Вы находитесь на странице: 1из 7

Chemical Engineering Science 155 (2016) 38–44

Contents lists available at ScienceDirect

Chemical Engineering Science


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

Gas–liquid mass transfer in a falling film microreactor: Effect of reactor


orientation on liquid-side mass transfer coefficient
David Lokhat n, Ashveer Krishen Domah, Kuveshan Padayachee, Aman Baboolal,
Deresh Ramjugernath
Reactor Technology Research Group, School of Engineering, University of KwaZulu-Natal, Durban 4041, South Africa

H I G H L I G H T S G R A P H I C A L A B S T R A C T

 Lower mass transfer rates in inclined


microchannels.
 Curved interface approximation
gives more realistic mass transfer
coefficients.
 Cellular convection in microchannels
responsible for enhanced mass
transfer.

art ic l e i nf o a b s t r a c t

Article history: Microreactors offer a unique platform for chemical syntheses and have been applied to numerous re-
Received 2 May 2016 action types including nitrations, fluorinations and hydrogenations. A key feature of falling film micro-
Received in revised form reactors is the comparably large specific surface area they afford compared to conventional reactors. The
20 July 2016
enhanced heat and mass transfer characteristics can be exploited for rapid and exothermic reactions.
Accepted 3 August 2016
Available online 4 August 2016
Adequate understanding of the mass transfer processes occurring within microchannels is necessary for
proper reactor design and optimization. In the current study the influence of reaction plate orientation
Keywords: and gas flowrate on liquid-side mass transfer coefficient was investigated via CO2 absorption experi-
Microreactor ments. Lower plate angles resulted in lower liquid-side mass transfer coefficients. At higher film velo-
Mass transfer
cities the rate of mass transfer was greater. The experimentally determined mass transfer coefficients
CO2 absorption
were at least twice as high as those predicted either by film or penetration theory. The enhancement in
Inclined plate
mass transfer is suggested to be due to cellular convection in the microchannels. For inclined reaction
plates, increasing the gas flowrate had a positive effect on the mass transfer characteristics due to in-
duced fluctuations of the gas–liquid interface.
& 2016 Elsevier Ltd. All rights reserved.

1. Introduction down a microstructured reaction plate under the action of gravity


and spreads to form an expanded thin film in each microchannel.
Microreactors are reactors with reaction channels of the order The gas flows over the liquid film contained within the micro-
of micrometers, at which diffusion is the dominant mixing me- channels. This design allows for specific gas–liquid interfacial areas
chanism rather than turbulent eddies. The interest in this tech- of up 20,000 m2/m3 to be achieved, which is noticeably higher
nology has been focused on fine and speciality chemical manu- than other gas–liquid contacting reactors (Hessel et al., 2000; Yue
facture (Fletcher et al., 2002). The falling film microreactor (FFMR) et al., 2007; Ziegenbalg et al., 2010). The high specific areas made
is the most popular type of gas–liquid contactor in this class of available by the reactor design allow for high heat and mass
reactors. Under normal operating conditions, the liquid flows transfer rates to be achieved which make FFMRs an efficient tool
for studying the kinetics of fast and highly exothermic gas–liquid
n
Corresponding author. reactions. Researchers have exploited the high heat and mass
E-mail address: lokhat@ukzn.ac.za (D. Lokhat). transfer rates of FFMRs for reactions such as catalysed

http://dx.doi.org/10.1016/j.ces.2016.08.002
0009-2509/& 2016 Elsevier Ltd. All rights reserved.
D. Lokhat et al. / Chemical Engineering Science 155 (2016) 38–44 39

Nomenclature QL liquid volumetric flowrate, m3  s 1


Re Reynold number
AL liquid cross sectional area in the microchannel, m2 t collection time, s
Cout liquid phase concentration of CO2 in the reactor outlet, Wch width of microchannel, m
mol  dm 3 We Weber number
Cn saturated concentration of CO2, mol  dm 3 VHCl volume of HCl dispensed between titration endpoints,
CHCl concentration of HCl, mol  dm 3 dm3
D CO2 diffusivity in the liquid phase, m2  s 1 x width dimension of microchannel
Fo Fourier number y depth dimension of microchannel
jL mean liquid velocity, m  s 1
kL liquid-side mass transfer coefficient, m  s 1 Greek symbols
kL, pen penetration model liquid-side mass transfer coeffi-
cient, m  s 1 δL thickness of falling film, m
kL, film film model liquid-side mass transfer coefficient, νL kinematic viscosity of liquid, m2  s 1

ms 1 θ angle of reaction plate, °


Larc arc length of parabola, m sL surface tension of liquid, N  m 1
Lch length of microchannel, m ρL density of liquid, kg  m 3
Nch number of channels

hydrogenation (Yeong et al., 2003), direct fluorination of aromatics fabrication imprecisions and liquid maldistribution lowered the
(Jähnisch et al., 2000) and direct synthesis of hydrogen peroxide reaction conversion by approximately 2%. However, none of these
(Inoue et al., 2007). previous studies have considered the effect of reactor orientation
The measurement of gas–liquid mass transfer in a falling film on the rate of mass transfer. The reactor orientation is considered
micro reactor has formed the basis for a number of published an additional operating parameter that can be used to adjust the
studies (Zanfir et al., 2005; Zhang et al., 2009). The reactive ab- performance of a FFMR. In particular it has been suggested that
sorption of carbon dioxide (CO2) and sulphur dioxide (SO2) in changes to the reactor orientation can be used to modify liquid
aqueous sodium hydroxide have both been used effectively to residence times for slower reactions (Hessel et al., 2008). There is a
characterize mass transfer efficiency in microreactors. For the need therefore to understand the effect of reactor orientation on
measurement of liquid phase mass transfer coefficients, CO2 ab- other operational characteristics of the apparatus, particularly the
sorption is preferred. The CO2 reacts rapidly with the NaOH at the rate of mass transfer.
gas–liquid interface and forms sodium carbonate. The analysis of In this work CO2 absorption measurements were used to
liquid reaction products can be done simply by titration and the quantify the effect of the degree of inclination of the reaction plate
reactants are inexpensive and non-toxic. For the measurement of and gas flowrate in a FFMR on the liquid-side mass transfer
gas phase mass transfer coefficient, the SO2 method is preferred. coefficient. For the specific geometry of the microchannels used in
Very low concentrations of the reactive gas-phase species must be this study, approximations for the shape and length of the liquid
used for measurement of the gas-phase mass transfer coefficient, menisci were used to compute the mass transfer coefficients.
to ensure that the process is limited by the gas phase mass transfer These results were compared to the case of a flat interface and the
(Commenge et al., 2011). An SO2 gas analyser is much more sen- relative differences were analysed.
sitive and selective at this concentration range compared to CO2
sensors which may encounter interference from the nitrogen
diluent. 2. Experimental
Applications of these methods are illustrated in the literature.
Zanfir et al. (2005) carried out an experimental and modelling The FFMR used in this study is a well-known and well-studied
study of the reactive absorption of carbon dioxide in a falling film microstructured reactor which was invented by the Institut für
microreactor. A two-dimensional model was formulated to re- Mikrotechnik Mainz GmbH and has successfully shown its ability
present the flow through the reactor and the experimental results to maintain thin liquid films of up to 100 mm in thickness (Al-
were compared to the model prediction in terms of carbon dioxide Rawashdeh et al., 2008). The FFMR consisted of a copper cooling
conversion. The model gave good agreement with the experi- plate, a stainless steel reaction plate with etched microchannels
mental data at low inlet NaOH concentration. Moreover, the and a viewing window, fabricated from glass, which enclosed the
parametric analysis of the model indicated that the major rate gas headspace. The device was fabricated from 316Ti stainless steel
limitation was on the liquid side and that the CO2 was consumed and was equipped with a microstructured panel containing 32
within a short distance from the gas–liquid interface. Zhang et al. microchannels of dimensions 641 mm  257 mm  66.4 mm. The
(2009) studied the hydrodynamics and mass transfer of gas–liquid experimental system is shown in Fig. 1.
flow in a FFMR. Once again the CO2 absorption method was em- Carbon dioxide gas (Afrox,499%) was supplied in a standard
ployed, with titration of the liquid phase product to determine the cylinder. The feed gas flowrate was controlled using a thermal
efficiency of the process. Al-Rawashdeh et al. (2008) developed a mass flow controller (Alicat). Inlet gas flowrates were fixed at ei-
three-dimensional computational fluid dynamic model of a falling ther 50 cm3  min 1 and 80 cm3  min 1. Deionized water (con-
film microreactor based on realistic channel geometry profiles. The ductivity 0.07 μS) was mixed with ethylene glycol (Riedel-de
model was validated experimentally by the absorption of CO2 in Haen, 499%) to produce a 12 wt% solution and was used as the
NaOH aqueous solution. The model allowed the authors to in- liquid feed. This ensured satisfactory wetting of the channel sur-
vestigate the effects of channel fabrication precision, liquid flow face due to a lower surface tension of the liquid. In all experiments
distribution, gas chamber height, and hydrophilic-hydrophobic the liquid flow into the microreactor was controlled using a HPLC
plate material on the conversion of CO2. It was found that pump. The feed pump supplies liquid to a header. The liquid
40 D. Lokhat et al. / Chemical Engineering Science 155 (2016) 38–44

Fig. 1. Experimental setup for CO2 absorption experiments.

overflows from the header through slits into the microchannels, 4Q L


ReL =
falling under gravity to the lower collection plate from where it is Wch νL (2)
withdrawn. The withdrawal pump is a peristaltic-type pump op-
erating at a flowrate higher than the inlet pump to ensure that For Reynolds numbers up to 25 the liquid flow is characterized
by a thin film with a smooth, flat interface and is referred to as the
there is no accumulation of fluid in the channels. Since the liquid
smooth laminar regime. Between Reynolds numbers of 25-1000
flow is gravity driven the high flow withdrawal pump does not
(laminar flow with rippling surface) the flow is characterized by a
completely empty the channel. The temperature in the micro-
wavy interface with fluctuating film thickness. Turbulent flow in
reactor (298 K) was maintained using an external water bath cir-
falling liquid films is usually encountered above a Reynolds
culator. In a typical experiment, the system was initially flooded
number of 1000. Flow patterns in FFMRs can be characterized as
with liquid. The withdrawal pump was then activated and the inlet
either corner rivulet flow, falling film with dry patches or complete
flow was set at the test value using the HPLC pump until a uniform falling film flow. The latter is the most appropriate for gas–liquid
liquid film was formed within each microchannel. The gas was mass transfer and reaction. For complete falling film flow char-
admitted and the outlet liquid was directed to a beaker containing acterized by a continuous liquid film without dry patches, laminar
a solution of NaOH (0.1 mol  dm 3). The CO2 absorbed by the flow conditions do prevail according to Nusselt theory. The validity
water was converted to Na2CO3. The experiment was allowed to of these regimes have been confirmed experimentally (Commenge
proceed for 20 min. The contents of the beaker was then titrated et al., 2011; Zhang et al., 2009).
against a 0.1 mol  dm 3 solution of HCl. The outlet liquid phase The first study of falling liquid films was conducted by Nusselt
concentration of CO2 was then given by: (1916). He assumed a steady state system, a flat interfacial region
and rectilinear flow (i.e. all streamlines are straight) and con-
VHCl CHCl
Cout = structed a shell momentum balance over a thin layer of the fluid.
Q Lt (1)
In the limit, the thickness of this layer was allowed to approach
where VHCl is the volume of HCl dispensed between the titration zero in order to generate a differential equation which could be
endpoints (representing hydroxide and carbonate species) and t is integrated to represent the momentum flux. The differential
the collection time (Zhang et al., 2009). Special wooden cradles at equation was solved to give the velocity profile of the fluid. The
60° and 75° were fabricated to hold the reactor unit for the in- boundary conditions used to solve these equations for falling li-
clined plate experiments. Each experiment was repeated three quid films were: (i) the velocity of the fluid at the wall must be
times and average results are presented for each set of conditions. zero relative to the velocity of the wall and (ii) the momentum
flux, or shear stress, and velocity gradient must be zero at the li-
quid-gas interface. By following the steps indicated above for a
falling film, equations describing the velocity profile, wall shear
3. Results and discussion stress and film thickness were derived. The flow through micro-
sized channels is characterized by a very low Reynolds number
Falling liquid films may be described through the use of three and is generally considered laminar. Commenge et al. (2011) state
progressive regimes, based on the Reynolds number of the flow. In that the film thickness in a FFMR can be related to the liquid
the original work on the subject by Nusselt (1916) such flow was flowrate, liquid properties and geometric parameters according to
restricted to flat plates and correction factors are introduced for Nusselt condensation theory:
structured plates. The Reynolds number is defined as four times
3Q L ν L
the volumetric flow rate per unit width of channel divided by the δL = 3
Nch Wch g (3)
kinematic viscosity.
D. Lokhat et al. / Chemical Engineering Science 155 (2016) 38–44 41

gδL2 channel grooves were thus assumed to have an elliptical shape


jL = and the liquid meniscus the form of a parabolic arc (cf. Fig. 2).
3ν L (4)
Based on the locus of the parabolic vertex and the radii of the
Zhang et al. (2009) also propose the use of these equations to ellipse the following functions were developed:
relate the liquid flowrate, mean velocity and film thickness. For
⎡ ⎤
exactly the same type of reactor as that used in this study, Al- ⎢ Dch − δL ⎥ ⎛ W ⎞2
Rawashdeh et al. (2008) state that the liquid flow is solely driven y1 = ⎢ 2⎥
× ⎜ x − ch ⎟ + δL
⎝ 2 ⎠
by gravity and found that Eqs. (3) and (4), developed for a free (
⎢⎣ Wch − Wch
2 ) ⎥⎦
(6)
falling film over a flat plate, are also valid for flow in the micro-
channels at liquid flowrates below 10 ml  min 1.

( x − ) ⎤⎥⎥
0.5
Wch 2
Liquid films in microchannels can exhibit morphologies other ⎢ 2 2
than the flat profile. Due to the action of capillary forces, the y2 = ⎢ 1 − × Dch + Dch
Wch 2
surface of the liquid film in microchannels usually takes the form ⎢⎣ ( ) ⎥⎦
2 (7)
of a flowing meniscus rather than a completely flat film (Zhang
et al., 2010). The shape of the film depends on the wetting ability In Fig. 2, y1 is the function representing the parabolic meniscus
of the liquid and aspect ratio of the channels (Al-Rawashdeh et al., and y2 is the function representing the elliptical channel. The area
2008). Under favourable conditions (low contact angle, low liquid in between these two curves (the liquid flow area) is given by:
flow-rate and proper aspect ratio) the liquid will form a meniscus Wch
that is pinned at the edges of the groove. The Weber number AL = ∫0 ( y1 − y2 ) dx
(8)
characterizes the ratio of liquid kinetic energy to interfacial en-
ergy. The volumetric flowrate through the microchannel is:

ρL δL jL2 Q L = jL AL Nch (9)


We =
σL (5) where the value of jL can be obtained from Eq. 3. For those ex-
3
At a sufficiently low Weber number ( o10 ), the interfacial periments that were conducted at plate orientations other than
energy dominates and a meniscus in the shape of a concave cir- 90°, the angular component of the gravitational acceleration was
cular arc can be assumed (Al-Rawashdeh et al., 2008). The mi- used to correct the velocity:
croreactor used in this study was equipped with an etched reac- sin θ × gδL2
tion plate consisting of a number of rounded microgrooves. The jL =
3ν L (10)
Weber number for each experiment (cf. Table 1) was calculated
based on the fluid properties and the film thickness and velocity At a completely horizontal configuration the liquid velocity
from Eqs. 3 and 4. In all cases the Weber number was below the becomes zero since the liquid from the pump merely fills the
critical value and a curved meniscus was extant. The film thickness header without overflow into the microchannels. Eventually when
and liquid velocity were then determined by an alternative enough liquid has accumulated in the header it will begin to
method based on approximate geometries of the channel and film spread into the microchannels and the gas cavity. This is an
meniscus. Anastasiou et al. (2013) carried out a systematic in- anomalous situation that is not encountered during normal op-
vestigation of the liquid phase hydrodynamics in FFMRs. The effect eration of the device, i.e. the basis for operation of the FFMR is that
of various fluid parameters as well as geometrical characteristics there is some inclination which induces flow through the micro-
and inclination angle of the conduit on the geometry of the liquid channels after overflow from the header.
phase was studied. For all cases it was found that the shape of the The film thickness is not known a priori. For each of the three
meniscus could be described accurately using a quadratic equation volumetric flowrates used in this study, the film thickness is found
representing a parabolic arc. A true circular arc is not possible from the simultaneous solution of Eqs. (6)–(10). That is, the vo-
since the meniscus is pinned to the outer edges of the channel and lumetric flowrate governs the thickness of the film and the velo-
the width of the channel is much greater than the depth. The city. The results are presented in Table 1. Fig. 3 shows the liquid

Table 1
Experimental results for gas–liquid mass transfer experiments a.

Experiment 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18

Plate angle / ° 90 90 90 90 90 90 75 75 75 75 75 75 60 60 60 60 60 60
QL /cm3  min 1 3 4 5 3 4 5 3 4 5 3 4 5 3 4 5 3 4 5
QG /cm3  min 1 50 50 50 80 80 80 50 50 50 80 80 80 50 50 50 80 80 80

Flat interface
jL / m  s 1,  102 2.60 3.15 3.66 2.60 3.15 3.66 2.51 3.04 3.53 2.51 3.04 3.53 2.25 2.73 3.17 2.25 2.73 3.17
δL / m,  104 0.94 1.03 1.11 0.94 1.03 1.11 0.94 1.03 1.11 0.94 1.03 1.11 0.94 1.03 1.11 0.94 1.03 1.11
We,  104 2.35 3.79 5.50 2.35 3.79 5.50 2.19 3.54 5.13 2.19 3.54 5.13 1.76 2.84 4.13 1.76 2.84 4.13
Fo 0.48 0.32 0.24 0.48 0.32 0.24 0.49 0.34 0.25 0.49 0.34 0.25 0.55 0.38 0.28 0.55 0.38 0.28
kL / m  s 1,  105 5.73 6.41 6.65 5.76 6.48 6.75 5.20 5.86 6.09 5.46 6.12 6.45 5.13 5.56 5.87 5.26 5.89 6.09

Curved interface
jL / m  s 1,  103 7.16 9.70 12.3 7.16 9.70 12.3 7.17 9.72 12.3 7.17 9.72 12.3 7.22 9.79 12.4 7.22 9.79 12.4
δL / m,  105 4.92 5.72 6.44 4.92 5.72 6.44 5.01 5.83 6.57 5.01 5.83 6.57 5.30 6.18 6.96 5.30 6.18 6.96
Larc / m,  104 7.91 7.81 7.72 7.91 7.81 7.72 7.90 7.80 7.71 7.90 7.80 7.71 7.86 7.75 7.66 7.86 7.75 7.66
Fo 6.29 3.43 2.13 6.29 3.43 2.13 6.05 3.30 2.05 6.05 3.30 2.05 5.36 2.91 1.81 5.36 2.91 1.81
kL / m  s 1,  105 4.64 5.26 5.52 4.67 5.32 5.60 4.22 4.81 5.07 4.43 5.03 5.36 4.18 4.60 4.91 4.29 4.87 5.10
kL, pen / m  s 1,  105 1.84 2.14 2.41 1.84 2.14 2.41 1.84 2.14 2.41 1.84 2.14 2.41 1.84 2.15 2.42 1.84 2.15 2.42
kL, film / m  s 1,  105 3.34 2.87 2.54 3.34 2.87 2.54 3.28 2.81 2.50 3.28 2.81 2.50 3.09 2.65 2.36 3.09 2.65 2.36

a 7 1
Average deviation on calculated mass transfer coefficients from repeated experiments was 72  10 ms
42 D. Lokhat et al. / Chemical Engineering Science 155 (2016) 38–44

Fig. 2. Approximations of channel and film geometries.

cross sectional profile in the channel at each of the three flowrates where L arc is the arc length of the parabola representing the top of
for the vertically oriented plate. In all cases the calculated film the liquid film:
thickness for the curved interface was lower than that for the flat Wch
interface. The film profiles obtained using these approximations L arc = ∫0 1 + y1 ′(x)2
(12)
compare favourably with the results of Al Rawashdeh et al. (2008).
The liquid-side mass transfer coefficient was calculated based Thus,
on the method developed by Zhang et al. (2009). A material bal-
QL ⎛ C* ⎞
ance over an elementary shell of the falling film for a flat interface kL = ln ⎜ ⎟
L arc × L ch ⎝ C * − Cout ⎠ (13)
yields the following:

QL ⎛ C* ⎞ Calculations showed that on average there is a 22% difference in


kL = ln ⎜ ⎟ the mass transfer coefficients obtained using the flat interface and
Wch L ch ⎝ C − Cout ⎠
* (11)
curved interface approximations.
For a curved meniscus the interfacial area is given by L arc × L ch , Fig. 4 shows the calculated liquid side mass transfer coefficients

Fig. 3. Liquid cross sectional profile in the channel for flowrates of (a) 3 cm3  min 1
, (b) 4 cm3  min 1
and (c) 5 cm3  min 1
. Reaction plate oriented at 90°.
D. Lokhat et al. / Chemical Engineering Science 155 (2016) 38–44 43

Fig. 5. Liquid-side mass transfer coefficient as a function of liquid film velocity for
Fig. 4. Liquid-side mass transfer coefficient as a function of liquid film velocity, all
vertically oriented reaction plate.
plate orientations.

at various liquid velocities, gas flowrates and plate angles. A de- trend showed more similarity to the penetration model predic-
crease in the plate angle gave lower mass transfer coefficients. tions than the film model. Similar results were obtained in the
There was also a distinct difference in the mass transfer coeffi- literature for CO2 absorption in a microreactor containing rectan-
cients at high and low gas flowrates for 60° and 75° angles. When gular microchannels (Zhang et al., 2009). Zhang et al. (2009)
the reactor was oriented vertically there was no significant dif- suggested that the discrepancy in the calculated and predicted
ference in the mass transfer coefficients obtained at the two gas mass transfer coefficients was due to the presence of a wavy gas–
velocities. liquid interface and a protracted liquid residence time as a result of
Several mass transfer models have been proposed to describe strong side-wall shear stress in the microchannels. Sobieszuk et al.
the absorption of gas into a liquid film (King, 1966). Zhang et al. (2010) on the other hand presented evidence of mass transfer
(2009) consider the film and penetration theories of interfacial enhancement through cellular convection (Marangoni effect) in
mass transfer for theoretical validation of the experimental results the microchannels of a falling film microreactor. This cellular
specifically due to the presence of a laminar flow regime. The film convection phenomenon gives behaviour similar to the motion of
theory assumes a stagnant film exists near the interface. In the turbulent eddies that is proposed to be the principle mechanism of
penetration theory, the interfacial region is imagined to be a very interphase mass transfer underlying the penetration theory, but in
thick film continuously generated by flow, which may be laminar. a much more intensified manner. This may be the reason why the
For both of these theories turbulence in the bulk is not a requisite. experimental data shows an increase in mass transfer coefficient
The surface renewal theory on the other hand requires both a thick as the liquid velocity increases. The enhancement in mass transfer
interfacial film region and a large well mixed bulk region. The due to cellular convections appears to be highest when the reac-
latter theory was therefore not considered in the analysis of the tion plate is vertically oriented. As the plate angle is reduced the
experimental results. effect is less pronounced, resulting in a lower mass transfer
The Fourier number, the ratio of liquid residence time to dif- coefficient.
fusion time, can be used to determine the applicability of a par- As shown by Zhang et al. (2009) the effect of gas flowrate on
ticular model. liquid-side mass transfer coefficient is not significant for a verti-
cally oriented plate (cf. Fig. 5). However the results of this study
L ch D also show that at lower plate angles distinctly higher mass transfer
Fo =
jL δL2 (14) coefficients are obtained at higher gas flowrates (cf. Fig. 4). This
can only be connected to the disturbance of the liquid interface as
If the Fourier number is much less than 1, then the penetration
induced by the higher gas flowrate and hence an increase in the
model is applicable and the mass transfer in the falling film is far
wavy region of the falling film (Zhang et al., 2009; Dukler, 1977).
from equilibrium. On the other hand if the Fourier number is much
greater than 1, then sufficient time is available to reach equili-
brium and the film model is applicable (Zhang et al., 2009). The
4. Conclusions
Fourier number was calculated for each data point using the flat
interface and curved interface approximations. For the former, all
CO2 absorption experiments in an FFMR have shown that the
Fourier numbers were lower than 1 whereas for the realistic liquid
angle of orientation of the reaction plate has a direct influence on
profiles the Fourier numbers were greater than 1 implying that the
the rate of mass transfer, i.e. lower plate angles give lower liquid-
film model was applicable. Mass transfer coefficients were pre-
side mass transfer coefficients. Mass transfer appears to proceed
dicted based on each of these fundamental models:
according to the penetration model but at significantly enhanced
1.5jL D rates. These observations appear to be the result of cellular con-
kL, pen = 2 vection which is more pronounced in the vertically oriented re-
πL ch (15)
action plate. Approximations for the channel and meniscus geo-
D metries based on elliptic and parabolic functions have yielded film
kL, film = thicknesses and velocities 50% lower than the flat interface ap-
δL (16)
proximation. These values compare favourably with literature (Al-
The results of these calculations are presented in Table 1 and Rawashdeh et al., 2008).
illustrated in Fig. 5 for the vertically oriented reaction plate. The effect of gas flowrate on liquid-side mass transfer coeffi-
Both models underpredicted the mass transfer coefficient by at cient is negligible at 90° plate angle but a significant difference is
least half the experimental value. Interestingly, the experimental observed over a narrow range of gas flowrates for 75° and 60°
44 D. Lokhat et al. / Chemical Engineering Science 155 (2016) 38–44

plate angles. Increased gas flow at lower plate angles has a Zhang, X., 2002. Microreactors: principles and applications in organic synthesis.
stronger effect on the hydrodynamics of the liquid phase, inducing Tetrahedron 58, 4735–4757.
Hessel, V., Ehrfeld, W., Herweck, T., Haverkamp, V., Lowe, H., Schiewe, J., Wille, C.,
rippling at the gas–liquid interface and extending the wavy in- Kern, T., Lutz, N., 2000. Gas/liquid microreactors: hydrodynamics and mass
terface region. transfer. In: Proceedings of the Fourth International Conference on Micro-
reaction Technology, Atlanta, USA, 5–9 March, pp. 174–186.
Hessel, V., Löb, P., Löwe, H., 2008. Gas-liquid reactions. In: Wirth, T. (Ed.), Micro-
reactors in Organic Synthesis and Catalysis. Wiley-VCH, Weinheim, p. 142.
Acknowledgement Jähnisch, K., Baerns, M., Hessel, V., Ehrfeld, W., Haverkamp, V., Loewe, H., Wille, C.,
Guber, A., 2000. Direct fluorination of toluene using elemental fluorine in gas/
This work was supported by the South African Research Chairs liquid microreactors. J. Fluor. Chem. 105, 117–128.
King, C.J., 1966. Turbulent liquid phase mass transfer at free gas-liquid interface.
Initiative of the Department of Science and Technology and the
Ind. Eng. Chem. Fund. 5, 1–8.
National Research Foundation. This work is based on the research Nusselt, W., 1916. Die oberflachenkondensation des wasserdamphes. VDI For-
supported in part by the National Research Foundation of South chungscheft 60, 541.
Africa for the grant, Unique Grant No. 99339. Sobieszuk, P., Pohorecki, R., Cygański, P., Kraut, M., Olschewski, F., 2010. Marangoni
effect in a falling film microreactor. Chem. Eng. Sci. 164, 10–15.
Yeong, K.K., Gavriilidis, A., Zapf, R., Hessel, V., 2003. Catalyst preparation and de-
activation issues for nitrobenzene hydrogenation in a falling film reactor. Catal.
References Today 81, 641–651.
Yue, J., Chen, G., Yuan, Q., Luo, L., Gonthier, Y., 2007. Hydrodynamics and mass
transfer characteristics in gas–liquid flow through a rectangular microchannel.
Anastasiou, A.D., Gavriilidis, A., Mouza, A.A., 2013. Study of the hydrodynamic Chem. Eng. Sci. 62, 2096–2108.
characteristics of a free flowing liquid film in open inclined microchannels. Zanfir, M., Gavriilidis, A., Wille, C., Hessel, V., 2005. Carbon dioxide absorption in a
Chem. Eng. Sci. 101, 744–754.
falling film microstructured reactor: experiments and modeling. Ind. Eng.
Al-Rawashdeh, M., Hessel, V., Lob, P., Mevissen, K., Schonfeld, F., 2008. Pseudo 3-D
Chem. Res. 44, 1742–1751.
simulation of a falling film microreactor based on realistic channel and film
Zhang, H., Chen, G., Yue, J., Yuan, Q., 2009. Hydrodynamics and mass transfer of
profiles. Chem. Eng. Sci. 63, 5149–5159.
gas–liquid flow in a falling film microreactor. AIChE J. 55, 1110–1120.
Commenge, J.M., Obein, T., Framboisier, X., Rode, S., Pitiot, P., Matlosz, M., 2011. Gas-
Zhang, H., Yue, J., Chen, G., Yuan, Q., 2010. Flow pattern and break-up of liquid film
phase mass-transfer measurements in a falling-film microreactor. Chem. Eng.
Sci. 66, 1212–1218. in single-channel falling film microreactors. Chem. Eng. J. 163, 126–132.
Dukler, A.E., 1977. The role of waves in two-phase flow: some new understandings. Ziegenbalg, D., Löb, P., Al-Rawashdeh, M.M., Kralisch, D., Hessel, V., Schönfeld, F.,
Chem. Eng. Ed. 11, 108–117. 2010. Use of ‘smart interfaces’ to improve the liquid-sided mass transport in a
Fletcher, P.D.I., Haswell, S.J., Pombo-Villar, E., Warrington, B.H., Watts, P., Wong, Y.F., falling film microreactor. Chem. Eng. Sci. 65, 3557–3566.

Вам также может понравиться