spontaneous liquid marble formation on packed porous beds

7
Spontaneous liquid marble formation on packed porous bedsCatherine P. Whitby, * Xun Bian and Rossen Sedev Received 2nd July 2012, Accepted 6th September 2012 DOI: 10.1039/c2sm26529j The encapsulation of aqueous and organic solvents with particles used to form liquid marbles implies there are attractive interactions between the particles and those different liquids. This is often masked, however, by the impact of the droplet kinetic energy on marble formation. We investigated droplet wetting and evaporation when drops were gently placed (without rolling or shaking) on beds of silanised glass beads. Particle coating of the drop surface occurred within seconds of liquid contact with the particle bed. This ruled out evaporation from causing the particles to appear to rise up the surface of the drop as it was reduced in volume. Particles attach to the fresh liquid surface created during the droplet oscillations immediately after contact. The further ordered advance of the particles over the drop surface and the close-packed arrangement of the particles revealed the role of capillary forces in the coating process. By minimising the kinetic energy of the droplet contact with the particles, we found that maximum particle coating occurs at liquid surface tensions just above the critical wetting tension of the beads. Introduction The wetting behaviour of fine particles at air–liquid surfaces 1 is important in powder and soil technologies where liquid droplets are in contact with granular or powdery surfaces. These include dispersion, granulation, filtration and water repellency. Liquid cannot penetrate into powders that consist of poorly wetting particles. The particles may encapsulate drops that are rolled across the bed, forming liquid core–particle shell structures (liquid marbles) that have large contact angles on any substrate. 2–4 Several studies have exploited ‘‘liquid marbles’’ for transporting small volumes of high surface tension liquids, 5 storing water in a powdered form 6,7 and for the encapsulation and delivery of active (water soluble) ingredients. 8,9 Recently Gao and McCarthy, 10 Xue et al. 11 and Matsukuma et al. 12 reported encapsulating liquids with lower surface tensions (down to 20 mN m 1 ) using fluorinated polymeric powders. All the liquid marbles had large contact angles. On the other hand, McHale et al. 13 showed that simply depositing water drops onto beds of hydrophobic particles causes the underside of the drops to be coated with particles, but no liquid marble is formed. These observations suggest that the particle surface energy influences liquid marble formation. They are complicated, however, by the contribution of the drop kinetic energy to the particle wetting and attachment processes. In this paper, we study the attachment and movement of particles over the surfaces of stationary drops of ethanol–water solutions. The particle wettability (q) was varied systematically from hydrophobic (<90 ) to hydrophilic (>90 ) by controlling the ethanol concentration and hence liquid surface tension (g l ). A liquid droplet (of radius, R and viscosity, m) placed on the surface of a bed of particles (of radius, r) may absorb into the bed (Fig. 1a) or evaporate (Fig. 1b), depending on the particle wettability. The average rate of imbibition of liquid into a porous substrate is determined by the pore geometry and q. 14 For steady flow conditions and small pores, the Washburn equation relates the capillary driving force of a liquid penetrating through a compact bed of particles and the viscous drag as d 2 t ¼ r eff g l cos q 2m (1) where d is the depth of liquid penetrating into the bed in time t. r eff is the effective pore radius. For the special case of powder beds that can be treated as uniform arrays of cylindrical pores, 15,16 r eff is calculated using r eff ¼ 2(1 f p )/f p r p A p (2) where A p is the specific surface area of the particles. f p , the porosity parameter, is determined by the ratio m p /V T r p , where m p is the mass of particles in the bed, V T is the bed volume and r p is the particle density. Assuming that the three phase contact line remains stationary as liquid drains into the pores (the constant drawing, or constant radius limiting case 17 ), the penetration or wicking time, s, is given by s ¼ 1:35 V o 2=3 1 f p 2 r eff m g l cos q (3) Ian Wark Research Institute, University of South Australia, Mawson Lakes, SA 5095, Australia. E-mail: [email protected] † Electronic supplementary information (ESI) available. See DOI: 10.1039/c2sm26529j This journal is ª The Royal Society of Chemistry 2012 Soft Matter Dynamic Article Links C < Soft Matter Cite this: DOI: 10.1039/c2sm26529j www.rsc.org/softmatter PAPER Downloaded by University of Massachusetts - Amherst on 04 October 2012 Published on 18 September 2012 on http://pubs.rsc.org | doi:10.1039/C2SM26529J View Online / Journal Homepage

Upload: rossen

Post on 11-Oct-2016

213 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Spontaneous liquid marble formation on packed porous beds

Dynamic Article LinksC<Soft Matter

Cite this: DOI: 10.1039/c2sm26529j

www.rsc.org/softmatter PAPER

Dow

nloa

ded

by U

nive

rsity

of

Mas

sach

uset

ts -

Am

hers

t on

04 O

ctob

er 2

012

Publ

ishe

d on

18

Sept

embe

r 20

12 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C2S

M26

529J

View Online / Journal Homepage

Spontaneous liquid marble formation on packed porous beds†

Catherine P. Whitby,* Xun Bian and Rossen Sedev

Received 2nd July 2012, Accepted 6th September 2012

DOI: 10.1039/c2sm26529j

The encapsulation of aqueous and organic solvents with particles used to form liquid marbles implies

there are attractive interactions between the particles and those different liquids. This is often masked,

however, by the impact of the droplet kinetic energy on marble formation. We investigated droplet

wetting and evaporation when drops were gently placed (without rolling or shaking) on beds of

silanised glass beads. Particle coating of the drop surface occurred within seconds of liquid contact with

the particle bed. This ruled out evaporation from causing the particles to appear to rise up the surface of

the drop as it was reduced in volume. Particles attach to the fresh liquid surface created during the

droplet oscillations immediately after contact. The further ordered advance of the particles over the

drop surface and the close-packed arrangement of the particles revealed the role of capillary forces in

the coating process. By minimising the kinetic energy of the droplet contact with the particles, we found

that maximum particle coating occurs at liquid surface tensions just above the critical wetting tension of

the beads.

Introduction

The wetting behaviour of fine particles at air–liquid surfaces1 is

important in powder and soil technologies where liquid droplets

are in contact with granular or powdery surfaces. These include

dispersion, granulation, filtration and water repellency. Liquid

cannot penetrate into powders that consist of poorly wetting

particles. The particles may encapsulate drops that are rolled

across the bed, forming liquid core–particle shell structures

(liquid marbles) that have large contact angles on any

substrate.2–4 Several studies have exploited ‘‘liquid marbles’’ for

transporting small volumes of high surface tension liquids,5

storing water in a powdered form6,7 and for the encapsulation

and delivery of active (water soluble) ingredients.8,9 Recently

Gao and McCarthy,10 Xue et al.11 and Matsukuma et al.12

reported encapsulating liquids with lower surface tensions (down

to 20 mN m�1) using fluorinated polymeric powders. All the

liquid marbles had large contact angles. On the other hand,

McHale et al.13 showed that simply depositing water drops onto

beds of hydrophobic particles causes the underside of the drops

to be coated with particles, but no liquid marble is formed. These

observations suggest that the particle surface energy influences

liquid marble formation. They are complicated, however, by the

contribution of the drop kinetic energy to the particle wetting

and attachment processes. In this paper, we study the attachment

and movement of particles over the surfaces of stationary drops

of ethanol–water solutions. The particle wettability (q) was

Ian Wark Research Institute, University of South Australia, MawsonLakes, SA 5095, Australia. E-mail: [email protected]

† Electronic supplementary information (ESI) available. See DOI:10.1039/c2sm26529j

This journal is ª The Royal Society of Chemistry 2012

varied systematically from hydrophobic (<90�) to hydrophilic

(>90�) by controlling the ethanol concentration and hence liquid

surface tension (gl).

A liquid droplet (of radius, R and viscosity, m) placed on the

surface of a bed of particles (of radius, r) may absorb into the bed

(Fig. 1a) or evaporate (Fig. 1b), depending on the particle

wettability. The average rate of imbibition of liquid into a porous

substrate is determined by the pore geometry and q.14 For steady

flow conditions and small pores, the Washburn equation relates

the capillary driving force of a liquid penetrating through a

compact bed of particles and the viscous drag as

d2

t¼ reffgl cos q

2m(1)

where d is the depth of liquid penetrating into the bed in time

t. reff is the effective pore radius. For the special case of powder

beds that can be treated as uniform arrays of cylindrical

pores,15,16 reff is calculated using

reff ¼ 2(1 � fp)/fprpAp (2)

where Ap is the specific surface area of the particles. fp, the

porosity parameter, is determined by the ratio mp/VTrp, where

mp is the mass of particles in the bed, VT is the bed volume and rpis the particle density. Assuming that the three phase contact line

remains stationary as liquid drains into the pores (the constant

drawing, or constant radius limiting case17), the penetration or

wicking time, s, is given by

s ¼ 1:35Vo

2=3

�1� fp

�2reff

m

gl cos q(3)

Soft Matter

Page 2: Spontaneous liquid marble formation on packed porous beds

Fig. 1 Behaviour of ethanol–water droplets (10 mL in volume) on

packed beds of hydrophobic glass spheres (98 mm in diameter). (a)

Ethanol-rich drops wet the particles and penetrate into the packed bed

within seconds, leaving a circular depression on the bed surface. (b)

Water-rich drops wet the particles only partially and remain trapped on

the packed beds; they evaporate over a few hours. (c) Water drops rolled

over the packed bed surface become encapsulated by a layer of particles,

i.e. form liquid marbles. The particle shell is robust and the liquid marble

can be rolled onto other surfaces without breaking or leaking. (d) A water

marble sitting on the packed bed. (e) Particles coat drops of intermediate

alcohol concentrations within a few seconds of drop impact on the bed.

The needle shown in the top image is 1.7 mm wide.

Dow

nloa

ded

by U

nive

rsity

of

Mas

sach

uset

ts -

Am

hers

t on

04 O

ctob

er 2

012

Publ

ishe

d on

18

Sept

embe

r 20

12 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C2S

M26

529J

View Online

where Vo is the initial liquid volume. Although suited to a large

volume of liquid, Marmur17 showed that this theory describes the

kinetics of liquid penetration from a single large drop (R/r >

�30).

Above a critical q, s/N and the droplet is effectively trapped

on the bed surface. For water drops on solid surfaces, q ¼ 90� isthe boundary between hydrophilic and hydrophobic solids.

Providing that porous solids behave like bundles of capillary

tubes, the 90� threshold will also apply. Ban et al.18 observed a

lower threshold, however, for powders of spherical particles wet

by ethanol–water solutions. The shape of a liquid surface inside a

cylindrical capillary does not change during imbibition. In

contrast, capillaries formed by the voids between close-packed

spherical particles do not possess parallel sides and the curvature

of the liquid surface keeps changing. Assuming that the radius of

the droplet is much larger than the size of the voids, Ban et al.18

Soft Matter

predicted that highest contact angle at which wetting occurs (qc)

is qc is 50.7�. Ban et al.18 found qc ¼ 50–58� for polymeric,

silanised glass and sulphur powders of various sizes. Compacted

powder beds were prepared by centrifuging polyethylene tubes

filled with powder. Shirtcliffe et al.19 measured critical contact

angles of 52–61� for fluorinated glass beads and 61–65� for

fluorinated sand. These analyses neglected the effect of the

hydrostatic pressure. Recently Extrand and Moon14 showed that

imbibition may be prevented even at very low q for drops

comparable in size to the pores between spherical particles.

A small drop of fluid trapped on a surface will adopt a

spherical cap shape, of contact radius, rd, and height, hd and

evaporate over time. Assuming that evaporation is due to

diffusion through the liquid vapour interface and that the vapour

concentration gradient is radially outwards, McHale et al.20

showed that the evaporation rate can be expressed as

dV/dt ¼ �lhd (4)

where l ¼ 2pD(cN � c0)/r. D is the vapour diffusion coefficient

and c0 and cN are the vapour concentrations at the drop surface

and far removed from the surface, respectively. For non-wetting,

pure liquids (q > 90�), the contact radius decreases while the

contact angle remains constant during evaporation.20 Under

these conditions20

rd2 ¼ C � 2lt sin2

q

pð1� cos qÞð2þ cos qÞ (5)

where C is a constant. The evaporation of binary liquid mixtures

has been examined experimentally. Sefiane et al.21 and Cheng

et al.22 observed that the more volatile ethanol evaporates before

the water from ethanol–water drops. A kind of ‘‘stick–slip’’

behaviour or transition regime is observed in between, as the

contact area de-pins and q increases to reflect the higher water

content of the drop. Comparison between roughened and

smooth substrates suggests that droplet pinning on textured

surfaces can slow evaporation.23

Solid particles strongly adhere to liquid surfaces. The depth to

which a particle immerses into the liquid (H) is given by

H ¼ r(1 + cos q) (6)

Thus as q/ 180� particles attached to droplet surfaces form a

barrier that prevents the liquid from contacting other surfaces.

Hence Aussillous and Quere24,25 showed that weak forces can

actuate motion of water and glycerol drops encapsulated by

silanised, naturally textured grains (lycopodium spores) with q �160 to 165�. Liquid marbles are formed by agitating drops on

very hydrophobic powders (hydrophobised silica, fluoropoly-

mers and graphite, Fig. 1c). Coatings formed by perfluoroalkyl

particles (q � 177�, up to 35 mm in size) are unusually robust and

long-lived.10 Water marbles move reversibly under electrowetting

conditions due to the high water surface tension and conform-

ability of the particle coating (silanised lycopodium spores).26

The rate of water evaporation from marbles (graphite micro-

powder) is substantially slower than from uncoated droplets.27

The drop profile remains spherical as the liquid volume

decreases, until the maximum surface packing density is reached

and the coating crumples into a compressed residue.27

This journal is ª The Royal Society of Chemistry 2012

Page 3: Spontaneous liquid marble formation on packed porous beds

Dow

nloa

ded

by U

nive

rsity

of

Mas

sach

uset

ts -

Am

hers

t on

04 O

ctob

er 2

012

Publ

ishe

d on

18

Sept

embe

r 20

12 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C2S

M26

529J

View Online

Liquid drops may be partially coated by hydrophobic parti-

cles. McEleney et al.28 observed that the mass of PMMA beads

(q� 120�, dp ¼ 42 mm) or copper powder (q� 157�, dp ¼ 9, 15 or

320 mm) collected by a water drop rolled on the powder bed was

proportional to the drop surface area. The ratio of the powder

mass to surface area varied with the particle density, size and

contact angle. Hapgood and Khanmohammadi29 observed liquid

marbles form prior to penetration of drops into beds of particles

(tens of micrometres in size) with q < 90�. The rate of penetrationby liquid sprayed onto a bed is delayed when particles adhere to

the drops.30 Eshtiaghi et al.31 observed that the impact of a drop

(of aqueous glycerol) on a bed of polytetrafluoroethylene spheres

(diameter, dp ¼ 100 mm) could cause liquid marble formation.

They found that the coated surface area depends on the drop

kinetic energy and m, but is less sensitive to gl.31 McHale et al.13

found that drops of alkanes of short chain lengths were partially

coated by coarse fluorinated silica beads under saturated vapour

conditions. Crucially, the mechanism of particle coating and the

roles of particle wettability and liquid absorption or evaporation

in the coating process remain unclear.

Here we present the results of a systematic study of particle

coating of alcohol–water drops placed onto porous beds of

silanised coarse glass spheres (98 mm in diameter). The kinetic

energy of the droplet impact and hence entrainment of the glass

spheres by liquid flow or oscillations was minimised. The kinetics

of the penetration and evaporation of alcohol–water drops

placed onto the porous beds were investigated to determine their

roles in the coating process. We show that the spheres coat the

drops within minutes. The attached particles form a hexagonally

close packed layer. Maximum coating is achieved when q is just

above the threshold for wetting of the powder bed.

Experimental section

Spherical glass beads (Potters Industries) of three different sizes:

(42 � 4) mm, (98 � 8) mm and (165 � 15) mm in diameter (as

measured from SEM images) were hydrophobised using tri-

chloro(1H,1H,2H,2H-perfluorooctyl)silane (97%, Sigma-

Aldrich). The particles were mixed with a solution of the silane

dissolved in 4 : 1 (v/v) mixture of hexadecane (99+%, Sigma-

Aldrich) and chloroform (99.8%, Chem-Supply) in a dry

nitrogen environment. Planar glass slides treated in the same

fashion exhibited a water contact angle of 108�.Particle beds were prepared by packing hydrophobised glass

beads into cylindrical containers, 25 mm in diameter and 11 mm

in height. The particles were packed gradually while tapping the

container to remove air pockets. A clean glass microscope slide

was used to smooth the surface of the bed. The effective pore

radius in the beds was derived frommeasurements of the pressure

required to prevent capillary flow of cylcohexane (a perfectly

wetting liquid) in beds of the particles (capillary pressure

technique).

Liquid drops (10 mL in volume) were gently placed on the

surfaces of the packed beds. The ratio of the drop volume to the

volumes of the pores in the beds was about 105, ensuring that

the penetration kinetics could be described by the Lucas–

Washburn theory.17 The liquids used were mixtures of ethanol

(99%, Chem-Supply) and ultrapure water, toluene (99%, Merck),

cyclopentanol (98%, Ajax), ethylene glycol (99%, Sigma

This journal is ª The Royal Society of Chemistry 2012

Aldrich), glycerol (99%, Chem-Supply) and methyl-

trioctylammonium trifluoroacetate (Merck). The evolution of

the drop shape during penetration or evaporation was recorded

using the video camera of a sessile drop setup (OCA20, Data-

Physics). The drop height, hd, drop radius, rd, and the height of

the particle coating, hp, were measured at different times from

images extracted from the video recording using ImageJ soft-

ware. The time for a drop to fully penetrate into the dry bed, s,was taken as the time from drop deposition to the moment when

no liquid remained on the porous surface, i.e. when hd ¼ 0.

Results

The possible behaviour of a liquid droplet deposited on a packed

bed is shown in Fig. 1. If the liquid wets the particles it will pene-

trate into the capillary channels under the contact area rather

quickly (Fig. 1a). If the particles are hydrophobic enough, the top

layer of particles behaves like a superhydrophobic surface. The

droplet sits on the particles but microscopically, a major portion of

the nominal contact area consists of separate liquid–air interfaces.

The droplet evaporates after a longer period of time (Fig. 1b). If the

droplet is rolled rather than deposited carefully on the surface, it

will collect many hydrophobic particles and form a liquid marble

(Fig. 1d). However, even a carefully deposited droplet may be

covered spontaneously with a particle layer, as shown in Fig. 1e.

When the drop impacts on the bed, it advances over the surface

to some extent. Hydrophobic particles attach on contact with the

freshly created liquid surface. Droplet spreading is opposed by

the poor wettability of the particles, causing the droplet to

retract. So the attached particles will be carried away from the

bed. If spherical, as in Fig. 1e, they form a hexagonal arrange-

ment; irregularly shaped particles form disordered arrays.

Particle coating during drop shape oscillations has been

described by Eshtiaghi et al.31 In the experiments described here,

however, the kinetic energy of the impact was minimised by

gentle deposition. The oscillations only last about 100 ms.

During this time the particle coating rises to about 66% of its

final height. Over the next few hundred milliseconds, the particle

array slowly advances to its final height over the now steady

surface. After 1 s there are only particle rearrangements within

the layer. In time, the liquid will evaporate from the armoured

droplet. The particle shell will sink back into the bed or remain

trapped on the bed surface, forming a crumpled residue.

For particles of a given size and shape, the area of the drop

that is coated by particles and the subsequent behaviour of the

drops (penetration into the bed, evaporation, or both) depends

on the surface tension of the liquid. Alcohol-rich drops (gl < 27

mN m�1) penetrate the packed bed within seconds. The kinetics

of penetration of a 10 mL drop is shown in Fig. 2. The drop

height, hd (solid line), decreases as liquid is quickly sucked into

the capillary space formed between the packed particles. As the

drop height decreases, particles coat the drop surface up to a

height hp (open symbols; hp# hd). The particle coating sinks with

the liquid as it is drawn into the bed. The time, s, taken for the

drop to penetrate completely into the porous substrate (with no

liquid remaining on the surface) is just over 1 s. During pene-

tration, the drop radius, rd (dashed line), increases very little as

particles on the surface of the bed pin the contact line quite

efficiently.

Soft Matter

Page 4: Spontaneous liquid marble formation on packed porous beds

Fig. 2 Droplet penetration. (a) Characteristic parameters of a sessile

water–ethanol drop (V ¼ 10 mL, gl ¼ 26.8 mN m�1) on a packed bed of

hydrophobic particles (98 mm in diameter). (b) Kinetic behaviour of the

drop: drop height, hd (solid line), drop radius, rd (dashed line) and height

of the particle coating, hp (open points) as functions of time. The drop

quickly penetrates the porous bed of hydrophobic glass spheres. The

penetration time, s, is the time taken for the whole droplet to disappear.

Fig. 3 Droplet evaporation. Kinetic behaviour of partially wetting

droplets (V ¼ 10 mL, gl ¼ 28.8 mN m�1) and slowly evaporate on (a) a

hydrophobic glass slide and (b) a packed bed (glass spheres, 98 mm in

diameter) of identical wettability. Parameters defined as in Fig. 2a: drop

height, hd (solid line), drop radius, rd (dashed line) and height of the

particle coating, hp (open points). (c) Profiles of an encapsulated drop

during evaporation. The droplet diameter in the upper image is 2.4 mm.

Dow

nloa

ded

by U

nive

rsity

of

Mas

sach

uset

ts -

Am

hers

t on

04 O

ctob

er 2

012

Publ

ishe

d on

18

Sept

embe

r 20

12 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C2S

M26

529J

View Online

For surface tensions lower than 26 mN m�1, the particle

coating does not appear to slow down liquid penetration. The

coating particles readily sink back into the bed as the liquid

drains out. However, at 26 <gl # 27 mN m�1, drops are rapidly

and fully coated by particles. This slows down penetration and sis now several seconds long. After penetration is complete, a

shallow depression is observed in the flattened packed bed.

The drop penetration time, s, was used to estimate the contact

angle of the ethanol–water solutions on the particles through eqn

(2). The derived angles were compared with those measured with

the same liquids on a flat glass plate silanised in the same manner

as the particles (Fig. S1 in ESI†). Except in the case of very short

penetration times (<0.2 s), the derived contact angles were

consistent with advancing contact angles on the flat surfaces.

The lowest contact angle at which a liquid drop sits on the

particle bed without any penetration is close to 84�. The criticalsurface tension of the particles was 15 mN m�1 as obtained from

linear plots of the cosine of the contact angle, cos q, against the

surface tension of the wetting liquid, gl (Zisman plot). The crit-

ical wetting angle is achieved by drops with gl > 27 mN m�1;

these poorly wet the particles, do not penetrate into the packed

bed, and evaporate.

Water-rich droplets evaporate within about 100 minutes of

being deposited on the particle bed, in a similar fashion to drop

evaporation on flat plates. Drops are slowly coated by particles

in the late stages of evaporation, when the droplet contact area

de-pins. In contrast, drops of intermediate alcohol concentra-

tions (27 mN m�1 < gl < 45 mN m�1) are coated by particles

within seconds after deposition (Fig. 1d). The average rate of

coating increases as the surface tension decreases. Coated

droplets evaporate at slower rates than drops placed on a smooth

hydrophobic surface.

Soft Matter

The kinetics of drop evaporation (at 27 mN m�1 < gl < 45

mN m�1) on flat hydrophobic surfaces and on porous bed of

hydrophobic spheres is shown in Fig. 3a and b, respectively.

During the first 100 seconds, the radius and height of drops on

the flat surface decrease very slightly. In the next stage (�400 s),

the radius decreases dramatically as the drop contact angle

increases. The drop height and radius then decrease simulta-

neously as evaporation proceeds (at t > 500 s). In contrast, the

radius of drops on the porous bed increases slightly while the

height of the drop decreases, during the first 100 seconds.

Particles coat the drop up to a height hp (<hd) during the first few

seconds of this period. Further coating of the drops is driven by

the subsequent reduction in the drop radius, apparently trapping

the drop at a relatively low contact angle. After about 200 s, the

drop height decreases while the drop radius remains roughly

constant. Fig. 3c shows that the encapsulated drop collapses in

the centre. After 100 min, there are no apparent further changes,

so that a crumpled shell remains on top of the packed bed

(Fig. 3c).

Comparisons of the height of the particle coating, hp, on

droplet surfaces were made one second after placing the drops on

the powder bed. This characteristic time, sc, was selected so that

the coating is fully formed but evaporation is rather insignificant

(Fig. 3b). The hp (t ¼ sc) increases dramatically as the particle

wettability approaches the angle corresponding to the surface

tension below which penetration occurs (Fig. 4a). Particle

coating was also observed with a variety of other liquids such as

alkanes, aromatics, esters and their mixtures. For low viscosity

solvents, the relation between hp and gl is essentially the same

(Fig. 4, solid line). For highly viscous solvents, such as glycerol

This journal is ª The Royal Society of Chemistry 2012

Page 5: Spontaneous liquid marble formation on packed porous beds

Fig. 4 Dependence of (a) the height of the particle coating, hp, nor-

malised with the droplet height, hd, and (b) droplet penetration time, s, onthe particle wettability, q, for drops placed on packed beds of hydro-

phobic glass spheres: 98 mm (circular points) and 42 mm (triangular

points). Drops (V ¼ 10 mL) consisted of aqueous ethanol solutions and

pure solvents (solvent type indicated in the graph). qc is the contact angle

of the particles that corresponds to their critical wetting tension. sc is thecharacteristic time for particle coating of the drops. The lines are guides

to the eye.

Dow

nloa

ded

by U

nive

rsity

of

Mas

sach

uset

ts -

Am

hers

t on

04 O

ctob

er 2

012

Publ

ishe

d on

18

Sept

embe

r 20

12 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C2S

M26

529J

View Online

(gl ¼ 64 mN m�1, m ¼ 1490 mPa s) or ionic liquids (e.g. meth-

yltrioctylammonium nitrate, gl ¼ 27 mN m�1, m ¼ 2316 mPa s)

the rate of particle coating is substantially slower.

Similar relations between drop penetration, evaporation and

particle coating and the liquid surface tension were observed with

hydrophobic particles between 90 and 180 mm in size. Longer

penetration times, s, was observed for alcohol-rich drops (gl < 25

mN m�1) on beds of 44 mm particles. Particles that coat drops

during penetration readily sink back into the bed and do not slow

down penetration. The complete coating of the drops occurs,

however, at slightly lower contact angle values for 44 mm sized

particles (Fig. 4a). Drops are coated within seconds of deposi-

tion. This slows down penetration and s ranges between 1 and

100 seconds (Fig. 4b).

Discussion

We have manipulated the behaviour of small droplets (10 mL)

gently deposited on packed beds of hydrophobic glass particles

This journal is ª The Royal Society of Chemistry 2012

(from 38 to 180 mm in size) by varying the liquid surface tension.

At low surface tension the liquid wets the particles well. Capillary

suction into the porous space formed between the packed

particles quickly drains the liquid from the droplet. At larger

surface tensions, the wettability is worse, i.e. the contact angle is

larger. Liquid penetration into the packed bed is slowed down or

even completely prevented by the upward action of the capillary

pressure. In the latter case the droplet sits on the surface of the

packed bed, which behaves like a superhydrophobic surface and,

over time, the liquid evaporates. A small depression is left on the

bed surface where the droplet was initially deposited. At inter-

mediate wettability (82� < q < 100�, which corresponds to 27 mN

m�1 < gl < 45 mN m�1) the rather large particles (�100 mm)

attach strongly to the liquid surface and climb up the droplet

forming a closely packed coating.

In order to model the relation between surface tension and

wettability we use here the surface of a glass slide silanised in the

same fashion as the particles as a model surface. The slide is very

hydrophobic (advancing and receding water contact angles of

120� and 110�, respectively) and contact angle obtained with

various liquids are shown in Fig. S2 in ESI.† For a circular

capillary vertically immersed in a liquid, the transition between

wetting (i.e. capillary rise) and dewetting (i.e. capillary depres-

sion) occurs exactly at a contact angle of 90�. On a packed bed,

however, the geometry of the surface plays a substantial role. In

tightly packed beds the contact angle characteristic of neutral

wettability is around 60�.19 Here the bed packing is looser, to

enable marble formation. The transition from penetration to

non-penetration occurs at about 80� for the 90–106 mm particles.

At lower surface tensions the droplets penetrate the porous

bed. The kinetics of penetration at constant drop diameter is

given by eqn (1). The droplet diameter stays nearly constant for a

period of time – see Fig. 2b, 3a and b. The surface of the packed

bed is essentially a rough hydrophobic surface and it is not

surprising that the contact line stays pinned until evaporation

destroys the droplet (Fig. 3a and b). The penetration time can be

used to estimate correctly the advancing contact angle (Fig. S1 in

ESI†). This is because the model captures the physics of the

process, i.e. the capillary suction is well described by the Wash-

burn equation, the contact area between the liquid and the solid

is approximately constant (the contact line is pinned), and

evaporation is negligible at such short times. The penetration

time for a droplet of a given surface tension is characteristic of

the wettability of the porous material and is commonly used for

wettability determination, e.g. in soil science.32–34 However, in

our case the times are very short (�1 s, Fig. 2b) and therefore

practically inconvenient.

Particle coating dominates drop wetting behaviour when q $

qc � 90�. Hydrophilic and hydrophobic particles will partition

between a polar aqueous phase and air which is effectively

hydrophobic. Particles of intermediate hydrophobicity will

attach very strongly to the interface between the two phases.35–37

Thus the particles coating the liquid drop (Fig. 1e) are in a more

favourable energetic situation than those in the powder bed, i.e.

dry. The interfacial free energy gained for single a spherical

particle transferred from the air to the liquid surface is given

by38,39

DF ¼ glprp2(1 + cos q)2 (7)

Soft Matter

Page 6: Spontaneous liquid marble formation on packed porous beds

Dow

nloa

ded

by U

nive

rsity

of

Mas

sach

uset

ts -

Am

hers

t on

04 O

ctob

er 2

012

Publ

ishe

d on

18

Sept

embe

r 20

12 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C2S

M26

529J

View Online

We use contact angles measured on planar silanised surfaces to

estimate DF. The adsorption free energy for a hemispherical

droplet (rd ¼ 1 mm) fully covered with particles (rp ¼ 50 mm and

724 particles) is shown in Fig. 5. It is clear that the driving force

for coating the droplet with particles decreases sharply as the

surface tension increases.

Our results show a direct correlation between the area of the

droplet surface coated by particles and the liquid surface tension

(cf. Fig. 4 and 5). Crucially, maximum particle coating of

droplets occurs at conditions just above those where penetration

into the bed becomes significant (q$ qc). For drops of liquid with

much higher surface tension (gl > 45 mN m�1) the bed surface is

effectively superhydrophobic. Particle coating does not occur

unless the drop is rolled over the bed surface.

There is also a decrease when the surface tension drops below

20.0 mNm�1, but this region is masked in our experiments by the

very rapid liquid penetration (Fig. 2b). While particles should be

partially wetted, penetration needs to be sufficiently slow for

particle coating to occur. Particle coating may slow liquid

imbibition into the bed. In this case, typically the particles tend to

sink back into the bed with the liquid, which accounts for the

ring-shaped depressions left in the bed. This behaviour is more

apparent for small particles where liquid takes several seconds to

drain into the smaller channels in the porous bed.

We also see a partial coating, i.e. hp < hd, on droplets of 30 mN

m�1 < gl < 40 mN m�1 (Fig. 5). The most intuitive explanation

seems to be that the potential energy of the climbing particles

increases. However, the estimate of the work done against

gravity (mghp, where m is the mass of a single particle) is about

10 pJ, i.e. at least one order of magnitude lower than the esti-

mated free energy of adsorption (Fig. 5). As the packing of the

adsorbed particles is dense – patches of hexagonally arranged

spheres are clearly seen – we speculate that friction between the

coating particles is the limiting factor.

In the simplified consideration above, we neglected the capil-

lary forces. Attached particles floating on droplet surfaces attract

each other to form a layer of close packed particle clusters.

Lateral capillary forces40,41 caused by deformation of the liquid

meniscus can be very strong. Paunov et al.42 showed that capil-

lary attraction between two particles of the same contact angle

Fig. 5 Theoretically predicted variation in the adsorption free energy for

a hemispherical droplet (rd ¼ 1 mm) fully covered with particles (rp ¼ 50

mm) with the liquid surface tension (gl) calculated using eqn (7).

Soft Matter

increases as the surface tension decreases. This will further

accentuate the experimental correlation shown in Fig. 4.

At very high gl, the droplets are not spontaneously coated by

particles. The gain in interfacial free energy due to particle

attachment is relatively low (Fig. 5). Evaporation dominates

droplet behaviour instead.Droplets of pure water (V¼ 10 mL, q¼108�) deposited on a particle bed evaporate in about the same time

as drops placed on a hydrophobic glass surface (�6000 s).

Consistent with the bed behaving like a superhydrophobic

surface, comparable evaporation times have been measured for

water droplets on polymeric substrates.20,21 For example, Sefiane

et al.21 found that 5 mLwater droplets deposited on PDMS-coated

silicon wafers evaporated within about 3000 seconds.

Evaporation of drops trapped on top of beds (q [ qc) can

cause particle coating due to the reduction in the droplet surface

area. Although the attached particles partly obscure the shape of

the drop, the time dependent variation in the shape of water–

ethanol binary drops shows several stages, corresponding to the

sequential evaporation of the different components.21,43 Ethanol

molecules diffuse to the droplet surface and evaporate first. The

drop radius then decreases as the contact line de-pins and the

contact angle increases to a value consistent with the water-rich

composition of the droplet surface. Coating occurs during this

period, perhaps driven by the stick–slip motion of observed for

drops at the late stages of drop evaporation.21,43 McHale et al.13

also observed drops on particle beds being coated by particles as

the droplet volume was reduced by evaporation.

The relative height of the particle coating that assembles

within minutes of contact for water–ethanol drops at q $ qc, is

however, significant (Fig. 4b). Under these conditions the rate of

particle coating is significantly larger than the rate of evapora-

tion. Typically, the coating is incomplete with the top portion of

the drop surface remaining exposed. The smaller area of the

droplet exposed to the atmosphere evaporates quickly. Reduc-

tion in the contact angle causes more coating and the drop radius

de-pins until the drop is fully coated by particles. Importantly,

the surface area of the hydrophobic particle coating does not

change over time. Instead the surface area of the drop decreases

until it is fully coated by particles. The presence of the complete

particle coating then causes significant retardation of droplet

evaporation.

We described here the spontaneous coating of liquid drops by

rather large solid particles. Similar to the contact angle depen-

dence of the colloidal structures formed by nanoparticle disper-

sions (foams, climbing films and marbles44), the height of the

particle coating depends on the particle contact angle and the

surface tension of the liquid. This selective process has potential

applications where mixed colloids are assembled into granular

structures. Foods such as milk that are spray dried into powders,

for example, are mixtures of fat droplets, proteins and sugars.

Using spontaneous particle coating as a precursor to drying

could control the spatial distribution of particles of different

wettability in the agglomerates and hence influence powder

rehydration processes (sinking, re-dispersion, dissolution).

Conclusions

We have described the behaviour of drops deposited on powder

beds as the wettability of the particles was varied systematically

This journal is ª The Royal Society of Chemistry 2012

Page 7: Spontaneous liquid marble formation on packed porous beds

Dow

nloa

ded

by U

nive

rsity

of

Mas

sach

uset

ts -

Am

hers

t on

04 O

ctob

er 2

012

Publ

ishe

d on

18

Sept

embe

r 20

12 o

n ht

tp://

pubs

.rsc

.org

| do

i:10.

1039

/C2S

M26

529J

View Online

by increasing the liquid surface tension, while keeping the

contribution of the droplet kinetic energy at a minimum. Liquid

drains into the pores of beds of wettable particles due to capillary

suction. The time taken for all the liquid to penetrate into the bed

increases with the surface tension until, just above the critical

wetting tension, it becomes very long. Particles attach to the

liquid surface as the drop spreads over the bed. The attached

particles form a hexagonally close-packed layer that encapsulates

the drop surface. The height of the particle coating decreases as

the liquid surface tension is increased. Thus drops with high

surface tension must be rolled forcefully on beds of low surface

energy particles for encapsulation to occur. Rolling causes liquid

circulation within the drops that carries attached particles away

from the bed so that the drops become fully coated with particles,

irrespective of the liquid surface tension (providing it is greater

that the critical wetting tension).

Acknowledgements

CPW acknowledges receipt of an Australian Research Council

Future Fellowship. This research was supported under the

Australia Research Council Linkage Project funding scheme

(project LP0667608). It was also supported by the Department of

Innovation, Industry, Science and Research (Australian

Government) through the Australia-India Strategic Research

Fund.

References

1 T. N. Hunter, G. J. Jameson, E. J. Wanless, D. Dupin andS. P. Armes, Langmuir, 2009, 25, 3400.

2 E. Bormashenko, Curr. Opin. Colloid Interface Sci., 2011, 16, 266.3 G. McHale and M. I. Newton, Soft Matter, 2011, 7, 5473.4 N. Eshtiaghi and K. P. Hapgood, Powder Technol., 2012, 223, 65.5 A. V. Rao,M.M. Kulkarni and S. D. Bhagat, J. Colloid Interface Sci.,2005, 285, 413.

6 B. P. Binks and R. Murakami, Nat. Mater., 2006, 5, 865.7 L. Forny, I. Pezron, K. Saleh, P. Guigon and L. Komunjer, PowderTechnol., 2007, 171, 15.

8 D. Dupin, S. P. Armes and S. Fujii, J. Am. Chem. Soc., 2009, 131,5386.

9 S. Fujii, S. Kameyama, S. P. Armes, D. Dupin, M. Suzaki andY. Nakamura, Soft Matter, 2010, 6, 635.

10 L. C. Gao and T. J. McCarthy, Langmuir, 2007, 23, 10445.11 Y. Xue, H. Wang, Y. Zhao, L. Dai, L. Feng, X. Wang and T. Lin,

Adv. Mater., 2010, 22, 4814.12 D. Matsukuma, H. Watanabe, H. Yamaguchi and A. Takahara,

Langmuir, 2011, 27, 1269.

This journal is ª The Royal Society of Chemistry 2012

13 G. McHale, N. J. Shirtcliffe, M. I. Newton, F. B. Pyatt andS. H. Doerr, Appl. Phys. Lett., 2007, 90, 054110.

14 C. W. Extrand and S. I. Moon, Langmuir, 2012, 28, 3503.15 S. Middleman, Modelling Axisymmetric Flows: Dynamics of Films,

Jets and Drops, Academic Press, San Diego, 1995.16 M. Denesuk, G. L. Smith, B. J. J. Zelinski, N. J. Kreidl and

D. R. Uhlmann, J. Colloid Interface Sci., 1993, 158, 114.17 A. Marmur, J. Colloid Interface Sci., 1988, 124, 301.18 S. Ban, E. Wolfram and S. Rohrsetzer, Colloids Surf., 1987, 22,

291.19 N. J. Shirtcliffe, G. McHale, M. I. Newton, F. B. Pyatt and

S. H. Doerr, Appl. Phys. Lett., 2006, 89, 094101.20 G.McHale, S. M. Rowan,M. I. Newton andM. K. Banerjee, J. Phys.

Chem. B, 1998, 102, 1964.21 K. Sefiane, S. David and M. E. R. Shanahan, J. Phys. Chem. B, 2008,

112, 11317.22 A. K. H. Cheng, D. M. Soolaman and H. Z. Yu, J. Phys. Chem. B,

2006, 110, 11267.23 C. Bourges-Monnier and M. E. R. Shanahan, Langmuir, 1995, 11,

2820.24 P. Aussillous and D. Quere, Nature, 2001, 411, 924.25 P. Aussillous and D. Quere, Proc. R. Soc. A, 2006, 462, 973.26 G. McHale, D. L. Herbertson, S. J. Elliott, N. J. Shirtcliffe and

M. I. Newton, Langmuir, 2007, 23, 918.27 M. Dandan and H. Y. Erbil, Langmuir, 2009, 25, 8362.28 P. McEleney, G. M. Walker, I. A. Larmour and S. E. J. Bell, Chem.

Eng. J., 2009, 147, 373.29 K. P. Hapgood and B. Khanmohammadi, Powder Technol., 2009,

189, 253.30 J. O.Marston, S. T. Thoroddsen, W. K. Ng and R. B. H. Tan, Powder

Technol., 2010, 203, 223.31 N. Eshtiaghi, J. S. Liu, W. Shen andK. P. Hapgood, Powder Technol.,

2009, 196, 126.32 J. Bachmann, S. K. Woche, M. O. Goebel, M. B. Kirkham and

R. Horton, Water Resour. Res., 2003, 39, 1353.33 S. H. Doerr, R. A. Shakesby and R. P. D. Walsh, Earth-Sci. Rev.,

2000, 51, 33.34 J. Letey, M. L. K. Carrillo and X. P. Pang, J. Hydrol., 2000, 231–232,

61.35 B. P. Binks, Curr. Opin. Colloid Interface Sci., 2002, 7, 21.36 E. Dickinson, Curr. Opin. Colloid Interface Sci., 2010, 15, 40.37 O. J. Cayre, J. Hitchcock, M. S. Manga, S. Fincham, A. Simoes,

R. A. Williams and S. Biggs, Soft Matter, 2012, 8, 4717.38 J. Diao and D. W. Fuerstenau, Colloids Surf., 1991, 60, 145.39 S. Levine, B. D. Bowen and S. J. Partridge, Colloids Surf., 1989, 38,

325.40 D. Y. C. Chan, J. D. Henry and L. R. White, J. Colloid Interface Sci.,

1981, 79, 410.41 P. A. Kralchevsky and K. Nagayama, Adv. Colloid Interface Sci.,

2000, 85, 145.42 V. N. Paunov, P. A. Kralchevsky, N. D. Denkov and K. Nagayama,

J. Colloid Interface Sci., 1993, 157, 100.43 K. Sefiane, L. Tadrist and M. Douglas, Int. J. Heat Mass Transfer,

2003, 46, 4527.44 P. D. I. Fletcher and B. L. Holt, Langmuir, 2011, 27, 12869.

Soft Matter