ArticlePDF AvailableLiterature Review

Theory and numerical simulation of droplet dynamics in complex flows - A review

Authors:
  • The Victor and Ruby Hansen Surface Professor of Molecular Modeling of Cancer, University of New Mexico

Abstract and Figures

We review theoretical and numerical studies and methods for droplet deformation, breakup and coalescence in flows relevant to the design of micro channels for droplet generation and manipulation.
Content may be subject to copyright.
Theory and numerical simulation of droplet dynamics in
complex flows—a review†
Vittorio Cristini‡ and Yung-Chieh Tan
Department of Biomedical Engineering, University of California, Irvine, REC 204, Irvine, CA
92697-2715, USA. E-mail: cristini@math.uci.edu; Fax: 949 824 1727; Tel: 949 824 9132
Received 3rd March 2004, Accepted 12th May 2004
First published as an Advance Article on the web 1st July 2004
We review theoretical and numerical studies and methods for droplet deformation, breakup and coalescence in flows
relevant to the design of micro channels for droplet generation and manipulation.
Introduction
Microscopic droplet formation and manipulation are key processes
(e.g., see ref. 1) in DNA analysis,
2–4
protein crystallization,
5,6
analysis of human physiological fluids,
7
and cell encapsulation.
8
Droplets are often used to either meter or mix small volumes of
fluids to improve significantly the mixing efficiency and allow
more complexity in chemical processing. For example, nanoliters
of reagents can be added (through droplet coalescence) and
removed (breakup) from a droplet leading to desired chemical
reaction rates and times. Droplet generation and manipulation using
immiscible flows,
9–18
has the advantage that generation is fast and
provides flexibility for the amount of reagent in each droplet, and
the cost of the micro channels is low, compared to electrowetting-
based devices. In addition, repeated fission and fusion of a single
stream of droplets at a bifurcating junction can be used to produce
parallel streams of smaller droplets containing a fraction of the
original chemical concentrations. Reactions of large volumes of
solution can be rapidly analyzed by splitting the solution into tens
or hundreds of smaller droplets.
10
This method can be applied to
studying protein conformations,
19
determining reaction kinetics,
20
and characterizing single molecule polymerase chain reactions.
21
For droplet systems controlled by external flows, the design of
the channel geometry is used to control the forces that create,
transport, split and fuse droplets. To our knowledge, droplet
generation in microfluidic circuits using flow was first demon-
strated (Fig. 1) by Thorsen et al.
14
Song et al.
11
were the first to
demonstrate droplet fission in microfluidic channels, and to show
that the channel walls, while significantly hindering droplet motion,
can improve the chemical mixing efficiencies by rotating the
droplets. Tan et al.
10
first demonstrated sorting and separation of
satellite droplets from droplets of larger sizes in microfluidics by
controlling the shear gradient at the bifurcating junction (Fig. 2),
Lab on a Chip special issue: The Science and Application of Droplets in
Microfluidic Devices.
Also at the Department of Mathematics.
Vittorio Cristini is currently an assistant professor of Bio-
medical Engineering and Mathematics at the University of
California, Irvine. He is also a member of the Chao Family
Comprehensive Cancer Center there. He earned his Ph.D.
degree in Chemical Engineering (2000) from Yale University,
where he also received a M.S. and a M.Phil. He earned his
degree of Dottore in Ingegneria Nucleare (Nuclear Engineer-
ing) from the University of Rome, Italy in 1994. Dr Cristini is an
expert in the field of multiphase flows and droplet and bubble
dynamics, where he has pub-
lished about twenty papers. For
his important work on droplet
breakup in laminar and turbu-
lent flows, Dr Cristini was the
first recipient of the “Andreas
Acrivos Dissertation Award in
Fluid Dynamics” from the
American Physical Society—
Division of Fluid Dynamics. Dr
Cristini has also published sev-
eral papers in the fields of
crystal and tumor growth.
Fig. 1 Control of droplet sizes by the pressure of the water-in-oil flow
(Fig. 4 from ref. 14 reproduced with permission from the American Physical
Society).
Fig. 2 Creation, sorting and collection of satellite droplets. (a) Satellite
droplets are created with larger primary droplets. (b) Primary droplets are
sorted towards the lower daughter channel while satellite droplets move into
the loop region of the sorter. (c) The channel collecting the satellite droplets
is free of primary droplets (data from ref. 10).
This journal is © The Royal Society of Chemistry 2004
MINIATURISATION FOR CHEMISTRY, BIOLOGY & BIOENGINEERING
DOI: 10.1039/b403226h
257Lab Chip, 2004, 4 , 257–264
and droplet fusion control using droplet focusing designs over a
variety of droplet sizes, numbers and generation speeds (Fig. 3).
Control of both the droplet generation speed
9
and size
9,16,18
has
been achieved in novel microfluidic designs (Figs. 4, 5 and 6). Tan
et al.
9
were successful in generating monodispersed pico- to
femtoliter sized droplets in a PDMS based microfluidic device by
controlling both the magnitude and type of flow in the micro
channels. Flow magnitude and type are controlled by varying the
difference between oil flow rates
9
and the difference between oil
and water flow rates,
9,16
in co-flowing streams at the droplet
generation site, and by varying the geometry and thus the flow
resistance of the channels at various junctions downstream.
10,18
Both designs lead to effectively varying the amount of straining of
the flow and thus provide finer control of the droplet sizes.
To control and optimize the creation, mixing, sorting, fusion and
splitting of droplets, droplet dynamics, and droplet breakup and
coalescence criteria and rates in 3-D flow and channel geometry
(e.g., in channel junctions and in non symmetric flow-focusing
streams) must be understood and controlled. The relative fluid
motions induced by the droplets, the detailed characteristics of the
dropletsolid contact angles, the interfacial tension at the droplet
liquid interface, the flow induced by the movement of the droplets,
and the shear force generated by the carrier flow around the droplets
affect the design. The current design is often empirical and not
founded on a rigorous study of the flow characteristics. Thus, there
is room for considerable improvements by using theory and direct
flow calculation.
Breakup and coalescence transitions are the result of the complex
interplay of viscous, inertial, capillary, Marangoni, electrostatic
and van der Waals forces over a wide variety of spatial and
timescales. Asymptotic (e.g. lubrication) theories describe the
limiting behaviors (e.g. pinching or film-drainage rates) of the
geometrical and field variables at the onset of these transitions
based on simplified assumptions on geometry (e.g. one-dimension-
ality), fluid and flow conditions. Numerical methods are capable of
describing transitions accurately and efficiently in simulations of a
variety of flow geometries (2-D and 3-D) and conditions where
interfaces deform significantly in principle without relying on
simplifying assumptions. A weakness of numerical simulations is
computational expense because the transition regions characterized
by small length- (and time-) scales need to be resolved by the
numerical discretization requiring a very large number of computa-
tional elements. Thus it is often practice to incorporate asymptotic
theories in numerical simulations, limiting direct numerical
solution to the larger scales.
Fig. 3 Fusion of large droplet slugs and free flowing droplets is controlled
by the external flow, which changes depending on the geometry and surface
properties of the walls (data from ref. 10). Flow rates are in ml min
21
.
Fig. 4 Two different scalings are shown for droplet generation (from data
in ref. 9). D/Di is the ratio of droplet diameter to the width of the orifice. Qo
with units of ml min
21
, is the flow rate of the oil phase. The flow rate of the
water phase is 0.5ml min
21
.
Fig. 5 Phase diagram for drop formation in flow focusing design (Fig. 3
from ref. 16 reproduced with permission from the American Institute of
Physics). For increasing oil flow rates (top-to-bottom) and increasing oil
flow rates relative to the water flow rate (left-to-right), water droplet sizes
decrease.
Fig. 6 Passive breakup at T-junctions (from Figs. 1 and 2 from ref. 18
reproduced with permission from the American Physical Society): by
varying the relative resistance (length) of the two side arms, the split droplet
volumes also vary. Bottom: the line dividing breaking and non-breaking
droplets is given by eqn. (2).
Lab Chip, 2004, 4 , 257–264258
In this article, we review recent theoretical and numerical
investigations of droplet dynamics in flow and we provide an
overview of numerical methods that can be applied to describe
breakup and coalescence phenomena in microfluidic design.
Computational accuracy and efficiency and the potential impact of
mesh adaptivity will also be discussed.
Theoretical and numerical investigations of
droplet breakup and coalescence
Drop breakup
Droplet generation in flow-focusing design, and breakup in flow,
are the result of the competition of viscous stresses associated with
the imposed flow field, and capillary stresses due to surface tension
between the two phases. In the flow-focusing design of water-in-oil
emulsions (e.g. see refs. 9 and 16), the shearing comes from the
relative magnitude of the co-flowing streams of oil and water. At a
channel T-junction,
18
droplets are sheared and extended as they
travel through the junction and can be split. It is necessary to know
under which conditions for the microphysical parameters and flow
type breakup occurs and at what rates, or droplets are stable (non
breaking), and the extent of droplet deformation.
Consider a drop of size (e.g. undeformed diameter) D, in a matrix
fluid of viscosity m undergoing flow with characteristic magnitude
G of the local velocity gradient, and with surface tension s. Viscous
stresses scale as mG, and extend a drop on a timescale
22
(1 + l)/G,
where l is the ratio of drop-to-matrix viscosities, taking equally
into account the contributions to friction coming from the fluid
viscosities. Surface tension s tends to relax a deformed drop back
to spherical. From a dimensional analysis, the capillary relaxation
velocity scales as s/m, thus the drop relaxation timescale is (1 + l)
mD/s, and capillary stresses scale as s/D. The relevant dimension-
less number is the capillary number Ca = mGD/s (assuming that
inertia is negligible), the ratio of viscous-to-capillary stresses, or
equivalently, the inverse ratio of the two corresponding timescales.
When Ca = O(1), viscous stresses deform the droplet significantly
and breakup may occur. The velocity gradient in a micro channel of
hydraulic diameter ~ D
i
and flow rate Q
o
can be estimated as G ~
Q
o
/D
i
3
and, under conditions of Ca = O(1), this gives a simple
relationship between generated droplet size and imposed flow
rate:
D ~sD
i
3
/mQ
o
(1)
Thus, the larger the flow rate, the smaller the droplet size. Using
9
s ~ 10
23
N m
21
, D
i
= 40 mm, m ~ 30 3 10
23
kg ms
21
and Q
o
= 12 ml min
21
, formula (1) gives a droplet size of D ~ 10mm (and
Re ~ 0.01) and is in agreement with the data in Fig. 4. These and
similar scalings have been employed to generate monodispersed
emulsions of controlled droplet sizes by Thorsen et al.,
14
Anna et
al.
16
and by Tan et al.
9
as illustrated in Figs. 4 and 5. Fig. 4 reveals
that droplet sizes fall in the theoretical scaling (1) when D < D
i
,
that is, droplets are small enough that the hydrodynamic forces
exerted by the channel walls are not important and breakup fully
relies on the straining of the imposed flow. When drops are large,
wall effects are dominant over the stresses directly imposed by the
flow, and the dependence of droplet sizes on flow rate is weaker.
These considerations are corroborated by the trend observed in Fig.
5. Similar scalings for breakup criteria have also been successfully
applied (Fig. 6) to the design of channel T-junction geometry and
inlet and outlet flow rates.
18
Interestingly, they found that the
capillary number above which droplets passively break at a T-
junction scales as
Ca ~e
0
(1/e
0
2/3
2 1)
2
(2)
where e
0
~ l
0
/w
0
, the ratio of length-to-width of the droplets in the
mother channel are upstream of the junction. Downstream of the
junction, split droplet volumes scale inversely with the lengths of
the side arms.
Experimental, theoretical and numerical studies of drop breakup
in imposed flows have been reviewed by Rallison
23
, Stone
24
and
Guido and Greco
25
(see also the review by Basaran
26
for jets).
Criteria for breakup were investigated experimentally (e.g. by
Bentley and Leal
27
) and analytically (e.g. by Navot
28
and
Blawzdziewicz et al.
29,30
). The distribution of drop fragments
resulting from breakup in shear flow was studied (e.g. see refs. 22
and 31). Numerical simulations have been developed (e.g. see refs.
22,3234). Emulsification is typically promoted using surfactants
that decrease surface tension on the droplet interfaces thus
favouring drop and jet breakup (see refs. 3542 and the review by
Maldarelli and Huang
43
). Methods for producing controlled micro
sized droplets were developed for shear flow,
22,31,4447
using co-
flowing streams (see ref. 48 and the recent microfluidics literature
listed above) and extrusion flow.
49
Monodispersed emulsions of
large numbers of droplets with controlled sizes were generated in
flow using the tip-streaming phenomenon due to redistribution of
surfactants to localized end caps on the drop interface.
50,51
Cristini
et al.
22
reported a study on the deformation and breakup of drops in
shear flow demonstrating that nearly bi-disperse emulsions of large
numbers of microscopic droplets of controlled sizes and generation
times can be achieved even without surfactants (Fig. 7). Inter-
estingly, the two sizes alternate (as also found by Tan et al.
9
). The
reason for this lies perhaps in the asymmetrical evolution of the
drop interface near the pinch-off region into cones with different
angles during the latest stages of pinch-off (Fig. 8, top), as
described by the theories of Blawzdziewicz et al.
52
and Lister and
Stone.
53
It was also found
22
that the breakup times have a non-
monotonic dependence on the capillary number, and have a (broad)
minimum corresponding to moderately supercritical shear rates.
This information can be used to optimize emulsification times.
It has been shown,
52,53
that during pinch-off of a thin liquid
thread under zero-Reynolds-number conditions (Fig. 8), the
thinning rate becomes asymptotically constant in time. Thus pinch-
off occurs in a finite time. As the thread thickness h
min
decreases,
viscous stresses mu/h
min
(u = 2dh
min
/dt is the thinning rate)
balance the capillary pressure : s/h
min
: mu/h
min
~s/h
min
. Thus the
neck pinching velocity u ~s/m and is (asymptotically) constant.
This is illustrated in Fig. 8 (bottom).
53
The axial curvature H’’ of
the thread (rescaled with h
min
(t)) at the minimum h
min
(t) is also
found to be asymptotically constant as h
min
? 0, thus revealing
self-similarity of the shape in the transition region. These scalings
for the capillary-driven pinch-off can be used when estimating drop
formation times in a micro channel. In a flow-focusing design
9,16
for water-in-oil emulsions, the time for accumulation of enough
water for one droplet is given by D
3
/Q
w
(the volume of the droplet
divided by the water flow rate). Capillary driven pinch-off time can
Fig. 7 Generation of a nearly bi-dispersed emulsion (From Figs. 11 and 12
from ref. 22 reproduced with permission from the American Institute of
Physics). The cumulative distribution of droplets of alternating sizes (inset)
formed from breakup of a mother drop in shear flow is shown. The droplet
sizes ¯a are rescaled with the maximum stable size in the flow.
Lab Chip, 2004, 4 , 257–264 259
be estimated as (1 + l) mD/s. Droplet generation time will be
determined by the larger of these two times. For droplets that are
smaller than the channel width (D < D
i
), the droplet size D is given
by the scaling (1), and thus the ratio of the two times is O(D/D
i
)
3
indicating that droplet generation time is dominated by the capillary
time for small droplets.
Drop coalescence
Coalescence rates and efficiencies between droplets depend on the
local dynamics of the fluid drainage in the near contact region
between the two approaching fluid interfaces. Asymptotic theories
and numerical simulations of film drainage between coalescing
fluidfluid interfaces,
5463
including surfactant effects,
6470
pro-
vide useful information such as the rates of dropdrop approach.
Two drops approaching each other trap a thin film of the
continuous phase between their interfaces. At small enough gaps
the hydrodynamic forces overcome capillarity and the drop
interfaces deform and often acquire a dimpled shape that traps more
fluid thus opposing coalescence.
55,62,69
In flow-driven drop
interactions, coalescence occurs for capillary numbers smaller than
a critical value such that the drop interaction time is larger than the
drainage time for the fluid trapped in the gap.
57
For sub-critical
capillary numbers, at sub-micron separations, van der Waals forces
become dominant leading to rapid coalescence (film rupture).
Surfactants (through Marangoni stresses, surface viscosity, Gibbs
elasticity, surface and/or bulk diffusivity and intermolecular forces)
can have a significant effect and stabilize emulsions by increasing
deformation and causing surface tension gradients (Marangoni
stresses, see the book by Edwards et al.
71
) that resist radial flow in
the gap and interfaceinterface approach thus preventing coales-
cence.
6870,72,73
Chesters and Bazhlekov
69
studied numerically the
axi-symmetric film drainage and rupture between two drops
approaching each other under a constant force in the presence of an
insoluble surfactant. For drops in the millimeter size range the
influence of surfactant diffusion is typically negligible. Drainage is
virtually unaffected by the presence of surfactants down to a film
thickness at which high Marangoni stresses and a transition to
immobile interfaces set in, leading to drainage orders of a slower
magnitude. These phenomena are illustrated in Fig. 9
69
for the
minimum film thickness h
min
. Analytical approximations for the
drainage time were derived.
69
For smaller drops the influence of
diffusion alleviates gradients in surfactant concentration reducing
the effect of surfactants. Yeo et al.
70
(see also ref. 73) using theory
and numerical simulations focused on constant approach velocity
collisions and highlighted the differences in dynamics that arise.
They showed that adding even a slight amount of insoluble
surfactant results in the immobilization of the interface. Three
regimes of drainage and possible rupture exist depending on the
relative magnitudes of the drop approach velocity and the van der
Waals interaction force: nose rupture, rim rupture, and film
immobilization and flattening. The possibility of forming secon-
dary droplets by encapsulating the continuous phase film into the
coalesced drop at rupture was also quantified.
Recent experimental
74,75
and numerical investigations
76
of flow
driven dropdrop coalescence with surfactants revealed that there is
a non-monotonic dependence of the critical capillary number for
coalescence on the surfactant coverage (Fig. 10). The critical
capillary number has a minimum for intermediate coverage due to
Marangoni stresses, and increases at high coverage (close to the
maximum packing of molecules on the interface) perhaps as a
consequence of interface immobilization and reduced deformation.
Fig. 8 Top: initial (dashed) and final (solid) shapes during pinch-off (Fig.
2 from ref. 53). Liquid threads separate large daughter drops and have nearly
conical shapes near the locations of minimum thickness. Bottom (Fig. 6
from ref. 53): constant thread-thinning velocity and shape self-similarity as
h
min
? 0. Reproduced with permission from the American Institute of
Physics.
Fig. 9 Schematic illustrating the effects of surfactants and transition to
immobilized interfaces and slow drainage (Fig. 13,
69
reproduced with
permission from Elsevier).
Fig. 10 Coalescence occurs for capillary numbers below the critical curve.
Curves are shown (data from ref. 76) as a function of surfactant coverage c
of the interface for insoluble surfactants and for a surfactant that is soluble
in the bulk fluids, for different values of a dimensionless van der Waals
force, indicating non-monotonic dependence on the surfactant coverage.
Lab Chip, 2004, 4 , 257–264260
These findings demonstrate that while much progress has been
made in the theoretical understanding of film drainage, there are
still open questions.
Numerical methods
The two main approaches to simulating multiphase and multi-
component flows are interface tracking and interface capturing.
Interface tracking methods
In interface tracking methods (or sharp-interface methods) the
computational mesh elements lay in part or fully on the fluidfluid
interfaces. Such methods, including boundary-integral meth-
ods,
32,33,34,70,7780
finite-element methods
8183
and immersed-
boundary methods,
84,85
are very accurate for simulating the onset of
breakup and coalescence transitions but have difficulties in
simulating through and past the transitions.
In boundary-integral methods, the flow equations are mapped
from the immiscible fluid domains to the sharp interfaces
separating them thus reducing the dimensionality of the problem
(the computational mesh discretizes only the interface). In finite-
element methods, the fluid domains are discretized by a volume
mesh and thus the dimensionality is not reduced. Both these
approaches lead to accurate and efficient solution of the flow
equations because the interface is part of the computational mesh
and the equations and interface boundary conditions are posed
exactly. In immersed-boundary methods, the interfacial forces are
calculated on a surface mesh distinct from the computational
volume mesh where the flow equations are solved and thus in
addition, interpolation onto the volume mesh is needed. A three-
dimensional boundary integral simulation of the onset of drop
breakup in simple shear flow is shown in Fig. 11 (top).
33
In the three
frames, the evolution of a drop towards breakup and the formation
of a thinning liquid thread separating two large daughter drops are
shown. The labels report the dimensionless time from breakup. The
calculated drop shapes (computational mesh) are compared to an
experiment (solid contour) by Dr Guido and coworkers (University
of Naples, Italy, personal communication), demonstrating the high
accuracy of the numerical method. These simulations used an
adaptive triangulated mesh.
34
In Fig. 12 a boundary-integral
simulation of the flow of deformable droplets through a channel is
shown.
86
In Fig. 13 an application of the finite-element method to
high-Reynolds-number satellite production from a jet is illus-
trated.
83
Cross sections of the computed evolution in time (top)
compare well with the experiments (bottom) (corresponding to
different initial conditions). The computational accuracy allowed
the authors to recover features of the evolutions such as capillary
waves (a)(c), overturning of the top and bottom of the satellite in
(f), spade-shaped profiles, (g) and (h), and spawning of a
subsatellite. In Fig. 14, immersed-boundary simulations of bubbles
interacting and coalescing are shown.
84,87
Near breakup and coalescence transitions, sharp interface models
break down because of the formation of singularities in flow
variables
88
and complex ad hoc cut-and-connect algorithms have
been employed
34,84,89
to change the topology of the meshes and
continue simulating through a transition. (A method to automat-
ically reconnect sharp interfaces has been recently developed by
Shin and Juric
85
). In the simulation in Fig. 14 (bottom), Nobari et
al.
87
reconnected the interfaces as their separation fell under a
prescribed value but noticed that the flow can depend sensitively on
the time at which the interface reconnections are performed.
Reconnection conditions based on asymptotic theories describing
liquid thread pinch-off and film drainage
52,53,9092
are also used to
extrapolate to the instant of breakup or coalescence, as illustrated in
Fig. 11 (bottom), where Cristini et al.
34
simulated the evolution of
the drop in shear after the pinch-off transition by employing a cut-
and-connect mesh algorithm after the asymptotically linear thin-
ning of the liquid thread
52,53
had been established. This allows
simulations to be continued past the transition while preserving
Fig. 11 Top: adaptive 3-D boundary-integral simulation of the onset of
drop breakup in shear flow (Fig. 1 from ref. 33 reproduced with permission
from the American Institute of Physics). Bottom: the simulation is
continued past the transition by reconnecting the computational mesh (Fig.
6 from ref. 34 reproduced with permission from Elsevier).
Fig. 12 Boundary-integral simulation of 3-D droplet motion through a
channel (Fig. 8 from ref. 86 reproduced with permission from Cambridge
University Press).
Fig. 13 Finite-element simulation (top) of satellite formation and
dynamics from a jet compared to an experiment (bottom) (Figs. 2 and 3 from
ref. 83 reproduced with permission from the American Institute of
Physics).
Lab Chip, 2004, 4 , 257–264 261
accurate information on breakup time (Fig. 11, top) and fragment
sizes.
Interface-capturing methods
Simulations through breakup and coalescence transitions using
interface capturing methods, i.e. lattice-Boltzmann and lattice-
gas,
9398
constrained-interpolation-profile,
99
level-set,
100
volume-
of-fluid,
101
coupled level-set and volume-of-fluid
102
and partial-
miscibility-model and phase-field methods,
99,103114
do not require
mesh cut-and-connect operations because the mesh elements do not
lay on the interface, but rather the interface evolves through the
mesh. The fluid discontinuities (e.g. density, viscosity) are
smoothed and the surface tension force is distributed over a thin
layer near the interface to become a volume force (surface tension
being the limit as the layer approaches zero thickness). Interface-
capturing methods are then ideal for simulating breakup and
coalescence in immiscible two-fluid systems (and the effect of
surfactants) and for three or more liquid components, and can be
especially powerful for micro channel design. Lattice-Boltzmann
methods are based on a particle distribution function and on
averaging to capture the macroscopic behavior. The constrained-
interpolation-profile, level-set and volume-of-fluid methods de-
scribe the macro scale directly and use auxiliary functions advected
by the flow (e.g. level-set, volume fraction, and color functions) to
mark the different fluid domains. In Fig. 15, we reproduce a
simulation of drop breakup in shear flow using a volume-of-fluid
method;
115
(see also refs. 116118). The drop is strongly stretched
in the supercritical flow leading to rupture into numerous fragments
of alternating sizes (see for comparison refs. 9 and 22). The diffuse-
interface (phase-field) approach is based on free-energy functionals
and uses chemical diffusion in narrow transition layers between the
different fluid components as a physical mechanism to smooth flow
discontinuities and to yield smooth evolution through breakup and
coalescence. Through the energy formulation, van der Waals
interactions, electrostatic forces and components with varying
miscibilities can be described. An example of simulation of jet
pinch-off using the partial-miscibility model is shown in Fig. 16 (J.
Lowengrub, U. C Irvine E. Longmire and U. Minnesota, personal
communications, see also refs. 111 and 112). The liquidliquid
interface and the axial velocity profiles during jet pinch-off from
the diffuse-interface simulation and from an experiment show good
agreement.
Mesh adaptivity
Accurate numerical simulations of multiphase flow and topology
transitions require the computational mesh to resolve both the
macro (e.g. droplet size, channel geometry) and micro scale where
pinch-off or coalescence occur, and to capture local interface
curvature, interfaceinterface separation, van der Waals forces,
surfactant distributions and Marangoni stresses. Adaptive mesh
algorithms greatly increase accuracy and computational efficiency
in boundary-integral,
34
finite element,
81,82
immersed boundary,
84
and interface capturing
119127
methods. In Fig. 17 level-set
simulations of drop coalescence in 2-D (top) and drop breakup in
3-D (bottom) under Stokes flow conditions are reported using novel
unstructured adaptive meshes.
34,127
The adaptive mesh algorithm
automatically imposes a mesh element size proportional to the
distance from the interface. As the interfaces deform, approach or
pinch-off, the mesh dynamically maintains accurate resolution of
the flow near the interface. This allows a simulation to recover the
Fig. 14 Top: 3-D front-tracking simulation of rising bubbles (Fig. 13 from
ref. 84 reproduced with permission from Elsevier). Bottom: axi-symmetric
front-tracking simulation of droplet coalescence (Fig. 13 from ref. 87
reproduced with permission from the American Institute of Physics).
Fig. 15 3-D volume-of-fluid simulation (Fig. 16 from ref. 115 reproduced
with permission from the American Institute of Physics) of drop breakup in
shear flow (view along the velocity gradient).
Fig. 16 Comparison between a diffuse-interface simulation (right) (J.
Lowengrub, U. C Irvine) and an experiment (left) (E. Longmire, U.
Minnesota, personal communications) of a liquidliquid jet pinch-off.
Lab Chip, 2004, 4 , 257–264262
lubrication forces that resist drop approach and delay coalescence
(top). The log-linear plot in the inset reports gap thickness history
for increasing mesh resolution (left to right) and is compared with
lubrication theory (straight line), that in 2-D predicts exponential
gap thinning and thus an infinite coalescence time when only
hydrodynamics are considered.
127
The simulation results converge
to the lubrication theory. Below a gap thickness that decreases with
increased resolution, numerical coalescence occurs. A 3-D simula-
tion (bottom) accurately describes drop breakup during retraction
of a previously elongated drop.
Conclusions
Theoretical considerations relevant to controlled droplet genera-
tion, breakup and coalescence in micro channels were discussed.
Theoretical and numerical investigations of droplet breakup and
coalescence in imposed flows were reviewed. Numerical methods
were reviewed that can be applied for directly simulating droplet
generation, dynamics, breakup and coalescence in micro channels.
The magnitude and type of flow are both important in determining
generation, breakup and coalescence times and droplet sizes. The
local velocity gradient imposed on a droplet is a function of position
in the micro channel, and depends strongly on the channel
geometry. In microfluidics design, channel geometry and flow
conditions can be optimised using (adaptive) simulations. Im-
portant issues are still open in the literature. In particular, one topic
where theoretical and numerical work has been limited and further
investigation is highly desirable in the near future is dropwall
interactions. For droplets of size comparable to the channel width,
the hydrodynamic forces exerted by the wall on the drop may
exceed those exerted by the imposed channel flow. Under these
conditions, droplet deformation and droplet breakup and coales-
cence criteria and rates may be strongly affected.
V. C. is grateful to John Lowengrub for the many enjoyable and
useful discussions on numerical methods, and acknowledges
support from the National Science Foundation.
References
1 mTAS, 7th International Conference on Micro Total Analysis Systems,
Proceedings of mTAS, ed. M. A. Northrup, K. F. Jensen and D. J.
Harrison, Squaw Valley, California, USA, 2003, Transducers Re-
search Foundation, San Diego, CA.
2 M. A. Burns, B. N. Johnson, S. N. Brahmasandra, K. Handique, J. R.
Webster, M. Krishnan, T. S. Sammarco, P. M. Man, D. Jones, D.
Heldsinger, C. H. Mastrangelo and D. T. Burke, Science, 1998, 282,
484.
3 M. G. Pollack, P. Y. Paik, A. D. Shenderov, V. K. Pamula, F. S.
Dietrich and R. B. Fair, 7th International Conference on mTAS, 2003,
619622.
4 S. Kaneda and T. Fujii, 7th International Conference on mTAS, 2003,
12791282.
5 M. Hirano, T. Torii, T. Higuchi, M. Kobayashi and H. Yamazaki, 7th
International Conference on mTAS, 2003, 473.
6 B. Zheng, L. S. Roach and R. F. Ismagilov, J. Am. Chem. Soc., 2003,
125, 1117011171.
7 V. Srinivasan, V. K. Pamula, M. G. Pollack and R. B. Fair, 7th
International Conference on mTAS, 2003, 12871290.
8 B. Schaack, B. Fouque, S. Porte, S. Combe, A. Hennico, O.
Filholcochet, J. Reboud, M. Balakirev and F. Chatelain, 7th Inter-
national Conference on mTAS, 2003, 669672.
9 Y. C. Tan, V. Cristini and A. P. Lee, Sens. Actuators, 2004, in
review.
10 Y. C. Tan, J. Fisher, A. I. Lee, E. Lin, V. Cristini and A. P. Lee, Lab
Chip, 2004, DOI: 10.1039/b403280m.
11 H. Song, J. D. Tice and R. F. Ismagilov, Angew. Chem., Int. Ed., 2003,
42(7), 767772.
12 J. S. Go, E. H. Jeong, K. C. Kim, S. Y. Yoon and S. Shoji, 7th
International Conference on mTAS, 2003, 12751278.
13 S. Sugiura, M. Nakajima, S. Iwamoto and M. Seki, Langmuir, 2001,
17, 5562.
14 T. Thorsen, W. R. Robert, F. H. Arnold and S. R. Quake, Phys. Rev.
Lett., 2001, 86, 4163.
15 T. Nisisako, T. Torii and T. Higuchi, Lab Chip, 2002, 2, 24.
16 S. L. Anna, N. Bontoux and H. A. Stone, Appl. Phys. Lett., 2003, 82,
364.
17 J. D. Tice, D. A. Lyon and R. F. Ismagilov, Anal. Chim. Acta, 2004,
507, 73.
18 D. R. Link, S. L. Anna, D. A. Weitz and H. A. Stone, Phys. Rev. Lett.,
2004, 92, 11781190, Art. 054503.
19 M. Kakuta, D. A. Jayawickrama, A. M. Wolters, A. Manz and J. V.
Sweedler, Anal. Chem., 2003, 75, 956960.
20 C. de Bellefon, N. Pestre, T. Lamouille, P. Grenouillet and V. Hessel,
Adv. Synth. Catal., 2003, 345(1 + 2), 190193.
21 M. Nakano, J. Komatsu, S. Matsuura, K. Takashima, S. Katsura and A.
Mizuno, J. Biotechnol., 2003, 102, 117124.
22 V. Cristini, S. Guido, A. Alfani, J. Blawzdziewicz and M. Loewenberg,
J. Rheol., 2003, 47, 1283.
23 J. M. Rallison, Annu. Rev. Fluid Mech., 1984, 16, 45.
24 H. A. Stone, Annu. Rev. Fluid Mech., 1994, 26, 65.
25 S. Guido and F. Greco, in Rheology Reviews, ed. D. M. Binding and K.
Walters, British Soc. Rheol. 2004, in press.
26 O. Basaran, AIChE J., 2002, 48, 1842.
27 B. J. Bentley and L. G. Leal, J. Fluid Mech., 1986, 167, 241.
28 Y. Navot, Phys. Fluids, 1999, 11, 990.
29 J. Blawzdziewicz, V. Cristini and M. Loewenberg, Phys. Fluids, 2002,
14, 2709.
30 J. Blawzdziewicz, V. Cristini and M. Loewenberg, Phys. Fluids, 2003,
15, L37.
31 V. Schmitt, F. Leal-Calderon and J. Bibette, Colloid Chem. II (Book
Series: Topics in current chemistry), EDP Sciences, Les Ulis, France,
2003, vol. 227, p. 195.
32 A. Z. Zinchenko, M. A. Rother and R. H. Davis, J. Fluid Mech., 1999,
391, 249.
33 V. Cristini, J. Blawzdziewicz and M. Loewenberg, Phys. Fluids, 1998,
10, 1781.
34 V. Cristini, J. Blawzdziewicz and M. Loewenberg, J. Comput. Phys.,
2001, 168, 445.
35 H. A. Stone and L. G. Leal, J. Fluid Mech., 1990, 220, 161.
36 W. J. Milliken, H. A. Stone and L. G. Leal, Phys. Fluids A, 1993, 5,
69.
37 W. J. Milliken and L. G. Leal, J. Colloid Interface Sci., 1994, 166,
275.
38 Y. Pawar and K. J. Stebe, Phys. Fluids, 1996, 8, 1738.
39 C. D. Eggleton, Y. Pawar and K. Stebe, J. Fluid Mech., 1999, 79,
385.
40 X. Li and C. Pozrikidis, J. Fluid Mech., 1997, 341, 165.
41 M. Siegel, SIAM J. Appl. Math., 1999, 69, 1998.
Fig. 17 Adaptive unstructured meshes of triangles (top) and tetrahedra
(bottom) maintain computational accuracy during simulations (data from
ref. 127). Some tetrahedra may appear skewed (bottom) as a result of
projecting the 3-D mesh onto the plane of the figure.
Lab Chip, 2004, 4 , 257–264 263
42 Y.-J. Jan and G. Tryggvason, in Proceedings of the Symposium on
Dynamics of Bubbles and Vortices Near a Free Surfaces,ed. Sahin and
Tryggvason, ASME, NY, 1991, vol. 119, 4659.
43 C. Maldarelli and W. Huang, in Flow particle suspensions. ed. U.
Schaflinger, CISM Courses and Lectures, Springer-Verlag, New York,
1996, vol. 370, p. 125.
44 T. G. Mason and J. Bibette, Langmuir, 1997, 13, 4600.
45 A. J. Abrahamse, R. van Lierop, R. G. M. van der Sman, A. van der
Padt and R. M. Boom, J. Membr. Sci., 2002, 204, 125.
46 W. L. Olbricht and D. M. Kung, Phys. Fluids A, 1992, 4, 1347.
47 W. G. P. Mietus, O. K. Matar, C. J. Lawrence and B. J. Briscoe, Chem.
Eng. Sci., 2001, 57, 1217.
48 P. B. Umbanhowar, V. Prasad and D. A. Weitz, Langmuir, 2000, 16,
347.
49 I. Kobayashi, M. Yasuno, S. Iwamoto, A. Shono, K. Satoh and M.
Nakajima, Colloids Surf., 2002, 207, 185.
50 R. A. de Bruijn, Chem. Eng. Sci., 1993, 277, 48.
51 C. D. Eggleton, T. M. Tsai and K. J. Stebe, Phys. Rev. Lett., 2001,
87.
52 J. Blawzdziewicz, V. Cristini and M. Loewenberg, Bull. Am. Phys.
Soc., 1997, 42, 2125.
53 J. R. Lister and H. A. Stone, Phys. Fluids, 1998, 10, 2758.
54 G. P. Neitzel and P. DellAversana, Annu. Rev. Fluid Mech., 2002, 34,
267.
55 S. G. Yiantsios and R. H. Davis, J. Fluid Mech., 1990, 217, 547.
56 S. G. Yiantsios and R. H. Davis, J. Colloid Interface Sci., 1991, 144,
412.
57 A. K. Chesters, Chem. Eng. Res. Des., 1991, 69, 259.
58 P. D. Howell, J. Eng. Math., 1999, 35, 271.
59 R. H. Davis, J. A. Schonberg and J. M. Rallison, Phys. Fluids A, 1989,
1, 77.
60 D. Li, J. Colloid Interface Sci., 1994, 163, 108.
61 E. Klaseboer, J. Ph. Chevaillier, C. Gourdon and O. Masbernat, J.
Colloid Interface Sci., 2000, 229, 274.
62 M. A. Rother and R. H. Davis, Phys. Fluids, 2001, 13, 1178.
63 J. Eggers, J. R. Lister and H. A. Stone, J. Fluid Mech., 1999, 401,
293.
64 G. Singh, G. J. Hirasaki and C. A. Miller, J. Colloid Interface Sci.,
1996, 184, 92.
65 D. Li, J. Colloid Interface Sci., 1996, 181, 34.
66 K. D. Danov, D. S. Valkovska and I. B. Ivanov, J. Colloid Interface
Sci., 1999, 211, 291.
67 D. S. Valkovska, K. D. Danov and I. B. Ivanov, Colloids Surf., A, 2000,
175, 179.
68 V. Cristini, J. Blawzdziewicz and M. Loewenberg, J. Fluid Mech.,
1998, 366, 259.
69 A. K. Chesters and I. B. Bazhlekov, J. Colloid Interface Sci., 2000,
230, 229.
70 L. Y. Yeo, O. K. Matar, E. S. P. de Ortiz and G. E. Hewitt, J. Colloid
Interface Sci., 2003, 257(1), 93107.
71 D. A. Edwards, H. Brenner and D. T. Wasan, Interfacial Transport
Processes and Rheology, ButterworthHeinemann, London, 1991.
72 J. Blawzdziewicz, V. Cristini and M. Loewenberg, J. Colloid Interface
Sci., 1999, 211, 355.
73 L. Y. Yeo, O. K. Matar, E. S. P. de Ortiz and G. F. Hewitt, J. Colloid
Interface Sci., 2001, 241, 233.
74 Y. T. Hu, D. J. Pine and L. G. Leal, Phys. Fluids, 2000, 12, 484.
75 J. W. Ha, Y. Yoon and L. G. Leal, Phys. Fluids, 2003, 15, 849.
76 H. Zhou, V. Cristini, C. W. Macosko and J. Lowengrub, Phys. Fluids,
in review.
77 C. Pozrikidis, Boundary Integral and Singularity Methods for
Linerarized Viscous Flow, Cambridge University Press, Cambridge,
1992.
78 A. Prosperetti and H. N. Oguz, Philos. Trans. R. Soc. London, 1997,
355, 491.
79 T. Y. Hou, J. S. Lowengrub and M. J. Shelley, J. Comput. Phys., 2001,
169, 302.
80 C. Pozrikidis, Eng. Anal. Bound. Elem., 2002, 26, 495.
81 E. D. Wilkes, S. Phillips and O. Basaran, Phys. Fluids, 1999, 11,
3577.
82 R. Hooper, V. Cristini, S. Shakya, J. Lowengrub, C. W. Macosko and
J. J. Derby, in Computational Methods in Multiphase Flow, ed. H.
Power and C. A. Brebbia, Wessex Institute of Technology Press, 2001,
vol. 29.
83 P. K. Notz, A. U. Chen and O. A. Basaran, Phys. Fluids, 2001, 13,
549.
84 G. Tryggvason, B. Bunner, A. Esmaeeli, D. Juric, N. Al-Rawahi, W.
Tauber, J. Han, S. Nas and Y.-J. Jan, J. Comput. Phys., 2001, 169,
708.
85 S. Shin and D. Juric, J. Comput. Phys., 2002, 180, 427.
86 C. Coulliette and C. Pozrikidis, J. Fluid Mech., 1998, 358, 1.
87 M. R. Nobari, Y.-J. Jan and G. Tryggvason, Phys. Fluids, 1996, 8,
29.
88 T. Y. Hou, Z. Li, S. Osher and H. Zhao, J. Comput. Phys., 1997, 134,
236.
89 N. Mansour and T. Lundgren, Phys. Fluids A, 1990, 2, 1141.
90 J. B. Keller and M. J. Miksis, SIAM J. Appl. Math., 1983, 43, 268.
91 J. Eggers and T. F. Dupont, J. Fluid Mech., 1994, 262, 205.
92 J. Eggers, Phys. Fluids, 1995, 7, 941.
93 D. H. Rothman and S. Zaleski, Lattice Gas Cellular Automata,
Cambridge University Press, Cambridge, 1997.
94 S. Chen and G. D. Doolen, Annu. Rev. Fluid Mech., 1998, 30, 329.
95 R. R. Nourgaliev, T. N. Dinh, T. G. Theofanous and D. Joseph, Int. J.
Multiphase Flow, 2003, 29, 117.
96 T. Watanabe and K. Ebihara, Comput. Fluids, 2003, 32, 823.
97 K. Sankaranarayanan, I. G. Kevrekidis, S. Sundaresan, J. Lu and G.
Tryggvason, Int. J. Multiphase Flow, 2003, 29, 109.
98 A. Lamura, G. Gonnella and J. M. Yeomans, Europhys. Lett., 1999, 45,
314.
99 T. Yabe, F. Xiao and T. Utsumi, J. Comput. Phys., 2001, 169, 556.
100 S. Osher and R. Fedkiw, J. Comput. Phys., 2001, 169, 463.
101 R. Scardovelli and S. Zaleski, Annu. Rev. Fluid Mech., 1999, 567.
102 M. Sussman and E. G. Puckett, J. Comput. Phys., 2000, 162, 301.
103 J. S. Lowengrub and L. Truskinovsky, Proc. R. Soc. Lond. Ser. A,
1998, 454, 2617.
104 J. Lowengrub, J. Goodman, H. Lee, E. Longmire, M. Shelley and L.
Truskinovsky, in Free boundary problems: theory and applications,
ed. I. Athanasopoulos, M. Makrakis and J. F. Rodrigues, CRC Press,
London, 1999 vol. 221.
105 D. M. Anderson, G. B. McFadden and A. A. Wheeler, Annu. Rev. Fluid
Mech., 1998, 30, 139.
106 D. Jacqmin, J. Comput. Phys., 1999, 155, 96.
107 D. Jamet, O. Lebaigue, N. Coutris and J. M. Delhaye, J. Comput. Phys.,
2001, 169, 624.
108 H. Lee, J. S. Lowengrub and J. Goodman, Phys. Fluids, 2002, 14,
492.
109 H. Lee, J. S. Lowengrub and J. Goodman, Phys. Fluids, 2002, 14,
514.
110 V. E. Badalassi, H. D. Ceniceros and S. Banerjee, J. Comput. Phys.,
2003, 190, 371.
111 J. S. Kim, K. Kang and J. S. Lowengrub, J. Comput. Phys., 2004, 193,
511.
112 J. S. Kim, K. Kang and J. S. Lowengrub, Commun. Math. Sci., 2004,
2, 5377.
113 L. Q. Chen, Annu. Rev. Mater. Res., 2002, 113, 32.
114 G. Patzold and K. Dawson, Phys. Rev. E: Stat. Phys., Plasmas, Fluids,
Relat. Interdiscip. Top., 1995, 52, 6908.
115 J. Li, Y. Y. Renardy and M. Renardy, Phys. Fluids, 2000, 12, 269.
116 Y. Y. Renardy, V. Cristini and J. Li, Int. J. Multiphase Flow, 2002, 28,
1125.
117 Y. Y. Renardy and V. Cristini, Phys. Fluids, 2001, 13, 7.
118 Y. Y. Renardy and V. Cristini, Phys. Fluids, 2001, 13, 2161.
119 G. Agresar, J. J. Linderman, G. Tryggvason and K. G. Powell, J.
Comput. Phys., 1998, 143, 346.
120 M. Sussman, A. Almgren, J. Bell, P. Colella, L. Howell and M.
Welcome, J. Comput. Phys., 1999, 148, 81.
121 N. Provatas, N. Goldenfeld and J. Dantzig, J. Comput. Phys., 1999,
148, 265.
122 O. Ubbink and R. I. Issa, J. Comput. Phys., 1999, 153, 26.
123 H. D. Ceniceros and T. Y. Hou, J. Comput. Phys., 172, 609, 2001.
124 J. H. Jeong, N. Goldenfeld and J. A. Dantzig, Phys. Rev. E: Stat. Phys.,
Plasmas, Fluids, Relat. Interdiscip. Top., 2001, 64, Art. 041602 Part
1.
125 J. H. Jeong, J. A. Dantzig and N. Goldenfeld, Metall. Mater. Trans. A,
2003, 34, 459.
126 I. Ginzberg and G. Wittum, J. Comput. Phys., 2001, 166, 302.
127 X. Zheng, A. Anderson, J. Lowengrub and V. Cristini, J. Comput.
Phys., in review.
Lab Chip, 2004, 4 , 257–264264
... The fluids can also be induced concurrently, for example, Takeuchi et al [105] fabricated the symmetric flow-focusing device by the cast molding method. In quasi-2D flow-focusing devices, the fluids flow in the same direction [106][107][108][109]. The minimum droplet size in a flow-focusing device can reach hundreds of nanometers, similar to those in co-flow devices [106,110]. ...
Article
Full-text available
In the last three decades, carbon dioxide (CO2) emissions have shown a significant increase from various sources. To address this pressing issue, the importance of reducing CO2 emissions has grown, leading to increased attention toward carbon capture, utilization, and storage (CCUS) strategies. Among these strategies, monodisperse microcapsules, produced using droplet microfluidics, have emerged as promising tools for carbon capture, offering a potential solution to mitigate CO2 emissions. However, the limited yield of microcapsules due to the inherent low flow rate in droplet microfluidics remains a challenge. In this comprehensive review, the high-throughput production of carbon capture microcapsules using droplet microfluidics is focused on. Specifically, the detailed insights into microfluidic chip fabrication technologies, the microfluidic generation of emulsion droplets, along with the associated hydrodynamic considerations, and the generation of carbon capture microcapsules through droplet microfluidics are provided. This review highlights the substantial potential of droplet microfluidics as a promising technique for large-scale carbon capture microcapsule production, which could play a significant role in achieving carbon neutralization and emission reduction goals.
... By introducing the microfluidics for the generation of emulsions, the interest in this field has increased continuously. It has led to the development of fluid flow devices at the scale of micro (Cristini and Tan, 2004;Whitesides, 2006;Zhu and Wang, 2017;Doufène et al., 2019). The advantages of using microfluidics include the low cost, high throughput analysis, the short processes of analysis, the precise control over conditions of reaction, the ability to automate and integrate, and the ability to consume low amounts of the samples (Haeberle and Zengerle, 2007;Zhang and Xing, 2007;Sohrabi and Moraveji, 2020). ...
Article
Droplet-based microfluidics, including the controlled manipulation of materials in the form of monodisperse droplets on the micrometer scale in microfluidic chips, has gained much attention due to their ability in biological assays. The capillary number of the continuous phase, as the ratio of viscous force and interfacial tension, is so important in the mechanism of droplet formation. This study aims to survey the type of continuous-phase fluid, the change in the parameters of the continuous phase, and the effect of the different flow rates ratios on the droplet formation dynamics, droplet size, distance between droplets and frequency of droplet formation. Three types of oil, including mineral, olive, and paraffin in a flow-focusing microfluidic device have been utilized numerically and experimentally. To characterize the parameters of the continuous phase affecting the droplet size, a microfluidic device has been designed to estimate of the droplet size for three different types of oil used in the assay. The results show that for a constant continuous phase velocity, the diameter of the droplets formed in olive oil has the lowest amount, followed by paraffin and mineral oil, respectively. Regarding the effects of oil type on the frequency of droplet formation and the distance between droplets, the rate of droplet formation in mineral oil, in the ratios of the variable flow rate and the constant continuous phase velocity, is the lowest and, the distance between the droplets has the highest level.
Chapter
As described in the previous chapter, tip streaming can be produced by accelerating the fluid in the tip of a mother drop. Stresses of hydrodynamic or electrohydrodynamic nature drive this acceleration to the extent of overcoming the resistance offered by the capillary and viscous forces. This phenomenon has modest practical (technological) applications due to the intrinsic unsteady character of the process. Besides, the experimenter has limited control over the flow outcome (for instance, the size of the emitted droplets), which can be achieved only by fixing the flow rate at which the dispersed phase is ejected. When the droplet is attached to a capillary and fed at the appropriate flow rate, the system eventually adopts either the microdripping or microjetting mode. Microdripping periodically emits tiny droplets of almost equal size from the droplet tip, while microjetting steadily ejects a thin, long, fluid thread. Microdripping and microjetting can produce droplets with the desired morphology, size, and electrical charge. This chapter describes the axisymmetric microfluidic configurations used to produce microdripping and microjetting. Specifically, we consider the cone-jet mode of electrospray and the coflowing, flow focusing, and confined selective withdrawal configurations. The characteristics of the selective withdrawal and electrified films are also discussed. We introduce and explain the meaning of the dimensionless numbers characterizing the corresponding flows. The numerical and experimental results will be discussed in the following chapters.
Article
Compared with the conventional emulsification method, droplets generated within microfluidic devices exhibit distinct advantages such as precise control of fluids, exceptional monodispersity, uniform morphology, flexible manipulation, and narrow size distribution. These inherent benefits, including intrinsic safety, excellent heat and mass transfer capabilities, and large surface‐to‐volume ratio, have led to the widespread applications of droplet‐based microfluidics across diverse fields, encompassing chemical engineering, particle synthesis, biological detection, diagnostics, emulsion preparation, and pharmaceuticals. However, despite its promising potential for versatile applications, the practical utilization of this technology in commercial and industrial is extremely limited to the inherently low production rates achievable within a single microchannel. Over the past two decades, droplet‐based microfluidics has evolved significantly, considerably transitioning from a proof‐of‐concept stage to industrialization. And now there is a growing trend towards translating academic research into commercial and industrial applications, primarily driven by the burgeoning demands of various fields. This paper comprehensively reviews recent advancements in droplet‐based microfluidics, covering the fundamental working principles and the critical aspect of scale‐up integration from working principles to scale‐up integration. Based on the existing scale‐up strategies, the paper also outlines the future research directions, identifies the potential opportunities, and addresses the typical unsolved challenges.
Article
Full-text available
The hydrodynamics and mass transfer in the Taylor-Couette mixing field are important activities for concurrent process intensification work for miniaturization of the chemical reactor volume and enhancing the energy input per unit volume. In the literature, many studies have been reported for mixing of conventional two-phase systems of industrial interest. However reported studies on mixing of PUREX aqueous organic pair in Taylor-Couette field are rare and few. Utilizing an indigenously developed Taylor-Couette column, hydrodynamic and mass transfer studies with PUREX aqueous-organic pair are reported in this study. Mass transfer and hydrodynamics runs were conducted with PUREX aqueous-organic pair. Aqueous dispersed conditions were employed for d→c mass transfer. Mass transfer performance was highly dependent on rotor speed till a critical value. Macro, meso and micromixing time values were evaluated for Taylor-Couette mixing. Prototyping of the system was completed with a special run with 0.1 M TODGA/n-dodecane nitric acid pair for mutual separation of Cs(I) and Sr(II).
Article
3D simulations have been achieved on a flow‐focusing geometry employing the VOF method to study the consequence of viscosity, surface tension, wettability, and geometry on drop generation for the dripping regime. Here the dispersed phase is the PDMS oil (polydimethylsiloxane), and the continuous phase is the water. Simulations were performed at different oil‐to‐water viscosity ratios of 3, 12, 27, and 50. The interfacial tension between PDMS oil and water is 0.0118 N/m. It has been abridged to 0.008 N/m, 0.005 N/m, and 0.002 N/m, and simulations were performed. The walls of the microchannel are considered to be PMMA surfaces. The contact angle of an oil droplet on the PMMA surface in the presence of water is 140°. The effect of wettability was shown at various contact angles (angle created by water droplet on the PMMA surface in the presence of oil) of 0°, 40°, 90°, 135° and 180°. The frequency of droplet generation (1/s), non‐dimensional droplet length (L/W c ), droplet volume (nl), and droplet velocity (m/s) have been calculated for each of the cases. A flow pattern map has been industrialized classifying the dripping and jetting regimes. A comparison between normal geometry and two constricted geometries (having different orifice lengths) based on the frequency of droplet, non‐dimensional drop length, drop volume, and drop velocity has been made for both dripping and jetting regimes. Prediction of simulated non‐dimensional droplet length has also been made using dimensional analysis.
Chapter
When bubbles or drops move and deform through a continuous liquid phase as in suspension flows, and a surface active species (surfactant) is present in the suspension, the surfactant absorbs onto the fluid particle interfaces. Adsorption lowers the surface tension in proportion to the surface concentration. Surfactant adsorbed onto the surfaces of moving fluid particles is partitioned by the surface flow and the interfacial deformation,creating surface tension gradients. The gradients exert a Marangoni traction on the continuous phase which affects the suspension flow. In this chapter we present a basis for the understanding and fluid mechanical modeling of the influence of Marangoni stresses. Attention is focused on the air/water inerface and the motion of bubbles. The chapter first describes the phase behavior of surfactant monolayers, the experimental methods used to identify phase polymorphism, and the principal phases for bulk soluble surfactants: gaseous, liquid expanded, liquid condensed and solid. Next, we describe the modeling of the state equations for the dependence of the tension on the surface concentration, and the rate equations for the exchange of surfactant between the surface and the adjoining sublayer. The measurement of the equation of state parameters and kinetic rate constants is described next, and we examine the surfactant systems which have been studied. The chapter concludes with a study of the buoyancy driven, inertialess motion of a bubble in a continuous liquid phase, as a model problem for the illustration of the inclusion of Marangoni stresses in a fluid particle flow. We compute the terminal velocity as a function of the surfactant transport properties.
Article
We describe an experimental technique for the production of highly monodisperse emulsions (with minimum achievable polydispersities <3%). The phase to be dispersed is introduced into a coflowing, surfactant-laden continuous phase via a tapered capillary. Drops detach from the capillary when the streamwise forces exceed the force due to interfacial tension. Drop size is a function of the capillary tip diameter, the velocity of the continuous phase, the extrusion rate, and the viscosities and interfacial tension of the two phases. Emulsions composed of a variety of fluids and with drop sizes ranging from 2 to 200 μm have been produced using this technique.
Article
We present a brief review of the application of boundary integral methods in two dimensions to multicomponent fluid flows and multiphase problems in materials science. We focus on the recent development and outcomes of methods which accurately and efficiently include surface tension. In fluid flows, we examine the effects of surface tension on the Kelvin–Helmholtz and Rayleigh–Taylor instabilities in inviscid fluids, the generation of capillary waves on the free surface, and problems in Hele-Shaw flows involving pattern formation through the Saffman–Taylor instability, pattern selection, and singularity formation. In materials science, we discuss microstructure evolution in diffusional phase transformations, and the effects of the competition between surface and elastic energies on microstructure morphology. A common link between these different physical phenomena is the utility of an analysis of the appropriate equations of motion at small spatial scales to develop accurate and efficient time-stepping methods.
Article
Time-dependent potential flows of a liquid with a free surface are considered, with surface tension the force that drives them. Two types of configuration are analyzed, in each of which the flow and the free surface are self-similar at all times. One is a model of a breaking sheet of liquid. The other is a model of the flow near the intersection of the free surface of a liquid with a solid boundary. In both flows, the velocities are found to be proportional to ( sigma / rho )**1**/**3, where sigma is the surface tension, rho is the liquid density and t is the time from the start of the motion. Each free surface is determined by converting the problem to an integrodifferential system of equations for the free surface and the potential on it. This system is discretized and solved numerically On the resulting surfaces there are waves, which are also calculated analytically.
Article
A spherical drop, placed in a second liquid of the same density and viscosity, is subjected to shear between parallel walls. The subsequent flow is investigated numerically with a volume-of-fluid continuous-surface-force algorithm. Inertially driven breakup is examined. The critical Reynolds numbers are examined for capillary numbers in the range where the drop does not break up in Stokes flow. It is found that the effect of inertia is to rotate the drop toward the vertical direction, with a mechanism analogous to aerodynamic lift, and the drop then experiences higher shear, which pulls the drop apart horizontally. The balance of inertial stress with capillary stress shows that the critical Reynolds number scales inversely proportional to the capillary number, and this is confirmed with full numerical simulations. Drops exhibit self-similar damped oscillations towards equilibrium analogous to a one-dimensional mass-spring system. The stationary drop configurations near critical conditions approach an inviscid limit, independent of the microphysical flow- and fluid-parameters.
Article
A spherical drop, placed in a second liquid of the same density, is subjected to shearing between parallel plates. The subsequent flow is investigated numerically with a volume-of-fluid (VOF) method. The scheme incorporates a semi-implicit Stokes solver to enable computations at low Reynolds number. Our simulations compare well with previous theoretical, numerical, and experimental results. For capillary numbers greater than the critical value, the drop deforms to a dumbbell shape and daughter drops detach via an end-pinching mechanism. The number of daughter drops increases with the capillary number. The breakup can also be initiated by increasing the Reynolds number.
Article
Previous long-wavelength analyses of capillary breakup of a viscous fluid thread in a perfectly inviscid environment show that the asymptotic self-similar regime immediately prior to breakup is given by a balance between surface tension, inertia, and extensional viscous stresses in the thread. In contrast, it is shown here that if viscosity in the external fluid, however small, is included then the asymptotic balance is between surface tension and viscous stresses in the two fluids while inertia is negligible. Scaling estimates for this new balance suggest that both axial and radial scales decrease linearly with time to breakup, so that the aspect ratio remains O(1) with time but scales with viscosity ratio like (μint/μext)1/2 for μint≫μext, where μint and μext are the internal and external viscosities. Numerical solutions to the full Stokes equations for μint=μext confirm the scalings with time and give self-similar behavior near pinching. However, the self-similar pinching region is embedded in a logarithmically large axial advection driven by the increasing range of scales intermediate between that of the pinching region and that of the macroscopic drop. The interfacial shape in the intermediate region is conical with angles of about 6° on one side and 78° on the other.
Article
Rising droplets with coalescence and breakup are numerically simulated using the lattice BGK method. It is shown for single droplet that the rising velocities are in good agreement with those obtained by two types of empirical correlations. The coalescence of droplets is shown when two droplets are placed vertically in line at different elevations. The breakup of droplet is observed in some cases after the coalescence. It is found that the breakup of coalesced droplet occurs when the Weber number at the coalescence exceeds a critical value, and the critical Weber number agrees well with that given by the empirical correlation.
Article
When two small bubbles approach each other, a dimpled thin liquid film is formed between them. A model is developed for the dynamics of the thinning film with mobile interfaces, in which the effects of surface tension gradient and surface viscosities upon the surface mobility, drainage, and rupture of a dimpled liquid film are investigated. The model predicts the coalescence time, which is the time required for the thinning and rupture of the liquid film, given only the radii of bubbles and the required physical properties of the liquid and interfaces, such as surface tension and surface viscosities, London–van der Waals constant, and surface diffusion coefficient. The predicted result is in good agreement with previous experimental results.
Article
When two small drops or bubbles approach one another, a dimpled thin liquid film is formed between them. A model is developed for the dynamics of the thinning film in which both London-van der Waals and electrostatic double layer forces are recognized. The model predicts the coalescence time, which is the time required for the thinning and rupture of the liquid film, given only the radii of the drops and the required physical properties of the fluids and interfaces. In the absence of the electrostatic double layer forces, a more general expression is derived for the coalescence time. The predicted result is in good agreement with previous experimental results.