Hydrodynamics of Liquid-Liquid Flows in Micro Chan
Hydrodynamics of Liquid-Liquid Flows in Micro Chan
Hydrodynamics of Liquid-Liquid Flows in Micro Chan
Review
Hydrodynamics of Liquid-Liquid Flows in Micro Channels and
Its Influence on Transport Properties: A Review
Arijit A. Ganguli 1,2, * and Aniruddha B. Pandit 2
1 School of Engineering and Applied Sciences, Ahmedabad University, Ahmedabad 380009, India
2 Chemical Engineering Department, Institute of Chemical Technology, Matunga, Mumbai 400019, India;
[email protected] or [email protected]
* Correspondence: [email protected] or [email protected]; Tel.: +91-8861817755
Abstract: Hydrodynamics plays a major role in transport of heat and mass transfer in microchannels.
This includes flow patterns and flow regimes in which the micro-channels are operated. The flow
patterns have a major impact the transport properties. Another important aspect is the pressure
drop in micro-channels. In the present review, the experimental and Computational Fluid Dynamics
(CFD) studies covering all the above aspects have been covered. The effect of geometrical parameters
like shape of channel, channel size, material of construction of channels; operating parameters like
flow velocity, flow ratio and fluid properties have been presented and analyzed. Experimental
and analytical work of different pressure drop models has also been presented. All the literature
related to influence of flow patterns on transport properties like volumetric mass transfer coefficients
(VMTC) and heat transfer coefficients (HTC) have been presented and analyzed. It is found that most
works in Liquid-Liquid Extraction (LLE) systems have been carried out in slug flow and T-junctions.
Models for coupled systems of flow and mass transfer have been presented and works carried out
for different coupled systems have been listed. CFD simulations match experimental results within
20% deviations in quantitative and qualitative predictions of flow phenomena for most research
Citation: Ganguli, A.A.; Pandit, A.B. articles referred in this review. There is a disparity in prediction of a generalized regime map and a
Hydrodynamics of Liquid-Liquid generalized regime map for prediction of flow patterns for various systems would need the help of
Flows in Micro Channels and Its Artificial Intelligence.
Influence on Transport Properties: A
Review. Energies 2021, 14, 6066. Keywords: microchannels; flow patterns; CFD; pressure drop; high speed camera; particle
https://doi.org/10.3390/en14196066
tracking velocimetry
as well as its effect on transport properties both in gas-liquid (GL) and liquid-liquid (LL)
systems. [5,6]. However, reviews on flow patterns of LL systems focus on major flow
patterns like slug-flow, parallel flow and transition flow. It is important to note that the
viscosity of the continuous and dispersed phase gives rise to various different flow pat-
terns [7,8]. Also, the flow patterns depend on the bubble formation and breakup during the
tubing, squeezing and dripping regimes and transitions between them. Further, reviews [5]
have focused mostly on the effect of geometrical and operating parameters, material of
construction of microchannels but the effect of surfactants, effect of hydrophobicity and
hydrophilicity on walls, effects of wall films on flow patterns have not been thoroughly
discussed. The present work focuses on reviewing all the prominent literature works
(including experimental, numerical works) on flow patterns and regime maps. First, state
of the art of experimental and CFD studies describing flow patterns in liquid-liquid flows
have been analyzed. This includes flow patterns in different geometries and geometrical
parameters like shape, size, material of construction, operating parameters like flow rates,
flow ratios, fluids properties of the fluids under consideration, measurement techniques
used by researchers to understand internal circulations in slugs of both phases are also
presented. Secondly, most important flow regime maps signifying the maps based on forces
and based on dimensionless groups (individual and combination of different dimensional
numbers have been presented. Finally, pressure drop in microchannels (experimental data)
of various authors has been presented for different two phase systems and geometries. The
corresponding models and their limitations have been presented. Finally, the review lists
out the various liquid-liquid reactive and non-reactive systems carried out by researchers.
Pressure drop studies has been covered well in recent review by Al-Azzawi et al. [5]. In
the present work, the most prominent works have been reported. The prominent works
showing effect of the flow patterns on transport properties (specifically Volumetric Mass
Transfer Ceofficients (VMTC)) have been included. Figures 1 and 2 show the entire struc-
ture of the manuscript. Figure 1 focuses on the experimental studies while Figure 2 focuses
on the CFD studies.
Energies 2021, 14, 6066 3 of 56
Figure 1. Flow structure of the experimental studies in the manuscript. Black arrows show interactions between flow patterns and other parameters andflow regimes.
Energies 2021, 14, 6066 4 of 56
2. Hydrodynamics in Microchannels
Micro-channels provide advantages like high surface to volume ratio that ensures
better control on size of the droplets. A typical experimental set-up for carrying out
experiments to study flow patterns, mass transfer, pressure drop etc. is shown in Figure 3.
The two immiscible liquids (aqueous and organic, from reservoirs (1 and 2) are introduced
by continuously operating high-precision piston pumps (3 and 4) to a symmetric 120◦
Y-junction or T-junction made of materials like Teflon, glass, steel etc. (5). The pumps are
controlled with the help of a computer to adjust the flow rate. The capillary contactor (6) is
made of poly-tetra-fluoro-ethylene (PTFE) and was attached directly downstream of the
Y-or T junction. The length of the capillary micro-channel can be adjusted easily. Thus, the
setup was very flexible with respect to the residence time used. At the end of the contacting
stage, a flow splitter (7), was used to separate the phases.
Figure 3. Schematic of the experimental setup (1,2) Reservoirs containing aqueous and organic
liquids respectively; (3,4) feed pumps; (5) Y-junction mixing element; (6) slug-flow capillary (7) flow
splitter and (8,9) sampling bottles. (Reproduced with permission from [9]) the two outlets of the
splitter were collected in the sampling bottles (8,9) for further analysis. With permission from Kashid
et al. [1].
Typical configurations for micro-channels are shown for in Figure 4. The micro-channel
geometries are different in cross sectional shape, main channel shape and mixing junction
configuration. Regarding the cross sectional shape of the micro-channel, micro-channels
can be classified into two main groups: circular and non-circular micro-channels, the latter
can be square, rectangular, trapezoidal, etc. Different types of junction configurations are
used at the micro-channel entrance. The most popular mixing junctions are T-junction,
Y-junction, cross-shaped junction, cross flow T-junction and concentric junction. The cross
sectional dimension of the inlet channels can be different from those of the main channel.
Different junction configurations can produce different stable flow regimes. For example, a
micro-channel with a Y-shaped junction is more appropriate for producing a stable liquid-
Energies 2021, 14, 6066 5 of 56
liquid parallel flow as compared to that with a T-shaped junction [9]. Droplet flow in a
micro-channel can be achieved by using a concentric inlet or cross-shaped junction. Further,
details into dispersion in micro-channels with different junction structures can be found in
the literature [10–12].
Figure 4. Types of micro-channels with different junctions 1 Inlet for continuous phase 2 Inlet for
dispersed phase 3 Outlet of the channel (a) Cross flow T-junction (b) Y-junction (c) Cross Junction
(d) Concentric junction (e) T-junction.
In the forthcoming sections, we will discuss about the flow patterns, pressure drop
in micro-channels using T and Y junctions and review the current experimental and CFD
modeling works.
Droplet size variation depends on Ca number. For example, for Ca = 0.019 droplet
size varies upto 15%. While the variation is lower for Ca = 0.011 is higher than 15% [22].
3. Transition regime: This regime is transition from dripping to sqeezing regime. In
transition regime the Ca numbers are in range (0.009 < Ca < 0.013) and Re < 1. Hence
inertial forces have no significant effect on flow patterns
4. Sqeezing regime: In this regime the inertial forces play a big role.
5. Jetting: In this type of flow the organic core thread falls in certain distance into the
outlet main micro-channel before it breaks up into slugs/droplets. At jet breakup,
interfacial shear is overcome by viscous shear. (Figure 5l).
In the upcoming subsections, the effect of different parameters like geometrical param-
eters like shape, size of microchannels, material of construction of channels, type of junction
for flow mixing and inlet, fluid properties of the dispersed and continuous phases like the
viscosity ratio, interfacial tension and the effect of other factors like addition of agents or
application of vacuum to stabilize flow patterns have been presented. Table 1 gives the
literature works on the flow patterns while Figures 6 and 7 gives the chart showing the
flow patterns due to the sensitivity of several parameters involved.
Energies 2021, 14, 6066 9 of 56
Equipment
Authors Channel (CH) Details Type of Study System Flow Regime
Used
CH Flow
CH Length
Geometry Type Diameter/Size CH Material Rate/Velocity Experimental/CFD
(mm)
(mm) (m/s)
Glycerol-
High Speed
kerosene; Oil
Square CMOS Camera
Su et al. [23] 0.6 × 0.6 2 NA 0.6 mL/min Experimental soluble Slug flow
microchannel system (Basler
surfactant Span
A504KC)
80
Slug flow;
T-square; 0.4 × 0.4 56
Kashid and Transition flow;
Y-rectangular; 0.5 × 0.4 75 5.8 × 10−3 High Speed
Kiwi-Minsker NA Experimental Water- Toluene Annu-
Trapezoidal; 0.25 × 0.292 40 m3 /s; 0.025 m/s Camera
[24] lar/Parallel
Concentric d = 0.5 200
flow
Plug flow; Plug
flow with
suspended High Speed
(2 < Qc < 6 Castor
Kovalev et al. 0.12 × 0.12 particles; Camera; Particle
T-junction 11.5; 22.5 SU-8 1.5 < Qc < 20) Experimental oil-Distilled
[8] 0.24 × 0.12 Droplet Flow; Tracking
µL/min water
Throat annular; Velocimetry
Slug flow;
Parallel Flow
0.6 × 0.6
Svensson et al. Square Qc = 8 mL/h; High Speed
0.4 × 0.4 5 Glass Experimental Water-n-hexane Slug flow
[25] microchannel Qd = 2 mL/h Camera
0.6 × 0.3
Water, dodecane,
5.8 × 10−3 <
nitric acid, High Speed
Vansteene et al. T, X Junction Qd< 0.033 Segmented/Slug
0.1 × 0.1 1.125 Glass Experimental Hydrochloric Camera mini
[26] microchannel mL/h; 0.07 < Qc flow
acid, AX 100
< 11
DMDBTDMA
40 < ud < 400
PDMS, Glass; TS-100S nikon
Li et al. [27] Y channel 0.24 × 0.008 2 mm µm/s; 100 < uc < Experimental Oleic acid-water Parallel flow
Quartz Japan
1000 µm/s
Energies 2021, 14, 6066 10 of 56
Table 1. Cont.
Equipment
Authors Channel (CH) Details Type of Study System Flow Regime
Used
CH Flow
CH Length
Geometry Type Diameter/Size CH Material Rate/Velocity Experimental/CFD
(mm)
(mm) (m/s)
55% Glycerol +
Li et al. [28] T-junction 0.3 × 0.15 5 mm Glass 0.5–5 µL/min Experimental Drop flow Micro-PIV
15%Span80
Non-reactive
systems:
Kerosene-water
Yagodnitsyna T-junction 0.2 × 0.2 All the different High Speed
22.5 SU-8 Experimental ParaffinOil-
et al. [7] microchannels 0.2 × 0.4 flow patterns Camera
water
castorOil-
ParaffinOil
Water-butanol; All flow
Rectangular Hydraulic dia: High Speed
Cao et al. [15] 105 Glass 0.2–150 mm/s Experimental Water-toluene; patterns
microchannelss 0.2, 0.4, 0.6 Camera
Water-hexane described
0.15 < uw< 580
Yagodnisyana 0.12; 0.24; 0.12 Ionic-liquid— High Speed
T-junctions 11.5; 22.5 Glass mm/s; 0.3 < uIL Experimental Plug Flow
et al. [29] 0.1; 0.4; 0.2 water Camera
< 14.5 mm/s
Kerosene-water;
Yagodnisyana 0.2 × 0.2; 0.2 × Flow rates not Parafin Oil
T-junctions 11.5; 22.5 Glass Experimental Plug Flow Micro-PIV
et al. [30] 0.4; provided water; castor Oil
water
Energies 2021, 14, 6066 11 of 56
Figure 6. Different flow patterns for different shapes of microchannels studied by authors.
Energies 2021, 14, 6066 12 of 56
Figure 7. Chart showing interactions between flow patterns and different geometric and operating parameters structures and shapes. VR represents viscosity ratio.
Energies 2021, 14, 6066 13 of 56
velocity on plug lengths and found that plug length increased with flowrate ratio increase
while decreased with increase in bulk flow.
Yagodnitsyna et al. [30] carried out investigations in a T-shaped microchannel using
3 sets of LL systems: Kerosene-water; paraffin-oil—water; castor-oil—water. The authors
obseved different flow patterns like rivulet flow instead of annular flow due to the wet-
tability at the walls by the liquids of systems investigated. Serpentile flow was observed
for higher Weber numbers. The authors concluded that flow pattern maps were different
for different fluidic systems. Further, absence of abundant information on influence of
parameters like density, viscosity, surface tension, contact angle on flow patterns were
highlighted by the authors. Most research works used systems where one of the liquids
did not wet the fluid walls. The authors concluded that rivulets had less velocities than
plugs based on a relation between velocity of plugs and bulk velocities. For the systems
considered the relation was U plug = 1.21Ubulk . Further, the circulations grow linearly with
plug velocity.
Viscosity of both phases play an important role in flow patterns both in the continuous
and dispersed phases (velocities inside droplets, slugs/plugs). Su et al. [23] found that
when dispersed phase viscosities are higher than continuous phase viscosities, faster flow
transition took place from slug flow to parallel flow and resulted in smaller slugs. Kovalev
et al. [8] carried out experimental investigations in T-junctions. Effects of continuous
phase viscosity on water slug lengths have been studied for very low capillary number
(2.32 × 10−3 –2.32 × 10−2 ) by Garstecki et al. [33]. They observed marginal change in the
slug length with an increase in silicon oil viscosity from 10 mPas to 100 mPas.
Energies 2021, 14, 6066 14 of 56
Kovalev et al. [8] found different flow patterns for low viscosity ratios upto 0.001.
The six distinct flow patterns were Plug flow, plug flow with suspended particles, droplet
flow, throat annular flow, parallel flow and slug flow (Figure 8). The definition of slug
flow (as defined by the authors) is different than the definition of literature works. Here
slug flow is periodical flow of dispersed phase segments which have assymetrical shapes
and wet at least one channel wall (as shown in Figure 8). According to the authors the
plug flow is the one which have diameter and length little less than the hydraulic diameter
of column. Further, the authors defined dropet flow as flow of almost spherical droplets.
Interesting features of plug flows have been reported by the authors (a) The plugs took
different shapes bullet like, dumble like shape with possibility of microdroplet breakoff. (b)
The velocity distribution inside the plugs were found to be three dimensional in nature
in contrast to two dimensional nature generally observed. (c) Flow structures evolved
downstream in contrast to that observed in literature. (d) Velocity distribution in plug-plug
interface was closely associated with each other. Dimensionless parameters (product of
flow ratio and Capillary number) were defined to define the plug flow.
Figure 8. Flow patterns of castoroil-water system with permission from Kovalev et al. [8]. (a) Plug
flow (b) plug flow with suspended particles (c) droplet flow (d) throat-annular flow (e) parallel flow
(f) slug flow.
Figure 9 shows how plugs and slugs are distinguished according to their lengths as
observed by Foroughi and Kawaji, [34]. If average length of the water segments was equal
to or less than 5 channel diameters the flow was classified as slug flow. The author have
kept the dispersed phase flow rate constant while increasing the flow rate of continuous
phase. With increase in continuous phase flow the flow patterns changed from drop to
plug to slug. With further increase in water flow sausage shaped annular patterns were
noticed. With highest flow rates, water droplets surrounding the oil core were observed.
The authors developed a correlation for slug length as a function of flow ratio and
capillary number and also related plug velocity as a function of bulk velocity. Vansteene
et al. [26], carried out experimental investigations in in T and Y junctions in glass mi-
crochannels in the dripping regime. The authors developed an empirical correlation for
normalized droplet diameter (NDD) for high viscosity ratios. NDD was a function of flow
ratio and capillary numbers. The empirical model was found to control flow parameters as
a means for optimization of droplet production frequency and specific interfacial area.
1, 14, x FOR PEER REVIEW 15 of 62
Energies 2021, 14, 6066 15 of 56
Figure 10. Plug shapes and flow structure at 5 mm, 10 mm and 15 mm from T-junction from castor oil
flow rate Q = 3 µL/min and several flowrates of water. Reproduced with permission from authors.
Energies 2021, 14, 6066 16 of 56
Li et al. [28] have studied internal flow in droplets for varying Bond numbers
(0.003 < Bo < 0.007) and Capillary numbers (4.05 x10−4 < Ca < 1.62 × 10−4 ) to see the effect
of viscosity ratio and flow on droplet size and droplet spacing and in turn on internal
circulations. The authors used µ-PIV in their investigations. The investigations were based
on dimensionless droplet size (ratio of droplet length to inlet channel width) and droplet
spacing (ratio of distance between two droplets to inlet channel width).
Role of Surfactants
Svensson et al. [25] found the effect of surfactants to two phase flows in square glass
microchannels. They reported that the surfactant molecules adsorb on the glass wall
hindering the flow between the organic phase slugs and the channel wall. This increases
the resistance and effects the transport properties (specifically mass transfer). However,
some authors have reported improvement in transport properties which will be discussed
Energies 2021, 14, 6066 17 of 56
Figure 11. (A) Regime map of Yagodnitsyna et al. [7]; (B) Regime map of Wu et al. [11]; (C) Regime map by Kashid and
Kiwi-Minsker [24]; slug flow; + slug-drop; × deformed interface; ◦ parallel-annular (D) Flow pattern map for silicon oil
water flow initially saturated with oil. Foroughi and Kawaji [34] (E) Regime map proposed by Zhang et al. [12]. Reproduced
with permission from authors.
Zhang et al. [12] carried out experimental investigations for flow patterns and mass
transfer using Toluene-water; toluene-sulfuric acid and ethylacetate-water systems. The au-
thors also carried out measurements to study the effect of temperature on the performance
of micro-channels for the different systems. The fluid systems were chosen so that they
could simulate reacting processes like nitration, sulfonation and acid catalysed reactions
and have different fluid properties enabling parametric studies. The authors observed a
lot of change in mixing behaviour for the operating range considered. Composite flow
regime maps (Figure 11E) were presented in the form of Cac Rec 0.5 v/s CaD 0.7 ReD 0.5 . The
experimental data as well as data from literature was compared with the regime map and
good prediction of flow patterns was found.
Table 2. Literature of studies of pressure drop in micro-channels for two phase systems.
∆P =
10 < Re < 150; L 8µ0 U plug aL 8µc Umix (1−α) L
(40%Ethylene +
Round micro- Fused 0.05–6 1000 ≤ ∆P ≤ LU
( R−h f ilm )
2 R2
Jovanovic Glycol)-
capillaries 0.248, 0.498 1000 silica mL/min Flow rate 10000 4 1 17, 18, 19 2
et al. [39] water- + LL σd C (3Ca) 3
Y-mixer h
toluene L plug
α= lU and
2
h f ilm = 1.3Ca 3
Y shaped Sulphuric
Rectangular Hydraulic 4.8–62.3 acid-sodium Flow rate; 18 < Re < 114; ∆P =
Fu [36] 20 PTFE 5 2 3, 4
microchan- dia = 0.174 mm3 /h bi-carbonate flow velocity 6 ≤ ∆P ≤ 30 ∆Pf + 21 ρ v2out + v2in + 2k ρv2
nel or water
Energies 2021, 14, 6066 21 of 56
Table 2. Cont.
Assumptions: 1. Constant dynamic contact angle 2 Constant slug length 3 Equal number of slugs of water and cyclohexane under similar operating conditions. 4 Neglecting end effects 5 For uniform micro-channels,
fully developed velocity profiles are obtained at x/L = 0.3. 6 In the flow rate calculations, water injection into the channel was assumed to have a negligible effect on the single-phase flow rate of oil through the
needle 7 Hagen-Poiseuille equation was assumed to be valid 8 the dynamic viscosity of the mixture depends on the viscosity of each phase, their volume fractions, the temperature, and the magnitude of the
dispersion of one phase in the other. Limitations: 1 CFD simulations for the geometries considered for experimental conditions have not been carried out 2 A single phase pressure drop instead of a two phase
pressure drop was used for validation with experiments Findings: 1 An expression for two phase pressure drop across a Y-junction micro-channel for slug flow conditions with and without film formation 2 Power
required per unit volume for most efficient conventional extractors was found to be around an order of magnitude higher than the micro-channels 3 An approximate single phase pressure drop model for
liquid-liquid mixtures was derived and validated with experimental results 4 Analytical velocity and pressure profiles for converging and diverging sections were derived and compared with numerical (CFD)
results. A good agreement was observed 5 Inlet configurations (T and Y junctions) did not affect the flow patterns or pressure drop 6 The flow patterns depend on the material of construction of the channels.
In glass micro-channels plug flow or intermittent flow with IL was observed while in Teflon micro-channels annular flow was observed at low volume fractions while plug and drop flow were observed at
higher volume fractions 7 Material also plays an important role in the determination of pressure drop. Two different Teflon micro-channels exhibit different pressure drops due to different plug length at same
operating conditions 8 Modeling film thickness is very important for good prediction of overall pressure drop. An improved film thickness model gave good predictions for pressure drop when compared with
experimental results 9 Proposed two models for evaluation of two phase pressure drop in square micro-channels. No film model and moving film model 10 The models showed a system specific behavior with
good agreement with water-toluene system while high deviations with SiliconOil-water system. Hence, the models work well with fluids with similar viscosities 11 Pressure drop increases with an increase 12
Lockhart Martinelli and homogeneous were used to interpret pressure drop in microchannels and it was observed that homogeneous models do not provide good fit with experimental data. 13 A new correlation
having two parameters that depend on microchannel type have been introduced 14 Pressure drops depend closely on type of micro channel material, the type of liquid that is introduced first and liquid flow rates
15 Prediction of flow regime map and flow patterns 16 Pressure drop could be predicted accurately using Lochart-Martinelli model for the system considered by the authors 17 Pressure drop is highly dependent
on slug size 18 It is important to correctly measure interfacial pressure drop 19 Film velocity has negligible influence on pressure drop.
Energies 2021, 14, 6066 22 of 56
The authors attributed this to lower film thickness of the film than the one assumed
for analytical model development [41]. For both systems considered there was an increase
of pressure drop with flow rate for both channels. For example, for silicon-oil-water system,
the pressure drop increased from 50 to 330 kPa for corresponding inlet velocities of 0.01 m/s
and 0.14 m/s for 0.2 mm channel diameters, the increase was restricted from 10 to 60 kPa
for corresponding inlet velocities of 0.01 m/s and 0.14 m/s for 0.4 mm channel diameters.
Similarly values for toluene-water systems for pressure drop were 30 and 225 kPa for
velocities of 0.02 and 0.27 m/s respectively for 0.2 mm channel diameters and 5 and 30 for
0.4 mm channel diameters.
Studies on two phase pressure drop carried by authors [37,38] show the following
findings:
1. Two phase pressure drops involve hydrodynamic and interfacial/capillary pressure
drops in microchannels
2. Both experimental investigations and analytical models show that two phase pressure
drop is higher than single phase pressure drop
3. Modelling two phase pressure drop involves careful consideration of two aspects (a)
film near the wall and (b) the movement of the film
4. Two phase pressure drop depends on (i) the fluids/system under consideration (ii)
volume fraction of dispersed phase (iii) material of construction of capillaries.
Most of the points mentioned above have been considered by the researchers [37] but
the model presented by the authors is valid only for fluids with similar viscosities and
show large deviations for fluids with large viscosity differences.
Figure 12A–C shows the experimental data of variation of single/two phase pressure
drops w.r.t Re for different authors. Figure 12A shows the variation of pressure drop as a
function of Re (5 ≤ Re ≤ 120) for water as working fluid. The channel diameter is 675 µm.
A linear relationship is observed. The pressure drop values range from 5 kPa to 100 kPa
respectively. Figure 12B shows the two pressure drop variation with respect to Reynolds
numbers. Systems with similar viscosities have been considered. Channel diameters range
from 400 to 500 µm respectively. It can be observed that the measured pressure drops
for [37–39] are similar till Re = 40 while deviations upto 50% between measurements
of [37,39] are observed after that. Deviations upto 50% can be observed. Probable reasons
are the type of junction being used and the difference in channel diameter magnitudes.
Figure 12C shows the pressure drop of Water-IL system by Tsaoulidis et al. [31]. The Re is
below 0.5 due to high viscosities of IL. The pressure drops are however higher ranging from
20 to 180 kPa. The researchers have validated their data with respective models as shown in
Table 2. Deviations of 10–20% have been shown by the authors for their respective models.
The pressure drop mode1 Jovanović et al. [39] was compared with another pressure drop
model of Kashid and Agar [38] and found that the latter over-predicted the experimental
data of former. Tsaoulidis et al. [31] have used both T and Y junctions but at very low
Reynolds numbers. They have also developed their own model which was seen to show
good agreement with experimental data with deviations of around 10%.
Figure 12. Pressure-drop variation with Reynolds number for equal flow rates of both phase liquids
(A) Single phase system with water as fluid and diameter 675 µm (B) Two phase systems N Water-
Cyclohexane system in Y-junction micro-channel of diameter 500 µm by Kashid et al. [1] Water-
Toluene system in Y junction by Jovanivić et al. [39] in Y junction micro-channel of diameter 498
µm Water- Toluene system by Ladosz and Rudolf von Rohr, [37] in Y junction micro-channel of
diameter 400 µm (C) Variation of pressure drop for Re < 0.5 for Water-Ionic-liquid system for channel
diameter of 270 µm by Tsaoulidis et al. [31].
easy and it is capable of resolving the interface accurately, it fails to describe significant
deformations of the interface. This is mainly because of the necessary redistribution of the
markers, due to deformation or, in case where the interface expands, the required addition
of new particles. Besides, this method can be very demanding in terms of computing
power and time for 3D simulations.
One of the most popular surface capturing method is the level set (LS) method [46,47].
Here, a function is used to locate the interface, which takes positive and negative values on
different sides of the interface and zero at the interface. The interface is therefore called
zero LS. The surface tension is modelled using the method by Brackbill et al. [48]. An
overview of the different LS methods is presented by Osher and Fedkiw, [49]. This method
is conceptually simple and easy to implement. Its main drawback is the loss of mass (or
volume), especially for significantly deformed interfaces van sind Annaland, [50].
The volume of fluid (VOF) method introduced by Hirt and Nichols, [51] represents
a typical volume capturing method. The basic idea of the VOF method is the definition
of a volume function which takes values between zero and one for the cells containing
the interface.
Energies 2021, 14, 6066 25 of 56
SOFTWARE
Authors CFD Details Assumptions Findings
USED
Single/Two VOF/ Levelset Size of
Dimension (D) Grid Details System Steady/Unsteady
Phase Method Microchannel
Free Surface
Kashid et al. In house code
Two phase 2D Coarse Model, VOF, Water-Toluene Unsteady state 1, 2, 3, 4 1, 2, 3, 4
[52] (FEATFLOW)
CSF, level set
Mesh Size–0.5 to
Kashid et al. 2 micrometer Ansys Fluent
Two phase 2D VOF Water-Toluene Unsteady state 1, 2, 3, 4 5
[44] Number of Ansys GAMBIT
cells-58,000
Medium Grid
Raj et al. [42] Two phase 3D VOF Water-Oil Unsteady state Ansys Fluent 1, 2, 3, 4 6, 7, 8
(72,960 Cells)
Yamasaki et al. Open source
Two phase 3D NA VOF 0.2 × 0.2 Water-Octane Unsteady state 1, 2, 3, 4
[53] code Gerris
Energies 2021, 14, 6066 26 of 56
Table 4. Cont.
SOFTWARE
Authors CFD Details Assumptions Findings
USED
Single/Two Dimension VOF/ Levelset
Grid Details Size of Microchannel System Steady/Unsteady
Phase (D) Method
984 < ρd< 1162
(kg/m3 ); 0.28 ×
Azarmanesh 10−3 < µd < 68 ×
Adaptive Mesh
and Farhadi, Two phase 3D VOF 0.1 × 0.1 10−3 kg/ms; 1240 Unsteady state Fluent 6.3 1, 2, 3, 4
refinement
[22] < ρc < 4130 kg/m3 ;
0.73 × 10−3 < µc <
3.46 × 10−3
0.57 to 7 micron
Kositanont
Two phase 2D size cells; 20,040 VOF 0.0956 Water-toluene Unsteady state CFX 5.6 1, 2, 3, 4
et al. [54]
cells
Unstructured
Modified 0.2 × 0.035 Inner tube;
Lan et al. [55] Two phase 3D mesh, 0.414 NA Unsteady state OpenFoam 1, 2, 3, 4
Levelset 0.4 outer tube
million cells
Main capillary width
= 0.4; Length = 10.8 Inhouse
Yang et al. [56] Two phase 2D 32,200 cells VOF Water-toluene Unsteady state 1, 2, 3, 4
vertical capillary = code
0.02; Length = 0.2
Densities and
DeMenech smallest cell 5 PhaseField
Two phase 3D 0.1 width viscosities Unsteady state Comsol 5.0 1, 2, 3, 4
et al. [57] micron Method
considered
Han and Chen, Levelset Ansys
Two phase 3D 58,000 cells 0.1 width; 1 length Water–Oil Unsteady state 1, 2, 3, 4
[58] Method Fluent 18.1
Sripadraja Open source
Two phase 2D 130,000 cells VOF 0.1 × 0.05 Oil-water Unsteady state 1, 2, 3, 4
et al. [59] code Gerris
* N.A.—Not Available. Findings: (1) Processes related to liquid-liquid slug flow can be studied numerically. Methodology presented for slug flow can capture slug flow generation, (2) The interface movement
and internal circulations were well predicted using numerical simulations, (3) The numerical model was tested under different flow conditions and fluid properties and found as a versatile tool for design of
micro-reactors, (4) Wall film provides lubricating action to the enclosed slug whereby the slug surrounded by the film moves with higher speed as compared to average velocity, (5) Slug length can be preserved
upto a channel diameter of 1 mm, (6) Wall adhesion had strong effect on slug/drop formation mechanism, (7) Predicted slug lengths were found to be almost independent of channel size and inlet distributor
type for a range of Capillary values in present work, (8) A thin liquid film was found in most of the simulations performed for capillary numbers higher than critical capillary numbers. However, for small values
dryout was observed. Limitations: (1) The present model is suitable for stationary interfaces and deformation of the interfaces is not considered, (2) The work is restricted to understanding flow patterns and slug
lengths while mass transfer model is not considered which is important for VMTC, (3) Geometries considered are very small. Practically, different geometries need to be considered which are used in applications.
Assumptions: (1) Flow is steady, (2) Flow is fully developed, (3) Flow is Newtonian 4 Flow is incompressible and miscible with each other.
Energies 2021, 14, 6066 27 of 56
A detailed review on the VOF methods is published by Rider and Kothe, [60]. This
method is capable of handling problems with significant interface topology change and
does not suffer from mass (or volume) losses. The extension of the method to 3D simu-
lations is straightforward and no specific algorithm is needed for the case of merging or
break up of the interface. However, the interface is smeared out and merges “numerically”.
The main drawback of this method is the inherent numerical smearing.
An alternative method is suggested by thegroup of researchers [61,62]. They used
a combination of both front tracking and front capturing methods, whereby a fixed grid
is used to describe the motion of the fluid flow and another moving grid with a low
dimension is used to track the interface. Since for both phases a fixed grid is used, they
are treated together, by solving a single set of governing equations for the whole flow
field. The complexity of this method is due to the necessary dynamic remeshing of the
moving sub-grid and mapping of the data coming from the moving grid to the fixed one.
Additionally, for the merging and breakage of the interface, a special sub-grid algorithm is
required. Most researchers have used the volume capturing (VOF) or surface capturing
(LS) methods with a combination of both. The equations and boundary conditions used for
VOF and Levelset equations along with continuity and momentum equations have been
given below.
Continuity Equation:
∇·v = 0 (1)
Momentum Equation:
∂u h n oi
+ ∇·(uu) = −∇ p + ∇· µ ∇·v + (∇·v)T + g + F (2)
∂t
where σ = surface tension, κn = surface curvature at the fluid interface computed from the
divergence of the unit surface normal.
F in the above equation is defined as:
" #
∅1 ρ1 + ∅2 ρ2
F = −σκn ρ +ρ (3)
1 1
2
1 n
κn = ∇·n̂ ·∇ |n| − (∇·n) (4)
|n| |n|
n
n̂ as n̂ = , n = ∇∅ = 0 (5)
|n|
VOF Equation:
∂f
+ ∇·(u f ) = 0 (6)
∂t
Level-Set Equation
∆∅
∂∅
+ ∇·(u∅) = γ∇ −∅(1 − ∅) + ε∆∅ (7)
∂t |∆∅|
The boundary conditions used for most of the research articles considered in the
present review are as follows:
1. Inlet- Velocity value
2. Wall–No slip Condition
3. Outlet–Constant Pressure
The mathematical equations used are as follows:
Inlet:
→
v = −vo n̂ (8)
Energies 2021, 14, 6066 28 of 56
The parametric sensitivity carried out by various authors in the subsections below:
investigated the mechanism of droplet breakup in squeezing, dripping and jetting regimes
using the PhaseField method. The breakup process was found to be driven chiefly by
buildup of pressure upstream of an emerging droplet. The capillary number was found to
have a negligible effect on droplet size and dynamics of breakup. The dripping regime was
influenced by the contrained geometry and modified the scaling law derived for size of
droplets from interfacial and viscous stress balance. The jetting regime was found to set at
high flow rates and higher values of Ca number.
The research works for CFD studies have considered flows in the slug flow regime.
A numerical model was developed to study the slug flow patterns under different flow
conditions and fluid properties Kashid et al. [14]. The authors validated their model with
the experimental results. The authors found that CFD can be a very efficient tool for design
of micro-extractors. Further, internal circulations inside the slugs (dispersed phase) as well
as outside the slugs (continuous phase in between two slugs) were studied.
VOF based CFD methodology to simulate slug flow in micro-channels was used by
some researchers Kashid et al. [44]. The authors considered two cases (i) when there is no
film between the wall and the slugs and (ii) when the slug is seperated from the wall by
a liquid film situations of slug flow were considered and validated. The authors found
that VOF method described slug flow generation and its characteristics very well. The slug
flow generation was found to depend on slug flow velocity and interfacial velocity. The
authors also concluded that slug length can be preserved upto a channel diameter of 1 mm.
3.1.1. Effect of Geometric, Operating Parameters, Wallfilms and Fluid Properties on VMTC
Kashid et al. [1] carried out studies in mass transfer systems for three different systems
1. Water-iodine-kerosene; 2. Kerosene-acetic-acid-water and 3. Water-succinic-acid-n-
butanol at constant flow velocities. The authors found that the VMTC’s varied from (0.32
to 0.98 s−1 ) for channel diameters of 500 µm while it dropped to (0.13–0.32 s−1 ) for channel
diameters of 1000 µm for water-iodine-kerosene system. The authors found that internal
circulations increase with flow velocity. Hence, VMTC increases due to higher internal
circulations. The authors, also, found that flow ratio (q) has a positive effect on VMTC. For
water-succinic-acid-n-butanol system the extraction efficiencies decreased by around 10%
with increase in slug flow velocity for constant values of internal diameter of the capillary.
Zhao et al. [16] have conducted experimental investigations with water/n-butanol/succinic
acid with micro-channels having internal diameter of 300–600 µm and length 45 mm. The
authors observed an increase of VMTC values to about 4 times when the channels size
was decreased from 600 µm to 300 µm. Further, this was observed for Re numbers in
the range of 500–600. Other researchers, [6,9] have operated their respective systems for
low Reynolds numbers (Re < 20) for toluene-trichloroacetic acid-water system for channel
diameter of 400 µm and found that the VMTC’s less than 1. Other researchers, Di Miceli
Raimondi et al. [70] have carried out measurements of VMTC’s in diameters of around
300 µm. The VMTC’s obtained by the authors were around 8.44 s−1 for 210 µm while it
decreased to around 2.65 s−1 for 300 µm micro-channels. The Re range in such systems
were less than 50.
Dessimoz et al. [9] have used T-shaped and Y-shaped micro-channels. The authors
used T-shaped micro-channel for generating slug flow while Y-shaped micro-channels
were used for generating parallel flow. The authors have used hexane-water system for
generation of slug flow while they added sodium hydroxide in the aqueous phase. Hexane-
trichloroacetic acid-water-sodium hydroxide system was used due to higher interfacial
forces responsible for the generation of slugs. Also addition of sodium hydroxide decreased
the viscous forces and interfacial forces further dominated the system due to which slug
flow could be formed. Xu et al. [69] carried out investigations on alkaline hydrolysis
reactions for T-junction for slug flow (similar to Dessimoz et al. [9] and found increase in
VMTC’s with increase in NAOH concentrations
Ghaini et al. [66] studied the effect of wall film on transport property. The authors
used the fast instantaneous reaction. Matsuoka et al. [73] studied the influence of wall
films during slug flow in micro-channels. The authors found the variation of VMTC with
Reynolds number for a fixed velocity and different hydraulic diameters (0.6 mm to 2 mm).
Maximum VMTC obtained was around 0.7 s−1 for a constant channel length. The VMTC
was found to be a strong function of slug flow velocity and decreased due to resistance due
to wall film.
Di Miceli Raimondi et al. [70] carried out measurements for dispersed phase VMTCs
using the water toluene system. The maximum inlet velocities measured were up to
300 mm/s and high values of transport properties are obtained.
Arsenjuk et al. [71] carried out experimental investigations for a kerosene-glycerol-
water system to see the effect of film thickness on transport property. The authors found
that film thicknesses were higher for systems having a transfer component than systems
not having transfer component. This influences the transport property. This study de-
serves special significance since the presence of transfer component affects two phase
investigations and affects the flow properties.
Energies 2021, 14, 6066 31 of 56
Sattari-Najafabadi et al. [74] have also performed investigations to study the effect
of shape of microchannels on the VMTC. Sattari-Najafabadi et al. [74] have found the
impact of geometric configuration i.e., change in channel width (W) and type of junction.
The authors used the T-junction, cross junction and cross T-junction to understand the
effect on VMTC. Further, the channel width (W) to depth (H) ratio (W/H) was varied
(1 ≤ W/H ≤ 2). The flow ratio (q) was varied from 0.5 to 2. The authors concluded that
best VMTC’s can be obtained for a cross-junction with flow ratio of 0.5.
Sattari-Najafabadi and Esfahany, [75] carried out experimental investigations to find
the effect of surfactant on VMTC during two phase flow of a LLE system. The authors
observed that VMTC obtained in presence of surfactant was 1.44 times better than VMTC
without surfactant.
Table 5. Literature of experimental works on flow patterns and its influence on transport properties.
Table 5. Cont.
Table 5. Cont.
Table 5. Cont.
Findings: 1. Liquid-liquid extraction is carried out in capillaries and VMTC’s were found to be two orders of magnitude higher than conventional equipment’s. 2. The improvement of mass transfer rates is due
to two phenomena. A. the large interfacial areas which enhances the mass transfer rates and b. the internal circulations inside the slugs which promote surface renewal at the phase interface. 3. The study showed
that capillary contactors can be a powerful tool for laboratory studies of liquid-liquid extraction 4. VMTC’s have a very strong dependence on parameters like slug flow velocity, flowrates of organic and aqueous
phases and capillary diameter. The VMTC’s are constant after a certain slug flow velocity is reached. 5. VMTC’s decreased with an increase in capillary diameter 6. Inlet locations play an important role in the
mass transfer process 7. Decrease in height/depth of the channel causes a increase in VMTC 8. Correlations of VMTC’s were developed 9. Found regime transition maps based on Capillary and Reynolds
numbers for parallel and slug flow systems 10. The authors found that the main parameter for the determination of VMTC’s for parallel flow was the initial concentration of water while for slug flow it was the
total linear velocity of fluids (dependent on total flow rate through the channel) 11. The authors focused on the wetting behavior at the capillary walls due to which the true interfacial area is different than the
Energies 2021, 14, 6066 36 of 56
one determined by the conventional snapshot method of determination. Hence, chemical methods need to be applied to such systems 12. Found the interfacial area for systems with and without wall films. Effect
of slug flow velocity on interfacial showed that for systems with film 13. An inert carrier gas can be used to enhance mass transfer under certain circumstances and to tune the separation behavior of a capillary
separator 14. Secondary flow patterns are developed due to the inert carrier gas in addition to the water slugs that exist in liquid-liquid flow which leads to an enhanced surface area between liquid phases 15.
Extraction efficiency and VMTC decreased with increased volume flow rate ratio, total flow velocity, and the inner diameter of the main channel. 16 Initial concentration of aqueous phase has an effect on
extraction efficiency. 17. VMTC decreases with channel length for an alkaline hydrolysis reaction for a micro-channel with slug flow 18. Overall extraction rate is a linear function of VMTC and reaction is mass
transfer controlled 19. Slug velocity is an important factor for enhancing the VMTC due to increase in internal circulations within the slugs 20 Determination of VMTC’s by experimental measurements and
validating a correlation obtained from numerical simulations 21 The flow patterns inside the droplets confined to the walls causes enhancement in mass transfer compared to droplets that are not confined 22.
Different flow patterns slug, annular, dispersed and inverted dispersed flow patterns were observed for experimental investigations of liquid-liquid extraction systems for three different flow orientations 23.
VMTC decreases with increase in capillary diameter 24. Vertical downflow orientation gives higher VMTC’s for both lower and higher diameter capillary than the other orientations 25. Suggested an innovative
co-current tubular contactor that is trumpet shaped with short horizontal section with long vertical up and downflow sections 26. Compared the performance of micro-channels with conventional extractors and
found that micro-channels have the potential to reduce solvent inventory and plant area due to miniaturization 27. The wall film in liquid–liquid slug flow is continuously renewed and thus actively participates
in mass transfer. Hence this should not be neglected while designing micro-reactors with liquid-liquid slug flow28. Experimental investigations in micro-channels with two different inlet junctions (T-type and
cross type) showed that droplet based microsystems provides higher VMTC’s between water and IL for micro-channel configurations 29. VMTC at droplet formation stage is 3–4 times higher than at the latter
stage due to strong disturbance inside droplet during formation 30. Found that under stable Taylor flow conditions in a 1 mm inner diameter glass channel, the organic phase (dodecane) formed liquid droplets
and the aqueous phase(water) formed liquid slugs with a liquid film enclosing the organic phase droplets 31. Compared the performance of micro-channels with batch extractors and found that micro-channels
can reduce the residence times to significant extent and make the processes continuous 32 Defined a criterion for parallelization for numbering up of several micro-channel units for higher feed rates 33 Defined a
universal flow map for flow patterns occurring in liquid-liquid fluid systems 34. Volume fractions of solvent played an important role in determining flow patterns and also extraction efficiency 35 Study of
hydrodynamics and mass transfer of micro-capillary with 0.6 mm internal diameter with a surface tension reducing agent (surfactant) revealed decrease in slug lengths resulting in enhancement in interfacial area
and corresponding increase in the VMTC. However, with increasing concentration of surfactant 36. The authors have developed a new correlation relating VMTC with Reynolds number, Capillary number and
SDS concentration 37 VMTC increases around 1.44 times by addition of surfactant 38. Studies were performed with rectangular micro-channel with variation of aspect ratio which showed that increasing the
aspect ratio gave increase in VMTCs due to enlargement of the dispersed phase slugs and reduction of wall film renewal 39. Studies with T-junctions and cross junctions for mixing was also carried out where
T-junctions showed improvement in VMTC’s 40 Extraction of rare earth metals can be done in micro-channels with low residence times (less than 20 s) 41 Comparison between conventional continuous and batch
processes and micro-channels shows that VMTC’s are increased by an order or two orders of magnitudes by using micro-channels and hence extraction efficiencies are better than conventional extractors 42. The
study was carried out to extract samarium (III) from hydrochloric acid using a T-mixer and rectangular channel. 43. Effect of pH and width on slug length was observed. Slug length increased 44. Increase in flow
rate increases VMTC’s while increase in width and length decreases VMTC’s. 45 Extraction of Uranium from leach solution is much better in microchannels with efficiencies over 90%. 46 Variation in temperature
as a variable has been considered during measurements and effect on flow patterns and mass transfer behavior have been studied 47. Studies were carried out with straight channels and curved bends and it was
found that VMTC increased with increase in viscosity of the continuous phase at large flow rates. Limitations: 1. No numerical or CFD studies have been done and validated with experimental data 2. No
mathematical model was developed for prediction of VMTC’s 3. The research is restricted to T-junction micro-channels and influence of geometric shapes on VMTC’s are not considered 4. Numerical simulations
have not been carried out to compare the flow patterns 5. The article does not consider the variation of VMTC due to change in channel diameter and type of inlet mixing (restricted to T-junction inlet). 6. The
article is restricted to slug flow patterns in microchannels 7. The article is restricted to experimental investigation of effect of wall film on mass transfer between continuous and dispersed phases 8. The article is
restricted to droplet flow regime and droplet formation only 9. The study does not consider the effect of wall film between the slugs which might affect the VMTC 10. The present article is restricted to prediction
of flow patterns while measurement of VMTC’s for these systems and operating parameters were not performed. 12. Pressure drop studies when the VMTC’s were increased due to flow rates are not reported in
the present study. Assumptions: 1 Outlet flow from steel outlet would be higher than the one from PTFE outlet due to higher affinity of aqueous phase for steel than affinity for organic phase of PTFE 2 The
combination of a flow splitter and capillary has been assumed to act as a mixer-settler device as in conventional liquid-liquid systems 3 Properties of two immiscible liquids are constant for all the experiments 4
For cases where there is no formation of wall film, the slugs were considered to be of rectangular shape for calculation of interfacial area 5 Fluid is Newtonian and incompressible.
Energies 2021, 14, 6066 37 of 56
coupled equations have been presented. They are a. Moving boundary method and b.
Front capturing method. Table 6 provides the geometric details and operating parameters
of the research works while Table 7 provides details of CFD softwares, mesh, methods used
for the research works considered.
A typical schematic of the moving boundary approach has been shown in Figure 14.
The velocity of the moving wall is same as the slug velocity. Equal stress condition across
the interface is considered for the momentum. Convection diffusion equations are solved
for each phase. No flux boundary conditions is implemented across moving walls while
the flux continuity and thermodynamic equilibrium boundary conditions are implemented
with the concentration equations.
Figure 14. Schematic of moving boundary approach domain for CFD simulations taken from Zhang et al. [86].
To evaluate the VMTC, the simulation process decoupled the hydrodynamics of two-
phase flow from the mass transfer phenomena and was then divided into two steps. First,
the steady-state flow field was established, considering the velocity to be time-invariant.
Then, all the components in the solution were initialized in the organic and aqueous
phases, respectively, and the flow time was reset to zero. The mass transfer equations were
calculated with a time step of 0.001 s and 10 iterations at each time step by analysing data
obtained with different time steps and iterations. For the initial conditions, the flow field
at t = 0 s calculated in the first step of the computational procedure is the steady-state
flow field. Since species 1 is only patched to a certain concentration value in phase 1, the
appropriate initial and final boundary conditions of the concentration field are provided.
Many industrial separation processes occur in fluid-fluid systems including droplets/
bubbles in a continuous phase (e.g., solvent extraction, gas scrubbing, waste water treat-
ment in bio reactors). For these processes interfacial mass transfer represents a crucial
phenomenon. To describe it properly, a coupled problem including momentum and mass
transport at and around moving interfaces must be solved. Such problems are very dif-
ficult, since the deformation of the interface influences interfacial mass transfer related
to the handling of boundary conditions at the moving interface. Due to the assumption
of the thermodynamic equilibrium, a concentration jump appears at the interface which
is numerically demanding and difficult to implement into the front capturing methods.
Besides, the component flux continuity should also be fulfilled at the interface. There are
just few papers dedicated to this problem.
Bothe et al. [87] used the VOF method to model a gas-liquid system (air bubbles rising
in water). They transformed the mass transfer equation to
∂C
e →
+ u ·∇Ce = D ∇Ce (12)
∂t
Energies 2021, 14, 6066 39 of 56
Table 6. Literature review of research work for coupled hydrodynamics and mass transfer in microchannels: Experimental details.
Dimensionless
Parameters Output
Authors Channel Details Type of Study VMTC, 1/s Number and
Varied Parameters
Range
Channel Channel
Geometry Flow Experimental/ Flow
Diame- Length System
Type Rate/Velocity CFD Regime
ter (mm)
Reactive system:
Flow patterns,
Kerosene-acetic acid-water; time, flow
Kashid Micro- 3 mm/s to VMTC,
0.5 × 0.5 2 CFD simulation Non-reactive system: Slug flow speed, slug 0.12–0.2 s−1 NA
et al. [65] channel 16 mm/s concentration
Water-succinic length
profiles
acid-nbutanol
Inlet velocity;
Gómez-
Y- 0.2; 0.04; 1 mm/s to Experimental Flourescin sodium salt in width to Seperation factor, Low Reynolds
Pastora 2 Stratified flow NA
microchannels 0.16 20.1 mm/s and CFD water-deionized water height ratio, flow patterns number
et al. [88]
phase ratio
Capillary
Di Miceli Square Reynolds VMTC,
0.06– Non-reactive system: number,
Raimondi micro- 100 CFD simulation Slug flow number, concentration 0.05–10 s−1
0.96 Water-nbutanol Reynolds
et al. [89] channel residence time profiles
number
Flow patterns, Reynolds
Rectangular
Cito et al. Sherwood number, number, 1–500;
micro- <1 10 CFD simulation Fluid with SC = 1000, water Slug flow Time
[90] wall mass transfer Sherwood
channel
rate number
Tsaoulidis
10,000– Nitric Extraction
and Micro- 0.117–5.67 Experimental and Residence time,
0.5–2 31,500 acid-UO2-TBP Slug flow efficiency, 0.049–0.29 s−1
Angeli capillary mL/min Numerical mixture velocity,
mm (IL) VMTC
[82]
Enhancement
factor,
Reactive system:
Zhang Circular Time, vertical concentration Hatta number
0.6 2.22 CFD simulation Water (NaOH)-aceticacid- Slug flow 0.1–0.3 s−1
et al. [86] capillary distance profiles, flow (0.32)
nbutanol
patterns, VMTC,
velocity profiles
Viscosity ratio, density
Ramji and ratio, diffusivity ratios; Viscosity ratio; Flow patterns, Reynolds
Rectangular
Puspa- water-acetone-toluene Reynolds velocity, number, 1–100,
microchan- NA NA NA CFD simulation Slug flow NA
vanam, system for model number; slug concentration Peclet number
nel
[91] validation with holdup profiles 1–5000
experiments in literature
Energies 2021, 14, 6066 40 of 56
Table 7. Literature review of research work for hydrodynamics and mass transfer in micro-channels: CFD Details.
Table 7. Cont.
Software
Authors CFD Details Assumptions Limitations Findings
Used
Unsteady state
Mass transfer governing
Gómez- FLOW 3D
only due to equations with
Pastora et al. Two phase 2D 400 × 100 VOF commercial 1 1 13, 14, 15
concentration convection-
[88] software
gradient diffusion
equations
Dimensionless In-house code
Unsteady state for solving
Ramji and Dimensionless governing dimensionless
Puspavanam, Two phase 2D 202 × 101 NA Transient equations with governing 1, 4, 5, 6, 7 1 16, 17
[91] Mass transfer convection- equations and
diffusion interface
equations movement
Findings: 1 The results showed that circulations inside and outside slugs keep on renewing the concentration at the interface and increase the concentration gradient between two slugs which enhances mass
transfer 2 The authors claimed that the model can be applied to any liquid-liquid system with fixed interfacial location 3 Found different flow structures as a function of Reynolds and Capillary numbers in a
square micro-channel 4 An empirical correlation was developed relating the VMTC as a function of Reynolds number and Capillary number 5 Wall mass transfer rates for fast first order chemical reactions which
occur on the wall can be predicted by this model 6 Recirculation’s in the bulk cause enhancement in mass transfer rates near the walls 7 Correlation is developed to predict VMTC as function of Re, contact angle
and time 8 The authors found that the dependence of the VMTC’s on Reynolds numbers agreed with typical laminar boundary layers and its dependence on time corresponded to unsteady mass transfer
diffusion problems 9 The extraction efficiency of the system under consideration decreases with decrease in mixture velocity and increase in internal diameter 10 For alkaline hydrolysis reactions the VMTC’s
show a sharp increase at the start of the process while reaching constant values after a certain time 11 Diffusivity of the fluid has an effect on the VMTC and increases when diffusivity of aqueous phase increases
12 Mass transfer process is intensified by the hydrolysis reaction with increase in velocity of the fluids which is depicted by the increase in enhancement factor values 13 Solute mass transfer in Y-Y micro-channels
has a coupled influence of both flow patterns and mass transfer kinetics 14 Applications where low organic to aqueous phase ratios and high separation and high solute flows are required can be achieved by a
specific geometrical configuration while those where ratios of diffusion times and residence times are high can be achieved by a different configuration 15 The rate of molecular diffusion determined for the first
time both by experimental and numerical approaches for asymmetric and symmetric devices showed that it was a strong function of ratio of diffusion to residence times 16 Rate of mass transfer can be enhanced
if viscosity of continuous phase is higher than dispersed phase 17 Fluid properties and flow parameters have a significant effect on mass transfer in slug flow when there is a thin film surrounding the slug.
Limitations: 1 The interface in the model is fixed and hence deformations occurring across interface during mass transfer could not be taken into account. 2 Approximation of slug movement is considered
using the moving boundary approach 3 Outcome of the simulations had no mention of values of VMTC 4 No free surface model was used to tract the deformation of the interface 5 The Sherwood number for
rectangular and circular micro-channels has been considered similar 6 Species transport equations have been used for mass transfer 7 VMTC’s obtained from numerical experiments were not validated with
experimental results. Assumptions: 1 Both liquids are Newtonian, viscous and incompressible 2 No effect of mass transfer and chemical reaction on shape and volume of slug 3 Flow is laminar and mass
diffusivity constant in both liquids considered in the domain for numerical simulations 4 The flow is two dimensional 5 Flow is fully developed and liquids are immiscible with each other 6 Liquid-liquid
interface is assumed to be flat for the sake of simplicity 7 Marangoni stresses at the interface are neglected and flow field is assumed to be independent of concentration field and the interfacial flux continuity
boundary condition to
Energies 2021, 14, 6066 42 of 56
Dl ∇ C
el ·n = HDg ∇C
eg ·n (13)
→
where n is the unit vector normal to the interface, C is concentration, u is velocity, D is
molecular diffusivity and
Cl
C=
e (14)
Cg /H
where indices l and g denote gas and liquid phases and H is the distribution coefficient
(Henry’s constant).
Yang and Mao, [92] carried out CFD simulations using the LS method for a rising
droplet in a continuous liquid phase (water–succinic acid–butanol). The equation for mass
transfer was modified in their work and is given below
∂Ĉ
+ û·∇Ĉ = ∇· D̂ ∇Ĉ (15)
∂t
where √
Ht i f ∅ ≥ 0
√
t̂(∅) = (16)
t/ H i f ∅ ≤ 0
√ √ √
D̂ (∅) = HD2 + D1 / H − HD2 Hε (∅) (17)
√ √ √
û(∅) = Hu + u/ H − Hu Hε (∅) (18)
where
Cl
Ĉ = (19)
Cg /H
Here, H is the distribution for the liquid-liquid system considered and Hε (∅) is the
regularized Heaviside function to avoid the numerical instabilities at the interface. It is
given by
0 i f ∅ < −ε
1
Hε (∅) = 1 + ε + sin(π ∅/ε)/π i f |∅| ≤ ε
∅ (20)
2
1i f ∅ > ε
Haroun et al. [93] suggested a VOF-based modelling approach for reactive absorption
systems. They considered the film flow of a liquid flowing along a corrugated surface
in counter-current with the laminar gas stream. The mass transport equation has been
modified to: →
∂Ci
+ ∇· u Ci = ∇·( Di ∇Ci + Φi ) + Wi (21)
∂t
where !
Cj 1 − Hj
Φi = D j ∇f (22)
f + H j (1 − f )
where, subscript c denotes continuous phase while subscript d denotes dispersed phase,
HD denotes the distribution coefficient, α1 , α2 are parameters, n is the unit vector normal to
the interface. The values of α1 , α2 are adjusted in such a way that the boundary conditions
are solved only at the interface. The authors have considered a droplet rising in an aqueous
phase in a water-acetic-acid-toluene system and mass transfer from toluene to water.
Two phase simulations with and without mass transfer have been carried out. CFD
simulations using a mass transfer model that predicts mass transfer across the interface
were carried out using COMSOL 3.5a. The model has been validated with liquid-liquid
and gas-liquid systems with satisfactory qualitative results.
Following are the boundary conditions for moving boundary approach:
1. Axial and radial velocities of the walls are equal to slug/plug velocity
2. No slip boundary condition at channel wall
3. Equal stresses at the interfaces between two phases
4. Flux continuity boundary condition at the interface
5. Zero flux set at the channel wall
6. Periodic boundary condition to satisfy flux continuity at the interface.
In Section 2.2, CFD studies were discussed in which governing equations for mass
and momentum were solved with interface tracking techniques like VOF. However, the
most important parameter for transport property (mass transfer) is the VMTC and hence
mass transfer from dispersed phase to continuous phase should be obtained by solving
concentration governing equations. However, boundary conditions of thermodynamic
equilibrium and flux continuity as given by Equations (2) and (3) should also need to be
incorporated while the system of equations is being solved. The approaches mentioned in
the previous subsection have been used by a few researchers and will be discussed in this
section. Details of investigations are provided in Tables 6 and 7.
Kashid et al. [65] performed simulations with the model provided by Yang and
Mao [92] and used the VOF method for defining the interface. The authors observed
internal circulations that keep on renewing at interface and increase the concentration
gradient between two slugs which enhances the rate of mass transfer. The authors claimed
that the model can be used for any liquid-liquid system with a fixed interface location.
No free surface model equations were used for tracking the interface. To summarize, the
authors have used the concept of moving boundaries for their numerical investigations.
Di Miceli Raimondi [89] have solved two dimensional simulations using their own
inhouse code. The simulations were carried out in two steps. First, the hydrodynamic
behaviour of the system is computed by means of the code. In the second step the con-
centration field of this solute present in the dispersed phase is computed by a code in
Matlab 7.4. The numerical method used by the authors is an interface capturing technique
without any interface reconstruction. The authors observed three flow structures (a) in-
ternal vortices in both continuous and dispersed phases and thin film that wets the wall,
(b) several recirculation nodes inside droplets and (c) without any recirculation loops in
continuous phase since droplet is not confined enough. The flow structures were found to
be a function of Reynolds and capillary numbers. Several simulations were performed for
different dispersed phase velocity, channel width, dispersed phase holdup. A correlation
has been developed for the computation of VMTC based on the data. The authors have
compared the correlation with other correlations [95,96] and the predictions were in good
agreement.
Zhang et al. [86] performed CFD simulations to predict mass transfer characteristics
with and without alkaline hydrolysis reactions during slug flow in micro-channels. The
species concentration equation was used for computing the concentration. The moving
boundaries approach was used. The predicted VMTC values were compared with ex-
perimental values and a close agreement was observed. The authors claimed that the
model could be applied to various liquid-liquid systems with fixed interfaces. The authors
also claimed that this model can be coupled with free surface models to capture complex
hydrodynamics near the interface.
Energies 2021, 14, 6066 44 of 56
Cito et al. [90] studied wall mass transfer rates generated by capillary driven flow
in a micro-channel. Simulations were performed for 2D rectangular and circular micro-
channels. The effect of Reynolds number, channel geometry and contact angle on average
wall mass transfer rates were observed. A correlation for Sherwood number Sherwood
and Wei [97] relating the above parameters was developed from the simulated data. The
effect of wall curvature was considered negligible in the present work due to high value of
Schmidt number. For Re < 500 the Sherwood number had a weak dependence on contact
time. The dependence on time corresponds to unsteady mass diffusion problems.
Tsaoulidis and Angeli, [82] carried out experimental investigations for extraction of Ura-
nium in straight microchannels. The internal diameter was varied from (0.5 ≤ D ≤ 2 mm). The
authors found the VMTC in small channels was in the range 0.049 to 0.29 s−1 . A correlation
for VMTC as a function of Re, Ca and (Internal dia (ID)/length of channel (D/L)) was
developed and predictions were compared with the experimental measurements as well as
other models available in the literature. The authors also investigated the effect of input
parameters like internal diameter and mixture velocity on efficiency and found that output
parameters like efficiency decreased with increase in ID. The authors have also carried out
numerical simulations with commercial software COMSOL 4.3. The domain taken by the
authors is a dispersed elongated structure which the authors mention as plug (instead of
the general definition of slug by previous authors) while the surrounding continuous phase
fluid is mentioned as the slug. Momentum and continuity equations were solved for this
domain along with convection-diffusion equations. The interface is considered stationary
and a moving wall boundary condition has been provided. The numerical results for the
velocity fields were compared with velocity patterns from Particle Image Velocimetry (PIV).
Images captured by High Speed camera were used to calculate output parameters like the
film thickness, length of unit cell, plug length, plug velocity. The data generated from the
images were used as boundary conditions of numerical simulations. Extraction efficiencies
were compared with numerical results and a good agreement was obtained.
Gómez-Pastora et al. [88] performed experimental and numerical investigations of
solute mass transfer in Y-Y micro-channels as a function of coupled influence of both flow
patterns and mass transfer kinetics. The influence of ratio between residence and diffusion
times and volumetric ratio between fluid phases was determined for 3 geometric variations
Symmetric device: Y junction inlet and outlet with equal inlet and outlet diameters for
phase 1 and phase 2. Configuration 1: Y junction inlet and outlet with inlet and outlet
diameter of phase 2 four times higher than the ones of phase 1 3. Configuration 2: Y
junction inlet and outlet with inlet and outlet diameter of phase 1 four times higher than
the ones of phase 2. Figure 15 shows the schematic of the micro-channels used by the
authors. Two factors (a) concentration factor (defined as the ratio of outlet concentration
in phase 2 to outlet concentration of solute in phase 1) and separation factor (similar to
conversion and defined as ratio of difference between inlet and outlet concentration of
phase 1 to inlet concentration of phase 1) were used to quantify performance. The authors
found that separation factor influenced the values of solute concentration in both phases
which influences mass flow of solute and channel outlet. Two configuration of Y-Y channels
were suggested for different applications. The authors recommended that systems with
separation factor less than 1 can be used for applications where high solute flows are
required (configuration 1 by authors can be used) while separation factors greater than 1
can be used for processes where high concentration factors are needed (configuration 2 can
be used).
Energies 2021, 14, 6066 45 of 56
Figure 15. (a) Symmetric device (b) Configuration 1: Asymmetric device H2 = 4H1 (c) Configuration
2: Asymmetric device H1 = 4H2 . Reproduced with permission from Gomez-Pastora et al. [88].
less numbers (Ca and We) were proposed by the authors. Dai et al. [100] also carried out
experimental investigations in a vertical 2 mm circular microchannel under constant wall
heat flux boundary conditions in an Re range 15 < Re < 48. The authors found an empirical
correlation for the same. Abdollahi et al. [101] carried out both experimental and numerical
investigations in microchannels with hydraulic diameter of 2 mm, a oil-water system for
constant heat flux boundary conditions. The authors determined correlations for friction
factor and heat transfer coefficients and found an increase of 700% in heat transfer rates
over single phase flow.
4.2. CFD Studies of Heat Transfer in Microchannels Involving Liquid-Liquid Two Phase Flow
A few numerical investigations have also been performed to study two phase flow.
Vivekanand and Raju [102] carried out 2D axisymmetric numerical investigations in 100 µm
channels using the VOF method. The investigation was carried out for both constant wall
heat flux and constant wall boundary conditions. The authors observed increase in Nusselt
(Nu) number upto 180% and 210% increase in two phase as compared to single phase.
Urbant et al. [103] have found increase in heat transfer coefficients for water droplets in
oil in circular microchannels of diameter 0.1 µm. The recirculations were seen in both
phases resulting in the increase in HTC. Fischer et al. [104] have observed 400% increase in
heat transfer rates for a Water-Silicon Oil system as compared to single phase fluids using
2D axisymmetric geometry of 0.1–1 mm diameter microcapillaries. Che et al. [105] did
simulations with 2D geometry for microchannels and found that heat transfer increases
with decrease in length of the slugs. Che et al. [106] carried out 3D large eddy simulations
(LES) for 0.2–0.8 mm (characteristic length) microchannels for water-mineral oil system.
An increase in Nu number was observed due to the two phase flow conditions due to
recirculations present in both continuous and dispersed phase. Bandara et al. [107] carried
out CFD simulations to investigate heat transfer characteristics (4.9 < Re < 21.9). The
authors concluded that the heat transfer rates depended on the Ca numbers, slug sizes, film
thickness and contact angles. The authors used correlations available in the literature (for
GL flows) to find the Nu number and concluded that correlations needed to be developed
specifically for LL flows. Presently two correlations for Nu number (one by numerical
studies by Dai et al. [100] and other by experimental studies by Giolla Eain et al. [99] exist
for LL flows. The equations are given in Table 8.
4.3. Other Aspects: Film Thickness, Peclet Number Effect, Systems Considered, Scope for
Different Geometries
During heat transfer in two phase LL flows a thin film is formed between the slugs
and the walls when flows are in slug flow regime. Techniques used in the literature consist
of Micro-PIV [98] and optical macroscopy [99]. The heat transfer rate has an inverse
relationship with film thickness and is decreased with an increase in film thikness. Most
authors have used in the past the correlations of GL flows in LL flows to determine film
thickness until Giolla Eain et al. [99] found a correlation for two phase flows based on
their experimental data. The authors defined film thickness in terms of Ca number and
We number
δ
= 0.35(Ca)0.35 (We)0.1 (25)
R
In case of rectangular channels, heat transfer from the wall to center of liquid slug
is further due to the space between the curvature of the liquid slug and the corners of
the rectangular channel since heat transfer in this region occurs due to conduction only.
Further, the heat transfer in LL flows will also depend on increase or decrease of flow
which will alter the size of the slugs and hence increase heat transfer rates.
Another important parameter affecting heat transfer is the Peclet (Pe) number which
is the ratio of heat transfer by convection to that by diffusion. Che et al. [105] in their
numerical investigations found that increasing Pe number increases the Nu number but
decreases the heat transfer process.
Energies 2021, 14, 6066 47 of 56
Authors Correlation
1
Nu TP =
2 2
1 LUC m 1 kC LUC 1 1
NuW + LD m +1 Nu FD kD + LS m +1 Nu FS
1.5
NuW = 5 + 2
Dai et al. [100] Ca 3
Nu FD = 4.364 + 0.0894
1 , L∗D = ReLDD Pr
/d
D
∗ ∗
LD +0.049LD 3
0.171 ∗ Ls /d
Nu FS = 4.364 + 1 , LS = ReTP PrC
L∗S +0.0663L∗S 3
4 4 1/4
Giolla Eain et al. [99] Nu FS = Nu x(S,Ent) + Nu x(S,Dev)
Fluid properties and size and shape of the microchannels would play a vital role
in heat transfer in LL flows. The literature review shows that mostly Oil-water systems
have been used for both experimental and numerical studies. Shape of the microchannel
would have an effect on the heat transfer process as seen in the description of rectangular
channels. Further, 3D simulations would give more opportunities for studying the shape
of the channels numerically.
5. Conclusions
The conclusions are summarized in Figure 16. The brief description is provided as
follows:
1. Flow patterns namely parallel, slug, annular, annular-slug, slug-dispersed have been
reported by authors but some new patterns like rivulet and serpentine patterns have
also been observed for liquid-liquid systems in main channels of microchannels
2. Flow patterns in junctions (cross or T-junctions) are also summarized as tubing,
dripping and jetting. Several works are available but a connection between the flow
patterns in the junctions and the main channels needs to be made while creation of
regime maps
3. Regime maps have been presented by authors with dimensionless numbers like
product of Reynolds numbers and Capillary numbers or product of Weber and
Ohnesorge numbers. However, there is no consensus on the particular regime maps
and no regime map is able to predict the works of the entire literature correctly
4. Pressure drop studies are in the form of experimental investigations and analytical
models. The pressure drop models are seen to predict well with the experimental
data for similar viscosity systems
5. Few CFD models focusing on flow patterns have been found for liquid-liquid sys-
tems. The models have shown good predictive capabilities with experimental results.
However, Capillary number plays an important role in the predictions. For certain
Capillary numbers the CFD models have been unable to predict the flow patterns as
those of experimental observations suggesting more amount of research works to be
carried out in this area.
6. Mass transfer studies in experimentation have shown the effect of operating variables
and properties and geometric dimensions. VMTC is found to increase with an increase
in flow velocity, presence of inerts, presence of surfactants, increase in flow ratio.
VMTC however decreases with increase in characteristic channel dimension (like
channel width or diameter) for specific channel length.
7. For cases of extraction of rare earths or uranium, the residence times are found to
be less than 20 s which is an order of magnitude less than conventional batch or
continuous processes
8. The VMTC’s for test systems involving conventional solvents or systems involving
IL’s or DES’s are two orders of magnitude higher than the VMTC’s of conventional
Energies 2021, 14, 6066 48 of 56
Figure 16. Summary of conclusions of the review (A) Experimental studies, (B) CFD studies.
6. Future Scope
1. Several works have been focused on understanding the flow patterns. Few researchers
have focussed on flow patterns at the junctions while many of them have focused
on the main channels only. In main channels, few researchers observed new flow
patterns which are different than the obvious flow patterns of parallel, slug, annular
and dispersed flow or a combination of these. These patterns need to be studied more
extensively by research fraternity similar to the work of researchers Yagodnitsyna
et al. [7], needs to be taken up by more researchers. Numerical simulations for these
cases needs to be carried out which for two phase flows for various systems with
different fluid properties and operating parameters and such models need to be
validated with the rich experimental data available in the literature both in junctions
and main channels
2. Commercial applications need to be found where LLE can be carried out in micro-
channels. The regime in which such operations would be carried out will pave
the way for efficient equipment design. Presently, from the reviewed literature,
extraction of rare earth metals seems to be the most promising application that can be
commercialized in micro-channels.
3. There have been very extensive and comprehensive studies for predictions of regime
maps by some researchers. A few researchers, claim universal regime maps. However,
the individual investigations in literature (for reactive and non-reactive systems) for
micro-channels does not quite comply with the flow regime maps except the regime
Energies 2021, 14, 6066 49 of 56
map by Zhang et al. [12]. This might be due to the use of different material of con-
structions for experimental investigations. Further, several authors claim combination
of different dimensional numbers. One of the recommendations would be to form a
consortium of experts for development of comprehensive regime maps which can be
used for any system which currently seems to be a challenge. Numerical simulations
with the power of volume capturing methods would play a vital role in preparing
such universal maps
4. Surprisingly, most of the experimental works compare the VMTC’s in micro-channels
with conventional equipment’s and the VMTCs are two orders of magnitude higher
than ones in conventional extractors. But inspite of such various advantages of mi-
crochannels, commercial application of micro-channels are not reported in literature.
Gaps in commercialization of micro-channel extraction units need to be found and
research work needs to be taken up in that direction
5. The advantages of CFD simulations in understanding flow patterns and mass trans-
fer in micro-channels have not be harnessed to its full potential. More number of
works of CFD with green solvents like ionic-liquids and Deep Eutetic Solvents (DES)
need to be taken up and validated against the experimental results both in terms of
hydrodynamics and mass transfer
6. Mass transfer models need to be used for predictions for VMTC’s with micro-channels
as well for the various systems that have been discussed in the present article (both
experimental and two phase CFD simulations). Surprisingly, mass transfer models
have not been used extensively for predictions of interfacial area and VMTC’s. Soft-
wares like Ansys Fluent/Comsol/in-house-code have been used by authors who
have developed CFD models. The surface is however considered constant in most of
the cases. Movement of the surface is important to capture the Marangoni instabilities
which enhances mass transfer.
7. Very few works in terms of coupled equations of front tracking methods (Level-
Set/VOF) with mass transfer models with Navier-Stokes equations exist in the lit-
erature for drops or bubbles moving in continuous liquids or liquid films on solid
surfaces. However, in these models the distribution coefficient is used as a transforma-
tion parameter and thus assumed to be constant. However, in reality, the distribution
coefficient depends on state variables and suffers significant changes. Hence, ro-
bust mass transfer models also needs to be developed before they can be applied
to microchannels for prediction of VMTC’s and interfacial areas. Ideally, a compre-
hensive mathematical model taking account of the interface deformation, interfacial
concentration jump and variable distribution coefficient needs to be developed.
8. More test systems need to be studied for understanding heat tranfer rates in liquid-
liquid systems with a wide range of sizes
9. Higher range of Re numbers need to be considered (100 < Re < 1000) for understanding
the effect of various parameters on heat transfer rates.
Author Contributions: Conceptualization, A.A.G. and A.B.P.; methodology, A.A.G.; software, None;
validation, None; formal analysis, A.A.G.; investigation, A.A.G.; resources, A.A.G.; data curation,
A.A.G.; writing—original draft preparation, A.A.G.; writing—review and editing, A.A.G.; visual-
ization, A.A.G.; supervision, A.B.P.; project administration, None; funding acquisition, A.A.G. All
authors have read and agreed to the published version of the manuscript.
Funding: This project received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Since this is review article the data is taken from the literature.
Conflicts of Interest: The authors declare no conflict of interest.
Energies 2021, 14, 6066 50 of 56
Abbreviations
l Liquid
L MTC
M Mean
Oil Oil
out Outlet
R Residence time
s Slug
sat Saturated
t Velocity
TP Two phase
UC Unit cell
W Wall
water Water
Greek Symbols
α Water phase fraction, -
β Dispersed phase fraction, -
δ Film thickness, m
εL Interface thickness in level-set equation, -
ε0 Constant in pressure drop equation
∅ Level-set function
λ Constant for correlation in Table 7, -
µc Viscosity of the continuous phase
µd Viscosity of the dispersed phase
η Variable in pressure drop equation
ρ Density
σ Surface tension
ϕg Gas phase hold up in Table 7, -
τ Contact time of aqueous and organic phases in micro-channel
References
1. Kashid, M.N.; Harshe, Y.M.; Agar, D.W. Liquid−liquid slug flow in a capillary: An alternative to suspended drop or film
contactors. Ind. Eng. Chem. Res. 2007, 46, 8420–8430. [CrossRef]
2. Cademartori, S.; Cravero, C.; Martini, M.; Marsano, D. CFD Simulation of the Slot Jet Impingement Heat Transfer Process and
application to a temperature control system for galvanizing line of metal band. Appl. Sci. 2021, 11, 1149. [CrossRef]
3. Cravero, C.; Domenico, D.; Leutcha, P.J.; Marsano, D. Strategies for the Numerical Modeling of Regenerative Preheating systems
for recycled glass raw material. Math. Model. Eng. Probl. 2019, 6, 324–332. [CrossRef]
4. Dixit, T.; Ghosh, I. Review of micro-and mini-channel heat sinks and heat exchangers for single phase fluids. Renew. Sustain.
Energy Rev. 2015, 41, 1298–1311. [CrossRef]
5. Al-Azzawi, M.; Mjalli, F.S.; Husain, A.; Al-Dahhan, M. A Review on the hydrodynamics of the Liquid–Liquid two-phase flow in
the microchannels. Ind. Eng. Chem. Res. 2021, 60, 5049–5075. [CrossRef]
6. Kashid, M.N.; Renken, A.; KiwiMinsker, L. Gas–liquid and liquid–liquid mass transfer in microstructured reactors. Chem. Eng.
Sci. 2011, 66, 3876–3897. [CrossRef]
7. Yagodnitsyna, A.A.; Kovalev, A.V.; Bilsky, A.V. Flow patterns of immiscible liquid-liquid flow in a rectangular microchannel with
T-junction. Chem. Eng. J. 2016, 303, 547–554. [CrossRef]
8. Kovalev, A.V.; Yagodnitsyna, A.A.; Bilsky, A.V. Flow hydrodynamics of immiscible liquids with low viscosity ratio in a rectangular
microchannel with T-junction. Chem. Eng. J. 2018, 352, 120–132. [CrossRef]
9. Dessimoz, A.L.; Cavin, L.; Renken, A.; Kiwi-Minsker, L. Liquid–liquid two-phase flow patterns and mass transfer characteristics
in rectangular glass microreactors. Chem. Eng. Sci. 2008, 63, 4035–4044. [CrossRef]
10. Bai, L.; Zhao, S.; Fu, Y.; Cheng, Y. Experimental study of mass transfer in water/ionic liquid microdroplet systems using micro-LIF
technique. Chem. Eng. J. 2016, 298, 281–290. [CrossRef]
11. Wu, Z.; Cao, Z.; Sunden, B. Flow patterns and slug scaling of liquid-liquid flow in square microchannels. Int. J. Mult. Flow. 2019,
112, 27–39. [CrossRef]
12. Zhang, Q.; Liu, H.; Zhao, S.; Yao, C.; Chen, G. Hydrodynamics and mass transfer characteristics of liquid–liquid slug flow in
microchannels: The effects of temperature, fluid properties and channel size. Chem. Eng. J. 2019, 358, 794–805. [CrossRef]
13. Kashid, M.N.; Gerlach, I.; Goetz, S.; Franzke, J.; Acker, J.F.; Platte, F.; Agar, D.W.; Turek, S. Internal circulation within the liquid
slugs of liquid-liquid slug flow capillary micro-reactor. Ind. Eng. Chem. Res. 2005, 44, 5003–5010. [CrossRef]
14. Kashid, M.N.; Rivas, D.F.; Agar, D.W.; Turek, S. On the hydrodynamics of liquid–liquid slug flow capillary microreactors. Asia-Pac.
J. Chem. Eng. 2008, 3, 151–160. [CrossRef]
Energies 2021, 14, 6066 53 of 56
15. Cao, Z.; Wu, Z.; Sattari-Najafabadi, M.; Sunden, B. Liquid-liquid flow patterns in micro-channels. In Heat Transfer Summer
Conference; American Society of Mechanical Engineers: Bellevue, WA, USA, 2017.
16. Zhao, Y.; Chen, G.; Yuan, Q. Liquid–liquid two-phase mass transfer in the T-junction microchannels. AlChE J. 2007, 53, 3042–3053.
[CrossRef]
17. Cubaud, T.; Mason, T.G. Capillary threads and viscous droplets in square microchannels. Phys. Fluids 2008, 20, 053302. [CrossRef]
18. Fu, T.; Wu, Y.; Ma, Y.; Li, H.Z. Droplet formation and breakup dynamics in microfluidic flow-focusing devices: From dripping to
jetting. Chem. Eng. Sci. 2012, 84, 207–217. [CrossRef]
19. Carrier, O.; Funfschilling, D.; Li, H.Z. Effect of the fluid injection configuration on droplet size in a microfluidic T junction. Phys.
Rev. E 2014, 89, 013003. [CrossRef]
20. Du, W.; Fu, T.; Ma, Y.; Li, H.Z. Breakup dynamics for high-viscosity droplet formation in a flow-focusing device: Symmetrical
and asymmetrical ruptures. AlChE J. 2016, 62, 325–337. [CrossRef]
21. Van Loo, S.; Soukatch, S.; Gilet, T. Droplet formation by squeezing in a microfluidic cross-junction. Microfluid. Nanofluidics 2016,
20, 146. [CrossRef]
22. Foroughi, H.; Kawaji, M. Viscous oil–water flows in a microchannel initially saturated with oil: Flow patterns and pressure drop
characteristics. J. Multiph. Flow 2011, 37, 1147–1155. [CrossRef]
23. Azarmanesh, M.; Farhadi, M. The effect of weak-inertia on droplet formation phenomena in T-junction microchannel. Meccanica
2016, 51, 819–834. [CrossRef]
24. Svensson, A.; Wu, Z.; Sunden, B. Effects of size and sodium dodecyl sulfate on mass transfer for liquid-liquid two-phase flow in
non-circular glass microchannels. In International Conference on Nanochannels, Microchannels, and Minichannels; American Societyof
Mechanical Engineers: Vienna, Austria, 2018; p. V001T01A001.
25. Kashid, M.N.; Renken, A.; Kiwi-Minsker, L. CFD modelling of liquid–liquid multiphase microstructured reactor: Slug flow
generation. Chem. Eng. Res. Des. 2010, 88, 362–368. [CrossRef]
26. Kashid, M.; Kiwi-Minsker, L. Quantitative prediction of flow patterns in liquid–liquid flow in micro-capillaries. Chem. Eng.
Process. Process Intensifi. 2011, 50, 972–978. [CrossRef]
27. Tsaoulidis, D.; Dore, V.; Angeli, P.; Plechkova, N.V.; Seddon, K.R. Flow patterns and pressure drop of ionic liquid–water two-phase
flows in microchannels. Int. J. Mult. Flow 2013, 54, 1–10. [CrossRef]
28. Sugiura, S.; Nakajima, M.; Oda, T.; Seki, M. Effect of interfacial tension on the dynamic behavior of droplet formation during
microchannel emulsification. J. Colloid. Interface Sci. 2004, 269, 178–185. [CrossRef]
29. Yagodnitsyna, A.A.; Kovalev, A.V.; Bilsky, A.V. Experimental study of ionic liquid-water flow in T-shaped microchannels with
different aspect ratios. In Journal of Physics: Conference Series; IOP Publishing: Mexican City, Mexico, 2017; p. 032026.
30. Yagodnitsyna, A.; Kovalev, A.; Bilsky, A. Experimental investigation of immiscible liquids flow in a T-shaped microchannel. In
Proceedings of the 10th Pacific Symposium on Flow Visualization and Image Processing, Naples, Italy, 15–18 June 2015; pp. 1–10.
31. Su, Y.; Chen, G.; Yuan, Q. Effect of viscosity on the hydrodynamics of liquid processes in microchannels. Chem. Eng. Technol. 2014,
37, 427–434. [CrossRef]
32. Garstecki, P.; Fuertsman, M.J.; Stone, H.A.; Whitesides, G.M. Formation of droplets and bubbles in a microfluidic T-junction—
scaling and mechanism of break-up. Lab Chip 2006, 6, 437–446. [CrossRef]
33. Vansteene, A.; Jasmin, J.; Cavadias, S.; Clarisse, M.; Cote, G. Towards chip prototyping: A model for droplet formation at both T
and X junctions in dripping regime. Microfluid. Nanofluidics 2018, 22, 1–14. [CrossRef]
34. Li, Y.; Wang, J.; Li, D.; Liu, R. Flow pattern diagrams of oil-water two-phase microflows and stable parallel flows obtained at low
Reynolds numbers. Int. J. Multiph. Flow 2018, 98, 139–146. [CrossRef]
35. Li, M.; Liu, Z.; Pang, Y.; Yan, C.; Wang, J.; Zhao, S.; Zhou, Q. Flow topology and its transformation inside droplets traveling in
rectangular microchannels. Phys. Fluids 2020, 32, 052009.
36. Liu, Z.; Li, M.; Pang, Y.; Zhang, L.; Ren, Y.; Wang, J. Flow characteristics inside droplets moving in a curved microchannel with
rectangular section. Phys. Fluids 2019, 31, 022004.
37. Fu, B.R. Liquid-liquid mixtures flow in micro-channels. In Transactions of the Canadian Society of Mechanical Engineering; Canadian
Science Publishing: Ottawa, ON, Canada, 2013; pp. 631–640.
38. Ladosz, A.; Von Rohr, P.R. Pressure drop of two-phase liquid-liquid slug flow in square micro-channels. Chem. Eng. Sci. 2018, 191,
398–409. [CrossRef]
39. Kashid, M.N.; Agar, D.W. Hydrodynamics of liquid–liquid slug flow capillary microreactor:flow regimes, slug size and pressure
drop. Chem. Eng. J. 2007, 131, 1–13.
40. Jovanović, J.; Zhou, W.; Rebrov, E.V.; Nijhuis, T.A.; Hessel, V.; Schouten, J.C. Liquid–liquid slug flow: Hydrodynamics and
pressure drop. Chem. Eng. Sci. 2011, 66, 42–54. [CrossRef]
41. Kashid, M.N.; Agar, D.W.; Turek, S. CFD modelling of mass transfer with and without chemical reaction in the liquid–liquid slug
flow microreactor. Chem. Eng. Sci. 2007, 62, 5102–5109. [CrossRef]
42. Salim, A.; Fourar, M.; Pironon, J.; Sausse, J. Oil–water two-phase flow in microchannels: Flow patterns and pressure drop
measurements. Can. J. Chem. Eng. 2008, 86, 978–988. [CrossRef]
43. Bretherton, F.P. The motion of long bubbles in tubes. J. Fluid Mech. 1961, 10, 166–188. [CrossRef]
44. Raj, R.; Mathur, N.; Buwa, V.V. Numerical Simulations of Liquid-Liquid Flows in Microchannels. Ind. Eng. Chem. Res. 2010, 49,
10606–10614. [CrossRef]
Energies 2021, 14, 6066 54 of 56
45. Harries, N.; Burns, J.R.; Barrow, D.A.; Ramshaw, C. A numerical model for segmented flow in a microreactor. Int. J. Heat Mass
Transfer 2003, 46, 3313–3322. [CrossRef]
46. Harlow, F.H.; Welch, J.E. Numerical Calculation of Time-Dependent Viscous Incompressible Flow of Fluid with Free Surface.
Phys. Fluids 1965, 8, 2182–2189. [CrossRef]
47. Olsson, E.; Kreiss, G. A conservative level-set method for two phase flow. J. Comp. Phys. 2005, 210, 225–246. [CrossRef]
48. Olsson, E.; Kreiss, G.; Zahedi, S. A conservative level-set method for two phase flow-II. J. Comp. Phys. 2007, 225, 785–807.
[CrossRef]
49. Brackbill, J.U.; Kothe, D.B.; Zemach, C.A. A continuum method for modeling surface tension. J. Comp. Phys. 1992, 100, 335–354.
[CrossRef]
50. Osher, S.; Fedkiw, R.P. Level Set Methods: An Overview and Some Recent Results. J. Comp. Phys. 2001, 169, 463–502. [CrossRef]
51. Van Sind Annaland, M.; Dijkhuisen, W.; Deen, N.G.; Kuipers, J.A.M. Numerical simulations of behavior of gas bubbles using a
3-D front-tracking method. AlChE J. 2006, 52, 99–110. [CrossRef]
52. Hirt, C.W.; Nichols, B.D. Volume of Fluid (VOF) method for the dynamics of free boundaries. J. Comp. Phys. 1981, 39, 201–225.
[CrossRef]
53. Kashid, M.N.; Platte, F.; Agar, D.W.; Turek, S. Computational modelling of slug flow in a capillary microreactor. J. Comp. Appl.
Math. 2007, 203, 487–497. [CrossRef]
54. Yamasaki, Y.; Kariasaki, A.; Morooka, S. Hydrophilic and hydrophobic modifications of microchannel inner walls for liquid-liquid
laminar layered flows. Int. J. Chem. React. Eng. 2010, 8, 1–16. [CrossRef]
55. Kositanont, C.; Tagawa, T.; Yamada, H.; Putivisutisak, S.; Assabumrungrat, S. Effect of surface modification on parallel flow in
microchannel with guideline structure. Chem. Eng. J. 2013, 215, 404–410. [CrossRef]
56. Lan, W.; Li, S.; Wang, Y.; Luo, G. CFD simulation of droplet formation in microchannels by a modified level set method. Ind. Eng.
Chem. Res. 2014, 53, 4913–4921. [CrossRef]
57. Yang, L.; Ladosz, A.; Jensen, K.F. Analysis and simulation of multiphase hydrodynamics in capillary microseparators. Lab Chip
2019, 19, 706–715. [CrossRef]
58. De Menech, M.; Garstecki, P.; Jousse, F.; Stone, S.A. Transition from squeezing to dripping in a microfluidic T-shaped junction. J.
Fluid Mech. 2008, 595, 141–161. [CrossRef]
59. Han, W.; Chen, X. New insights into the pressure during the merged droplet formation in the squeezing time. Chem. Eng. Res.
Des. 2019, 145, 213–225. [CrossRef]
60. Sripadaraja, K.; Umesh, G.; Satyanarayan, M.N. Simulation studies on picolitre volume droplets generation and trapping in
T-junction microchannels. SN Appl. Sci. 2020, 2, 1413. [CrossRef]
61. Rider, W.J.; Kothe, D.B. Reconstructing volume tracking. J. Comp. Phys. 1998, 141, 112–152. [CrossRef]
62. Unverdi, S.O.; Trygvasson, G. A front-tracking method for viscous, incompressible, multi-fluid flows. J. Comp. Phys. 1992, 100,
25–37. [CrossRef]
63. Koynov, A.; Khinast, J.G.; Tryggvason, G. Mass transfer and chemical reactions in bubble swarms with dynamic interfaces. AlChE
J. 2005, 51, 2786–2800. [CrossRef]
64. TeGrotenhuis, W.E.; Cameron, R.J.; Butcher, M.G.; Martin, P.M.; Wegeng, R.S. Microchannel devices for efficient contacting of
liquids in solvent extraction. Sep. Sci. Technol. 1999, 34, 951–974. [CrossRef]
65. Burns, J.R.; Ramshaw, C. The intensification of rapid reactions in multiphase systems using slug flow in capillaries. Lab Chip 2001,
1, 10–15. [CrossRef] [PubMed]
66. Ghaini, A.; Kashid, M.N.; Agar, D.W. Effective interfacial area for mass transfer in the liquid–liquid slug flow capillary microreac-
tors. Chem. Eng. Process. Process Intensif. 2010, 49, 358–366. [CrossRef]
67. Assmann, N.; Von Rohr, P.R. Extraction in microreactors: Intensification by adding an inert gas phase. Chem. Eng. Process. Process
Intensif. 2011, 50, 822–827. [CrossRef]
68. Ufer, A.; Sudhoff, D.; Mescher, A.; Agar, D.W. Suspension catalysis in a liquid–liquid capillary microreactor. Chem. Eng. J. 2011,
167, 468–474. [CrossRef]
69. Xu, B.; Cai, W.; Liu, X.; Zhang, X. Mass transfer behavior of liquid–liquid slug flow in circular cross-section microchannel. Chem.
Eng. Res. Des. 2013, 91, 1203–1211. [CrossRef]
70. Di Miceli Raimondi, N.; Prat, L.; Gourdon, C.; Tasselli, J. Experiments of mass transfer with liquid–liquid slug flow in square
microchannels. Chem. Eng. Sci. 2014, 105, 169–178. [CrossRef]
71. Arsenjuk, L.; Kaske, F.; Franzke, J.; Agar, D.W. Experimental investigation of wall film renewal in liquid–liquid slug flow. Int. J.
Mult. Flow 2016, 85, 177–185. [CrossRef]
72. Li, Q.; Angeli, P. Intensified Eu (III) extraction using ionic liquids in small channels. Chem. Eng. Sci. 2016, 143, 276–286. [CrossRef]
73. Matsuoka, A.; Noishiki, K.; Mae, K. Experimental study of the contribution of liquid film for liquid-liquid Taylor flow mass
transfer in a micro-channel. Chem. Eng. Sci. 2016, 155, 306–313. [CrossRef]
74. Sattari-Najafabadi, M.; Nasr Esfahany, M.; Wu, Z.; Sunden, B. Hydrodynamics and mass transfer in liquid-liquid non-circular
microchannels: Comparison of two aspect ratios and three junction structures. Chem. Eng. J. 2017, 322, 328–338. [CrossRef]
75. Sattari-Najafabadi, M.; Esfahany, M.N.N. Intensification of liquid-liquid mass transfer in a circular microchannel in the presence
of sodium dodecyl sulfate. Chem. Eng. Process. Process Intensif. 2017, 117, 9–17. [CrossRef]
Energies 2021, 14, 6066 55 of 56
76. Singh, K.K.; Renjith, A.U.; Shenoy, K.T. Liquid–liquid extraction in microchannels and conventional stage-wise extractors: A
comparative study. Chem. Eng. Process. 2015, 98, 95–105. [CrossRef]
77. Sahu, A.; Vir, A.B.; Molleti, L.N.S.; Ramji, S.; Pushpavanam, S. Comparison of liquid-liquid extraction in batch systems and
micro-channels. Chem. Eng. Process. Process Intensif. 2016, 104, 190–200. [CrossRef]
78. He, Y.; Chen, K.; Srinivasakannan, C.; Li, S.; Yin, S.; Peng, J.; Guo, S.; Zhang, L. Intensified extraction and separation Pr(III)/Nd(III)
from chloride solution in presence of a complexing agent using a serpentine microreactor. Chem. Eng. J. 2018, 354, 1068–1074.
[CrossRef]
79. Jiang, F.; Yin, S.; Srinivasakannan, C.; Peng, J. Separation of lanthanum and cerium from chloride medium in presence of
complexing agent along with EHEHPA (P507) in a serpentine microreactor. Chem. Eng. J. 2018, 334, 2208–2214. [CrossRef]
80. Sen, N.; Darekar, M.; Sirsat, P.; Singh, K.K.; Mukhopadhyay, S.; Shirsath, S.R.; Shenoy, K.T. Recovery of uranium from lean streams
by extraction and direct precipitation in microchannels. Sep. Purif. Technol. 2019, 227, 115641. [CrossRef]
81. Tang, J.; Zhang, X.; Cai, W.; Wang, F. Liquid–liquid extraction based on droplet flow in a vertical micro-channel. Exp. Therm. Fluid
Sci. 2013, 49, 185–192. [CrossRef]
82. Tsaoulidis, D.; Angeli, P. Effect of channel size on mass transfer during liquid–liquid plug flow in small scale extractors. Chem.
Eng. J. 2015, 262, 785–793. [CrossRef]
83. Ani, Z.; Wahabi, T.A.; Mjalli, F.; Hashmi, A.A.; Idayil, B. Flow of deep eutectic solvent-simulated fuel in circular channels: Part
II—Extraction of dibenzothiophene. Chem. Eng. Res. Des. 2017, 119, 294–300. [CrossRef]
84. He, Y.; Zeng, T.; Chen, K.; Ullah, E.; Li, S.; Zhang, L.; Yina, S.; Guo, S.; Jin, B.; Rene, S. Solvent extraction performance of Sm(III)
using a T-junction microreactor with 2-ethylhexyl phosphonic acid mono-2-ethylhexyl (EHEHPA). Chem. Eng. Process. Process
Intensif. 2019, 136, 28–35. [CrossRef]
85. Yao, C.; Ma, H.; Zhao, Q.; Liu, Y.; Chen, G. Mass transfer in liquid-liquid Taylor flow in a microchannel: Local concentration
distribution, mass transfer regime and the effect of fluid viscosity. Chem. Eng. Sci. 2020, 223, 115734. [CrossRef]
86. Zhang, Y.; Zhang, X.; Xu, B.; Cai, W.; Wang, F. CFD simulation of mass transfer intensified by chemical reactions in slug flow
microchannels. Can. J. Chem. Eng. 2015, 93, 2307–2314. [CrossRef]
87. Bothe, D.; Koebe, M.; Wielage, K.; Warnecke, H.J. VOF simulations of mass transfer from single bubbles and bubble chains rising
in aqueous solutions. In Proceedings of the 4th ASME/JSME Joint Fluids Summer Engineering Conference, Honolulu, HI, USA,
6–10 July 2003; pp. 423–429.
88. Gómez-Pastora, J.; González-Fernández, C.; Fallanza, M.; Bringas, E.; Oritz, I. Flow patterns and mass transfer performance of
miscible liquid-liquid flows in various microchannels: Numerical and experimental studies. Chem. Eng. J. 2018, 344, 487–497.
[CrossRef]
89. Di Miceli Raimondi, N.; Prat, L.; Gourdon, C.; Cognet, P. Direct numerical simulations of mass transfer in square micro-channels
for liquid–liquid slug flow. Chem. Eng. Sci. 2008, 63, 5522–5530. [CrossRef]
90. Cito, S.; Pallares, J.; Fabregat, A.; Katakis, I. Numerical simulation of wall mass transfer rates in capillary-driven flow in
microchannels. Int. Comm. Heat Mass Transfer 2012, 39, 1066–1072. [CrossRef]
91. Ramji, S.; Puspavanam, S. Liquid-liquid extraction in laminar two-phase stratified flows in capillary microchannels. Chem. Eng.
Sci. 2019, 195, 242–249. [CrossRef]
92. Yang, C.; Mao, Z.S. Numerical simulation of interphase mass transfer with the level set approach. Chem. Eng. Sci. 2005, 60,
2643–2660. [CrossRef]
93. Haroun, Y.; Legendre, D.; Raynal, L. Volume of fluid method for interfacial reactive mass transfer: Application to stable liquid
film. Chem. Eng. Sci. 2010, 65, 2896–2909. [CrossRef]
94. Kenig, E.K.; Ganguli, A.A.; Atmakidis, T.; Chasanis, P. A novel method to capture mass transfer phenomena at free fluid-fluid
interfaces. Chem. Eng. Process. Process Intensif. 2011, 50, 68–76. [CrossRef]
95. Skelland, A.H.P.; Wellek, R.M. Resistance to mass transfer inside droplets. AlChE J. 1964, 10, 491–496. [CrossRef]
96. Van Baten, J.M.; Krishna, R. CFD Modeling of a Bubble Column Reactor carrying out a consecutive A → B → C Reaction. Chem.
Eng. Technol. 2004, 27, 398–406. [CrossRef]
97. Sherwood, T.K.; Wei, J.C. Interfacial phenomena in liquid extraction. Ind. Eng. Chem. 1957, 49, 1030–1034. [CrossRef]
98. Asthana, A.; Zinovik, I.; Weinmueller, C.; Poulikakos, D. Significant Nusselt number increase in microchannels with a segmented
flow of two immiscible liquids: An experimental study. Int. J. Heat Mass Transfer 2011, 54, 1456–1464. [CrossRef]
99. Giolla Eain, M.; Egan, V.; Punch, J. Local Nusselt number enhancements in liquid–liquid Taylor flows. Int. J. Heat Mass Transfer
2015, 80, 85–97. [CrossRef]
100. Dai, Z.; Guo, Z.; Fletcher, F.; Haynes, B.S. Taylor flow heat transfer in microchannels—unification of liquid–liquid and gas–liquid
results. Chem. Eng. Sci. 2015, 138, 140–152. [CrossRef]
101. Abdollahi, A.; Norris, S.E.; Sharma, R.N. Fluid flow and heat transfer of liquid-liquid Taylor flow in square microchannels. Appl.
Therm. Eng. 2020, 172, 115123. [CrossRef]
102. Vivekanand, S.V.B.; Raju, V.R.K. Numerical study of the hydrodynamics and heat transfer characteristics of liquid–liquid Taylor
flow in microchannel. Heat Transfer-Asian Res. 2018, 47, 794–805. [CrossRef]
103. Urbant, P.; Leshansky, A.; Halupovic, Y. On the forced convective heat transport in a droplet-laden flow in microchannels.
Microfluid. Nanofluidics 2008, 4, 533–542. [CrossRef]
Energies 2021, 14, 6066 56 of 56
104. Fischer, M.; Juric, D.; Poulikakos, D. Large convective heat transfer enhancement in microchannels with a train of coflowing
immiscible or colloidal droplets. J. Heat Transfer 2010, 132, 112402. [CrossRef]
105. Che, Z.; Wong, T.N.; Nguyen, N.T. Heat transfer enhancement by recirculating flow within liquid plugs in microchannels. Int. J.
Heat Mass Transfer 2012, 55, 1947–1956. [CrossRef]
106. Che, Z.; Wong, T.N.; Nguyen, N.T. Heat transfer in plug flow in cylindrical microcapillaries with constant surface heat flux. Int. J.
Therm. Sci. 2013, 64, 204–212. [CrossRef]
107. Bandara, T.; Cheung, S.C.P.; Rosengarten, G. Slug flow heat transfer in microchannels: A numerical study. Comput. Therm. Sci.
2015, 7, 81–92. [CrossRef]