 — — — —
H-H 0.0000 0.0000 +0.4238 — — — —
O-H 0.0000 0.0000 — — — — —
O-Au 2.8675 0.0214 — 60.0 — — —
O-Au 2.8675 0.0266 — 47.0 — — —
O-Au 2.8675 0.0312 — 42.0 — — —
O-Au 2.8675 0.0340 — 32.0 — — —
H-Au 0.0000 0.0000 — — — — —
Au-Au — — — — 1.4475 0.1269 2

Fig. 1. The initial configuration of the simulated system. Gold atoms are yellow, oxygen atoms are blue, and hydrogen atoms are pink.

to real spreading processes. This is because cylindrical and tem, so that an additional force fe,i = qiE is exerted on each charged
spherical droplets follows the same spreading law. A water nan- atom. After the initial configuration is constructed, simulations are
odroplet with radius of R0 = 45 Å (3500 water molecules) and a implemented using the LAMMPS (large-scale atomic/molecular
gold substrate (23496 gold atoms) decorated with nanogrooves massively parallel simulator). The PPPM (particle-particle
are mounted in the box. The thickness of the cylindrical nan- particle-mesh) method is employed to solve the long-range elec-
odroplet is taken as 16 Å based on the requirement of periodic trostatic interactions with an accuracy of 10-4 [32], and the
boundary conditions. The gold substrate has a thickness of 24 Å, velocity-Verlet algorithm is employed to solve the Newtonian
and the nanogrooves in the gold substrate have a constant aspect motion equations of particles at a time step of t = 1 fs [33]. The con-
ratio of h/w = 1 (height h = 1.2 nm and width w = 1.2 nm). tact angle and spreading radius are calculated based on a free-
The water nanodroplet and gold substrate are respectively surface fitting method. The nanodroplet above the nanogrooves
relaxed in the NVT ensemble at 298 K to reach equilibrium. After is divided into many layers with a layer thickness of 3 Å. The num-
that, a uniform electric field in the z-direction is applied to the sys- ber density of water molecules in each layer is calculated as a func-
3
Ben-Xi Zhang, Shuo-Lin Wang, X. He et al. Journal of Molecular Liquids 333 (2021) 116039

tion of the distance from the center of mass of the layer. The radius hydrogen atoms facing towards the gold surface but negatively
R(t) in each layer is determined by the number density, namely charged oxygen atoms against the surface [24]. Owing to this
98% of water molecules in the layer are contained. The profile of directional arrangement of water molecules, the nanodroplet
the nanodroplet is fitted by the radius in each layer, and the con- would be elongated along the field direction, which is referred to
tact angle is calculated by the fitted profile. as an electro-stretching effect [34,35]. Thus, the nanodroplet shape
is controlled by both electro-stretching and interfacial tensions,
forming an ellipsoid in the early stage of spreading, as shown in
3. Results and discussion Fig. 2(b). Because only a small field strength of E = 0.05 V Å1
is employed in this case, the electro-stretching effect is relatively
3.1. Electrowetting of nanodroplet on nanostructured surface weak, and hence, the droplet finally evolves into a hemisphere,
as the same as that in the absence of the electric field.
Simulations are first implemented with the following fixed con- Besides the electro-stretching effect, applying an external elec-
ditions to understand the statics and dynamics of electrowetting tric field to nanodroplets also produces an electro-wettability
on nanostructured surfaces. The aspect ratio of nanogrooves is h/ effect. The directional arrangement of water molecules caused by
w = 1, and the intrinsic wettability of the nanostructured surface electric fields may break hydrogen bonds formed between water
is h0 = 60°, corresponding to the energy parameter of eO- molecules [24]. The bond breaking promotes the migration of
Au = 0.0214 eV. A downward electric field with field strength of water molecules to liquid and solid surfaces because they are no
E = 0.05 V Å1 is applied to the nanodroplet. Fig. 2 presents the longer constrained by hydrogen bonds. As a consequence of the
snapshots of a water nanodroplet spreading over the nanostruc- migration, water molecules more readily infiltrate nanogrooves
tured surface. As shown in Fig. 2(a), when no electric field is and wet solid surfaces, which is referred to as the electro-
applied, the bottom liquid first infiltrates nanogrooves to form a wettability effect [36]. It is worth noting that the static contact
composite surface composed of nanopillars and collapsed liquid, angle of the nanodroplet subjected to the electric field is larger
whereas the bulk liquid remains nearly stationary until three than that in the absence of the electric field. This result is com-
nanogrooves are infiltrated by the collapsed liquid at t = 70 ps. It pletely opposite to that of macrodroplets, indicating that the
should be noted that although the nanodroplet is in the Wenzel electro-stretching effect outweighs the electro-wettability effect.
state, it still remains nearly spherical. Subsequently, accompanied The force balance in the horizontal direction at the triple line
with the further infiltration, the bulk liquid starts to spread over can be expressed as clvcoshe = Few + Fc-Fp [37], where Few is the
the composite surface, and the droplet evolves into a hemisphere. electro-wettability force, Fc = csv-csl is the capillary force, Fp is
Finally, when six nanogrooves are fully infiltrated, the spreading the pinning force, csv is the solid-gas interfacial tension, and csl is
stops and the nanodroplet has a static contact angle of he = 53°. the solid-liquid interfacial tension. In general, the pinning force
However, as shown in Fig. 2(b), in the presence of the electric field, can be ignored for macrodroplets when the droplet size is much
the bottom liquid fills up three nanogrooves at t = 60 ps, and the larger than the microstructure size, so that the equation is reduced
nanodroplet evolves into an ellipsoid instead of the sphere in the to clvcoshe = Few + Fc, which can be considered as a variant of the
absence of the electric field. When the nanodroplet reaches equi- classical Young-Lippmann equation. However, the pinning effect
librium, only five nanogrooves are fully infiltrated, leading a larger becomes significantly strong and hence cannot be ignored for nan-
static contact angle of he = 67°. odroplets. Daub et al. [22] presented that the solid-liquid interfa-
Hydrogen and oxygen atoms have different electronegativity cial tension for nanodroplets can be related to the average
values, and hence, water molecules are polar with a partial positive number of hydrogen bonds per molecule, and a larger hydrogen
charge for each hydrogen atom and a partial negative charge for bond number means a lower solid-liquid interfacial tension.
each oxygen atom. Water molecules can be approximately consid- Fig. 3 shows the average number of hydrogen bonds, <nHB>, per
ered as point dipoles because of their small sizes. In the absence of water molecule in the vicinity of the solid-liquid interface. The fol-
an external electric field, point dipoles are randomly oriented lowing three conditions are used to distinguish whether a hydro-
owing to thermal motions, so that the nanodroplet shape is gov- gen bond forms between two water molecules. First, the
erned only by interfacial tensions, as shown in Fig. 2(a). However, distance, RO-O, between the oxygen atoms of two water molecules
when an electric field is applied to the nanodroplet, electric field is smaller than the critical one, RO-O,c. Second, the distance, RO-H,
forces make point dipoles reorient themselves. Under the down- between the hydrogen atom accepting a lone pair of electrons
ward electric field, the reorientation leads to positively charged and the oxygen atom donating a lone pair of electrons is less than

(a)

0 ps 26 ps 70 ps 170 ps 512 ps 1666 ps 2000 ps

(b)

0 ps 30 ps 60 ps 202 ps 524 ps 1574 ps 2000 ps


Fig. 2. Morphology evolution of a water nanodroplet on a nanostructured surface: (a) in the absence of an electric field and (b) in the presence of a downward electric field
with field strength of E = 0.05 V Å1.

4
Ben-Xi Zhang, Shuo-Lin Wang, X. He et al. Journal of Molecular Liquids 333 (2021) 116039

3.20 Downward Applying an electric field also affects the dynamics of spreading
Upward [41–44]. The evolution of spreading radius for the water nan-
3.18
odroplet with or without an electric field is shown in Fig. 5. The
3.16 spreading can be divided into two stages. In an early spreading
3.14 stage, the spreading is not affected by the electric field, whereas
the spreading is deaccelerated in the later spreading stage when
<nHB>

3.12
a downward electric field is applied to the droplet.
3.10
3.08 3.2. Effect of field strength
3.06
3.04 The effect of field strength on the statics and dynamics of wet-
ting for a water nanodroplet on a nanostructured surface is inves-
3.02
tigated under the following conditions. The nanogrooves have an
0 0.03 0.05 0.07 0.09 aspect ratio of h/w = 1 and intrinsic wettability of h0 = 60°. A down-
ward electric field is applied to the nanodroplet with field
E (V Å-1) strengths of E = 0.03, 0.05, 0.07, and 0.09 V Å1. Figs. 2
and 6 show the snapshots of the nanodroplet with and without
Fig. 3. Average number of hydrogen bonds per water molecule in the vicinity of the
solid-liquid interface in the presence of an upward or a downward electric field. the electric field during spreading. When no electric field is
applied, six nanogrooves are eventually infiltrated by the collapsed
liquid; however, there are only five for the field strengths of
the critical one, RO-H,c. Third, the OOH bond angle, /, is less than a
E = 0.03, 0.05, and 0.07 V Å1. Interestingly, when the field
critical one, /c. The values of the critical parameter are taken as RO-
strength further increases to E = 0.09 V Å1, the number of infil-
O,c = 3.6 Å, RO-H,c = 2.4 Å and /c = 30° [26]. The thickness of water
trated nanogrooves increases to six, as the same as that in the
molecular layer is 5 Å, which is larger than the hydrogen bond
absence of the electric field.
length of 2.8 Å. As shown in Fig. 3, <nHB> significantly decreases
Fig. 7 shows the static contact angle as a function of field
at the field strength of E = 0.05 V Å1, as compared with that in
strength for the water nanodroplet on the nanostructured surface
the absence of the electric field. Therefore, the solid-liquid interfa-
in the presence of a downward electric field. To verify the reliabil-
cial tension increases, which leads to a reduced Fc and hence an
ity of data, the static contact angles of droplets with radii of R0 = 30
increased static contact angle.
and 51 Å in the presence of a downward electric field are also sim-
The pinning effect hinders the movement of the triple line,
ulated and drawn in Fig. 7. The data show good agreement for the
which makes water molecules accumulate in the vicinity of the tri-
three nanodroplets. As expected, the static contact angle increases
ple line, thereby leading to a high local density of water molecules.
when the field strength changes from E = 0 to 0.05 V Å1. How-
Fig. 4 shows the number density of water molecules in the vicinity
ever, an abnormal result occurs when further increasing the down-
of the triple line with and without an electric field. Here, the sim-
ward field strength. The static contact angle starts to decreases
ulation box is divided into multiple small cells with dimensions of
when the field strength varies from E = 0.05 to 0.09 V Å1.
0.3 nm (x)  0.3 nm (z), the number density in each cell is calcu-
The static contact angle is 35.5° at E = 0.09 V Å1, which
lated. A significantly higher number density is observed at E = 0
decreases by 18° as compared with that in the absence of the elec-
.05 V Å1 than that in the absence of the electric field, which agrees
tric field.
with the previous reports [24,38]. The increased number density
Because the mechanisms for static contact angles at low down-
provides evidence for the enhanced pinning effect, which was also
ward field strengths have been discussed in Section 3.1, only the
experimentally reported by Ren et al. [39] and Baviere et al. [40].
abnormal variation of static contact angles at high field strengths
Therefore, the enhancement of pinning effect is another reason
are analyzed here. As shown in Fig. 3, the average number of
for the increased static contact angle of nanodroplets on nanos-
hydrogen bonds per water molecule in the vicinity of the solid-
tructured surfaces when an external electric field is applied.
liquid interface increases when the downward field strength is lar-
ger than 0.05 V Å1, which causes a reduced solid-liquid interfacial

Downward
14 100
Upward
90
80
12 70
60
50
(N Å-3 )

10
40
R (Å)

8
30
E= 0 V Å-1
6
20 E= -0.03 V Å-1
E= -0.05 V Å-1
4 E= -0.07 V Å-1
E= -0.09 V Å-1
0 0.03 0.05 0.07 0.09
10
0.1 1 4
E (V Å ) -1

t (ns)
Fig. 4. The number density of water molecules in the vicinity of the triple line on
the nanostructured surface in the presence of an upward or a downward electric Fig. 5. Evolution of spreading radius of a water nanodroplet on a nanostructured
field. surface in the presence of a downward electric field.

5
Ben-Xi Zhang, Shuo-Lin Wang, X. He et al. Journal of Molecular Liquids 333 (2021) 116039

(a)

0 ps 32 ps 70 ps 122 ps 220 ps 1418 ps 2000 ps

(b)

0 ps 30 ps 60 ps 310 ps 344 ps 1320 ps 2000 ps

(c)

0 ps 30 ps 60 ps 322 ps 404 ps 1052 ps 2000 ps


Fig. 6. Morphology evolution of a water nanodroplet on a nanostructured surface in the presence of a downward electric field at: (a) E = 0.03 V Å1; (b) E = 0.07 V Å1; and
(c) E = 0.09 V Å1.

90 2.0
E= 0 VÅ
-1
30 Å Downward
1.8 E = -0.03 V Å
-1

80 45 Å Downward
E = -0.05 V Å
-1
51 Å Downward 1.6
E = -0.07 V Å
-1

70 45 Å Upward 1.4
substrate

E = -0.09 V Å
-1

1.2
θ (°)

-3

60

1.0

50 0.8
0.6
40 0.4
0.2
30
0.0
0 0.03 0.05 0.07 0.09 0 20 40 60 80 100 120 140 160

Z (Å )
3
E (V Å-1)
Fig. 8. The number density of water molecules in the z-direction in the presence of
Fig. 7. The static contact angle as a function of field strength in the presence of an
a downward electric field.
upward or a downward electric field.

increase in field strength, namely it first decreases and then


tension and hence an increased capillary force Fc. Thus, the increases with increasing the downward field strength. As shown
decreased static contact angle for the downward field strength lar- in Fig. 4, when the downward field strength is larger than
ger than 0.05 V Å1 can be partially attributed to the increase in the 0.05 V Å1, the local number density of water molecules in the
average number of hydrogen bonds per water molecule in the vicinity of the triple line starts to decrease. This result indicates a
vicinity of the solid-liquid interface. This increased hydrogen bond weaker pinning effect, which enhances infiltration of the sus-
number at a strong downward electric field can be explained as fol- pended liquid into nanogrooves and thus promotes spreading of
lows. When field strengths are large enough, more hydrogen bonds the nanodroplet. Accordingly, the weaker pinning also contributes
in the bulk nanodroplet break owing to the directional arrange- to the decreased static contact angle at high downward field
ment of water molecules. The water molecules losing hydrogen strengths.
bonds are more likely move to the solid-liquid interface and form
hydrogen bonds again, leading to the increased hydrogen bond 3.3. Effect of field direction
number in the vicinity of the solid-liquid interface. The averge
number density of water molecules in the z-direction is also It is widely recognized that electrowetting of macrodroplets
extracted. Here, the nanodroplet is divided into multiple layers does not depend on the field direction. However, a recent study
with a 3 Å layer thickness, and the number density in each layer has shown that with the same field strength, static contact angles
is calculated. As shown in Fig. 8, the number density near the of a nanodroplet on a smooth surface are not equal when upward
solid-liquid interface follows the same trend as that with an and downward electric fields are applied [23]. In this section, the
6
Ben-Xi Zhang, Shuo-Lin Wang, X. He et al. Journal of Molecular Liquids 333 (2021) 116039

effect of the field direction is further examined on a nanostruc- bulk liquid to the liquid-gas or solid-liquid interfaces is hardly
tured surface. Fig. 9 shows the snapshots of nanodroplet spreading affected by the electric field. The reduction in the number of hydro-
in the presence of an upward electric field. The comparison among gen bonds in the vicinity of the solid-liquid interface arises only
Figs. 2, 6 and 9 indicates that there is no significant difference from applying the electric field. Therefore, the average number of
between upward and downward electric fields at low field hydrogen bonds in this region are almost the same under the
strengths. Five nanogrooves are infiltrated by the equilibrated nan- upward and downward electric fields when the field strength
odroplet at E = 0.03 and 0.05 V Å1 regardless of the field direction. remains constant, as shown in Fig. 3. At high field strengths, more
However, the difference starts to emerge at high field strengths, hydrogen bonds break in the bulk liquid, increasing the chance of
especially at E = 0.09 V Å1 six nanogrooves are infiltrated under migration of water molecules to the interfaces. The difference
the downward electric field, whereas only four nanogrooves are between the upward and downward electric fields lies that water
infiltrated under the upward electric field. Another significant dif- molecules in the bulk liquid more likely migrate to the solid-
ference at E = 0.09 V Å1 lies in that the equilibrated nanodroplet is liquid interface and forms hydrogen bonds again under the down-
a hemisphere under the downward electric field, whereas it is ward field, but to the liquid-gas interface under the upward field.
elongated vertically, forming a hemi-ellipsoid under the upward As the result of different movement trajectory of water molecules,
electric field. the average number of hydrogen bonds increases at high field
As shown in Fig. 7, unlike under the downward electric field, the strengths under the downward electric field but decreases under
static contact angle always increases with the field strength under the upward electric field, as shown in Fig. 3. Thus, the solid-
the upward electric field. Almost the same static contact angles are liquid interfacial tension always increases with the field strength
observed under both electric fields when E  0.05 V Å1, above under the upward electric field, leading to the continuously
which a larger static contact angle occurs under the upward elec- increased static contact angles. Moreover, the local number den-
tric field. The static contact angle is 76.5° at E = 0.09 V Å1 under sity, as shown in Fig. 4, indicates that the pinning effect is always
the upward electric field, 41° higher than that under the down- enhanced under the upward field, and the pinning force monoto-
ward electric field. nously increases with the field strengths, which is another reason
The difference at high field strengths between the upward and for the continuously increased static contact angles under the
downward electric fields can be partially attributed to different upward electric field.
movement trajectory of water molecules. At low field strengths, Fig. 10 shows the evolution of spreading radius for the water
only a small number of hydrogen bonds break in the bulk liquid, nanodroplet with or without an upward electric field. The spread-
and hence, the number of water molecules that migrate from the ing rate in the early spreading stage are almost the same at various

(a)

0 ps 38 ps 72 ps 360 ps 478 ps 558 ps 2000 ps

(b)

0 ps 30 ps 44 ps 72 ps 338 ps 1028 ps 2000 ps

(c)

0 ps 34 ps 154 ps 352 ps 542 ps 1358 ps 2000 ps

(d)

0 ps 30 ps 70 ps 318 ps 966 ps 1448 ps 2000 ps


Fig. 9. Morphology evolution of a water nanodroplet on a nanostructured surface in the presence of an upward electric field at: (a) E = 0.03 V Å1; (b) E = 0.05 V Å1; (c)
E = 0.07 V Å1; and (d) E = 0.09 V Å1.

7
Ben-Xi Zhang, Shuo-Lin Wang, X. He et al. Journal of Molecular Liquids 333 (2021) 116039

100 of h0 = 50°, 42°, and 32° in the presence of a downward electric


90 field with E = 0.05 V Å1.
80
70 Fig. 11 shows the snapshots of nanodroplet spreading for the
60 three intrinsic contact angles. As the surface becomes more wet-
50 table, more nanogrooves are infiltrated by the nanodroplet. Only
40 five nanogrooves are infiltrated at h0 = 60°, whereas the number
R (Å)

increases to 7, 8, and 9 at h0 = 50°, 42°, and 32°, respectively. As


30 mentioned before, the reorientation of polar water molecules trig-
E= 0 V Å-1 gered by electric fields may make hydrogen bonds break. However,
20 E= -0.03 V Å-1 the enhanced intrinsic wettability means stronger interactions
E= -0.05 V Å-1 between gold atoms and water molecules, which hinders the reori-
E= -0.07 V Å-1 entation of water molecules in the vicinity of the solid-liquid inter-
E= -0.09 V Å-1 face, thereby preventing hydrogen bonds from breaking. As shown
10 in Fig. 12, the average number of hydrogen bonds in the vicinity of
0.1 1 4 the solid-liquid interface increases when the surface becomes
t (ns) more wettable, leading to a reduced solid-liquid interfacial tension
and hence an enhanced capillary force, which promoting spreading
Fig. 10. Evolution of spreading radius of a water nanodroplet on a nanostructured
surface in the presence of an upward electric field.

3.23

field strength and is equal to that under the downward electric 3.22
field, indicating that the spreading kinetics is not affected by the
field direction. In the later spreading stage, different spreading 3.21
kinetics is observed between both fields. The spreading rates under
3.20
<nHB>
the upward electric field are always lower than those under the
downward electric field. This result again confirms that spreading
3.19
of the nanodroplet is suppressed by the upward electric field and
the suppression becomes more significant at high field strengths. 3.18
E= 0 V Å
-1

3.17
E= -0.05 V Å
-1

3.4. Effect of intrinsic wettability


3.16
It is worth noting that the intrinsic wettability may influence 30 32 34 36 38 40 42 44 46 48 50 52
the formation of hydrogen bonds as well as the pinning, electro-
wettability, and electro-stretching effects, especially for a nan-
odroplet spreading on a nanostructured surface. In this section, Fig. 12. Average number of hydrogen bonds per water molecule in the vicinity of
simulations are implemented on the nanostructured surface with the solid-liquid interface on a nanostructured surface with various intrinsic
the same aspect ratio of h/w = 1 but different intrinsic wettability wettability in the presence of a downward electric field at E = 0.05 V Å1.

(a)

0 ps 22 ps 166 ps 542 ps 770 ps 1106 ps 2000 ps

(b)

0 ps 24 ps 116 ps 298 ps 938 ps 1308 ps 2000 ps

(c)

0 ps 28 ps 46 ps 316 ps 894 ps 1356 ps 2000 ps


Fig. 11. Morphology evolution of a water nanodroplet on a nanostructured surface with various intrinsic wettability in the presence of a downward electric field at
E = 0.05 V Å1: (a) h0 = 50°; (b) h0 = 42°; and (c) h0 = 32°.

8
Ben-Xi Zhang, Shuo-Lin Wang, X. He et al. Journal of Molecular Liquids 333 (2021) 116039

θ0=32 when applying an upward electric field, the spreading is


100 θ0=42 always hindered, and the static contact angle monotonously
θ0=50 increases with the field strength. The increased solid-liquid
interfacial tension and the enhanced pinning effect are
found to be responsible for the increased static contact angle
R (Å)

for both electric fields at low field strengths. The difference


at high field strengths between both electric fields is par-
tially attributed to different movement trajectory of water
10 molecules in the bulk liquid. As a result, the average number
of hydrogen bonds increases at high field strengths under
the downward electric field but decreases under the upward
electric field. Moreover, the pinning effect is always
0.01 0.1 1 4 enhanced under the upward field, whereas it is weakened
t (ns) at high field strengths under the downward field, which is
another reason for different statics and dynamics of wetting
Fig. 13. Evolution of spreading radius of a water nanodroplet on a nanostructured at high field strengths between upward and downward elec-
surface with various intrinsic wettability in the presence of a downward electric tric fields.
field at E = 0.05 V Å1. (3) A small intrinsic contact angle means strong interactions
between the nanodroplet and surface, which suppresses
the breaking of hydrogen bonds in the presence of an electric
of the nanodroplet on the nanostructured surface. On the basis of field, thereby reducing the solid-liquid interfacial tension
an analysis of the local number density, it is found that the and promoting the spreading. The small intrinsic contact
enhanced intrinsic wettability also decreases the pining force in angle also reduces the pinning force at the triple line as well
the triple line region. Moreover, our previous study has demon- as the energy barrier of the Cassie-Wenzel wetting transi-
strated that the enhanced intrinsic wettability reduces the energy tion, and hence, more nanogrooves are infiltrated by the
barrier for the wetting transition from the Cassie state to Wenzel nanodroplet, leading a reduced static contact angle.
state [41]. Therefore, the reduced pinning and energy barrier for
the wetting transition are responsible for the fact that more nano-
grooves are infiltrated by the nanodroplet. Fig. 13 shows the evolu- Declaration of Competing Interest
tion of spreading radius for the three intrinsic contact angles. In the
late spreading stage, the enhanced intrinsic wettability promotes The authors declare that they have no known competing finan-
the nanodroplet spreading. Likewise, this result can be explained cial interests or personal relationships that could have appeared
by the reduced static contact angle. to influence the work reported in this paper.

4. Conclusions
Acknowledgments

In this study, the spreading of a water nanodroplet over a


This study was partially supported by the State Key Program of
nanostructured surface in the presence of a vertical electric field
National Natural Science of China (No. 51936004), Science Fund for
are investigated via MD simulations. The field strength, field direc-
Creative Research Groups of the National Natural Science Founda-
tion, and intrinsic wettability of the surface are altered to under-
tion of China (No. 51821004), and the Fundamental Research
stand the statics and dynamics of wetting of the nanodroplet.
Funds for the Central Universities (No. 2020MS063).
The main conclusions are as follows.

(1) In the presence of an electric field, electric field forces make References
polar water molecules reorient themselves, forming a direc-
[1] G. Lu, X.D. Wang, Y.Y. Duan, A critical review of dynamic wetting by complex
tional arrangement along the field direction. The directional fluids: from Newtonian fluids to non-Newtonian fluids and nanofluids, Adv.
arrangement induces an electro-stretching effect and hence Colloid Interface Sci. 236 (2016) 43–62.
hinders the spreading. The directional arrangement also [2] C. Huh, L.E. Scriven, Hydrodynamic model of steady movement of a solid/
liquid/fluid triple line, J. Colloid Interface Sci. 35 (1) (1971) 85–101.
leads to the breaking of hydrogen bonds both in the bulk liq- [3] T.D. Blake, J.M. Haynes, Kinetics of liquid/liquid displacement, J. Colloid
uid and in the vicinity of the solid-liquid interface. Owing to Interface Sci. 30 (3) (1969) 421–423.
the breaking of hydrogen bonds in the bulk liquid, water [4] P.G.D. Gennes, Wetting: statics and dynamics, Rev. Mod. Phys. 57 (3) (1985)
827–863.
molecules are more likely to migrate to the solid-liquid or [5] P.G.D. Gennes, Deposition of langmuir-blodgett layers, Colloid Polym. Sci. 264
liquid-gas interface, accelerating the spreading, referred to (5) (1986) 463–465.
as an electro-wettability effect. The breaking of hydrogen [6] F.B. Wyart, G. Debregeas, P.G.D. Gennes, Spreading of viscous droplets on a non
viscous liquid, Colloid Polym. Sci. 274 (1) (1996) 70–72.
bonds in the vicinity of the solid-liquid interface increases [7] T.D. Blake, Dynamic contact angles and wetting kinetics, Wettability 251
the solid-liquid interfacial tension, thereby decreases capil- (1993).
lary force and hindering the spreading. Moreover, the elec- [8] D. Bonn, J. Eggers, J. Indekeu, J. Meunier, E. Rolley, Wetting and spreading, Rev.
Mod. Phys. 81 (2) (2009) 739–805.
tric field also alters the pinning force in the vicinity of the [9] T.D. Blake, J.D. Coninck, The influence of solid–liquid interactions on dynamic
triple line. Thus, the electro-stretching, electro-wettability, wetting, Adv. Colloid Interface Sci. 96 (1–3) (2002) 21–36.
modified solid-liquid interfacial tension, and pinning jointly [10] E. Celia, T. Darmanin, E.T.D. Givenchy, S. Amigoni, F. Guittard, Recent advances
in designing superhydrophobic surfaces, J. Colloid Interface Sci. 402 (2013) 1–
affect the statics and dynamics of wetting of nanodroplets
18.
on nanostructured surfaces. [11] A.B.D. Cassie, Contact angles, Discuss. Faraday Soc. 3 (1948) 11–16.
(2) When a downward electric field is applied to the nan- [12] R.N. Wenzel, Resistance of solid surfaces to wetting by water, Ind. Eng. Chem.
odroplet, the spreading is first hindered and then is pro- 28 (8) (1936) 988–994.
[13] T.H. Yen, Wetting characteristics of nanoscale water droplet on silicon
moted with increasing the field strength, leading to a first substrates with effects of surface morphology, Mol. Simul. 37 (9) (2011)
increased and then decreased static contact angle. However, 766–778.

9
Ben-Xi Zhang, Shuo-Lin Wang, X. He et al. Journal of Molecular Liquids 333 (2021) 116039

[14] D.Y. Zong, Z. Yang, Y.Y. Duan, Wetting kinetics of nanodroplets on lyophilic [31] S.M. Foiles, M.I. Baskes, M.S. Daw, Embedded-atom-method functions for the
nanopillar-arrayed surfaces: a molecular dynamics study, Chem. Phys. Lett. fcc metals Cu Ag, Au, Ni, Pd, Pt, and their alloys, Phys. Rev. B 33 (12) (1986)
685 (2017) 27–33. 7983–7991.
[15] Q.Z. Yuan, Y.P. Zhao, Wetting on flexible hydrophilic pillar-arrays, Sci. Rep. 3 [32] J.V.L. Beckers, C.P. Lowe, S.W.D. Leeuw, An iterative PPPM method for
(1) (2013) 1–6. simulating Coulombic systems on distributed memory parallel computers,
[16] Q.Z. Yuan, Y.P. Zhao, Multiscale dynamic wetting of a droplet on a lyophilic Mol. Simul. 20 (6) (1998) 369–383.
pillar-arrayed surface, J. Fluid Mech. 716 (2013) 171–188. [33] W.C. Swope, H.C. Andersen, P.H. Berens, K.R. Wilson, A computer simulation
[17] G. Lippmann, Relations entre les phénomènes électriques et capillaires, method for the calculation of equilibrium constants for the formation of
Gauthier-Villars 35 (1875) 494–548. physical clusters of molecules: application to small water clusters, J. Chem.
[18] F.H. Song, L. Ma, J. Fan, Q.C. Chen, L. Zhang, B.Q. Li, Wetting behaviors of a Phys. 76 (1) (1982) 637–649.
nano-droplet on a rough solid substrate under perpendicular electric field, [34] S. Sun, J.T.Y. Wong, T.Y. Zhang, Molecular dynamics simulations of phase
Nanomaterials 8 (5) (2018) 340. transition of lamellar lipid membrane in water under an electric field, Soft
[19] F.H. Song, B.Q. Li, C. Liu, Molecular dynamics simulation of the electrically Matter 7 (1) (2011) 147–152.
induced spreading of an ionically conducting water droplet, Langmuir 30 (9) [35] S.N. Reznik, A.L. Yarin, A. Theron, E. Zussman, Transient and steady shapes of
(2014) 2394–2400. droplets attached to a surface in a strong electric field, J. Fluid Mech. 516
[20] G.H. Hu, A.J. Xu, Z. Xu, Z.W. Zhou, Dewetting of nanometer thin films under an (2004) 349–376.
electric field, Phys. Fluids 20 (10) (2008) 102101. [36] T. Steinel, J.B. Asbury, J. Zheng, M.D. Fayer, Watching hydrogen bonds break: a
[21] Y.P. Zhao, Y. Wang, Fundamentals and applications of electrowetting, Reviews transient absorption study of water, J. Phys. Chem. A 108 (50) (2004) 10957–
of Adhesion and Adhesives 1 (1) (2013) 114–174. 10964.
[22] C.D. Daub, D. Bratko, K. Leung, A. Luzar, Electrowetting at the nanoscale, J. [37] J.M. Oh, S.H. Ko, K.H. Kang, Analysis of electrowetting-driven spreading of a
Phys. Chem. C 111 (2) (2007) 505–509. drop in air, Phys. Fluids 22 (3) (2010) 032002.
[23] D.Y. Zong, Z. Yang, Y.Y. Duan, Wettability of a nano-droplet in an electric field: [38] D. Chakraborty, S. Pathak, M. Chakraborty, Molecular investigation of triple
A molecular dynamics study, Appl. Therm. Eng. 122 (2017) 71–79. line movement in electrowetted nanodroplets, Langmuir 36 (42) (2020)
[24] F.H. Song, B.Q. Li, C. Liu, Molecular dynamics simulation of nanosized water 12580–12589.
droplet spreading in an electric field, Langmuir 29 (13) (2013) 4266–4274. [39] H. Ren, R.B. Fair, M.G. Pollack, E.J. Shaughnessy, Dynamics of electro-wetting
[25] Y.P. Zhao, Q.Z. Yuan, Statics and dynamics of electrowetting on pillar-arrayed droplet transport, Sens. Actuators, B 87 (1) (2002) 201–206.
surfaces at the nanoscale, Nanoscale 7 (6) (2015) 2561–2567. [40] R. Baviere, J. Boutet, Y. Fouillet, Dynamics of droplet transport induced by
[26] T.H. Yen, Investigation of the effects of perpendicular electric field and surface electrowetting actuation, Microfluid. Nanofluid. 4 (4) (2008) 287–294.
morphology on nanoscale droplet using molecular dynamics simulation, Mol. [41] B.X. Zhang, S.L. Wang, X.D. Wang, Wetting transition from the Cassie-Baxter
Simul. 38 (6) (2012) 509–517. state to the Wenzel state on regularly nanostructured surfaces induced by an
[27] J.P. Ryckaert, G. Ciccotti, H.J.C. Berendsen, Numerical integration of the electric field, Langmuir 35 (3) (2019) 662–670.
cartesian equations of motion of a system with constraints: molecular [42] A. Bateni, S. Laughton, H. Tavana, S.S. Susnar, A. Amirfazli, A.W. Neumann,
dynamics of n-alkanes, J. Comput. Phys. 23 (3) (1977) 327–341. Effect of electric fields on contact angle and surface tension of drops, J. Colloid
[28] H.J.C. Berendsen, J.R. Grigera, T.P. Straatsma, The missing term in effective pair Interface Sci. 283 (1) (2005) 215–222.
potentials, J. Phys. Chem. 91 (24) (1987) 6269–6271. [43] D.Y. Zong, H. Hu, Y.Y. Duan, Viscosity of water under electric field: Anisotropy
[29] L. Chen, C. Li, N.F.A. Vegt, G.K. Auernhammer, E. Bonaccurso, Initial induced by redistribution of hydrogen bonds, J. Phys. Chem. B 120 (21) (2016)
electrospreading of aqueous electrolyte drops, Phys. Rev. Lett. 110 (2) (2013) 4818–4827.
026103. [44] L. Chen, E. Bonaccurso, Electrowetting-From statics to dynamics, Adv. Colloid
[30] D. Vanzo, D. Bratko, A. Luzar, Nanoconfined water under electric field at Interface Sci. 210 (2014) 2–12.
constant chemical potential undergoes electrostriction, J. Chem. Phys. 140 (7)
(2014) 02B616–11.

10

You might also like