You are on page 1of 9

Chemical Engineering Science 62 (2007) 3259 3267

www.elsevier.com/locate/ces

Pressure drop measurements and modeling on SiC foams


Maxime Lacroix a , Patrick Nguyen b , Daniel Schweich c , Cuong Pham Huu a , Sabine Savin-Poncet d ,
David Edouard a,
a LMSPC, UMR 7515 du CNRS, ECPM, Universit Louis Pasteur-ELCASS, 25 rue Becquerel, 67087 Strasbourg, France
b SiCat Technical Center, 1 rue du Brotsch, 67700 Otterswiller, France
c LGPC, CPE 43, bd du 11 novembre 1918, BP 2077 69616 Villeurbanne Cedex, France
d TOTAL, Centre Scientique et Technique Jean Feger, Avenue Larribau, 64000 Pau, France

Received 22 December 2006; received in revised form 17 March 2007; accepted 19 March 2007
Available online 23 March 2007

Abstract
Foam-structured beds are likely to be the next generation of catalyst supports due to their interesting specic properties (large exchange area,
low pressure drop, easy control of external porosity, etc.). Nevertheless, chemical engineering parameters of this new catalyst support types
are still not completely clear for the scientic community and many approaches are attempted to solve this problem. SiC foams offer the dual
advantages of the interesting properties of structured beds and the intrinsic thermal and mechanical properties of silicon carbide as a catalytic
support. In the present work, the problem of pressure drops along foam beds is studied with a new simplistic geometrical model as a rst
step in the understanding of the peculiar hydrodynamic behavior of SiC foams in chemical processes. The proposed model was successfully
validated by experimental results on a relatively large range of parameters which fully conrm the validity of the model.
2007 Elsevier Ltd. All rights reserved.
Keywords: Pressure drop; Foam; Cubic lattice model; Catalysis; Silicon carbide; Ergun equation

1. Introduction
Ceramic-based (Al2 O3 , cordierite, SiC, etc.) and metal-based
(aluminum, copper, etc.) cellular foams are widely used in a
large range of applications, especially employed in thermalinsulation applications (Lu et al., 1998)). These peculiars structures are commonly applied for packaging of food, disposable
hot-drinks cup (Gibson and Ashby, 1997), packed cryogenic
microsphere insulations (Beavers and Sparrow, 1969; Tien and
Vafai, 1990; Collishaw and Evans, 1994), solar energy utilization, transpiration cooling (Beavers and Sparrow, 1969), cavity wall insulation, (dwellings, aircraft, submarine cabins, etc.)
and more recently employed as catalyst or structured catalytic
supports (Twigg and Richardson, 1994; Groppi and Tronconi,
2000; Richardson et al., 2000; Ismagilov et al., 2001; Twigg
and Richardson, 2002; Richardson et al., 2003a,b; Pesant, 2005;

Corresponding author. Tel.: +33 390 242 675.

E-mail address: David.Edouard@ecpm.u-strasbg.fr (D. Edouard).


0009-2509/$ - see front matter 2007 Elsevier Ltd. All rights reserved.
doi:10.1016/j.ces.2007.03.027

Alie et al., 2006; Schimmoeller et al., 2006; Wine et al., 2006;


Zamaro et al., 2006). The reason why reticulated ceramic foams
are so attractive is the promising of an increase in transport
properties (Maestri et al., 2005); the high external surface areas (high porosity respecting an important geometric surface
area per volume of solid ac , m2 /m) lead to high external mass
transfer rates. Another reason is a much lower pressure drop
in a reactor lled with foams compared to a classical packed
particles bed reactor (i.e., extrudates and pellets). The low pressure drop of foams is probably the main advantage to take in
account for this new class of catalyst carrier. This permits to
perform catalytic reactions at very high space velocity (low
contact time) and in this way, increase in some cases the selectivity and/or use of a high length over diameter ratio reactors,
mostly found in highly exothermic or endothermic processes
(Richardson et al., 2003b). It is therefore important to understand the pressure drop characteristics of foam matrices before using these supports in catalytic reactions. Pressure drop
measurement presented in the literature, (Richardson et al.,
2000; Bhattacharya et al., 2002; Fourie and Plessis, 2002; Giani

3260

M. Lacroix et al. / Chemical Engineering Science 62 (2007) 3259 3267

Fig. 1. SiC foam prepared by shape memory synthesis.

et al., 2005) conrmed that foam matrices follow the


Forscheimer relationship. The pressure drop is the sum of
viscous and inertial terms and Ergun model may simulate
these results. But, the major problem in the foam pressure
drop estimation is to reliably dene structural properties of
the cellular medium to dene the equivalent particle diameter (dp ) necessary to use the well known Erguns equation.
Several approaches have been proposed to tackle this problem
(Richardson et al., 2000; Fourie and Plessis, 2002), unfortunately, the complexity of the geometric shape of foam materials, made that no general equation exists in the literature
for the calculation of the pressure drop through different foam
matrices.
Actually, the foams studied in the literature and available
on the market are metallic as well as ceramic. In some cases
metallic foams have to be passivated in order not to meet
reactivity toward the reactants or the products with the foam
material itself. Ceramic foams often present a low BET surface area and need to be covered with a washcoat in order to
increase their specic surface area The insulator character of
these ceramic foams also induces thermal transfer phenomena
which could be detrimental for the nal selectivity of the process. In addition, in some cases, the wash-coat could react with
either the reactants or products leading to the deactivation of
the catalyst. Silicon carbide foams studied in this work (Fig. 1)
are crystallized in the cubic structure (-SiC). They exhibit a
high thermal conductivity, a high resistance toward oxidation,
a high mechanical strength, a low specic weight and chemical inertia, a series of properties that renders -SiC a good
heterogeneous catalyst support (Ledoux et al., 1988; Ledoux
and Pham-Huu, 2001; Pham-Huu et al., 2002). The -SiC with
medium surface area (1520 m2 /g) can also be synthesized
in a foam monolithic structure according the shape memory
synthesis (Patents US 5,429,780; US 5,449,654; EP 0 624 560;

EP 0 880 406 B1; US 5,958,831; US 6 251 819; FR2860992,


FR2860993, US20050159292, FR2834655). -SiC exhibits
a higher specic area compared to -SiC preshaped foams
(SBET = 1 m2 /g) and can be shaped in a wide panel of geometric forms without the need for applying a wash-coat. Basically,
open-cell polyurethane foams are impregnated by a homogeneous mixture of silicon, charcoal and a phenolic resin which
acts as a binder and oxygen supplier. After polymerization of
the resin, the material is then carburized at 1300 C under inert
atmosphere. The nal -SiC foams are the replica of the
open-cell polyurethane starting foam. This method allows the
synthesis of silicon carbide pieces with controlled macroscopic
shapes respecting the PPI (pore per linear inch) of the starting
polyurethane material.
In this study, the problem of pressure drops along SiC foam
beds was studied with a new simplistic geometrical model,
without a priori knowledge of the gas ow behavior through
the SiC foam. The objective of the present work is to evaluate
the applicability of analogy between the traditional spherical
particles bed and the SiC foam in order to predict the pressure
drops in SiC foams through Erguns equation. The estimated
pressure drop results were compared with those obtained experimentally under air ow on SiC foams with 11003700 m
cell diameter and porosity between 0.76 and 0.92.

2. Experimental
Experimental pressure drop (P ) across each sample was
measured using the apparatus schematized in Fig. 2. Gas velocity is measured with anemometer Testo 435-1 equipped with
hot wire probe (020 m/s). Hot wire probe, due to its small
diameter, was chosen to limit the gas ow perturbation. Pressure
drop was measured with differential pressure sensor (Keller
Druckmesstechnik PD-41 (0200 mbar)). Pressure drop was
measured on 0.1 m long foams varying the gas velocity in the
05 m/s range.
Many authors and foam manufacturers characterize the foam
dimensions with the PPI (pore per inch) parameter. This characterization of the material is made by counting the number
of pores present on linear distance unit. The relation between
the cell size and the PPI parameter is given by the manufacturer. In our work, the average cell size is in the range
 = 1100.3700 m 50 m.
The foam cells can be modeled by a dodecahedron (12 pentagonal faces, 20 vertices and 30 edges or struts) (Fig. 3). This
peculiar geometry is a consequence of the forming process of
the polyurethane foam, base material of the SiC foam. One of
the characteristics of the foam is the cell size (i.e., the diameter of the dodecahedron) obtained directly by Fig. 3. But this
is not the pertinent parameter for pressure drop, as the smallest channel through which the gas has to pass is the pentagonal window (pentagonal side of the dodecahedron, Fig. 3). The
principal parameter for pressure drop estimation is then the size
of the window (a). A simple geometrical approach shows that
the ratio of the cell diameter () over the window diameter (a)
is around 2.3.

M. Lacroix et al. / Chemical Engineering Science 62 (2007) 3259 3267

3261

Fig. 2. Pressure drop measurement apparatus.

Fig. 3. Modeling of foam cell (left); optical picture of SiC foam: dotted circle shows the dodecahedral cell diameter (), whereas white arrow shows one of
the pentagonal windows (a) of the dodecahedron (right).

The second important parameter of the foam is the foam


porosity  which is the volume available to the uids ow
through the open-cell structure. Thus,  can be calculated on the
basis of mass and volume measurements using the expression
 = 1 g /s , where g is the foam apparent density and s
is the SiC density of the struts, which can be porous.
The SiC strut bulk density s and porosity are obtained with
the porous distribution delivered by mercury intrusion. This
measure gives the silicon carbide foam skeletal density from
which the cumulated volume < 30 m (volume equal to the
empty strut volume cf. Fig. 4) must be withdrawn, giving then
the SiC strut bulk density s . The ratio of mass to apparent
volume of SiC foams gives g .
For each sample of SiC foam studies in this work, the size
of the window (a), the porosity () and the measurements of
pressure drop are given in the following section.

3. Pressure drop results


3.1. Pressure drop of foams with different cell size
Three foams with porosity near to 0.91 were tested and compared (see Table 1). The cell size varies from 17501800 to
36503700 m. For these samples, the apparent density remains
constant around 130 g/L.
As cell size increases (Fig. 5):
PPI decreases,
number of struts (solid phase) per length unit decreases
strut diameter increases in order to respect the constant
apparent density around 130 g/L.
In Fig. 6, we can see that an increase in the cell size implies
a decrease of the pressure drop. Indeed, the channel through

3262

M. Lacroix et al. / Chemical Engineering Science 62 (2007) 3259 3267

which gas has to ow becomes larger and the uid frictions get
weaker, inducing smaller pressure drops.

3.2. Pressure drop of foams with different density


In Fig. 7, the samples studied present the same cell size
(1750 m) but with different densities. These differences are
due to foam preparation modications. Polyurethane foams
were rst inltrated with the silicon/resin mix. After drying inltration process was repeated two or three times resulting in a
signicant increase of the foam apparent density. Porosities ()
change as a consequence of these different densities. We can
see that as foam density (g ) increases, porosity () decreases
and pressure drop increases.
In this case, PPI parameter (equivalent to 1750 m cell size)
is constant and refers only to the number of cell per length unit
but not to the size of the strut. Fig. 7 shows that foams with

Fig. 4. SEM micrograph of the empty volume of a strut. This volume (obtained
by the porous distribution delivered by mercury intrusion) must be withdrawn
from the skeletal density.

Table 1
Apparent density and porosity of the studied SiC foams
Foam cell size (m)

Apparent density (g/L)

Porosity ()

17501800
26502700
36503700

142
136
121

0.915
0.910
0.914

Fig. 6. Pressure drop measurement versus gas velocity for different cell sized
SiC foams and estimated results.

Fig. 5. Optical pictures of 17501800 m (A) and 36503700 m (B) SiC foams; as the cell size increases, the number of struts per space unit decreases and
strut diameter increases in order to respect constant apparent density.

M. Lacroix et al. / Chemical Engineering Science 62 (2007) 3259 3267

3263

Fig. 8. Cubic cell model.


Fig. 7. Pressure drop measurement versus gas supercial velocity for different
foam densities with same cell size and estimated results.
Table 2
Comparison of calculated and measured strut diameter for foams with different
cell sizes

same PPI and different apparent density and porosity present


different pressure drop properties. This shows that not only the
PPI parameter is important for the foam description but also
the foam density and consequently the material porosity.
These results are in agreement with those of the literatures
(Richardson et al., 2000; Bhattacharya et al., 2002; Fourie and
Plessis, 2002; Giani et al., 2005).
The pressure drops follow the known Forchheimers equation
given by
P
= a0 ug + a1 u2g ,
L

(1)

where P /L is the pressure drop per unit length, ug is the


supercial velocity and a0 and a1 are constant.
In the following paragraph, we will see how starting from the
PPI and the porosity it is possible to estimate through Erguns
equation the pressure drop for standard structure foams and
thus developing a new model accounting for such new type of
support which calls for increasing development.

4. Model of pressure drop based on cubic cell


4.1. Model presentation
Erguns (1952) equation has been successfully employed in
the literature to predict the pressure drop of granular media.
Erguns equation for an uncompressible uid, through a rigid
and homogeneous porous medium is given by
(1 )u2g
(1 )2 ug
P
+
E
,
= E1
2
L
3 dp2
3 dp

(2)

Foam cell size (m)

Calculated ds (m)

Measured ds (m)

17501800
36503700

178
377

143
390

Table 3
Calculated versus measured strut diameters for SiC foams with different
apparent densities
Foam
Calculated ds (m) Measured ds (m) Measured nod (m)
characteristics
142 g/L
 = 0.91
184 g/L
 = 0.89
320 g/L
 = 0.81

178

143

160

209

103

230

301

110

320

where ug is the uid velocity (m/s), P the pressure drop


(Pa), L the length of the porous medium (m),  and  are the
uid viscosity and uid density, respectively, and E1 and E2
are Ergun constants,  is the bed porosity and dp is the mean
particle diameter of the granular medium.
The major problem in the foam pressure drop estimation is
to reliably dene structural properties of the cellular medium
to replace the equivalent particle diameter (dp ) in Eq. (2).
Different geometrical models for the foam structure are presented in the literature (Richardson et al., 2000; Bhattacharya
et al., 2002; Fourie and Plessis, 2002; Buciuman and
Kraushaar-Czarnetzki, 2003; Giani et al., 2005; Leong and Jin,
2006), but the models are very complex and empirical equation
must be used to t the data. Consequently, no general equation
is proposed and the Erguns parameters E1 and E2 are changed
by the author to correctly t the experimental data.

3264

M. Lacroix et al. / Chemical Engineering Science 62 (2007) 3259 3267

Fig. 9. Optical image of SiC foam of 142 g/L (A), 184 g/L (B) and 320 g/L (C); as apparent density increases, the nod size increases and more cell are closed.

In this work, we use the cubic cell model presented by


Giani et al. (2005). Richardson used external surface areas
(ac ) to compare the pressure drop of the ceramic foams with
the spherical particles (Richardson et al., 2000). In the present
work we use the cubic cell model with the same surface area
and same porosity as in the equivalent particle model in order
to develop a direct analogy between foams and beds made of
spherical particles. In a cubic cell model, we assume that the
cellular structure is made of solid cylindrical laments (struts)
connected in the three dimensions as a regular cubic lattice (see
Fig. 8).
Considering the unit cubic cell (see Fig. 8) and the fact that
each strut is shared among four cells, the total volume V0 is
given by
V0 = b3 .

(3)

The overall volume of the struts Vs can be calculated as a


function of the foam void fraction ():
Vs = (1 )V0 .

(4)

However, Vs can be calculated as the overall volume of the


cylinders (struts) included in the unit cell. Assuming that the
intersection volume of the strut is of second order, Vs is well
estimated by
 
12 ds 2
Vs = 
b.
(5)
4
2
Combining Eqs. (3) and (4), the following expression is obtained for ds from the pitch (a) and porosity ():
ds =

a[(4/3)(1 )]1/2
,
1 [(4/3)(1 )]1/2

(6)

with a = /2.3 and  the diameter of the dodecahedron presented in Fig. 3.


The external specic surface area per unit cell volume is then
computed as
ac =

4
(1 ).
ds

(7)

For comparing foams and particles, the external specic surface area per unit bed volume of spherical particles is used.
ac is given by the following equation:
6
ac =
(1 ).
dp

(8)

Finally, substitution of (7) in (8) gives the relationship between strut and particle diameter for a porous medium with the
same porosity :
6
(9)
d p = ds .
4
Eqs. (6) and (9) lead to the particle diameter equivalent to the
cellular structure based only on its mean windows size (a) and
the foam porosity. Once the porosity  is quantied, Erguns
equation (Eq. (2)) can be used to estimate the pressure drop in
the foam.
This analogy between foams and bed of spherical particles
has no physical meaning due to the high void inside the virtual packed bed. In other words, it is not possible to obtain a
bed composed of spherical particles with such porosity (typical packed bed porosity near of 0.39). However, this analogy
allows the direct determination of the estimation of the pressure drops in the SiC foams in a very simple and reliable way
for further studies. In the following paragraph, this new correlation is validated in the range of foam porosity from 0.75 to
0.92 and from uid velocities until 6 m/s.
4.2. Model validation
Ergun (1952) proposed the following value for packed
columns made of spheres: E1 = 150 and E2 = 1.75.
In practical cases, the Ergun constants depend on the packing properties (spheres, extrudates, slabs, etc. (see for instance
Iliuta et al., 1998)). They can be determined by measuring
single-phase gas ow pressure drops (Holub et al., 1992) or
estimated from literature correlations (Iliuta et al., 1998). With
the analogy developed in this work, it should not be necessary
to optimize the Ergun constants. So, the values of 150 and 1.75,
respectively, for E1 and E2 proposed by Ergun (1952) are used
in Eq. (2) for the estimation of the pressure drop in the SiC
foam. In other words, the results of simulation presented in the
following paragraphs do not involve tted parameters. The gas
used for pressure drop measurement is air and the inlet pressure
is considered close to 1 atm. In simulation, air properties are
assumed constant for the whole range of considered pressure
and the values are given by
gas = 1.2 kg/m3
gas = 1.838 105 Pa s.

M. Lacroix et al. / Chemical Engineering Science 62 (2007) 3259 3267

3265

4.2.1. Validation for different cell foam sizes


With the presented analogy pressure drop parameters are
correctly estimated and follow closely the experimental data
as shown in Fig. 6. Thus it is not necessary to modify Erguns constants as did the different research groups, who justify this new approach: foam cubic model spherical
particles bed with same porosity and same exchange area as
foams.
The window size (a) and strut diameters (ds ) of the foams
are measured by optical microscopy (Wild Mikroskop M420).
These values are measured on about 20 cells or struts and
the mean size is given with 5% standard deviation. The
mean window size given by the manufacturer is compared
with image analysis before use in order to estimate the particles equivalent diameter used in Erguns equation. The
calculated strut diameters are in the range of measured ones
(cf. Table 2).
4.2.2. Validation for different foam densities
In the following pressure drop simulations, PPI is constant
and only foam density and porosity change. The starting SiC
foam (142 g/L) has been re-inltrated to obtain SiC foam with
184 g/L and re-inltrated twice to obtain SiC with 320 g/L.
Fig. 7 shows good agreement of simulated pressure drops with
experimental data. Pressure drop increases with increasing
apparent density (g ) and decreasing foam porosity (). Indeed, porosity change induces strut diameter change following
Eq. (6).
Measured strut diameter is close to the estimated ds for
142 g/L foam but not in the cases of foams with 184 and 320 g/L
(cf. Table 3). Fig. 9 and Table 3 show that as apparent density
increases, the nod size (i.e., strut intersections) and closed cell
number increases. Due to the preparation method, the real strut
diameter seems to remain constant as apparent density increases
but not the nod size: SiC precursor seems to be deposited more
at the intersections of the polyurethane foam than all along the
strut when using repeated inltration. As a consequence, the
strut diameter remains unchanged, but the nod (struts intersection) size increases. Moreover, more cells become closed and
thus inaccessible to the gas ow.
The cubic cell model does not take into account the fact that
matter can be non-homogeneously distributed along the struts.
Thus, in the model extra solid volume added by re-inltration
is distributed all along the strut but not in the reality. The
observation of bigger nods has already been reported by
Bhattacharya et al. (2002). The observed differences between
calculated and measured strut diameters are then explained
by the reinforcement of the strut intersection rather than the
strut itself. Moreover, although the measured and estimated
strut diameters are not in good agreement in this case of reinltration, the model estimates correctly the pressure drop
properties of such re-inltrated foams (Fig. 7).
4.2.3. Validation for different foam materials
In the literature, no pressure drop measurement through SiC
foams has ever been reported. Usually, authors use alumina-

Fig. 10. Pressure drop measurement for different foam materials, SiC and
Al2 O3 and estimated results.

mullite or metallic foams (Richardson et al., 2000; Bhattacharya


et al., 2002; Fourie and Plessis, 2002; Giani et al., 2005; Leong
and Jin, 2006). Pressure drops have been measured in our experimental set-up on alumina-type foams with cell size between
1100 and 1150 m ( = 0.75) to validate the model with other
types of material. As can be seen in Fig. 10, the simulated and
experimental data are in good agreement and again conrm the
developed model for further applications.
4.2.4. Comparison of pressure drop through foam and
spherical particles xed bed
In catalytic reactors, external surface area is an important
parameter as it dictates the balance between the molecular
gaseous reactants passing in and the available active sites on
the unit volume of the catalytic bed. Fig. 11 shows the comparison of pressure drop between spherical xed bed and foam
presenting the same exchange surface area (2520 m2 /m3 or
m1 ). In order to compare the same mass of catalyst support,
pressure drops are divided by (1 ) on Fig. 11. As shown
in this gure, the pressure drop difference is quite important
and becomes larger as the gas velocity increases. With a gas
velocity of 1 m/s, pressure drop recorded on foam was about
0.25 105 Pa/m, whereas it was about 0.5 105 Pa/m for the
xed bed made of spheres; at a higher gas velocity, for example
4 m/s, pressure drop was still around 1.5105 Pa/m for foams,
whereas it was higher than 4 for the spheres bed. Richardson
et al. quote a factor 10 between the pressure drop recorded
between foam-type supports and spherical particles xed bed
(Richardson et al., 2000). These studies clearly show that pressure drops of foams are much lower than the ones of packed
bed of spheres and that foams are appropriate support for catalytic reactions with short contact times (i.e., high reactant
ows).

3266

M. Lacroix et al. / Chemical Engineering Science 62 (2007) 3259 3267

Acknowledgment
The author would like to thank SiCat and Total SA for, respectively, technical and nancial support.
References

Fig. 11. Comparison of pressure drop between foam and real packed bed.

5. Conclusion
A new pressure drop model based on a cubic lattice approach
of the foam structure has been developed. Although no physical
reason can be invoked in principle to explain the extension of
the Ergun equation, the model is in good agreement with experimental data derived from a large number of parameters. The
model is applicable with different PPI, different foam porosities and also with foams made of different materials using the
standard Ergun parameters. Measurement and estimated pressure drop conrm that the window size and the strut diameter
are the appropriate characteristics for simulating the pressure
drop in matrices foam. The hydrodynamic properties of foams
make them suitable as catalyst support due to their controlled
high porosity and the low pressure drop induced by the gas ow
compared to classical packed bed. This new catalysts support
will be used to control the catalytic reaction, for instance under
trickle ow condition, and this will be the subject of a future
article.
Notation
a
ac
dp
ds
E1,2
L

window diameter, m
external specic surface, m1
particle diameter, m
strut diameter, m
Ergun constants, (dimensionless)
length of foam, m

P

pressure drop, Pa

Greek letters
s
g


strut bulk density, g/L


apparent density, g/L
cell diameter (given by manufactured), m

Alie, C., Ferauche, F., et al., 2006. Pd-Ag/SiO2 xerogel catalyst forming by
impregnation on alumina foams. Chemical Engineering Journal 117 (1),
1322.
Beavers, G.S., Sparrow, E.M., 1969. Journal of Applied Mechanics 36, 711.
Bhattacharya, A., Calmidi, V.V., et al., 2002. Thermophysical properties of
high porosity metal foams. International Journal of Heat and Mass Transfer
45 (5), 10171031.
Buciuman, F.C., Kraushaar-Czarnetzki, B., 2003. Ceramic foam monoliths
as catalyst carriers. 1. Adjustment and description of the morphology.
Industrial & Engineering Chemistry Research 42 (9), 18631869.
Collishaw, P.G., Evans, J.R.G., 1994. An assessment of expressions for the
apparent thermal conductivity of cellular materials. Journal of Materials
Science V 29 (2), 486498.
Ergun, S., 1952. Fluid ow through packed columns. Chemical Engineering
Progress 48 (2), 8994.
Fourie, J.G., Plessis, J.P.D., 2002. Pressure drop modelling in cellular metallic
foams. Chemical Engineering Science 57 (14), 27812789.
Giani, L., Groppi, G., et al., 2005. Mass-transfer characterization of metallic
foams as supports for structured catalysts. Industrial and Engineering
Chemistry Research 44, 49935002.
Gibson, L.J., Ashby, M.F., 1997. Cellular Solids. Cambridge University Press,
Cambridge.
Groppi, G., Tronconi, E., 2000. Design of novel monolith catalyst supports
for gas/solid reactions with heat exchange. Chemical Engineering Science
55 (12), 21612171.
Holub, R.A., Dudukovic, M.P., et al., 1992. A phenomenological model for
pressure drop, liquid holdup, and ow regime transition in gasliquid
trickle ow. Chemical Engineering Science 47 (911), 23432348.
Iliuta, I., Larachi, F., et al., 1998. Pressure drop and liquid holdup in trickle
ow reactors: improved Ergun constants and slip correlations for the slit
model 37, 45424550.
Ismagilov, Z.R., Pushkarev, V.V., et al., 2001. A catalytic heat-exchanging
tubular reactor for combining of high temperature exothermic and
endothermic reactions. Chemical Engineering Journal 82 (13), 355360.
Ledoux, M.-J., Pham-Huu, C., 2001. Silicon carbide a novel catalyst support
for heterogeneous catalysis. Cattech 5 (4), 226246.
Ledoux, M.-J., Hantzer, S., et al., 1988. New synthesis and uses of highspecic-surface SiC as a catalytic support that is chemically inert and has
high thermal resistance. Journal of Catalysis 114 (1), 176185.
Leong, K.C., Jin, L.W., 2006. Characteristics of oscillating ow through a
channel lled with open-cell metal foam. International Journal of Heat
and Fluid Flow 27 (1), 144153.
Lu, T.J., Stone, H.A., et al., 1998. Heat transfer in open-cell metal foams.
Acta Materialia 46 (10), 36193635.
Maestri, M., Beretta, A., et al., 2005. Comparison among structured and
packed-bed reactors for the catalytic partial oxidation of CH4 at short
contact times. Catalysis Today 105 (34), 709717.
Pesant, L., 2005. Elaboration dun systme catalytique base de carbure de
silicium (b-SiC) pour la combustion des suies issues des automobiles
moteur diesel. LMSPC, Strasbourg, Louis Pasteur.
Pham-Huu, C., Keller, N., et al., 2002. Le carbure de silicium. Un nouveau
support pour la catalyse htrogne. Actualits Chimiques 257, 818.
Richardson, J.T., Peng, Y., et al., 2000. Properties of ceramic foam catalyst
supports: pressure drop. Applied Catalysis A: General 204 (1), 1932.
Richardson, J.T., Garrait, M., et al., 2003a. Carbon dioxide reforming with Rh
and Pt-Re catalysts dispersed on ceramic foam supports. Applied Catalysis
A: General 255 (1), 6982.
Richardson, J.T., Remue, D., et al., 2003b. Properties of ceramic foam catalyst
supports: mass and heat transfer. Applied Catalysis A: General 250 (2),
319329.

M. Lacroix et al. / Chemical Engineering Science 62 (2007) 3259 3267


Schimmoeller, B., Schulz, H., et al., 2006. Ceramic foams directly-coated
with ame-made V2O5/TiO2 for synthesis of phthalic anhydride. Journal
of Catalysis 243 (1), 8292.
Tien, C., Vafai, K., 1990. Advances in Applied Mechanics 27, 225.
Twigg, M.V., Richardson, J.T., 1994. Sixth International Symposium on the
Scientic Bases for the Preparation of Heterogeneous Catalysts. Elsevier,
Amsterdam, Netherlands.
Twigg, M.V., Richardson, J.T., 2002. Theory and applications of ceramics
foam catalysts. IChemE 80 (Part A), 183189.

3267

Wine, G., Tessonnier, J.-P., et al., 2006. Beta zeolite supported on a


[beta]-SiC foam monolith: a diffusionless catalyst for xed-bed FriedelCrafts reactions. Journal of Molecular Catalysis A: Chemical 248 (12),
113120.
Zamaro, J.M., Ulla, M.A., et al., 2006. Growth of mordenite on monoliths
by secondary synthesis: effects of the substrate on the coating structure
and catalytic activity. Applied Catalysis A: General 314 (1), 101113.

You might also like