Bansal 2015

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

PHYSICAL REVIEW E 92, 042304 (2015)

Universal buckling kinetics in drying nanoparticle-laden droplets on a hydrophobic substrate

Lalit Bansal,* Ankur Miglani,* and Saptarshi Basu*,†


Department of Mechanical Engineering, Indian Institute of Science, Bangalore, India-560012
(Received 3 June 2015; revised manuscript received 18 September 2015; published 6 October 2015)
We provide a comprehensive physical description of the vaporization, self-assembly, agglomeration, and
buckling kinetics of sessile nanofluid droplets pinned on a hydrophobic substrate. We have deciphered five distinct
regimes of the droplet life cycle. Regimes I–III consists of evaporation-induced preferential agglomeration that
leads to the formation of a unique dome-shaped inhomogeneous shell with a stratified varying-density liquid
core. Regime IV involves capillary-pressure-initiated shell buckling and stress-induced shell rupture. Regime V
marks rupture-induced cavity inception and growth. We demonstrate through scaling arguments that the growth
of the cavity (which controls the final morphology or structure) can be described by a universal function.

DOI: 10.1103/PhysRevE.92.042304 PACS number(s): 64.70.fm, 47.55.D−, 47.52.+j, 82.70.Dd

I. INTRODUCTION the buckling front leading to enhanced flow shear and packing
of nanoparticles resulting in a compact final microstructure
Vaporization [1,2] and self-assembly [3] of nanosuspen-
[Figs. 5(a) and 5(c)]. In particular, the buckling transition in
sions are ubiquitous phenomena that are pivotal to a plethora
regimes IV and V occurs in the last 30% of the droplet lifetime
of natural and industrial processes like food processing [4], fire
(i.e., tI V + tV ∼ 0.1te − 0.3te : total droplet lifetime) (Fig. 1).
suppression [5], formation of microcapsules [6], drug delivery
In the past decade, shell formation during drying of
[7], production of functional coatings [8,9] and combustion of
colloidal droplets and its subsequent buckling have been
nanofuels [10]. The life cycle of a drying nanofluid droplet
investigated for different configurations ranging from sessile
[11–13] in sessile or levitated mode exhibits a series of tightly
[14,26,27] to pendant [28] and thin films sandwiched between
coupled physicochemical processes occurring at multiple
glass slides [29] to levitation [17,19]. However, the latter
spatiotemporal scales. For instance, recent studies [14–18] on
stages of shell rupture and volumetric cavity growth (buckling
drying colloidal droplets have shown that evaporation-induced
front kinetics) have not been observed previously in sessile
internal flow (∼mm s−1 ) leads to formation of a porous crust
droplets. For example, Evans and co-workers [30] qualitatively
of randomly packed [pore size ∼O(10–100 nm)] nanoparticles
described a particle agglomeration process in a sessile droplet
that mimics a thin elastic-shell ∼O(10 μm). This elastic shell
that leads to a similar hollow precipitate but the dynamics of
eventually buckles at a critical capillary pressure causing
cavity growth were not captured. We elucidate that buckling-
severe deformations, thereby resulting in microstructures with
induced cavity growth inside a drying sessile nanofluid droplet
distinct morphologies (bowl, rings, and doughnuts) [15,19].
is controlled by the solvent evaporation rate through the pores
However, there is a dearth of experimental data and theoretical
and can be described by a universal function independent of
analyses on the kinetics of buckling front except in a recent
the initial concentration of nanoparticles.
study by Miglani and Basu [17]. Such understanding for sessile
droplets is crucial for multiple interdisciplinary application
domains like deposition of photonic crystals [20], formation II. MATERIALS AND METHODS
of DNA microarrays [21], surface patterning [22], and cell
The buckling front kinetics is investigated for an aque-
biology [23,24].
ous Ludox TM-40 colloidal silica suspension (from Sigma
In this work, we analyze the motion of a buckling front
Aldrich, average particle diameter d = 2a = 25 ± 2 nm, dis-
in a drying nanosilica droplet dispensed on a hydrophobic
- ∼ 0.15). Experiments are conducted for a 3 μl test
persity D
substrate. The droplet life cycle can be divided into five
droplet dispensed on a porous hydrophobic substrate (static
regimes as shown in Figs. 1 and 2 for an initial nanoparticle
contact angle θc ∼ 125◦ , initial droplet height hi ≈ 1.45 mm
loading of 40 wt % (the corresponding transition for an initial
and diameter Di ≈ 1.7 mm) at an ambient temperature of
20 wt % particle loading is shown in Fig. 3): (I) Volumetric
300 K and relative humidity (RH) of approximately (45 ± 5)%
shrinkage in constant contact radius (CCR) mode. (II) Particle
with the initial particle loading rate in the range ϕo =
transport leading to a dome-shaped structure with nonuniform
20−40 wt % (in increments of 5 wt %). The evaporation and
agglomeration along the periphery [Movie 1 [25] and Figs. 1(a)
subsequent buckling are captured using a MotionScope NR3S1
and 1(b)]. (III) Formation of an inhomogeneous porous
camera coupled with a 1× Navitar lens (spatial resolution
shell with minimum thickness at the top [Fig. 1(b)]. (IV)
∼22 μm/pixel and temporal resolution 100 ms).
Capillary–pressure-induced shell buckling at the pole forming
a primary cavity (PC) and subsequent rupturing resulting in
a daughter cavity (DC) [Movies 2 and 3 [25] and Figs. 1(c), III. RESULTS
1(d), 2(c), and 4]. (V) Advection and topological dynamics of Abkarian et al. [31] reported destabilization of the colloidal
particle layer on an oil-water interface due to the gravi-
tational force.
√ However, in the current study the capillary
*
All authors contributed equally to this work. length l = σ/pg ≈ 2.7 mm is greater than the droplet

Corresponding author: sbasu@mecheng.iisc.ernet.in radius (∼l/2). Thus gravity does not induce any droplet

1539-3755/2015/92(4)/042304(6) 042304-1 ©2015 American Physical Society


LALIT BANSAL, ANKUR MIGLANI, AND SAPTARSHI BASU PHYSICAL REVIEW E 92, 042304 (2015)

FIG. 1. (Color online) Droplet life cycle. (a) Recirculation flow pattern inside the droplet (see Video 1 [25]) using laser sheet of 175 microns
in thickness. (b) Shell inhomogeneity at the onset of buckling; TZ is the top zone, MZ is the middle zone, and BZ is the bottom zone. (c)
Buckling onset (formation of parent cavity PC) and shell rupture due to stretching (formation of daughter cavity DC). (d) Evaporation through
the porous shell, air invasion though the punctured PC, and DC growth in the r-yc direction.

deformation. The mass diffusion Péclet number for the droplet, near the three-phase contact line, thereby suppressing the
Pem = tdiff /tevap = (Ri2 /Da )/te is in the range of 15 < Pem < local evaporative flux [Fig. 1(a)]. In particular, the distribution
30 (Da ∼ 5.6 × 10−11 m2 s−1 calculated using the Stokes- of solvent vaporization flux across the droplet surface is
Einstein relation). With Pem  1, i.e., in the fast-drying given by j (t) = −Dv [(∂c/∂r) cos β + (∂c/∂z) sin β] [32],
regime, the rate of droplet surface regression is an order which is time variant since the droplet contact angle changes
higher compared to the particle diffusion rate. This leads to [where Dv is the vapor diffusion coefficient, c is the vapor
nanoparticle (NP) accumulation at the receding liquid-vapor concentration, and β is the angle as shown in Fig. 1(a)]. Below
interface thereby forming a viscoelastic shell [Fig. 1(b)]. horizontal planes with β = 0 there is a region of high vapor
concentration with a marked reduction in j compared to that
A. Inhomogeneous shell formaon at the top. At the initial time instant t = t0 (θc ∼ 125◦ ), β = 0
Evans et al. [30] proposed that for a sessile droplet on a corresponds to the plane hs ∼ 0.35hi where the reduction in
hydrophobic substrate, a region of high humidity builds up j is ∼70%, while below this plane j/jtop < 0.3. However,

FIG. 2. (Color online) Dynamic morphological transformation of a drying n-SiO2 dispersion droplet (40 wt %) from buckling onset until
the formation of the final precipitate. Top row depicts the front view while the bottom row shows the corresponding top view. (a) Buckling
onset, (b) stretching of the parent cavity, (c) shell puncturing and formation of daughter cavity, (d)–(g) radial-longitudinal cavity expansion,
and (h) final microstructure. Images are at time instants 0, 20, 64, 76, 96, 128, 184, and 236 s. Scale bar equals 1.2 mm.

042304-2
UNIVERSAL BUCKLING KINETICS IN DRYING . . . PHYSICAL REVIEW E 92, 042304 (2015)

FIG. 3. (Color online) Dynamic deformation of the droplet from buckling onset until the formation of the final precipitate structure for
initial particle loading of 20 wt %. (a) Droplet at the onset of buckling, (b) buckling onset marked by the stretch marks on the top surface,
(c) stretching of the shell, (d) shell punctures forming a cavity, (e)–(h) radial-longitudinal cavity spreading, and (i) final precipitate structure.
The images are at time instants 0, 25, 33, 53, 61, 67, 92, 117, and 135 s. Scale bar equals 1.1 mm.

at buckling onset (θc ∼ 90◦ ), this plane shifts downwards Due to this internal recirculation resulting from nonuniform
to hs ∼ 0.25hi at 40 wt % and hs ∼ 0.1hi at 20 wt % evaporative flux, there is preferential particle deposition near
with j/jtop ∼ 0.75 and 0.65, respectively. To satisfy mass the three-phase contact line. This results in an inhomogeneous
conservation, this differential evaporation induces solvent flow dome-shaped structure with an encapsulated liquid domain
towards the apex, thereby generating a poloidal recirculation that can be divided into a three-layered density-stratified
pattern (flow velocity ∼0.015 mm s−1 ) [Fig. 1(a) and Movie 1 medium [Fig. 1(b) with distinct changes in refractive index].
[25]). This is in contrast to the capillary flow (towards pinned The dimensionless dome height hB /hi at buckling onset
edges) observed by Deegan et al. [33] in droplets drying is ∼1.6 times higher at 40 wt % compared to 20 wt % while
on hydrophilic substrates (the coffee ring phenomenon). the dimensionless buckling radius RB /Ri is ∼1.35 times
larger. This is shown as the inset in Fig. 6. The time delay to
buckling onset (tB = tI + tII + tIII ) is however 2.2 times shorter
at 40 wt %. This is because higher concentrations favor fast
particle aggregation (shorter agglomeration time scale) thereby
leading to an early and hence a larger dome structure formation
(note that for the shell to buckle, the formation of a dome
structure is a prerequisite step). Specifically, the orthokinetic
gelation time scale given by tg = 4γ̇ (3−D) πD
(ϕo−1 − 1) [34] is
2.25 times higher at 20 wt % compared to 40 wt % (where
D ∼ 1.5 is the fractal dimensionality and γ̇ is the velocity
gradient) which agrees directly with tB . The dome formation
is not sustained for ϕ0 < 8 wt % (due to the lower available
mass of the particles). However, in the absence of substrate
(contact-free configuration) buckling and cavity growth have
been observed even at very dilute concentrations (0.01 vol % in
Leidenfrost droplets [19] and 2 wt % in acoustically levitated
droplets [17]).
The resultant shell is inhomogeneous with maximum
FIG. 4. (Color online) Dynamic three-stage growth process of thickness τ at the base and gradually narrowing towards the
the daughter cavity. top with minimum τ at the apex [Fig. 1(b)]. This nonuniform

042304-3
LALIT BANSAL, ANKUR MIGLANI, AND SAPTARSHI BASU PHYSICAL REVIEW E 92, 042304 (2015)

FIG. 5. (Color online) SEM micrographs of the final precipitate: (a) Overall view for ϕo = 20 wt %, (c) ϕo = 40 wt % and cut sections
showing the inhomogeneous shell structure (b) ϕo = 20 wt %, (d) ϕo = 40 wt %. Scale bar equals 200 μm. Magnified view showing closely
packed NPs at (e) the equatorial plane and (f) the base. Scale bar equals 100 nm.

shell formation is evident from scanning electron microscope due to the flow divergence. This weak, thin-walled elastic
(SEM) micrographs of the cut-sections of the final precipitate shell with Foppl–von Karman number α ≈ 10 ( RτBB )2 ∼ O(104 )
(Fig. 5). Spatially varying shell thickness can be defined by eventually undergoes inversion of curvature, marking the
an inhomogeneity parameter 1 − τ/τbase . Clearly, γ → 0 at buckling onset [Fig. 1(c)].
the base and γ → 1 at the top. At 20 wt %, γ varies from
∼0.57 in the bending strip region surrounding the buckled
top sector to ∼0.45 at the equatorial plane, while at 40 wt %, B. Two-stage invagination process
the corresponding values are ∼0.35 and ∼0.27, respectively. Initiation of buckling from the top is also corroborated
Thus, the shell top is ∼15% thinner compared to the equatorial by the fact that the critical buckling pressure Pcr ≈ 10Y /α
belt and ∼35%−60% thinner compared to the base. As such, (linear thin shell buckling theory [35]) is least at this location.
the top sector becomes a weak soft spot and acts as a preferred Pcr represents the threshold external pressure beyond which
buckling site. Clearly, the minimum thickness at the apex is the inverted cap [Fig. 1(c)] is in unstable equilibrium and
small deformations grow spontaneously when the inversion
depth y ∼ τB [35]. Based on this condition and given that
r  RB , τB is estimated to be O(12 μm). As shown in SEM
images at the NP scale [Figs. 5(e) and 5(f)], particle close
packing is the same across the shell. Thus, assuming a constant
shell’s Young’s modulus Y , Pcr varies as τB2 and hence Pcr
required for buckling onset is minimum at the top. This
critical–buckling load is provided by the capillary pressure
drop PB = μJ τ/k ≈ 7 kPa in accordance with Darcy’s law,
where k is the permeability obtained from the Carman-Kozeny
relation, μ is the dynamic viscosity of the solvent (variation in
the solvent viscosity is ∼2%). Thus, with Pcr ∼ PB , Young’s
modulus Y is estimated to be O(107 Pa).
Following buckling onset, as the inverted cap (parent
cavity) grows, the energy localized in the bending strip
becomes enormous, and as such another buckling pathway
becomes energetically favored [29]. This features an increase
in indentation depth (y) with constant lateral width (2rtip )
thereby leading to in-plane stretching (tensile stress) of the
FIG. 6. (Color online) Temporal variation of nondimensional PC. This stretching causes significant weakening of the shell,
daughter cavity volume. Inset shows the droplet height and radius leading to PC rupture [Fig. 1(c)]. For the shell to rupture, the
at buckling onset as a function of NP concentration (wt%). tensile stress in the stretched part of the shell due to the relative

042304-4
UNIVERSAL BUCKLING KINETICS IN DRYING . . . PHYSICAL REVIEW E 92, 042304 (2015)

length change constant and the solvent vaporization proceeds such that
dVPC /dT + dVDC /dt = −dV /dt (i.e., VPC + VDC + V =
l f − l0 larc(ab c) − larc(abc) const = VB , the droplet volume at buckling onset). Note that
δ= =
l0 larc(abc) due to the nonuniform growth of the DC in the third stage (BZ)
(Figs. 2 and 4), subsequent analysis is restricted only to the
[Fig. 1(c)] must overcome the cohesive stress holding the NPs top and middle zones where the DC dilates uniformly.
together, i.e., Y δ ∼ σ . Here, the critical stress (σ ) required In regime V, the net evaporative flux across the droplet
to separate the particles placed at an equilibrium separation surface sustains the volumetric growth of the daughter
of Z ≈ 0.2 nm can be calculated based on the van der Waals cavity:
−20
interaction as σ ∼ ( 12Z Aa
2 )/a [29], where A ∼ 0.83 × 10
2
J
is the Hamaker constant and a is the NP radius ∼11 nm. dVDC
Based on these calculations, δ is estimated to be ∼0.13, ≈ J Sε, (1)
dt
which is close to the value determined from experiments,
δ ∼ 0.16. Puncturing of the parent cavity creates a hole in where S = π (4RB2 (ARB ) − rc2 ) is the droplet surface area
the shell (∼0.1 mm), thereby causing a five-order decrease (variable but constant for a given concentration) and ε is
in the capillary pressure due to an increase in the radius the shell porosity (which depends on particle packing). The
from ∼O(10−9 m) (pore radius) to ∼10−4 m (hole radius) solvent evaporation rate per unit surface area at the air-water
(capillary pressure is proportional to 1/r). According to the interface is diffusion limited and is given by J ∼ Dv ∇Pv ≈
invasion percolation theory [36,37] for transport in a saturated 1
nL kB T
Dv Pv
L
≈ JL [38], where Dv (∼3 × 10−5 m2 s−1 ), ∇Pv
porous medium, air invades through the largest pore on and nL are respectively the vapor diffusion coefficient, the
the surface where capillary pressure is minimum. Thus, air vapor pressure gradient, and the number density of water
ingression through this punctured site leads to the formation molecules. J = DnLv kBPTv is a constant (independent of the
of a daughter cavity [Fig. 1(c)]. Subsequently, the daughter initial concentration) that depends only on the thermophysical
cavity undergoes a sequential three-stage radial-longitudinal properties of the solvent and the operating conditions. L is
expansion (r−yc ) with a constant volumetric growth rate the diffusion length scale. It is known that the evaporative
dVDC /dt ∼ O(10−3 mm3 s−1 ), marking regime V. Daughter flux is spatially nonuniform; however, the net evaporative flux
cavity growth proceeds in three stages with continuous averaged over the droplet surface is constant from a scaling
increase in the cavity depth (yc ) and radius along the plane of or order of magnitude point of view. Temporal variation of
maximum lateral width rc, max . For calculating the volumetric evaporative flux (for natural drying) is largely negligible during
cavity growth rate the cavity is assumed to be a surface of the cavity growth process.
revolution of different geometries depending on its aspect In regime V, given the initial condition that VDC = 0 at
ratio k = yc /rc, max , i.e., (1) V = π6 krc3 (3 + k 2 ) for a spherical puncturing onset (t = 0), Eq. (1) reduces to
cap (k < 1), (2) V = 2π r 3 for a hemisphere (k = 1), and (3)
3 c J
V = 3 k rc, max (3 − k) for a spherical dome (k > 1). This is
π 2 3
VDC = εS t. (2)
shown in Fig. 4. The sequential three-stage process shows the L
following features: (1) Rapid, in-tandem radial-longitudinal Based on SEM micrographs of the final precipitate [shown
expansion with ṙc ≈ ẏc ∼ O(0.015 mm s−1 ) which occurs as in Figs. 5(e) and 5(f)], the average void fraction of NPs
the DC enters the top zone (TZ), which is primarily in the is estimated to be ε ∼ 0.4 [averaged spatially over multiple
solution phase, and the cavity is assumed to be a spherical locations as marked in Fig. 5(d)]. As such, during DC growth,
cap with rycc < 1.(2) As the cavity enters a much more viscous assuming that random close packing (1 − ε) ranges from
middle zone (MZ) cavity growth is characterized by a constant ∼0.4 to 0.6, the diffusion length scale L is estimated to be
rim radius rc , but with a marginal difference in spreading of the same order as the shell thickness, ∼O(0.2 mm) at
rates along the longitudinal direction and laterally along the all concentrations. Since at all concentrations rc  RB ∼ hB
plane of maximum lateral width (2rc,max ), i.e. dy dt
c
∼ 0.004 and assuming ε ≈ 0.4 and L ≈ 0.2 mm, Eq. (1) suggests
drc, max −1
and dt ∼ 0.0025 mm s . Thus, spreading along these that the volumetric growth of the daughter cavity exhibits
directions is an order of magnitude slower compared to the a linear time dependence and is primarily a function of the
first stage; however, the volumetric growth rate dVDC /dt ∼ droplet dimensions at buckling onset (i.e., RB and hB ), which
O(10−3 mm3 s−1 ) remains constant in both stages. In addition, in-turn depend on the initial concentration (inset in Fig. 6).
as the cavity growth progresses, nanoparticles are pushed to Thus, normalizing the cavity volume by a factor εS(J /L)tm
the periphery as the flow diverges at the progressing front face and time by the cavity growth time scale tm shows that the
of the cavity (along the droplet centerline). This results in high transient profiles of the daughter cavity volume across all
shear and gradual consolidation of NPs along the periphery concentrations indeed collapse into a universal trend (Fig. 6).
(Movie 1 [25]), thereby, resulting in a close packing. (3) Here, tm is the time from puncturing onset until the end of
Finally, as the cavity penetrates the gelatinous bottom zone DC growth in the MZ [Figs. 1(b) and 4] and calculated as
(BZ) it grows randomly. the difference between the droplet vaporization time scale
(tevap ) and the time delay to puncturing onset (tP ). Therefore,
the growth of the daughter cavity can be predicted via the
C. Universal buckling features of cavity global length scales (RB and hB ) and time scales (tevap
During the buckling transition characterized by regimes and tP ), both being strong functions of the initial particle
IV and V, the droplet aspect ratio (ARB = hB /2RB ) remains loading.

042304-5
LALIT BANSAL, ANKUR MIGLANI, AND SAPTARSHI BASU PHYSICAL REVIEW E 92, 042304 (2015)

IV. CONCLUSION leading to local suppression of evaporation. This alters both the
location of the buckling site and the buckling pathway. Similar
A drying sessile nanofluid droplet on a hydrophobic
observations were also reported by Pauchard and Couder [14],
surface exhibits a series of thermophysical changes at multiple
where they suppressed evaporation locally by keeping a solid
spatiotemporal scales. We have described the complete droplet
plate near the droplet. Although these aforementioned studies
life cycle from deployment to the final precipitate formation,
have not captured the cavity growth process dynamically at
featuring a sequential process of preferential particle agglom-
high speed, their results are indicative that the location of
eration, shell formation, onset of buckling instability, and the
the buckling site, the time delay to buckling onset, and the
subsequent two-stage growth of the resulting cavity. Based
buckling pathway can be controlled by altering the droplet
on solvent mass balance and scaling analysis, we conclude
evaporation. Thus, our findings clearly indicate that the cavity
that the volumetric cavity growth rate during the second stage
growth is indeed an evaporation-driven phenomenon and the
exhibits a universal trend independent of the initial particle
global surface-averaged evaporation flux (primarily a function
loading. Chen and Evans [26] have also shown that if two
of surface temperature for given RH) is a good normalization
droplets are placed in close proximity to each other the air
parameter for predicting the universal trend in cavity growth
in their vicinity quickly gets saturated with vapor, thereby
for sessile droplets.

[1] R. H. Chen, T. X. Phuoc, and D. Martello, Int. J. Heat Mass [22] Q. Li, Y. T. Zhu, I. A. Kinloch, and A. H. Windle, J. Phys. Chem.
Transfer 53, 3677 (2010). B 110, 13926 (2006).
[2] J. R. Moffat, K. Sefiane, and M. E. R Shanahan, J. Nano Res. 7, [23] D. Vella, A. Ajdari, A. Vaziri, and A. Boudaoud, J. R. Soc.
75 (2009). Interface 9, 448 (2012).
[3] P. J. Yoo, K. Y. Suh, S. Y. Park, and H. H. Lee, Adv. Mater. 14, [24] D. A. Fletcher and R. D. Mullin, Nature (London) 463, 485
1383 (2002). (2010).
[4] S. Rogers, Y. Fang, S. X. Qi Lin, C. Selomulya, and X. Dong [25] See Supplemental Material at http://link.aps.org/supplemental/
Chen, Chem. Eng. Sci. 71, 75 (2012). 10.1103/PhysRevE.92.042304 for videos of the internal recircu-
[5] S. Chandra, M. Di Marzo, Y. M. Qiao, and P. Tartarini. Fire lation and top and side views of the daughter cavity expansion.
Safety J. 27, 141 (1996). [26] L. Chen and J. R. G. Evans, Langmuir 25, 11299 (2009).
[6] A. Fery and R. Weinkamer, Polymer 48, 7221 (2007). [27] T. A. Nguyen, M. A. Hampton, and A. V. Nguyen, J. Phys.
[7] P. Thybo, L. Hovgaard, J. S. Lindelov, A. Brask, and S. K. Chem. C 117, 4707 (2013).
Andersen, Pharm. Res. 25, 1610 (2008). [28] C. Sadek, H. Tabuteau, P. Schuck, Y. Fallourd, N. Pradeau,
[8] G. Bertrand, P. Roy, C. Filiatre, and C. Coddet, Chem. Eng. Sci. C. Le Floch-Fouéré, and R. Jeantet, Langmuir 29, 15606
60, 95 (2005). (2013).
[9] Y. O. Popov, Phys. Rev. E 71, 036313 (2005). [29] F. Boulogne, F. Giorgiutti-Dauphiné, and L. Pauchard, Soft
[10] A. Miglani, S. Basu, and R. Kumar, Phys. Fluids 26, 032101 Matter 9, 750 (2013).
(2014). [30] Y. Zhang, S. Yang, L. Chen, and J. R. G. Evans, Langmuir 24,
[11] M. A. Hampton, T. A. Nguyen, A. V. Nguyen, Z. P. Xu, L. Huang, 3752 (2008).
and V. Rudolph, J. Colloid Interface Sci. 377, 456 (2012). [31] M Abkarian, S. Protiere, J. M. Aristoff, and H. A. Stone, Nat.
[12] R. H. Chen, T. X. Phuoc, and D. Martello, Int. J. Heat Mass Commun. 4, 1895 (2013).
Transfer 54, 2459 (2011). [32] S. Semenov, V. M. Starov, R. G. Rubio, and M. G. Velarde,
[13] S. M. Murshed and C. A. N. de Castro, J. Nanosci. Nanotechnol. Colloids Surf., A 372, 127 (2010).
11, 3427 (2011). [33] R. D. Deegan, O. Bakajin, T. F. Dupont, G. Huber, S. R. Nagel,
[14] L. Pauchard and Y. Couder, Europhys. Lett. 66, 667 (2004). and T. A. Witten, Nature (London) 389, 827 (1997).
[15] J. Bahadur, D. Sen, S. Mazumder, S. Bhattacharya, H. Frieling- [34] L. G. Bremer, P. Walstra, and T. van Vliet, Colloids Surf., A 99,
haus, and G. Goerigk, Langmuir 27, 8404 (2011). 121 (1995).
[16] J. Paulose and D. R. Nelson, Soft Matter 9, 8227 (2013). [35] L. Landau and E. Lifshitz, Theory of Elasticity, 3rd ed.
[17] A. Miglani and S. Basu, Soft Matter 11, 2268 (2015). (Butterworth-Heinemann, London, 1986).
[18] B. Sobac and D. Brutin, Colloids Surf., A 448, 34 (2014). [36] D. Wilkinson and J. F. Willemsen, J. Phys. A 16, 3365
[19] N. Tsapis, E. R. Dufresne, S. S. Sinha, C. S. Riera, J. W. (1983).
Hutchinson, L. Mahadevan, and D. A. Weitz, Phys. Rev. Lett. [37] M. Prat, Int. J. Multiphase Flow 19, 691 (1993).
94, 018302 (2005). [38] E. R. Dufresne, E. I. Corwin, N. A. Greenblatt, J. Ashmore, D. Y.
[20] M. Egen and R. Zentel, Chem. Mater. 14, 2176 (2002). Wang, A. D. Dinsmore, J. X. Cheng, X. S. Xie, J. W. Hutchinson,
[21] R. Blossey and A. Bosio, Langmuir 18, 2952 (2002). and D. A. Weitz, Phys. Rev. Lett. 91, 224501 (2003).

042304-6

You might also like