Sie sind auf Seite 1von 12

Field-induced control of ferrofluid emulsion rheology and droplet break-up in shear

flows
Lucas H. P. Cunha, Ivan R. Siqueira, Taygoara F. Oliveira, and Hector D. Ceniceros

Citation: Physics of Fluids 30, 122110 (2018); doi: 10.1063/1.5055943


View online: https://doi.org/10.1063/1.5055943
View Table of Contents: http://aip.scitation.org/toc/phf/30/12
Published by the American Institute of Physics

Articles you may be interested in


Sliding behavior of droplet on a hydrophobic surface with hydrophilic cavities: A simulation study
Physics of Fluids 30, 122006 (2018); 10.1063/1.5063857

Transient evolution of the heat transfer and the vapor film thickness at the drop impact in the regime of film
boiling
Physics of Fluids 30, 122109 (2018); 10.1063/1.5059388

Deformation of a ferrofluid droplet in simple shear flows under uniform magnetic fields
Physics of Fluids 30, 092002 (2018); 10.1063/1.5047223

Dynamics of drop formation from submerged orifices under the influence of electric field
Physics of Fluids 30, 122104 (2018); 10.1063/1.5063913

Deformation of a surfactant-laden viscoelastic droplet in a uniaxial extensional flow


Physics of Fluids 30, 122108 (2018); 10.1063/1.5064278

Numerical investigation of the breakup behavior of an oscillating two-phase jet


Physics of Fluids 30, 072101 (2018); 10.1063/1.5029772
PHYSICS OF FLUIDS 30, 122110 (2018)

Field-induced control of ferrofluid emulsion rheology and droplet


break-up in shear flows
Lucas H. P. Cunha,1,2,a) Ivan R. Siqueira,2,b) Taygoara F. Oliveira,1,c)
and Hector D. Ceniceros3,d)
1 Department of Mechanical Engineering, Universidade de Brası́lia, Brası́lia, DF 70910-900, Brazil
2 Department of Chemical and Biomolecular Engineering, Rice University, Houston, Texas 77005, USA
3 Department of Mathematics, University of California Santa Barbara, Santa Barbara, California 93106, USA

(Received 12 September 2018; accepted 27 November 2018; published online 28 December 2018)

Ferrofluid droplets have been widely used in a number of cutting-edge applications in microfluidics,
biomedicine, and microrheology. In many cases, the droplet is simultaneously subjected to a hydro-
dynamic flow and an external magnetic field. However, the response of a ferrofluid droplet under
these forces in terms of deformation, inclination, and potential breakup into smaller droplets is not yet
fully understood. In this work, we present a numerical study of the dynamics of a two-dimensional
ferrofluid droplet suspended in a non-magnetic, immiscible liquid when the two-phase fluid undergoes
a simple shear flow under the action of an external, uniform magnetic field. The model consists of
magnetostatic Maxwell’s equations and the incompressible Navier-Stokes equations with additional
terms that take into account both magnetic and capillary forces on the droplet. The resulting system
of fully coupled, non-linear equations is accurately solved with the projection method together with
the level set method to capture the droplet interface. Our results show that the external magnetic field
strongly affects the droplet deformation and inclination relative to the flow. We investigate the effects
of the external field-induced droplet distortion on the viscosity of the resulting complex fluid when
the two-phase liquid is viewed as a dilute emulsion of ferrofluid droplets. Notably, the viscosity of
the ferrofluid emulsion can be either dramatically increased or decreased depending on the inten-
sity and direction of the external magnetic field. We also analyze for the first time the effects of the
external magnetic field on the breakup process of ferrofluid droplets undergoing large deformations.
Remarkably, the external magnetic field can be adjusted to control the droplet breakup process both
in terms of time to break up and size of satellite droplets. These new insights indicate the potential of
external magnetic fields as tunable tools to control the rheology of ferrofluid emulsions and topology
of ferrofluid droplets in shear flows. Published by AIP Publishing. https://doi.org/10.1063/1.5055943

I. INTRODUCTION immersed ferrofluid droplet have been proposed for measuring


the viscosity and surface tension of liquids.10,11 Thus, a fun-
Ferrofluid emulsions are two-phase mixtures consisting
damental understanding of the behavior of ferrofluid droplets
of ferrofluid droplets suspended in a non-magnetic, immis-
under the combined action of hydrodynamic flows and external
cible liquid. Their macroscopic behavior is dictated by com-
magnetic fields is essential for improving existing applications
plex physical phenomena that occur at the microscopic level,
and developing new technologies.
such as droplet deformation, rotation, and breakup, which,
Since the pioneering studies of Sherwood12 and Séro-
in turn, are ruled by the combined action of hydrodynamic
Guillaume et al.,13 the response of ferrofluid droplets to
and magnetic effects. As a consequence, suspensions of fer-
external magnetic fields has been substantially investigated.
rofluid droplets are complex fluids whose response under flow
Fundamental studies on the equilibrium shape of ferrofluid
can be controlled by external magnetic fields. As recently
droplets under applied fields14–18 and on the field-induced
reviewed by Torres-Dı́az and Rinaldi,1 there is a wide range of
motion of ferrofluid droplets suspended in non-magnetic liq-
applications supporting the growing research on the behavior
uids19–21 abound. It has also been shown that external magnetic
of ferrofluid droplets as magnetically controllable fluids. For
fields play an important role in formation and breakup of
instance, their interaction with external magnetic fields has
ferrofluid droplets in microfluidic systems,22–26 viscous fin-
been extensively explored in microfluidics2,3 and biomedical
gering phenomena in Hele-Shaw cells,27,28 self-assembly of
processes, such as targeted drug delivery4–6 and restorative
ferrofluid droplets into ordered structures,29 and spreading of
treatment of retinal detachment.7–9 Recently, new microrhe-
ferrofluid droplets.30,31
ology techniques based on the field-induced motion of an
Notwithstanding, most of the research to date has centered
on the interaction of ferrofluid droplets and magnetic fields in
a) E-mail: lh36@rice.edu
b) E-mail:
a quiescent flow. One exception is the recent numerical work
irs@rice.edu
c) E-mail: taygoara@unb.br presented by Jesus, Roma, and Ceniceros32 on the dynamics
d) E-mail: hdc@math.ucsb.edu of a three-dimensional ferrofluid droplet in shear flows under

1070-6631/2018/30(12)/122110/11/$30.00 30, 122110-1 Published by AIP Publishing.


122110-2 Cunha et al. Phys. Fluids 30, 122110 (2018)

external magnetic fields. However, due to the high computa- neutrally buoyant, ferrofluid droplet. The continuous phase has
tional cost of the simulations, the authors did not address large viscosity η and magnetic permeability µ0 , which is assumed
droplet deformations and breakup conditions. Other studies to be equal to the magnetic permeability of the free space (that
that deserve special attention are the ones presented by Hassan, is, µ0 = 4π × 10−7 H · m−1 ). The ferrofluid droplet, which
Zhang, and Wang33 and Capobianchi, Lappa, and Oliveira,34 is initially circular with radius a and centered at the channel
which focused on the deformation, inclination, lateral migra- center, has viscosity λη and magnetic permeability ζ µ0 . Both
tion, and transient relaxation of ferrofluid droplets in simple phases are assumed to be Newtonian and incompressible, and
shear flows under the influence of external magnetic fields. the two-phase fluid can be seen as a dilute emulsion of fer-
However, the effects of the external field on the viscosity of rofluid drops. The system is subjected to a simple shear flow
the resulting ferrofluid emulsion and droplet breakup process with shear rate γ̇ caused by the movement of the channel walls
due to large deformations were not explored either. in opposite directions. Notice that the flow is symmetric with
In this work, we present a detailed numerical study on the respect to the droplet initial position and the velocity along
response of a two-dimensional ferrofluid droplet suspended in the channel centerline is zero. Furthermore, the system is also
a non-magnetic, immiscible liquid when the system undergoes subjected to an external, uniform external magnetic field H0
the combined action of a simple shear flow and an external, uni- that can be either parallel or perpendicular to the main flow
form magnetic field. We are able to explore large deformations direction (that is, the x-direction). Figure 1 shows a sketch of
and droplet breakup by employing a highly accurate interface the problem in which the external magnetic field is parallel
capturing method to compute the topological changes of the to the flow direction; the ferrofluid droplet is confined in the
droplet surface when the rupture occurs. The mathematical region Ωi (bounded by Γi ) and the continuous phase occupies
modeling couples magnetostatic Maxwell’s equations and the the region Ωo (bounded by Γo ).
incompressible Navier-Stokes equations with additional terms
related to magnetic and capillary forces. The model is solved B. Governing equations for the magnetic
with the project method together with the level set method to and hydrodynamic problems
capture the dynamics of the droplet interface. Our results indi- The applied magnetic field affects the magnetization M
cate that the magnetic field can drastically affect the droplet of the ferrofluid, which, in turn, contributes to the magnetic
deformation and inclination in the flow, which, in turn, play a induction as B = µ0 (H + M), where H is the magnetic field. In
key role in the effective viscosity of the resulting dilute emul- the absence of electric currents, H and B satisfy magnetostatic
sion of ferrofluid droplets. We also show that the intensity and Maxwell’s equations, ∇ · B = 0 and ∇ × H = 0. We fur-
direction of the external magnetic field can be adjusted to con- ther assume that the ferrofluid is superparamagnetic such that
trol the droplet breakup process both in terms of breakup time M = χH, where χ is the magnetic susceptibility. This assump-
and size of the satellite droplet. tion leads to B = µ0 ζH, where µ0 ζ = µ0 (1 + χ) is the magnetic
The remainder of this article is organized as follows. In permeability of the dispersed ferrofluid phase. Since H is an
Sec. II, we present the mathematical modeling of the problem, irrotational field, there is a scalar potential ψ that satisfies
including the governing equations and dimensionless parame- H = −∇ψ. Then, the magnetic problem is reduced to
ters. Section III summarizes the numerical methodology used  
in the simulations. The numerical results are presented and ∇ · µ0 ζ(x) ∇ψ = 0, (1)
discussed in Sec. IV, and concluding remarks are given in
Sec. V. Further details of the numerical method are given in where x is the position vector. In Eq. (1), the permeability ratio
the Appendix. ζ = ζ(x) is extended to the entire fluid domain, being equal to
the unit for the continuous phase.
The two-phase fluid flow is governed by the incompress-
II. MATHEMATICAL MODEL ible Navier-Stokes equations as follows:
A. Problem formulation ∇ · u = 0, (2)
Consider a channel between two parallel plates filled Du f  g
with a non-magnetizable liquid in which there is a suspended, ρ = −∇p + ∇ · λ(x)η ∇u + ∇uT + Fm + Fs , (3)
Dt

FIG. 1. Sketch of the problem (not to scale).


122110-3 Cunha et al. Phys. Fluids 30, 122110 (2018)

where u is the velocity field, ρ is the density, t is the time, is defined as36 D E
p is the pressure field, Fm is the magnetic force (density) on σxy − η γ̇
the ferrofluid phase, and Fs is the capillary force (density) on [η] = , (6)
η γ̇ β
the droplet interface. Notice that the viscosity ratio λ = λ(x) where hσ xy i is the volume-averaged shear stress on the moving
is also extended to the entire domain, being equal to one walls and β is the volume fraction of the dispersed phase.
for the continuous phase. The magnetic force is given by35
Fm = µ0 M · ∇H = µ0 (ζ(x) − 1)H · ∇H, and the capillary C. Level set formulation for droplet dynamics
force is defined as Fs = −σκδs n, where σ is the interfacial
Due to the droplet shape, position, and even topology
tension, κ is the curvature, δs is a Dirac delta distribution on
change in time, we are dealing with a free-boundary prob-
the liquid-liquid interface, and n is the unit normal vector out-
lem. Here, we capture the droplet dynamics with the level
ward the droplet surface. Notice that the interaction between
set method.37 The interface between the liquids is represented
the droplet and the basis liquid is purely hydrodynamic; that
at all times as the zero-level set of a scalar-valued function
is, except for capillary effects, there is no additional stress
φ(x, t) such that, for each time t, the droplet surface is the
induced by the ferrofluid phase in the absence of an external
set of points x such that φ(x, t) = 0. The level set function is
magnetic field.
advected by the flow field according to
We non-dimensionalize the governing equations by using
the following dimensionless variables: u∗ = u/γ̇a, t ∗ = t γ̇, ∂φ
+ u · ∇φ = 0. (7)
x∗ = x/a, p∗ = p/(ρa2 γ̇ 2 ), and H∗ = H/||H0 ||. Dropping the ∂t
superscript ∗ to alleviate the notation, the dimensionless form The level set function also serves as an indicator or maker
of Eqs. (2) and (3) becomes function to evaluate the piecewise constant λ(x) and ζ(x) such
that the discontinuities across the liquid-liquid interface are
∇ · u = 0, (4)
replaced by smooth transitions occurring in a thin region of
size ε. To this end, we consider a mollified Heaviside function
Du 1 f  g
= −∇p + ∇ · λ(x) ∇u + ∇uT defined as
Dt Re
Camag 1  0, if φ < −ε
κδs n.

+ (ζ(x) − 1)H · ∇H − (5)


φ 1  πφ 
 " #
Ca Re Ca Re 1


Hε (φ) =  1 + + sin , if |φ| ≤ ε . (8)
Three dimensionless numbers appear in the formulation: (i) 


 2 ε π ε
the Reynolds number, Re = ργ̇a2 /η, which represents the

if φ > ε

 1,

ratio of inertial to viscous forces; (ii) the capillary number,
Ca = ηaγ̇/σ, which denotes the ratio of viscous to capil- The piecewise constants λ(x) and ζ(x) are then replaced by
lary forces; and (iii) the magnetic capillary number, Camag the smooth functions λ ε (φ) = λ + (1 − λ)H ε (φ) and ζ ε (φ)
= µ0a ||H0 ||2 /σ, which represents the ratio of magnetic to cap- = ζ + (1 − ζ)H ε (φ). The level set function is also used to
illary forces. Notice that the magnetic capillary number plays define a mollified Dirac delta distribution,
the same role as the magnetic Bond number recently used by 0, if |φ| > ε
∂Hε (φ) 


Jesus et al.32 and Hassan et al.33 to study the dynamics of fer- δ ε (φ) = =

1 πφ , (9)
∂φ
  
rofluid droplets in shear flows under external magnetic fields. 1 + cos , if |φ| ≤ ε


ε

 2ε
These three dimensionless groups together with the viscosity
and magnetic permeability ratios, λ and ζ, respectively, are as well as the unit normal vector outward the droplet surface,
the main parameters of the model. For the reader interested n = ∇φ/|∇φ|, and the droplet curvature, κ = ∇ · (∇φ/|∇φ|).
in typical dimension values for the fluid properties, external Those definitions are then used to couple the magnetic and
field intensity, and droplet size used in experiments and prac- hydrodynamic governing equations with the level set function,
tical applications, we suggest the review by Torres-Dı́az and leading to a model valid throughout the entire domain of the
Rinaldi1 and the references therein. two-phase fluid flow.
Boundary conditions are needed to solve Eqs. (1), (4), and
(5) as follows: (i) for the magnetic potential field, we impose III. NUMERICAL METHODOLOGY
a constant flux condition at the four boundaries to obtain an
A. Numerical scheme
external, uniform magnetic field; notice that these fluxes can
be zero or non-dependent on the direction of the external field, The numerical methodology employed in this work con-
which can be either parallel or perpendicular to the main flow sists of three main components: (i) a second-order method for
direction; (ii) for the velocity field, we consider the no-slip the elliptic problem governing the magnetic potential field, ψ;
condition at the channel walls together with periodic bound- (ii) a second-order non-stiff projection method for the hydro-
ary conditions in the flow direction; and (iii) for the pressure dynamic problem governing the velocity and pressure fields,
field, we apply the no-flux condition at the channel walls and u and p; and (iii) a high-order Total Variation Diminishing
periodic boundary conditions in the flow direction. (TVD) scheme to advance the level set function, φ. A uniform
Finally, it is worth mentioning that the two-phase system staggered grid is used for the discretization of the problem
can be seen as a dilute emulsion of ferrofluid droplets. To ana- domain.
lyze the effect of the external magnetic field on the emulsion A standard second-order finite difference is used for the
rheology, we consider the emulsion reduced viscosity, which spatial discretization of Eqs. (1), (4), and (5). The elliptic
122110-4 Cunha et al. Phys. Fluids 30, 122110 (2018)

initialized as a signed distance to the droplet surface. However,


a well-known difficulty associated with the level set method is
the progressive distortion of the level set function as the prob-
lem evolves. Because not all fluid particles move with the same
velocity as the interface velocity, after some time, φ(x, t) no
longer represents a signed distance function to the droplet sur-
face, i.e., |∇φ| , 1, which leads to inaccurate representations
of the droplet interface and its associated geometric quanti-
ties (e.g., normal and tangent vectors and curvature).41,42 To
overcome this issue, we employ the reinitialization method of
Sussman, Smereka, and Osher.43
In all simulations, we set ε = 1.5h for the interfacial
thickness, where h is the mesh size. p The time step is chosen
to satisfy the condition ∆t = 0.05h/ 2(fcm + fcc ), where fcm
FIG. 2. Comparison with the experimental results of Flament et al.14 for the = Camag (ζ − 1)/CaRe is a characteristic magnetic force and
droplet equilibrium form factor, e, as a function of magnetic capillary number, fcc = 1/CaRe is a characteristic capillary force. Formally, the
Camag . Simulations performed with λ = 1 and ζ = 3.2.
overall method is second-order accurate in space and time, but
due to the regularization of discontinuous material quantities
problem for the magnetic potential field is solved efficiently (effectively a diffuse interface approach), the accuracy near
with a preconditioned conjugate gradient method. The fluid the interface is expected to be only first order. A mesh conver-
flow problem is solved with a splitting projection method38 gence test for the overall numerical scheme is presented in the
with the semi-implicit modification introduced by Badalassi, Appendix.
Ceniceros, and Banerjee,39 such that
3u∗ − 4un + un−1 λ 2 ∗ B. Validation
= ∇ u + G(û, φn , Hn ) + Fs (φn ), (10)
2∆t Re
We first validate the model and numerical methodology
3 with two benchmark problems. First, we consider the deforma-
(un+1 − u∗ ) = −∇pn+1 , (11)
2∆t tion of a ferrofluid droplet under an external, uniform magnetic
where field in a quiescent flow. The droplet equilibrium shape, which
λ 2 is approximately ellipsoidal, is dictated by a balance between
G(û, φn , Hn ) = −û · ∇û − ∇ û magnetic and capillary forces. The experiments of Flament
Re
1   et al.14 in a Hele-Shaw cell provide an appropriate set of data
+ ∇ · λ ε (φn )(∇û + ∇ûT ) for comparison with our two-dimensional numerical predic-
Re
Camag tions. We use λ = 1 and ζ = 3.2 in a 12 × 12 square domain
+ (ζ ε (φn ) − 1)Hn · ∇Hn , (12) discretized with 256 × 256 cells. Figure 2 shows a compar-
Ca Re
ison of the experimental results of Flament et al.14 with our
∇φn ∇φn
!
1 numerical predictions for the equilibrium droplet form factor e,
Fs (φn ) = − ∇· n
δ ε (φn ) , (13)
Ca Re |∇φ | |∇φn | defined here as the ratio of the length of minor and major axes,
λ = max{1, λ ε (φn )}, and û = 2un − un−1 . The pressure pn+1 respectively, as a function of Camag . As can be seen, there is an
3 excellent agreement for the general physical trend. We observe
is then obtained by solving ∇2 pn+1 = ∇ · u∗ . an average deviation of only 4.66% for the full range of Camag
2∆t
Finally, the advection of the level set function given in explored here. This small difference might be due to experi-
Eq. (7) is integrated with a combined third-order TVD Runge- mental factors and/or to the two-dimensional assumption of the
Kutta method for the temporal discretization and a fifth-order Hele-Shaw cell flow. Note that the droplet equilibrium shape is
Weighted Essentially Non-Oscillator (WENO) scheme for the also well predicted by the numerical simulations, as displayed
computation of the spatial derivative.40 The level set function is in Fig. 3.

FIG. 3. Comparison of the ferrofluid droplet equilibrium


shape obtained experimentally by Flament et al.14 (black
and white background) and our numerical predictions
(yellow contour) for (a) Camag = 0.5, (b) Camag = 1.8,
(c) Camag = 2.7, and (d) Camag = 4.3. Simulations per-
formed with λ = 1 and ζ = 3.2. Reprinted with permission
from Flament et al., “Measurements of ferrofluid sur-
face tension in confined geometry,” Phys. Rev. E 53,
4801–4806 (1996). Copyright 1996 American Physical
Society.
122110-5 Cunha et al. Phys. Fluids 30, 122110 (2018)

FIG. 4. Comparison with the phase field results of Ghigliotti et al.44 for (a) droplet inclination, θ, and (b) reduced viscosity, [η], as functions of the viscosity
ratio, λ. Simulations performed with Re = 0.01 and Ca = 0.3.

Second, we consider a droplet undergoing a simple shear model differences, we note that there is a good agreement in
flow in the absence of an external magnetic field. We have the droplet equilibrium shape, as demonstrated in Fig. 5 for
compared our numerical results with those of Ghigliotti, λ = 2.
Biben, and Misbah,44 who analyzed a similar problem using a
phase field (PF) model under the Stokes flow regime (that is,
Re = 0). The shear flow deforms and rotates the droplet such IV. RESULTS AND DISCUSSIONS
that its equilibrium shape is usually not spherical and not A. Droplet inclination and ferrofluid emulsion
aligned with the flow direction. Figure 4 shows the droplet viscosity under external magnetic fields
inclination angle relative to the main flow direction, θ, and
As recently elucidated by Jesus et al.32 and Hassan et al.,33
the emulsion reduced viscosity, [η], as functions of the viscos-
external magnetic fields can significantly affect the deforma-
ity ratio, λ, for Re = 0.01 and Ca = 0.3. Although we see a
tion and inclination of ferrofluid droplets in shear flows. This,
good qualitative agreement between our predictions and those
in turn, has a direct effect on the reduced viscosity of the two-
of Ghigliotti et al.,44 there is a relative difference of 4.8%
phase fluid when it is seen as a dilute emulsion of ferrofluid
for the reduced viscosity and about 2.2% for the inclination
droplets. Here, we analyze this phenomenon using the follow-
angle. This quantitative discrepancy can be attributed to iner-
ing procedure. First, we compute the steady state response of
tial effects—our simulations were performed with Re = 0.01
a ferrofluid droplet under shear in the absence of an external
and the PF model of Ghigliotti et al.44 is for Re = 0—and,
magnetic field. Then, we turn on the magnetic field and com-
more importantly, to the dependence of the capillary number
pute the new droplet equilibrium shape and the corresponding
on the interface thickness in the PF model.45 Despite these
reduced viscosity. In this study, we fix Re = 0.01, Ca = 0.1,
λ = 2, and ζ = 2, and consider external magnetic fields parallel
and perpendicular to the main flow direction. With no applied
field, 2θ/π = 0.41, [η] = 1.66, and the droplet does not break.
All simulations were performed in a 12 × 12 square domain
discretized with 400 × 400 cells.
Figure 6 shows the droplet inclination and reduced vis-
cosity as functions of the external field intensity (expressed in
terms of Camag ). As the field intensity increases, the droplet
major axis tends to become more aligned with the external
field direction. Thus, when the external magnetic field is par-
allel to the flow direction, θ decreases with Camag and tends
asymptotically to zero for sufficiently large Camag . As the
droplet gets more aligned with the flow, the streamlines get
less distorted around the droplet surface, as shown in Fig. 7(a).
As a consequence, [η] decreases with Camag . For instance, at
Camag = 20, [η] = 0.62, which represents a decrease of 63%
FIG. 5. Comparison of the ferrofluid droplet equilibrium shape obtained with relative to the case where there is no applied field, and the incli-
the Stokes-PF model of Ghigliotti et al.44 (black and white background with nation is 2θ/π = 0.002. On the other hand, when the external
the velocity field) and our numerical predictions (red contour). Simulations field is perpendicular to the flow direction, θ increases with
performed with Re = 0.01, Ca = 0.3, and λ = 2. Reproduced with permission
from Ghigliotti et al., “Rheology of a dilute two-dimensional suspension of
Camag , as expected. However, as the field induces a droplet
vesicles,” J. Fluid Mech. 653, 489–518 (2010). Copyright 2010 Cambridge deformation in the vertical direction, it increases the shear
University Press. effects which tend to rotate the droplet back to the extensional
122110-6 Cunha et al. Phys. Fluids 30, 122110 (2018)

FIG. 6. Effects of the external magnetic field on the (a) droplet inclination, θ, and (b) reduced viscosity, [η], as a function of the magnetic capillary number,
Camag . Simulations performed with Re = 0.01, Ca = 0.1, λ = 2, and ζ = 2.

quadrant of shear. The final droplet equilibrium shape depends droplet equilibrium shape for different values of Camag when
critically on this competition between field-induced stretch- the external field is parallel and perpendicular to the flow
ing and shear stress. The misalignment of the droplet with direction.
the flow requires that the streamlines deflect significantly to
conform to the droplet shape, as shown in Fig. 7(b), which B. Field-induced control of droplet breakup
leads to a higher reduced viscosity. Therefore, [η] strongly
increases with Camag when the applied field is perpendicu- External force fields offer the possibility to control emul-
lar to the flow. At Camag = 20, [η] = 9.04, which represents sion rheology in shear flows by inducing or hindering topo-
an increase of almost one order of magnitude with respect logical changes in the suspended droplets. Recent studies
to the zero-field case, and 2θ/π = 0.91. Figure 8 shows the have investigated this effect for emulsions in which the
droplets undergo small deformations under external electrical
fields.46–48 Here, we focus on the effects of external magnetic
fields on the rupture of a ferrofluid droplet in shear flows and
do not restrict our study to small deformations.
First, we consider a case where the sheared droplet breaks
up in the absence of an external magnetic field (that is, Camag
= 0) and examine how an external field, parallel or perpendic-
ular to the flow direction, affects the rupture. We take Re = 1,
Ca = 0.5, λ = 1.2, and ζ = 2 for this study. The channel is a
12 × 6 rectangle discretized with 300 × 150 cells. Figure 9
shows snapshots of the droplet deformation and its eventual

FIG. 7. Streamlines of the flow for Camag = 10 when the external magnetic
field is (a) parallel and (b) perpendicular to the flow direction.

FIG. 8. Droplet equilibrium shape for Camag = 0, 2.5, 5, 7.5, and 10 when FIG. 9. Evolution of the droplet shape in the absence of an external magnetic
the external magnetic field is (a) parallel and (b) perpendicular to the flow field (Camag = 0). Simulations performed with Re = 1, Ca = 0.5, λ = 1.2, and
direction. ζ = 2.
122110-7 Cunha et al. Phys. Fluids 30, 122110 (2018)

rupture into three smaller, daughter droplets. The smallest −1 − Ca∗−1 )0.359 .
the satellite droplet size scales like Ss ∼ (Camag mag
daughter droplet is usually referred to as satellite droplet, being The external field increases the droplet deformation in the flow
very common in the droplet breakup process. direction and contributes to its rotation from the extensional to
We first analyze the effect of an external magnetic field the compressional quadrant of shear. This delays the rupture
parallel to the flow direction. Snapshots of the droplet evo- process and reduces the amount of liquid in the droplet neck,
lution for Camag = 2, 4, 8, and 12 appear in Fig. 10. As the which rips up and gives rise to the satellite droplet. It is worth
intensity of the external field increases, the breakup is delayed mentioning that the critical magnetic capillary number Camag ∗

and eventually arrested for large enough Camag . Figure 11(a) is likely dependent on the dimensionless parameters Re, Ca,
displays the estimated rupture time, τ b , as a function of Camag −1
λ, and ζ. However, the study of this four-dimensional phase
and clearly indicates that the breakup time increases nonlin- space on the breakup dynamics is a challenging problem and
early with the magnetic field intensity. This result also indicates will be addressed in a future study.
that there is a critical magnetic capillary number Camag ∗ above A strikingly different behavior takes place when the
which rupture does not occur. A power-law fitting suggests applied magnetic field is perpendicular to the flow direction.

that Camag ≈ 7.15 and the droplet breakup time scales like Figure 12 shows snapshots of the droplet shape evolution for
τb ∼ (Camag − Camag
−1 ∗−1 )−0.245 . For Ca
mag > Camag , the droplet is
∗ this case. For instance, at Camag = 2, the time to rupture and
so aligned with the flow direction that the shear stress acting on the size of the satellite droplet increase relative to the corre-
its surface is not strong enough to induce rupture. We also note sponding values for the case where Camag = 0, as shown in
that for Camag < Camag ∗ the size of the satellite droplet relative Fig. 12(a). At an intermediate field intensity, Camag = 4, the
to the initial droplet size, S s , decreases with increasing Camag , applied field prevents breakup and the droplet attains a sta-
as demonstrated in Fig. 11(b). A power-law fitting yields that ble, steady shape, as displayed in Fig. 12(b). Surprisingly, as

FIG. 10. Evolution of the droplet shape under the action of an external magnetic field parallel to the flow direction for (a) Camag = 2, (b) Camag = 4, (c) Camag
= 8, and (d) Camag = 12. Simulations performed with Re = 1, Ca = 0.5, λ = 1.2, and ζ = 2.
122110-8 Cunha et al. Phys. Fluids 30, 122110 (2018)

FIG. 11. Effect of an external magnetic field parallel to the flow direction on the breakup of ferrofluid droplets in terms of (a) time to break up, τ b , and
−1 . Notice that the point corresponding to Ca−1 = 1/7 does not appear in (b) because the size of
(b) relative size of the satellite droplet, S s , as a function of Camag mag
the satellite droplet in this case is so small that it cannot be measured due to computational limitations.

Camag increases, rupture occurs again. The time to break up when breakup occurs again and in a spectacular form, appears
decreases and the size of the satellite droplet increases with in Figs. 12(c) and 12(d). This complex behavior results from
Camag . The droplet evolution for Camag = 8 and Camag = 12, the competition between two distinct effects caused by the

FIG. 12. Evolution of the droplet shape under the action of an external magnetic field perpendicular to the flow direction for (a) Camag = 2, (b) Camag = 4,
(c) Camag = 8, and (d) Camag = 12. Simulations performed with Re = 1, Ca = 0.5, λ = 1.2, and ζ = 2.
122110-9 Cunha et al. Phys. Fluids 30, 122110 (2018)

FIG. 13. Evolution of the droplet shape


under the action of an external magnetic
field perpendicular to the flow direc-
tion with Camag = 20. Simulations per-
formed with Re = 0.1, Ca = 0.3, λ = 1.2,
and ζ = 2.

perpendicular external field. On the one hand, it deforms the V. CONCLUDING REMARKS
droplet along the velocity gradient direction, which increases
the shear stress acting on its surface and facilitates an eventual We presented a numerical study on the effects of an exter-
breakup. On the other hand, it also makes the droplet neck nal, uniform magnetic field on the dynamics and breakup of a
flatter, which increases the amount of liquid in this region and, two-dimensional supraparamagnetic ferrofluid droplet in sim-
as a consequence, hinders the rupture process. Notice that for ple shear flows. We also explored the role of the external field
a high enough field intensity as Camag = 12, the long satel- on the rheology of the two-phase fluid system viewed as a
lite droplet undergoes the same process as the initial droplet model for a dilute emulsion of ferrofluid droplets. The math-
until it pinches off at the center and gives rise to a secondary ematical model consists of the incompressible Navier-Stokes
rupture. equations, with the added magnetic and capillary forces, cou-
Finally, we analyze the effect of an external magnetic field pled with magnetostatic Maxwell’s equations. The dynamics
on a droplet that does not break up under the shear action of the droplet surface, including the topological transition dur-
only. To this end, we take Re = 0.1, Ca = 0.3, λ = 1.2, and ing breakup, was captured through the evolution of a level set
ζ = 2. Again, we consider external fields both parallel and function.
perpendicular to the flow direction, but now we take three dif- The results show a clear influence of the applied magnetic
ferent field intensities: Camag = 1, 10, and 20. For the cases field on the droplet deformation and inclination in shear flows,
where the external field is parallel to the flow direction, the which, in turn, contribute to substantial changes in the emul-
droplet does not break and rapidly achieves a stable configu- sion viscosity. When the external field is parallel to the flow
ration. Indeed, as discussed so far, for the range of parameters direction, the droplet strongly aligns with the flow streamlines,
explored here, the field-induced distortion in the flow direction yielding a reduction in the emulsion viscosity. In turn, when
acts as a restorative mechanism of the droplet shape and does the external field is perpendicular to the flow direction, the
not induce rupture. A different result is observed when the droplet is forced to align with the field and leads to a large dis-
external field is perpendicular to the flow direction. At Camag tortion of the flow streamlines. As a result, there is a dramatic
= 1 and Camag = 10, the field intensity is not strong enough to increase in the two-phase liquid viscosity.
induce rupture and the droplet still does not break. However, Our predictions also demonstrate that an external mag-
at Camag = 20, the field-induced distortion in the velocity gra- netic field can induce or hinder droplet rupture depending on
dient direction is so dramatic that the shear stress on the highly its orientation and intensity (relative to capillary forces). In
elongated droplet is strong enough to induce the breakup. particular, external fields parallel to the flow direction delay
Figure 13 displays the corresponding droplet evolution for this the breakup process and reduce the size of the satellite droplet.
case. Notably, this result indicates that strong enough exter- Remarkably, it was found that there is a critical magnetic capil-
nal magnetic fields perpendicular to the flow direction can be lary number above which the droplet becomes so aligned with
applied to induce droplet breakup in shear flows in which the the flow that the breakup process is totally avoided. When the
droplet would not rupture. external field is perpendicular to the flow direction, there is a
122110-10 Cunha et al. Phys. Fluids 30, 122110 (2018)

TABLE I: Results for the mesh convergence test. 6 Q. Liu, H. Li, and K. Y. Lam, “Optimization of deformable magnetic-
sensitive hydrogel-based targeting system in suspension fluid for site-
N 2θ/π [η] specific drug delivery,” Mol. Pharm. 15, 4632–4642 (2018).
7 J. P. Dailey, J. P. Phillips, C. Li, and J. S. Riffle, “Synthesis of silicone
100 0.8572 4.4197 magnetic fluid for use in eye surgery,” J. Magn. Magn. Mater. 194, 140–148
200 0.8622 4.6397 (1999).
8 P. A. Voltairas, D. I. Fotiadis, and C. V. Massalas, “Elastic stability of sil-
300 0.8637 4.6926
400 0.8642 4.7115 icone ferrofluid internal tamponate (SFIT) in retinal detachment surgery,”
J. Magn. Magn. Mater. 225, 248–255 (2001).
9 O. T. Mefford, R. C. Woodward, J. D. Goff, T. P. Vadala, T. G. S. Pierre,

J. P. Dailey, and J. S. Riffle, “Field-induced motion of ferrofluids through


delicate competition between the field-induced stretching in immiscible viscous media: Testbed for restorative treatment of retinal
detachment,” J. Magn. Magn. Mater. 311, 347–353 (2007).
the velocity gradient direction, which facilitates the breakup, 10 M. Backholm, M. Vuckovac, J. Schreier, M. Latikka, M. Hummel, M. B.
and the increase in the amount of liquid in the droplet neck, Linder, and R. H. A. Ras, “Oscillating ferrofluid droplet microrhe-
which hinders the rupture. This competition leads to a strik- ology of liquid-immersed sessile droplets,” Langmuir 33, 6300–6306
ing result: the intensity of the applied field can be adjusted (2017).
11 C. Khokhryakova (Bushueva), K. Kostarev, and A. Shmyrova, “Deforma-
to either delay, prevent, or induce droplet breakup. Moreover, tion of ferrofluid floating drop under the action of magnetic field as method
we also observed that very strong external magnetic fields that of interface tension measurement,” Exp. Therm. Fluid Sci. 101, 186–192
are perpendicular to the flow direction can be used to induce (2019).
12 J. D. Sherwood, “Breakup of fluid droplets in electric and magnetic fields,”
droplet breakup in situations where there would be no rupture
J. Fluid Mech. 188, 133–146 (1988).
under the action of the shear flow only. 13 O. E. Séro-Guillaume, D. Zouaoui, D. Bernardin, and J. P. Brancher,
In summary, the results presented here highlight the enor- “The shape of a magnetic liquid drop,” J. Fluid Mech. 241, 215–232
mous potential of external magnetic fields as a tool to con- (1992).
14 C. Flament, S. Lacis, J.-C. Bacri, A. Cebers, S. Neveu, and R. Perzynski,
trol the topology of ferrofluid droplets and the rheology of “Measurements of ferrofluid surface tension in confined geometry,” Phys.
ferrofluid emulsions. Rev. E 53, 4801–4806 (1996).
15 S. Banerjee, M. Fasnacht, S. Garoff, and M. Widom, “Elongation of con-

fined ferrofluid droplets under applied fields,” Phys. Rev. E 60, 4272–4279
ACKNOWLEDGMENTS (1999).
16 V. Bashtovoi, S. Pogirnitskaya, and A. Reks, “Dynamics of deformation of

We would like to acknowledge the financial support from magnetic fluid flat drops in a homogeneous longitudinal magnetic field,”
the Brazilian Research Council (CNPq) through Project No. J. Magn. Magn. Mater. 201, 300–302 (1999).
17 S. Afkhami, A. J. Tyler, Y. Renardy, M. Renardy, T. G. S. Pierre, R. C. Wood-
830015/2005-1. H.D.C. gratefully acknowledges support from ward, and J. S. Riffle, “Deformation of a hydrophobic ferrofluid droplet
the National Science Foundation through Grant No. DMS suspended in a viscous medium under uniform magnetic fields,” J. Fluid
1317684. Mech. 663, 358–384 (2010).
18 P. Rowghanian, C. D. Meinhart, and O. Campàs, “Dynamics of ferrofluid

drop deformations under spatially uniform magnetic fields,” J. Fluid Mech.


APPENDIX: MESH CONVERGENCE TEST 802, 245–262 (2016).
19 S. Afkhami, Y. Renardy, M. Renardy, J. S. Riffle, and T. S. Pierre, “Field-

We present here a mesh convergence test for the overall induced motion of ferrofluid droplets through immiscible viscous media,”
numerical scheme employed in this work. A representative set J. Fluid Mech. 610, 363–380 (2008).
20 D. Shi, Q. Bi, and R. Zhou, “Numerical simulation of a falling ferrofluid
of parameters is adopted for this test: Re = 0.01, Ca = 0.1, Camag droplet in a uniform magnetic field by the VOSET method,” Numer. Heat
= 10, λ = 2, ζ = 2, and a 12 × 12 square domain. The solution Transfer, Part A 66, 144–164 (2014).
21 D. Shi, Q. Bi, Y. He, and R. Zhou, “Experimental investigation on falling
was computed up to the time t = 1.3 using four different meshes
consisting of N × N cells. The accuracy of the numerical results ferrofluid droplets in vertical magnetic fields,” Exp. Therm. Fluid Sci. 54,
313–320 (2014).
was assessed by computing the droplet inclination angle, θ, and 22 J. Liu, Y. F. Yap, and N.-T. Nguyen, “Numerical study of the formation
the emulsion effective viscosity, [η]. The results are presented process of ferrofluid droplets,” Phys. Fluids 23, 072008 (2011).
23 Y. Wu, T. Fu, Y. Ma, and H. Z. Li, “Active control of ferrofluid droplet
in Table I. As can be seen, the four meshes yield very close
breakup dynamics in a microfluidic T-junction,” Microfluid. Nanofluid. 18,
results. Increasing the number of cells in each direction from
19–27 (2015).
N = 200 to N = 400 changes the effective viscosity in 0.2% 24 V. B. Varma, A. Ray, Z. M. Wang, Z. P. Wang, and R. V. Ramanujan,
and the droplet inclination in 1.6%. This analysis supports the “Droplet merging on a lab-on-a chip platform by uniform magnetic fields,”
fact that the results presented are independent of the mesh Sci. Rep. 6, 37671 (2016).
25 H. Li, Y. Wu, X. Wang, C. Zhu, T. Fu, and Y. Ma, “Magnetofluidic control of
refinement. the breakup of ferrofluid droplets in a microfluidic Y-junction,” RSC Adv.
6, 778–785 (2016).
1 I. Torres-Dı́az and C. Rinaldi, “Recent progress in ferrofluids research: 26 M. Aboutalebi, M. A. Bijarchi, M. B. Shafii, and S. K. Hannani, “Numerical

Novel applications of magnetically controllable and tunable fluids,” Soft investigation on splitting of ferrofluid microdroplets in T-junctions using
Matter 10, 8584–8602 (2014). an asymmetric magnetic field with proposed correlation,” J. Magn. Magn.
2 R.-J. Yang, H.-H. Hou, Y.-N. Wang, and L.-M. Fu, “Micro-magnetofluidics Mater. 447, 139–149 (2018).
in microfluidic systems: A review,” Sens. Actuators, B 224, 1–15 (2016). 27 C. Flament, G. Pacitto, J.-C. Bacri, I. Drikis, and A. Cebers, “Viscous fin-
3 P. Zhu and L. Wang, “Passive and active droplet generation with microflu- gering in a magnetic fluid. I. Radial Hele-Shaw flow,” Phys. Fluids 10,
idics: A review,” Lab Chip 17, 34–75 (2017). 2464–2472 (1998).
4 Q. A. Pankhurst, N. T. K. Thanh, S. K. Jones, and J. Dobson, “Progress in 28 G. Pacitto, C. Flament, and J.-C. Bacri, “Viscous fingering in a magnetic

applications of magnetic nanoparticles in biomedicine,” J. Phys. D: Appl. fluid. II. Linear Hele-Shaw flow,” Phys. Fluids 13, 3196–3203 (2001).
Phys. 42, 224001 (2009). 29 J. V. I. Timonen, M. Latikka, L. Leibler, R. H. A. Ras, and O. Ikkala,
5 I. K. Puri and R. Ganguly, “Particle transport in therapeutic magnetic fields,” “Switchable static and dynamic self-assembly of magnetic droplets on
Annu. Rev. Fluid Mech. 46, 407–440 (2014). superhydrophobic surfaces,” Science 341, 253–257 (2013).
122110-11 Cunha et al. Phys. Fluids 30, 122110 (2018)

30 A. Ahmed, B. A. Fleck, and P. R. Waghmare, “Maximum spreading of 39 V. E. Badalassi, H. D. Ceniceros, and S. Banerjee, “Computation of mul-
a ferrofluid droplet under the effect of magnetic field,” Phys. Fluids 30, tiphase systems with phase field models,” J. Comput. Phys. 190, 371–397
077102 (2018). (2003).
31 A. Ahmed, A. J. Qureshi, B. A. Fleck, and P. R. Waghmare, “Effects of mag- 40 G.-S. Jiang and D. Peng, “Weighted ENO schemes for Hamilton-Jacobi

netic field on the spreading dynamics of an impinging ferrofluid droplet,” equations,” SIAM J. Sci. Comput. 21, 2126–2143 (2000).
J. Colloid Interface Sci. 532, 309–320 (2018). 41 C. Min, “On reinitializing level set functions,” J. Comput. Phys. 229,
32 W. C. Jesus, A. M. Roma, and H. D. Ceniceros, “Deformation of a sheared 2764–2772 (2010).
magnetic droplet in a viscous fluid,” Commun. Comput. Phys. 24, 332–355 42 F. Gibou, R. Fedkiw, and S. Osher, “A review of level-set methods and some

(2018). recent applications,” J. Comput. Phys. 353, 82–109 (2018).


33 M. R. Hassan, J. Zhang, and C. Wang, “Deformation of a ferrofluid droplet in 43 M. Sussman, P. Smereka, and S. Osher, “A level set approach for com-

simple shear flows under uniform magnetic fields,” Phys. Fluids 30, 092002 puting solutions to incompressible two-phase flow,” J. Comput. Phys. 114,
(2018). 146–159 (1994).
34 P. Capobianchi, M. Lappa, and M. S. N. Oliveira, “Deformation of a 44 G. Ghigliotti, T. Biben, and C. Misbah, “Rheology of a dilute two-

ferrofluid droplet in a simple shear flow under the effect of a constant dimensional suspension of vesicles,” J. Fluid Mech. 653, 489–518 (2010).
magnetic field,” Comput. Fluids 173, 313–323 (2018). 45 H. D. Ceniceros, R. L. Nós, and A. M. Roma, “Three-dimensional, fully
35 R. E. Rosensweig, Ferrohydrodynamics (Cambridge University Press, adaptive simulations of phase field fluid models,” J. Comput. Phys. 229,
1985). 6135–6155 (2010).
36 G. K. Batchelor, “The stress system in a suspension of force-free particles,” 46 P. M. Vlahovska, “On the rheology of a dilute emulsion in a uniform electric

J. Fluid Mech. 41, 545–570 (1970). field,” J. Fluid Mech. 670, 481–503 (2011).
37 S. Osher and J. A. Sethian, “Fronts propagating with curvature dependent 47 S. Mandal and S. Chakraborty, “Effect of uniform electric field on the drop

speed: Algorithms based on Hamilton-Jacobi formulations,” J. Comput. deformation in simple shear flow and emulsion shear rheology,” Phys. Fluids
Phys. 79, 12–49 (1988). 29, 072109 (2017).
38 G. E. Karniadakis, M. Israeli, and S. A. Orszag, “High-order splitting meth- 48 S. Mandal and S. Chakraborty, “Uniform electric-field-induced non-

ods for the incompressible Navier-Stokes equations,” J. Comput. Phys. 97, Newtonian rheology of a dilute suspension of deformable Newtonian
414–443 (1991). drops,” Phys. Rev. Fluids 2, 093602 (2017).

Das könnte Ihnen auch gefallen