droplet formation in a t-junction microfluidic device in...

9
DROPLET FORMATION IN A T-JUNCTION MICROFLUIDIC DEVICE IN THE PRESENCE OF AN ELECTRIC FIELD Zeeshan Ahmad Department of Mechanical Engineering Indian Institute of Technology Delhi Hauz Khas, New Delhi 110016, India Email: [email protected] Rattandeep Singh Department of Mechanical Engineering Indian Institute of Technology Delhi Hauz Khas, New Delhi 110016, India Email: [email protected] Supreet Singh Bahga Department of Mechanical Engineering Indian Institute of Technology Delhi Hauz Khas, New Delhi 110016, India Email: [email protected] Amit Gupta 1 Department of Mechanical Engineering Indian Institute of Technology Delhi Hauz Khas, New Delhi 110016, India Email: [email protected] 1 Corresponding author ABSTRACT In this work, the effect of applying an electric field on droplet formation in a T-junction microfluidic device is examined by simulations based on a recent technique known as lattice Boltzmann method (LBM). The electric field is applied in the main channel just beyond the confluence of the continuous and dispersed phases. A combined electrohydrodynamics-multiphase model that can simulate the flow of immiscible fluids in the presence of an electric field is developed and validated. The same model is then applied to study the droplet formation process in a T-junction microfluidic device at a capillary number of 0.01 and at different dispersed to continuous phase flow rate ratios. Results show that there is a decrease in the droplet size and an increase in formation frequency as the electric field is increased. The interplay of the electric and interfacial forces on droplet formation is investigated. NOMENCLATURE B droplet length perpendicular to electric field direction C color function Ca / c c u , capillary number D droplet deformation d discriminating function E electric field e i lattice speed of particles moving in direction i F force L droplet length along electric field direction n normal vector at the interface p pressure Q flow rate ratio- dispersed to continuous phase q volume charge density R inner to outer fluid conductivity ratio Re Reynolds number r o initial radius of the droplet S inner to outer fluid permittivity ratio U electric potential u velocity w i lattice weights Greek Symbols relaxation time gradient interfacial tension Proceedings of the ASME 2015 International Technical Conference and Exhibition on Packaging and Integration of Electronic and Photonic Microsystems and ASME 2015 13th International Conference on Nanochannels, Microchannels, and Minichannels InterPACKICNMM2015 July 6-9, 2015, San Francisco, California, USA InterPACK/ICNMM2015-48388 1 Copyright © 2015 by ASME

Upload: hoangminh

Post on 24-May-2018

218 views

Category:

Documents


2 download

TRANSCRIPT

Page 1: Droplet Formation in a T-Junction Microfluidic Device in ...azeeshan/files/InterPACKICNMM2015...Supreet Singh Bahga Department of Mechanical Engineering ... examined by simulations

DROPLET FORMATION IN A T-JUNCTION MICROFLUIDIC DEVICE IN THE PRESENCE OF AN ELECTRIC FIELD

Zeeshan Ahmad Department of Mechanical Engineering

Indian Institute of Technology Delhi Hauz Khas, New Delhi 110016, India

Email: [email protected]

Rattandeep Singh Department of Mechanical Engineering

Indian Institute of Technology Delhi Hauz Khas, New Delhi 110016, India

Email: [email protected]

Supreet Singh Bahga Department of Mechanical Engineering

Indian Institute of Technology Delhi Hauz Khas, New Delhi 110016, India

Email: [email protected]

Amit Gupta1 Department of Mechanical Engineering

Indian Institute of Technology Delhi Hauz Khas, New Delhi 110016, India

Email: [email protected]

1 Corresponding author

ABSTRACT

In this work, the effect of applying an electric field on

droplet formation in a T-junction microfluidic device is

examined by simulations based on a recent technique known as

lattice Boltzmann method (LBM). The electric field is applied

in the main channel just beyond the confluence of the

continuous and dispersed phases. A combined

electrohydrodynamics-multiphase model that can simulate the

flow of immiscible fluids in the presence of an electric field is

developed and validated. The same model is then applied to

study the droplet formation process in a T-junction microfluidic

device at a capillary number of 0.01 and at different dispersed

to continuous phase flow rate ratios. Results show that there is a

decrease in the droplet size and an increase in formation

frequency as the electric field is increased. The interplay of the

electric and interfacial forces on droplet formation is

investigated.

NOMENCLATURE

B droplet length perpendicular to electric field direction

C color function

Ca /c cu , capillary number

D droplet deformation

d discriminating function

E electric field

ei lattice speed of particles moving in direction i

F force

L droplet length along electric field direction

n normal vector at the interface

p pressure

Q flow rate ratio- dispersed to continuous phase

q volume charge density

R inner to outer fluid conductivity ratio

Re Reynolds number

ro initial radius of the droplet

S inner to outer fluid permittivity ratio

U electric potential

u velocity

wi lattice weights

Greek Symbols

relaxation time

gradient

interfacial tension

Proceedings of the ASME 2015 International Technical Conference and Exhibition on Packaging and Integration of Electronic and Photonic Microsystems and

ASME 2015 13th International Conference on Nanochannels, Microchannels, and Minichannels InterPACKICNMM2015

July 6-9, 2015, San Francisco, California, USA

InterPACK/ICNMM2015-48388

1 Copyright © 2015 by ASME

Page 2: Droplet Formation in a T-Junction Microfluidic Device in ...azeeshan/files/InterPACKICNMM2015...Supreet Singh Bahga Department of Mechanical Engineering ... examined by simulations

density

parameter controlling the width of the interface

curvature

permittivity

conductivity

dynamic viscosity

kinematic viscosity

source term

Subscripts

B blue fluid

R red fluid

i index

c continuous phase

d dispersed phase

INTRODUCTION

The area of droplet-based microfluidics has emerged out of

the need to overcome the shortcomings of continuous-flow

microfluidics such as contamination due to diffusion, surface

adsorption [1] and dispersion, which lead to varying reaction

times affecting throughput of microfluidic devices [2]. These

effects become more prominent as the size of the microfluidic

device is decreased. Droplets provide an ideal platform for

isolating the species within and dealing with very small amounts

of reagents, thereby opening huge opportunities in drug

discovery and biology. The size of droplets formed in

microfluidic channels is affected by the flow rates and

properties of the two immiscible fluids used and geometry of

the microchannel. If these parameters are fixed, it is difficult to

exercise control over the size and frequency of formation of the

droplets. One of the ways to secure greater flexibility over

droplet parameters is by setting up an electric field in the

microchannel.

Electrohydrodynamics (EHD) is the study of fluid motion

in the presence of an external electric field. Depending upon the

physical properties of the fluid like conductivity and

permittivity, a droplet can get deformed into different shapes in

the presence of an electric field. For static fluid, if a droplet acts

as a perfect insulator or is more conductive than the

surrounding fluid, it gets deformed into prolate shape. In some

experimental results, oblate shape of droplets was also noted

and to support those results, Taylor proposed the leaky

dielectric model [3, 4, 5]. According to this model, both the

droplet and the surrounding fluids are considered to have finite

conductivities and when an electric field is applied, charge gets

accumulated at the droplet interface. Due to this charge,

tangential stresses are generated on the surface of the droplet

resulting in an oblate shape. Many theoretical and experimental

studies have been carried out to determine the effect of an

electric field on two-fluid systems [6, 7, 8]. Fernandez et al. [9]

used the finite volume method to study the effect of electrostatic

forces on the distribution of droplets in a channel flow. Zhang

and Kwok [10] analyzed the change in the shape of 2D droplets

in the presence of electric field using the lattice Boltzmann

method (LBM). Tomar et al. [11] studied droplet deformation

using volume-of-fluids approach. All these studies have used

leaky dielectric model for simulating the effect of electric field

on the droplet.

In this paper, the leaky dielectric model is used and

simulations are performed using the lattice Boltzmann method.

LBM is an attractive alternative to conventional computational

fluid dynamics to simulate multicomponent flows analyzed in

this paper as it does not require separate treatment of the

interface. Using LBM, we have developed a coupled

electrohydrodynamics [12]-multiphase [13, 20] model and the

results obtained are validated with the theoretical models. The

same model is then applied to study the droplet formation

process in a T-junction microfluidic device.

MODEL DEVELOPMENT

LBM has been applied to simulate multiphase flow

(silicone oil + water) in a T-junction microfluidic channel in the

presence of an electric field. At the interface of the two phases,

stress boundary condition is imposed by incorporating a source

term in the collision step [14]:

( )( , ) ( , )

eq

i ii i i

f

f ff t t f t

x x (1)

An interfacial force acts normal to the interface with

magnitude proportional to the gradient of the phase field

( ) / ( )R B R BC or the color function [19]:

1

2C F (2)

The equilibrium distribution eq

if in D2Q9 is given by:

29 3[1 3 . ( . ) . ]

2 2

eq

i if w i i i i i ie u e u u u (3)

For the D2Q9 lattice used for simulation, the weights

iw are given by:

4 9 ; 0

1/ 9 ; 1 4

1/ 36 ; 5 8

i

i

w i

i

(4)

The lattice velocities ie in the above equations are given

by Eq. (5) and are also shown in Fig.1.

0 ; i=0

1 1(cos ,sin ) ; 1 i 4

2 2

2 9 2 92(cos ,sin ) ; 5 i 8

4 4

i i

i i

ie (5)

2 Copyright © 2015 by ASME

Page 3: Droplet Formation in a T-Junction Microfluidic Device in ...azeeshan/files/InterPACKICNMM2015...Supreet Singh Bahga Department of Mechanical Engineering ... examined by simulations

FIGURE 1: LATTICE ARRANGEMENT FOR D2Q9

To sharpen the interface, minimize spurious velocities and

eliminate lattice-pinning, the post-collision distributions of the

two color fluids, red and blue are computed as [14]:

,R R B

i R i i

R B R B

f f w

ie n (6)

,B R B

i B i i

R B R B

f f w

ie n (7)

where is the anti-diffusion parameter , with a value of 0.7.

This keeps spurious velocities fairly low and maintains a sharp

interface between the two phases [20].

Implementation of the leaky dielectric model with diffuse

interface

The electric forces experienced by fluid particles can be

written in terms of E, the electric field, , the local fluid

permittivity and q, the free charge density. Electrostriction

forces have been neglected since the fluid is considered to be

incompressible.

21

2q EF E E (8)

The first term corresponds to the coulombic force while the

second one corresponds to dielectrophorectic force. The free

charge density q is given by:

q E (9)

Electric field is given as:

U E (10)

The charge conservation is governed by the electrostatic law:

.( )=0U (11)

To solve this equation using LBM, particle distributions hi are

defined such that:

i

i

U h (12)

The evolution of these particles is described by the following

equation:

( )( , ) ( , )

eq

i ii i

h

h hh t t h t

x x (13)

where relaxation time is given as:

3 0.5h (14)

and the equilibrium distributions eq

ih are given as:

4 ; i=0

9

1 ; 1 i 4

9

1 ; 5 i 8

36

eq

i

U

h U

U

(15)

To obtain smooth functions for permittivity and

conductivity, the mixture properties have been assumed to

follow [10]:

R R B B

R B

(16)

R R B B

R B

(17)

In the collision step of f particle distributions, the source

terms due to both the interfacial and electric forces are

incorporated:

interfacial electric

i i i (18)

RESULTS AND DISCUSSION

Model Validation

The validity of the multiphase model has been established

by conducting the static droplet test. Two-dimensional

simulations were run with the uniform initial pressure

throughout a 101 101 periodic domain as the initial condition

for different initial radii of the bubble. The relation between

pressure inside and outside the bubble is given by Laplace’s

Law:

0

in outp p pr

(19)

3 Copyright © 2015 by ASME

Page 4: Droplet Formation in a T-Junction Microfluidic Device in ...azeeshan/files/InterPACKICNMM2015...Supreet Singh Bahga Department of Mechanical Engineering ... examined by simulations

Thus a linear dependence of p with 1/ or will indicate the

model correctly incorporates the physics involved at the

interface. The same was observed in the present LBM

simulations. The plot of p vs. 1/ or obtained from the

simulations is shown in Fig. 2.

0 0.05 0.1 0.15 0.2 0.250

0.5

1

1.5

2

2.5x 10

-3

1/ro

p

Analytical

LBM

FIGURE 2: DIFFERENCE IN THE INSIDE AND OUTSIDE

PRESSURE p (LB units) FOR STATIC DROPLET TEST

WITH =0.01 FOR DIFFERENT DROPLET RADII or (LB

units). THE LINEAR VARIATION SHOWS A CONSTANT VALUE OF INTERFACIAL TENSION GIVEN BY THE SLOPE OF THE GRAPH

In the presence of an electric field, the droplet shape

changes to either prolate spheroid or oblate spheroid and, for

small values, the deformation D is governed by [18]:

2 2

0

2

( 1 3 )

3 (1 )

inR R S rL BD

L B S R

E (20)

The sign of the discriminant 2 1 3d R R S determines

the final shape of the drop. To validate the electrohydrodynamic

model, droplet deformations obtained using Eq. (20) were

compared with those obtained from LBM simulations. These

simulations were conducted for a droplet in presence of an

electric field on a domain of size 128x128. For fluid particles,

periodic boundary conditions were implemented. The non-

dimensional density of the two fluids was set to 1.0. A uniform

potential was imposed at the right and left boundaries (Fig. 3)

and linearly varying potential from the right to left at the top

and bottom boundaries. At the boundary nodes, three out of the

nine particle distributions ih will be unknown. These have been

calculated using the fact that the sum of particle distributions

along opposite directions is equal to the sum of their

equilibrium distributions [17]. For example, for the left

boundary, the unknown distributions 1 5 8, and h h h can be

calculated using:

, , ,

eq eq

i unknown i opposite i i oppositeh h h h (21)

For d>0, the drop deforms into a prolate spheroid with

lengthening of the axis along the electric field and shortening of

the axis perpendicular to it. For d<0, the droplet axis

perpendicular to the electric field lengthens.

FIGURE 3: PROLATE SHAPE ATTAINED BY DROP IN THE PRESENCE OF AN ELECTRIC FIELD

Variation of Deformation with different parameters

Figure 4 shows the variation of deformation D with E,

,or and in . The deformation from simulations was

calculated using its definition (Eq. (20)). L and B (Fig. 3) were

calculated by finding the coordinates of points on the vertical

and horizontal axis where the color function C becomes zero.

The variation is nearly linear with E2, 0 ,r 1/ and in as

suggested by Eq. (20). When D<<1, the deformation from our

simulations closely mimics the theoretical predictions. However

at higher D, LBM predicts a larger deformation compared to

theory since Eq. (20) was derived under the assumption of small

deformation and hence is only valid for small deformations.

This deviation has also been observed in earlier studies [10, 17,

22]. The difference in theoretical and LBM predicted

deformation at small radii, however, is due to error in finding

the exact value of L and B. Since the coordinates of the point

where color function C exactly becomes zero are determined by

linear interpolation between adjacent grid values, the

percentage error in calculating L and B is larger when the radius

is small.

4 Copyright © 2015 by ASME

Page 5: Droplet Formation in a T-Junction Microfluidic Device in ...azeeshan/files/InterPACKICNMM2015...Supreet Singh Bahga Department of Mechanical Engineering ... examined by simulations

0 0.02 0.04 0.06 0.08 0.1 0.12 0.14 0.16 0.18 0.20

0.02

0.04

0.06

0.08

0.1

0.12

0.14

0.16

0.18

0.2

E2

D

Theoretical

LBM

0 5 10 15 20 25 300

0.01

0.02

0.03

0.04

0.05

0.06

0.07

ro

D

Theoretical

LBM

0 50 100 150 200 2500

0.02

0.04

0.06

0.08

0.1

0.12

0.14

1/

D

Theoretical

LBM

0 0.005 0.01 0.015 0.02 0.025 0.03 0.035 0.04 0.045 0.050

0.02

0.04

0.06

0.08

0.1

0.12

0.14

in

D

Theoretical

LBM

FIGURE 4: VARIATION OF DROPLET DEFORMATION D WITH (a) ELECTRIC FIELD |E|, (b) DROPLET RADIUS or , (c)

INTERFACIAL TENSION and (d) INNER FLUID PERMITTIVITY in (ALL PROPERTIES IN LB UNITS) - THEORETICAL [18] &

USING LBM SIMULATIONS. ( 2, 25, | | 0.235, 0.02, 0.01 )o inR S r E UNLESS THE PARAMETER IS VARIED

Induced Velocity Field

Since tangential stresses cannot be supported by the fluid

interface, the tangential electric stresses at the drop interface set

up a velocity field in order to balance them with viscous

stresses. The theoretical flow field equations lead to the

following solution:

3

,

3

,

3

,out

3

,out

cos 2

2 sin 2

1/ 1/ cos 2

(1/ )sin 2

r in

in

r

u r r

u r r

u r r

u r

(22)

Here the radial distance r is expressed in units of ro and

velocities are in units of characteristic velocity u*:

2

*

2

02 (1 )

in o

in ut

rR Su

S R

E (23)

The nature of the velocity field depends on the sign of R-S

(Eq. (23)). If R>S, the flow enters from a direction

perpendicular to the electric field and is expelled along the

equatorial axis. If R<S, the flow occurs in the reverse direction.

The radial and azimuthal components of the velocity along the

equator are shown in Fig. 5(a). A distinct feature of the velocity

field is that the azimuthal component of velocity at equatorial

axis is identically zero. Clearly the velocities obtained by LBM

simulation and the theoretical predictions are in good agreement

with each other inside the drop ( r < or ) under the low

deformation limit. Outside the drop, the LBM predicted

5 Copyright © 2015 by ASME

Page 6: Droplet Formation in a T-Junction Microfluidic Device in ...azeeshan/files/InterPACKICNMM2015...Supreet Singh Bahga Department of Mechanical Engineering ... examined by simulations

velocities deviate from the theoretical values since the

theoretical analysis assumes an infinite domain with no

boundaries for the outer fluid whereas in LBM simulation, the

outer fluid is bounded (max 03r r in LBM simulation as shown

in Fig. 5). The velocities in LBM simulation vanish to zero at

the boundaries but the theoretical velocities do not.

From the theoretical flow field equations (Eq. (22)), it is

clear that at the interface (or r ), the radial velocity vanishes.

The flow velocities at the interface were computed in LBM

simulation by noting that at the interface the color function

C=0. These velocities are compared with the theoretical values

in Fig. 5(b). The LBM predictions are fairly close to the

theoretical values and the LBM predicted radial velocity is very

close to zero.

-4 -3 -2 -1 0 1 2 3 4-4

-3

-2

-1

0

1

2

3

4x 10

-4

r/ro

u

ur

u

ur (theoretical)

FIGURE 5(a): FLOW FIELD ALONG EQUATOR COMPARED WITH THEORETICAL

0 20 40 60 80 100 120 140 160 180-1

-0.8

-0.6

-0.4

-0.2

0

0.2

0.4

0.6

0.8

1x 10

-3

u

ur

u

u (theoretical)

FIGURE 5(b): FLOW FIELD ALONG THE INTERFACE COMPARED WITH THEORETICAL

Flow in T-junction microfluidic device

The developed model was used to simulate droplet

formation in a T-junction in the presence of an electric field.

The two fluids used were silicone oil and water. Electrodes

positioned at specific locations, as shown in Fig. 6, in the T-

junction set up the electric field required. The boundary

conditions used were constant potential at the electrodes and

vanishing normal component of electric field at the walls of T-

junction (i.e. E n 0 ). A constant pressure was maintained at

the exit, with fully developed velocity profiles at the inlets.

FIGURE 6: SIDE VIEW OF A T-JUNCTION

To convert physical quantities such as T-junction width

( 50 m ), interfacial tension ( 0.045 N/m ) and density

(31000 kg/m ) to LB units, first a set of reference quantities

were chosen to non-dimensionalize them. The width of main

channel was taken as reference length, the density of water as

reference density and outlet pressure as reference pressure.

These non-dimensional quantities were then converted to LB

units by discretizing the reference length and the reference time

[21] such that the maximum Mach number<0.3. Simulations

were carried out for Ca=0.01, Re=0.225 at three different flow

rate ratios Q=1/5, 1/10 and 1/20. For each case, droplet size

and frequency of generation were recorded both with and

without the electric field. The droplet size is simply the distance

between the two ends of the droplet when it is in the main

channel. The frequency of generation was measured by

observing the change in color function C at a point near the exit

of the T-junction over time.

Figure 7 and 8 show the variation of droplet size and

formation frequency with applied electric field at different flow

rate ratios Q.

6 Copyright © 2015 by ASME

Page 7: Droplet Formation in a T-Junction Microfluidic Device in ...azeeshan/files/InterPACKICNMM2015...Supreet Singh Bahga Department of Mechanical Engineering ... examined by simulations

FIGURE 7: VARIATION OF DROPLET SIZE WITH APPLIED ELECTRIC FIELD (LB UNITS) AT DIFFERENT FLOW RATE RATIOS

FIGURE 8: VARIATION OF FREQUENCY OF DROPLET FORMATION WITH ELECTRIC FIELD (LB UNITS) AT DIFFERENT FLOW RATE RATIOS

On applying an electric field, generally a reduction in

droplet size and increase in formation frequency was observed.

Reduction in the size of droplets and increase in frequency of

formation is due to an early pinch-off of the dispersed fluid in

the presence of electric field. As the dispersed fluid approaches

the point of confluence with the continuous phase in the main

channel, the applied electric field causes polarization at the

interface leading to coulombic and dielectrophorectic forces on

the interface. Since water (dispersed phase) is attracted to the

high electric field region [15], it moves towards the edge of the

upper electrode (Fig. 6) where electric field is maximum. As a

result, the neck of the dispersed fluid gets thinned, assisting in

an early pinch off (Fig. 10(b)). Figures 10(a) and (b) show the

difference in the mechanism of droplet formation in a T-junction

channel in the absence and presence of electric field. Without

an electric field, the droplet formation is due to the effect of

pressure forces exerted by the continous phase on the dispersed

phase alone (Fig. 10(a)). In the presence of electric field,

electric forces also help in the breaking up of the fluid thereby

forming droplets.

From Fig. 7 (resp. 8), it can be seen that there is a local

minima (maxima) in the plot at an intermediate value of electric

field (|E|~0.1). This is explained as follows. With increase in

electric field above a threshold value, the dispersed phase starts

getting pinned to the electrodes as it enters the main channel

(before breakup). On comparing figures 9(i) and (iii), it can be

seen that the pinning at the upper electrode results in

accumulation of more volume of dispersed phase in the main

channel before droplet formation. The same pinning effect

causes a decrease in frequency of droplet formation near the

threshold electric field. However, this deviation in behavior is

only noticed while transitioning from low electric field (no

pinning) to high electric field (when pinning begins).

i. |E|=0.1 (No pinning)

ii. |E|=0.1 (No pinning)

iii. |E|=0.2 (Pinning)

iv. |E|=0.2 (Pinning)

FIGURE 9: COMPARISION OF DROPLET DYNAMICS BEFORE AND AFTER BREAKUP NEAR THRESHOLD ELECTRIC FIELD (THE ARROWS SHOW ELECTRIC FIELD VECTORS)

7 Copyright © 2015 by ASME

Page 8: Droplet Formation in a T-Junction Microfluidic Device in ...azeeshan/files/InterPACKICNMM2015...Supreet Singh Bahga Department of Mechanical Engineering ... examined by simulations

i

ii

iii

iv

v

vi

vii

FIGURE 10(a): STAGES OF THE DROPLET FORMATION MECHANISM IN A T-JUNCTION WITHOUT ELECTRIC FIELD

i

ii

iii

iv

v

vi

vii

FIGURE 10(b): STAGES OF THE DROPLET FORMATION MECHANISM IN A T-JUNCTION IN THE PRESENCE OF AN ELECTRIC FIELD

CONCLUSIONS

A lattice Boltzmann fluid dynamics framework has been

developed which incorporates the effects of both

electrohydrodynamic and interfacial forces. This model has

been validated by comparing the results obtained for

benchmarking problems like static droplet test and a leaky

dielectric drop in an electric field. This was then successfully

employed to the problem of interest.

Results show that it is possible to increase droplet

formation frequency and reduce its size in a T-junction by

applying an electric field. This arms us with tools to achieve

better control over the droplet parameters without changing the

flow rate ratio, capillary number and geometry of the T-

junction.

8 Copyright © 2015 by ASME

Page 9: Droplet Formation in a T-Junction Microfluidic Device in ...azeeshan/files/InterPACKICNMM2015...Supreet Singh Bahga Department of Mechanical Engineering ... examined by simulations

ACKNOWLEDGMENT

The authors are grateful to Mr. Nipun Arora for providing

assistance with the numerical framework.

REFERENCES

[1] Link, D.R., Grasland-Mongrain, E., Duri, A., Sarrazin, F.,

Cheng, Z., Cristobal, G., Marquez, M., and Weitz, D.A.,

2006, “Electric Control of Droplets in Microfluidic

Devices,” Angewandte Chemie International Edition, Vol.

45, pp. 2556–2560.

[2] Huebner, A., Sharma, S., Srisa-Art, M., Hollfelder, F., Edel,

J.B., and deMello, A.J., 2008, “Microdroplets: A sea of

applications?” Lab on a Chip, Vol. 8, pp. 1244-1254.

[3] Saville, D.A., 1997, “Electrohydrodynamics: The Taylor-

Melcher Leaky Dielectric Model,” Annual Reviews of

Fluid Mechanics, Vol. 29, pp. 27-64.

[4] Melcher, J.R., and Taylor, G.I., 1969,

“Electrohydrodynamics: A Review of the Role of

Interfacial Shear Stresses,” Annual Reviews of Fluid

Mechanics, Vol. 1, pp. 111-146.

[5] Taylor, G.I., 1964, “Disintegration of Water Drops in an

Electric Field,” Proceedings of the Royal Society of

London, Series A, Mathematical and Physical Sciences,

Vol. 280, No. 1382, pp. 383-397.

[6] Cheng, K.J., and Chaddock, J.B., 1984, “Deformation and

stability of drops and bubbles in an electric field,” Physics

Letters, Vol. 106A, pp. 51-53.

[7] Cheng, K.J., 1985, “Capillary Oscillations of a drop in an

electric field,” Physics Letters, Vol. 112A(8), pp. 392-396.

[8] Herman, C. and Iacona, E., 2004, “Modeling of bubble

detachment in reduced gravity under the influence of

electric fields and experimental verification,” Heat and

Mass Transfer, Vol. 40, pp. 943-957.

[9] Fernandez, A., Tryggvason, G., Che, J. and Ceccio, S.L.,

2005, “The effects of electrostatic forces on the distribution

of drops in a channel flow: Two-dimensional oblate drops,”

Physics of Fluids, Vol. 17, 093302.

[10] Zhang, J., and Kwok, D.Y., 2005, “A 2D lattice Boltzmann

study on electrohydrodynamics drop deformation with the

leaky dielectric theory,” Journal of Computational Physics,

Vol. 206, pp. 150-161.

[11] Tomar, G., Gerlach, D., Biswas, G., Alleborn, N., Sharma,

A., Durst, F., Welch, S.W.J., and Delgado, A., 2007, “Two-

phase electrohydrodynamic simulations using a volume-of-

fluid approach,” Journal of Computational Physics, Vol.

227, pp. 1267-1285.

[12] He, X., and Li, N., 2000, “Lattice Boltzmann simulation of

electrochemical systems,” Computer Physics

Communications, Vol. 129, pp. 158-166.

[13] Gupta, A., and Kumar, R., 2010, “Flow regime transition at

high capillary numbers in a microfluidic T-junction:

Viscosity contrast and geometry effect,” Physics of Fluids,

Vol. 22, 122001.

[14] Lishchuk, S.V., Care, C.M. and Halliday, I., 2003, “Lattice

Boltzmann algorithm for surface tension with greatly

reduced microcurrents,” Physical Review E, Vol. 67,

03670.

[15] Wang, W., Yang, C., and Li, C.M., 2009, “Efficient On-

Demand Compound Droplet Formation: From

Microfluidics to Microdroplets as Miniaturized

Laboratories,” Small, Vol. 5, No. 10, pp. 1149-1152.

[16] Gong, S., Cheng, P. and Quan, X., 2010, “Lattice

Boltzmann simulation of droplet formation in

microchannels under an electric field,” International

Journal of Heat and Mass Transfer, Vol. 53, pp. 5863–

5870.

[17] Ha, J.-W., and Yang, S.-M., 2000, “Deformation and

breakup of Newtonian and non-Newtoian conducting drops

in an electric field,” Journal of Fluid Mechanics, Vol. 405,

pp. 131-156.

[18] Feng, J.Q., 2002, “A 2D Electrohydrodynamic Model for

Electrorotation of Fluid Drops,” Journal of Colloid

Interface Science, Vol. 246, pp. 112-120.

[19] Gupta, A., Matharoo, H.S., Makkar. D., and Kumar, R,

2014, “Droplet formation via squeezing mechanism in a

microfluidic flow-focusing device,” Computers and Fluids,

Vol. 100, pp. 218-226.

[20] Gupta, A., and Kumar, R., 2010, “Effect of geometry on

droplet formation in the squeezing regime in a microfluidic

T-junction”, Microfluidics and Nanofluidics, Vol. 8, pp.

799-812.

[21] Latt, J., 2008, “Choice of units in lattice Boltzmann

simulations,” wiki.palabos.org/_media/howtos:lbunits.pdf

(Accessed Jan. 2015).

[22] Feng, J.Q. and Scott, T.C., 1996, “A computational analysis

of a leaky dielectric drop in an electric field,” Journal of

Fluid Mechanics, Vol. 311, pp. 289-326.

9 Copyright © 2015 by ASME