You are on page 1of 12

chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697

Contents lists available at ScienceDirect

Chemical Engineering Research and Design

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

Hydrodynamics and mass transfer in a tubular


reactor containing foam packings for
intensification of G-L-S catalytic reactions in
co-current up-flow configuration

Julien Lévêque a , Régis Philippe a,∗ , Marie-Line Zanota a , Valérie Meille a ,


Flavie Sarrazin b , Loïc Baussaron c , Claude de Bellefon a,∗
a Univ Lyon, CNRS, CPE Lyon, UCBL, Laboratoire de Génie des Procédés Catalytiques, UMR 5285, 43 bd du 11
novembre 1918, F-69616 Villeurbanne, France
b Laboratory of Future, UMR 5258, CNRS-Solvay, 178 av. du Dr Albert Schweitzer, F-33600 Pessac, France
c Solvay Research & Innovation Center, 85 av. des Frères Perret, F-69192 Saint-Fons, France

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

Article history: Stainless steel open-cell solid foams of various linear pore densities (20, 40, 60 PPI) are
Received 31 July 2015 wash-coated with a commercial catalyst and are evaluated as internals for G-L-S reactions
Received in revised form 8 March in co-current up-flow configuration. Hydrodynamics parameters such as liquid mean res-
2016 idence time, axial dispersion and pressure drop have been determined for an air/water
Accepted 14 March 2016 system with superficial velocities between 0.8 mm/s and 25 mm/s for liquid and between
Available online 19 March 2016 100 and 900 mm/s for gas. A generic piston-dispersion model represents well the liquid
hydrodynamics and is used to estimate axial Péclet number and mean residence time for
Keywords: this phase. The Péclet number appears to increase with liquid velocity and foam linear pore
Open cell solid foam (OCSF) density (5 < Pe < 60 for 20 PPI, 10 < Pe < 140 for 40 PPI and 60 < Pe < 200 for 60 PPI). Determined
Mass transfer liquid holdups (0.4 < εL < 0.8) are always higher than those encountered in conventional up-
Hydrodynamics flow fixed beds under comparable flow conditions. Liquid superficial velocity and foam linear
Gas-liquid up-flow pore density appear to be the most influent parameters while the gas superficial velocity
Catalytic hydrogenation possesses a less pronounced impact. Pressure drop measurements may indicate the exist-
ence of two different flow regimes (bubbly and pulsed regime) and globally the total pressure
drop remains low in the experimental domain tested with a maximum value of 0.2 bar/m,
comparable to literature data. The overall external mass transfer efficiency was determined
through the gas/liquid/solid catalytic reaction of ␣-methylstyrene (AMS) hydrogenation and
is compared to theoretical values obtained through correlations for conventional up-flow
fixed beds. Very high mass transfer coefficients, in the range of 0.2–0.9 s−1 were obtained at
low Re numbers, which is one order of magnitude higher than in up-flow fixed beds. A set of
correlations is derived for the calculation of the gas/liquid and liquid/solid contributions
to external mass transfer and allows explaining the unique bell-shaped Re dependence
displayed by the overall mass transfer coefficient.
© 2016 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.


Corresponding author.
E-mail addresses: regis.philippe@lgpc.cpe.fr (R. Philippe), claude.debellefon@lgpc.cpe.fr (C. de Bellefon).
http://dx.doi.org/10.1016/j.cherd.2016.03.017
0263-8762/© 2016 The Institution of Chemical Engineers. Published by Elsevier B.V. All rights reserved.
chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697 687

Nomenclature
 viscosity (Pa s−1 )
a specific surface area (m2 m−3 ) density (kg m−3 )
aGL effective gas–liquid specific surface area conductivity (S m−1 )
(m2 m−3 ) L surface tension (N m−1 )
aLS effective liquid–solid specific surface area contact time (s)
(m2 m−3 )
Bo Bodenstein number (–) Subscripts and superscripts
C molar concentration (mol m−3 ) 0 initial or inlet
C(t) time dependant tracer concentration in RTD * saturation
experiments (mol m−3 ) G gas
Csu hydrogen concentration at the external surface H hydrogen
of the catalyst (mol m−3 ) L liquid
dcell mean cell diameter (m) P catalyst particle or layer
dstrut mean strut diameter (m) su surface
Dm molecular diffusion coefficient in the liquid
mixture (m2 s−1 ) Abbreviations
Dax,L liquid axial dispersion coefficient (m2 s−1 ) AMS ␣-methylstyrene
Deff effective diffusivity of H2 in the catalyst layer SEM scanning electron microscopy
(m2 s−1 ) RTD residence time distribution
dR reactor diameter (m) PPI pore per inch
E(t) residence time distribution function (–)
Ea activation energy (J mol−1 )
F(t) cumulative residence time distribution func- 1. Introduction
tion (–)
k0 pre-exponential factor (mol s−1 g−1 Pd
) Since many years, the trend towards intensification leads to
KH adsorption equilibrium constant of hydrogen study and develop new kinds of structured catalysts and reac-
(m3 /mol) tors dedicated to G-L-S reactions with the aim to achieve better
kL gas–liquid mass transfer coefficient (m s−1 ) performance in momentum, heat and mass transfer as well
Kov overall external mass transfer coefficient (s−1 ) as potentially good production outputs (Cybulski and Moulijn,
kS liquid–solid mass transfer coefficient (m s−1 ) 2006). Several structured packings such as monoliths, Sulzer
L liquid specific mass flow-rate (kg s−1 m−2 ) type elements or fibres have been investigated as an alterna-
Lc coating characteristic length (m) tive to classical catalysts in form of rings, pellets or spherical
P pressure (bar) particles (Pangarkar et al., 2008). This has resulted into signif-
Pe Péclet number (–) icant improvements in terms of hydrodynamics, conversion
Q volumetric flow-rate (m3 s−1 ) and selectivity. While possessing high external specific sur-
r intrinsic reaction rate (mol s−1 g−1
Pd
) face areas, these new supports often exhibit a poor internal
r̄p apparent reaction rate (mol s−1 m3p−1) porosity and a high voidage that limits the catalyst hold-up.
R perfect gas constant (J mol−1 K−1 ) This specific point can be considered as the major drawback in
ReG gas Reynolds number (–) comparison with packed beds and restricts these new reactors
ReL liquid Reynolds number (–) to quite demanding and fast chemistries where mass, heat
ScL liquid Schmidt number (–) and momentum transfer issues are the main bottlenecks to
Sh Sherwood number (–) solve in order to obtain more efficient reactors.
T temperature (K) Open cell solid foam (OCSF) can be considered as a promis-
t time (s) ing alternative to fixed beds and monoliths due to their low
t̄s mean residence time (s) pressure drop (high voidage in the range of 90%), their good
u superficial velocity (m s−1 ) thermal characteristics (for SiC or metallic foams), their good
Vp active catalyst volume (m3 ) radial mixing and their high external specific surface area.
VR reactor volume (m3 ) These were the main drivers to divert OCSF from their ini-
WeL liquid Weber number (–) tial applications like heat exchanger internals or light weight
XG Lockhart–Martinelli parameter (–) mechanical structures, to use OCSF as structured catalyst sup-
Z height of the foam bed (m) ports mainly for gas–solid reactions (Richardson et al., 2003;
Bianchi et al., 2012; Giani et al., 2005; Groppi et al., 2007; Incera-
Greek letters Garrido and Kraushaar-Czarnetzki, 2010; Lacroix et al., 2007;
ˇL liquid saturation (–) Montebelli et al., 2014; Patcas et al., 2007). However, these
P/Z linear pressure drop (Pa m−1 ) potential advantages can also be of great interest in reactive
ε foam open porosity three-phase systems.
εL liquid holdup (m3L /m3R ) Few studies on three-phase G-L-S systems using tubular
ϕ Thiele modulus (–) foam reactors have been published. The pioneering ones were
 liquid hold up correlation parameter (–) those of the Schouten’s group for a plate reactor containing
 catalyst effectiveness factor foams. They mostly focused on hydrodynamics measure-
 reduced time in RTD (–) ments with investigation of pressure drop and axial dispersion
under various feeding configurations: co-current down-flow
688 chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697

(Stemmet et al., 2008), co-current up-flow (Stemmet et al., no bypass and good contacting between foams and with the
2007, 2008) and counter-current flow (Stemmet et al., 2005, tube wall.
2006). The potentialities of counter-current flow for reactive
distillation were also investigated by Lévêque et al. (2009) and 2.2. Catalyst coatings
Grosse and Kind (2011, 2012). In all these works, first rough
estimations and measurements of G-L mass transfer were also The catalyst used for the coating of these foams is a commer-
provided and discussed. In the meantime, Topin et al. (2006) cial 2 wt.% Pd/Al2 O3 from Johnson Matthey. Catalyst coating
also investigated deeply this contactor albeit not for cataly- and characterizations are made according to a methodol-
sis but rather for the development of convective boiling heat ogy developed within the laboratory (Meille et al., 2005). The
exchangers. quantity of catalyst anchored on the foam is determined by
At the same time, the group of Edouard gave more exper- weighting. Then, the mean thickness of the catalyst layer
imental measurements and insights in modelling pressure can be estimated using the specific area measured and mak-
drop, liquid holdup and axial dispersion behaviours in co- ing assumption that coating thickness is homogeneous on all
current configuration for a tubular reactor in down-flow mode the surface support. This aspect has been controlled qual-
(Edouard et al., 2008a,b) and in a millimetre-scale plate reactor itatively through Scanning Electron Microscopy (Hitachi TM
in horizontal flow (Saber et al., 2012a,b). Since 2013, the group 1000 tabletop). Adherence of the catalyst layer is checked with
of Lange has started to publish on foam tubular reactors for mechanical testing standardized in-home methods (Meille
G-L-S applications with detailed G-L hydrodynamic studies in et al., 2005). Moreover, the catalyst activity is checked before
a pilot reactor (100 mm i.d.). Special care was addressed to the and after the coating in a batch reactor for the reaction of inter-
G-L distribution system, start up procedures and the resulting est concerned thereafter, showing that the coating procedure
possible hydrodynamic multiplicity (Mohammed et al., 2013, does not alter the catalyst activity.
2015). This group also investigated the external L-S mass
transfer by the electrochemical method (Mohammed et al., 2.3. Hydrodynamic measurements
2014).
Concerning the reactive media, Wenmakers et al. (2010a,b) Fig. 2 shows a global view of the setup used for the hydro-
were the first ones to publish data on a real G-L-S system in dynamic study. It consists in a tubular reactor with an inner
plate and a tubular foam reactors respectively. They studied diameter of 21 mm made of 5 sections of 20 cm height for a
and quantified the influence of carbon nanofiber coatings on total reactor height of 1 m. This tube is filled with 40 stacked
the L-S mass transfer. More recently, Tourvieille et al. (2015a,b) foam elements (diameter of 20 mm and height of 25 mm each).
investigated a new millimetre-scale horizontal foam reac- This contactor was operated at room temperature and atmo-
tor in the pulsed regime and demonstrated some benefits spheric pressure with a co-current up-flow of air and water.
brought by the confinement in a milli-channel on hydrody- These fluids were supplied through a T-junction at the bot-
namics and overall mass transfer through measurements with tom of the reactor and are mixed in a small packed bed of glass
the ␣-methyl styrene model reaction. Finally, G-L-S stirred beads (4 mm diameter) to ensure a good fluid distribution. The
tanks using foam pieces as stirrer have also been the subject liquid flow-rate is delivered by a gear-pump (Tuthill DGS 99)
of several recent works (see for example Tschentscher et al., and regulated by a Coriolis mass-flow controller (Bronkhorst
2010; Truong-Phuoc et al., 2014; Leon et al., 2014; Lali et al., Cori-flow). The gas flow-rate is regulated and fed through
2015) but this contact mode being too different from the fixed a thermal mass-flow controller (Brooks 5851 E-series). The
bed investigated here, it was chosen to not focus on them. superficial velocities investigated range between 0.8 mm/s and
In this work, a new contribution to the evaluation of cat- 25 mm/s for the liquid and between 100 mm/s and 900 mm/s
alytic foams as internal for intensification of tubular G-L-S for the gas. This operating window was chosen to be realistic of
up-flow reactors is proposed. A hydrodynamic study and the potential industrial reactive applications (hydrogenations,
overall external mass transfer measurements under reac- oxidations, etc.). The pressure drop is monitored using a differ-
tive conditions are provided. Hydrodynamics are investigated ential pressure sensor (Keller PD-33XEi) and liquid residence
through residence time distribution (RTD), pressure drop is time distribution (RTD) acquisition is done with a calibrated
measured and the overall external mass transfer performance conductivity cell (Radiometer Analytical) using an aqueous
is evaluated through the very fast AMS catalytic hydrogena- KCl solution as the tracer. Step injections of tracer solutions
tion reaction under various experimental conditions. were made at the reactor inlet using a syringe pump (Har-
vard Apparatus PHD4400) and successive measures of inlet
2. Materials and method and outlet conductivity signals are carried out.
Before each experiment, the column is completely filled
2.1. Foam internals with water to eliminate possible initial dry zones and trapped
bubbles, then the required gas and liquid flow-rates are set
Solid foams used in this study are made of stainless steel and for a period of flow stabilization before tracer injection and
are commercially available from Porvair Corp. Three grades conductivity measurement. The tracer concentration being
are used, commercially defined by their apparent linear pore linearly proportional to the measured conductivity ( ), the
density as 20, 40 and 60 pores per inch (PPI). More precise cumulative residence time distribution function F(t) is directly
morphological parameters such as global voidage, pore size obtained after normalization (by max which is the conductiv-
distribution and mean strut diameter are estimated with X- ity of the tracer solution at the concentration C0 contained in
ray micro tomography and 3-D image reconstruction followed the syringe). The residence time distribution function E(t) is
by image analysis with Imorph (Brun, 2009). A typical 2D tomo- directly and easily calculated through derivation (Eq. (1)).
graphy foam slice and the corresponding 3-D reconstruction
are exemplified in Fig. 1. Foams were precisely cut with an elec- (t) C(t) dF(t)
F(t) = = and E(t) = (1)
tro arc discharge technique (precision of ␮m) in order to ensure max C0 dt
chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697 689

Fig. 1 – X-ray tomography examples: (a) typical 2D acquisition from X-ray tomography; (b) foam 3D reconstruction.

with k0 = 8.5 × 106 mol s−1 g−1


Pd
, Ea = 38.7 kJ mol−1 , and
−2
KH = 1.4 × 10 m /mol.
3

Because the effective reaction order is H2 is close to 1 (0.75),


a first order reaction is assumed. Taking into account the
internal mass transfer limitation with ordinary Thiele mod-
ulus estimations (Eq. (4)), the AMS conversion, measured by
gas chromatography, allows determining an apparent particle
reaction rate used to compute the concentration of hydrogen
at the catalyst surface.


tanh(ϕ) rp (Csu )
rp = r(wt.%Pd ) p with  = and ϕ = Lc
ϕ Deff Csu
(4)

where  is the catalyst effectiveness factor, p is the cata-


lyst density, wt.%Pd is the catalyst loading with palladium
(expressed in mass), Lc is the wash-coat thickness, Deff is the
effective diffusivity of H2 in the catalyst porous network and
rp is expressed in mol s−1 m3catalyst .
Fig. 2 – Schematic view of the experimental set-up used in The overall external mass transfer coefficient, Kov (Eq. (5))
hydrodynamic studies. expressed in s−1 , is computed simultaneously to the determi-
nation of Csu by an iterative process using the mass balance
at steady state (Eq. (6)):
2.4. Overall external mass transfer measurements
1 1 1
= + (5)
2.4.1. Principle Kov kL aGL kS aLS
The catalytic hydrogenation of ␣-methylstyrene (AMS) on a
Pd/Al2 O3 catalyst is known to be one of the fastest reactions Kov (C∗ − Csu )VR = rp Vp (6)
and appears often strongly limited by hydrogen transfer from
the gas phase to the reaction sites (Eq. (2)). This reaction is where C* is the liquid concentration of H2 at saturation,
used to measure the global mass transfer performance of the calculated under working T and P conditions, according to
foam contactor. This method has been developed in the lab Herskowitz et al. (1978). VR is the reactive bed volume (gas and
and used in previous works on characterization of different liquid and solid foam catalyst) in m3 and Vp is the wash-coated
G-L-S reactors (Meille et al., 2004; Tourvieille et al., 2013, 2015b). catalyst volume in m3 . Because the rate law for the reaction
is zeroth order with respect to the liquid reagent (AMS) and
because the H2 pressure can be considered as constant over
the reactor volume (pure hydrogen, negligible pressure drop
and negligible solvent vaporization) then, Eq. (6) is valid what-
ever the flow behaviour (ideal plug flow, dispersive plug flow,
CSTR, etc.). More details can be found in Tourvieille et al.
(2013).
(2)
2.4.2. Set up and experimental procedure
The three phase reactive experiments have been carried out in
The intrinsic kinetic law of this reaction was determined
the experimental set-up used for hydrodynamics with some
in a previous study of Meille et al. (2002) and is given as a
adaptations (see Fig. 3). Catalytic foams are placed in the
reminder in Eq. (3).
medium part of the reactor for a total catalytic bed height of
 E  KH CH
12.5 cm, all the reactor being filled with the same inert foam
a −1 −1
r = k0 · exp −
RT
·  √ 2 (mol s gPd ) (3) material to ensure an identical hydrodynamic behaviour. The
1 + KH CH
liquid mixture is composed of 90 wt.% of methyl-cyclohexane
690 chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697

Concerning the active phase loading, it increases logically


while decreasing the mean cell diameter due to the increase in
external specific surface area available for coating. One inter-
esting thing is that the order of magnitude of active phase is
similar to those obtained in classical fixed bed of “egg-shell”
beads (820 gPd /m3 for beads of 2.5 mm diameter is taken as a
reference point for comparison). The good homogeneity and
the textural quality of the coatings are checked qualitatively
by SEM (Fig. 4) and their good adherence is quantitatively
demonstrated by the negligible weight loss encountered after
successive treatments in an ultrasonic bath of water and hep-
tane.
Concerning the activity evaluation of the deposited cata-
lysts, the different tests carried out in a well characterized
batch reactor with scratched powder after deposit on a plane
substrate of the same stainless material confirmed that the
initial activity of the commercial catalyst remain unchanged
by the coating procedure.

3.2. Residence time distribution


Fig. 3 – Schematic view of the experimental set up adapted
for three phase catalytic reaction under pressurized Measurements and modelling with piston-dispersion model
conditions. of inlet signals for all the conditions considered were first real-
ized. Little deviations from ideal step injections are obtained
(solvent) and 10 wt.% of the AMS substrate. The gaseous for all the tested conditions as shown in typical examples of
phase consists in pure hydrogen. Downstream to the reactor, E() and F() curves in Fig. 5. For the most unfavourable condi-
the gas/liquid mixture is separated in a gas/liquid separa- tions (namely the lowest liquid and gas velocities, see Fig. 5a),
tor allowing liquid sampling for GC analysis and where the a Péclet number as high as 300 is obtained which is commonly
outlet pressure is regulated. All the experiments have been admitted to be close to ideal plug flow conditions. When the
performed at a pressure of almost 3 bar and ambient tem- liquid velocity increases (Fig. 5b and c), the corresponding inlet
perature (20–25 ◦ C). The axial temperature profile inside the signals tend to stiffen and approach the ideal step.
reactor is followed with several thermocouples and the max- Before making the assumption of ideal steps as inlet sig-
imum temperature is taken for Kov calculations in order to nals to lighten the treatment, an additional verification of its
be conservative. When steady state is reached (stable pres- low impact has been done on random experiments. A typi-
sure, temperature and outlet composition), liquid samples are cal Fourier analysis of inlet and outlet raw signals followed
collected for GC analysis and AMS conversion determination. by a deconvolution in this space and an inverse Fourier trans-
form get access to the experimental transfer function E(t). This
3. Results and discussions latter was then used to fit a classical one-dimension piston-
dispersion function (Eq. (7)).
3.1. Foam support characterization and catalyst    2

coating 1 Pe Pe(t̄S − t) us,L Z
E(t) = exp − with Pe = (7)
2 t̄S t 4t̄S t Dax,L
The mean morphological characteristics of the different foams
used in this study are determined by X-ray tomography and are A comparison of the results obtained by this treatment with
listed in Table 1. This table contains also the data concerning those made with the ideal inlet step signals and the same
catalyst coatings (determined by weighing). 1-D model for various examples shows little and acceptable
These features clearly illustrate the main differences in deviations (Table 2).
terms of specific area and morphological parameters (cell The piston-dispersion model represents properly the RTD
diameters, etc.) for the 3 different foam grades while keep- curves of the liquid phase (Fig. 6). The experimental signal
ing a very high open porosity (around 90%) for all the foams. perturbations encountered are due to gas phase and could

Fig. 4 – Qualitative SEM pictures of the catalyst coating on the 20 PPI stainless steel foam.
chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697 691

Table 1 – Summary of foam support characteristics and catalyst wash-coats.


Foam characteristics Commercial grades

20 PPI 40 PPI 60 PPI

Open porosity, ε (–) 0.94 0.92 0.88


Mean cell diameter, dcell (␮m) 2470 1450 680
Mean strut diameter, dstrut (␮m) 320 233 195
Specific surface area, a (m2 /m3foam ) 773 1495 3188
Palladium loading (gPd /m3foam ) 618 805 1402
Catalyst layer estimated thickness (␮m) 40 27 22

a) b)
1 35 1 35
F experiment
F experiment
F model
F model 30 30
0.8 0.8 E model
E model
25 25
Pe ≈ 6000
0.6 0.6

E(θ) (-)

E(θ) (-)
F(θ) (-)
F(θ) (-)

20 20

0.4 15 0.4 15

10 10
0.2 0.2
Pe ≈ 300 5 5

0 0 0 0
0 0.25 0.5 0.75 1 1.25 1.5 0 0.25 0.5 0.75 1 1.25 1.5
θ (-) θ (-)

c) d)
1 35 35
F experiment uL = 8.0 mm/s
F model Pe ≈ 15000 30 30 uL = 2.4 mm/s
0.8 E model uL = 0.8 mm/s
25 25
0.6
E(θ) (-)

E(θ) (-)
F(θ) (-)

20 20

0.4 15 15

10 10
0.2
5 5

0 0 0
0 0.25 0.5 0.75 1 1.25 1.5 0.75 0.875 1 1.125 1.25
θ (-) θ (-)

Fig. 5 – RTD measurements of inlet signals for uG = 0.08 m s−1 with (a) uL = 0.8 mm/s; (b) uL = 2.4 mm/s; (c) uL = 8 mm/s; and (d)
comparison of inlet E curves.

increase the deviation on the determination of the RTD param- be very dependant of the foam mean cell size (or linear pore
eters. density) with highest Péclet numbers being obtained for the
No by-pass or dead zones induced by the stacks of foam smallest cell sizes (5 < Pe < 60 for 20 PPI, 10 < Pe < 140 for 40 PPI
internals (or the system itself) could be detected in the exper- and 60 < Pe < 350 for 60 PPI). These findings are in agreement
iments. The estimated mean residence times are mostly with the idea that a lower characteristic size tends to stabilize
influenced by the liquid velocity and the mean cell size and the liquid flow. These rather high Péclet numbers support an
have a little dependence with the gas velocity (Fig. 7a). The ideal plug flow for a wide range of experimental conditions
obtained axial dispersion coefficients are represented by a especially for the high liquid velocities or the low mean cell
Péclet number (Fig. 7b). diameters. In comparison to conventional packed beds, the
For all the foams, the Péclet number significantly increases liquid axial dispersion coefficient appears quite higher with
with increasing liquid velocity whereas it is quite insensitive the foams. This is probably due to an effect of the higher
to the gas velocity. The range of Péclet number appears to void fraction of foams. The same behaviour has been noticed

Table 2 – Comparison of Péclet number and mean residence time obtained with different models.
Operating conditions Ideal inlet signal Deconvolution of inlet and
outlet signals
Pe t̄s (s) Pe t̄s (s)
−1
20 PPI, uL = 0.8 mm/s, uG = 0.08 m s 6 662 7 667
40 PPI, uL = 4.8 mm/s, uG = 0.08 m s−1 33 156 36 154
60 PPI, uL = 2.4 mm/s, uG = 0.08 m s−1 82 279 81 268
692 chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697

0.8

0.6

εL (-)
0.4

60 PPI uL=2.4mm/s 60 PPI uL=16mm/s


0.2 40 PPI uL=2.4mm/s 40 PPI uL=16mm/s
20 PPI uL=2.4mm/s 20 PPI uL=16mm/s
40 PPIuL=20mm/s
40 PPI uL=20mm/s Stemmet
Stemmet

0
Fig. 6 – Illustration of the good match between 0 200 400 600 800 1000
experimental and calculated outlet curves for the 40 PPI uG (mm/s)
foam with uL = 4 mm/s and uG = 80 mm/s.
Fig. 8 – Comparison of experimental liquid holdup for 20,
by Saber et al. (2012a) in a co-current down flow configura- 40 and 60 PPI foams for various conditions. (Dashed lines
tion. The only work performed with an up-flow configuration are only guides for the eyes.)
is the study of Stemmet et al. (2007) where the characteriza-
tion of axial dispersion involved a modified Bodenstein liquid liquid flowrates. A correlation for the liquid hold-up in OCSF is
number using the interparticle velocity (Eq. (8)). Thus this Bo proposed, inspired from the work of Larachi et al. (1991) devel-
number and the Pe number used in this study could be com- oped to predict liquid saturation (ˇL ) for conventional packed
pared (Fig. 7b). beds of spherical particles in trickle flow configuration (Eqs.
(10) and (11)).
uL Z Pe
Bo = = (8)
Dax,L εL εL εL WeBL
ˇL = = 1 − 10− with  = A C
(10)
ε XG ReD
L
Stemmet et al. (2007) obtained a lower Péclet number with
a 10 PPI foam. This is in agreement with the trend concerning
where
the impact of foam mean cell diameter. The liquid holdup is
estimated from the mean residence time (Eq. (9)): 
uG G L uL dcell L uL dcell
XG = , ReL = and WeL =
uL L L L
QL
εL = (9) (11)
VR

The liquid holdup increases with the liquid velocity and


decreases as the gas velocity increases (Fig. 8). A higher liq- The optimization led to the values of parameters A, B, C
uid holdup is also obtained for smaller mean cell diameters and D of 218, 0.5, 0.3 and 1.16 respectively. This correlation
(or higher PPI). A physical explanation could lie in the greater is able to predict satisfactorily the hold up with a low mean
importance of capillary forces observed within more finely deviation of less than 10% (Fig. 9). Additional measurements
structured foams. Liquid holdups, determined experimentally with liquids of various properties, especially higher viscosities
by Stemmet et al. (2007) and their dependence on gas and would be of great interest to extend the validity of this corre-
liquid flowrates appear similar and are also represented for lation and to approach conditions encountered with reactive
comparison (Fig. 8). The range of liquid holdup observed can systems. Finally, it is noticeable that the Weber number pos-
be considered as quite high even at a high ratio of gas to sesses a stronger influence in comparison to the correlation of

a) b)
1000 1000
60 PPI 60 PPI
40 PPI 40 PPI
20 PPI 20 PPI
10 PPI Stemmet

100
Pe (-)
τ (s)

100

10

10 1
0 15 30 45 60 0 15 30 45 60 75
ReL (-) ReL (-)

Fig. 7 – Comparison of (a) mean residence times and (b) liquid Péclet numbers obtained for 20, 40 and 60 PPI foams. (Dashed
lines are only a guide for the eyes.)
chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697 693

1 25
+10% 60 PPI uL = 2.4 mm/s 60 PPI uL = 24 mm/s
60 PPI 40 PPI uL = 2.4 mm/s 40 PPI uL = 24 mm/s
40 PPI 20 PPI uL = 2.4 mm/s 20 PPI uL = 24 mm/s

0.8 20 PPI 20
-10%

Δp/L (kPa/m)
0.6 15
εL, calc (-)

0.4 10

0.2 5

0 0
0 0.2 0.4 0.6 0.8 1 0 200 400 600 800 1000
εL,exp (-) uG (mm/s)

Fig. 9 – Comparison of experimental and calculated values Fig. 11 – Comparison of global pressure drop for 20, 40 and
of liquid holdup. 60 PPI foams at 2 liquid superficial velocities. Error bars
represent the signal amplitude fluctuations during
measurements.
Tourvieille et al. (2015a) developed for horizontal foam reac-
tors with a high degree of confinement.
It is closely linked with the slow liquid holdup evolution at
these superficial velocities. Indeed, the static pressure term
3.3. Pressure drop
continues to decrease but more slowly and the dynamic pres-
sure drop term significantly increases. This behaviour might
The global linear pressure drop (dynamic + static) has been
also indicate that two distinct flow regimes exist – for example
measured at various superficial gas and liquid velocities and
bubbly and pulsed regimes – as already proposed by Stemmet
for all the OCSF. For the sake of conciseness, only the measure-
et al. (2007). However, this cannot be confirmed without fur-
ments for the 20 PPI foam are presented first as typical results
ther experiments. Comparison of pressure drops measured
(Fig. 10).
for several foams and various gas and liquid velocities is also
A noticeable discontinuity concerning the evolution of
given in Fig. 11.
pressure drop dividing the curves in two distinct zones is
Based on the evolution of the pressure drop encountered
observed. For a given liquid velocity, the linear pressure drop
for 20 and 40 PPI foams, bubble and pulsing regimes have been
decreases first with a gas velocity increase and increases after
proposed. The observed steep increase in the noise to signal
this transition near uG = 0.2 m s−1 . At lowest gas and liquid
ratio upon a gas velocity increase may further support the idea
velocities, the foam beds are mostly full of liquid and the
of reaching a pulse regime. For the 60 PPI foam, a different
observed pressure drop is essentially due to the liquid static
behaviour is encountered: pressure drop always increases as
contribution. In the first part of the curve, the decrease of
the gas velocities increases but with a much lower noise to
the liquid holdup with an increase in the gas velocity con-
signal ratio, likely indicating the preservation of the supposed
tributes to a decrease of static pressure and of total pressure
bubbly flow over the whole range of operating conditions for
drop because it is not compensated by a comparable or higher
that tight foam structure. Overall, the trend is that measured
dynamic pressure drop term. After the transition, the pressure
pressure drops increase with the liquid velocity and the foam
drop increases significantly when the gas flow-rate increases.
grade but are kept far below those found in fixed beds of
spherical particles. In the work of Stemmet et al. (2007), the
15 boundary between bubble and pulse regimes has been found
uL=2.4mm/s at the same gas velocity of 0.3 m s−1 for both 10 and 40 PPI
uL=4.8mm/s
uL=16mm/s foams. This value is comparable to the range of values found
12.5
uL=24mm/s in this study (0.1–0.3 m s−1 ). The relatively weak deviation of
behaviour between 40 and 60 PPI foams may be also explained
10 by possible intrinsic differences between foams. Indeed, flow
Δp/L (kPa/m)

behaviour and pressure drop may be greatly influenced by


7.5 the presence of imperfections and variations in the quality
of foam manufacturing. Thus, more quantitative comparisons
5 or discussions appear unproductive. It can just be said that
the range of observed pressure drops (near 0.05–0.1 bar/m and
always lower than 0.2 bar/m) appears in agreement with the
2.5
work of Stemmet et al. (2007) on up-flow and is much lower
than pressure drops encountered in denser packed beds. To
0 go further, measured pressure drops can be corrected by the
0 200 400 600 800 1000
static pressure contribution (knowing the liquid hold up) and
uG (mm/s)
get access to the dynamic (or frictional) pressure drop term
Fig. 10 – Global linear pressure drop measurements for the which could be of interest to define more properly the pos-
20 PPI foam. sible transition between bubble and pulsing flow regimes. A
694 chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697

a) 1 b)
1
0.16 m/s 0.16 m/s
0.32 m/s 0.32 m/s
0.42 m/s 0.42 m/s
0.8 0.8

0.6 0.6
Kov (s-1)

Kov (s-1)
0.4 0.4

Fixed bed uG = 0.32 m/s


0.2 0.2 Fixed bed
dp = 0.70 mm uG = 0.32 m/s
uG = 0.16 m/s dp = 1.45 mm
uG = 0.16 m/s
0 0
0 2 4 6 8 10 0 2 4 6 8 10
uL (mm/s) uL (mm/s)
c) d)
1 1
0.16 m/s 60 PPI
0.22 m/s 40 PPI
0.32 m/s 20 PPI
0.8 0.42 m/s 0.8

0.6 0.6

Kov (s-1)
Kov (s-1)

0.4 0.4

0.2 Fixed bed


0.2
dp = 2.50 mm uG = 0.32 m/s
uG = 0.16 m/s
0 0
0 2 4 6 8 10 0 2 4 6 8 10
uL (mm/s) uL (mm/s)

Fig. 12 – Evolution of overall mass transfer coefficients with liquid superficial velocity for (a) 60 PPI, (b) 40 PPI, (c) 20 PPI
foams and (d) comparison of the 3 foams for a gas superficial velocity of 0.16 m s−1 .

detailed study focused on this topic may be of great inter- confirms the low effect of cell diameter on overall mass trans-
est, especially if it is coupled with quantitative considerations fer coefficient values. The “bell shaped” profile of the curves
about the quality of the foam structure. seems more pronounced for the highest cell diameter which
is likely related to the hydrodynamic behaviours observed in
3.4. Global mass transfer the pressure drop measurements.
Regarding the comparison with packed bed of spheres
The experimental overall external mass transfer coefficients (Fig. 12a–c), the mass transfer coefficients obtained (between
determined with AMS hydrogenation are presented in func- 0.2 and 0.9 s−1 ) appear one order of magnitude higher than the
tion of liquid velocity for the 3 foams (Fig. 12a–c). values estimated for co-current fixed bed reactors operating at
The overall external mass transfer coefficients for the OCSF similar gas and liquid superficial velocities and Reynolds num-
are compared with computed values for fixed beds of spheres bers. Another difference with dense packed bed is that in the
calculated from the correlations of Lara-Marquez et al. (1993) chosen operating window, no bell shape curves are obtained
for the G-L contribution and Delaunay et al. (1980) for the L-S for Kov in packed beds which exhibit a constant increase of
contribution. This comparison has been done under compa- overall mass transfer with increasing superficial liquid veloc-
rable flowing conditions with the mean cell diameter used as ity. The substantial improvement in terms of external mass
particle diameter for calculations. transfer coupled to pressure drop reduction brought by solid
Whatever the foam is, the overall mass transfer coefficients foams is confirmed under reactive conditions in this study
exhibit a similar “bell-shaped” profile upon increase of the liq- for low liquid Reynolds numbers (0.5–20). The “bell-shaped”
uid superficial velocity. Moreover, the flow velocity at which profile was not encountered by Stemmet et al. (2007) who
the maximum mass transfer coefficients of 0.5–0.9 s−1 are performed gas–liquid mass transfer coefficient measurements
reached appears quite independent from the foam struc- in the up-flow configuration and with a physical method.
ture and lies in the range 3–5 mm/s. This behaviour was first Surprisingly, the determination of overall mass transfer coef-
noticed by Tourvieille et al. (2015b) but with higher Kov val- ficient with oxidation of sodium formate saturated in oxygen
ues (up to 2 times higher). The present study tends to confirm by Wenmakers et al. (2010b) leads to a similar “bell-shaped”
these results, especially the possible enhancements brought profile. The authors provided an explanation based on a pos-
by the combination of inlet pulsing flow and confinement in a sible deactivation of the catalyst. This cannot be in the present
milli-channel. As in the work of Tourvieille et al. (2015b), the work, because the absence of deactivation has been checked.
overall mass transfer appears greatly influenced by the liquid An explanation in agreement with the present studies and
velocity for all 3 foams and in a less extent by the gas veloc- the former published data Wenmakers et al. (2010b) would lie
ity, especially for 60 PPI foam. A comparison of overall mass in the possible transition between a regime with a strong L-S
transfer coefficients for all foams as a function of liquid veloc- limitation to a regime mainly driven by the G-L one at higher
ity for a constant gas velocity is also presented (Fig. 12d). It liquid velocities due to a large reduction of the specific surface
chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697 695

1 +30%
60 PPI
40 PPI
0.8 20 PPI
Kov, calc (s-1)

0.6 -30%

0.4

0.2

0
0 0.2 0.4 0.6 0.8 1
Kov,exp (s-1)

Fig. 13 – Comparison of experimental and calculated values


of overall mass transfer coefficients.

area for G-L contacting mass transfer (strongly dependant of


the liquid hold up and its evolution with fluid velocities). This
theory has already been discussed by Tourvieille et al. (2015b).
In this sense, a similar attempt of modelling using the same
correlation forms is proposed. The resulting final correlations
are given in Eqs. (12)–(13) and the parity plot for experimental
and theoretical overall mass transfer coefficients is proposed
(Fig. 13).

kL aGL d2cell d 1.49


ShGL aGL dcell = = 1211Re−0.29
L Re0.71
G ScL
0.33 cell
Dm dR
(12)

kS dcell
 1.49
0.33 dcell
ShLS = = 932Re0.49
L ScL (13)
Dm dR

The proposed set of correlations could be considered as


satisfactory whereas the dispersion of values is quite signifi-
cant (mean deviation near 25%). This correlation set is a first
attempt of modelling and has certainly to be improved and
confirmed by performing additional experiments, especially
with other characterization methods and a wider experimen-
tal domain including other reactor diameters. Nevertheless,
this model can be used to estimate and discuss gas/liquid and
liquid/solid separate transfer contributions (see Fig. 14).
Fig. 14 – Illustration of correlation set behaviour: (a)
It is interesting to notice the small dependence of Kov and
gas/liquid and liquid/solid separated mass transfer
kL aGL with the foam structure (Fig. 14a and b) as already
contributions for 20, 40 and 60 PPI (constant uG = 0.1 m s−1 );
observed (Stemmet et al., 2007; Tourvieille et al., 2015b).
(b) corresponding Kov evolution and (c) impact of gas
The correlation set indicates that for a given gas superfi-
velocity on the separate contributions for the 60 PPI foam.
cial velocity, the L-S mass transfer contribution increases
with increasing liquid velocity whereas the G-L mass trans-
fer tends to decrease. This is related to a large diminution
of the gas/liquid effective area which counterbalances a contribution presents contrasts with the constant increase in
shorter stagnant liquid film at the G-L interface. This opposite kL aGL with increasing liquid velocities as reported by Stemmet
behaviour of each external mass transfer resistances leads to et al. (2007). No clear explanation can be proposed and addi-
an optimum value of global mass transfer coefficient and qual- tional separate investigations of the G-L mass transfer and
itatively explains the “bell-shaped” profiles. Quantitatively, hydrodynamics in both reactor configurations (plate and tubu-
the decrease in theoretical Kov appears less pronounced than lar reactors) are mandatory.
in the experiments probably due to the significant deviations
observed (Fig. 13). Finally, the correlations allow illustrating 4. Conclusion
the dependence of the maximum Kov value with gas superfi-
cial velocity (Fig. 14c) as observed experimentally (Fig. 12). The Open cell solid foams have been further characterized as
L-S mass transfer coefficient variation is supported by previ- catalyst carrier structures for alternative co-current up-flow
ous measurements (Mohammed et al., 2014). However, the G-L G-L-S tubular fixed bed reactors. Stainless steel foams have
696 chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697

been successfully wash-coated with a commercially avail- Groppi, G., Giani, L., Tronconi, E., 2007. Generalized correlation for
able Pd/Al2 O3 catalyst without any loss in intrinsic activity. gas/solid mass-transfer coefficients in metallic and ceramic
Residence Time Distribution measurements have shown that foams. Ind. Eng. Chem. Res. 46, 3955–3958.
Grosse, J., Kind, M., 2011. Hydrodynamics of ceramic sponges in
the classical piston-dispersion approach is able to represent
counter-current flow. Ind. Eng. Chem. Res. 50, 4631–4640.
perfectly the behaviour of the liquid phase with axial Péclet Grosse, J., Kind, M., 2012. A contribution to the experimental
numbers increasing with liquid velocity and linear pore den- determination of mass transfer in counter-current flow
sity while the gas superficial velocity has little impact in the through ceramic sponges. Chem. Eng. Process. 55, 29–39.
process window tested. Plug flow behaviour could be assumed Herskowitz, M., Morita, S., Smith, J.M., 1978. Solubility of
reasonably for the liquid phase in most of the experiments and hydrogen in alpha-methylstyrene. J. Chem. Eng. Data 23,
a predictive correlation for liquid hold-up has been derived. 227–228.
Incera-Garrido, G., Kraushaar-Czarnetzki, B., 2010. A general
Very high liquid holdups are encountered, even for the high
correlation for mass transfer in isotropic and anisotropic solid
gas to liquid flow-rate ratio. Global pressure drop measure- foams. Chem. Eng. Sci. 65, 2255–2257.
ments indicate the possible existence of two different flow Lacroix, M., Nguyen, P., Schweich, D., Pham-Huu, C., Savin Poncet,
behaviours and confirm the very attractive low pressure drop S., Edouard, D., 2007. Pressure drop measurements and
induced by solid foams in comparison to fixed bed operated modelling on SiC foams. Chem. Eng. Sci. 62, 3259–3267.
at the same fluid superficial velocities. Overall external mass Lali, F., Böttcher, G., Schöneich, P.-M., Haase, S., Hempel, S., Lange,
R., 2015. Preparation and characterization of Pd/Al2 O3
transfer coefficient estimations have been performed with a
catalysts on aluminum foam supports for multiphase
reactive chemical system and demonstrate the highest per-
hydrogenation reactions in rotating foam reactors. Chem.
formance of foams relatively to packed beds operated under Eng. Res. Des. 94, 365–374.
identical flow conditions. Combined to the pressure drop per- Larachi, F., Laurent, A., Midoux, N., Wild, G., 1991. Experimental
formance, this study confirms the large potential for process study of a trickle-bed reactor operating at high pressure:
intensification brought by these open structures for G-L-S up- two-phase pressure drop and liquid saturation. Chem. Eng.
flow operation, especially at low Reynolds numbers. A first Sci. 46, 1233–1246.
Lara-Marquez, A., Wild, G., Laurent, A., Midoux, N., 1993.
attempt of modelling of the specific “bell-shaped” curves evo-
Coefficient de transfert de matière gaz-liquide dans un
lution with fluid superficial velocity has been proposed and réacteur catalytique à lit fixe fonctionnant à co-courant vers
discussed. The transition between two mass transfer limited le haut de gaz et de liquide en présence de systèmes
regimes: a first one dominated by the L-S resistance and a sec- coalescents ou inhibiteurs de coalescence et/ou visqueux.
ond one by the G-L resistance is proposed. The correlations Récents Progrès en Génie des Procédés 30, 67–72.
need to be further tested over a larger experimental window Leon, M.A., Nijhuis, T.A., van der Schaaf, J., Schouten, J.C., 2014.
and to be extended with specific experiments on each separate Residence time distribution and reaction rate in the horizontal
rotating foam stirrer reactor. Chem. Eng. Sci. 117, 8–17.
external resistances. However, the very attractive mass trans-
Lévêque, J., Rouzineau, D., Prevost, M., Meyer, M., 2009.
fer behaviour demonstrated under reactive conditions for low Hydrodynamic and mass transfer efficiency of ceramic foam
energy consumption in one tube can reasonably reflect the packing applied to distillation. Chem. Eng. Sci. 64, 2607–2616.
one obtained in an industrial multi-tubular reactor of higher Meille, V., de Bellefon, C., Schweich, D., 2002. Kinetics
capacity. of␣-methylstyrene hydrogenation on Pd/Al2 O3 . Ind. Eng.
Chem. Res. 41, 1711–1715.
Meille, V., Pestre, N., Fongarland, P., de Bellefon, C., 2004.
Acknowledgements Gas-liquid mass transfer in small laboratory batch reactors:
comparison of methods. Ind. Chem. Eng. Res. 43, 924–927.
Meille, V., Pallier, S., Santa Cruz, G.V., Roumanie, M., Reymond,
Solvay is gratefully acknowledged for the participation to the
J.P., 2005. Deposition of gamma-Al2 O3 layers on structured
financial support of this study and the postdoctoral position supports for the design of new catalytic reactors. Appl. Catal.
of Dr J. Lévêque. A: Gen. 286, 232–238.
Mohammed, I., Bauer, T., Schubert, M., Lange, R., 2013.
Hydrodynamic multiplicity in a tubular reactor with solid
References foam packings. Chem. Eng. J. 231, 234–244.
Mohammed, I., Bauer, T., Schubert, M., Lange, R., 2014.
Bianchi, E., Heidig, T., Visconti, C.G., Groppi, G., Freund, H., Liquid–solid mass transfer in a tubular reactor with solid
Tronconi, E., 2012. An appraisal of the heat transfer properties foam packings. Chem. Eng. Sci. 108, 223–232.
of metallic open-cell foams for strongly exo-/endo-thermic Mohammed, I., Bauer, T., Schubert, M., Lange, R., 2015. Chem.
catalytic processes in tubular reactors. Chem. Eng. J. 198-199, Eng. Proc.: Proc. Intens. 88, 10–18.
512–528. Montebelli, A., Visconti, C.G., Groppi, G., Tronconi, E., Kohler, S.,
Brun, E., 2009. PhD thesis of University Aix-Marseille. 2014. Optimization of compact multitubular fixed-bed
Cybulski, A., Moulijn, J.A., 2006. Structured Catalysts and reactors for the methanol synthesis loaded with highly
Reactors, 2nd ed. Taylor & Francis. conductive structured catalysts. Chem. Eng. J. 255,
Delaunay, G., Storck, A., Laurent, A., Charpentier, J.C., 1980. 257–265.
Electrochemical study of liquid-solid mass transfer in packed Pangarkar, K., Schildhauer, T.J., Ruud van Ommen, J., Nijenhuisb,
beds with upward co-current gas-liquid flow. Ind. Eng. Chem. J., 2008. Structured packings for multiphase catalytic reactors.
Proc. Des. Dev. 19, 514–521. Ind. Eng. Chem. Res. 47, 3720–3751.
Edouard, D., Lacroix, M., Pham, C., Mbodji, M., Pham-Huu, C., Patcas, F.C., Incera-Garrido, G., Kraushaar-Czarnetzki, B., 2007. CO
2008a. Experimental measurements and multiphase flow oxidation over structured carriers: a comparison of ceramic
models in solid SiC foam beds. AIChE J. 54, 2823–2832. foams, honeycombs and beads. Chem. Eng. Sci. 62, 3984–3990.
Edouard, D., Lacroix, M., Pham-Huu, C., Luck, F., 2008b. Pressure Richardson, J.T., Garrait, M., Hung, J.K., 2003. Carbon dioxide
drop modelling on solid foam: state-of-the art correlation. reforming with Rh and Pt–Re catalysts dispersed on ceramic
Chem. Eng. J. 144, 299–311. foam supports. Appl. Catal. A: Gen. 255, 69–82.
Giani, L., Groppi, G., Tronconi, E., 2005. Mass-transfer Saber, M., Truong-Huu, T., Pham-Huu, C., Edouard, D., 2012a.
characterization of metallic foams as supports for structured Residence time distribution, axial liquid dispersion and
catalysts. Ind. Eng. Chem. Res. 44, 4993–5002. dynamic–static liquid mass transfer in trickle flow reactor
chemical engineering research and design 1 0 9 ( 2 0 1 6 ) 686–697 697

containing ␤-SiC open-cell foams. Chem. Eng. J. 185–186, Tourvieille, J.N., Bornette, F., Philippe, R., de Bellefon, C., 2013.
294–299. Mass transfer characterization of a microstructured falling
Saber, M., Pham-Huu, C., Edouard, D., 2012b. Axial dispersion film at pilot scale. Chem. Eng. J. 227, 182–190.
based on the residence time distribution curves in a Tourvieille, J.N., Philippe, R., de Bellefon, C., 2015a. Milli-channel
millireactor filled with ␤-SiC foam catalyst. Ind. Eng. Chem. with metal foams under an applied gas–liquid periodic flow:
Res. 51, 15011–15017. flow patterns, residence time distribution and pulsing
Stemmet, C.P., Jongmans, J.N., van der Schaaf, J., Kuster, B.F.M., properties. Chem. Eng. Sci. 126, 406–426.
Schouten, J.C., 2005. Hydrodynamics of gas–liquid Tourvieille, J.N., Philippe, R., de Bellefon, C., 2015b. Milli-channel
counter-current flow in solid foam packings. Chem. Eng. Sci. with metal foams under an applied gas–liquid periodic flow:
60, 6422–6429. external mass transfer performance and pressure drop.
Stemmet, C.P., van der Schaaf, J., Kuster, B.F.M., Schouten, J.C., Chem. Eng. J. 267, 332–346.
2006. Solid foam packings for multiphase reactor: modelling Truong-Phuoc, L., Truong-Huu, T., Nguyen-Dinh, L., Baaziz, W.,
of liquid holdup and mass transfer. Chem. Eng. Res. Des. 84, Romero, T., Edouard, D., Begin, D., Janowska, I., Pham-Huu, C.,
1134–1141. 2014. Silicon carbide foam decorated with carbon nanofibers
Stemmet, C.P., Meeuwse, M., van der Schaaf, J., Kuster, B.F.M., as catalytic stirrer in liquid-phase hydrogenation reactions.
Schouten, J.C., 2007. Gas–liquid mass transfer and axial Appl. Catal. A: Gen. 469, 81–88.
dispersion in solid foam packings. Chem. Eng. Sci. 62, Tschentscher, R., Nijhuis, T.A., van der Schaaf, J., Kuster, B.F.M.,
5444–5450. Schouten, J.C., 2010. Gas–liquid mass transfer in rotating solid
Stemmet, C.P., Bartelds, F., van der Schaaf, J., Kuster, B.F.M., foam reactors. Chem. Eng. Sci. 65, 472–479.
Schouten, J.C., 2008. Influence of liquid viscosity and surface Wenmakers, P.W.A.M., van der Schaaf, J., Kuster, B.F.M., Schouten,
tension on the gas–liquid mass transfer coefficient for solid J.C., 2010a. Liquid-solid mass transfer for co-current gas–liquid
foam packings in co-current two-phase flow. Chem. Eng. Res. up-flow through solid foam packings. AIChE J. 56, 2923–2933.
Des. 86, 1094–1106. Wenmakers, P.W.A.M., van der Schaaf, J., Kuster, B.F.M., Schouten,
Topin, F., Bonnet, J.-P., Madani, B., Tadrist, L., 2006. Experimental J.C., 2010b. Enhanced liquid–solid mass transfer by carbon
analysis of multiphase flow in metallic foam: flow laws, heat nanofibers on solid foam as catalyst support. Chem. Eng. Sci.
transfer and convective boiling. Adv. Eng. Mater. 8, 890–899. 65, 247–254.

You might also like