You are on page 1of 7

chemical engineering research and design 8 8 ( 2 0 1 0 ) 263–269

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

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

Bubble lengths in the gas–liquid Taylor flow in


microchannels

Paweł Sobieszuk ∗ , Paweł Cygański, Ryszard Pohorecki


Faculty of Chemical and Process Engineering, Warsaw University of Technology, Warynskiego 1, PL 00-645 Warsaw, Poland

a b s t r a c t

Experimental results of measurements of the bubble and slug lengths in Taylor (slug) flow are presented. The exper-
iments were carried out using 3 different straight microchannels (microreactor with square cross-section made of
polydimethyloxosilane (PDMS); microreactor with circular cross-section made of glass; microreactor with rectangu-
lar cross-section made of polyethylene terephthalate modified by glycol (PETg)) and 4 different liquids (water, ethanol
propanol and heptane). The results have been compared with the available literature correlations. It is concluded,
that the values obtained from the correlation proposed by Laborie et al. [Laborie, S., Cabassud, C., Durant-Bourlier, L.,
Laine, J.M., 1999. Characterization of gas–liquid two-phase flow inside capillaries. Chem Eng Sci 54, 5723–5735] do not
agree with the results of measurements, while the agreement of these results with the predictions obtained using the
correlation proposed by Qian and Lawal [Qian, D., Lawal, A., 2006. Numerical study on gas and liquid slugs for Taylor
flow in a T-junction microchannel. Chem Eng Sci 61, 7609–7625] is good. New, corrected values of the pre-exponential
constant and the exponents in the Qian and Lawal [Qian, D., Lawal, A., 2006. Numerical study on gas and liquid slugs
for Taylor flow in a T-junction microchannel. Chem Eng Sci 61, 7609–7625] correlation are proposed.
© 2009 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.

Keywords: Microreactor; Bubble length; Liquid slug length; Taylor flow

1. Introduction Closed microchannels are microtunnels through which


both contacted phases flow simultaneously.
Modern structured process equipment usually involves Closed microchannels are much more commonly used
microstructures. Microstructures ensure significant inten- in microstructures, the flow structure in these channels is
sification of mass and heat transfer owing to short more complicated and definitely less understood than the
diffusion/conduction paths, large concentration/temperature flow structure in the open channels. Consequently, the closed
gradients and high surface to volume ratio. These features channels have been chosen as the main object of investigation
warrant intensification of physical, chemical and biochemical in our investigation program.
processes, especially in multiphase (gas–liquid, liquid–liquid, This paper concerns two-phase gas–liquid flow in closed
gas–solid, liquid–solid or gas–liquid–solid) systems. microchannels. A number of different flow patterns (regimes)
The basic elements of microstructures are microchannels. may be distinguished in the case of such flow. They
There are two kinds of microchannels used to contact two include:
different phases:
• Bubble flow.
• Open microchannels. • Slug (Taylor) flow.
• Closed microchannels. • Annular flow.
• Spray flow.
Open microchannels have the form of grooves in a flat • Foam flow.
plate, opened from one side to ensure contact with the other
phase. and transition regimes between those listed above.


Corresponding author. Tel.: +48 22 234 63 19.
E-mail address: sobieszuk@ichip.pw.edu.pl (P. Sobieszuk).
Received 12 December 2008; Received in revised form 5 July 2009; Accepted 7 July 2009
0263-8762/$ – see front matter © 2009 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
doi:10.1016/j.cherd.2009.07.007
264 chemical engineering research and design 8 8 ( 2 0 1 0 ) 263–269

rates. It also seems to be the most suitable regime to carry out


Nomenclature the gas–liquid reactions.
To quote the opinion of Hessel et al. (2005): “Bubble and
A, B, C, D, E constants
slug lengths in microchannels mainly depend on the inlet con-
Ca = uTP L−1 capillary number
ditions, as the significance of surface tension forces means
d channel (hydraulic) diameter (m)
that once the Taylor bubbles form, little change to their size
Eo = (L − G )d2 gL−1 Eotvos number
is expected within the channel as a result of breakup or
LG bubble length (m)
coalescence. Although there is abundant literature on the for-
LL liquid slug length (m)
mation of small bubbles, usually in an unconfined liquid, very
Re = uTP dL −1
L Reynolds number for two-phase flow
little information is available on Taylor-bubble formation in
Re(uG ) = uG dL −1
L Reynolds number in Laborie et al.
microchannels”.
(1999) correlation
Indeed, only a few references describing the bubble and
Re(uS ) = uS dL −1
L Reynolds number in Laborie et al.
slug lengths have been found in the literature. Laborie et al.
(1999) correlation
(1999) measured bubble and slug lengths in vertical glass cap-
u superficial velocity (m s−1 )
illaries of 1, 2, 3 and 4 mm inner diameter. They proposed the
uS gas bubble velocity (m s−1 )
following correlations.
uTP = uG + uL superficial two-phase velocity (m s−1 )
For the bubble length:
We = u2TP d −1 Weber number
0.63
LG [Re(uS )]
Greek letters = 0.0878 (1)
d Eo1.26
ε hold-up
 viscosity (Pa s) where from:
 density (kg m−3 )
 surface tension (N m−1 ) LG ∼d−0.89 (2)
 diameter (m)
For the slug length:
Subscripts
G gas −1.27
LL [Re(uG )]
L liquid = 3451 (3)
d Eo1.27

where from:
The flow pattern depends on the liquid and gas flow rates,
on the properties of the fluids, channel geometry and its LL ∼d−2.81 (4)
material. Several diagrams, called flow maps, have been pro-
posed in the literature, to depict regions in which a given It follows from these correlations, that LG should be
flow pattern occurs (e.g. Jayawardena et al., 1997; Triplett et approximately inversely proportional to d while LL should be
al., 1999; Waelchli and von Rohr, 2006). An example of such a approximately inversely proportional to the third power of d.
map is shown in Fig. 1. There were also numerous efforts to In another study, Qian and Lawal (2006) performed numer-
develop an universal flow map, valid for different microchan- ical simulations of the Taylor flow, using commercial CFD
nels and different gas–liquid systems. Although these efforts package FLUENT (release 6.1.22.2003). In this package the
had limited success, they all confirm, that the slug (Taylor) volume of fluid (VOF) model is used to simulate two-phase
flow occupies a large part of any flow map, and is obtainable fluid–fluid flows. As a result of 148 numerical simulations for
in the most practically interesting ranges of gas and liquid flow channel widths ranging from 0.25 to 1 mm, they proposed the
following correlations:

(LG + LL ) −1.05
= 1.637ε−0.893
G (1 − εG ) Re−0.075 Ca−0.0687 (5)
d

LG −1.05
= 1.637ε0.107
G (1 − εG ) Re−0.075 Ca−0.0687 (6)
d

LL −0.05
= 1.637ε−0.893
G (1 − εG ) Re−0.075 Ca−0.0687 (7)
d

They also concluded, that for bigger channel diameters (2.0


and 3.0 mm) the above correlations are not applicable.
Pohorecki and Kula (2008) tried to explain the mecha-
nism of gas bubble and liquid slug formation in Taylor flow.
They proposed a simple mechanistic “switching” mechanism,
which explains the form of the correlations proposed by Qian
and Lawal (2006).
It should be pointed out, that bubble and slug lengths are
of paramount importance for mass transfer and back mixing
in Taylor flow. The existing experimental or CFD-based cor-
Fig. 1 – Flow map (own results obtained for glass relations usually relate the mass transfer coefficient to the
microreactor and ethanol–nitrogen system). bubble and/or slug lengths (Bercic and Pintar, 1997; Vandu et
chemical engineering research and design 8 8 ( 2 0 1 0 ) 263–269 265

Table 1 – Characteristics of microreactor used.


Material Shape Channel dimensions (mm)

Width Height Length Hydraulic diameter

PDMS polydimethylsiloxan Y 0.15 0.15 25 0.15


Glass Y  = 0.4 100 0.4
PETg–polyethylene terephthalate modified by glycol Y 0.2 0.55 55 0.29

al., 2005; van Baten and Krishna, 2005). In Taylor flow, if the
liquid exhibits good wettability of the channel wall, the wall is
covered by a thin liquid layer, and the gas bubbles are sliding
over this “lubricating” layer (Schwartz et al., 1986; Bretherton,
1961). The interfacial gas–liquid area is therefore composed of
two parts: the lateral film part and the perpendicular cap part.
Pohorecki (2007) proposed a criterion for the activity of the lat-
eral part of the interfacial area in the mass transfer processes.
The criterion also includes bubble length in a microchannel.
A comprehensive paper on the hydrodynamics of bubble
formation in microbubble columns with single and multiple
channels was published by Haverkamp et al. (2006). Fol-
lowing an earlier paper by Haverkamp (2002), they report
a distribution of bubble size, observed both in single and
multiple channel columns—a phenomenon only occasionally
observed in our work. A very interesting study on CFD simula-
tions of hydrodynamics of slug flow in curved microchannels Fig. 2 – PDMS microreactor, Y shaped.
was published by Kumar et al. (2007). The authors deter-
mined numerically (using the control volume finite difference
method, CVFDM) the gas bubble and liquid slug length for both
straight and curved microchannels, and found good agree-
ment of their results for straight microchannels with Qian
and Lawal (2006) correlations. Bubble and slug lengths were
measurement experimentally (using photographic method) by
Warnier et al. (2008). The authors concentrated on the thick-
ness of the liquid film surrounding a bubble, and did not
report direct information on bubble and slug lengths. Compu-
tational modelling and results of experimental measurements
of the slug size were also reported for liquid–liquid slug flow by
Kashid et al. (2007), and Kashid and Agar (2007). The results for
liquid–liquid flow cannot, however, be directly compared with
those for gas–liquid flow. It should be pointed out, that both in
numerical simulations and experiments performed using the
photographic method, the bubble and slug lengths are usually
measured along the channel axis. Only in a few papers (e.g. Fig. 3 – Glass microreactor, Y shaped.
Heiszwolf et al., 2001) these parameters were measured using
the conductivity method, which makes the results difficult to (PETg) by micromachining method and is shown in Fig. 4.
compare directly with those obtained from the photographs. Four liquids were used for the investigations, namely: water,
In this paper we report experimental values of gas bubble propanol and ethanol (PETg microreactor), ethanol and hep-
and liquid slug lengths, obtained using 3 different microreac- tane (PDMS microreactor), ethanol (glass microreactor). The
tors and 4 different liquids, with nitrogen as the gas phase. physico-chemical properties of the liquids used are shown in
Table 2.
2. Experiments The experimental set-up is shown in Fig. 5. The gas (nitro-
gen) was supplied from a cylinder through a reducing valve
Three microreactors were used. The microreactors were
designed for visual investigations of the flow in two-phase
Table 2 – Physico-chemical properties of the liquid
gas–liquid system. Table 1 shows shapes, dimensions and employed.
materials of microreactors. The first microreactor with square
Density Viscosity Surface tension
cross-section was built using polydimethyloxosilane (PDMS)
(kg m−3 ) (Pa s) (N m−1 )
by photolithography method and it is shown in Fig. 2. The
second microreactor with circular cross-section was made of Ethanol 790 0.00119 0.0225
Heptane 685 0.00041 0.0203
glass (three capillaries were joined together) and is shown in
Propanol 804 0.00220 0.0239
Fig. 3. The third microreactor with rectangular cross-section
Water 998 0.00103 0.0729
was built of polyethylene terephthalate modified by glycol
266 chemical engineering research and design 8 8 ( 2 0 1 0 ) 263–269

Fig. 6 – Measurement of the bubble and slug lengths.


Fig. 4 – PETg microreactor, Y shaped.
of the bubble caps. In order to calculate the length of a cylinder
of equivalent volume, one has to subtract suitable correction
(about 2d/3) from the bubble length, and add similar correction
to the slug length. For every set of experimental conditions
about 100 bubbles and slugs lengths were measured.
The superficial gas velocity was in the range: 0.02–1.29 m/s
(PETg); 0.02–1.1 m/s (PDMS); 0.02–0.93 m/s (glass). The super-
ficial liquid velocity was in the range: 0.01–0.51 m/s (PETg);
0.05–0.7 m/s (PDMS); 0.04–0.48 m/s (glass).

3. Results and discussion

Microreactor polydimethyloxosilane (PDMS) was used


for measurements with the ethanol–nitrogen and
Fig. 5 – The experimental set-up: 1, gas cylinder; 2, heptane–nitrogen systems.
electronic gas flowmeter; 3, soap-bubble flow meter; 4, The glass microreactor was used for measurements with
microreactor; 5, syringe pump, 6, buffer tank; 7, rotameters; the ethanol–nitrogen system.
8, liquid cylinder; 9, high speed camera. Microreactor PETg (polyethylene terephthalate modified by
glycol) was used for measurements with the ethanol–nitrogen,
and a needle valve. The gas flow to the microreactor was mea- water–nitrogen and propanol–nitrogen systems.
sured by an electronic gas flowmeter (made by Brooks) and a Exemplary experimental data of bubble and slug lengths
“soap film” meter. The gas flowmeter measured the volume are shown in Fig. 7.
flow reduced to standard conditions, subsequently corrected All the results have been compared with the available lit-
for real pressure. Pressure was measured at the gas inlet and erature correlations. A comparison of the results obtained
ranged from 0.2 to 50 kPA. The liquid was supplied by a syringe using the glass and PDMS microreactors is shown in Fig. 8.
pump via a buffer tank (to damp flow oscillations) and rotame- As can be seen, the results of experiments are in good agree-
ters. To further eliminate oscillations a syringe with teflon ment with the predictions obtained from the correlation of
piston was used. At the outlet from the microreactor the flow Qian and Lawal (2006), and in complete disagreement with
of liquid was once more controlled by volumetric method. the predictions obtained using the correlation of Laborie et
The liquid slugs and gas bubbles lengths were measured as
well as their dependence on the gas and liquid flow rates. To
this end, the flow in the microreactor was photographed using
CMOS high speed camera (1200 hs made by pco). The system
allowed to record frames with the resolution 1280 × 1024 pix-
els and speed in the range from 500 to 4000 frames per second.
The pixel size in the pictures obtained was about 10 ␮m. The
microreactor was lit using the halogen lamp 500 W using light
diffuser. All the data were measured at room temperature.
To eliminate possible channel heating, the halogen lamp was
used only for several seconds.
To speed up the measurements, an automatic computer
system was developed. This system enables measurements
of the bubbles and slug lengths directly from the registered
(photographed or recorded) flow pictures (Fig. 6). It should be
pointed out, that the system measures the bubble and slug
lengths along the channel axis as it customary in the literature, Fig. 7 – Glass microreactor, bubble and slugs length for
and therefore does not account for the semi-spherical shape ethanol–nitrogen system and uL = 0.36 m/s.
chemical engineering research and design 8 8 ( 2 0 1 0 ) 263–269 267

Fig. 10 – CFD simulation of the Taylor flow in PETg


microreactor: volumetric gas flow, 8.2 × 10−9 m3 /s;
volumetric liquid flow, 3.3 × 10−8 m3 /s.

this reason the results obtained for the PET microreactor have
been excluded from the further analysis.
The good agreement between our results and the Qian and
Lawal (2006) correlation convinced us, that the form of this
correlation is suitable for the description of Taylor flow in
microchannels. Further support of this form of correlation fol-
Fig. 8 – Comparison of the experimental results of bubble lows from the “switching” mechanism, proposed by Pohorecki
length and those calculated from Qian and Lawal (2006) and and Kula (2008). This form of correlation was thus used to
Laborie et al. (1999) correlations for PDMS and glass describe our results. Table 3 shows the values of the pre-
microreactor. exponential constants and exponents obtained for the results
from the PDMS and the glass microreactors. The correlations
were obtained using the least squares method.
For the PDMS reactor and heptane–nitrogen system, we
obtained the correlation:

LG −0.946
= 1.301ε0.086
G (1 − εG ) Re−0.142 Ca−0.152 (8)
d

For the PDMS reactor and ethanol–nitrogen system, we


obtained the correlation:

LG −1.009
= 1.298ε0.106
G (1 − εG ) Re−0.009 Ca0.013 (9)
d

For the glass reactor and ethanol–nitrogen system, we


obtained the correlation:

LG −1.076
= 1.306ε0.028
G (1 − εG ) Re−0.143 Ca−0.186 (10)
d

Figs. 12–14 show comparisons of experimental LG values with


those calculated from the correlations ((8)–(10)), respectively.
Fig. 9 – Comparison of the bubble length LG calculated from
The agreement is very good. For the results shown in Fig. 12
Qian and Lawal correlation with the experimental values
obtained in the PDMS and glass microreactor.

al. (1999). Detailed comparisons of the results obtained with


individual microreactors for individual systems with the pre-
dictions obtained using the correlation of Qian and Lawal
(2006) are shown in Fig. 9. All the comparisons show rea-
sonable (in the case of the glass microreactor—very good)
agreement with the predictions obtained using the Qian and
Lawal (2006) correlation. Larger differences are observed for
the PET microreactor. To find out the possible reasons of this
disagreement numerical simulations of the bubble formation
and flow in this microreactor were performed using FLUENT
6.3 CFD software (volume of fluid method). These simulations,
described in detail elsewhere (Rudniak and Pohorecki, 2007)
revealed that, due to the high value of the height to width ratio
of the PET channel, the channel was not always completely
filled by the gas bubbles, in some cases the bubbles even clung Fig. 11 – CFD simulation of the Taylor flow in PETg
to the upper wall of the channel, giving the impression of Tay- microreactor: volumetric gas flow, 2.2 × 10−9 m3 /s;
lor flow when photographed from the top (Figs. 10 and 11). For volumetric liquid flow, 3.3 × 10−8 m3 /s.
268 chemical engineering research and design 8 8 ( 2 0 1 0 ) 263–269

Table 3 – Correlation coefficients.


C
LG /d = AεBG (1 − εG ) ReD CaE A B C D E

Qian and Lawal correlation (6) 1.637 0.107 −1.05 −0.075 −0.0687
PDMS reactor (heptane–nitrogen) 1.301 0.086 −0.946 −0.142 −0.152
PDMS reactor (ethanol–nitrogen) 1.298 0.106 −1.009 −0.009 0.013
Glass reactor (ethanol–nitrogen) 1.306 0.028 −1.076 −0.143 −0.186
Proposed general correlation 1.302 0.073 −1.010 −0.098 −0.108

Fig. 14 – Comparison of the bubble length LG calculated


Fig. 12 – Comparison of the bubble length LG calculated
from correlation (10) with the experimental values obtained
from correlation (8) with the experimental values obtained
in the glass microreactor, ethanol–nitrogen system.
in the PDMS microreactor, heptane–nitrogen system.

multiple regression method produces a number of set of con-


stants and exponents, which describe the results with similar
precision, but are quite different.
Fig. 15 shows a comparison of the experimental bubble
lengths (for all the systems) with the values calculated using
correlation (11). The agreement is good. It may be noticed, that
the exponents on the Reynolds and capillary numbers are very
close. This suggests, that the proper parameter to be used in
the correlation is the Weber number:

We = Re · Ca

Taking this into account, and truncating the values of the


constants (as the results are only about 10% accurate), one can

Fig. 13 – Comparison of the bubble length LG calculated


from correlation (9) with the experimental values obtained
in the PDMS microreactor, ethanol–nitrogen system.

average deviation is 14.0%, for the results shown in Fig. 13 aver-


age deviation is 7.7%, for the results shown in Fig. 14 average
deviation is 5.8%.
On the basis of the above results, we propose a general
correlation with averaged values of the constants from the
correlations ((8)–(10)):

LG −1.01
= 1.302ε0.073
G (1 − εG ) Re−0.098 Ca−0.108 (11)
d

It has to be pointed out, that this procedure seems more


suitable that the multiple regression method performed using Fig. 15 – Comparison of the bubble length LG calculated
all the data available (i.e. the data obtained for all microchan- from correlation (11) with the experimental values obtained
nels and all liquids used). With a large number of variables, the in the PDMS and glass microreactors.
chemical engineering research and design 8 8 ( 2 0 1 0 ) 263–269 269

Bretherton, F.P., 1961, The motion of long bubbles in tube. J Fluid


Mech, 12: 2367–2371.
Haverkamp, V., Hessel, V., Löwe, H., Menges, G., Warnier, M.J.F.,
Rebrov, E.V., de Croon, M.H.J.M., Schouten, J.C. and Liauw,
M.A., 2006, Hydrodynamics and mixer-introduced bubble
formation in micro bubble columns with single and multiple
channels. Chem Eng Technol, 29: 1015–1026.
Haverkamp, V., 2002, Charakterisierung einer Mikroblasensäule
zur Durchführung stofftransportlimitierter und/oder
hoch-exothermer Gas/Flüssing Reactionen. Fortschritt,
Berichte VDI 771, Reiche 3 Verfahrenstechnik, 1–153.
Heiszwolf, J., Kreutzer, M.T., van den Eijnden, M.G., Kapteijn, F.
and Moulijn, J.A., 2001, Gas–liquid mass transfer of aqueous
Taylor flow in monoliths. Catal Today, 69: 51–55.
Hessel, V., Angeli, P., Gavriilidis, A. and Lőwe, H., 2005, Gas–liquid
and gas–liquid–solid microstructured reactors: contacting
principles and applications. Ind Eng Chem Res, 44: 9750–9769.
Jayawardena, S.S., Balakotaiah, V. and Witte, L., 1997, Flow
pattern transition maps for microgravity two-phase flows.
Fig. 16 – Comparison of the bubble length LG calculated AIChE J, 43: 1637–1640.
from correlation (12) with the experimental values obtained Kashid, M.N., Platte, F., Agar, D.W. and Turek, S., 2007,
in the PDMS and glass microreactors. Computational modeling of slug flow in a capillary
microreactor. J Comp Appl Math, 203: 487–497.
Kashid, M.N. and Agar, D.W., 2007, Hydrodynamics of
simplify the above correlation to the form: liquid–liquid slug flow capillary microreactor: flow regimes,
slug size and pressure drop. Chem Eng J, 131: 1–13.
LG −1.01 Kumar, V., Vashisth, S., Hoarau, Y. and Nigam, K.D.P., 2007, Slug
= 1.3ε0.07
G (1 − εG ) We−0.1 (12) flow in curved microreactors: hydrodynamic study. Chem Eng
d
Sci, 62: 7494–7504.
Fig. 16 shows a comparison of the experimental bubble lengths Laborie, S., Cabassud, C., Durant-Bourlier, L. and Laine, J.M., 1999,
(for all the microchannels and systems) and with those cal- Characterization of gas–liquid two-phase flow inside
capillaries. Chem Eng Sci, 54: 5723–5735.
culated from correlation (12). The agreement is also good.
Pohorecki, R., 2007, Effectiveness of interfacial area for mass
The average deviations for (11) and (12) have similar val-
transfer in two-phase flow in microreactors. Chem Eng Sci, 62:
ues: 17.2% and 16.8%, respectively. We recommend correlation 6495–6498.
(12) because of its simplicity. The correlation is valid in the Pohorecki, R. and Kula, K., 2008, A simple mechanism of bubble
ranges: We = 0.1–26; uG = 0.02–1.2 m s−1 ; uL = 0.004–0.7 m s−1 ; and slug formation in Taylor flow in microchannels. Chem
εG = 0.06–0.85. Eng Res Des, 86: 997–1001.
Qian, D. and Lawal, A., 2006, Numerical study on gas and liquid
slugs for Taylor flow in a T-junction microchannel. Chem Eng
4. Conclusions Sci, 61: 7609–7625.
Rudniak, L. and Pohorecki, R., 2007, Numerical simulation of a
• The experimental values of the bubble lengths in the Taylor gas bubble formation in a microreactor, In Proceedings of the
(slug) gas–liquid flow in microchannels are presented. XIX Polish Conference of Chemical and Process Engineering
• The results totally disagree with the predictions obtained Rzeszow, Poland, Part III, , pp. 183–186 (in Polish)
using the correlation proposed by Laborie et al. (1999). Schwartz, L.W., Princen, H.M. and Kiss, A.D., 1986, On the motion
of bubbles in capillary tubes. J Fluid Mech, 172: 259–275.
• The agreement of the results with the predictions obtained
Triplett, K.A., Ghiaasiaan, S.M., Abdel-Khalik, S.I. and Sadowski,
using the correlation proposed by Qian and Lawal (2006) is
D.L., 1999, Gas–liquid two-phase flow in microchannels. Part I.
generally good. The form of this correlation may be justified Two-phase flow pattern. Int J Multiphase Flow, 25: 377–394.
theoretically by the “switching” mechanism, proposed by van Baten, J.M. and Krishna, R., 2005, CFD simulations of wall
Pohorecki and Kula (2008). mass transfer for Taylor flow in circular capillaries. Chem Eng
• New, corrected values of the pre-exponential constant and Sci, 60: 1117–1126.
the exponents in the correlation of the form developed by Vandu, C.O., Liu, H. and Krishna, R., 2005, Mass transfer from
Taylor bubbles rising in single capillaries. Chem Eng Sci, 60:
Qian and Lawal (2006) are proposed.
6430–6437.
• A new, simplified correlation is proposed. Waelchli, S. and von Rohr, P.R., 2006, Two-phase flow
characteristics in gas–liquid microreactors. Int J Multiphase
References Flow, 32: 791–806.
Warnier, M.J.F., Rebrov, E.V., de Croon, M.H.J.M., Hessel, V. and
Schouten, J.C., 2008, Gas hold-up and liquid film thickness in
Bercic, G. and Pintar, A., 1997, The role of gas bubbles and liquid
Taylor flow in rectangular microchannels. Chem Eng J, 135S:
slug lengths on mass transport in the Taylor flow through
S153–S158.
capillaries. Chem Eng Sci, 52: 3709–3719.

You might also like