You are on page 1of 13

Chemical Engineering Science 245 (2021) 116808

Contents lists available at ScienceDirect

Chemical Engineering Science


journal homepage: www.elsevier.com/locate/ces

Dynamics of drop formation, growth and pinching phenomena from


a submerged nozzle
Abhik Majumder a, Dibyendu Ghosh b, Prasanta Kumar Das b,⇑
a
National Institute of Technology Agartala, Department of Mechanical Engineering, Agartala, Jirania 799046, India
b
Indian Institute of Technology Kharagpur, Department of Mechanical Engineering, Kharagpur 721302, India

h i g h l i g h t s

 Drop formation from submerged nozzles in an immiscible liquid is reported.


 Transition from dripping to jetting, shape oscillation, neck radius fluctuation investigated.
 Dependents of pinching on velocity and fluid properties investigated.
 Different regimes of drop formation identified.

a r t i c l e i n f o a b s t r a c t

Article history: Drop formation is studied through a comprehensive experimentation, injecting different lighter fluids in
Received 23 December 2020 an immiscible liquid, over a range of velocities, from submerged nozzles of four different diameters. The
Received in revised form 22 March 2021 process involves a number of unique interfacial developments, namely inception and growth, neck for-
Accepted 26 May 2021
mation and pinching, drop departure, etc. The study emphasizes on the dynamical aspects like shape
Available online 29 May 2021
oscillation on the nozzle tip, neck radius fluctuation, length variation of the leading edge (Lv ) etc. At a
low flow rate, the retractile motion of the remnant mass is found responsible for shape alteration. The
Keywords:
process of drop pinching exhibits radial fluctuation of the neck and depends on both flow rate and vis-
Drop generation
Immiscible liquid
cosity. Moreover, in an unconfined system, drops may be generated with increased frequency within
Necking and pinching the limit of the dripping regime while keeping Lv approximately unchanged. Different regimes of drop
Dripping formation have also been identified.
Jetting Ó 2021 Elsevier Ltd. All rights reserved.

1. Introduction injected liquid mass may take a shape of a liquid jet of finite length
at the nozzle outlet. Eventually, drops are produced from the tip of
The formation of liquid drops by injecting one liquid into the jet. The drop formation in the dripping and the jetting regimes
another immiscible liquid through an orifice or a nozzle is widely are illustrated in Fig. 1. The phenomenon of drop formation at the
used in many industries. By regulating the operating parameters, nozzle end is termed as dripping irrespective of upward or down-
drops of the desired volume can be formed at a required frequency ward movement of the drops and the formation of a jet at the noz-
by the breakup of the injected liquid mass. For convenience, let us zle tip is commonly described as jetting (Homma et al., 2006). As
term the drop-forming liquid as the dispersed phase and the other the dispersion of one liquid into the other, in the form of droplets,
as the bulk liquid phase. If the dispersed phase is lighter than the provides a large interfacial area, these basic arrangements find
bulk phase, the opening of the nozzle or the orifice is made to face wide applications in a variety of processes and equipment like a
upward, then the dispersed phase liquid drops move up. When the liquid–liquid contractor, chemical reactors, distillation, pathologi-
dispersed phase is heavier, the nozzle is made to face down and cal investigations, extraction and emulsification etc. (Chaurasia
drops fall through the bulk liquid after formation. Further, irre- et al., 2015; Hayworth and Treybal, 1950; Wang et al., 2009;
spective of the upward or downward facing orientation of the noz- Zhang, 1999). In recent years, various unique applications of drop
zles, the drop may form immediately at the nozzle outlet or the formation in a host of an immiscible liquid have been proposed
for microfluidic devices and different flow focusing techniques
⇑ Corresponding author. have also been postulated (Cramer et al., 2012; Nie et al., 2008;
E-mail address: pkd@iitkgp.ac.in (P. Kumar Das). Utada et al., 2007).

https://doi.org/10.1016/j.ces.2021.116808
0009-2509/Ó 2021 Elsevier Ltd. All rights reserved.
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

Nomenclature

As Drop Surface area, mm2 t Time, s


a Major axis of a drop, mm tp width of the elemental disk, mm
b Minor axis of a drop, mm Un Flow velocity through nozzle, m/s
a=b Aspect ratio Vd Final volume of drop, mm3
Di Inner diameter of nozzle, mm Vavg Average volume of drop, mm3
DF Diameter of the detached drop, mm Vp Primary volume of a drop, mm3
di the diameter of the ith elemental disk, mm
dw Wetting diameter, mm Greek symbol
g Gravitational acceleration, m/s2 q Density, kg/m3
ID Inner diameter, mm qd Density of dispersed phase, kg/m3
Lbf Break off length, mm ri Interfacial tension, N/m
Lv Leading edge length of a drop, mm l Viscosity, mPas
Oh Ohnesorge number lr Ratio of viscosity,
Ql Volume flow rate, ml/min h Instantaneous Contact angle, degree
Re Reynolds number Dq Density difference, kg/m3
Rn Neck radius, mm kc Capillary length, mm
R0 Nozzle outer radius, mm s Non dimensional time

The majority of the literature on drop formation can be catego- the effect of viscosity ratio on the mechanism of pinching, mecha-
rized into two groups, viz.: unconfined drop formation in quiescent nism of transition from dripping to jetting (Chaurasia et al. 2015),
or flowing liquid and confined drop formation in microfluidic chan- apparent contact angle behavior (Wang et al. 2009), the effect of
nel configurations (Utada et al. 2007; Shah et al. 2008; Meyer and dispersed liquid property and nozzle diameter on the transition
Crocker 2009; Cramer et al. 2012; Josephides and Sajjadi 2015; from dripping to jetting etc.
Bertrandias et al. 2017). The unconfined arrangement of drop for- Over the years, a variety of research investigations have been
mation is employed over the ages due to its simplicity, carried out on the formation of liquid drops in the host of an
buoyancy-driven upward or downward motion of drop based on immiscible liquid through a submerged nozzle (Harkin and
specific gravity (Hayworth and Treybal, 1950). Drops of macro, Brown, 1919; Smith and Moss, 1917; Hayworth and Treybal,
micro, and nanometric dimensions from such systems may also 1950; Null and Johnson, 1958). The prediction of drop shape and
find application in process equipment, artificial cell, bio- estimation of drop size has been the prime attention of such work
encapsulation, drug delivery, and membrane separation (Chang due to the importance of the interfacial area in the transport mech-
2004; Bremond et al. 2010; Chaurasia et al. 2015). Though drop anism of different industrial processes. The review work of Bogy,
formation in liquid–liquid systems enjoys diverse applications, (1979), Lin and Reitz, (1998) and Eggers and Villermaux, (2008)
the phenomenon is still left with many unresolved issues such as provides good detail of numerous investigations carried out on
drop formation in liquid, air, or gas phase. In the past, in addition
to experimental investigation for drop size estimation under varied
external fluid properties, several attempts have also been made to
develop correlation for drop volume (Rao et al.1966; Scheele and
Meister, 1968; Heertjes et al. 1970, Izard 1972) as a function of
fluid properties.
On the other hand, a good number of numerical simulation
works were also found to investigate the drop formation from a
submerged nozzle. For example, Homma et al. (2006)employing
the front tracking method computationally explored the immisci-
ble liquid jet formation for axisymmetric condition and predicted
a regime map of dripping to jetting transition while varying the
Weber number and the viscosity ratio of the phases. Lakdawala
et al. (2014, 2015) numerically investigated the formation of a drop
in the immiscible system and showed periodic and non-periodic
formation patterns both in dripping and jetting regime. More
recently, using a coupled level-set and VOF method, Borthakur
et al. (2017) showed a drop interface overturning at a critical
Ohnesorge number.
In a unique experimental study, Wang et al. (2009) studied the
change in current contact angle with time until the drop detached
from a nozzle of outer diameter 0.508 mm using the organic liquid
in stagnant water. They found that the slope of the current contact
angle coincides with the drop formation stages except for the
detachment stage. However, to date, no experimental data
reported on the effect of nozzle radius, dispersed phase viscosity,
and flow velocity on neck radius, pinching process, etc. More
Fig. 1. Dripping and jetting of immiscible dispersed fluid in buoyancy assisted recently, considering the effect of dynamic interfacial tension on
unconfined system. drop volume for a nozzle inner diameter of 0.23 mm, Chaurasia
2
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

et al. (2015) investigated the effect of injection velocity of dis- and viscosity of the injected fluid. Further, a radial fluctuation of
persed phase on polydispersity and pinching time for three differ- the neck was also noted for fluids of high Ohnumber. A wide range
ent alkanes. However, the effect of nozzle diameter and dispersed of data for four different dispersed liquids injected through nozzles
phase viscosity on transition remains mostly unexplored. The of four different diameters is presented.
necking and subsequent motion of a separated drop of aqueous
glycerol in silicon and mineral oil were investigated by Patrascu
et al. (2017). They found that neck filament rupture time increases 2. Experimental facility and procedure
with the viscosity of the dispersed liquid and detached drop
rapidly attains constant velocity. The present investigation on drop formation is primarily
The majority of the semi-empirical or correlation-based experimental. A schematic of the experimental facility is presented
approaches for the prediction of drop characteristics use either in Fig. 2. The main constituent is an acrylic (poly-acrylic-
force balance type mechanistic and oversimplified models while methacrylate) rectangular container of dimension 100 mm 
ignoring the effect of the contact line, wetting phenomena, an 100 mm  300 mm with its top open to the atmosphere. An over-
approximate model for drag force, etc. On the other hand, the flow hole of 4 mm diameter is placed in one of the container faces,
numerical works are also handicapped from the limitations of suf- thus a constant height of 250 mm liquid is maintained. At the center
ficient computational resources, absence of precise contact angle of the bottom surface of the container, a steel nozzle is fitted verti-
model, wetting model, limitations of axisymmetric consideration, cally on top of an adjustable mount. A steel tube of 2 mm diameter
etc. In view of the above, the role of experimental investigation is connected to the other end of the mount. Through this steel tube
is beyond doubt for such a complex phenomenon. lighter liquid (dispersed fluid) is supplied at a constant flow rate
Most of the earlier investigations attempted to predict the drop from a Harvard make infusion pump (accuracy 0.35%). A Phantom
volume, Sauter diameter, jet length, etc. under different flow con- make V1211 high-speed camera with Nikon 60 mm micro zoom
ditions. However, a very limited number of studies considered the lens (maximum resolution of 1280  800 pixel) is used to record
effect of dispersed liquid viscosity, interfacial tension and nozzle the evolution of the drop at the nozzle tip at frame rates up to
diameter on resulting drop dynamics and mechanism of neck for- 8000 frames per second. The number of the representative pixel
mation and neck pinching. The neck pinching process is a very corresponding to the nozzle outer diameter is counted to obtain
important stage of drop separation and significantly influences the scale of physical dimension from the captured images. The scale
the drop dynamics. Moreover, the effect of properties of the dis- of such physical dimension is shown in the respective images.
persed phase and nozzle diameter on drop characteristics such as Before each of the experimental runs, the acrylic container is
drop size, drop leading edge length are not well understood. In thoroughly cleaned and is filled with de-ionized water up to a
the present work a comprehensive study is undertaken to under- height of 250 mm. The small air bubbles attached to the con-
stand the effect of (1) nozzle diameter, (2) dispersed phase viscos- tainer wall are carefully removed to establish a clear path for
ity and (3) interfacial tension on (1) drop volume, (2) drop leading recording the images. The overflow opening allows the drainage
edge length at detachment and (3) neck radius and pinching pro- of the accumulated lighter fluid on top of the heavier fluid during
cess. In the present investigation dynamical aspects of drop forma- the experiment and maintains a constant level of the heavier fluid
tion such as shape oscillation on nozzle tip, neck radius fluctuation, on top of the nozzle tip throughout the experimental run and for
drop volume, and leading-edge length of a drop are revisited with all the experiments, ensuring uniform and constant hydrostatic
the parametric variation of nozzle diameter and viscosity ratio. The head at the open end of the nozzle. The lighter fluid is carefully
process of drop pinching is found to depend on both the flow rate filled in a new syringe and the selected flow rate is maintained

Fig. 2. Schematic of the experimental facility.

3
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

Table 1
Properties and range of flow rate of the fluids used.

Type of Phase Fluid Viscosity (mPas) Density (kg/m3) Interfacial Tension (N/m) Refractive Index Flow rate (ml/min)
Dispersed kerosene 1.640 810.00 0.0324* 1.440 0.525–14.70
Dispersed cyclohexane 0.860 778.10 0.0480* 1.426 0.525–16.80
Dispersed n-heptane 0.336 679.50 0.0491* 1.385 0.525–12.39
Dispersed n-hexane 0.300 645.80 0.0494* 1.375 0.525–12.39
Bulk Deionized water 0.890 998.00 0.0729** 1.330 Stationary
*
Interfacial Tension of the dispersed liquid with water.
**
Surface tension of water with air.

by using the infusion pump. The length of the nozzle is kept 25 images. Care is taken for proper illumination and camera position.
times the inner diameter of the nozzle. This acts as a flow All the fluids used in the experiments are almost colorless and
straightener and renders the flow axial at the nozzle exit. Four transparent enabling the capture of images only due to the relative
nozzle of inner diameter (ID) 1.0, 0.6, 0.4 and 0.3 mm respec- difference of refractive index between the host and dispersed flu-
tively, are selected to study the effect of diameter. Also, four dif- ids. No die/additive is added to any fluid to avoid contamination
ferent lighter fluids are carefully selected for the present work (at and deviation in fluid properties.
a temperature of 25 °C of all the fluids used in the investigation Images are recorded two minutes after the appearance of the
are given in Table 1). These fluids were chosen for their water first drop to allow the internal flow field of the heavier fluid to
immiscibility as well as a variation of properties such as viscosity, develop and settle in response to the injected lighter fluid motion.
density, and interfacial tension etc. The experiments are con- This is also done to confirm that the bulk liquid is pre-saturated
ducted at room temperature, which was close to 25 °C and no and the top surface of the heavier fluid in the container obtains a
appreciable variation in properties was present. The immiscibility thin layer of lighter fluid. This ensures the drainage of lighter fluid
with the bulk fluid and density lower than it (water in this case) becomes continuous, thereby endorse minimum disturbance in the
were two important criteria for the selection of the dispersed operating conditions. The repeatability of the experimental results
fluid. It may further be noted that, only readily available natural has been checked carefully. At a low flow rate (Re ¼ 27 for all the
fluids have been considered. Among the available options a selec- fluids), the experiments were repeated at least four times. At other
tion has been made so that some variation is there in the trans- conditions, random check of repeatability have been made from
port properties of the dispersed fluids. time to time. Careful observations are made to understand drop
Before switching over from one dispersed fluid to the other, the formation and evolution, pinching, post-pinch-off development of
steel tube and nozzle are thoroughly cleaned and a fresh syringe is remnant mass, detachment, etc. through the processing of the
used. The high-speed video camera is placed at the desired location recorded images. The images are recorded at a speed range of
to obtain the best quality images through optimally choosing the 2000 to 8000 frames per second. The captured images are analyzed
focal length of the lens. Sufficient backlight is provided by an using P.C.C., MATLAB, and the software Image-Pro Plus. The details
LED array placed behind a diffusion screen to capture the drop of image processing are given in the subsequent section.

Fig. 3. Steps of image processing (a) to (e). (f) Unit disk representation of a typical fictitious drop. (g) Schematic of elemental frustum .

4
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

The original image is first converted to a greyscale image fol- separated droplets are more or less spherical and waves are not
lowed by cropping operation to remove the nozzle from the image. seen in the stem or the column of the lighter liquid emanating from
Initial wiener filter is applied on the greyscale image to reduce the nozzle end. In snapshot 6 the formation of the wave in the liq-
noise. A threshold method with a user-defined fudge-factor value uid jet is discernable and the pinching point corresponds to a node
of 0.8 is applied to identify the drop borderline. To eliminate the of the wave. Further, the separated droplet is rather elongated indi-
unwanted spots from the image a secondary noise reduction filter cating an early pinch-off at a particular node before the liquid mass
is also applied. The resultant image is then converted into a binary above the node has enough time to assume a spherical shape. The
image as shown in Fig. 3 (a) to (e). The final binary image is pro- presence of the wave becomes more prominent in the succeeding
cessed to obtain the coordinates of elemental frustum as shown snapshots, and interestingly, the separated mass of the lighter liq-
in Fig. 3 (f) and (g) to calculated the drop surface area and volume uid depicts a tadpole-like shape with an oval head and a narrow
using following equations (Sabliov et al., 2002). thread-like tail (snapshot 8 and 9). Nodes and anti-nodes are also
vffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
)ffi visible in the elongated structure severed from the liquid jet. The
u( 2
X
n
p u di diþ1 severed liquid structures are seen to generate a number of drops
As ¼ ðdiþ1 þ di Þ t  þ tp 2
ð1Þ
2 2 2 depending upon their length. Further, the separated drops may
i¼1
be seen colliding with one another downstream of the jet tip where
the succeeding drops are forming. Thus the process of collision
X
n
pt p  2 2

may affect the formation of succeeding drops (snapshot 9). With
Vd ¼ diþ1 þ di þ diþ1 di ð2Þ
i¼1
12 the increase of flow rate, the jet becomes sinuous, particularly at
the upper end, as can be seen in snapshot 11.

3. Result and discussion


3.2. Evolution of remnant mass into a new drop
3.1. Overview of submerged drop formation
The detachment of a drop from the nozzle is an important stage
To provide an overall idea of the evolution of the dripping pro- of drop formation. The evolution of the prevailing interface drasti-
cess through submerged nozzles, we have selected some typical cally varies with the change in flow rates. The remnant of the
experimental runs involving the flow of cyclohexane as the dis- pinched drop attached to the nozzle tip serves as the nucleus for
persed fluid through a nozzle ID of 1.0 mm. Initially, at a very the succeeding drop. The separation of a drop at the near nozzle
low flow rate, nearly monodispersed drops are produced as can and far nozzle is discussed in the preceding sections. After a suc-
be observed in snapshots 1–3 of Fig. 4. Furthermore, the drops cessful drop separation, interfacial force driven recoiling of a liquid
are also formed at very regular intervals. The flow rate is gradually interface attached to the nozzle tip is observed at a low flow rate.
increased, and eventually, the lighter liquid issues in the form of a Immediately after a successful pinching, the remnant liquid mass
jet and a liquid column of constant height appear at the nozzle exit attached to the nozzle tip undergoes retracting motion due to
with a shift of the drop formation process from the nozzle tip to the the interfacial force acting to minimize the elongated conical inter-
tip of the jet. The uniformity in drop diameter is ultimately lost at facial area against the inertial and viscous force of the dispersed
higher flow rates. liquid. In due course, these motions establish a greater degree of
In general, as the flow rate increases, drop generation frequency interface sphericity through a transition of interfacial shape
increases along with a decrease in drop size. The break-off length (Zhang, 1999). The pre-pinching flow momentum can significantly
of the liquid jet, as marked with an arrow in snapshot 6, increases affect the recoiling shape (Sierou and Lister, 2004). At low and
gradually (snapshot 1–5) with the increase in flow rate. Interest- intermediate flow rates, it is noticed that this retracting recoil
ingly, a steep increase in break-off length, as noticed in snapshot velocity is sufficient enough to overcome the combined instanta-
5, reveals a probable onset of transition from dripping to the jetting neous buoyancy and internal flow inertia resistance. The present
regime. The onset of the transition state is identified by quantifying investigation also reveals that such retracting motion can intro-
this break-off length, which is found approximately three times the duce oscillations and readjustment of shape (such as prolate to
nozzle’s outer diameter (Chaurasia et al. 2015). Till snapshot 3, oblate) as shown in Fig. 5.

Fig. 4. Cyclohexene drop formation in stagnant water at various flow rates from a nozzle of the inner diameter of 0.99 mm. Starting from snap 1–11, corresponding flow rates
are (1) 3.07 ml/min; (2) 8.200 ml/min; (3) 10.250 ml/min; (4) 12.300 ml/min; (5) 14.350 ml/min; (6) 16.400 ml/min; (7) 18.450 ml/min; (8) 22.550 ml/min; (9) 26.65 ml/min;
(10) 30.750 ml/min; and (11) 34.850 ml/min respectively. The bar represents the outer diameter of the nozzle.

5
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

increases, we observe a relatively lesser variation in aspect ratio.


This may be primarily due to the effect of gravitational force (large
liquid mass attached to bigger nozzle tip) in addition to the viscous
and inertial resistance responsible for the rapid dwindling of the
retracting momentum. At higher flow rates, the drop forms by sep-
aration from a liquid jet far away from the nozzle tip. The forma-
tion of node and anti-nodes are clearly visible on the liquid
column. Post-separation of a drop, the retracting motion induced
by interfacial force at this stage causes backward motion of the
interface which may assist in the formation of neck as illustrated
in Fig. 6 with a series of snapshots taken for cyclohexane at a flow
rate of 8.25 ml/min. The phenomena are also observed for other
dispersed fluids.

3.3. Process of pinching of a drop

The process of pinching is the last stage of drop formation. At


this stage, the drop finally separates from the nozzle tip. At the
dripping regime, the quantitative estimation of the duration of
drop formation not only helps to understand the pinching process
but also brings out the effect of viscosity ratio (lr ) on the pinching.
Fig. 7 shows a comparison of cyclohexane, kerosene and n-hexane
drop formation at an identical flow rate of 3.07 ml/min through a
nozzle ID of 1.0 mm. The transport properties of the dispersed liq-
uids are not identical. Since the flow rate is the same for all the flu-
Fig. 5. Shape oscillation of a the remnant mass after a successful pinch-off; ids presented in series a, b and c in Fig. 7; the comparison depicts
dispersed fluid cyclohexane, flow rate of 1.06 ml/min, nozzle ID of 0.3 mm. The first the fluid properties on drop generation. Wang et al. (2009) showed
shape reversal may be noted at 1.5 ms (approx.).
the dynamics of wetting radius for a first drop formation through a
nozzle and reported that the dispersed fluid interface may be
pinned at the nozzle outer radius for subsequent drop formation.
Fig. 5 illustrates the shape oscillation of remnant mass for a typ- For the present analysis, the neck thinning and pinching of a drop
ical cyclohexane drop formed through nozzle ID of 0.3 mm; the are reported where the drop interface is pinned at the outer radius
aspect ratio a=b decreases initially due to the retracting motion. of the nozzle. The photographic evidence from Fig. 7 shows that
At 1.5 ms, the initial shape alteration of interface configuration is the nozzle is completely wetted by all the fluids suggesting an
due to retraction, as shown in Fig. 5. Thereafter, the aspect ratio equality of wetting radius. Therefore, considering the same wetting
reaches 1.29 at 3.62 ms, again only to assume a value approxi- radius, the variation in pinching time can be explained by compar-
mately unity at 6.0 ms. Here, aand b are the longitudinal and lateral ing thelr (viscosity ratio). In view of that, the drop evolution time is
dimensions of the remnant liquid mass respectively. During the divided into two intervals; namely, pre necking time or drop-
period of shape oscillation, the three-phase contact line may also growth time (primary growth time) and post necking time or
undergo advancing and retracting motion and may modify the pinching time.
wetting perimeter. Subsequently, the oscillation of the drop is seen Snapshot series d1 to d6 shows the evolution of neck radius of a
to dampen with elapsed time. However, as the nozzle diameter kerosene drop for a flow rate of 11.13 ml/min. An Oscillatory

Fig. 6. The formation of a cyclohexane drop from jet column and retracting motion of the interface induced by interfacial force. The formation of neck at a flow rate of
8.25 ml/min through a nozzle of ID 0.3 mm. The bar represents 0.56 mm.

6
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

Fig. 7. Drop formation of cyclohexane (a1-a5), kerosene (b1-b5) and n-hexane (c1-c5) at a flow rate of 3.07 ml/min. Snapshots a3, b3, and c3 depict a typical drop
configuration (h  p=2). Snapshots 5 for all the three fluids depict pinch off, nozzle inner diameter 1.0 mm.

evolution of neck radius is shown in d4 and d5. The bar represents kerosene drop for a flowrate of 11.13 ml/min. After an initial thin-
1.264 mm. ning attempt at time 0.278 sec, the neck radius reaches to
Here, necking time signifies the time elapsed starting from the 0.486 mm. However, with the additional support of viscous resis-
inception of a drop to reach the state of neck formation, identified tance of the dispersed liquid, the relatively high momentum flux
by h  p=2(see a3, b3, and c3 in Fig. 7). Whereas, the pinching time intends to resist the breakup by acting against the strong capillary
indicates the separation time of a drop from the state of h  p=2. force driving the breakup process (Patrascu et al., 2017). Such
The evaluation of photographic data shows that even though the resistance may even lead to expansion of the neck as shown in
dispersed fluid flow rate and nozzle diameter are the same for all Fig. 7 (d5) at 0.291 sec. Finally, the drop pinch-off takes place at
the liquids, the estimated drop generation frequencies differ 0.306 sec.
(0.729, 1.081 and 1.159 for cyclohexane, kerosene and n-hexane
respectively). For instance, for a flow rate of 3.07 ml/min, for a typ- 3.4. Effect of nozzle diameter and dispersed fluid property on the drop
ical cyclohexane drop, as shown in Fig. 7 (series a1-a3), the time of volume
neck formation is 1.371 sec. For the same flow rate, this value is
much larger than those recorded for kerosene and n-hexane. On For efficient transport of species and heat transfer enhance-
the other hand, for the same flow rate of 3.07 ml/min, the pinching ment, surface area and volume of a drop are considered the most
time (0.091 sec) for kerosene is found higher than that of cyclohex- important influencing parameters. Therefore, the study of these
ane (0.073 sec) and n-hexane (0.064 sec). The high viscosity of the significant drop characteristics is necessary. Fig. 9 (a) depicts the
dispersed liquid resists the gravitational acceleration (Longmire variation of average volume of cyclohexane drop and standard
et al., 2001) causing the thinning of the drop stem during a final deviation of drop sizes against the injection velocity in nozzle ID
pinch; the pinching time is found large in case of higherlr fluids. of 1.0 mm. In general, the average drop volumes are noted to
Initially, for the same flow rate, the pinch-off of a drop from the increase at the beginning up to injection velocity of 0.092 m/s;
nozzle tip through the process of thin liquid thread formation is beyond this velocity, a decrease in average drop volume is
found similar for all the fluids. The length of the liquid thread is observed. As the injection velocity of the dispersed liquid
small and the neck radius is seen to decrease monotonically as increases, fluctuations are noted in the process of drop formation.
shown in Fig. 7 for a flow rate of 3.07 ml/min. Here, the computa- As a result, at a particular injection velocity, a range of drop sizes
tion of pinch-off time begins from the position (h  p=2) as shown (around a mean drop) are noted. It is thought that the standard
in a3, b3, c3 and d3 of Fig. 7. The pinch-off time is found dependent deviation of the drop volume could be a good indicator of the
on dispersed phase viscosity and increases with an increase in vis- hydrodynamics o the drop formation. The standard deviation of
cosity. One can easily note the largest pinching time (0.113 sec) of drop volumes is found insignificant for low injection velocities;
kerosene (highestlr ) at a flow rate of 1.06 ml/min compared to however, it increases considerably as the velocity increases. Based
that of the other fluids as shown in Fig. 8. In this range, as the iner- on the values of standard deviation (SD), the process of drop gen-
tia force is less, the buoyancy force pulls the liquid interface eration can be divided into three regimes viz. regime 1 (0 < SD < 0.
against the interfacial tension. However, close to the end of the 2), regime 2 (0.2 < SD < 1.5) and regime 3 (1.5 < SD < 6.0).
dripping regime (flow rates 10.6–12.72 ml/min), such pinch-off In regime 1, the drop formation is essentially dripping in nature.
process is found to occur from longer liquid threads and is sub- The drop formation process is periodic, symmetric, and mono-
jected to an oscillatory radial reduction of the neck as shown in dispersed. In regime 2, with intermediate standard deviation,
Fig. 8 (a) and (b). Also, the pinching time in these cases is found uneven drop formation is noted. The process of drop formation is
to increase due to the high viscous resistance during neck filament still dripping in nature; however, variation in drop volume and
thinning. Such phenomena are not observed for dispersed fluids of asymmetry in drop shape at the instant of detachment is often
low lr as noted in Fig. 8 (c) and (d). Fig. 7 (series d1-d5) provides a observed. In this regime, we also noted the formation of the long
photographic sequence of the oscillatory reduction of the neck of a stem at the nozzle tip (See d1 and d6 in Fig. 7). This is the regime
7
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

Fig. 8. Change of non-dimensional Neck radius with pinching time. The inner diameter of the nozzle is 1.0 mm.

where the transition from dripping to jetting takes place. In regime drop volume variation for all the fluids generated in nozzle ID of
3, the drop formation is essentially jetting in nature. The drops are 1.0 mm, 0.6 mm, and 0.4 mm respectively. In this plot, only the
now formed from a steady liquid column far away from the nozzle drop volume of the primary drops (largest) are considered. This
tip. In this regard, it is essential to know the drop volume distribu- does not affect the results as the primary drop volume is of the
tion to ascertain the effect of injection velocity. In regime 1, the order of mean drop volume. From these plots, it is clearly discern-
standard deviation of the volume is very low implying uniform able that for a particular nozzle, the drop volume differs as the
drop generation. injected fluid is varied in all three regimes identified earlier. More-
Fig. 9 (b) shows the probability distribution of drop volume at over, as the nozzle size decreases, the drop volume also decreases
an injection velocity of 0.275 and 0.321 m/s in regime 2; and irrespective of injected dispersed fluid. Further, it is noticed that at
0.367, 0.413 and 0.504 m/s in regime 3. In regime 2, the probability any particular flow rate within the regime 1 and 2, the cyclohexane
distribution exhibits a narrow range compared to regime 3, essen- drop volume is the highest compared to other fluids (see Fig. 9 (c),
tially signifying a lesser variation of drop volume in regime 2. Also, (d) and (e)). This difference in drop volume reduces after the tran-
as the injection velocity increases, the left arm of the bell curve sition. It has been observed that with the change in dispersed liq-
shifts towards the lower end of the range. This indicates the forma- uid properties such as interfacial tension, density, and viscosity;
tion of smaller drops at a higher velocity. significant variations in the final drop diameter and volume are
Though experiments have been conducted with four different noted particularly in the dripping regime.
nozzles, systematic variations of flow rates over a considerable At a particular nozzle, the generated drop sizes are measured
range could be made only in nozzle IDs of 1.0 mm, 0.6 mm and for variation in dispersed fluid properties and injection velocity.
0.4 mm. Nozzle ID of 0.3 mm, which has the lowest inner diameter, Such an assessment may help understand and compare the influ-
could not be operated over a wide range of flow rates. Within the ence of interfacial tension and the viscosity ratio on overall drop
operational limits, efforts have been made to collect the maximum size. In Fig. 10 the drop volume of cyclohexane, kerosene, n-
number of experimental data to facilitate a good comparison and heptane, and n-hexane are plotted against the injection velocity
cover a wide range of flow rates. Fig. 9 (c), (d), and (e) show the of the dispersed phase for nozzle ID 1.0 mm. Within the range of
8
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

Fig. 9. (a) Average drop size vitiation with an increase in the flow rate of cyclohexane in nozzle 1; (b) Probability density of cyclohexane drop volume formed at nozzle ID of
1.0 mm with an increase in injection velocity; (c), (d) and (e) drop volume variation with an increase in injection velocity in nozzle ID of 1.0 mm, nozzle ID of 0.6 mm and
nozzle ID of 0.4 mm respectively.

reported injection velocity, the variations in drop formation regime For high flow rates, we find a large deviation between experimen-
like dripping (a, d, g, and j), intermediate (b, e, h, and k), and jetting tal data and the correlation. It is factual to mention here that the
(c, f, i, and l) are also noted (see the snapshots of individual dis- proposed correlation is limited to low injection velocities only.
persed fluid in Fig. 10). In the present case, the drop volume is Nonetheless, cyclohexane is found to generate the largest drop
obtained by averaging ten consecutive drops since an uneven drop among the groups, especially in the dripping regime. It is reported
formation is present in the intermediate and jetting regime. To in the literature (Byakova et al., 2003; Gnyloskurenko et al., 2003)
compare the present experimental result, the proposed correlation that the final drop volume is dependent on factors such as three-
of Scheele and Meister,(1968) is also plotted in the same figure for point contact line position, the wetting characteristics of the liq-
each fluid separately. uids, the contact angle hysteresis, the thickness of the nozzle wall
2 !1=3 3 (Clanet and Lasheras, 1999), etc. We propose to probe the issue fur-
pri Di 20lQ l Di 4qd Q l U n Q 2l D2i qd ri ther. While recognizing the difficulty to practically separate out the
V d ¼ F4 þ 2  þ 4:5 5 ð3Þ
g Dq D F g Dq 3g Dq ðg DqÞ2 effect of all the above factors, at a very low flow rate, we assume
and define capillary length as, kc  ðri =Dqd g Þ1=2 by comparing
At a very low injection velocity, we find a very well match the interfacial tension with gravitational force and used it as the
between the present experimental results except for kerosene. characteristic length. The value of kc is 4.69 mm, 4.14 mm,
The reason for the deviation in the case of kerosene is not apparent. 3.95 mm and 3.84 mm for cyclohexane, kerosene, n-heptane, and
Although, the correlation is found to predict the trend correctly. n-hexane respectively. Moreover, to understand the effect of
9
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

Fig. 10. Comparison of drop volume of cyclohexane, kerosene, n-heptane, and n-hexane are plotted against the injection velocity; the nozzle ID is 1.0 mm. The bar represents
1.264 mm.

dispersed phase viscosity on drop formation, the combined role of ascertained since the observation from images gives a perception
viscous, inertia, and interfacial tension forces needs to be ascer- of complete wetting of the top surface.
tained. With kc a characteristic length, we define Ohnesorge (Oh) It is also found that, for kerosene, and cyclohexane, initially, the
number as (Mckinley et al., 2011) drop volume slightly increases with flow rate and reaches the max-
!1=2 imum value, beyond which a further increase in flow rate results in
 
t Viscocapillary ld k c q k3 l a decrease in drop volume. Recently Chaurasia et al. (2015) also
Oh ¼ ¼ = d c ¼ pffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi
d
ð4Þ
tRayleigh ri ri ðqd kc ri Þ observed such phenomena for Octane, Dodecane, and Hexadecane;
but they attributed this change to dynamic interfacial tension (due
The calculated values of Oh are 4.469  103, 10.170  103, to the added surfactant). Since no surfactant is added to the dis-
2.125  103, and 1.676  103 for cyclohexane, kerosene, n- persed fluid in the present work, the reason for a local maximum
heptane, and n-hexane, respectively. Clearly, there are two groups, in drop volume in low injection velocity may not be a surfactant.
viz. higher (cyclohexane and kerosene) and lower Oh (n-heptane The higher Ohvalue fluids (cyclohexane and kerosene) have high
and n-hexane). The Oh number essentially signifies the ratio visco-capillary time, signifying a longer time for neck and final
between visco-capillary time tViscocapillary (which describes the thin- pinching of the drop (as already discussed in Section 3.3). There-
ning mechanics of liquid thread under visco-capillary effect) and fore, a small window may be present in which an increase in flow
Rayleigh time tRayleigh (which characterizes the breakup of inviscid rate essentially adds more liquid in the forming drop while pinch-
liquid thread or jet) (Eggers, 1993). Since, for a fluid with a large ing takes place, yielding a larger drop. Accordingly, no sign of local
Ohnumber, visco-capillary time is large, hence the time required maximum in drop volume in the case of n-heptane and n-hexane
for liquid necking is more, which would eventually feed more liq- (which are part of the low Ohgroup) is found as shown in Fig. 9
uid into the drop during the drop growth stage as constant flow (c), (d) and (e).
rate condition is maintained at nozzle end.
The above comparison shows that cyclohexane has the highest 3.5. Effect of dispersed fluid property on drop leading edge length
value of kc and an intermediate value of Ohnumber. This combina-
tion of Ohand kc completes the picture of all forces excluding iner- From the preceding section, we observed that drop formation in
tia force, which can be neglected in the dripping regime. For the an immiscible environment might exhibit phenomenological dif-
present study, cyclohexane with the highest kc (largest capillary ference in drop separation process, such as drop formation on tip
length) and the intermediateOh number may yield the largest of the nozzle and drop formation from tip of liquid jet. To study
drop, compared to other dispersed liquids irrespective of nozzle the influence of nozzle geometry, injection velocity and fluid vis-
diameter as experimentally evident from Fig. 9 (c), (d), and (e). cosities etc. on drop geometry, one unifying factor is drop leading
Even though kerosene shows the largestOh value, its capillary edge length. Drop leading edge length (Lv ) is the distance between
length is smaller than cyclohexane. Therefore, kerosene forms the farthest point of the drop interface in the direction of flow
smaller drops compared to those formed by cyclohexane, espe- injection and the nozzle top surface. This dimension is also an
cially in the dripping regime. However, the exact role of nozzle important characteristic length (Zhang, 1999) to investigate the
top surface wettability by different dispersed fluids could not be effect of dispersed fluid properties and flow rate on drop growth.
10
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

Fig. 11. Variation of Lv against time at different lighter phase Reis illustrated. At the inception of a drop, the seeding height is small for low Re; however, it increases with the
flow rate. The subplot (a) and (b) for nozzle ID of 0.3 mm (smallest); (c) and (d) for nozzle ID of 1.0 mm (largest). The capillary length is shown in solid lines.

With the decrease in nozzle diameter, in the present experiments, of dispersed phase Reynolds number (Re)as shown in Fig. 11 for
the drop volume decreases all through dripping to jetting. How- nozzle ID 0.3 mm (smallest) and 1.0 mm (largest). For both the
ever, a wide distribution of drop size exists. To understand the nozzles, Lv remains constant over the entire range of Re, as illus-
effect of nozzle geometry on drop formation, kc is compared with trated in Fig. 11. However, at the higher end of Re, a sudden
Lv , and both are plotted against drop formation time over a range increase in Lv may be witnessed. For nozzle ID of 0.3 mm (smallest),

Fig. 12. Non-dimensional plot of Lv and time for cyclohexane drop (a) nozzle ID of 0.3 mm; (b) nozzle ID of 1.0 mm.

11
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

Lv is found very close to kc , for all the fluids, only cyclohexane and Based on the values of standard deviation (SD), the process of
n-hexane are shown in (a) and (b) of Fig. 11, respectively. drop generation is divided into three regimes viz. regime 1
It is observed that, in the dripping regime, irrespective of dis- (0 < SD < 0.2), regime 2 (0.2 < SD < 1.5) and regime 3 (1.5 < SD < 6.
persed fluid, Lv is found very close to kc , particularly in nozzle ID 0). The probability distribution exhibits a narrow range in regime 2
of 0.3 mm. As the nozzle diameter is increased, a wide difference compared with regime 3, essentially signifying a lesser variation of
in kc and Lv is noticed. One such comparison is presented for nozzle drop volume in regime 2. Also, in the dripping regime, the effect of
ID of 1.0 mm in (c) and (d) of Fig. 11. As the nozzle diameter dispersed liquid properties such as interfacial tension, density, and
increases, Lv also increase. However, for a wide range of Re, Lv viscosity is found significant and large variations in the final drop
remains the same. This is shown with a dotted line in Fig. 11 for volume are noted. Whereas in regime 3 the effect of dispersed fluid
cyclohexane and n-hexane. This implies that in the dripping properties is insignificant. It is also found that barring the initial
regime, even though the inertial force of the dispersed fluid shape reversal, Lv increases monotonically as the drop grows to
increases with increases in Re, the drop formation transpires fol- exhibit three distinct stages in the dripping regime. Also in the
lowing a consistent geometric pattern (see in Fig. 11). Whereas, dripping regime, with an increase in Re, even though with
the drop volume found to decrease with the increased generation increased frequency and decreased drop volume the drop forma-
frequency. The corresponding snapshot of a typical drop at the tion ensures a consistent geometric pattern.
instant of pinch-off is also shown in the plots.
The initial point in the plot of Lv in Fig. 10 shows the inception of
CRediT authorship contribution statement
a typical drop. The last point of any individual curve represents the
pinching of a drop from the nozzle tip. Barring the initial shape
Abhik Majumder: Methodology, Conceptualization, Investiga-
reversal (as discussed in Fig. 5), Lv increases monotonically as the
tion, Formal analysis, Writing - original draft, Writing - review &
drop grows to exhibit three distinct stages: initial rapid growth
editing. Dibyendu Ghosh: Software, Validation, Writing - review
phase during which the drop acquires the stable spherical shape,
& editing. Prasanta Kumar Das: Resources, Conceptualization,
steady growth phase up to a state of equilibrium between resistive
Supervision, Project administration, Writing - review & editing.
and disruptive force while the drop retains its spherical symmetry;
and finally rapid necking phase followed by a detachment of the
drop. These three stages are not very clearly discernable for high Declaration of Competing Interest
Reflow shown in Fig. 11; however, the non-dimensional plot
shown in Fig. 12 clearly illustrates these stages I, II, and III. This sig- The authors declare that they have no known competing finan-
nifies that in an unconfined system with an increase in flow rate cial interests or personal relationships that could have appeared
through the nozzle, drops may be generated with increased fre- to influence the work reported in this paper.
quency keeping the Lv approximately the same.
Acknowledgement

4. Summary and conclusions The authors of this manuscript acknowledge the help and sup-
port received from IIT Kharagpur and NIT Agartala.
Drop generation in a host quiescent immiscible liquid through a
submerged nozzle is one of the simplest mechanism, yet involves
complex and intriguing interfacial evolution owing to the interplay References
among counteractive forces. Although the fundamental theories Bertrandias, A., Duval, H., Casalinho, J., Giorgi, M.L., 2017. Dripping to jetting
underlining the mechanism are reasonably well-advanced; even transition for cross-flowing liquids. Phys. Fluids 29, 044102–44109. https://doi.
so, this requires improvement to apprehend some crucial aspects org/10.1063/1.4979266.
Bogy, D.B., 1979. Drop formation in a circular liquid jet. Annu. Rev. Fluid Mech. 11,
of the process. Experiments are conducted to study the dynamics
207–228.
of drop formation in an unconfined system where the dispersed Borthakur, M.P., Biswas, G., Bandyopadhyay, D., 2017. Formation of liquid drops at
liquid is injected through a submerged nozzle into bulk immiscible an orifice and dynamics of pinch-off in liquid jets. Phys. Rev. E 96, 013115–
liquid. The main focus of the study is to understand the effect of 13211. https://doi.org/10.1103/PhysRevE.96.013115.
Bremond, N., Santanach-Carreras, E., Chu, L.-Y., Bibette, J., 2010. Formation of liquid-
injection velocity and dispersed fluid property on the behavior of core capsules having a thin hydrogel membrane: liquid pearls. Soft Matter 6,
remnant mass, pinching, drop characteristics, and leading-edge 2484. https://doi.org/10.1039/b923783f.
length variations. Four different nozzles and dispersed fluid are Byakova, A.V., Gnyloskurenko, S.V., Nakamura, T., Raychenko, O.I., 2003. Influence of
wetting conditions on bubble formation at orifice in an inviscid liquid
used to study the effect of some important parametric variation. Mechanism of bubble evolution. Colloids Surf. A Physicochem. Eng. Asp. 229,
High-speed imaging is used to capture the dynamics of different 19–32. https://doi.org/10.1016/j.colsurfa.2003.08.009.
drop formation stages. In the dripping regime, the measured drop Chang, T.M.S., 2004. Artificial cell bioencapsulation in macro, micro, nano, and
molecular dimensions: keynote lecture. Artif. Cells. Blood Substit. Immobil.
volume compared well with the existing correlation. Immediately Biotechnol. 32, 1–23. https://doi.org/10.1081/BIO-120028665.
after a successful drop detachment, the remnant liquid mass Chaurasia, A.S., Josephides, D.N., Sajjadi, S., 2015. Buoyancy - driven drop generation
attached to the nozzle tip undergoes retracting motion. The pre- via microchannel revisited. Microfluid. Nanofluidics 18, 943–953. https://doi.
org/10.1007/s10404-014-1484-x.
sent investigation reveals that such retracting motion can intro- Clanet, C., Lasheras, J.C., 1999. Transition from dripping to jetting. J. Fluid Mech. 383,
duce oscillations and shape readjustment of the remnant mass. 307–326.
The process of pinching is the last stage of drop formation. In order Cramer, C., Studer, S., Windhab, E.J., Fischer, P., 2012. Periodic dripping dynamics in
to study the effect of lr on the pinching process, the pinching time a co-flowing liquid-liquid system. Phys. Fluids 093101, 093101–93114. https://
doi.org/10.1063/1.4752477.
for all the fluids is compared at the same flow rate. The quantita- Eggers, J., 1993. Universal pinching of SD axisymmetric free-surface flow. Phys. Rev.
tive measurement of the time scale of the drop formation process Lett. 71, 3458–3460.
at the dripping regime shows the pinching time is large in higherlr Eggers, J., Villermaux, E., 2008. Physics of liquid jets. Reports Prog. Phys. 71. https://
doi.org/10.1088/0034-4885/71/3/036601.
fluids. Whereas close to the end of the dripping regime, such a Gnyloskurenko, S.V., Byako, A.V., Raychenko, O.I., Nakamura, T., 2003. Influence of
pinch-off process occurs from longer liquid threads and is sub- wetting conditions on bubble formation at orifice in an in v iscid liquid.
jected to radial oscillation and subsequent reduction of the neck Transformation of bubble shape and size. Colloids Surfaces A Physicochem. Eng.
Asp. 218, 73–87. https://doi.org/10.1016/S0927-7757(02)00592-7.
radius. The pinching time in these cases is found to increase partic- Harkin, W.D., Brown, F.E., 1919. The determination of surface tension (free surface
ularly for high lr fluids. energy), and the weight of falling drops: the surface tension of water and

12
A. Majumder, D. Ghosh and P. Kumar Das Chemical Engineering Science 245 (2021) 116808

benzene by the capillary height method. J. Am. Chem. Soc. 41, 499–524. https:// device: effect of the viscosities of the liquids. Microfluid. Nanofluid. 5, 585–594.
doi.org/10.1021/ja01461a003. https://doi.org/10.1007/s10404-008-0271-y.
Hayworth, C.B., Treybal, R.E., 1950. Drop formation in two-liquid-phase systems. Patrascu, C., Chiriac, E., Balan, C., 2017. Drop dispensing in a viscous outer liquid.
Ind. Eng. Chem. 42, 1174–1181. https://doi.org/10.1021/ie50486a030. INCAS Bull. 9, 101–110. https://doi.org/10.13111/2066-8201.2017.9.4.9.
Homma, S., Koga, J., Matsumoto, S., Song, M., Tryggvason, G., 2006. Breakup mode of Rao, E.V.L.N., Kumar, R., Kuloor, N.R., 1966. Drop formation studies in liquid—liquid
an axisymmetric liquid jet injected into another immiscible liquid. Chem. Eng. systems. Chem. Eng. Sci. 21, 867–880. https://doi.org/10.1016/0009-2509(66)
Sci. 61, 3986–3996. https://doi.org/10.1016/j.ces.2006.01.029. 85081-9.
Josephides, D.N., Sajjadi, S., 2015. Microfluidic method for creating monodisperse Sabliov, C.M., Boldor, D., Keener, K.M., Farkas, B.E., 2002. Image processing method
viscous single emulsions via core – shell templating. Microfluid Nanofluid 18, to determine surface area and volume of axi-symmetric agricultural products.
383–390. https://doi.org/10.1007/s10404-014-1439-2. Int. J. Food Prop. 5, 641–653. https://doi.org/10.1081/JFP-120015498.
Lakdawala, A.M., Gada, V.H., Sharma, A., 2014. A dual grid level set method based Scheele, G.F., Meister, B.J., 1968. Drop formation at low velocities in liquid-liquid
study of interface-dynamics for a liquid jet injected upwards into another systems: Part I. Prediction of drop volume. AIChE J. 14, 9–15. https://doi.org/
liquid. Int. J. Multiph. Flow 59, 206–220. https://doi.org/10.1016/j. 10.1002/aic.690140105.
ijmultiphaseflow.2013.11.009. Shah, R.K., Shum, H.C., Rowat, A.C., Lee, D., Agresti, J.J., Utada, A.S., Chu, L., Kim, J.,
Lakdawala, A.M., Thaokar, R., Sharma, A., 2015. Break-up of a non-Newtonian jet Fernandez-nieves, A., Martinez, C.J., Weitz, D.A., 2008. Designer emulsions using
injected downwards in a Newtonian liquid. Sadhana - Acad. Proc. Eng. Sci. 40, microfluidics. Mater. Today 11, 18–27. https://doi.org/10.1016/S1369-7021(08)
819–833. https://doi.org/10.1007/s12046-015-0349-7. 70053-1.
Lin, S.P., Reitz, R.D., 1998. Drop and spray formation from a liquid jet. Annu. Rev. Sierou, A., Lister, J.R., 2004. Self-similar recoil of inviscid drops. Phys. Fluids 16,
Fluid Mech. 30, 85–105. 1379–1394. https://doi.org/10.1063/1.1689031.
Longmire, E.K., Norman, T.L., Gefroh, D.L., 2001. Dynamics of pinch-off in liquid/ Smith, S.W.J., Moss, H., 1917. Experiments with mercury jets. Proc. R. Soc. A Math.
liquid jets with surface tension. Int. J. Multiph. Flow 27, 1735–1752. https://doi. Phys. Eng. Sci. 93, 373–393.
org/10.1016/S0301-9322(01)00030-1. Utada, A.S., Fernandez-Nieves, A., Stone, H.A., Weitz, D.A., 2007. Dripping to jetting
Mckinley, G.H., Renardy, M., Tech, V., 2011. Wolfgang von Ohnesorge. Phys. Fluids transitions in coflowing liquid streams. Phys. Rev. Lett. 99, 1–4. https://doi.org/
23, 127101–127106. https://doi.org/10.1063/1.3663616. 10.1103/PhysRevLett.99.094502.
Meyer, R.F., Crocker, J.C., 2009. Universal dripping and jetting in a transverse shear Wang, W., Ngan Ho, K., Gong, J., Angeli, P., 2009. Observations on single drop formation
flow. Phys. Rev. Lett. 102, 194501–194504. https://doi.org/10.1103/ from a capillary tube at low flow rates. Colloids Surfaces A Physicochem.
PhysRevLett.102.194501. Eng. Asp. 334, 197–202. https://doi.org/10.1016/j.colsurfa.2008. 10.011.
Nie, Z., Seo, M., Xu, S., Lewis, P.C., Mok, M., Kumacheva, E., Whitesides, G.M., Zhang, X., 1999. Dynamics of drop formation in viscous flows. Chem. Eng. Sci. 54,
Garstecki, P., Stone, H.A., 2008. Emulsification in a microfluidic flow-focusing 1759–1774.

13

You might also like