You are on page 1of 29

Numerical study of deformation and breakup of a

multi-core compound droplet in simple shear flow

Tri-Vien Vua, Truong V. Vu*,b, Dang Thanh Buic


a
Modeling Evolutionary Algorithms Simulation and Artificial Intelligence, Faculty of
Electrical & Electronics Engineering, Ton Duc Thang University, Ho Chi Minh City, Vietnam
Email: vutrivien@tdtu.edu.vn
b
School of Transportation Engineering, Hanoi University of Science and Technology, 01 Dai
Co Viet, Hai Ba Trung, Hanoi, Vietnam
Email: vuvantruong.pfae@gmail.com;
*
Corresponding author
Tel.: +84-24-3869-2984; Fax: +84-4-3868-4945
c
Institute for Control Engineering and Automation, Hanoi University of Science and
Technology, 01 Dai Co Viet, Hai Ba Trung, Hanoi, Vietnam

ABSTRACT
In this paper, we present the numerical simulation results of deformation and breakup of a

multi-core compound droplet. The method used is a two-dimensional front-tracking method

with a modification for multi-layer droplets. The droplet is placed at the center of the domain

whose top and bottom boundaries move in the opposite directions to create the shear flow. We

vary the values of various parameters in specific ranges to investigate their effects on the

deformation and breakup of the multi-core compound droplet in shear flow. We mainly focus

on the two-core compound droplets. Basing on the numerical results, we reveal various

patterns (i.e. modes) of deformation and breakup: non-breakup types 1 and 2, breakup types 1,

2, 3 and 4. For the first non-breakup type (type 1), the outer interface is deformed with two

inner droplets accumulating at the center whereas in the other non-breakup type (type 2), the

inner droplets move to the two furthest ends. The non-breakup type 1 generally changes to the

breakup type 4, at some critical parameters, in which the compound droplet breaks up into

simple droplets at its ends with a smaller daughter compound droplet in between. The
1
non-breakup type 2 transits to the breakup type 1 or type 2, depending on the flow conditions,

with the formation of two smaller daughter compound droplets at the ends. The remaining

breakup mode (type 3) is a mixed breakup mode that is a combination of the breakup types 2

and 4. These deformation and breakup patterns are mapped onto a Re–Ca diagram. We also

propose some other diagrams based on the variations of the inner droplet location. In addition,

we investigate the effect of the number of the inner droplets encapsulated in the compound

droplet on its deformation and breakup.

Keywords: Front-tracking; multi-core compound droplet; deformation; breakup; numerical


simulation

1. Introduction
Compound liquid droplets exist in various industrial applications such as food processing

[1], cosmetics, drug delivery, biotechnology [2,3], lab on chip systems [4] and so on. The

compound droplets consist of an outer interface encapsulating one, i.e. single-core compound

droplet, or many inner droplets, i.e. multi-core compound droplet [5]. To produce such

droplets, we can use the liquid jet breakup [1,6,7] or membrane homogenizers [2]. Generally,

these compound droplets move in an externally continuous phase flow, and thus they

experience deformation and might break up into smaller droplets. Accordingly, the

deformation and breakup (i.e. rheological behavior) of the compound droplet plays an

important role, and understanding its mechanism is a key to control and advance the

above-mentioned applications. Accordingly, the dynamic behavior of the compound droplet

has been drawn much attention recently.

Concerning single-core compound droplets, Chen et al. [8] experimentally investigated

the compound droplets moving through an orifice of a nozzle. The nozzle was tapered to

induce the droplet breakup. A phase diagram of the Capillary number Ca versus
2
droplet-to-orifice diameter ratio was also proposed. In another work, Li et al. [9] also

considered the breakup of the compound droplets flowing through a tapered nozzle, but with

focusing on the effect of the tilt angle of the nozzle. The authors found that the nozzle with a

tilt angle of 90 induced no breakup. Numerically, Kan et al. [10] used a combined

Eulerian-Lagrangian method to study the deformation characteristics of a compound droplet

in a symmetrically external flow. Chen et al. [11] combined the experiments with the

numerical simulations done by a volume of fluid method to study the effect of the inner

droplet on deformation of the compound droplet under simple shear flow. In another work

[12], the authors proposed a diagram, of the Capillary number versus the radius ratio, of two

types of transient deformation: “over damped” and “under damped” oscillations. Luo et al.

[13] also studied the deformation dynamics of a compound droplet under simple shear by the

front-tracking-based simulations. The authors focused on the effect of the radius ratio of the

inner to outer interfaces with various Capillary numbers. Hua et al. [14] used two- and

three-dimensional immersed boundary methods to investigate the finite deformation of the

compound droplet with attention to the effects of the size ratio of the inner to outer interfaces,

the interfacial tension ratio, and the inner droplet location. However, the results in the

above-mentioned studies were limited to the non-breakup regime.

Considering only the breakup regions, Chen et al. [15] used volume of fluid-based

simulations to reveal various patterns (or modes) of the compound droplet breakup in shear

flow: necking, end pinching, and capillary instability. These modes were recognized in the

phase diagrams based on the Capillary number Ca, dynamic viscosity ratios and radius ratio.

Recently, we used a front-tracking method to study both deformation and breakup of the

compound droplet in shear flow under the effects of Ca and the Reynolds number Re. We

proposed a diagram based on the Capillary number and the Reynolds number, on which the

transition from the non-breakup to breakup regimes appears.

However, in the above-mentioned investigations, the authors considered only the


3
single-core compound droplets (i.e. double compound droplets or double emulsions). Not

limited to single-core compound droplets, Wang et al. [16–18] developed a boundary integral

method to investigate the rheological behavior of a two-dimensional (2D) compound droplet

with one and multiple cores encapsulated. The authors considered the effect of the number

and asymmetrically internal structure of the inner droplets on the deformation of the outer

interface that was exerted to external flows [16]. In another work of Wang and coworkers [19],

the authors studied the effect of the asymmetrically internal structure of the inner droplets on

their shift within the outermost interface. The authors also proposed three phase diagrams

based on the structural asymmetric parameter, Ca and viscosity ratio. Recently, Xu et al. [20]

used the same numerical method reported in [16] to investigate the oriented shift and inverse

of the compound droplets with two different-sized inner droplets. The authors found a critical

Capillary number beyond which the compound droplet moves to the side of the larger inner

droplet, otherwise it moves to the side of the other inner droplet. However, in these studies the

authors have not considered the breakup region of the multi-compound droplet.

It is evident that the former works considered both the deformation (i.e., non-breakup) and

the breakup only for single-core compound droplets. For multi-core compound droplets,

detailed numerical simulations have not been found in the literature. None of the

above-mentioned works considered the transition from non-breakup to breakup of the

compound droplet with two or more inner droplets. Thus, the effects of some parameters on

the transition and breakup modes have not been done. Filling out these missing gaps is the

main purpose of the present study. In this study, we use a 2D front-tracking/finite difference

method [21,22] to investigate the deformation and breakup of a multi-core compound droplet

exerted to simple shear flow. We focus on the effect of various parameters including Ca, Re,

the radius ratio, the interfacial tension ratio, and so on. We also propose some phase diagrams

based on some investigated parameters.

Even though it is well-known that multi-core droplets are three-dimensional (3D), and 3D
4
effects should be considered in the study [23], to our knowledge the present study is a

pioneering work providing a valuable picture of both deformation and breakup of multicore

compound droplets in simple shear flow.

2. Mathematical formulation and numerical method


Fig. 1 shows the configuration of the investigated problem, in which a 2D compound

droplet with two inner droplets is immersed at the center of the computational domain. The

top and bottom move in the opposite directions with a velocity magnitude U to create a shear

rate  = 2U H where H is the domain height. We assume that initially the inner and outer

interfaces are circular. To simplify the problem, unless stated otherwise, we assume that two

inner droplets are identical and located symmetrically to the center of the outer droplet. The

distance from the center of the first inner droplet (i.e. inner droplet 1) to the center of the outer

droplet is , and this inner droplet is located at an angle , as shown in Fig. 1. The radii of the

inner and outer droplets are respectively R1 and R2. The inner, middle and outer fluids are

denoted by 1, 2 and 3, respectively. In terms of one-fluid representation, the governing

equations [21] for immiscible, incompressible and Newtonian fluids give

Fig. 1. A two-core compound droplet in shear flow.


5
 u  u 2  uv p  u   u v 
+ + = − + 2 +   +  +  g x + Fx (1)
t x y x x x y  y x 
 v  uv  v 2 p   u v   v
+ + = − +   +  + 2 +  g y + Fy (2)
t x y y x  y x  y y
D  D
= =0 (3)
Dt Dt
Here, u = (u,v) and p are the velocity vector and the pressure, respectively.  and  are the

interfacial tension coefficient and curvature, respectively. The fluid properties including

density  and viscosity  are assumed constant in each fluid. In this study, we neglect the

effects of gravity, and the gravitational acceleration g = (gx,gy) is thus zero. F = (Fx,Fy) is the

interfacial tension force acting on the interface separating different fluids [21,24]

 ( )
F =  x − x f n f dS (4)
f

where δ(x − xf) is the Dirac delta function that is zero everywhere except for a unit impulse at

the interface xf with f denoting interface.  and  are the interfacial tension coefficient and

curvature, respectively. nf is the unit normal to the interface.

To solve the above-mentioned governing equations, we use a front-tracking/finite

difference method [21,25]. The interface between two different fluids is represented by a

chain of straight lines (elements) in which the length of each element is in the range of

(0.2–0.8)h. Here, h is the grid spacing of the fixed, rectangular grid on which the front

element coordinates are updated by

xn+1 = xn + Vn t (5)

Here, n and n+1 are the current and next time levels. Vn is the front velocity interpolated

from the fixed grid velocities that are solved from the momentum equations by approximating

the spatial derivatives by second ordered central difference and by using the

predictor-corrector scheme for the time integration on the staggered grid. The value of the

material properties  and , indicated by , at every location in the domain is given as

 = 1Ii + 2 (1 − Ii ) I o + 3 I o (6)


6
where Ii and Io are the indicator functions reconstructed from the location of the inner and

outer interfaces respectively:

1, within the inner interfaces


Ii = 
0, other
(7)
1, within the outer interface
Io = 
0, other

This calculation technique for the fluid property fields has been widely used in our previous

works for three phase computations (e.g. [26,27]). Fig. 2 shows the distribution of the density

field, whose value is calculated from Eq. (6), along the horizontal line through the center of

the outer droplet with 1/3 = 3 and 2/3 = 2. The figure confirms the accuracy of the

technique for the calculation of the fluid properties of multi-core compound droplets. The

boundary conditions are shown in Fig. 1.

Concerning the topology change during the deformation, breakup or merging might

appear. Numerically, breakup occurs as the distance between two certain elements on the

same front less than one half of a grid space [22,28]. However, unlike breakup, merging

requires more physical aspects [28], and in some cases two droplets of the same fluid do not

merge even though they are very close to each other [6,7]. In this study, we simply assume no

merging in accordance with the experimental observations [6] and our previous study [7]. In

Fig. 2. A double-core compound droplet with (a) front location and (b) variation of the density
field along the dash line in (a) with C3(xc3, yc3) - the center of the outer droplet.
7
addition, it is worth noting that the breakup process of droplets frequently happens along with

capillary instability that would affect the results [29].

3. Numerical parameters, grid refinement and method validation


Basing on the properties of the outer fluid and R2, we have the following dimensionless

parameters governing the dynamics of the problem:

3 R22 3 R2 R   2


Re = , Ca = , R12 = 1 ,  12 = 1 ,  0 = , 0 = , (8)
3 2 R2 2 R2 t =0
 t =0

1 2 1 2
13 = , 23 = , 13 = , 23 = (9)
3 3 3 3
where Re and Ca are respectively the Reynolds and Capillary numbers. R12 and 12 are the

radius and interfacial tension ratios of the inner and outer interfaces. 0 and 0 is the

normalized distance and angle of the inner droplet 1 at t = 0. 13 and 23 are the density ratios.

13 and 23 are the viscosity ratios. The dimensionless time is  = t.

In the following figures, C1 (xC1,yC1), C2 (xC2,yC2) and C3 (xC3,yC3) are the center of mass

of the first and second inner droplets and the outer droplet. We also use C(xC,yC), as an

alternative for C3, defined as

xC = 0.5 ( xC1 + xC 2 ) ; yC = 0.5 ( yC1 + yC 2 ) (10)

Fig. 3. Grid refinement study with Re = 1.0, Ca = 0.5, R12 = 0.3, 12 = 13 =23 =13 =23
=1.0, 0 = 0.5, and 0 = 1.0: (a) temporal variation of the deformation parameters of the inner
and outer droplets, and (b) and (c) droplet profiles at  = 14.4 and 36 (nearly steady state).
8
Accordingly, we can define the angle  of the first inner droplet during the droplet

deformation and breakup as the angle in between CC1 and the x axis. We also define L as the

distance between the inner droplets for the two-core compound droplets.

The height H of the computational domain is chosen as 5R2 while the width W of the

domain depends on the flow conditions. But the typical domain is WH = (205)R2. Basing

on this domain, we perform a grid refinement study with Re = 1.0, Ca = 0.5, R12 = 0.3, 12 =

13 =23 =13 =23 =1.0, 0 = 0.5, and 0 = 1.0. Three grid resolutions 256064, 512128 and

1024256 were used. Fig. 3a shows the deformation parameters of the outer and inner

droplets yielded from three grid resolutions. The results show that the results from the two

finer grids are in almost agreement why the coarsest grid results in some difference. Fig. 3b

compares the compound droplet profiles at middle way through deformation and at nearly

final state of deformation with complete agreement. Accordingly, for the rest of the

computations presented in this paper, we use 128 grid points in the y direction while the

number of grid points in the x direction is varied according to the width of the domain, e.g.

512128 for WH = (205)R2.

To validate the method used in the present study, we perform a simulation of the

deformation of a single-core compound droplet in shear flow and compared the results with

Fig. 4. Comparison of the computational results (solid lines) with the numerical predictions of
Hua et al. [14] (symbols) for the deformation of a single-core compound droplet in shear flow
at  = 10. R1 and R2 are respectively the inner and outer droplet radii, and (xc, yc) are the
coordinates of the droplet center. Other parameters are shown in the text.
9
other numerical predictions [14], as shown in Fig. 4. To perform this simulation, we just

remove one core, and place the remaining core at the center of the outer droplet. We then set

the computational domain with the dimensions and parameters same as that of Hua et al. [14].

Three cases of R12 = 0.3, 0.5 and 0.7 with Re = 1.0, Ca = 0.25 are computed. The other

parameters are set to unity. Fig. 4 shows that our results are in good agreement with that of

Hua et al. [14], supporting the accuracy of the numerical method used in this study. Other

validation can be found in our previous work [22].

Fig. 5. Non-breakup type 1: Re = 1.0, Ca = 0.4, R12 = 0.3, 0 = 0.5, 12 = 13 =23 =13 =23
=1.0, and 0 = 1.0. (a) Temporal deformation of a double-core compound droplet to the steady
state. (b) Temporal variation of the deformation parameter To of the outer droplet and the
angle of the first inner droplet. (c) Temporal variation of the distance L between the inner
droplets. The circles in (c) indicates the moments shown in (a). The velocity in (a) is
normalized by Uc.
10
4. Results and discussion
Fig. 5a shows the deformation of the compound droplet with Re = 1.0, Ca = 0.4, R12 = 0.3,

12 = 13 =23 =13 =23 =1.0, 0 = 0.5, and 0 = 1.0 at different moments of deformation. The

moments correspond to the circles indicated in Fig. 5c. The compound droplet with an initial

circular outer interface (i.e. circular outer droplet) is then deformed because of the shear flow

induced in the outer liquid. This results in a clockwise circulation flow, within the outer

interface, that drives the inner droplets at  = 1.8 (Fig. 5a). At this time, the deformation

parameter of the outer interface To (Fig. 5b) is increasing, and no circulations appear within

each inner droplet. Accordingly, the inner droplets move away from each other, resulting in an

increase in L (the distance between the inner droplets, Fig. 5c) and a decrease in the inner

angle  of the first inner droplet (Fig. 5b). At  = 10.8 (Fig. 5a), the inner droplets are moving

closer to each other (Fig. 5c), and a clockwise circulation presents in each inner droplet.

While the inner droplets are still moving, the outer interface reaches almost the steady state

shape (To in Fig. 5b). The inner droplets then again move away from each other at a later time

 = 23.4 (Fig. 5), resulting an increase in L (Fig. 5c). However, the inner angle changes

slightly because of the minor change in the shape of the outer interface. As time progresses,

the inner droplets just move slightly within the outer droplet ( = 36 and 64.8 in Fig. 5a), and

no breakup occurs at later time. This type of non-breakup is called “non-breakup type 1” with

Fig. 6. (a) Steady state shape of a compound droplet at  =140 (i.e. non-breakup type 2). (b)
Temporal variations of the distance L between the inner droplets and of the inner droplet angle
. Re = 0.8, Ca = 0.8, R12 = 0.3, 0 = 0.5, 12 = 13 =23 =13 =23 =1.0, and 0 = 1.0.
11
the two inner droplets almost accumulating at the center.

Non-breakup also occurs for a case shown in Fig. 6 with Re = 0.8, Ca = 0.8. However, in

this case, the outer interface is much more stretched as compared with the case shown in Fig.

5, and each inner droplet moves to each opposite end in accordance with the shear flow,

forming a bone-like shape. This causes the inner angle to reach nearly zero (Fig. 6b). We call

this non-breakup mode “non-breakup type 2”. We also observe from our numerical

simulations that the droplet interface itself rotates, and combines with its stretching results in

an oscillation [29].

Unlike the case of non-breakup shown in Fig. 6, in the case of breakup shown in Fig. 7 the

inner droplets move far away from each other (see the droplet profile at  = 14.4 in Fig. 7a

with the distance between two droplets in Fig. 7c) with a higher deformation parameter of the

Fig. 7. Breakup type 2 with Re = 1.0, Ca = 1.0, R12 = 0.3, 0 = 0.5, 12 = 13 =23 =13 =23
=1.0, and 0 = 1.0. (a) Temporal deformation and breakup of a double-core compound
droplet. (b) Temporal variation of the deformation parameter To of the outer droplet and the
angle of the first inner droplet. (c) Temporal variation of the distance between the inner
droplets. The circles in (c) indicates the moments shown in (a).
12
outer droplet as compared to that in Fig. 5. After reaching the maximum distance, the inner

droplets move back to a distance that, however, is much longer than the initial one ( = 46.8,

Fig. 7a). Because of the interfacial tension force, the outer interface performs necking at its

ends ( = 266.4, Fig. 7a) and then breaks up into two smaller single-core compound droplets

with a remaining elongated simple filament in between. Fig. 7b indicates that during breaking

up, the angle of the first inner droplet keeps unchanged with a very slight increase in the

distance between two inner droplets. After breakup, two daughter single-core compound

droplets again move away from each other to reach their new positions, because of the shear

flow, which lie almost on the horizontal line through the domain center. Each compound

droplet oscillates a bit before reaching the final position on the horizontal center line. We call

this breakup case “compound droplet end-pinching” (or breakup type 2) since the initial

compound droplet exhibits an end-pinching with two newly formed compound droplets at its

ends and simple filaments in between them. Another breakup mode called “breakup type 1” is

also observed from our computational results in which the compound droplets are formed at

two ends but with no simple droplets in between (this breakup type 1 will be shown in a

diagram of Ca versus Re discussed later).

Next, we consider two other breakup modes as shown in Fig. 8a and b (Re = 1.0, Ca = 1.0,

R12 = 0.3, 12 = 13 =23 =13 =23 =1.0, and 0 = 1.0) in which the compound droplet breaks

up into simple droplets from its ends. In Fig. 8a (with 0 = 0.3), at the beginning stage of the

deformation, two inner droplets move away from each other to a distance of about 1.0R2.

They then move back and do not split anymore. This is caused by the outer interface breakup

into simple droplets and a daughter compound droplet consisting of these two inner droplets.

However, unlike the case shown in Fig. 5, two inner droplets rotate around the center because

of the circulation flow appearing within that compound droplet (Fig. 8a). Because of the

rotation and deformation, the distance and angle vary in a sine-function-like shape. Increasing

the distance to 0 = 0.4 results in a different breakup pattern in which there are two simple
13
droplets formed at two ends and a daughter double-core compound droplet at the center (Fig.

8b). We call these breakup modes “simple droplet end-pinching” (or breakup type 4).

Another mode of breakup observed from our numerical calculations is shown in Fig. 8c

(Re = 1.0, Ca = 1.0, R12 = 0.4, 0 = 0.5, 12 = 13 =23 =13 =23 =1.0, and 0 = 1.0) in which

the compound droplet breaks up into simple droplets with daughter compound droplets in

between. Similarly to the case shown in Fig. 8b, the compound droplet with two inner

droplets initially in a circular shape is deformed and stretched by shear flow. It first exhibits

an end-pinching mode to produce a simple droplet at each end (at  = 93.6, Fig. 8c), i.e.

breakup type 4. The remaining droplet encapsulating the inner droplets is then elongated and

necks at its two ends. At later time ( = 403.2, Fig. 8c), it breaks up into two smaller, daughter

compound droplets in a mode shown in Fig. 7a (i.e. breakup type 2). Accordingly, the

compound droplet in this situation breaks up into both compound and simple droplets, and we

thus call it “mixed pinching” or breakup type 3. It is worth to note here that even though the

Fig. 8. (a), (b): Simple droplet end-pinching modes (i.e. breakup type 4) with temporal
variations of the distance between the inner droplets and the angle of the first inner droplet.
(c) Droplet shape at two moments of breakup in a mixed pinching mode (i.e. breakup type 3).
14
problem is fully resolved by the front-tracking method (i.e. a direct numerical method)

representing the interface by connected points, the method numerically handles breakup as

previously discussed. Accordingly, the simulated results of the deformation and breakup are

numerically reproducible. However, as reported by Hinch and Acrivos [29], the capillary

instability frequently associated with the breakup process is as a source causing the

experimentally measured data to be largely scattered and difficult to be reproducible when

droplets are extendedly long in the simple shear flow. This instability would influence the

breakup patterns observed in experiments.

Next, we investigate the effects of some parameters on the deformation and breakup of the

multi-core compound droplets in the shear flow as well as on the modes of breakup.

4.1. Effect of the initial distance between the inner droplets

Fig. 9 shows the effects of the initial distance of the first inner droplet to the center of the

compound droplet on the breakup of the compound droplet with Re = 1.0, Ca = 1.0, R12 = 0.3,

12 = 13 =23 =13 =23 =1.0, and 0 = 1.0. Since the inner droplets are symmetric to the

center, the initial distance between them is 20 (Fig. 1). For 0 = 0.3, the inner droplets are

very close to each other, then the circulation flow within the outer interface makes them rotate

Fig. 9. Effect of the distance between the inner droplets in terms of 0 on the breakup of the
compound droplet: (a) droplet shape and (b) temporal variation of the distance between the
inner droplets. Re = 1.0, Ca = 1.0, R12 = 0.3, 12 = 13 =23 =13 =23 =1.0, and 0 = 1.0.
15
while attaching to each other (Fig. 9b). In the meantime, the shear flow causes the outer

interface to break up into simple droplets (Fig. 9a). Increasing the value of 0 to 0.4 pushes the

inner droplets far away from each other, thus the distance between them becomes increasing

in time. During this time, the outer interface performs breakup to form droplets at its ends (i.e.

breakup type 4). However, the shear flow is not strong enough to make the outer interface

split into compound droplets. At higher values of 0 = 0.5, 0.6 and 0.7, the inner interfaces are

close to the outer interface, the shear flow induces droplets in a compound droplet

end-pinching mode, i.e. breakup type 2. As shown in Fig. 9b, increasing the value of 0 makes

the compound droplet to sooner breaks up into smaller compound droplets (since after

breaking up the distance between the inner droplets has a sudden jump in its value).

4.2. Effect of the radius ratio

Fig. 10 depicts the shape of the compound droplet with the temporal variation of the

distance between two inner droplets for different ratios R12 of the inner droplet to the outer

droplet. The parameters for these computations are Re = 1.0, Ca = 1.0, 0 = 0.5, 12 = 13 =23

=13 =23 =1.0, and 0 = 1.0. R12 = 0.1 yields very small inner droplets, and thus the inner

droplets are easily driven by the flow within the outer interface. Accordingly, the inner

Fig. 10. Effect of the radius ratio of the inner to outer interfaces R12 the breakup of the
compound droplet with Re = 1.0, Ca = 1.0, 0 = 0.5, 12 = 13 =23 =13 =23 =1.0, and 0 =
1.0: (a) droplet shape and (b) temporal variation of the distance between the inner droplets.
16
droplets come close to each other and rotate at the center (Fig. 10b), and no breakup (i.e.

non-breakup type 1) occurs (Fig. 10a). Increasing the value of R12 to two time higher, i.e. R12

= 0.2, a contrary situation, in which each inner droplet moves to each end of the compound

droplet, occurs. However, no breakup still presents under the mode of non-breakup type 2. At

more higher values of R12, the phenomenon completely changes with the presence of breakup,

i.e. breakup type 2. It is evident that the size of the inner droplet plays an important role in the

shape of the compound droplet in the shear flow, in which an increasing in its size enhances

the compound droplet breakup. In addition, Fig. 10b indicates that the breakup occurs sooner

at higher R12.

4.3. Effect of the Capillary number Ca

The Capillary number is the ratio of the viscous force to the interfacial tension force, and

thus a low Ca (i.e. Ca = 0.2) corresponding to a high interfacial tension force that tends to

hold the droplet in a circular shape yields a slightly deformed droplet as shown in the top

frame of Fig. 11a (i.e. non-breakup type 1). Increasing Ca to a higher value, i.e. Ca = 0.6,

corresponds to decreasing the interfacial tension force, and thus the droplet is stretched with

two inner droplets at its ends. However, similarly to the case of the lower value, no breakup

Fig. 11. Effect of the Capillary number with Re = 1.0, R12 = 0.5, 0 = 0.5, 12 = 13 =23 =13
=23 =1.0, and 0 = 1.0: (a) droplet shape and (b) temporal variation of the distance between
the inner droplets.
17
(with a mode of non-breakup type 2) presents, and thus the distance between two inner

droplets keeps constant at  > 100 (Fig. 11b). At Ca = 1.2 corresponding to a very week

interfacial tension force, the shear flow induces a breakup in a mixed pinching mode (bottom

frame of Fig. 11a), i.e. breakup type 3. Fig. 11b also indicates that an increase in Ca  1.0

causes the compound droplet to break up sooner.

4.4. Effect of the Reynolds number Re

The effects of the Reynolds number on the shape of the compound droplet with Ca = 1.0,

R12 = 0.3, 0 = 0.5, 12 = 13 =23 =13 =23 =1.0, and 0 = 1.0 are shown in Fig. 12. At Re =

0.4, the compound droplet with an elongated outer shape does not breakup as shown in the top

frame of Fig. 12a. Increasing its value to 0.8 corresponding to increasing the shear force

results in a more stretched droplet with two inner droplets at a father distance (Fig. 12b).

However, the shear force at these Reynolds numbers (i.e. Re  0.8) is not strong enough to

make the droplet break up. Thus, to induce breakup, the shear force should be increased by

increasing the Reynold number to a value higher than one, e.g. Re = 1.2 or 1.4 as shown in the

bottom frames of Fig. 12a. These Reynolds numbers produce a compound droplet with an

end-pinching mode, i.e. breakup type 2. Concerning the distance between the inner droplets

Fig. 12. Effect of the Reynolds number with Ca = 1.0, R12 = 0.5, 0 = 0.5, 12 = 13 =23 =13
=23 =1.0, and 0 = 1.0: (a) droplet shape and (b) temporal variation of the distance between
the inner droplets.
18
shown in Fig. 12b, we see that an increase in Re causes the inner droplets to move father in

the non-breakup regimes, but in the breakup regimes, an increase in Re causes the breakup to

occur sooner.

4.5. Effect of the viscous ratios

Fig. 13 shows the effects of the viscous ratios of the inner to outer fluids and of the middle

to outer fluids. Starting from a unity ratio that induces a compound droplet end-pinching

mode, i.e. breakup type 2, an increase in 13 in the range of 1.0 to 4.0 does not change the

mode of breakup (Fig. 13a). However, a decrease in this ratio to a value of 0.25 makes the

droplet change from the breakup to non-breakup mode (i.e. from breakup type 2 to

non-breakup type 2). This is understandable since viscosity represents for fluid resistance.

Thus, increasing the viscosity of the inner corresponds to reducing the rotation of the inner

droplets while enhancing their movements far away from each other. As a result, increasing

Fig. 13. Effect of the viscous ratios with Re = 1.0, Ca = 1.0, R12 = 0.5, 0 = 0.5, 12 = 13 =23
=1.0, and 0 = 1.0: (a) droplet shape and (b) temporal variation of the distance between the
inner droplets.
19
this ratio makes two inner droplets more separated (Fig. 13b), and thus promotes breakup

because of interfacial tension.

The breakup mode of the compound droplet is also strongly affected by the viscosity ratio

of the middle to the outer as shown in Fig. 13c and d. When decreasing the viscosity of the

middle, the compound droplet end-pinching mode (corresponding to 23 = 1.0), i.e. breakup

type 2, changes to the simple droplet end-pinching mode, i.e. breakup type 4, with the inner

droplets attaching to each other. In contrast, increasing this viscosity ratio to a value of 2.0

corresponds to making the middle fluid more viscous than the others, thus enhances the

droplet breakup [15]. Consequently, the simple elongated filament in the two daughter

compound droplets experiences further breakup to produce simple droplets, as the viscosity

ratio of the middle to outer fluids increases. Concerning the distance between the inner

droplets, Fig. 13d indicates two distinguished regions corresponds to two breakup modes.

4.6. Effect of the density ratio

Fig. 14 shows the effect of the density ratios on the deformation and breakup of the

Fig. 14. Effect of the density ratios with Re = 1.0, Ca = 1.0, R12 = 0.5, 0 = 0.5, 12 = 13 =23
=1.0, and 0 = 1.0: (a) droplet shape and (b) temporal variation of the distance between the
inner droplets.
20
compound droplet, where decreasing the density ratio to a value less than one enhances the

droplet breakup. That is, the breakup happens sooner as the inner or middle fluid is less dense

than the outer. However, a heavier inner or middle fluid slightly slows down the breakup. This

can be understood as the density ratios increase the heavier inner or middle fluid causes the

necking process to happen longer.

4.7. Effect of the interfacial tension ratio

The effects of the interfacial tension ratio 12 are shown in Fig. 15 with Re = 1.0, Ca = 1.0,

R12 = 0.5, 0 = 0.5, 13 =23 =13 =23 =1.0 and 0 = 1.0. Since the interfacial tension

represents for a force keeping the droplet in a circular shape thus at 12 = 0.2 (the top frame of

Fig. 15a), corresponding to a low interfacial tension force acting on the inner droplet interface,

the inner droplets are deformed with a shape in accordance with the outer interface

encapsulating them. Accordingly, the influence of the inner droplets on the behavior of the

outer interface is minor, and thus the outer interface does not break up at this ratio. At a little

higher value, i.e. 12 = 0.4, the effect of the inner droplets becomes stronger with the thinner

necks at two ends of the compound droplet. However, no breakup occurs (Fig. 15). The

situation becomes different with a compound end-pinching mode, i.e. breakup type 2, when

Fig. 15. Effect of the interfacial tension ratio with Re = 1.0, Ca = 1.0, R12 = 0.5, 0 = 0.5, 13
=23 =13 =23 =1.0 and 0 = 1.0: (a) droplet shape and (b) temporal variation of the distance
between the inner droplets.
21
the interfacial tension of the inner interface is higher than that of the outer interface. This is

understandable since an increase in 12 corresponding to making the inner interfacial tension

force stronger results in more circular inner droplets and thus induces breakup as shown in

two frames at the bottom of Fig. 15a. Fig. 15b also indicates that the increase in 12 causes the

compound droplet to break up sooner. This is consistent with the former finding for

compound droplets with a single inner droplet in shear flow [22].

4.8. Effect of the number of the inner droplets

Fig. 16 indicates that the number of the inner droplets encapsulated in the outer interface

remarkably affects the breakup of the compound droplet. Except for the top case in which

there is only one droplet at the center of the compound droplet, i.e. single-core compound

droplet, the inner droplets are distributed uniformly around the center of the compound

droplet, and their centers are at a distance of 0.5R2, i.e. 0 = 0.5. The first inner droplet is

located at 0 = 1.0 (left frames of Fig. 16). Starting from the typical case with two inner

droplets, decreasing the number of the inner droplets to one induces no breakup. In contrast,

increasing the number of the inner droplets enhances the droplet breakup. Accordingly,

Fig. 16. Effect of the number of the inner droplets with Re = 1.0, Ca = 1.0, R12 = 0.5, 0 = 0.5,
12 = 13 =23 =13 =23 =1.0 and 0 = 1.0: (a) single core, (b) double cores, (c) triple cores
and four cores. In (a) – (d), the left shows the initial shape, and the right shows the
corresponding shape at the steady state [(a) and (b)] or at a certain moment after final breakup
[(c) and (d)].
22
increasing the inner droplet number from one to four makes the compound droplet experience

from non-breakup to mixed breakup modes.

Fig. 16 also shows that the inner droplets that are nearest the top and bottom boundaries

are pushed furthest and thus make the outer interface to break up into daughter compound

droplets encapsulating them, while the remaining droplets tend to accumulate at the center to

form another daughter compound droplet encapsulating them. This is understandable since the

shear flow is strongest at the top and bottom boundaries and is decreased to the center, and

thus the droplets far from the center are easier driven. As a result, these droplets are always

located furthest from the center after breakup (Fig. 16d). Paying little more attention to the

triple-core compound droplet (Fig. 16c), three inner droplets are not located symmetrically to

the horizontal line through the center of the compound droplet. This asymmetry results in an

asymmetrical breakup unlike the other cases shown in Fig. 16. This is understandable since

the inner droplets with their asymmetrical distribution to the horizontal centered line leads to

the asymmetrical circulation and pressure within the outer interface [20] and thus causes the

outer interface to produces droplets asymmetrically.

Fig. 17. Phase diagram of Re versus Ca of the breakup mode of the multi-core compound
droplet with R12 = 0.5, 12 = 13 = 23 = 13 = 23 = 1.0 and 0 = 1.0.
23
4.9. Phase diagrams and effects of the inner droplet angle

Fig. 17 shows a phase diagram of Re versus Ca in which there exist various modes of the

multi-core compound droplet deformation and breakup. It is evident that a low shear rate

(corresponding to low values of Re) or high interfacial tension (corresponding to low values

of Ca) induces non-breakup with two types: non-breakup type 1 and 2. For Ca = 0.2

(bottommost row), increasing the value of Re in the range of 0.2 to 1.6 only induces

non-breakup type 1 in which the two inner droplets are close to each other. However, for Re =

0.2 (leftmost column), increasing Ca in the range of 0.2 to 1.6, the non-breakup mode changes

from type 1 to type 2 with two inner droplets located at two ends of the outer interface.

Fig. 18. Deformation and breakup mode diagrams under the effect of the angle 0 with R12 =
0.5, 12 = 13 = 23 = 13 = 23 = 1.0.
24
Increasing either Ca or Re promotes the droplet breakup with the formation of a mixed

breakup mode, i.e. breakup type 3, at high Re and Ca, i.e. Re  1.0 and Ca  1.2. This phase

diagram presents all deformation and breakup patterns presented in this paper.

So far, we have considered the cases in which the initial angle of the inner droplet 1 is π/2,

i.e. 0 = 1.0. Accordingly, to give a more general picture of the compound droplet

deformation and breakup, we also vary this initial angle of the inner droplet 1 and see how it

affects.

Here we fix other parameters: 12 = 13 = 23 = 13 = 23 = 1.0. The typical case is 0 =

1.0 with Re = 1.0, Ca = 1.0 and 0 = 0.5, in which the compound droplet breaks up into

droplets in the breakup type 2. Then we vary 0 in the range of 0–2.0 with corresponding Re

(in the range of 0.2–1.4), Ca (in the range of 0.2–1.6), 0 (in the range of 0.3–0.7), R12 (in the

range of 0.2–0.5) as shown in Fig. 18. These diagrams indicate that in the cases of

non-breakup, the variation of 0 has almost no effect except for Ca = 0.6 and 0.8 in which the

mode changes from the non-breakup type 1 to the breakup type 4 at low and high 0.

Increasing 0 in the range of 0.8–1.2 slightly changes the breakup mode.


5. Conclusion
We have presented the two-dimensional simulation results of deformation and breakup of

a multi-core compound droplets in simple shear flow by a front-tracking/finite difference

method. The main focus is the deformation and breakup of the compound droplet with two

inner droplets initially located symmetrically to the center. We also consider the effect of the

number of the inner droplets on the transition from breakup to non-breakup modes.

The numerical results, for the two-core compound droplets, yielded from variations of

various parameters (such as the Capillary number Ca, the Reynolds number Re, the density

and viscosity ratios, the interfacial tension ratio 12 of the inner to the outer interfaces, the

radius ratio R12 of the inner to outer and so on) reveal various deformation and breakup

patterns: non-breakup type 1, non-breakup type 2, breakup types 1–4. Increasing the value of

25
Ca from 0.2 to 1.6, the compound droplet changes from the non-breakup type 1 (two inner

droplets accumulating at the center) to a non-breakup type 2 (two inner droplets located at the

droplet ends) and to a mixed breakup mode (i.e. breakup type 3). Similarly, an increase in Re

(in the range of 0.2–1.4), 12 (in the range of 0.2–3.2) or R12 (in the range of 0.1–0.5)

enhances the compound breakup. Five diagrams based on these parameters are proposed to

show the deformation and breakup modes and the transition. The numerical results on the

number of the inner droplets indicate that an increase in the quantity of the inner droplets

encapsulated in the compound droplets enhances the droplet breakup.

To our knowledge, this is the first work providing a general picture of deformation and

breakup of a multi-core compound droplet in simple shear flow. Recently, many 2D

simulation works have been done for compound droplet deformation, e.g. [17,18,22] , in some

other configurations. In addition, many droplets have been formed and experienced

deformation and/or breakup in 2D microfluidic devices with partial wettability. Thus, our 2D

results are no doubt valuable and applicable.

However, there are some issues that need further studies. Even though the ideal shear flow

is 2D, but the flow within the outer droplet might not be 2D. Therefor, the inner droplets

might move out of the 2D plane. In addition, as mentioned in Hoang and co-workers’ work

[23], the breakup for some critical condition is affected by 3D capillary. Accordingly, 3D

calculations would be better to address these issues and overcome the limitation of the present

2D results. What will happen if the size of the inner droplets is different from droplet to

droplet. What does the rheological behavior of the compound droplet look like if the fluids are

contaminated or the interfacial tension is varied with the presence of temperature field. One

more issue should be considered for simulations to more accurately reflect the instability

associated with the breakup process in simple shear, which was mentioned in [29].

26
Acknowledgments

This research is funded by Vietnam National Foundation for Science and Technology

Development (NAFOSTED) under grant number 107.03-2017.01. The authors are grateful to

Prof. John C. Wells at Ritsumeikan University, Japan for facilitating computing resources.

The authors would like to kindly acknowledge the reviewers for their valuable suggestions

and comments to improve the quality of the paper.

References
[1] A.A. Maan, K. Schroën, R. Boom, Spontaneous droplet formation techniques for
monodisperse emulsions preparation – Perspectives for food applications, J. Food Eng.
107 (2011) 334–346.
[2] D.J. McClements, Advances in fabrication of emulsions with enhanced functionality
using structural design principles, Curr. Opin. Colloid Interface Sci. 17 (2012) 235–245.
[3] M. Abkarian, M. Faivre, A. Viallat, Swinging of red blood cells under shear flow, Phys.
Rev. Lett. 98 (2007) 188302.
[4] S.-Y. Teh, R. Lin, L.-H. Hung, A.P. Lee, Droplet microfluidics, Lab Chip 8 (2008)
198–220.
[5] T.V. Vu, J.C. Wells, H. Takakura, S. Homma, G. Tryggvason, Numerical calculations of
pattern formation of compound drops detaching from a compound jet in a co-flowing
immiscible fluid, J. Chem. Eng. Jpn. 45 (2012) 721–726.
[6] A.S. Utada, E. Lorenceau, D.R. Link, P.D. Kaplan, H.A. Stone, D.A. Weitz,
Monodisperse double emulsions generated from a microcapillary device, Science 308
(2005) 537–541.
[7] T.V. Vu, S. Homma, G. Tryggvason, J.C. Wells, H. Takakura, Computations of breakup
modes in laminar compound liquid jets in a coflowing fluid, Int. J. Multiphase Flow 49
(2013) 58–69.
[8] H. Chen, J. Li, H.C. Shum, H.A. Stone, D.A. Weitz, Breakup of double emulsions in
constrictions, Soft Matter 7 (2011) 2345–2347.
[9] J. Li, H. Chen, H.A. Stone, Breakup of double emulsion droplets in a tapered nozzle,
Langmuir 27 (2011) 4324–4327.
[10] H.-C. Kan, H.S. Udaykumar, W. Shyy, R. Tran-Son-Tay, Hydrodynamics of a compound
drop with application to leukocyte modeling, Phys. Fluids 10 (1998) 760–774.
[11] Y. Chen, X. Liu, C. Zhang, Y. Zhao, Enhancing and suppressing effects of an inner
droplet on deformation of a double emulsion droplet under shear, Lab Chip 15 (2015)

27
1255–1261.
[12] Y. Chen, X. Liu, Y. Zhao, Deformation dynamics of double emulsion droplet under shear,
Appl. Phys. Lett. 106 (2015) 141601.
[13] Z.Y. Luo, L. He, B.F. Bai, Deformation of spherical compound capsules in simple shear
flow, J. Fluid Mech. 775 (2015) 77–104.
[14] H. Hua, J. Shin, J. Kim, Dynamics of a compound droplet in shear flow, Int. J. Heat
Fluid Flow 50 (2014) 63–71.
[15] Y. Chen, X. Liu, M. Shi, Hydrodynamics of double emulsion droplet in shear flow, Appl.
Phys. Lett. 102 (2013) 051609.
[16] J. Wang, J. Liu, J. Han, J. Guan, Effects of complex internal structures on rheology of
multiple emulsions particles in 2D from a boundary integral method, Phys. Rev. Lett.
110 (2013) 066001.
[17] J. Wang, J. Liu, J. Han, J. Guan, Rheology investigation of the globule of multiple
emulsions with complex internal structures through a boundary element method, Chem.
Eng. Sci. 96 (2013) 87–97.
[18] J. Wang, H. Jing, Y. Wang, Possible effects of complex internal structures on the
apparent viscosity of multiple emulsions, Chem. Eng. Sci. 135 (2015) 381–392.
[19] J. Wang, X. Wang, M. Tai, J. Guan, Oriented shift and inverse of the daughter droplet
due to the asymmetry of grand-daughter droplets of multiple emulsions in a symmetric
flow field, Appl. Phys. Lett. 108 (2016) 021603.
[20] G. Xu, X. Wang, S. Xu, J. Wang, Asymmetric rheological behaviors of double-emulsion
globules with asymmetric internal structures in modest extensional flows, Eng. Anal.
Bound. Elem. 82 (2017) 98–103.
[21] G. Tryggvason, B. Bunner, A. Esmaeeli, D. Juric, N. Al-Rawahi, W. Tauber, J. Han, S.
Nas, Y.-J. Jan, A front-tracking method for the computations of multiphase flow, J.
Comput. Phys. 169 (2001) 708–759.
[22] T.V. Vu, L.V. Vu, B.D. Pham, Q.H. Luu, Numerical investigation of dynamic behavior of
a compound drop in shear flow, J. Mech. Sci. Technol. 32 (2018) 2111–2117.
[23] D.A. Hoang, L.M. Portela, C.R. Kleijn, M.T. Kreutzer, V. van Steijn, Dynamics of
droplet breakup in a T-junction, J. Fluid Mech. 717 (2013) R4.
[24] V.N. Duy, T.V. Vu, A numerical study of a liquid drop solidifying on a vertical cold wall,
Int. J. Heat Mass Transfer 127 (2018) 302–312.
[25] S.O. Unverdi, G. Tryggvason, A front-tracking method for viscous, incompressible,
multi-fluid flows, J. Comput. Phys. 100 (1992) 25–37.
[26] T.V. Vu, Three-phase computation of solidification in an open horizontal circular
cylinder, Int. J. Heat Mass Transfer 111 (2017) 398–409.
[27] T.V. Vu, Deformation and breakup of a pendant drop with solidification, Int. J. Heat
Mass Transfer 122 (2018) 341–353.

28
[28] G. Tryggvason, R. Scardovelli, S. Zaleski, Direct numerical simulations of gas-liquid
multiphase flows, Cambridge University Press, Cambridge; New York, 2011.
[29] E.J. Hinch, A. Acrivos, Long slender drops in a simple shear flow, J. Fluid Mech. 98
(1980) 305–328.

29

You might also like