You are on page 1of 23

Droplet dynamics in mixed flow conditions: Effect of

shear/elongation balance and viscosity ratio


Elia Boonen, Peter Van Puyvelde, and Paula Moldenaers

Citation: Journal of Rheology (1978-present) 54, 1285 (2010); doi: 10.1122/1.3490661


View online: http://dx.doi.org/10.1122/1.3490661
View Table of Contents: http://scitation.aip.org/content/sor/journal/jor2/54/6?ver=pdfcov
Published by the The Society of Rheology

Articles you may be interested in


Dynamics of collapsed polymers under the simultaneous influence of elongational and
shear flows
J. Chem. Phys. 135, 014902 (2011); 10.1063/1.3606392

Effects of viscosity ratio and three dimensional positioning on hydrodynamic interactions


between two viscous drops in a shear flow at finite inertia
Phys. Fluids 21, 103303 (2009); 10.1063/1.3253351

Effect of confinement and viscosity ratio on the dynamics of single droplets during
transient shear flow
J. Rheol. 52, 1459 (2008); 10.1122/1.2978956

Effects of shear flow on a polymeric bicontinuous microemulsion: Equilibrium and steady


state behavior
J. Rheol. 46, 529 (2002); 10.1122/1.1446883

Three-dimensional shape of a drop under simple shear flow


J. Rheol. 42, 395 (1998); 10.1122/1.550942

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
Droplet dynamics in mixed flow conditions: Effect of
shearÕelongation balance and viscosity ratio

Elia Boonen, Peter Van Puyvelde,a) and Paula Moldenaers

Department of Chemical Engineering and Leuven MRC, Katholieke Universiteit


Leuven, Willem de Croylaan 46, 3001 Leuven, Belgium

(Received 14 July 2009; final revision received 12 August 2010;


published 5 October 2010兲

Synopsis
The dynamics of single droplets dispersed in a second, immiscible liquid undergoing a controlled
mixture of shear and elongational flow has been studied using a home made eccentric cylinder
device. The model system consists of poly共dimethyl siloxane兲 droplets in a poly共isobutylene兲
matrix, both Newtonian liquids at room temperature. In continuation of previous work 关Boonen
et al., Rheol. Acta 48, 359–371 共2009兲兴, the effect of changing the balance of shearing and
elongational flow components and varying viscosity ratio on the deformation and orientation of
the droplets has been systematically investigated under sub-critical flow conditions. The
experimental results obtained from optical microscopy are compared with theoretical predictions of
the phenomenological model by Maffettone and Minale 关J. Non-Newtonian Fluid Mech. 78,
227–241 共1998兲兴, obtained using the transient form of the model and incorporating a flow type
parameter that accounts for the relative amount of extension in the flow. Overall, a fair agreement
was found between the model predictions and the experimental results for all sub-critical mixed
flows applied and all viscosity ratios investigated here. This work provides an experimental
reference data set which can be used to guide future modeling efforts. © 2010 The Society of
Rheology. 关DOI: 10.1122/1.3490661兴

I. INTRODUCTION
The behavior of droplet dispersions is of interest for many industrial applications, e.g.,
food emulsions, cosmetics, pharmaceuticals, and polymer blending. An understanding of
the droplet dynamics in such systems helps explaining the rheological behavior of the
flowing emulsions, such as viscoelasticity 共even if the components are Newtonian兲 and a
shear-dependent viscosity. Conversely, from the rheology of the dispersion substantial
information about the morphology development can be obtained 关e.g., Vinckier et al.
共1996兲兴. Since the pioneering work of Taylor 共1932, 1934兲, a large number of experimen-
tal and theoretical studies on droplet dispersions have been carried out. These are well
described in several recent reviews 关Guido and Greco 共2004兲; Ottino et al. 共1999兲; Stone
共1994兲; Tucker and Moldenaers 共2002兲兴. Most studies have focused on the structure
development in either purely elongational flow or simple shear flow. However, since most

a兲
Author to whom correspondence should be addressed; electronic mail: peter.vanpuyvelde@cit.kuleuven.be

© 2010 by The Society of Rheology, Inc.


J. Rheol. 54共6兲, 1285-1306 November/December 共2010兲 0148-6055/2010/54共6兲/1285/22/$30.00 1285

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
1286 BOONEN, VAN PUYVELDE, and MOLDENAERS

FIG. 1. 共a兲 Schematic of eccentric cylinder device. 共b兲 Sketch of top view of eccentric cylinder geometry.

industrial processing operations consist of complex mixtures of shear and elongation, a


systematic investigation of the microstructure evolution in such “mixed flow” conditions
is considered to be very relevant.
In our previous work 关Boonen et al. 共2009兲兴, a summary has been given of the prior
research conducted on the morphology development of droplet dispersions in controlled
complex flow fields. Important to recall here is the work of Leal and co-workers 关Bentley
and Leal 共1986兲; Stone et al. 共1986兲; Stone and Leal 共1989兲兴, who explored a whole range
of two-dimensional 共2D兲 flows using a computer controlled four-roll mill. They studied
droplet deformation and break-up in different types of steady flow, ranging from simple
shear to pure 2D-elongation, and provided a systematic set of data for the droplet dy-
namics, in these complex flow types. Godbille and Picot 共2000兲 and Khayat et al. 共2000兲
studied the influence of shear and elongation on drop deformation and break-up in
convergent-divergent channels. They found that the initial droplet diameter is of major
importance for the droplet deformation in this specific geometry, and different break-up
mechanisms, driven by shear and extensional flows, respectively, were identified. An
alternative set-up to study the morphology development in controlled mixed flow condi-
tions is the so called eccentric Couette system, as shown in Fig. 1. This type of flow has
previously been studied extensively in fluid mechanics 关e.g., Ballal and Rivlin 共1976兲;
Diprima and Stuart 共1972兲; Wannier 共1950兲兴 and has received renewed interest from
Windhab and co-workers 关Feigl et al. 共2003兲; Kaufmann et al. 共2000兲; Windhab et al.
共2005兲兴. Feigl et al. 共2003兲, for instance, investigated, through experiments and numerical
simulations, the drop deformation and break-up when only the inner cylinder rotates at a
constant speed. However, they neglected the wall effects in their study, which may be
questionable, in view of the ratio of droplet diameter to gap size used 关see Van Puyvelde
et al. 共2008兲; Vananroye et al. 共2006兲兴. Finally, Egholm et al. 共2008兲 used a rather similar
geometry to explore droplet dynamics in complex flow conditions. Their flow channel
consists of two concentric cylinders with toothed walls as a model for extruding flow.
They reported that for small deformations, the relation between the time-averaged drop
deformation and a time-averaged apparent shear rate can be described by Taylor’s small
deformation theory. Also, numerical simulations agreed fairly well with the experimental
results, although the calculations predict a somewhat higher deformation than experimen-
tally observed.
Here, we use a newly designed eccentric cylinder device 共ECD兲 to study the defor-
mation and orientation of single Newtonian droplets dispersed in a Newtonian matrix
undergoing controlled mixed flow conditions. Details of the cell can be found in Boonen
et al. 共2009兲. For such a system with matching fluid densities and Newtonian compo-

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
DROP DYNAMICS IN MIXED FLOWS 1287

TABLE I. Physical properties of component fluids.

␩0 共23 ° C兲 p = ␩PDMS / ␩PIB Activation energy Ea


Material 共Pa s兲 共23 ° C兲 共kJ/mole兲

PIB 51.6 Matrix 60.1


Glissopal 1300
PDMS 60.4 1.17 15.4
Rhodorsil 47V60.000
PDMS 501.6 9.72 15.3
Rhodorsil 47V500.000
PDMS 5.10 0.10 15.2
Rhodorsil 47V5000

nents, only three dimensionless parameters play a role for the dynamics in slow flows: the
viscosity ratio p = ␩d / ␩m, where ␩d is the viscosity of the drop fluid and ␩m is the
viscosity of the matrix, the capillary number Ca which represents the ratio of viscous to
interfacial tension stresses, and the flow type which can be represented by the flow type
parameter ¯␣ 共see Sec. II兲. Ca can be defined as 共␩mR0E兲 / ⌫, where R0 is the radius of the
undistorted spherical droplet, ⌫ is the interfacial tension, and E is the flow intensity
which equals 冑4␧˙ 2 + ␥˙ 2 for 2D flows. In Boonen et al. 共2009兲, the experimental results for
a viscosity ratio of p = O共1兲 have been compared with predictions obtained using the
transient form of the model of Maffettone and Minale 共1998兲 adapted to complex flows
关see also Sec. II兴. Under sub-critical flow conditions, good agreement was found between
model predictions and experimental data, providing a quantitative assessment of drop
shape predictions in controlled complex flows. In the present study, a systematic inves-
tigation is made of the effect of viscosity ratio on the droplet dynamics for a wide range
of viscosity ratios 共ranging from 0.1 to 10兲. In addition, we will explore more extreme
flow conditions with an enhanced extensional contribution as compared to the previous
study 关Boonen et al. 共2009兲兴.

II. EXPERIMENT

A. Materials
The materials used in this work are poly共isobutylene兲 共PIB兲 共Glissopal from BASF兲
and poly共dimethyl siloxane兲 共PDMS兲 共Rhodorsil from Rhodia兲. Since Guido et al. 共1999兲
reported that PIB is slightly soluble in PDMS, PIB is chosen as the matrix phase. In order
to vary the viscosity ratio, three different grades of PDMS have been selected as the
droplet phase. Table I gives the zero shear viscosities ␩0 and the corresponding viscosity
ratios p at 23 ° C. All fluids are transparent at room temperature and exhibit Newtonian
behavior over the range of strain rates investigated here. The interfacial tension ⌫ for the
PIB/PDMS system was measured by three methods 共small deformation theory, droplet
retraction measurements, and pendant drop technique兲 and was found to be
2.8⫾ 0.1 mN/ m. Moreover, this value was measured to be independent of the molecular
weight for the three grades of PDMS used here. All experiments were performed at
ambient temperature 共⬃23 ° C兲; as the viscosity of PIB is very sensitive to temperature,
the temperature of the sample was directly monitored by immersing a fine thermocouple
needle in the continuous phase. The viscosities ␩m and ␩d, and the resulting viscosity
ratio p could be back-calculated using an Arrhenius equation.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
1288 BOONEN, VAN PUYVELDE, and MOLDENAERS

The density difference between both polymers is quite small 共about 0.08 g / cm3兲 so
that in combination with the high viscosity of the matrix, gravitational effects can be
neglected. Furthermore, the difference in refractive indices is high enough to have good
contrast for observations by optical microscopy.

B. Methods
The experiments are performed using a home built ECD which is equipped with a
microscope and camera system to enable visualization of droplet dynamics in controlled
complex flows. A detailed discussion of the experimental set-up has been given in
Boonen et al. 共2009兲. In Fig. 1共a兲, a schematic of the basic ECD is shown. In brief, the
apparatus allows applying controlled “mixed flow” conditions, i.e., a combination of
shear and elongational components, in the narrowing and expanding areas of the gap
between the rotating, eccentrically positioned cylinders. Figure 1共b兲 shows a 2D sketch of
the geometry, consisting of an inner and an outer cylinder with radii Ri and Ru and
angular velocities ␻i and ␻u, respectively. The axes of these cylinders can be displaced by
a certain distance e, the eccentricity. The eccentricity ratio X of the system is then defined
as e / 共Ru − Ri兲. This is an important, adjustable parameter that influences the relative, as
well as the absolute magnitudes of shear and elongational strain rates along the stream-
lines of the flow field. In addition, the flow configuration, i.e., inner cylinder rotating,
outer cylinder rotating, counter-rotating, or co-rotating, also has a profound effect on the
type of mixed flow obtained 共see further兲. For the experiments conducted in this study,
the ECD has been used in two configurations: configuration A in which only the outer
cylinder is rotating with a constant speed, and configuration B where both cylinders are
co-rotating with the ratio of outer to inner cylinder velocity ␻u / ␻i equal to 1.5. For both
configurations the eccentricity ratio X is set to 0.2. In configuration A, a shear dominated
type of flow is obtained, while configuration B produces a flow with an enhanced exten-
sional contribution, which allows exploring more extreme flow conditions.
Single PDMS droplets with initial radii R0 in the range of 300– 700 ␮m were intro-
duced in the matrix material by using a home made injection system. The spherical drop
is injected in the widest part of the gap and positioned so that its center is at a radial
distance around 30 mm 共half the clearance兲 from the axis of the inner cylinder. Upon
start-up of the flow, the droplet deformation and orientation is visualized during several
revolutions using optical microscopy 共Olympus SZ61-TR stereo microscope兲. The micro-
scope, equipped with a CCD camera 共Basler A301f兲, is mounted on top of the flow cell
so that images are captured in the velocity-velocity gradient plane 共see Fig. 2兲 using the
Streampix Digital Video Recording Software 共NORPIX兲. Due to the limited field of view
of the microscope and difficulties in illumination, images of the deforming drop could, in
a single experiment, only be captured in the first two 共I and II兲 or the last two 共III and IV兲
quadrants of the Cartesian coordinate system, defined in Fig. 1共b兲. In order to obtain the
shape of the droplet, an ellipse is fitted to the recorded images after performing basic
image processing, including a convolution operation and threshold using IMAGEJ software
for Windows 关Rasband 共1997兲兴. Throughout the experiments, the flow intensity is kept
below the critical conditions for breakup, so that no significant deviations from the
ellipsoidal shape are present. Using this approach, it is then possible to obtain the defor-
mation and orientation of the droplets, i.e., the length L of the long axis, the length B of
the short axis, and the angle ␪ between the long axis and the flow direction, as shown in
Fig. 2.
The experimental results are compared with the predictions of the Maffetone–Minale
model 关Maffettone and Minale 共1998, 1999兲兴, which has been applied to complex flows

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
DROP DYNAMICS IN MIXED FLOWS 1289

FIG. 2. Schematic view of a deformed droplet.

关Boonen et al. 共2009兲兴. This phenomenological model assumes that during flow the drop
shape is ellipsoidal at all times and that its volume is preserved. The drop shape can
therefore be described by a symmetric, positive definite second order tensor S, which
eigenvalues represent the square semi-axes of the ellipsoid. The evolution equation of S,
resulting from the competing actions of the hydrodynamic forces and the interfacial
tension, is given by 关Maffettone and Minale 共1998兲兴

dS f MM
− 共W · S − S · W兲 = − 1 共p兲共S − gMM共S兲I兲 + f MM
2 共p,Ca兲共D · S + S · D兲, 共1兲
dt ␶
where the characteristic time ␶ equals ␩mR0 / ⌫, I is the second rank unit tensor, and D
and W are the deformation rate and vorticity tensor of the flow field. The characteristic
time ␶ will be used to define a dimensionless time tⴱ = t / ␶ in which t is the time since the
start-up of the flow. The scalar function g MM is required to preserve drop volume, while
the functions f 1MM and f 2MM have been determined to recover the asymptotic analytic
limits 关Maffettone and Minale 共1998兲; Taylor 共1934兲兴:

40共p + 1兲
f MM
1 =
共3 + 2p兲共16 + 19p兲

5 3Ca2 1
f 02 = , f c2 = ,
3 + 2p 2 + 6Ca2+␦ 1 + ␧p2

f MM
2 = f 02 + f c2 ,

3IIIS
gMM = . 共2兲
IIS
Here, IIS is the second scalar invariant of S, and ␦ and ␧ are the small positive numbers
which, for all practical purposes when Ca is not too large and p is far from infinity, are
set to zero. The flow field generated in the ECD can be represented by a deformation rate
and vorticity tensor of the form

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
1290 BOONEN, VAN PUYVELDE, and MOLDENAERS

冢 冣 冢 冣
1 − 兩¯␣兩 1 − 兩¯␣兩
¯␣ 0 0 0
2 2
D = Ḡ 1 − 兩¯␣兩 , W = Ḡ − 1 + 兩¯␣兩 . 共3兲
− ¯␣ 0 0 0
2 2
0 0 0 0 0 0
Here, the flow type parameter ¯␣ expresses the relative amount of elongation present in
the flow and is defined as 关Feigl et al. 共2003兲兴

␧˙ ␧˙
¯␣ = = , 共4兲
兩␥˙ 兩 + 兩␧˙ 兩 Ḡ

with Ḡ the sum of absolute shear and elongation rates and −1 ⱕ ¯␣ ⱕ 1. Mixed flow
conditions between eccentric cylinders can thus be represented by the flow strength Ḡ共t兲
and the flow parameter ¯␣共t兲 which in general vary in time. In this work, we will then refer
to the MM model as the model predictions according to the evolution equation, Eq. 共1兲,
with the parameters given by Eq. 共2兲 and with the flow field represented by Eq. 共3兲.

III. RESULTS AND DISCUSSION


In Sec. III A and III B, the experimental results on droplet deformation obtained in
two types of sub-critical complex flows with different relative contributions for shear and
elongation, and for three viscosity ratios, are presented and discussed. These results will
be compared to the model predictions of the MM model as described above to help
elucidate the effect of flow type and viscosity ratio in these mixed flow conditions.

A. Effect of shear/elongation balance


First, the dynamics of droplets with a constant viscosity ratio of about 1.2 in two
different types of mixed flow are considered: shear dominated flow in configuration A of
the ECD 共Figs. 3 and 4兲 and a flow field with an enhanced extensional component
obtained in configuration B 共Figs. 5 and 6兲. In both cases, the transient droplet deforma-
tion is observed in the first two quadrants of the flow field 关Fig. 1共b兲兴 for a number of
revolutions during start-up of the flow. Furthermore, by varying the velocities of the
cylinders and changing the initial droplet radius R0, a range of capillary numbers can be
explored.

1. Configuration A: Shear dominated flow


Due to the motion through the ECD, the drop is subjected to a periodic flow field.
Hence a new characteristic time-scale, the oscillation period of the imposed flow, comes
into play. The droplet dynamics as reported in this paper are not influenced by the
presence of this additional time-scale. For instance, in the experiments presented in Fig.
3, the droplet relaxation time is about 0.7 s, whereas the ECD rotation period is 250 s.
These time-scales are well separated indicating that the droplet can rapidly respond to the
applied deformation and that the droplet shape is fully determined by the deformation at
that instant.
For the shear dominated flow of configuration A, the results at a viscosity ratio of 1.2
have been discussed in detail elsewhere 关Boonen et al. 共2009兲兴. Typical profiles for the
strain rates and the corresponding flow parameters Ca共t兲 and ¯␣共t兲 for a PDMS drop with
a representative initial radius of 375 ␮m are shown in Fig. 3. The shear rate is seen to

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
DROP DYNAMICS IN MIXED FLOWS 1291

FIG. 3. Example of flow parameters for configuration A with ␻i = 0, ␻u = 0.025 rad/ s and eccentricity ratio
X = 0.2, along streamline starting at x = 30 mm from the axis of the inner cylinder: 共a兲 strain rate profiles, shear
共–兲 and elongation 共–兲; 关共b兲 and 共c兲兴 evolution of Ca and ¯␣ for a PDMS drop with initial radius of 375 ␮m; 共d兲
example of a deforming droplet at different points along the trajectory.

reach a sharp maximum when the gap is at its narrowest 共point C兲 and displays a broad
minimum in the widest part of the gap 共point A兲. These features also show up in the
Ca-profile shown in Fig. 3共b兲. Here, the flow conditions are mostly shear dominated,
which is also evident from the values of the flow type parameter ¯␣ in Fig. 3共c兲. The
average amount of elongation ¯␣avg,I−II over the first two quadrants 关I–II in Fig. 1共b兲兴 is

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
1292 BOONEN, VAN PUYVELDE, and MOLDENAERS

FIG. 4. Droplet dynamics as a function of dimensionless time tⴱ for the flow conditions of Fig. 3 and for a
viscosity ratio of p ⬇ 1.2. The symbols represent the experimental results, whereas the lines are the predictions
according to the Maffettone–Minale model.

about 10% of the sum of strain rates 关Ḡ兴. The flow type parameter ¯␣ logically follows the
same profile as the elongational strain rate, shown in Fig. 3共a兲: both exhibit a symmetric
profile with a positive 共stretching兲 maximum and an equal but opposite, negative 共com-
pression兲 minimum in the converging and diverging parts of the gap. The results for
droplet deformation under these flow conditions are summarized in Fig. 4 together with
the model predictions according to the MM model for complex flows 关Eqs. 共1兲–共3兲兴.
It is observed that the experimental deformation parameters L / 2R0 and B / 2R0, and the
orientation angle ␪ show a steady oscillation in response to the applied, time periodic
capillary numbers Ca共t兲 共see Fig. 3ii兲. Furthermore, good quantitative agreement is found
between the experimental results and the model predictions. Quantitative deviations only
start to occur at Ca-numbers that are expected to be near the critical conditions for
break-up. It should be noticed that the experiments are characterized by the average
capillary number Caavg over one rotation rather than by the maximum and minimum
values Camax and Camin, as in Boonen et al. 共2009兲.

2. Configuration B: Enhanced extensional contribution


For configuration B, in which both cylinders are co-rotating with the ratio ␻u / ␻i set to
a value of 1.5, the flow field comprises an enhanced extensional contribution. Here, the
flow conditions are drastically different from the one previously discussed, as shown in
Fig. 5. In this case, the shear rate 关see Fig. 5共a兲兴 displays a minimum peak at the point
where the gap is the narrowest 共point C兲 and a broad maximum in the widest part of the
gap 共point A兲. For this specific flow configuration, in some parts of the gap, the velocity
profiles in the radial direction appear the show a nonmonotic behavior reaching an ex-
tremum value. This could explain this counterintuitive finding that the shear rate is

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
DROP DYNAMICS IN MIXED FLOWS 1293

FIG. 5. Example of flow parameters for configuration B with ␻i = 0.02, ␻u = 0.03 rad/ s and eccentricity ratio
X = 0.2, along streamline starting at x = 35 mm from the axis of the inner cylinder: 共a兲 strain rate profiles, shear
共–兲 and elongation 共–兲; 关共b兲 and 共c兲兴 evolution of Ca and ¯␣ for a PDMS drop with initial radius of 500 ␮m.

FIG. 6. Droplet dynamics as a function of dimensionless time tⴱ for the flow conditions of Fig. 5 and for a
viscosity ratio of p ⬇ 1.2. The symbols represent the experimental results, whereas the lines are the predictions
according to the Maffettone–Minale model.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
1294 BOONEN, VAN PUYVELDE, and MOLDENAERS

maximum in the widest part of the gap. The elongation rate on the other hand shows the
same type of symmetric profile as in Fig. 3共a兲, being positive in the converging part and
negative in the diverging part of the gap. The main difference as compared to the previ-
ous configuration is that the extensional strain rate has the same order of magnitude as the
shear rate during a substantial part of the profile. This results in relatively large values for
the flow type parameter ¯␣ 关shown in Fig. 5共c兲兴. This parameter follows the same sym-
metric profile as in Fig. 3共c兲, but with significantly higher maximum values of about
⫾0.50 in quadrants II and III. As a consequence, the average amount of elongation
¯␣avg,I−II in the first two quadrants is now increased to approximately 30% of the sum of
strain rates, as compared to only 10% for the previous configuration. The profile for the
capillary number Ca also differs substantially from that in Fig. 3, as can be seen in Fig.
5共b兲. The Ca-number is more or less constant in the first quadrant of the flow field,
subsequently drops rapidly to reach a sharp minimum at the end of quadrant II 共point C兲,
and symmetrically continues in quadrant III and IV. It will be verified now to what extent
the drastically changed flow field will affect the droplet dynamics.
Figure 6 shows the results for droplet deformation observed in quadrants I–II during
start-up, as well as the model predictions according to the MM model. It can be seen that
the deformation parameters L / 2R0 and B / 2R0, and the orientation angle ␪ again seem to
follow an oscillating pattern for subsequent revolutions, in nice agreement with the model
predictions. In this case, however, the deformation strongly relaxes near the end of the
second quadrant as shown by the decrease and increase 共especially for the lower Ca
values兲 of the major axis L and the minor axis B of the droplet, respectively. Furthermore,
the orientation angle starts to increase at the end of the second quadrant, indicating the
droplet rotates away from the velocity direction. This decreased deformation and orien-
tation corresponds to the steep decline of the capillary number Ca at the end of quadrant
II 关see Fig. 5共b兲兴. As the Ca-number drops, the driving force for deformation and orien-
tation of the droplets decreases, and due to the action of the interfacial tension the droplet
relaxes to a less deformed and less oriented state. For the orientation angle, the model
predictions even indicate 共for this range of Ca-numbers兲 that the droplet rotates past 45°,
which is the asymptotic limit for small Ca in simple shear flow 关Taylor 共1934兲兴. This
peculiar effect cannot be explained entirely by the evolution of Ca. To investigate this
phenomenon more thoroughly, experiments were also performed in the third and fourth
quadrants of the flow field, the droplet still being injected at point A.
The results are shown in Fig. 7; the relaxation of the droplet is clearly confirmed in the
evolutions of L / 2R0 and B / 2R0 depicted in Figs. 7共a兲 and 7共b兲. In addition, the orienta-
tion angle quickly rises from the velocity direction toward the perpendicular direction to
a maximum value above 45°, as predicted by the MM model. This effect can be attributed
to the local change in flow type the droplet experiences in this flow field. At the end of
the second quadrant the elongation rate changes sign and becomes negative 共compression
flow兲, while the shear rate remains quite low for some time 关see Fig. 5共a兲兴. As a conse-
quence a local flow field with a large, negative ¯␣-value exists for some time. This tends
to rotate the drop toward the 90° limit, resulting in a maximum orientation angle above
45°. In summary, the balance of elongational and shear components in the flow field
clearly affects droplet dynamics in transient mixed flow conditions. Furthermore, it has
been shown that this effect can be correctly predicted by the MM model.

B. Effect of viscosity ratio


It is known that the viscosity ratio is a crucial parameter for droplet dynamics in
simple flow conditions. Here, we want to investigate the effect of the viscosity ratio p on

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
DROP DYNAMICS IN MIXED FLOWS 1295

FIG. 7. Droplet dynamics as a function of dimensionless time tⴱ for the flow conditions of Fig. 5 and a
viscosity ratio of p ⬇ 1.2, with observations made in quadrants III and IV of the flow field. The symbols
represent the experimental results, whereas the lines are the predictions according to the Maffettone–Minale
model.

droplet dynamics in mixed flow conditions. Hereto, droplets with viscosity ratios of
O共10兲 and O共0.1兲 have been observed during the type of complex flow obtained in
configuration B of the ECD with enhanced extensional contribution 共see Fig. 5兲.

1. High viscosity ratio drops


Figure 8 shows the droplet dynamics for drops with a high viscosity ratio of p ⬇ 10.
Note that observations have only been made in the first two quadrants of the flow field,
and results are for low to intermediate Ca 共higher Ca-numbers will be discussed further
on兲. Qualitatively, the same behavior is observed as for a viscosity ratio around 1.2 共see
Fig. 6兲. The deformation and orientation increase in the first part of the flow field and
start to relax again near the end of the second quadrant, when the capillary number
quickly drops to a lower value 共see Fig. 5兲. It seems though that the relaxation of the
droplet starts somewhat later for this high viscosity ratio case as compared to p ⬇ 1.2.
This is more clearly illustrated in Fig. 9 where the droplet dynamics for p ⬇ 1.2 and p
⬇ 10 are compared for about the same average Ca-number. As expected, the high vis-
cosity ratio drops are less deformed 共lower L, higher B兲. They are also slightly more
oriented toward to the flow direction 共lower ␪兲. Furthermore, nice agreement is obtained
between the model predictions and the experimental results for the low to intermediate
Ca-range for these high viscosity ratio drops.
For higher Ca, as depicted in Fig. 10, deviations with respect to the predictions of the
MM model start to appear for the major and minor axes of the drop, while the predictions
for the angles are still reasonable. Furthermore, for these high capillary numbers the
effect of p on the orientation angle becomes clearer, as shown in Fig. 11. Here, the
experimental results for the orientation of droplets with p around 10 and 1.2 are com-

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
1296 BOONEN, VAN PUYVELDE, and MOLDENAERS

FIG. 8. Droplet dynamics as a function of dimensionless time tⴱ for the flow conditions of Fig. 5 for a viscosity
ratio of p ⬇ 10 and intermediate capillary numbers Ca. The symbols represent the experimental results, whereas
the lines are the predictions according to the Maffettone–Minale model.

pared for an average capillary number of about 0.33. A high viscosity ratio obviously
promotes orientation toward the flow direction, something that is also seen in simple
shear flow 关e.g., Rumscheidt and Mason 共1961兲兴. If we apply even higher Ca-numbers, as
the ones shown in Fig. 12, quantitative deviations between model predictions and experi-
mental results grow even larger for L / 2R0 and B / 2R0, while the angle predictions still
coincide with the experimental data very well. Qualitatively, the predictions for the major
and minor axes still describe the experimental results. In addition, Fig. 12 illustrates that
at this high viscosity ratio, the amplitude of the oscillating deformation parameters varies
in time, as also seen in the model predictions. This peculiar transient oscillatory behavior
for high p in mixed flow conditions is similar to the damped oscillatory behavior ob-
served during start-up of simple shear flow for high viscosity ratios 关e.g., Torza et al.

FIG. 9. Comparison of droplet deformation and orientation for the flow conditions of Fig. 5, for a viscosity
ratio of p ⬇ 10 and p ⬇ 1.2, respectively, for the same average capillary number Caavg. The symbols represent
the experimental results, whereas the lines are the predictions according to the Maffettone–Minale model.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
DROP DYNAMICS IN MIXED FLOWS 1297

FIG. 10. Droplet dynamics as a function of dimensionless time tⴱ for the flow conditions of Fig. 5 for a
viscosity ratio of p ⬇ 10 and high capillary numbers Ca. The symbols represent the experimental results,
whereas the lines are the predictions according to the Maffettone–Minale model.

共1972兲兴 and is not fully understood at present. Note that in Fig. 12 in addition to the MM
model predictions 共represented by the solid lines兲 also model predictions according to the
so called “Minale model” 共corresponding to the dashed lines兲 are shown. The Minale
model 关Minale 共2004兲兴 is an extension of the Maffettone–Minale model to account for
elastic effects in the matrix and/or droplet phase. It has an extra term in the evolution
equation of the droplet shape tensor S with the coefficient f 3. This extra term allows for
oblate droplet configurations, and for instance, enables this model to predict the “widen-
ing” behavior seen for p Ⰶ 1 in transient simple shear flow 关Cristini et al. 共2002兲兴. In the

FIG. 11. Comparison of droplet orientation for the flow conditions of Fig. 5, for viscosity ratios of p ⬇ 10 and
p ⬇ 1.2, respectively, for high capillary numbers Ca.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
1298 BOONEN, VAN PUYVELDE, and MOLDENAERS

FIG. 12. Droplet dynamics as a function of dimensionless time tⴱ for the flow conditions of Fig. 5 for a
viscosity ratio of p ⬇ 10 and very high capillary numbers. The symbols represent the experimental results,
whereas the lines are the predictions according to the Maffettone–Minale model 共solid line兲 and the Minale
model 共dashed line兲.

case of Newtonian components studied here, Minale 共2004兲 showed that the predictions
for steady simple shear flow almost coincide with the MM model over a large range of
Ca and only differ at very high capillary numbers. The evolution equation for S for
Newtonian components is given as
dS f MM
− 共W · S − S · W兲 = − 1 共p兲共S − gMinale共S兲I兲 + f Minale共p,Ca兲共D · S + S · D兲
dt ␶ 2

+
R0

f Minale
3
共p,Ca兲共D · S · S + S · S · D兲 − 共D · S + S · D兲
S:I
3
, 册
where the extra coefficient f 3 depends on the applied Ca and the viscosity ratio p 关note
that gMinale is again a term to preserve the volume and f 2Minale is a slightly adapted form
of f 2MM using a weighting factor for f 2c, as defined in Eq. 共2兲兴. In our study at high
viscosity ratio and high capillary number, the extra term in the Minale model seems to
give different predictions as compared to the MM model, as shown in Fig. 12. From this
figure, it can be concluded that the Minale model predictions show somewhat better
agreement with the experimental profiles of L / 2R0 and B / 2R0, although not quite satis-
factorily yet. For the orientation angle on the other hand, the predictions of the Minale
model and the MM model are practically on top of each other and agree very well with
the experimental values.
Considering both models in more detail, one might suspect that the discrepancy is
related to the behavior of another 共almost兲 mutual coefficient of both models: f 2. The

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
DROP DYNAMICS IN MIXED FLOWS 1299

FIG. 13. Effect of parameter ␧ on droplet dynamics for the flow conditions of Fig. 5, a viscosity ratio of p
⬇ 10 and large capillary number Ca. The symbols represent the experimental results, whereas the lines are the
predictions according to the Maffettone–Minale model 共solid line兲, the Minale model 共dashed line兲, and the
Minale model with ␧ = 0.001 共dotted line兲 and ␧ = 0.01 共dash-dotted line兲.

parameters f 2MM and f 2Minale depend on p as well as Ca and consists of two terms 关see
Eqs. 共2兲兴. One term, f 20, is determined to recover the asymptotic limits of Taylor 共1934兲.
The other term, f 2c, is an empirical factor added by Maffettone and Minale to improve the
predictions in simple shear flow and used in the Minale model with a weighting factor of
0.75. The factor f 2c contains two empirical constants ␧ and ␦, which for all practical
purposes, when Ca and p are far from infinity, have been set to zero by Maffettone and
Minale. In theory though, these parameters have to be small, positive numbers to be able
to recover the asymptotic limits of p → ⬁, and of affine behavior for p = 1 and Ca → ⬁,
respectively. This could mean that, in our case of high p, a small but nonzero value of the
parameter ␧ has to be chosen in order to reflect the behavior of the high viscosity ratio
droplets. In Fig. 13, we have verified the effect of the parameter ␧ on the predictions of
droplet dynamics by the Minale model for the high viscosity ratio drops 共p ⬇ 10兲 for a
large Ca. By assigning a small positive value to ␧ 共10−3 and 10−2 in Fig. 13兲, the
predictions for L / 2R0 and B / 2R0 are indeed improved as compared to the “normal” MM
and Minale model predictions. For the predictions of the orientation angle, on the other
hand, hardly any effect of ␧ is observed. This indicates that it is the high viscosity ratio
effect that causes deviations from the MM model predictions because we are reaching the
asymptotic limit for high p for which a small, nonzero value of the parameter ␧ has to be
set in the coefficient f 2MM of the model.
Furthermore, it is known 关e.g., Grace 共1982兲兴 that in simple shear flow when p ⬎ 4 a
steady deformation limit, independent of Ca, exists for large Ca. In that case, the defor-
mation parameter D of the droplet, defined as 共L − B兲 / 共L + B兲, tends toward a constant
value equal to 5 / 共4p兲 and independent of Ca 关Taylor 共1934兲兴. To check if similar behav-
ior can occur in transient mixed flows, the deformation results for several large, but
different Ca-number experiments for p = O共10兲, are compared in Fig. 14. Comparing the

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
1300 BOONEN, VAN PUYVELDE, and MOLDENAERS

FIG. 14. Limiting behavior of droplet deformation for the flow conditions of Fig. 5, a viscosity ratio of p
⬇ 10 and large capillary numbers Ca: 共a兲 Dimensionless droplet axes L / 2R0 and B / 2R0. 共b兲 Droplet deforma-
tion parameter D = 共L − B兲 / 共L + B兲. 共c兲 Average deformation parameter Davg as a function of the number of
cycles.

results for two experiments for subsequent revolutions, the deformation, in terms of L and
B 关see Fig. 14共a兲兴, appears to level off to the same value, independent of Ca, similar to
the limiting behavior seen in steady simple shear flow for p ⬎ 4. In the case of mixed flow
conditions, Fig. 14共b兲 demonstrates the evolution of the deformation parameter D, for the
same two experiments shown in Fig. 14共a兲. Here, two additional experiments have been
included with about the same average capillary numbers as the first two ones, to illustrate
the reproducibility. Inspecting Fig. 14共b兲, the deformation indeed seems to level off to a
steady value. To show this more clearly, the average values of the experimental defor-
mation parameter Davg were calculated 关using 共兰t0D共t兲dt兲 / t兴 for each revolution after
start-up of the flow. In Fig. 14共c兲, Davg is plotted as a function of the number of cycles.
Obviously, apart from the initial start-up transients, the average deformation for all four
experiments appears to go to a constant value, although the observations were only made
for a limited number of cycles. To confirm this steady limiting behavior for high p, the
experiments were repeated for a larger number of revolutions, the results of which are
shown in Fig. 15. In this figure, clear evidence is given for the limiting behavior for high
viscosity ratios 关p = O共10兲兴 in mixed flow conditions. Note that for steady mixed flows
with a constant value of the flow type parameter ¯␣, we can also derive a steady defor-
mation limit for p → ⬁ from the analytical solution of the Maffettone–Minale model 共see
Appendix兲. This is represented by the solid line in Fig. 15 for a steady 2D flow with a
constant value of ¯␣ equal to the average value ¯␣avg present in the first two quadrants of
the flow field. Clearly, the experimental deformation limit is close to the limit for steady
2D flows obtained from the MM model. The existence of a steady deformation limit for
high viscosity ratios in mixed flow conditions was already reported by Bentley and Leal
共1986兲 for steady flows with a constant ¯␣-value. For instance, they found that for ¯␣

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
DROP DYNAMICS IN MIXED FLOWS 1301

FIG. 15. Average droplet deformation for a viscosity ratio of p ⬇ 10 and very high Ca for the flow conditions
of Fig. 5 as function of the number of cycles. The solid line is the deformation limit for p → ⬁ for steady mixed
flows with a constant, positive value of ¯␣ according to the Maffettone–Minale model.

⬇ 0.36, a limiting deformation exists for p-values larger than about 27. In the present
case, however, this steady limiting behavior is already observed for p ⬇ 10. Probably, this
difference is related to the transient nature of the mixed flow applied with the eccentric
cylinder device. Nevertheless, the average deformation observed for high p in the tran-
sient complex flow of the ECD is approximately the same as for a steady mixed flow with
the same 共average兲 value of the flow type parameter ¯␣.

2. Low viscosity ratio drops


In the last part of this study, the effect of a low viscosity ratio on the droplet dynamics
in complex flows was investigated, with p-values of about 0.1 at 23 ° C 共see Table I兲. The
resulting dynamics for these low viscosity ratio drops are shown in Fig. 16for low to
intermediate capillary numbers Ca and in Fig. 17 for high Ca. Again, qualitatively the
same profiles as for p ⬇ 1.2 are obtained in Fig. 16 关compare with Fig. 6兴; the deforma-
tion and orientation increase during the first quadrant of the flow field and subsequently
relax near the end of the second quadrant. As before, good quantitative agreement is
found between experimental results and predictions according to the MM model. Devia-
tions only start to appear for higher Ca, as illustrated in Fig. 17. In this Ca-range we are
near the critical conditions for droplet break-up where the MM model obviously fails. In
fact, for the highest Ca explored here, it was found experimentally that the droplet indeed
breaks up after many revolutions 共outside the time scale of Fig. 17兲, as indicated by the
arrows in the L / 2R0-profile of Fig. 17.
To quantify the effect of viscosity ratio in more detail, the results for p ⬇ 0.1 and p
⬇ 1.2 are compared in Fig. 18. Note the difference in time-scales which is due not only
to differences in revolution period but also to differences in R0 and ␩m for the different
experiments. These differences only become visible 共and pronounced兲 at larger t 共more
revolutions兲. First, Fig. 18共a兲 shows the comparison for intermediate Ca 共Caavg ⬇ 0.2兲;
here, no significant effect of the low viscosity ratio is visible, and the deformation and
orientation are approximately at the same level for both viscosity ratios. In Fig. 18共b兲
results are given for a higher Ca 共Caavg ⬇ 0.4兲, where some differences start to appear.
The low viscosity ratio drops exhibit a somewhat larger deformation 共e.g., visible in the
evolution of the major axis L兲 and slightly less orientation 共e.g., a larger maximum value

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
1302 BOONEN, VAN PUYVELDE, and MOLDENAERS

FIG. 16. Droplet dynamics as a function of dimensionless time tⴱ for the flow conditions of Fig. 5, for a
viscosity ratio of p ⬇ 0.1 and intermediate capillary numbers Ca. The symbols represent the experimental
results, whereas the lines are the predictions according to the Maffettone–Minale model.

FIG. 17. Droplet dynamics as a function of dimensionless time tⴱ for the flow conditions of Fig. 5, for a
viscosity ratio of p ⬇ 0.1 and high capillary numbers Ca. The symbols represent the experimental results,
whereas the lines are the predictions according to the Maffettone–Minale model.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
DROP DYNAMICS IN MIXED FLOWS 1303

FIG. 18. Comparison of droplet dynamics for the flow conditions of Fig. 5, for viscosity ratios of p ⬇ 0.1 and
p ⬇ 1.2, respectively, for 共a兲 intermediate capillary number Ca, 共b兲 high capillary number Ca, and 共c兲 very high
capillary number Ca.

for ␪ at the beginning of first quadrant兲. These observations are also in agreement with
the model predictions according to the MM model 共not shown in Fig. 18兲. Finally, Fig.
18共c兲 compares the data at p ⬇ 0.1 and p ⬇ 1.2 for a large capillary number 共Caavg
⬇ 0.5兲. Clearly, the low viscosity ratio drops are more deformed, especially the major
axis of the droplet L. In addition, they show significantly less orientation toward the flow
direction as evident from the average level of the orientation angle ␪. In summary, the
effect of low viscosity ratio 关p = O共0.1兲兴 in transient mixed flow conditions as applied
with the ECD only becomes apparent for high Ca, leading to increased deformation and
decreasing orientation, as compared to the case for p ⬇ 1.2.

IV. CONCLUSIONS
In this work, we have studied the dynamics of single Newtonian droplets suspended in
an immiscible Newtonian matrix, which is subjected to transient mixed flow conditions.
The controlled complex flows are applied using a home built ECD, as described in
Boonen et al. 共2009兲. By exploring different operation modes for the ECD, and varying
the viscosity ratio p over a range of 2 decades 共from 0.1 to 10兲, the effect of shear/
elongation balance and viscosity ratio has been systematically explored. In addition to the
experimental results on droplet dynamics obtained from optical microscopy, model pre-
dictions have been used for comparison. To this end, the model of Maffettone and Minale
共1998兲 was used, adapted to complex flows by incorporating a flow type parameter that
accounts for the relative amount of elongation in the flow field.
For the different types of sub-critical flow applied and all viscosity ratios explored in
this work, good agreement is found between the experimental results and model predic-
tions. The effect of varying the balance of shear and elongational components is pre-
dicted, as evident from the comparison between two different flows obtained in the ECD,
using two different configurations 共shear dominated versus enhanced extension兲. Next,
the experimentally observed effect of a high 关p = O共10兲兴 and low 关p = O共0.1兲兴 viscosity
ratio, respectively, as compared to the base case of p ⬇ 1.2, is also in agreement with the
model predictions. Quantitative deviations only start to appear for higher capillary num-
bers Ca, for p ⬇ 0.1 and p ⬇ 1.2 because we are near the critical conditions for droplet

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
1304 BOONEN, VAN PUYVELDE, and MOLDENAERS

break-up, where the MM model obviously is not valid anymore. For p ⬇ 10, on the other
hand, a deformation limit exists, independent of the applied Ca. This limiting behavior
can only be incorporated into the MM model by assigning a nonzero value to the em-
pirical factor ␧ in the definition of the coefficient f 2MM in the model.
All these findings provide evidence that the rather simple phenomenological of Maf-
fettone and Minale performs quite well in predicting single droplet dynamics for arbitrary
sub-critical mixed flow fields. Hence, it can be regarded as a useful tool when predicting
the morphology development in more complex systems 共e.g., incorporating elastic and
concentration effects, and “real” processing flows兲. The extensive data set provided here
can serve as a reference to guide and evaluate future modeling efforts.

ACKNOWLEDGMENTS
The authors would like to thank Ir. Bart Caerts for his help with the design and
construction of the ECD. This work has been financially supported by the Onderzoeks-
fonds KULeuven 共Grant Nos. GOA 03/06 and GOA 09/002兲.

APPENDIX
The non-dimensional form of the MM model 关Maffettone and Minale 共1998兲兴 is given
by

dSⴱ
− Ca共Wⴱ · Sⴱ − Sⴱ · Wⴱ兲 = − f MM ⴱ
1 共p兲共S − g
MM ⴱ
共S 兲I兲 + Caf MM ⴱ
2 共p,Ca兲共D · S

dtⴱ
+ Sⴱ · Dⴱ兲, 共A1兲

where Sⴱ = S / R02, tⴱ = t / ␶, and D and W have been made dimensionless with the flow
intensity E:

冢 冣
1 − 兩¯␣兩
¯␣ 0
2
1
Dⴱ = 1 − 兩¯␣兩 ,
冑4¯␣2 + 共1 − 兩¯␣兩兲2 − ¯␣ 0
2
0 0 0

冢 冣
1 − 兩¯␣兩
0 0
2
1
Wⴱ = − 1 + 兩¯␣兩 . 共A2兲
冑4¯␣2 + 共1 − 兩¯␣兩兲2 0 0
2
0 0 0

The steady state values for the dimensionless axes L / 2R0 and B / 2R0 can then be calcu-
lated from Eqs. 共A1兲 and 共A2兲 to be

冉 冊L
2R0
2
=
f 21 + Ca2共1 − 兩¯␣兩兲2 + Caf 2冑关f 21 + Ca2共1 − 兩¯␣兩兲2兴关共1 − 兩¯␣兩兲2 + 4¯␣2兴
关f 21 + Ca2共1 − 兩¯␣兩兲2兴1/3关共共1 − f 22兲共1 − 兩¯␣兩兲2 − 4¯␣2 f 22兲Ca2 + f 21兴2/3
,

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
DROP DYNAMICS IN MIXED FLOWS 1305

冉 冊B
2R0
2
=
f 21 + Ca2共1 − 兩¯␣兩兲2 − Caf 2冑关f 21 + Ca2共1 − 兩¯␣兩兲2兴关共1 − 兩¯␣兩兲2 + 4¯␣2兴
关f 21 + Ca2共1 − 兩¯␣兩兲2兴1/3关共共1 − f 22兲共1 − 兩¯␣兩兲2 − 4¯␣2 f 22兲Ca2 + f 21兴2/3
. 共A3兲

Thus, the deformation parameter D = 共L − B兲 / 共L + B兲 equals

冑 f 21 + Ca2共1 − 兩¯␣兩兲2 − 冑 f 21 + Ca2共1 − 兩¯␣兩兲2 − Ca2 f 22关共1 − 兩¯␣兩兲2 + 4¯␣2兴


D= . 共A4兲
Caf 2冑共1 − 兩¯␣兩兲2 + 4¯␣2
Taking the limit for 1 / p → 0 where f 1 goes to 20/ 共19p兲 and f 2 to 5 / 共2p兲 关see Eqs. 共2兲 for
␧, ␦ ⬎ 0兴, finally we obtain

5 冑共1 − 兩¯␣兩兲2 + 4␣2


Dlim = . 共A5兲
4p 共1 − 兩¯␣兩兲
Hence, Dlim represents the steady deformation limit for steady 2D mixed flows with a
constant ¯␣. For ¯␣ = 0 Dlim equals 5 / 共4p兲, the Taylor limit for simple shear flow 关Taylor
共1934兲兴; for ¯␣ = 1, on the other hand, no steady limit exists as a pure extensional flow can
break-up a drop of any p.

References
Ballal, B. Y., and R. S. Rivlin, “Flow of a Newtonian fluid between eccentric rotating cylinders—Inertial
effects,” Arch. Ration. Mech. Anal. 62, 237–294 共1976兲.
Bentley, B. L., and L. G. Leal, “An experimental investigation of drop deformation and break-up in steady
two-dimensional linear flows,” J. Fluid Mech. 167, 241–283 共1986兲.
Boonen, E., P. Van Puyvelde, and P. Moldenaers, “Droplet dynamics in sub-critical complex flows,” Rheol. Acta
48, 359–371 共2009兲.
Cristini, V., R. W. Hooper, C. W. Macosko, M. Simeone, and S. Guido, “A numerical and experimental
investigation of lamellar blends morphologies,” Ind. Eng. Chem. Res. 41, 6305–6311 共2002兲.
Diprima, R. C., and J. T. Stuart, “Flow between eccentric rotating cylinders,” ASME J. Lubr. Technol. 94,
266–274 共1972兲.
Egholm, R. D., P. Fischer, K. Feigl, E. J. Windhab, R. Kipka, and P. Szabo, “Experimental and numerical
analysis of droplet deformation in a complex flow generated by a rotor-stator device,” Chem. Eng. Sci. 63,
3526–3536 共2008兲.
Feigl, K., S. F. M. Kaufmann, P. Fischer, and E. J. Windhab, “A numerical procedure for calculating droplet
deformation in dispersing flows and experimental verification,” Chem. Eng. Sci. 58, 2351–2363 共2003兲.
Godbille, F. D., and J. J. C. Picot, “Drop break-up in combined shear and elongational flow conditions,” Adv.
Polym. Technol. 19, 14–21 共2000兲.
Grace, H. P., “Dispersion phenomena in high-viscosity immiscible fluid systems and application of static mixers
as dispersion devices in such systems,” Chem. Eng. Commun. 14, 225–277 共1982兲.
Guido S., and Greco F., “Dynamics of a liquid drop in a flowing immiscible liquid,” Rheology Reviews 2004,
The British Society of Rheology 共2004兲, pp. 99–142.
Guido, S., M. Simeone, and M. Villone, “Diffusion effects on the interfacial tension of immiscible polymer
blends,” Rheol. Acta 38, 287–296 共1999兲.
Kaufmann, S. F. M., P. Fischer, and E. J. Windhab, “Investigation of droplet dispersing processes in shear and
elongational flow,” Second International Symposium on Food Rheology and Structure, Laboratory of Food
Process Engineering, Zurich, Switzerland 共2000兲, edited by P. Fischer, I. Marti, and E. J. Windhab, pp.
404–405.
Khayat, R. E., A. Luciani, L. A. Utracki, F. D. Godbille, and J. J. C. Picot, “Influence of shear and elongation
on drop deformation in convergent-divergent flows,” Int. J. Multiphase Flow 26, 17–44 共2000兲.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56
1306 BOONEN, VAN PUYVELDE, and MOLDENAERS

Maffettone, P. L., and M. Minale, “Equation of change for ellipsoidal drops in viscous flow,” J. Non-Newtonian
Fluid Mech. 78, 227–241 共1998兲.
Maffettone, P. L., and M. Minale, “Equation of change for ellipsoidal drops in viscous flows 共vol 78, pg 227,
1998兲,” J. Non-Newtonian Fluid Mech. 84, 105–106 共1999兲.
Minale, M., “Deformation of a non-Newtonian ellipsoidal drop in a non-Newtonian matrix: Extension of the
Maffetone-Minale model,” J. Non-Newtonian Fluid Mech. 123, 151–160 共2004兲.
Ottino, J. M., P. De Roussel, S. Hansen, and D. V. Khakhar, “Mixing and dispersion of viscous liquids and
powdered solids,” Adv. Chem. Eng. 25, 105–204 共1999兲.
Rasband, W. S., IMAGEJ, U. S. National Institutes of Health, Bethesda, MD 共1997兲.
Rumscheidt, F. D., and S. G. Mason, “Particle motions in sheared suspensions. XII. Deformation and burst of
fluid drops in shear and hyperbolic flow,” J. Colloid Sci. 16, 238–261 共1961兲.
Stone, H. A., “Dynamics of drop deformation in viscous liquids,” Annu. Rev. Fluid Mech. 26, 65–102 共1994兲.
Stone, H. A., B. L. Bentley, and L. G. Leal, “An experimental study of transient effects in the break-up of
viscous drops,” J. Fluid Mech. 173, 131–158 共1986兲.
Stone, H. A., and L. G. Leal, “The influence of initial deformation on drop break-up in sub-critical time-
dependent flows at low Reynolds numbers,” J. Fluid Mech. 206, 223–263 共1989兲.
Taylor, G. I., “The viscosity of a fluid containing small drops of another fluid,” Proc. R. Soc. London, Ser. A
138, 41–48 共1932兲.
Taylor, G. I., “The formation of emulsions in definable fields of flow,” Proc. R. Soc. London, Ser. A 146,
501–523 共1934兲.
Torza, S., R. G. Cox, and S. G. Mason, “Particle motions in sheared suspensions. XXVII. Transient and steady
deformation and burst of liquid drops,” J. Colloid Interface Sci. 38, 395–411 共1972兲.
Tucker, C. L., and P. Moldenaers, “Microstructural evolution in polymer blends,” Annu. Rev. Fluid Mech. 34,
177–210 共2002兲.
Van Puyvelde, P., A. Vananroye, R. Cardinaels, and P. Moldenaers, “Review on morphology development of
immiscible blends in confined shear flow,” Polymer 49, 5363–5372 共2008兲.
Vananroye, A., P. Van Puyvelde, and P. Moldenaers, “Structure development in confined polymer blends:
Steady-state shear flow and relaxation,” Langmuir 22, 2273–2280 共2006兲.
Vinckier, I., P. Moldenaers, and J. Mewis, “Relationship between rheology and morphology of model blends in
steady shear flow,” J. Rheol. 40, 613–631 共1996兲.
Wannier, G. H., “A contribution to the hydrodynamics of lubrication,” Q. Appl. Math. 8, 1–32 共1950兲.
Windhab, E. J., M. Dressler, K. Feigl, P. Fischer, and D. Megias-Alguacil, “Emulsion processing—From
single-drop deformation to design of complex processes and products,” Chem. Eng. Sci. 60, 2101–2113
共2005兲.

Redistribution subject to SOR license or copyright; see http://scitation.aip.org/content/sor/journal/jor2/info/about. Downloaded to IP:


69.26.46.21 On: Sun, 11 May 2014 15:18:56

You might also like