1 s2.0 S0009250906005604 Main

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 17

Chemical Engineering Science 61 (2006) 7609 – 7625

www.elsevier.com/locate/ces

Numerical study on gas and liquid slugs for Taylor flow


in a T-junction microchannel
Dongying Qian ∗ , Adeniyi Lawal
New Jersey Center for Micro Chemical Systems, Department of Chemical, Biomedical and Materials Engineering, Stevens Institute of Technology,
Castle Point on Hudson, Hoboken, NJ 07030, USA

Received 4 May 2006; received in revised form 26 July 2006; accepted 23 August 2006
Available online 14 September 2006

Abstract
The rapid development of microfabrication techniques creates new opportunities for applications of microchannel reactor technology in
chemical reaction engineering. The extremely large surface-to-volume ratio and the short transport path in microchannels enhance heat and
mass transfer dramatically, and hence provide many potential opportunities in chemical process development and intensification. Multiphase
reactions involving gas/liquid reactants with a solid as a catalyst are ubiquitous in chemical and pharmaceutical industries. The hydrodynamics
of the flow affects the reactor performance significantly; therefore it plays a prominent role in reactor design. For gas/liquid two-phase flow
in a microchannel, the Taylor slug flow regime is the most commonly encountered flow pattern. The present study deals with the numerical
simulation of the Taylor flow in a microchannel, particularly on gas and liquid slugs. A T-junction empty microchannel with varying cross-
sectional width (0.25, 0.5, 0.75, 1, 2 and 3 mm) served as the model micro-reactor, and a finite volume based commercial computational fluid
dynamics (CFD) package, FLUENT, was adopted for the numerical simulation. The gas and liquid slug lengths at various operating and fluid
conditions were obtained and found to be in good agreement with the literature data. Several correlations in the T-junction microchannel were
developed based on the simulation results. The slug flows for other geometries and inlet conditions were also studied.
䉷 2006 Elsevier Ltd. All rights reserved.

Keywords: Fluid mechanics; Multiphase flow; Hydrodynamics; Numerical simulation; Microchannel

1. Introduction packed bed microreactor, single phase microreactor and multi-


phase microreactor (Jensen, 2001).
Microchannel reactors have many advantages over conven- Multiphase reaction in microchannels finds good applications
tional macroreactors. With its small transverse dimensions in many industrially important chemical processes, e.g. direct
(submillimeter), microchannel reactors possess extremely high synthesis of hydrogen peroxide, gas–liquid hydrogenation and
surface to volume ratios, and consequently exhibits enhanced immiscible liquid–liquid nitration. The hydrodynamic charac-
heat and mass transfer rates. Scale-up for high throughput is teristics of the flow plays a very important role in the reactor
achieved by simple replication of microreactor units; which design. Since the flow in a microchannel is typically laminar,
eliminates costly reactor redesign and pilot plant experiments, the mass transfer between the phases is mainly dominated by
thus shortening the development time from laboratory to com- diffusion, and turbulent convective flow normally relied upon
mercial production. Microchannel reaction technology has for rapid and effective mixing in macroreactors is non-existent.
numerous potential applications in chemical engineering, and Mixing is therefore a critical issue in the design of multi-
different types of reactor are often required for various re- phase microreactors. Many investigators have characterized
actions, i.e., thin-wall microreactor, membrane microreactor, gas–liquid flows in a microchannel (Chen et al., 2002; Chung
and Kawaji, 2004; Triplett et al., 1999; Yang and Shieh, 2001;
Fukano and Kariyasaki, 1993; Coleman and Garimella, 1999;
∗ Corresponding author. Tel.: +1 201 216 5539; fax: +1 201 216 8306. Akbar et al., 2003). Bubbly, slug, churn and annular flow are
E-mail address: dqian@stevens.edu (D. Qian). among the observed flow regimes. However, over a wide range
0009-2509/$ - see front matter 䉷 2006 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2006.08.073
7610 D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625

of operating conditions, the flow in a multiphase microchannel solid mass transfer using CFX package. Vandu et al. (2005)
is typically the so-called Taylor flow regime (Heiszwolf et al., obtained a dimensionless expression for liquid to solid mass
2001). The Taylor flow is a special case of slug flow where the transfer after experimentally studying oxygen absorption dy-
bullet-shaped bubbles (Taylor bubbles) are separated by free- namics. Kreutzer et al. (2005b) measured the pressure drop in
gas-entrained liquid slugs. The elongated bubble has a charac- a capillary and also performed CFD studies with FIDAP code.
teristic capsular shape with equivalent diameter larger than the They correlated the apparent friction factor f with the following
channel width. There is a very thin film between the gas bub- expression:
ble and the channel wall. Taylor slug flow has been shown to    
increase transverse heat and mass transfer compared to single 16 d Re 1/3  UT P d
f= 1 + 0.07 , Re = L , (5)
phase laminar flow because of the recirculation within the liq- Re LL Ca L
uid slugs and the reduction of axial mixing between the liquid
slugs (Bercic and Pintar, 1997; Kreutzer et al., 2001; Irandoust where Re is Reynolds number. This empirical correlation holds
and Andersson, 1989a; Thulasidas et al., 1997). for Weber number W e = L UT2 P d/ > 0.1. The constant 0.17
A number of studies on Taylor flow in a microchannel have should be used instead of 0.07 if there are impurities in the
been carried out both experimentally and numerically. Kreutzer experiment, because the impurities give rise to surface tension
et al. (2005c) have summarized the work from the literature. gradient, thus resulting in Marangoni effect. Liu et al. (2005)
Bretherton (1961) theoretically studied the liquid film thick- presented another form of correlation for the friction faction
ness by applying lubrication theory. Irandoust and Andersson from their experiments. Bercic and Pintar (1997), Thulasidas
(1989b) measured the film thickness  in a capillary with spec- et al. (1995b, 1999), and Kreutzer et al. (2005a) measured res-
trometric method and obtained an empirical formula: ident time distribution (RTD) in microchannels and found a
dependence of RTD on the slug length.
/d = 0.18[1 − exp(−3.1Ca 0.54 )], Ca = L UT P /, (1) All the results available in the literature indicate that mass
transfer, pressure drop and RTD for slug flows in microchan-
where d is diameter of the capillary, and Ca is Capillary num- nels are highly liquid film and slug length dependent. Unlike
ber. Thulasidas et al. (1995a) measured the bubble size and other flow parameters such as superficial velocities and fluid
shape, bubble velocity with flow visualization method. Taha properties, the liquid film thickness and the slug length can-
and Cui (2004, 2006a,b) performed CFD studies on bubble not be determined a priori. Flow visualization, conductivity,
size and shape, bubble velocity and velocity field inside the or other advanced tomography techniques like magnetic reso-
slugs in the microchannels by applying the VOF model in nance imaging (MRI) are commonly used to determine these
FLUENT package and found good agreement with the pub- quantities. While the film thickness has been extensively stud-
lished experimental data. For slug flows in capillaries, at very ied experimentally and numerically, there are limited studies
low Ca, the film thickness tends to be zero, and the bubble on the slug length. Laborie et al. (1999) visualized the slug
velocity is approximately equal to the sum of the liquid su- flows in a 1–4 mm capillary for water, glycerol, ethanol—air
perficial velocity and the gas superficial velocity. The veloc- systems, and obtained a dimensionless expression for the gas
ity inside the liquid slug has a parabolic profile with respect and liquid slug lengths:
to the solid wall except for very short liquid slugs. There is a
clear recirculation relative to the bubble movement. The mass LG Re0.63  Vb d  d 2g
= 0.0878 V1.26 , ReV = L , Eo = L , (6)
transfer for the Taylor flow in a microchannel has been inves- d Eo L 
tigated intensively in the literature. Bercic and Pintar (1997)  1.27
proposed the following mass transfer correlations for gas to LL 1  UG d
= 3451 , ReG = L , (7)
liquid and liquid to solid after measuring the physical absorp- d ReG Eo L
tion of methane in water and dissolution of benzoic acid into
water: where Eo is Eotvos number. Liu et al. (2005) and Vandu
et al. (2004) attempted to correlate their experimental data from
kGL aGL = 0.111UT1.19 0.57
P /LL , (2) a 0.9–3.0 mm circular or square channel for water, ethanol–air
systems by adopting an approximate form of Eq. (2):
kLS aLS = 0.069UT0.63
P /(LL − 0.105LG )
0.44
, (3)
L = 0.088ReG ReL ,
UT P /L0.5 ReG = G UG d/G ,
0.72 0.19

where kGL aGL , kLS aLS are mass transfer coefficients, LL and ReL = L UL d/L . (8)
LG are slug lengths, and UT P is unit cell velocity. For the
mass transfer of gas to solid through the liquid film, Kreutzer However, all the data parities for Eqs. (6)–(8) are not satis-
et al. (2001) considered a stagnant layer and provided the mass factory. Mantle et al. (2002) measured the gas slug lengths in
transfer coefficient kGS aGS , as 1.2 mm square channels within a ceramic monolith using MRI
method, and found a wide range of bubble size distributions.
kGS aGS = D/, (4) Heiszwolf et al. (2001) and Kreutzer et al. (2005d) detected
slug frequency using conductivity probes in 1.56 mm square
where D is diffusion coefficient. Van Baten and Krishna (2004, channels within a monolith, from which they calculated the gas
2005) performed CFD studies for gas to liquid and liquid to and liquid slug lengths. They found that the slug length was
D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625 7611

normally between 2 and 10 times the channel diameter. They water


reported that the slug length was highly dependent on the
geometry of the inlet distributors. There was a considerable
variation of the gas and liquid slug lengths, even for the same d
experimental setup; 30% standard deviation was not uncom- 6d
mon. They proposed to measure the pressure drop and then
compute the slug length for the short slugs. 60 d
The slug length is an important parameter for Taylor flows,
however, there is a significant inconsistency in the literature air
studies, therefore, a systematic study on the gas and liquid
y
slug lengths in microchannels is warranted. As Kreutzer et al.
(2005c) pointed out in their review paper, CFD studies will
0 x
provide the mechanistic insight for the slug flow in microchan-
nels. Current CFD studies mainly focus on one unit slug cell, Fig. 1. Model T-junction microreactor used in the simulation.
which presume that the gas and liquid slug lengths are known.
Researchers either used single phase model with void gas represented by d. The mixing zone has a length of 6d while the
(Van Baten and Krishna, 2004, 2005; Kreutzer et al., 2005b; reactor zone has a length of 60d. “Cold” flow without any chem-
Irandoust and Andersson, 1989a) or implemented two phase ical reaction is considered. The whole system is maintained at
VOF model (Taha and Cui, 2004, 2006a,b) to investigate room temperature, and the pressure is atmospheric at the exit.
bubble shape, bubble velocity, film thickness, mass transfer, The superficial gas velocity as well as superficial liquid veloc-
pressure drop and velocity profile inside the liquid slug. Some ity varies from 0.01 to 0.25 m/s. According to the flow regime
commercial software CFD packages such as FLUENT, FIDAP maps by Chung and Kawaji (2004), Triplett et al. (1999), Yang
and CFX have proven to be appropriate to perform these stud- and Shieh (2001), Fukano and Kariyasaki (1993), Coleman and
ies. The aim of this paper is to study numerically the gas and Garimella (1999) and Akbar et al. (2003), at these operating
liquid slugs for Taylor flow in microchannels. A T-junction conditions, the flow falls within the Taylor slug regime in mi-
microchannel serves as a model geometry. The development of crochannels.
the gas and liquid slugs in the model geometry with different
inlet configurations at various operating and fluid conditions is 2.2. Governing equations
investigated.
A finite volume based commercial CFD package, FLUENT
2. Analysis (Release 6.1.22, 2003), was used to perform the numerical sim-
ulations. In order to track the interface between the gas and
2.1. Model geometry liquid slugs, one of the general multiphase models in FLUENT,
the volume of fluid (VOF) model, was adopted to simulate the
The hydrodynamic characteristics of the flow in microchan- Taylor slug flows in the T-junction microchannel. Among sev-
nels are different from those encountered in ordinarily large- eral general multiphase models currently in use, the VOF model
size channels. The threshold of hydraulic channel diameter is is the only multiphase model that enables identification of the
about 1 mm with fluid properties similar to air and water (Akbar interface clearly. The VOF model in FLUENT package has been
et al., 2003). In microchannels, the buoyancy effect is negligi- successfully applied in various applications: drop ejection from
ble in comparison to surface tension, which in effect renders the nozzle of a printhead in an ink jet printer, the shape of free
the flow characteristic independent of channel orientation with surface change due to a sudden change of gravity, three phase
respect to gravity (Triplett et al., 1999). In the studies of chem- free surface flow over a weir, flow of a gas bubble moving in
ical reaction in microchannels, a T-junction microreactor with the surrounding liquid, etc. (FLUENT application briefs and
millimeter to submillimeter range is typically chosen to inves- technical notes, 2002a–d). Taha and Cui (2004, 2006a,b) have
tigate the hydrodynamics or kinetics for multiphase reactions. also demonstrated that the VOF model in FLUENT correctly
The present work considers an empty T-junction microchannel predicts the Taylor flow behavior in a microchannel.
reactor (not packed with catalyst) with cross-sectional width The governing equations of the VOF formulations on multi-
of 0.25, 0.5, 0.75, 1, 2 and 3 mm for the studies of gas–liquid phase flows are as follows:
hydrodynamics. This type of reactor is applicable to homoge- Equation of continuity:
neous reaction or surface reaction with thin-film catalyst coated
j
on the wall. + ∇ · (
v ) = 0. (9)
Fig. 1 is the model T-junction microchannel reactor used in jt
our simulation. The microchannel reactor comprises a vertical Equation of motion:
inlet mixing zone and a horizontal reaction zone. A stream of
j(v)
water and a stream of air are fed separately into the two inlets + ∇ · (
v v) = − ∇p + ∇ · [(∇ v + ∇ vT )]
of the mixing zone, and then enter the reaction zone. The cross- jt
sectional dimensions of the channel are the same, which are g + F .
+  (10)
7612 D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625

Volume fraction equation: where G is the curvature computed from the divergence of the
unit surface normal, and  is the surface tension.
jG
+ v · ∇G = 0, (11) The model geometry was meshed with the preprocessor
jt GAMBIT software (part of FLUENT package), then imported
where into processor FLUENT for calculation. The simulation results
were either analyzed by FLUENT integrated postprocessor or
 = L L + G G ,  = L L + G G . (12) exported into data files for other postprocessor packages. A
segregated time dependent unsteady solver in FLUENT was
The flow is treated as incompressible since the pressure drop used. The boundary conditions were: velocity inlet for the gas
along the reactor is small. Compared to single phase flow sim- and liquid feeds, specified as uniform entrance velocity, and
ulations, the VOF method accomplishes interface tracking by outflow for the reactor outlet which treats the flow as fully de-
solving additional continuity-like equation(s) for the volume veloped at the exit. The other options selected in FLUENT were
fraction (n − 1 equations with n as the number of phases) as follows: the PRESTO! (pressure staggering option) scheme
(Hirt and Nichols, 1981). For gas–liquid two phase flows, the for the pressure interpolation, the PISO (pressure-implicit
volume fraction of gas G is obtained by numerically solving with splitting of operators) scheme for the pressure–velocity
Eq. (11), and the volume fraction of liquid L is simply com- coupling, second-order up-wind differencing scheme for the
puted from 1 − G . However, the implementation of the VOF momentum equation, the geometric reconstruction scheme for
model is numerically challenging. First, both the density  and the interface interpolation, implicit body force treatment for
viscosity  in Eqs. (9) and (10) are mixture properties, which the body force formulation, and Courant number 0.25 for the
vary within the flow domain and are computed by the volume volume fraction calculation. In addition, the time step, the max-
fraction weighted average in Eq. (12). Second, the interface imum number of iterations per time step, and the relaxation
needs to be constructed based on the computed value of the vol- factors need to be carefully adjusted to ensure convergence.
ume fraction with the application of interpolation schemes for
the identification of the interface. Third, surface tension along
3. Results and discussion
the interface between each pair of phases, and wall adhesion
between the phases and the walls become very important for
3.1. Slug flow development
some cases, especially when the gravitational effect is negli-
gible. All these complexities make VOF multiphase flow sim-
The development of the slug flow was initially investigated
ulation computationally expensive, and convergence difficult
in the model T-junction microchannel as well as in a straight
to achieve when compared to its single phase counterpart. In
microchannel. For the straight channel, the feed was a mathe-
fact, multiphase flow is perhaps the most difficult topic in the
matically well pre-mixed gas/liquid mixture. The microchannel
CFD simulation. Theory and simulations agree with experiment
was first filled with water, and then the streams were introduced
with a degree of accuracy rarely found in multiphase reactor
at time zero. Figs. 2 and 3 show the slug flow development at
engineering (Kreutzer et al., 2005c).
different time steps in the T-junction geometry and the straight
microchannel respectively. It can be seen that, the gas and
2.3. Numerical approach

The flow inside the microchannel is essentially laminar. Typ-


ically, surface tension force dominates over the gravitational
effect, and wall adhesion force is also important. In FLUENT,
the surface tension model is the continuum surface force (CSF)
model proposed by Brackbill et al. (1992). With this model,
the addition of surface tension to the VOF calculation results
in a source term in the momentum Eq. (10) (FLUENT, 2003).
In the case of wall adhesion, a contact angle needs to be spec-
ified. Rather than impose the boundary condition at the wall
itself, the contact angle that the fluid is assumed to make with
the wall is used to adjust the surface normal in cells near the
wall. This so-called dynamic boundary condition results in the
adjustment of the curvature of the surface near the wall, and
this curvature is then used to adjust the body force term in the
surface tension calculation (FLUENT, 2003). For gas–liquid
two phase flows, the source term in Eq. (10) that arises from
surface tension can be computed from the following:
2G ∇G
F =  , (13) Fig. 2. Slug flow development in the model geometry (red——gas slug,
(G + L ) blue——liquid slug, d = 0.5 mm, UG = UL = 0.02 m/s).
D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625 7613

Fig. 3. Slug flow development in a straight microchannel (red——gas slug,


blue——liquid slug, d = 0.5 mm, UG = UL = 0.02 m/s).

Table 1
Slug length in a 3D tube and a 2D duct with a pre-well-mixed inlet stream

d UG UL LG LG LL LL
(mm) (m/s) (m/s) (mm) (mm) (mm) (mm)
Fig. 4. Laplace pressure in a 3D tube and a 2D duct.
(3D) (2D) (3D) (2D)

0.25 0.05 0.05 0.53 0.54 0.37 0.47


0.50 0.05 0.05 0.96 0.90 0.62 0.80
0.75 0.05 0.05 1.57 1.55 1.34 1.40 that of a 2D case since it involves both radii of curvature with
1.0 0.05 0.05 2.10 2.10 1.85 1.80 a spherical bubble. This Laplace pressure term, which results
2.0 0.05 0.05 4.98 4.73 3.60 3.70 from the surface tension, was computed from the average pres-
3.0 0.05 0.05 7.70 7.45 6.00 5.85 sure drop across the front interface and the back interface, with
0.5 0.02 0.08 0.39 0.36 1.56 1.78
0.5 0.08 0.02 2.57 2.40 0.33 0.31
several randomly selected points for each tube size. From the
figure, one can see that the Laplace pressure from the simu-
lations satisfies the Young–Laplace equation quite well. Some
discrepancy is caused by the difference between the tube size
liquid slugs develop and propagate either at the Tee intersec- and the real bubble shape and size since the resolution in sim-
tion or near the entrance of the channel. After a slug develops ulations is not able to capture adequately the bubble shape and
and detaches, it travels downstream. The slug is generally sta- film thickness.
ble; it seldom breaks or coalesces or collapses unless it is a very In the following discussions, 2D simulation results are shown
short gas slug. Both a uniform velocity inlet and a parabolic because it takes an enormous amount of time to perform a single
velocity inlet have the same slug lengths for these geometries. 3D run in order to collect sufficient statistic data in comparison
It is also worth noting that even though the flow conditions are to a 2D run. The focus of this work is the slug length, and there
the same for these two cases, the slug lengths are different on is not much difference between a 3D tube and a 2D duct based
account of the difference of the inlet geometry. on the simulation results.
In the simulations, the Ca is on the order of 10−3 , so the film
3.2. Slug flow parameters thickness is about 0.01d, and the bubble velocity is about 3%
faster than the bulk superficial velocity due to the slip. Thus
Both 3D simulations for a circular tube and 2D simulations the gas slug velocity Vb will be in use approximately equal to
for a duct at the same operating conditions were first performed. UG + UL in the following discussions. From the simulation re-
The purpose of this exercise was to determine how the third di- sults, the gas slug velocity was computed and the approximation
rection affects the slug length. In this simulation, a straight tube was confirmed to be valid. The velocity across a sectional line
and a microchannel with a pre-well-mixed inlet stream were inside the bulk slug has a parabolic profile, with a slightly big-
selected. The results of the slug length are listed in Table1. One ger gas velocity than the liquid velocity; this is always true in
observes that between 3D and 2D geometries, there is not much the simulation as long as the slug is not too short. Flow recircu-
difference in terms of slug length. The little difference between lation inside the liquid slug has been observed on the reference
the liquid slug lengths especially for smaller size channels is frame of the gas slug. Fig. 5 shows the pressure profile along
probably due to the lower resolution of 3D simulations, that is, the axial direction. The upper lines are in the gas slugs while
the impact of artificially finite gas/liquid interface is stronger the lower ones are in the liquid slugs. The pressure drop can be
for the smaller size slugs. The Laplace pressure with respect decomposed into a viscous term due to friction and a Laplace
to the tube size is plotted in Fig. 4, where the lines represent term due to surface tension. The pressure difference between
the Laplace theory. The Laplace pressure of a 3D case is twice the adjacent gas and liquid slug is the Laplace pressure. As one
7614 D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625

Fig. 6. Slugs at different mesh resolutions (d =0.5 mm, UG =UL =0.02 m/s).
Number of quadrilateral cells (a) 1665 (b) 3087 (c) 6600 (d) 13202 (e) 26400.

Fig. 5. Pressure profile along axial direction (d=0.5 mm, UG =UL =0.02 m/s).

can see, the pressure drop across the back interface is always
higher than that across the front interface, so one can expect
that the shape of the interface is different at the bubble head and
the bubble end, with the bubble end having a bigger curvature.

3.3. Comparison of experimental data with numerical


predictions

The simulations were carried out on Linux PC cluster with


2.8 GHz Pentium IV or Xeon processors. A few thousand to ten
thousand time steps were required for each run. Twenty to a
hundred slugs were collected, from which the sample averages
and other statistical information were computed. Unless oth-
erwise stated, the data presented in the following sections are Fig. 7. (f Re) as a function of the dimensionless group (LL /d)∗ (Ca/Re)0.33 .
for an empty bed with surface tension of 0.072 N/m, a liquid
viscosity of 0.001 kg/ms and wall contact angle of liquid of 0◦ .
The study of mesh dependence on the simulations was ini- Kariyasaki, 1993; Coleman and Garimella, 1999), no corre-
tially performed. A uniform rectangular grid was used for 2D sponding slug length data were presented. Laborie et al. (1999)
simulations. In the beginning, a series of mesh cells with dif- presented some experimental data about the gas and liquid slug
ferent fineness were chosen until the results did not show sig- lengths inside capillaries. However, the air was introduced into
nificant changes. Fig. 6 shows the slug contours at different the water through a porous membrane, rather than through a
mesh resolutions. The significant difference on these contour Tee mixer. According to our simulation, the degree of premix-
plots occurs only on the interface. The higher the resolution, ing of the gas and the liquid and the inlet geometry strongly
the sharper the interface is. When the resolution is too low, part affect the slug length, while the length of the mixer or the reac-
of the interface is twisted so the shape is not always a circular. tor has no effect on the slug length as long as it is long enough
The data presented in this work were based on 6600 quadrilat- to obtain a fully developed flow.
eral cells in the model geometry. Fig. 7 plots the dimensionless apparent friction factor for
A number of code validation exercises have been carried out. air–water and air–decane (surface tension 0.024 N/m) systems
The experimental data are taken from the literature. As ear- compared with experimental data by Kreutzer et al. (2005b).
lier stated, most of the literature studies characterized the flow The pressure drop and liquid slug length were the average
patterns, and presented the flow regime maps inside the mi- values of randomly selected points. The simulations are in
crochannels. Although there were a few pictures of the slug good agreement with the experiments. The scattering of the
flows in these studies (Chen et al., 2002; Chung and Kawaji, points are due to the non-uniformity of the slugs. Vandu et al.
2004; Triplett et al., 1999; Yang and Shieh, 2001; Fukano and (2004) conducted air–water flow visualization experiments in
D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625 7615

Table 2
Comparison between numerical simulations and experiments in a 3 mm vertical circular capillary

UG UL LG LGmin (mm) LGmax (mm) LL LLmin (mm) LLmax (mm)


(m/s) (m/s) (mm) LGmin (exp.) LGmax (exp.) (mm) LLmin (exp.) LLmax (exp.)

0.05 0.02 15.38 13.95 16.65 7.28 6.30 8.10


13.00 15.00 4.95 9.45
0.05 0.05 12.80 12.15 13.50 12.47 11.25 13.50
12.00 13.00 12.00 13.00
0.05 0.10 6.22 3.60 9.45 9.39 3.15 18.90
3.50 9.50 6.00 18.50
0.05 0.15 4.56 2.25 9.45 9.10 2.25 20.25
3.00 8.50 3.00 25.00
0.05 0.20 3.30 1.35 8.10 7.48 1.35 22.05
2.00 9.50 1.50 38.00
0.10 0.02 19.19 6.75 32.85 3.98 1.80 7.65
20.50 31.00 3.00 5.00
0.10 0.05 15.39 9.45 18.90 5.64 2.25 8.55
17.00 18.50 6.50 7.50
0.10 0.10 5.55 4.50 13.50 3.22 1.35 8.10
10.00 13.50 7.50 11.00
0.10 0.15 4.98 3.60 10.35 4.50 2.25 9.90
3.50 13.00 2.00 15.50
0.10 0.20 4.93 2.25 8.55 6.42 1.35 12.60
3.00 9.00 1.50 24.00

a 3.0 mm vertical capillary through a T-junction mixer. From


their movies, the slug length information was obtained. Table 2
lists our simulation results and their experimental data. It can be
seen that the simulations are consistent with the experiments. In
Table 2, LG and LL are sample average values (time and space
mean values). Both simulation results and experimental data
indicate that the slug length is not uniform throughout the chan-
nel because of the time-dependent toroidal vortices generated
when the two streams meet at the mixing section, and which
subsequently propagate throughout the length of the channel.
This non-uniformity becomes more pronounced as the superfi-
cial gas or liquid velocity increases. Fig. 8 shows five snapshot
flow patterns at different liquid flow rates. The non-uniformity
of the slugs can be clearly seen in this figure. The simulations
confirm that, for all the above operating conditions, the flow
in each case is in the Taylor slug regime. Fig. 9 is the bar plot
of the gas and liquid slug length distributions corresponding to
these cases, where N is the number of slugs at a particular size
range, and N0 is the total number of slugs. As one observes,
the slug length variation becomes wider as the liquid flow rate
increases. Fig. 8. Contour plots of volume fraction of air in a 3 mm vertical capillary
(gravity direction: −x).
Heiszwolf et al. (2001) and Kreutzer et al. (2005b) studied
the gas–liquid flows in a monolith with 0.5–5 mm channels. The
superficial velocity is in the range of 0.02 m/s < UG < 0.3 m/s the dimensionless gas slug length LG /d and liquid slug length
and 0.02 m/s < UL < 0.2 m/s. They observed slug flows in LL /d as a function of liquid hold-up L . Here the liquid flow
these operating conditions, and measured the slug length with rate volume fraction UL /(UG + UL ) is approximately equal to
conductivity probes. Different flow distributors were used and the liquid hold-up L . As one observes, the simulation results
the experiments showed that the slug lengths were depen- have the same trend as, and agree favorably with the exper-
dent on the inlet geometry. They also reported strong non- imental data. The scattered data indicate that the slug length
uniformity of the slugs. Our simulations were compared with should also depend on other parameters such as flow rate and
their experiments on the coarse distributor because we think fluid properties. The data variations in big diameter channels
that the behavior of the Tee mixer should closely approximate or high flow rates are bigger because of the high variation
the coarse distributor. Fig. 10 depicts the relationship between of the slug length at these conditions. For the considerable
7616 D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625

Fig. 9. Slug length distributions in a 3 mm vertical capillary.


D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625 7617

Fig. 10. Dimensionless slug length against liquid hold-up.

Fig. 11. Gas slugs and liquid slugs at different inlet mixing levels (d = 0.5 mm, UG = UL = 0.02 m/s).

range of the liquid hold-up (0.25 < L < 0.75), the dimension- Fig. 12 shows the contour of gas slugs and liquid slugs at
less liquid slug length is between 1 and 10, which is well in three different inlet Tee orientations. As one can see, there is
agreement with the literature. not much difference in the slug length between (a) and (b),
however, the slug length of (c) is much longer. It is thus better
3.4. Slug length for different inlet geometries to introduce gas and liquid feeds head to head or perpendicular
to each other with the liquid stream parallel to the microchan-
As stated above, the slug length is highly dependent on the nel. Perpendicular feeds with the gas stream parallel to the mi-
inlet geometry. In this work, the slug flows with different inlet crochannel should be avoided because this case has smoother
configurations or different mixing levels have been investigated. stream at the mixing section, and the slug is more difficult to
Fig. 11 shows the slugs in a straight channel at different inlet detach. Fig. 13 is the contour of gas slugs and liquid slugs for
mixing levels. This can be modeled as if there are alternating three different inlet Tee sizes. Although the microchannel ge-
multilaminated channels before the inlet zone. There are 1, 2, ometry and flow rate are same, the slug length is totally differ-
4 and infinite laminating levels corresponding to these cases. ent: the smaller the diameter of the mixing zone, the shorter is
The result indicates that the degree of the inlet premixing alone the slug length. This is due to the fact that the smaller mixing
will influence the slug length: the better the inlet premixing, zone generates larger velocity, hence bigger toroidal vortices
the shorter the slug length. are produced at this region and make it easier for the slug to
7618 D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625

Fig. 12. Gas slugs and liquid slugs at different tee orientations (d = 0.5 mm, UG = UL = 0.02 m/s).

detach. This is a conclusion that is generally valid, i.e., the gen-


eration of big toroidal vortices in the mixing zone reduces the
slug length.

3.5. Gravitational effect on slug flow

In conventional-size tubes as well as millimeter order tubes,


two-phase flow patterns are dominated in general by gravity
with surface tension effects relatively insignificant. On the other
hand, in microchannels of the order of a few microns to a few
tens of microns, two-phase flow is influenced mainly by sur-
face tension, viscous and inertial effects, with gravity playing
a minor role. In order to investigate the gravitational effect in
microchannels, two sets of runs were performed; one with grav-
itational effect included while in the other it is ignored. Fig. 14
shows the comparison of gas and liquid slug lengths with and
without gravity, where the subscript “no” represents no gravity
condition, and Bo is Bond number (Bo = L gd 2 /). One can
see that, as the Bo number increases, the difference between the
slug lengths with and without gravity grows bigger. Since most Fig. 13. Gas slugs and liquid slugs at different inlet tee sizes (d = 0.5 mm,
data here at high Bo number are from bigger channels, gravity UG = UL = 0.02 m/s).
has more influence in millimeter channels. The scatter in the
data in the same size channels demonstrates that the degree of
the gravitational effect also depends on other flow parameters superficial liquid velocity UL , surface tension , liquid viscos-
like Re and Ca. ity L and wall contact angle of liquid  have been performed.
The influence of gas and liquid superficial velocities on the
3.6. Influence of gas and liquid superficial velocities gas and liquid slug lengths was first studied. Fig. 15 shows
the relationship between the gas slug length and the super-
The previous studies show that the slug length depends upon ficial velocities, while Fig. 16 shows the corresponding be-
flow velocity, inlet geometry and fluid properties. Now the havior for the liquid slug length. The gas slug grows with
slug flows in the model T-junction microchannel are focused to increase in superficial gas velocity, and decrease in superfi-
study. A total of about 500 simulation runs with varying param- cial liquid velocity. Similar trend is observed for the liquid
eter values: channel diameter d, superficial gas velocity UG , slug length, i.e., it increases with increase in superficial liquid
D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625 7619

Fig. 14. Comparison of gas and liquid slug lengths with/without gravity.

velocity, and decrease in superficial gas velocity. This trend 3.8. Influence of wall surface adhesion
is consistent with experimental observations of Laborie et al.
(1999) who reported a linear relationship between LG and UG The influence of wall surface adhesion was also investigated.
in their experimental region. Here, power-law fits for the sam- As stated above, the wall surface adhesion plays an equally
ple data were used, which are represented as the solid curves important role as the surface tension. Here, the contact angle of
in Figs. 15 and 16. For cross-sectional dimensions of 1 mm the liquid to the wall was varied from 0◦ to 180◦ , and the results
and lower, the data are well described by one curve. For the are shown in Fig. 19. It is seen that both the gas and liquid
computational geometries considered in this study, the 1 mm slug lengths decrease slightly until they are almost the same
and submillimeter channels exhibit more uniform slugs than after the liquid becomes non-wetting. With the change of the
millimeter-size channels. The simulation data show more scat- wettability of the fluid on the wall from fully wetting (contact
ter for the 2 and 3 mm channels in comparison to the submil- angle 0◦ ) to fully non-wetting (contact angle 180◦ ), the shape
limeter channels. The effect of superficial liquid velocity on of the gas–liquid interface changes from concave in to convex
gas slug length, and the effect of superficial gas velocity on out from the liquid side accordingly. It is speculated that there
liquid slug length are very remarkable. The higher the ratio of is a thin film around the gas slug for the fully wetting cases.
superficial gas velocity to superficial liquid velocity, the longer Most literature sources report a thin liquid film around the gas
are the gas slugs. Similar observation also applies to the liquid slug on the order of several to tens of microns. Unfortunately
slugs. Figs. 15 and 16 also show the influence of the channel our grid resolution is not fine enough to capture the liquid film.
width on the gas and liquid slug lengths: at the same superfi- For fully wetting fluid, one can expect the gas to liquid slug
cial gas and liquid velocities, a wider channel has longer gas length ratio to be slightly higher than the volume flow rate
and liquid slugs. ratio because of the existence of the thin film, and the fact that
the slug length was computed across the central line, with the
shape of the interface affecting the values. Our data show that
3.7. Influence of fluid properties there is always a 20–30% difference between the slug length
ratio and volume flow rate ratio.
A number of simulations with varying liquid properties in
submillimeter channels were performed. Fig. 17 shows the in-
3.9. Correlations
fluence of surface tension, while Fig. 18 shows the influence
of liquid viscosity. In our computational region, the slug length
Eqs. (9)–(11) can be rewritten in dimensionless form using
slightly increases with the increase of surface tension; however,
dimensional analysis.
there is almost no influence of liquid viscosity. These results in-
Equation of continuity:
dicate that the surface tension effect is stronger than the viscous
effect for this system. When the surface tension is very small, it
is difficult to generate slug flow. For example, when the surface j∗
+ ∇ · (∗ v∗ ) = 0. (14)
tension is close to zero, the reality flow is the stratified flow. jt ∗
7620 D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625

Fig. 15. Influence of gas and liquid superficial velocities on gas slug length.

Equation of motion: Volume fraction equation:

j(∗ v∗ ) jG


+ ∇ · (∗ v∗ v∗ ) + v∗ · ∇G = 0, (16)
jt ∗ jt ∗
1
= −∇p ∗ + ∇ · [∗ (∇ v∗ + ∇ v∗ )]
T
where
Re
1 ∗ ∗ F ∗ ∗ = /L = G G /L + (1 − G ),
+  g + . (15)
Fr Re · Ca ∗ = /L = G G /L + (1 − G ), (17)
D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625 7621

Fig. 16. Influence of gas and liquid superficial velocities on liquid slug length.

tV b p g velocity Vb . For the conditions used in the simulation, Vb ≈


t∗ = , v ∗ = v/Vb , p∗ = , g∗ = , (18)
d L Vb2 g UG + UL and G ≈ UG /(UG + UL ), so UG = G Vb and UL =
L Vb = (1 − G )Vb . From Eqs. (14)–(16) and the correspond-
L Vb d L Vb Vb2
Re = , Ca = , Fr = . (19) ing dimensionless boundary conditions, one can solve for the
L  gd three dependent variables p∗ , v∗ , G as a function of the inde-
Here the dimensionless numbers are defined based on chan- pendent variables t ∗ , x∗ and the parameters Re(=L Vb d/L ),
nel width d, liquid properties L , L and gas slug moving Ca(=L Vb /), F r(=Vb2 /gd), G /L , G /L , G , , where
7622 D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625

Fig. 17. Influence of surface tension on gas and liquid slug length (UG = 0.1 m/s, UL = 0.1 m/s).

Fig. 18. Influence of liquid viscosity on gas and liquid slug lengths (UG = 0.1 m/s, UL = 0.1 m/s).

G is introduced from the velocity inlet boundary conditions, value (time and space average), one can conclude that the
and  comes from the wall surface adhesion boundary condi- dimensionless slug length will have the following dependence:
tion. Then the slug length can be obtained from the gas vol-
ume fraction map. Since the simulated slug length is a mean L∗ = L∗ (Re, Ca, F r, G /L , G /L , G , ). (20)
D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625 7623

dimensionless slug length falls within a narrow range even for


wide operating conditions.

4. Conclusions

The characteristics of Taylor slug flow in microchannels have


been elucidated using CFD simulations. The results show that
the slug length is not uniform throughout the channel especially
for channel cross-sectional dimensions exceeding 1 mm. When
the gas or liquid flow rate increases, the slug non-uniformity
becomes more pronounced. The slug length is also highly de-
pendent on the inlet configuration. From the studies of gas and
liquid slugs in a T-junction microchannel at various operating
conditions, the following conclusions are drawn:

• The gas slug length increases with increase in superficial gas


velocity, and decrease in superficial liquid velocity.
• The liquid slug length increases with increase of superficial
liquid velocity, and decrease of superficial gas velocity.
Fig. 19. Influence of contact angle on gas and liquid slug lengths (d=0.50 mm, • The dimensionless slug length is mainly determined by the
UG = 0.1 m/s, UL = 0.1 m/s). phase hold-up, with a slight effect of Re and Ca. Therefore,
wider channels have longer slug length at the same superficial
gas and liquid velocities.
• Gravitational effects can be ignored in microchannels.
From the simulations, we have confirmed that gravity does
• The effects of fluid density and viscosity are also negligible.
not have much of an effect in slug flow in microchannels, so
• The surface tension and wall surface adhesion moderately
the Froude number can be ignored. The viscosity of the gas
impact the slug lengths.
is negligible compared to that of the liquid; hence G /L can
be dropped. We have observed this from the simulations. The
After a dimensionless analysis from the governing equa-
same argument applies to the density ratio; therefore, G /L is
tions, correlations for the dimensionless unit slug length, gas
ignored. Furthermore we do not consider the influence of the
slug length and liquid slug length in the model T-junction mi-
contact angle, and finally Eq. (20) reduces to the form:
crochannel have been obtained. These correlations can be used
to predict the slug length based on the operating condition,
L∗ = L∗ (Re, Ca, G ). (21) and subsequently some important transport parameters such as
pressure drop, heat transfer and mass transfer coefficients can
Regression of the simulation data provides the following cor- be estimated. Based on this numerical study of Taylor flow in
relations for the dimensionless slug length: microchannels, we can begin to understand, and quantify heat
and mass transfer enhancement in microchannels, which will
(LG + LL )/d = 1.637−0.893
G (1 − G )−1.05 be the subject of our next study.
× Re−0.075 Ca −0.0687 , (22)
Notation
LG /d = 1.6370.107
G (1 − G )−1.05 Re−0.075 Ca −0.0687 , (23)
a mass transfer surface area, m2 /m3
LL /d = 1.637−0.893
G (1 − G )−0.05 Re−0.075 Ca −0.0687 . (24) d internal channel width, m
D diffusion coefficient, m2 /s
A total of 148 sets of numerical simulation data were f apparent friction factor
correlated with channel width 1 mm and less for parame- F body force, N
ter values in the range: 0.09 < G < 0.91, 15 < Re < 1500, g gravity, kg/m2
0.000278 < Ca < 0.01. Eq. (23) is obtained by multiplying k mass transfer coefficient, m/s
Eq. (22) by G , while Eq. (24) is obtained by multiplying L mean slug length, m
Eq. (22) by 1 − G , for LG /(LG + LL ) ≈ G and LL /(LG + L̃ in-situ slug length, m
LL ) ≈ 1 − G in our simulation conditions. The average dif- N number of slugs
ference between the correlated data and the simulated data is N0 total number of slugs
about 10%. These correlations show that the dimensionless slug p system pressure, Pa
length is mainly determined by the phase hold-up, while the t time, s
Re and Ca have a slight effect. This result can explain why the U superficial velocity, m/s
7624 D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625

UT P unit cell velocity (=UG + UL ), m/s Coleman, J.W., Garimella, S., 1999. Characterization of two-phase flow
v velocity, m/s patterns in small diameter round and rectangular tubes. International
Journal of Heat and Mass Transfer 42, 2869–2881.
Vb gas slug velocity (≈ UG + UL ), m/s
FLUENT application briefs, 2002a. Drop ejection from a printhead nozzle,
EX147, Fluent Incorporated, Lebanon, New Hampshire.
Greek letters FLUENT application briefs, 2002b. Free surface change during free-fall,
EX168, Fluent Incorporated, Lebanon, New Hampshire.
 volume fraction FLUENT application briefs, 2002c. Three phase free surface flow, EX184,
 liquid film thickness, m Fluent Incorporated, Lebanon, New Hampshire.
 hold-up FLUENT technical notes, 2002d. Unsteady deformation and internal
 contact angle, ◦ circulation of a gas bubble in a zero gravity uniform flow. TN 282, Fluent
Incorporated, Lebanon, New Hampshire.
 curvature, 1/m
FLUENT 6.1 documentation, 2003. Fluent Incorporated, Lebanon, New
 molecular viscosity, kg/m s Hampshire.
 density, kg/m3 Fukano, T., Kariyasaki, A., 1993. Characteristics of gas–liquid two–phase
 surface tension, N/m flow in a capillary tube. Nuclear Engineering and Design 141, 59–68.
Heiszwolf, J.J., Kreutzer, M.T., van den Eijnden, M.G., Kapteijn, F., Moulijn,
Dimensionless numbers J.A., 2001. Gas–liquid mass transfer of aqueous Taylor flow in monoliths.
Catalysis Today 69, 51–55.
Bo Bond number (=d 2 g/) Hirt, C.W., Nichols, B.D., 1981. Volume of fluid (VOF) method for the
dynamics of free boundaries. Journal of Computational Physics 39,
Ca capillary number (=U/)
201–225.
Eo Eotvos number (=d 2 g/) Irandoust, S., Andersson, B., 1989a. Simulation of flow and mass-transfer
Fr Froude number (=U 2 /gd) in Taylor flow through a capillary. Computers Chemical Engineering 13
Re Reynolds number (=U d/) (4/5), 519–526.
We Weber number (=U 2 d/) Irandoust, S., Andersson, B., 1989b. Liquid film in Taylor flow through a
capillary. Industrial and Engineering Chemistry Research 28, 1684–1688.
Jensen, K.F., 2001. Microreaction engineering—is small better? Chemical
Subscripts and superscripts
Engineering Science 56, 293–303.
Kreutzer, M.T., Du, P., Heiszwolf, J.J., Kapteijn, F., Moulijn, J.A., 2001.
G gas Mass transfer characteristics of three-phase monolith reactors. Chemical
L liquid Engineering Science 56, 6015–6023.
S solid Kreutzer, M.T., Bakker, J.J.W., Kapteijn, F., Moulijn, J.A., Verheijen,
min minimum P.J.T., 2005a. Scaling-up multiphase monolith reactors: linking residence
max maximum time distribution and feed maldistribution using isobars. Industrial and
Engineering Chemistry Research 44 (14), 4898–4913.
exp experiment
Kreutzer, M.T., Kapteijn, F., Moulijn, J.A., Kleijn, C.R., Heiszwolf, J.J.,
* dimensionless value 2005b. Inertial and interfacial effects on pressure drop of Taylor flow in
capillaries. A.I.Ch.E. Journal 51 (9), 2428–2440.
Kreutzer, M.T., Kapteijn, F., Moulijn, J.A., Heiszwolf, J.J., 2005c. Multiphase
Acknowledgments monolith reactors: chemical reaction engineering of segmented flow in
microchannels. Chemical Engineering Science 60, 5895–5916.
The authors gratefully acknowledge the financial support Kreutzer, M.T., van der Eijnden, M.G., Kapteijn, F., Moulijn, J.A., Heiszwolf,
from Department of Energy-Industrial Technologies Program J.J., 2005d. The pressure drop experiment to determine slug lengths in
monoliths. Catalysis Today 105, 667–672.
(DOE-ITP) under Grant Nos. DE-FC36-021D14427 and DE-
Laborie, S., Cabassud, C., Durand-Bourlier, L., Laine, J.M., 1999.
FC36-03G013156. We also thank Dr. Kreutzer for providing Characterization of gas–liquid two-phase flow inside capillaries. Chemical
some raw experimental data. Engineering Science 54, 5723–5735.
Liu, H., Vandu, C.O., Krishna, R., 2005. Hydrodynamics of Taylor flow in
vertical capillaries: flow regimes, bubble rise velocity, liquid slug length,
References and pressure drop. Industrial and Engineering Chemistry Research 44 (14),
4884–4897.
Akbar, M.K., Plummer, D.A., Ghiaasiaan, S.M., 2003. On gas–liquid two- Mantle, M.D., Sederman, A.J., Gladden, L.F., 2002. Dynamic MRI
phase flow regimes in microchannels. International Journal of Multiphase visualization of two-phase flow in a ceramic monolith. A.I.Ch.E. Journal
Flow 29, 855–865. 48 (4), 909–912.
Bercic, G., Pintar, A., 1997. The role of gas bubbles and liquid slug lengths on Taha, T., Cui, Z.F., 2004. Hydrodynamics of slug flow inside capillaries.
mass transport in the Taylor flow through capillaries. Chemical Engineering Chemical Engineering Science 59, 1181–1190.
Science 52 (21/22), 3709–3719. Taha, T., Cui, Z.F., 2006a. CFD modelling of slug flow in vertical tubes.
Brackbill, J.U., Kothe, D.B., Zemach, C., 1992. A continuum method for Chemical Engineering Science 61, 676–687.
modeling surface tension. Journal of Computational Physics 100, 335–354. Taha, T., Cui, Z.F., 2006b. CFD modelling of slug flowinside square capillaries.
Bretherton, F.P., 1961. The motion of long bubbles in tubes. Journal of Fluid Chemical Engineering Science 61, 665–675.
Mechanics 10, 166–188. Thulasidas, T.C., Abraham, M.A., Cerro, R.L., 1995a. Bubble-train flow in
Chen, W.L., Twu, M.C., Pan, C., 2002. Gas–liquid two-phase flow in micro- capillaries of circular and square cross section. Chemical Engineering
channels. International Journal of Multiphase Flow 28, 1235–1247. Science 50 (2), 183–199.
Chung, P.M.-Y., Kawaji, M., 2004. The effect of channel diameter on adiabatic Thulasidas, T.C., Cerro, R.L., Abraham, M.A., 1995b. The monolith froth
two-phases flow characteristics in microchannels. International Journal of reactor: residence time modelling and analysis. Chemical Engineering
Multiphase Flow 30, 735–761. Research and Design 73, 314–319.
D. Qian, A. Lawal / Chemical Engineering Science 61 (2006) 7609 – 7625 7625

Thulasidas, T.C., Abraham, M.A., Cerro, R.L., 1997. Flow patterns in liquid Van Baten, J.M., Krishna, R., 2005. CFD simulation of wall mass transfer
slugs during bubble-train flow inside capillaries. Chemical Engineering for Taylor flow in circular capillaries. Chemical Engineering Science 60,
Science 52 (17), 2947–2962. 1117–1126.
Thulasidas, T.C., Abraham, M.A., Cerro, R.L., 1999. Dispersion during Vandu, C.O., Liu, H., Krishna, R., 2004. Taylor bubble rise in circular and
bubble-train flow in capillaries. Chemical Engineering Science 54 (1), square capillaries. University of Amsterdam, Amsterdam, The Netherlands,
61–76. the 20th of July 2004 http://ct-cr4.chem.uva.nl/singlecapillary/.
Triplett, K.A., Ghiaasiaan, S.M., Abdel-Khalik, S.I., Sadowski, D.L., 1999. Vandu, C.O., Liu, H., Krishna, R., 2005. Mass transfer from Taylor
Gas–liquid two-phase flow in microchannels Part I: two-phase flow bubbles rising in single capillaries. Chemical Engineering Science 60,
patterns. 25, 377–394. 6430–6437.
Van Baten, J.M., Krishna, R., 2004. CFD simulation of mass transfer from Yang, C.-Y., Shieh, C.-C., 2001. Flow pattern of air–water and two-phase
Taylor bubbles rising in circular capillaries. Chemical Engineering Science R-134a in small circular tubes. International Journal of Multiphase Flow
59, 2535–2545. 27, 1163–1177.

You might also like