You are on page 1of 19

Breakup modes of fluid drops in

confined shear flows


Cite as: Phys. Fluids 28, 073302 (2016); https://doi.org/10.1063/1.4954995
Submitted: 19 January 2016 • Accepted: 18 June 2016 • Published Online: 06 July 2016

Nilkamal Barai and Nibir Mandal

ARTICLES YOU MAY BE INTERESTED IN

Droplet deformation and breakup in shear flow of air


Physics of Fluids 32, 052109 (2020); https://doi.org/10.1063/5.0006236

Numerical simulation of breakup of a viscous drop in simple shear flow through a


volume-of-fluid method
Physics of Fluids 12, 269 (2000); https://doi.org/10.1063/1.870305

Deformation and breakup of a confined droplet in shear flows with power-law


rheology
Journal of Rheology 61, 741 (2017); https://doi.org/10.1122/1.4984757

Phys. Fluids 28, 073302 (2016); https://doi.org/10.1063/1.4954995 28, 073302

© 2016 Author(s).
PHYSICS OF FLUIDS 28, 073302 (2016)

Breakup modes of fluid drops in confined shear flows


Nilkamal Barai1,2 and Nibir Mandal1,a)
1
High Pressure and Temperature Laboratory, Jadavpur University, Kolkata 700032, India
2
Department of Physics, Jadavpur University, Kolkata 700032, India
(Received 19 January 2016; accepted 18 June 2016; published online 6 July 2016)

Using a conservative level set method we investigate the deformation behavior of


isolated spherical fluid drops in a fluid channel subjected to simple shear flows,
accounting the following three non-dimensional parameters: (1) degree of confine-
ment (Wc = 2a/h, where a is the drop radius and h is the channel thickness); (2)
viscosity ratio between the two fluids (λ = µ d /µ m , where µ d is the drop viscosity
and µ m is the matrix viscosity); and (3) capillary number (Ca). For a given Wc , a drop
steadily deforms to attain a stable geometry (Taylor number and inclination of its long
axis to the shear direction) when Ca < 0.3. For Ca > 0.3, the deformation behavior
turns to be unsteady, leading to oscillatory variations of both its shape and orientation
with progressive shear. This kind of unsteady deformation also occurs in a condition
of high viscosity ratios (λ > 2). Here we present a detailed parametric analysis of
the drop geometry with increasing shear as a function of Wc , Ca, and λ. Under a
threshold condition, deforming drops become unstable, resulting in their breakup into
smaller droplets. We recognize three principal modes of breakup: Mode I (mid-point
pinching), Mode II (edge breakup), and Mode III (homogeneous breakup). Each of
these modes is shown to be most effective in the specific field defined by Ca and λ.
Our study also demonstrates the role of channel confinement (Wc ) in controlling the
transition of Mode I to III. Finally, we discuss implications of the three modes in
determining characteristic drop size distributions in multiphase flows. Published by
AIP Publishing. [http://dx.doi.org/10.1063/1.4954995]

I. INTRODUCTION
Fluid drops suspended in another immiscible fluid undergo deformation and subsequently
breakup into smaller droplets when the bulk fluid is subjected to strong shear flows. The mechanism
of such droplet formation has recently emerged as a subject of multidisciplinary research owing to
its relevance to various natural systems, e.g., bubbly magma flows, gas-oil flows in porous geolog-
ical layers, and air-water interactions in oceanic environments as well as a range of engineering
applications, like production and processing of emulsions and polymer blends, aerosols, and drug
delivery systems. Following the pioneering work by Taylor,1,2 the dynamics of drop deformations
turned to be a major focus of theoretical and experimental investigations in multiphase fluid me-
chanics.3,4 More notably, understanding the drop breakup mechanism became a major challenge
for the scientists and technologists, especially engaged in modeling the drop size distributions
(DSD) for the formation of dispersions and emulsions5–7 (for a detailed review, see Rallison,3 and
Tucker and Moldenaers8). Using experiments earlier workers3 recognized two major drop breakup
mechanisms-end pinching and capillary instability to predict the DSD in multiphase flows. They
estimated the critical capillary number (Cac ), required for the drop breakup as a function of vis-
cosity ratio (λ). The end pinching mechanism dominates when Cac ≤ Ca ≤ 2Cac . This process of
drop breakup is replaced by the capillary instability mechanism as Ca > 2Cac . Several workers
have demonstrated that the end pinching mechanism is mediated by a process, called elongative end
pinching,7 in contrast to retractive end pinching reported in earlier studies.10 Despite a significant

a) Electronic mail: nmandal@geology.jdvu.ac.in

1070-6631/2016/28(7)/073302/18/$30.00 28, 073302-1 Published by AIP Publishing.


073302-2 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

progress in recent time, the drop dynamics is still a challenging front of the theoretical work in order
to better understand the breakup mechanisms.
Our present study aims to characterize the pre- and post-breakup deformation behavior of drops
in varying confined shear flow conditions. Using numerical simulations we show that they can
undergo breakup in different ways depending upon the degree of channel confinement, in addition
to the effects of Ca and λ. We recognized three breakup modes as- Mode I: mid-point pinching,
Mode II: edge breakup, and Mode III: homogeneous breakup (Fig. 1). These three modes give rise
to contrasting DSD in multiphase flows. Our study also shows the steady and unsteady deformations
of non-breaking drops with progressive shear flow.
Based on the conservative level-set method we adopted a computational fluid dynamic (CFD)
approach to simulate the deformation of an isolated drop suspended in another immiscible fluid,
subjected to Couttes flows within a confined channel. The most challenging step in simulation
studies concerns an efficient mathematical formulation to track the interfaces between two fluid
phases that are essentially transient, undergoing intense geometrical modifications due to large
deformations and subsequent breakup. Earlier workers have developed a range of interface track-
ing methods, such as front-tracking (FT),11,12 volume-of-fluid (VOF),13–19 and level set (LS)20–24
methods. All these methods treat two-phase flows with the density and viscosity varying smoothly
across the moving interface, captured either in an Eulerian (VOF and LS) or a Lagrangian frame-
work (FT). However, their numerical implementation vastly differs from one another, where each
method has its own advantages and disadvantages. We used the Level Set (LS) method, which is
one of the most efficient front capturing methods introduced by Osher and Sethian20 in late 1980s,
and subsequently favored by researchers dealing in multiphase fluid mechanics over the last couples
of decades.25–27 The advantage of this method is that it takes care of coalescence and breakup of
the phase interface automatically and does not require any explicit reconstruction unlike the VOF
method. However, there are some disadvantages of LS in ensuring the conservation of mass.25,28 For
a standard LS method, a mass loss/gain results mainly from two reasons: (1) discretization of the
LS equation, leading to numerical dissipation of the masses and (2) re-initialization of the phase
space. Using different numerical formulations it has been possible to improve the mass conservation
issue of the LS method.29–31 Recently, Olsson and Kreiss32 and Olsson et al.33 have proposed a

FIG. 1. Three principal breakup behaviour of a viscous drop hosted in another immiscible viscous fluid subjected to shear
flows (arrows indicate the shear sense). Mode I (Mid-point pinching): the drop undergoes breakup through mid-point
pinching, producing two droplets of nearly equal size. Mode II (Edge breakup): the drop becomes unstable preferentially
at their edges, leading to detachment of smaller droplets from the intact central segment. Mode III (Homogeneous breakup):
the drop breaks apart into a large number of droplets, with equal sizes.
073302-3 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

conservative level-set (CLS) method, which remarkably helps to overcome the mass conserva-
tion problem. Moreover, this approach offers a great benefit for automatic handling of topology
changes and efficient parallelization, without incurring much additional costs. This study employs
an improved CLS method to solve our incompressible two-phase shear flow problem based on
unstructured meshes. The biggest advantage of working on the CLS platform is that the method can
track an interface with great accuracy, restoring a good mass conservation property.
We present this paper in the following order. Sec. II provides a mathematical basis of the LS
method used in our numerical simulations and highlights different model parameters used in the
simulations. We then describe the simulation results in Section III. The first part of this description
shows the progressive deformation behavior of drops without any breakup under confined shear
flows. The rest deals with a detailed parametric analysis of the breakup process, giving an account
of the hydrodynamic conditions for the three breakup modes. We conclude our principal findings in
Section IV.

II. COMPUTATIONAL APPROACH


A. Level-set method: The mathematical framework
Consider a simple closed curved Γ(t) in a two-dimensional space to define the interface be-
tween the two fluids phases, attributed to a velocity field, v = (v x , v y ). A level set (LS) function ϕ
is chosen, where the zero-level set (ϕ = 0) detects the evolving interface Γ(t). The LS function is
restricted to a domain as
ϕ = ϕ(r,t), r ∈ Ω, (1)
where Ω is the domain of interest, defined by the physical problem. The time dependent expression
of ϕ corresponding to the moving interface is given by the advection equation
∂t ϕ = v · ∇ϕ. (2)
Considering the two fluids separated by the interface Γ in the domain Ω as incompressible, the
velocity field must satisfy the divergence equation, ∇ · v = 0. Eq. (2) can be then written as
∂t ϕ + ∇ · (vϕ) = 0. (3)
This represents a continuity equation of the level set function.
The numerical approach approximates the solution to the time-dependent initial value problem
of the associated level set function, where the zero-level set can always locate the moving interface
at an instant. For a standard level set method, an indicator of the LS function ϕ is defined as
assigned distance to the interface Γ and the smallest distance between a given point in the domain
and the interface,
|ϕ (X, t = 0)| = min X Γ ∈Γ (|X − X Γ |) , (4)
is used to delineate the interface. ϕ > 0 represents one side of the interface, whereas ϕ < 0 the other
side, while the zero LS of ϕ indicates the interface
Γ = {X |ϕ(X,t) = 0|}. (5)
The conservative level set method, on the other hand, employs a Heaviside function as

 0, ϕ < 0

H(ϕ) = 
 1, ϕ > 0. (6)

However, an abrupt jump in the field, as defined by the Heaviside function in Eq. (6) causes
instabilities in the computational operations, especially when the materials on either side of the
interface have strong contrasts in their physical properties. To overcome this problem a smeared out
Heaviside function is used in the conservative LS method in the form
073302-4 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)


 0, ϕ < 0
1 ϕ ( πϕ )

1


H (ϕ) =  + + , −ε ≤ ϕ ≤ ε,

 sin (7)
 2 2ε 2π ε
1, ϕ > 0




where ε denotes half the thickness of the interface. The interface thickness depends on the grid size
chosen for meshing. We can now define a new LS function
Φ(r) = Hsm(ϕ(r)). (8)
The LS function in Eq. (8) has an advantage in the sense that the function describes continuous
variations of the material properties across the interface. For example, the density and the viscosity
of the two fluids can be expressed as
ρ (r) = ρ1 + (ρ2 − ρ1) Φ(r), (9)
µ (r) = µ1 + (µ2 − µ1) Φ(r), (10)
where ρi and µi are the density and the viscosity of phases, i = 1, 2. The interface is attributed to
the 1/2-level set, as defined in Eq. (8). While operating with interfaces it is necessary to define the
normal vector and the curvature of the interface, which are obtained from the distance function,
∇ϕ
n= , (11)
|∇ϕ|
∇ϕ
κ (ϕ) = −∇ · n = −∇ · . (12)
|∇ϕ|
It is noteworthy that the shape of ϕ can be complexly distorted due to a strongly heterogeneous
velocity field, leading to inevitable numerical errors or artificial diffusion and the numerical prob-
lem can turn worse. To overcome this hurdle, Olsson et al.32,33 added an a re-initialization step to
maintain the shape and width of the interface by using the following equation:
ϕ (1 − ϕ) ∇ϕ
( )
∂tϕ + ∇ · = ∇ · (ε(∇ϕ · n)n) . (13)
|∇ϕ|
In Eq. (13), the diffusion in the normal direction is balanced by the compressive term, where the
shape and the width of the interface are conserved. These interface conditions are compulsory in
order to maintain the conservative properties of the interface.

B. Governing equations
Implementation of the numerical simulations employs a set of governing equations based
on the conservation of momentum and mass. The conservation of momentum is described by
Navier–Stokes equations on a spatial-time domain Ω × [0,T] with boundary ∂ Ω as
∂t (ρk vk ) + ∇ · (ρk vkvk) = ∇ · Sk + ρk g in Ω k × [0,T], (14)
where
Sk = −pk I + µk (∇vk + (∇vk)T ). (15)
Imposing the condition of incompressibility for the fluid, we have
∇ · vk = 0 in Ω k × [0,T], (16)
where, Ω = Ω1UΩ2UΓ, k = {1, 2} denote the sub domains corresponding to the two fluid phases
and Γ = ∂ Ω1 ∩ ∂ Ω2 represents the fluid interface. In Equations (14)–(16) ρ and µ stand for the
density and dynamic viscosity of the fluid, v is the velocity field, g is the acceleration due to gravity,
and p, S, and I stand for pressure, stress tensor, and identity tensor, respectively.
Assuming no mass transfer between the fluids, we can impose a continuous velocity condition
at the interface
v1 = v2 in Γ × [0, T]. (17)
073302-5 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

Neglecting variations of the surface tension coefficient σ, the interface is subjected to the following
boundary condition for momentum conservation:
(S1 − S2) · n = σκn in Γ × [0, T], (18)
where n is the unit normal vector outward to ∂ Ω1 and κ is the interface curvature. By combining
Eqs. (13)–(15) and (17) and (18), we have
∂t (ρv) + ∇ · (ρvv) = −∇ p + ∇ · µ∇v + ∇ · µ(∇v)T + ρg + σκnδ Γ in Γ × [0, T]. (19)
Since the Navier–Stokes equation (Eq. (19)), based on the finite-element integration, involves dis-
cretization, the aforesaid problems can be resolved by converting the singular term σκnδ Γ , to a
volume force as
σκnδΓ = σκ (ϕ) ∇ (ϕ) , (20)
where κ(ϕ) is given by Eq. (11) and Eq. (12). The most significant computational advantage of this
method is that an explicit tracking of the interface is not necessary. In this way the level-set method
is coupled to the Navier–Stokes equations for incompressible fluids.

C. Two-phase flow modeling


Our model simulates a Couette flow with zero gravity, excluding the effects of density in
drop deformations. The initial model contains a circular liquid drop with an initial radius, a, and
viscosity µ d , suspended in another liquid of viscosity, µ m . The interfacial surface tension between
the drop and the ambient fluid, σ is held constant. We performed numerical simulations considering
various physical quantities in terms of the following two dimensionless numbers: capillary number
Ca = γ̇ µ m a/σ (a dimensionless measure of the balance between viscous and interfacial forces)
and viscosity ratio λ = µ d /µ m . We use the channel confinement, normalized to drop diameter
(Wc = 2a/h) as another dimensionless parameter in our simulations. In our multiphase fluid model
the simple shear flow is simulated by imposing a steady shear velocity, u∞ = (γ̇ y, 0, 0) at the model
boundary in a Cartesian frame, where γ · is the imposed shear rate (Fig. 2(a)). Fluid drops undergo
deformation under the shear flow, where their deformation occurs through the action of two oppo-
site forces developed by (1) the viscous shear stress exerted by the matrix fluid and (2) surface ten-
sion force (responsible for drop shape relaxation). Their mathematical expressions are respectively,
τγ̇ = µm γ̇, (21)
σ
τσ = . (22)
a

FIG. 2. Two-phase shear flow model used for numerical simulations of drop deformations based on the CLS method.
(a) Choice of boundary conditions and other model parameters (see text for details). (b) Consideration of the semi-axes
dimensions (L, B) and the orientation (θ) of the long axis with the shear direction to describe the deformed drop geometry.
Arrows indicate the sense of shear.
073302-6 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

TABLE I. Numerical results of deformed drop geometries with varying grid


refinement at fixed Ca = 0.3, λ = 2, and Wc = 0.5.

h maxa 1.18 1.12 0.52


δb 1.36 1.34 1.33
θc 26.16 28.64 29.30

ah is the grid element size.


max
bδ is the elongation parameter, L/a, where a is the initial radius of circular
drops and 2L is the length dimension of their deformed shapes.
c θ is inclination of the long axis of deformed drops to the shear direction.

We investigate the effects of the capillary number, Ca, and the viscosity ratio, λ in controlling the
drop deformations and the breakup mechanisms, depending upon the degree of channel confinement
(Wc ). For very low Reynolds numbers (Stokes flow), the parameters used to measure the steady
shape deformation are: Taylor deformation parameter, D = (L − B)/(L + B), where L and B are the
semi-major axis and semi-minor axis of the deformed drop, respectively. The ratio of maximum
elongation to initial drop diameter (δ = L/a) is used to characterize the drop deformation when
their steady shape is no longer ellipsoidal. The second parameter is the inclination of drop’s long
axis (θ) to the shear flow direction (Fig. 2(b)).

D. Validation of mesh refinement


We have performed a grid refinement analysis of our simulation results, considering both the
accuracy level of model results as well as the computational run time. The full set of dimensionless
parameters (Re, Ca, λ, m) remains unchanged even when the resolution is increased. We have
refined the mesh by multiplying factors of β = 1.3 and 3.5 times, starting from normal meshing
with a maximum grid element size, hmax = 1.8 mm, where the interface controlling parameter,
ϵ = hmax/2. The drop elongation and orientation angle for varying grid sizes are presented in
Table I. The difference in the values of elongation (δ) for hmax = 1.12 and 0.52 is less than 1.1%,
whereas that of drop orientation is less than 2.2% (Table I). Further grid refinement yields no
significant difference in the results.

III. RESULTS AND DISCUSSIONS


A. Drop deformations without breakup
1. Effects of channel confinement
For specific Ca and λ values (discussed later), fluid drops undergo continuous deformation
without any breakup even for large finite shear. We systematically investigate the deformation
behavior of such non-breaking drops as a function of the channel confinement. It is noteworthy that
the capillary number, Ca determines the effects of bulk shear rates on drop deformations. To eval-
uate the influence of channel confinement independently we thus ran simulations, keeping Ca fixed
at a given value so that the effects of shear rates remain constant. The confinement parameter, Wc ,
is varied on a wide range (0.125–1.0) to simulate almost an unconfined to an extreme confinement
condition, where the channel width equals the drop diameter. We present here a set of simulations
run with Wc = 0.125, 0.25, 0.5, 0.8, 1, keeping fixed Ca = 0.3 and λ = 2 (Fig. 3). The confinement
has virtually no effects when Wc ≤ 0.25. Circular drops progressively deform into elliptical shapes,
which attain a stable geometry (both δ and θ) with ongoing shear flow (Figs. 3(a) and 3(b)).
The confinement effects become significant as Wc exceeds 0.25. The drops show increasing
flattening (δ) with increasing Wc (>0.25), and they are deformed not to typical ellipsoidal shapes,
as observed in weak confinement (Wc = 0.125), but to sigmoidal shapes (Figs. 3(c) and 3(d)).
Under extreme confinement (Wc = 1) drops cannot retain the spherical shapes. Table II shows a
quantitative variation of the stable elongation, δ s with Wc for non-breaking drops.
073302-7 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

FIG. 3. Progressive stages of drop deformations in confined shear flows for varying confinement parameter, Wc (values on
top left corners). (a)–(c) Low confinement conditions (Wc ≤ 0.5): Steady deformation of drops into stable elliptical shapes.
(d) High confinement condition (0.5 < Wc ≤ 0.8): Deformation into non-elliptical shapes. (e) Extreme confinement condition
(Wc > 0.8): Breakup of drops into smaller daughter drops. Numerical values on the bottom right corners are the normalized
time steps (t ∗ = γ̇t). The Capillary number and the viscosity ratio are kept constant (Ca = 0.3 and λ = 2).

Increasing channel confinement (i.e., Wc = 0.25 to 0.8) increases δ by a factor of 1.5 (Fig. 4(a)).
Similarly, the stable drop orientation, θ s shows a large variation with Wc . For example, θ s ∼ 25◦ for
Wc = 0.25, which decreases to ∼10◦ when Wc = 0.8 (Fig. 4(b)).
Our simulations results described suggest that increasing channel confinement strengthens the
capacity of ambient viscous medium to deform drops under a given shear flow. We investigated the
underlying hydrodynamics of confined shear flows that promotes drop deformations. For a given
Ca, our simulations show that increasing confinement intensifies the local hydrostatic pressure in

TABLE II. Stable values of different geometrical parameters as a function of wall confinement (Wc), capillary number (Ca),
and viscosity ratio (λ).

Ca = 0.3, λ = 2 Wc = 0.5, λ = 2 Ca = 0.3, Wc = 0.5

Wc Ca λ

0.125 0.25 0.5 0.8 1.0 0.10 0.30 0.45 0.1 1 3


δa 1.32 1.36 1.70 2.10 Breakup 1.34 2.2 Unstable 1.03 1.6 Unstable
Db 0.27 0.26 0.42 0.67 ... 0.26 0.69 ... 0 0.42 ...
θc 24.9 24 19.6 10.2 ... 30.5 10.8 ... 37.4 21.4 ...

aδ is the elongation parameter, L/a, where a is the initial radius of circular drops and 2L is the length dimension of their
deformed shapes.
b D is the Taylor deformation parameter, (L − B)/(L + B), where L and B are the major and the minor semi-axes of elliptical

deformed shapes.
c θ is inclination of the long axis of deformed drops to the shear direction.
073302-8 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

FIG. 4. (a) Computed plots of elongation (δ) versus normalized shear flow time (t ∗) variations. δ tends to stabilize with
progressive flow, where the stable value is larger for larger Wc . (b) Orientation (θ) versus t ∗ plots. The drop orientation
fluctuates, but later tends to attain stable inclinations with the shear direction. Increasing Wc lowers their stable inclination.
Ca = 0.3 and λ = 2.

the neighborhood of drops (Fig. 5(b)), which in turn strengthens the ambient viscous flow to deform
the drops.

2. Effects of capillary number


We varied the shear rate in channel flows to change the capillary number (Ca), keeping the
channel confinement and the viscosity ratio constant (Wc = 0.5 and λ = 2). With increasing Ca the
drops breakup when Ca exceeds a critical value (Cac = 0.45), as reported in earlier studies.3,4,34 We
present here a set of three simulations with Ca = 0.1, 0.3, and 0.45 to demonstrate the effects of
increasing capillary number on drop deformations (Fig. 6).
For Ca = 0.1, drops deform into an elliptical shape, which readily attains a stable geometry
(δ s = 1.34 and θ s = 30.5◦) due to ambient shear flow (Fig. 7). In contrast, they do not stabilize, but
undergo large elongation (δ > 2.2) when Ca = 0.3 (Fig. 7(a)). Furthermore, they also continuously
rotate towards the shear plane. After attaining a maximum δ of 2.8, their flattened shapes start to
relax and reduce their elongation. During such relaxation phase the drops rotate in the backward
direction (i.e., against the shear sense) (Fig. 7(b)). These two processes: elongation and relaxation,
coupled with forward and backward rotation occur periodically in progressive shear. However, the
drops finally stabilize their geometry (δ s = 2.2 and θ s = 10.7◦) after a large finite shear. Their un-
steady behavior escalates further with increasing Ca (Fig. 7(c)). For Ca = 0.45, drops show a large
elongation (δmax = 4.7) during their forward rotation; but the deformed shapes subsequently relax to
greatly reduce the elongation (δ = 2.3) on their back rotation. This oscillating variation of periodic
elongation followed by shortening in δ for Ca = 0.45 by far exceeds that for Ca = 0.3 by almost
10 times in magnitude. Secondly, the oscillation tends to decay with progressive shear, but for large

FIG. 5. Normalized pressure distribution around a drop in confined shear flows. (a) Simulation with Wc = 0.5 and λ = 2. (b)
Simulation high confinement (Wc = 0.8). (c) Simulation with a high viscosity ratio (λ = 5). Note that increase in either
confinement or viscosity ratio intensifies the local pressure around the drops. Capillary number was kept constant (Ca
= 0.47).
073302-9 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

FIG. 6. Effects of capillary number (Ca) on drop deformations, where λ = 2 and Wc = 0.5. (a) Development of a stable
geometry at Ca = 0.10. (b) and (c) Unsteady deformations at higher Ca (0.30 and 0.45), characterized by alternate elongation,
and relaxation of drop shapes with ongoing shear flow. Notice that the drops also show oscillatory forward and backward
rotational motion when Ca is high.

values of Ca (i.e., 0.45) the decay persists notably even after a large finite shear flow (Fig. 7(c)).
A summary of the simulation results is presented in Table II. For Ca = 0.1, the maximum value of
Taylor deformation parameter, D = 0.2, which increases to ∼0.7 when Ca = 0.3. Also the stable
inclination of deformed drops with the shear direction, for Ca = 0.1 is θ ≈ 30◦, which decreases to
∼10◦ for Ca = 0.45 (Fig. 7(b)).
Our simulations thus reveal that the drop deformations occur in an unsteady state when Ca is
large. They undergo stretching and contraction in a cyclic fashion with ongoing shear, and at the

FIG. 7. Variations of (a) drop elongation (δ) and (b) orientation (θ) with shear flow time (t ∗) for Ca = 0.10, 0.30, and 0.45.
Deformed drop geometry readily stabilizes under low Ca conditions. Increase in Ca causes the deformation to turn unsteady,
resulting in pulsating variations of δ and θ. λ = 2 and Wc = 0.5.
073302-10 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

same time their long axes rotate forward and backward with the shear plane maintaining the same
phase with stretching and contraction respectively. To explain this unsteady dynamics, we account
for the two principal forces acting on a drop: (1) viscous drag that causes the drop to stretch and
to align it along the shear direction and (2) surface tension that offers a resistance to the stretching,
and thereby restore it back to the initial shape (i.e., circular) and orientation (i.e., close to 45◦ with
the shear direction). For low Ca shear flows the surface tension is strong enough to counter the
action by viscous drags, leading to a stable deformed geometry of the drop with progressive shear.
Conversely, high Ca shear flows produce large viscous drags that force the drops to continuously
stretch and align them along the shear direction. The instantaneous bulk stretch rate along the drop
axis is 0.5 γ̇ sin 2θ, where θ is the inclination of drop axis to the shear direction, and γ̇ is the
bulk shear rate. The viscous force associated with stretching thus decreases as the drop axes rotate
towards the shear plane and is taken over by the action of surface tension to restore the initial shape
and orientation, resulting in back rotation of the drops. When the drops rotate back, the viscous
drags become again important, resulting in stretching and forward rotation of the drops. These two
onward and backward processes of drop deformation, driven by a competition between viscous drag
forces and surface tension keep the drops in an unsteady state with ongoing shear flows when Ca is
large.

3. Effects of viscosity ratio


Our simulations show that the viscosity ratio (λ) greatly influences the deformation behavior
of fluid drops in a confined shear flow. To show this, we present here another set of simulations
run with λ = 0.1, 1, and 3, keeping a fixed capillary number (Ca = 0.3) and confinement param-
eter (Wc = 0.5) (Fig. 8). Increasing λ has two main effects on the drop deformation: (1) stable to
unstable transition of the drop shape and (2) increasing flattening and rotation towards the shear
direction. Low viscosity ratio (λ = 0.1–1) drops undergo relatively weak flattening, and they readily
attain the stable geometry. For λ = 0.1, the Taylor deformation parameters D is close to 0, and
θ = 38◦, whereas for λ = 1, D = 0.42, and θ = 21◦ in the stable state.

FIG. 8. Effects of the viscosity ratio (λ) on drop deformations: λ = 0.1 (a), 1.0 (b), and 3.0 (c). Ca and Wc are kept constant
(Ca = 0.3 and Wc = 0.5). Notice that (1) increasing λ causes drops to flatten more and (2) there occurs a transition from
steady to unsteady deformation as λ exceeds a threshold value (λ = 1).
073302-11 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

FIG. 9. (a) Plots of (a) elongation (δ) and (b) orientation (θ) of deformed drops with shear flow time (t ∗) as a function of the
viscosity ratio (λ). Ca = 0.3 and Wc = 0.5. It is noteworthy both δ and θ are stable for λ < 3, but unstable as λ ≥ 3.

For large viscosity ratios (λ = 3.0), both finite elongation (δ) and inclination (θ) remain un-
steady (Fig. 9), varying from maximum values of 4.4 and 12◦ to minimum values of 2 and 2◦,
respectively (Table II). We obtain a threshold viscosity ratio (λ = 3) beyond which drops remain no
longer intact during their deformation, but break up into smaller daughter droplets.
It follows from the simulation results that the degree of drop deformation holds an inverse
relation with viscosity ratio λ. However, one would expect rigid –like behavior of drops if they
have large viscosity ratios. In a Sec. III A 1 we have discussed that the channel confinement builds
local pressures that promote drop deformations. For a given channel confinement, the magnitude of
such local pressure multiplies with increasing viscosity ratios (Fig. 5(c)) and thereby strengthens
the capacity of the ambient viscous medium to deform drops of high viscosity. Interestingly, earlier
experimental studies39 have also shown that drops can breakup in shear flows when they have large
viscosity ratios (λ > 10).

B. Modes of drop breakup


Predicting the critical conditions in which fluid drops become unstable and break up into
smaller droplets is a major challenge in both theoretical and experimental studies of multiphase
fluid flows, as discussed in the first section of this paper. To investigate the breakup process,
earlier studies dealt with mainly two dynamic parameters: capillary number and viscosity ratio,
and showed how the breakup behavior can change depending upon the relative values of these
parameters.9 Using the level set method we characterize the breakup process accounting the channel
confinement as an additional parameter. We recognize three principal modes of breakup: Mode I
(mid-point pinching), Mode II (edge breakup), and Mode III (homogeneous breakup) (Fig. 1). In
Mode I, drops undergo flattening into an elongate shape, pinch out preferentially at its mid-point,
and subsequently break up into two equal daughter drops (Fig. 10(a)). Mode II, in contrast, leads
to a break up selectively at the drop edges, leaving the central region intact (Fig. 10(b)). In Mode
III the break up process affect the entire drop, forming homogeneously distributed smaller daughter
drops (Fig. 10(c)). The simulation results presented above show specific dynamics conditions for
these three modes.
Fluid drops within a channel subjected to shear flow break up when the capillary number
exceeds a critical value (Cac ). Cac depends on the viscosity ratio (λ) as well as the channel
confinement (Wc ). For a given Wc (Wc = 0.5), Cac varies nonlinearly with λ, showing broadly an
exponential decrease with increasing λ (Fig. 11(a)).
The Cac versus λ variation obtained from our simulations differs from earlier results. For
example, Zhao9 has shown that the critical value of Cac decreases with λ, but shows a nonlinear
increase as λ exceeds 1. We ran our computation up to λ = 102, but did not find any appreciable
increase of Cac with λ. This difference perhaps originates due to the additional effect of channel
confinement. For a given λ, increasing Wc lowers Cac (Fig. 11(b)), and thereby promotes the
073302-12 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

FIG. 10. Simulations of drop breakup modes in numerical models. (a) Mode I: Drops develop midpoint pinching in the
course of their progressive elongation, leading to a breakup along the central line and generation of two daughter drops of
nearly equal size, often coupled with small satellite droplets between the two drops. (b) Mode II: Drops show preferentially
edge breakup, leaving their central part intact, and produce a train of smaller droplets on either side of the deformed primary
drops. (c) Mode III: Drops undergo homogeneous breakup, affecting their entire length, and produce uniformly distributed
second generation droplets.

breakup process. This is in agreement with the findings of Bentley and Leal,10 who have demon-
strated from physical experiments that a high-viscosity drop (λ > 4) can breakup under confined
channel flows. We systematically varied Ca and λ to study the three breakup modes for varying
channel confinement (Wc ). Here we present a set of specific simulations run with Wc = 0.5 to
discuss different features of these breakup modes.
For Ca > Cac , fluid drops undergo continuous elongation to develop a thread-like shape, and
reach a stage when further elongation ceases to take place with ongoing shear flow. At this critical
stage they start to breakup following any of the three modes, depending on the capillary number,
viscosity ratio, and the wall confinements. We show three specific simulation results to demonstrate
the breakup processes in Mode I, II, and III. The breakup initiates through localization of pinching
at the mid-point of deformed drops when λ = 2 and Cac = 0.47 and eventually gives rise to two
daughter drops of almost equal size (Mode I; Fig. 10(a)). Similar breakup style has been widely
reported from numerical simulations as well as laboratory experiments which run with low capillary
number.34–37 These secondary drops relax themselves to stabilize their geometry and do not break
up further with ongoing shear. The Mode I breakup process often yields a small satellite drop
between two drops of much larger size. There occurs a remarkable change in the breakup mech-
anism as we increase the capillary number to high values. When Ca = 1, the pinching instability

FIG. 11. Computed plots of the critical capillary number (Cac ) for drop breakup as a function of (a) viscosity ratio (λ) and
(b) channel confinement (Wc ).
073302-13 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

FIG. 12. Numerical simulations of drop breakup in shear flows with confinement: (a) Wc = 0.5, (b) Wc = 0.8, and (c)
Wc = 1.0. Increasing confinement causes midpoint pinching to occur over a broad zone, leading to a transition from Mode I
to III breakup (left to right).

develops uniformly over the entire length of strongly stretched drops, leading to a homogeneous
breakup (Mode III). This mode produces a large number of daughter drops of nearly equal size.
These secondary droplets remain stable and increase their spacing with progressive shear. For large
shear they are homogeneously distributed in the matrix fluid, leaving no impression of their parent
drop (Fig. 10(c)). With increasing viscosity ratio there is also a transition in the break up modes.
For λ > 3, the elongated drops undergo fragmentation preferentially at their edges, leaving the
central regions intact (i.e., Mode II; Fig. 10(b)). Consequently, the production of daughter droplets
is heterogeneous, localizing at the two flanks of the parent drop. The large drop in the central
region relaxes to stabilize its geometry. With further increase in λ daughter drops multiply in num-
ber, but the breakup mode remains unchanged. The channel confinement has a large effect on the
drop breakup dynamics, particularly in Mode I. We have seen that this mode involve pinching at
the mid-point, forming two equal daughter drops under low confinement (Fig. 12(a)). Increasing
confinement (Wc ) promotes pinching over a wide zone in the central part, instead of sharp pinching
at the mid-point in case of low confinement. Mode I under high Wc thereby produces a number
of satellite droplets between large drops (Fig. 12(b)). These smaller satellite droplets multiply in
number with increasing Wc . Secondly, their size contrast with the principal droplets is reduced. In
overall, Mode I undergoes a transition to Mode III with increasing confinement (Fig. 12(c)).
The three breakup modes discussed above give rise to contrasting drop size distributions (DSD)
in multiphase flows (Fig. 13). Mode I breakup results in varying DSD depending on the channel
confinement. The distribution is nearly unimodal for low confinement, but tends to be bimodal (0.05
and 0.65) when the confinement is large (Fig. 13(a)). The peak value of small-size fraction results
from a larger production of satellite droplets of a smaller size between larger drops under large
confinement. Mode II develops typically bi-modal size distributions with two frequency peaks of
contrasting drop sizes (0.08 and 0.55); the smaller size fraction peak far exceeds the peak at a larger
size fraction (Fig. 13(b)). In contrast, Mode III involves homogeneous fragmentation, producing
multiple droplets of similar sizes. After large amounts of shear the droplets occur in isolation,
homogeneously distributed in the matrix fluid. Their DSD displays a narrow spread of size distribu-
tion, characterized by a well-defined unimodal frequency (Fig. 13(c)). Our estimate shows that the
most frequent drop size is around 0.1a, where a is the initial drop diameter.

C. Streamline patterns
Drops perturb confined Couette flows with characteristic streamline patterns. Such streamlines
are generally transient, varying systematically with ongoing drop breakup processes. This section
presents an analysis of the streamlines to show the mode of interfacial instabilities,38 leading to
breakup in the three modes discussed in Sec. III B. The flow perturbations onset with a pair of
symmetrical vortices with their centers located along the shear-parallel central line, where the
073302-14 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

FIG. 13. Histograms of the characteristic drop size distributions (DSD) corresponding to three breakup modes. (a) Mode I: a
bimodal distribution with the dominant frequency peak at a normalized drop size (a ∗) of 0.6 (nearly half the initial drop size,
as expected) and a minor peak of relatively much smaller drop size (a ∗ = 0.05). (b) Mode II: a bimodal DSD, but showing an
inversion of the peak locations, where the dominant peak lies at the finer size fraction. (c) Mode III: a unimodal DSD with
the peak at 0.1, suggesting an equilibrium size of the secondary droplets.

vortices undergo a transformation with progressive shear, leading to complex streamline patterns
depending upon the breakup modes involved in drop deformations (Fig. 14). In case of Mode I,
the paired vortices are replaced by single vortices inside the drops as the latter are deformed into
elongate shapes (Fig. 14(a)). The flows subsequently turn to become unstable, producing secondary
vortex formation in the central zone of the deformed drops. This transformation of streamlines is a
proxy to the concomitant interfacial instability, which grows at the highest rates in the mid points,
073302-15 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

FIG. 14. Streamline patterns of the multiphase flows associated with drop deformation and subsequent breakup in progressive
shear. (a) Mode I: Formation of flow vortices during drop deformations, and split of a single vortex into multiple secondary
smaller vortices following the interfacial instability growth dominantly in the midpoint, and finally their coalescence into
single vortices after the breakup. (b) Mode II: Growth of interfacial instability preferentially at the drop edges, leading to a
number of secondary vortices. (c) Mode III: Entire breakup of a single vortex into a train of secondary vortices.
073302-16 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

forcing the drop to split into two daughter drops. The instability, however, leaves the drop edges
unaffected, as revealed from the two regular peripheral vortices. Each daughter drop has a pair of
vortices; but they merge to form a single vortex as the daughter drops tend to relax and stabilize
their elongate shapes after the mid-point breakup event. The vortex streamlines finally conform to
the deformed drop geometry. Mode II breakup involves transformation of the streamline patterns in
a different fashion. With progressive elongation drops develop an interfacial mechanical instability
preferentially at their edges, leaving the central zone virtually unaffected (Fig. 14(b)). The single
vortex within the drop becomes unstable in the peripheral region and eventually gives rise to a
number of secondary vortices at the drop edges. On the other hand the interfacial instability growth,
coupled with secondary vortex generation leads to a breakup of the drop into a train of smaller
droplets by Mode II. After this edge breakup event the primary drop relaxes its elongate shape with
a concomitant change in the streamline pattern. This relaxation arrests any interfacial instability
to occur with progressive shear. The streamlines show a similar transformation from a single to
a multiple vortex patterns in the Mode III breakup (Fig. 14(c)). However, the secondary vortex
is widespread, covering the entire elongate drop length. The interfacial instability develops at the
edge, but propagates towards the central region, leading to a homogeneous breakup of the drop in
the course of elongation.

D. Breakup fields
We integrate results from a large number of simulations to construct a field diagram defined by
the capillary number (Ca) and viscosity ratio (λ) and show the specific hydrodynamic conditions
required for the three breakup modes (Fig. 15).
According to our simulations results, the critical capillary number required for drop breakup
decreases with increasing viscosity ratio, as reported from earlier experiments (Bentley and Leal10).
The experimental results show that the critical Ca value remains low (∼0.1) even the viscosity ratio
is increased to a large value (∼100). Our calculated critical Ca approaches a value of the same order
as the λ > 5. The channel confinement is a crucial factor in determining the Cac versus viscosity
ratio variation. Our simulations suggest that a strong confinement condition can cause drop breakup
at low Cac even when the viscosity ratio is large. Previous experimental studies by Vannroye et al.39
have demonstrated that Cac can increase with viscosity ratio when the confinement is low; but
this increase does not occur when the confinement is strong. Their study also reveals that, under
high confinement conditions Cac becomes extremely low (0.1), which do not increase significantly

FIG. 15. A field diagram of the three breakup modes in a Ca versus λ space. Note that Mode I (midpoint pinching) occurs
in a narrow field where the capillary number is high, but the viscosity ratio is low, whereas Mode III can be active in a wide
range of Ca and λ. The curve fits of the transition lines are based on threshold Ca values computed from simulation runs
(Wc = 0.5).
073302-17 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

even the viscosity ratio is large. Our simulation results are thus consistent with these experimental
findings.
The diagram suggests that the breakup process onsets at a critical capillary number (Cac ),
which exponentially drops with increasing λ. For example, a drop can break at Cac = 0.47 for
λ = 2, whereas for viscosity contrast λ = 5, Cac = 0.2. This variation is in agreement with the
experimental data reported by previous workers.40 Increasing viscosity ratio causes the non-breakup
field to narrow down and thereby promotes the breakup process. There is a large field of Mode III
breakup, which is separated from the non-breakup field by a narrow band, delineating the fields for
Mode I and II. The plot shows that drops undergo homogeneous breakup (Mode III) for high Ca
or large λ values. Mid-point pinching (Mode I) occurs preferentially in low λ conditions (λ < 2.5),
which is replaced by Mode III and II with increasing Ca and λ, respectively. The Mode II (edge
breakup) field widens as λ becomes large. To summarize, Mode I is the dominant breakup mecha-
nism in conditions with low viscosity ratios, whereas Mode II in conditions of large viscosity ratios.
On the other end, Mode III breakup is the dominant mechanism in conditions of low viscosity
ratios, but with a high capillary number or vice versa.

IV. CONCLUSIONS
The conservative level set (CLS) method can track a two-phase fluid interface with great accu-
racy, restoring a good mass conservation property. We use the CLS method to characterize drop
deformations in confined shear flows. For a given flow confinement, drops undergo deformation
without breakup unless the capillary number (Ca) and the viscosity ratio (λ) exceed threshold
values. Under sub-critical Ca, deforming drops tend to stabilize both their shapes and orientations
with progressive shearing in a steady state only when λ < 3 and Ca < 0.3. Otherwise, they show
oscillatory deformations before attaining their stable geometry. For large values of Ca (>0.4) or λ
(>3) their ongoing deformations remain unsteady even after large finite shear flow. The stable Tay-
lor factor (i.e., degree of flattening) increases nonlinearly with increasing λ. On other hand, increas-
ing Ca lowers the stable angle of deformed drops to the shear direction. The flow confinement factor
(Wc ) has little effects on the steadiness of drop deformation, albeit Wc greatly influences the stable
drop geometry. In this study we show three possible modes of drop breakup: Mode I (mid-point
pinching), Mode II (edge breakup), and Mode III (homogeneous breakup), each with characteristic
drop size distribution (DSD). Mode I produces two equal smaller drops, often associated with a
few droplets of much smaller size between the daughter drops. This mode leads to a DSD with
a unimodal, or a bi-modal, pattern, but with a weak peak for the smaller size fraction. In case of
Mode II, the DSD shows a well-defined bimodal pattern with a large contrast in the modal values.
Mode III breakup again gives rise to a unimodal DSD; however, the distribution is dominated by
small droplets. Increasing capillary number results in a Mode I to III transition, whereas for large
viscosity ratio Mode II is the dominant breakup modes. The mid-point pinching occurs on a broad
zone to produce smaller droplets in the central zone, and the break process occurs practically by
Mode III under a high confinement condition.

ACKNOWLEDGMENTS
We thank two anonymous reviewers for their insightful comments on this work. This work was
carried out under a research project supported by the SERB, Department of Sciences and Technol-
ogy, India. N.M. and N.B. acknowledge the Department of Sciences and Technology for awarding
them with the J. C. Bose Fellowship and the Inspire Fellowship, respectively.
1 G. I. Taylor, “The viscosity of a fluid containing small drops of another fluid,” Proc. R. Soc. A 138, 41–48 (1932).
2 G. I. Taylor, “The formation of emulsions in definable fields of flow,” Proc. R. Soc. A 146, 501–523 (1934).
3 J. M. Rallison, “The deformation of small viscous drops and bubbles in the shear flows,” Annu. Rev. Fluid Mech. 16, 45

(1984).
4 H. A. Stone, “Dynamics of drop deformation and breakup in viscous fluids,” Annu. Rev. Fluid Mech. 26, 65 (1994).
5 S. Torza, R. G. Cox, and S. G. Mason, “Particle motions in sheared suspensions. XXVII transient and steady deformation

and burst of liquid drops,” J. Colloid Interface Sci. 38, 395–411 (1972).
073302-18 N. Barai and N. Mandal Phys. Fluids 28, 073302 (2016)

6 H. P. Grace, “Dispersion phenomena in high viscosity immiscible fluid systems and application of static mixers as dispersion
devices in such systems,” Chem. Eng. Commun. 14, 225–277 (1982).
7 C. R. Marks, “Drop breakup and deformation in sudden onset strong flows,” Ph.D. thesis, University of Maryland, College

Park, 1998.
8 C. L. Tucker and P. Moldenaers, “Microstructural evolution in polymer blends,” Annu. Rev. Fluid Mech. 34, 177–210 (2002).
9 X. Zhao, “Drop breakup in dilute Newtonian emulsions in simple shear flow: New drop breakup mechanisms,” J. Rheol.

51, 367–392 (2007).


10 B. J. Bentley and L. G. Leal, “An experimental investigation of drop deformation and breakup in steady, two-dimensional

linear flows,” J. Fluid Mech. 167, 241–283 (1986).


11 S. Unverdi and G. Tryggvason, “A front-tracking method for viscous, incompressible, multifluid flows,” J. Comput. Phys.

100, 25–37 (1992).


12 G. Tryggvason, B. Bunner, A. Esmaeeli, D. Juric, N. Al-Rawahi, W. Tauber, J. Han, S. Nas, and Y. J. Jan, “A front-tracking

method for the computations of multiphase flow,” J. Comput. Phys. 169, 708–759 (2001).
13 R. DeBar, “Fundamentals of the KRAKEN code,” Technical Report UCIR-760, LLNL, 1974.
14 D. L. Youngs, Time-Dependent Multi-Material Flow with Large Fluid Distortion in Numerical Methods for Fluid Dynamics

(Academic, New York, 1982), p. 27.


15 J. Li, “Numerical resolution of Navier-Stokes equation with reconnection of interfaces. Volume tracking and application to

atomization,” Ph.D. thesis, University of Paris VI, 1996.


16 U. Dutta, S. Sarkar, and N. Mandal, “Ballooning versus curling of mantle plumes: Views from numerical models,” Curr.

Sci. 104, 893–903 (2013).


17 U. Dutta, S. Sarkar, A. Baruah, and N. Mandal, “Ascent modes of jets and plumes in a stationary fluid of contrasting

viscosity,” Int. J. Multiphase Flow 63, 1–10 (2014).


18 R. Scardovelli and S. Zaleski, “Direct numerical simulation of free surface and interfacial flow,” Annu. Rev. Fluid Mech.

31, 567–603 (1999).


19 C. Hirt and B. Nichols, “Volume of fluid (VOF) method for the dynamics of free boundaries,” J. Comput. Phys. 39, 201–225

(1981).
20 S. Osher and J. A. Sethian, “Fronts propagating with curvature-dependent speed: Algorithms based on Hamilton–Jacobi

formulations,” J. Comput. Phys. 79, 175–210 (1988).


21 M. Sussman, P. Smereka, and S. Osher, “A level-set approach for computing solutions to incompressible two-phase flow,”

J. Comput. Phys. 114, 146–159 (1994).


22 M. Sussman, E. Fatemi, and P. Smereka, “An improved level set method for incompressible two-phase flows,” Comput.

Fluids 27, 663–680 (1998).


23 M. Sussman and E. Fatemi, “An efficient, interface-preserving level set redistancing algorithm and its application to inter-

facial incompressible fluid flow,” SIAM J. Sci. Comput. 20, 1165–1191 (1999).
24 D. Peng, B. Merriman, S. Osher, H. Zhao, and M. Kang, “A PDE-based fast local level-set method,” J. Comput. Phys. 155,

410–438 (1999).
25 J. A. Sethian, Level Set Methods and Fast Marching Methods (Cambridge University Press, Cambridge, 1999).
26 M. Quecedo and M. Pastor, “Application of the level set method to the finite element solution of two-phase flows,” Int. J.

Numer. Methods Eng. 30, 645–664 (2001).


27 M. I. Herreros, M. Mabssout, and M. Pastor, “Application of level-set approach to moving interfaces and free surface

problems in flow through porous media,” Comput. Methods Appl. Mech. Eng. 195, 1–25 (2006).
28 T. H. WSheu, C. H. Yu, and P. H. Chiu, “Development of level set method with good area preservation to predict interface

in two-phase flows,” Int. J. Numer. Methods Fluids 67, 109–134 (2011).


29 S. P. van der Pijl, A. Segal, C. Vuik, and P. Wesseling, “A mass-conserving level-set method for modelling of multi-phase

flows,” Int. J. Numer. Methods Fluids 47, 339–361 (2005).


30 L. Zhao, X. Bai, T. Li, and J. J. R. Williams, “Improved conservative level set method,” Int. J. Numer. Methods Fluids 75,

575–590 (2014).
31 O. Desjardins, V. Moureau, and H. Pitsch, “An accurate conservative level set/ghost fluid method for simulating turbulent

atomization,” J. Comput. Phys. 227, 8395–8416 (2008).


32 E. Olsson and G. Kreiss, “A conservative level set method for two phase flow,” J. Comput. Phys. 210, 225–246 (2005).
33 E. Olsson, G. Kreiss, and S. Zahedi, “A conservative level set method for two phase flow II,” J. Comput. Phys. 225, 785–807

(2007).
34 J. Li, Y. Y. Renardy, and M. Renardy, “Numerical simulation of breakup of a viscous drop in simple shear flow through a

volume-of-fluid method,” Phys. Fluids 12, 269 (2000).


35 V. Sibillo, G. Pasquariello, M. Simeone, V. Cristini, and S. Guido, “Drop deformation in microconfined shear flow,” Phys.

Rev. Lett. 97, 054502 (2006).


36 A. E. Komrakova, O. Shardt, D. Eskin, and J. J. Derksen, “Lattice Boltzmann simulations of drop deformation and breakup

in shear flow,” Int. J. Multiphase Flow 59, 24–43 (2014).


37 R. Croce, M. Griebel, and A. Schweitzer, “Numerical simulation of bubble and droplet deformation by a level set approach

with surface tension in three dimensions,” Int. J. Numer. Methods Fluids 62, 963–993 (2010).
38 A. De Wit, P. De Kepper, K. Benyaich, G. Dewel, and P. Borckmans, “Hydrodynamic instability of spatially extended

bistable chemical systems,” Chem. Eng. Sci. 58, 4823 (2003).


39 A. Vananroye, P. van Puyvelde, and P. Moldenaers, “Effect of confinement on droplet breakup in sheared emulsions,” Lang-

muir 22, 3972 (2006).


40 P. J. A. Janssen, A. Vananroye, P. Van Puyvelde, P. Moldenaers, and P. D. Anderson, “Generalized behavior of the breakup

of viscous drops in confinements,” J. Rheol. 54, 1047 (2010).

You might also like