Vous êtes sur la page 1sur 12

Fundamentals and Applications of Structured Ceramic Foam Catalysts

Martyn V. Twigg

and James T. Richardson*


,
Catalytic Systems DiVision, Johnson Matthey, Royston, Herts SG8 5HE, United Kingdom, and Department of
Chemical and Biomolecular Engineering, UniVersity of Houston, Houston, Texas 77204-4004
This paper reviews the use of ceramic foams as structured catalyst supports. They are open-cell ceramic
structures that may be fabricated in a variety of shapes from a wide range of materials, and they exhibit very
high porosities with good interconnectivity. These characteristics result in a lower pressure drop than that
observed with packed beds and high convection in the tortuous megapores, which, in turn, enhances mass
and heat transfer. They are easily coated with high-surface-area catalytic components, using well-established
techniques. Research in the past decade has produced a large amount of fundamental information that elucidates
the desirable properties of ceramic foams. In addition, many applications involving important reactions have
appeared in the open and patent literature, especially for catalytic processes that suffer certain limitations,
such as those encountered in relieving high pressure drop with low-contact-time reactions at high space
velocities or with narrow reactors in heat-transfer-limited systems and in controlling axial and radial temperature
profiles in highly exothermic and endothermic reactions. These important contributions are discussed, and
the advantages and shortcomings of using ceramic foams as structured catalyst supports to benefit commercial
operations are considered.
Introduction
In their book on modern chemical process technology,
Moulijn et al.
1
related many technical problems that can be
solved with structured catalyst supports replacing conventional
packed-bed reactors. Methods for accomplishing this have been
detailed in two reviews that were devoted solely to the subject
of structured catalysts.
2,3
Many types of structures were
examined; however, the use of foamed supports is only now
gaining momentum in the research literature. This paper reviews
the fundamentals and potential applications of ceramic foams
as catalyst supports, with an emphasis on comparisons with
packed beds of catalyst particles.
Cellular structures in natural materials (wood, bone, coral,
etc.) have long been known and studied. Significantly, nature
has optimized certain mechanical properties, such as stiffness,
strength, and mass, in an efficient manner; however, potential
engineering applications did not evolve until 50 years ago,
when manufacturing methods for making synthetic materials
were developed. This inspired investigation into the structure
and properties of cellular materials, and this was reviewed by
Gibson and Ashby.
4,5
Many useful practical applications fol-
lowed, e.g., insulating materials and lightweight structures. For
ceramic foams, their high-temperature characteristics led to use
as molten metal filters and kiln furniture, which, today, are the
largest applications. Later uses include burner enhancers, soot
filters for diesel engine exhausts, catalyst supports, and biomedi-
cal devices.
6-9
Ceramic foams are characterized as closed-cell and open-
cell, and, accordingly, the two characterizations have vastly
different properties. Closed-cell foams are composed of polyhedra-
like cells connected via solid faces, i.e., with no interconnectivity
between them, whereas open-cell structures have solid edges
and open faces, with fluid flow possible from one cell to another.
Figure 1 illustrates the open-cell structure with three compo-
nents: struts, which are composed of solid ceramic material;
cells, which are approximately spherical voids enclosed by
struts; and windows, which are openings connecting the cells
to each other. The open structure is sponge-like, with pore
densities of typically 10-100 PPI (pores per inch), creating an
interconnecting porosity in the range of 75%-90% or even
higher. Consequently, they have low resistance to fluid flow
and considerable turbulence is generated by the tortuous flow
paths. This review is restricted to open-cell ceramic foams,
because they have obvious implications for catalysis. Metal
foams, which have similar properties and applications but are
manufactured differently and have unique features, are also not
included, except where structural similarities add insight into
foam behavior.
The present authors reviewed the state-of-the art for catalytic
applications of ceramic foams and emphasized the following
advantages:
10,11
(1) the ability to preshape ceramic foam
cartridges to match complicated reactor configurations and
simplify reactor loading; (2) a lower pressure drop, relative to
packed beds, with savings in energy efficiency and low-contact-
time operation; (3) high geometric surface areas that improve
mass transport and lead to higher effectiveness factors for
processes that otherwise require large pellets; and (4)
* To whom correspondence should be addressed. E-mail: jtr@uh.edu.
Tel.: +1 (713) 743-4324. Fax: +1 (713) 743-4323.

Johnson Matthey.

University of Houston.
Figure 1. Open-cell, retriculated ceramic foam cell structure.
4166 Ind. Eng. Chem. Res. 2007, 46, 4166-4177
10.1021/ie061122o CCC: $37.00 2007 American Chemical Society
Published on Web 02/24/2007
enhanced radial convection, which improves heat transfer and
allows better reactor stability for highly exothermic reactions.
These previous reviews listed emerging applications at that time
as evidence for the benefits of improved performance. Now,
over a decade following initial interest, we cover advances in
improved characterization, modeling, and process applications.
In addition, we discuss potential weaknesses that could affect
commercial implementation.
Fabrication of Ceramic Foam-Supported Catalysts
Most commercial ceramic foams are prepared using a
reticulation method that was first patented by Schwartzwalder
and Somers.
12
The process starts (Figure 2) with polymer foam.
This foam is usually polyurethane, although other organic
polymers are equally effective. Polyurethane is readily available
in a wide range of cell sizes that determine the cell density of
the resulting ceramic.
The pores of the polymer foam are filled with an aqueous
slurry of ceramic (e.g., R-Al
2
O
3
, ZrO
2
, etc.) that is composed
of 0.1-10 m diameter particles, together with appropriate
amounts of wetting agents, dispersion stabilizers, and viscosity
modifiers.
13-15
In the most common approach, low-viscosity
suspensions are used and excess slurry is removed by blowing
air through the foam or by squeezing and kneading. The wet
foam is dried and calcined in air at temperatures of 1000-1700
C. With careful control, the plastic vaporizes and the ceramic
particles sinter, forming a ceramic replica or positive image of
the plastic. The surface of the struts surrounding the faces or
windows comprises the area available for supporting catalysts.
Alternatively, the viscosity of the slurry is increased by adding
thickening agents
16,17
and excess slurry is removed only at the
external surface of the structure, so the ceramic slurry remains
in the pores. After calcination and removal of the plastic,
ceramic replaces the voids in the original organic material,
giving a negative image of the plastic. Pore diameters are
smaller, bulk densities are higher, and porosities are lower;
however, the mechanical strength is much higher than that of
material prepared via the other method.
Preformed shapes, such as cylinders, rings, rods, or custom-
designed configurations, are fabricated by machining the plastic
before or after soaking and drying. Dispersion of the catalytic
agents onto the foam surface follows conventional techniques,
18
such as single or multiple impregnations of salts or adsorption
of ionic precursors from solution, followed by heat treatment
at moderate temperatures.
Measured Brunauer-Emmett-Teller (BET) surface areas of
the foams are low (1-2 m
2
/g) but may be increased by adding
a high-area washcoat, such as -Al
2
O
3
. Typically, a dilute slurry
containing the washcoat precursors (such as boehmite, with
binders and small amounts of viscosity-modifying agents) is
added to the dry foam, which is then drained and the excess
slurry is removed. This is followed by calcination at moderate
temperatures (<500 C), to preserve the -Al
2
O
3
phase. The
desired surface area is achieved by repeating the procedure, and
the BET surface area increases linearly with the washcoat
loading (see Figure 3).
19
Back-electron imaging (Figure 4) shows
a porous washcoat of -Al
2
O
3
firmly adhering to R-Al
2
O
3
, with
a thickness of 50 m.
20
Washcoated foams have been
successfully loaded with metals, oxides, zeolites, and carbon
using this technique.
10
Structure and Characterization of Ceramic Foams
Structural Parameters. Table 1 shows the properties avail-
able with commercial ceramic foams produced via the reticula-
tion method. These properties are pore density (PPI), pore size
(d
p
, the diameter of the pore opening between each cell), bulk
density (F
B
, mass per unit volume of foam), porosity (, the
fraction of the pore volume between the solid struts (i.e., the
open megaporosity subject to fluid flow and not including closed
space within the struts)), external surface area (S
v
, the geometric
surface area per unit volume of solid), and equivalent spherical
diameter (d
s
, the diameter of solid spheres with an external
surface area of S
v
). Table 1 shows that, with pore densities and
diameters varying by almost a factor of 10, bulk
Figure 2. Steps in the fabrication of ceramic foam using the reticulation
technique.
Figure 3. Surface areas obtained from multiple loading of -Al2O3 on
30-PPI R-Al2O3 foam (from ref 19).
Figure 4. Backscattered electron image of the ceramic foam strut.
Ind. Eng. Chem. Res., Vol. 46, No. 12, 2007 4167
densities and porosities remain approximately the same. As
expected, S
v
increases and d
s
decreases as d
p
decreases. Low
values of d
s
in Table 1 emphasize that ceramic foams in this
size range exhibit similar transport and surface reaction behavior
to packed beds of extremely small particles.
Pore density (PPI) is an approximate commercial character-
istic designated by the manufacturer of the organic foam
precursor. However, pore diameter is a measured property.
Typically, cylindrical foam pellets are cut along axial and radial
dimensions into thin slices and images are taken of each slice
with a digital camera.
21-23
Each image is scanned and, after
converting each pore area to the diameter of an equivalent circle,
data from all segments are combined to give a pore-size
distribution. Typical results are shown in Figure 5 for a 30-PPI
R-Al
2
O
3
foam.
22
Normal and log-normal distributions give
equally good fits, from which the average pore or window size
(d
p
) is determined. Washcoats of up to 15 wt % do not seem to
have any effect on the distribution, although decreases in d
p
are consistent with the thicknesses of the washcoat layer.
22
Bulk density (F
B
) is determined from the mass of the ceramic
foam structure divided by the geometrical volume, which
includes the volume of the solid material, the open volume, and
any closed volume inside the solid struts. The latter is formed
when the organic polymer is vaporized, leaving behind a hollow
space, which is dependent on the exact manufacturing condi-
tions, and is not normally accessible to the fluid flow. To
compensate for this, solid density (F
S
) is measured with a helium
pycnometer and the open porosity is calculated from
Usually, only the parameters d
p
, F
B
, and F
S
are measured, with
being calculated from eq 1 and S
v
being determined from a
model that relates S
v
to d
p
and . Recently, more-sophisticated
X-ray topography microscopy has been applied that provides
direct measurement of the porosity; however, the results seem
to be less than those determined from conventional methods.
23-25
Gibson and Ashby
4,5
reviewed models of foam structures as
regular packing of various polyhedra, such as cubic cells,
triangular prisms, rectangular prisms, hexagonal prisms, rhombic
dodecahedra, pentagonal dodecahedra, and tetrakaidecahedra.
These authors preferred tetrakaidecahedra, because they are most
consistent with the observed properties. Figure 6 shows a unit
tetrakaidecahedron and the ideal representation of the strut
architecture.
22
Each cell comprises 14 faces or windows (pores)
connected with neighboring cells to form a three-dimensional
structure. The windows (8 hexagonal and 6 square) are bordered
by triangular struts. Gibson and Ashby
4
derived geometric
relationships for the unit cell, and Richardson et al.
22
used these
relationships to derive an expression for S
v
:
Buciuman and Kraushaar-Czarnetzki
26
reaffirmed this approach
and, while developing similar equations, discussed the ap-
proximate nature of this model that is due to complex physical
differences among the struts arising from manufacturing vari-
ables. The same type of equations may be derived for other
structures suggested for the foams. For example, Richardson et
al.
22
developed equations for the pentagonal dodecahedron,
which is another favorite model, and determined that they were
very similar to those of the tetrakaidecahedron. Other models
examined with studies on metal foams include cubic cells,
27
dodecahedra,
28
and a regular packing of polyhedra.
29
Richardson et al.
22
considered the hydraulic diameter model
commonly applied to packed beds, in which pores are uniform,
parallel cylinders, and derived the following relationship:
These authors concluded that values for S
v
calculated from
commonly used models are approximately the same, within the
precision permitted by the model approximations. For example,
a 30-PPI foam (Table 1) gives values of 2.60 10
4
, 3.38
10
4
, and 2.52 10
4
m
2
/(m
3
solid) for the hydraulic, cubic cell,
and tetrakaidecahedron models, respectively. Therefore, the
hydraulic model was adopted because of its simplicity and
common acceptance. However, Giana et al.
24
advocated the
cubic cell, especially for metal foams, which generally have
higher porosities than their ceramic counterparts.
In regard to describing transport and other properties of foams,
there is a wide variation in the literature for the characteristic
Table 1. Properties of Commercial Ceramic Foam Supports (92%
r-Al2O3, 8% Mullite) without Washcoat
a
pore density
(PPI)
dp
(mm)
bulk density
(g/cm
3
) porosity
Sv 10
-4
(m
2
/(m
3
solid))
ds
b
(mm)
10 1.52 0.51 0.87 1.76 0.34
20 0.94 0.61 0.85 2.41 0.25
30 0.75 0.66 0.83 2.60 0.23
45 0.42 0.65 0.84 5.00 0.12
65 0.29 0.70 0.81 6.88 0.09
80 0.21 0.67 0.83 9.30 0.06
a
Data supplied by Hi-Tech Ceramics, Alfred, NY.
b
Diameter of a sphere
with the same value of Sv.
Figure 5. Pore-size distribution for 30-PPI R-Al2O3 foam, unwashcoated
(from ref 22).
) 1 -
F
B
F
S
(1)
Figure 6. (a) Ideal tetrakaidecahedron structure, (b) strut structure around
the faces, and (c) cross section of the strut.
S
v
)
12.979[1 - 0.971(1 - )
0.5
]
d
p
(1 - )
0.5
(2)
S
v
)
4
d
p
(1 - )
(3)
4168 Ind. Eng. Chem. Res., Vol. 46, No. 12, 2007
length parameter used to compute Reynolds numbers and mass-
or heat-transfer coefficients. These are usually d
p
or d
s
.
Richardson et al.
22
argued that d
s
in packed-bed models come
from imposing a spherical geometry onto the more-fundamental
S
v
, and they recommended using the latter for dimensionless
parameters, such as the Reynolds number:
where F is the fluid density, the viscosity, and V the superficial
velocity. Unfortunately, S
v
is dependent on d
p
, , and the choice
of model. Because the struts or ceramic elements of foams
contain cracks and fissures, due to manufacturing methods,
surface areas measured via the BET method vary from 1 m
2
/g
to 2 m
2
/g (i.e., 2 orders of magnitude higher than geometric
surface areas) and are not suitable. At this time, there is no
reliable, independent method for measuring S
v
.
Pressure Drop. Pressure drop is important in reactor design,
and the most widely accepted interpretation is the Forscheimer
equation, which was adapted by Ergun and Orning
30
for packed
particle beds:
where R and are parameters that are dependent on the
geometry and packing of the particles. (Note that eq 5 is the
original form derived by Ergun and Orning; the more-familiar
expression involving d
s
results from substituting the relationship
S
v
) 6/d
s
). Cellular materials such as ceramic foams do not
exhibit the same geometry as packed beds of particles (e.g.,
void boundaries are concave rather than convex); yet, most
authors follow the same procedures in developing pressure drop
correlations for foams. Many studies have been reported
22,31-47
that show the pressure drop following the Forscheimer equation,
but with a wide variation in the parameters derived.
Richardson et al.
22
determined that the Ergun parameters (R
and ) are dependent on d
p
and , so that R decreases and
increases as the PPI increases (Table 2). This is not surprising,
because Ergun and Orning
30
reported R and values in the range
of 3.6-20 and 0.14-0.70, respectively, but they recommended
values of R ) 4.17 and ) 0.292 as universal constants.
Macdonald et al.
48
confirmed these conclusions but preferred
values of R ) 5.00 and ) 0.300 for smooth particles and
) 0.667 for rough particles. Most authors agree that R and
are not constants but rather parameters that are dependent on
the medium. For example, Macdonald et al.
48
concluded that a
value of R ) 10.6 is better for consolidated media. Richardson
et al.
22
developed empirical expressions for R and that are
useful for reactor modeling; however, there is no fundamental
rationale for the dependence.
Further research by Richardson et al.
22
showed that both R
and increase as washcoat is added to the surface. Because
the pore-size distribution does not change appreciably with
added washcoat, this was believed to be due to the increased
texture of the surface, rather than to blocking of the pores. This
is similar to the effect of roughness in pipes and suggests
that the pressure drop of ceramic foams is dependent not only
on d
p
and but also on the physical characteristics of the surface,
implying that the parameters for pressure drop correlations
should be tested for each application and that permeability
measurements should be a routine part of foam characterization.
Heat-Transfer Correlations. The second important foam
property that affects catalytic applications is enhanced heat
transfer produced by convection due to greater turbulence in
tortuous pores and radiation between struts. Most work with
ceramic foams has focused on the measurement of volumetric
heat-transfer coefficients.
49-57
Most authors correlated their
results using
in which h is the heat-transfer coefficient of the film, Re is the
Reynolds number (based on d
p
or d
s
), and C
1
and m are empirical
constants, although Zumbrunnen et al.
49
advocated the addition
of a radiation term, whose effect increases with porosity.
However, mathematical simulation of reactors with one-
dimensional (1-D) reactor models requires an appropriate radial
heat-transfer correlation, and this was developed by Peng and
Richardson
54
from an axial temperature profiles in a reactor
containing a 30-PPI R-Al
2
O
3
foam. The resulting correlation
was
where h
e
is the overall heat-transfer coefficient and
g
is the
thermal conductivity of the fluid. The first term in eq 7
represents the combined solid thermal conductivity and the
radiation component; the second term is the convective heat
transfer. Furthermore, these authors found the second term in
eq 7 varied with washcoat loading, e.g., the value of the constant
for 6 and 8 wt % loading increased to 0.0601 and 0.0708,
respectively, although the first term was essentially unchanged.
Using this expression, it is possible to compare equivalent
(i.e., same S
v
value) beds of foam and particles, and this is
demonstrated in Figure 7. The ratio of the overall radial heat-
transfer coefficient of the foam to the particles increases from
5 to 10 at low and high Reynolds numbers, respectively, and
to a lesser extent, from low to high temperatures.
Equation 7 is suitable for 1-D modeling of endothermic
processes; however, exothermic reactions at lower temperatures
are more demanding. Average bed temperatures derived from
Table 2. Ergun Equation Parameters for 99.5 wt % r-Al2O3 Foams
without Washcoat
a
PPI R
10 8.05 0.0338
30 7.72 0.0838
45 4.85 0.0872
65 2.88 0.111
a
Data taken from ref 22.
Figure 7. Comparison of heat transfer between equivalent ceramic foam
and particle beds (from ref 54).
h ) C
1
Re
m
(6)
h
e

g
S
V
) (3.43 10
-11
)T
3
+ 0.0340Re
S
(7)
Re
s
)
FV
S
v

(4)
dP
L
)
[
RS
v
2
(1 - )
2

3 ]
V +
[
S
v
F(1 - )

3 ]
V
2
(5)
Ind. Eng. Chem. Res., Vol. 46, No. 12, 2007 4169
1-D models are not sufficient for reactions that release large
amounts of heat, because radial temperature profiles can be
severe. Two-dimensional (2-D) pseudo-homogeneous models
are more appropriate. This requires two parameters:
e
, which
is the effective radial conductivity of the bed, and R
W
, which is
the heat-transfer coefficient at the wall. By fitting radial and
axial temperature profiles in 30-PPI R-Al
2
O
3
foam, Peng and
Richardson
57
determined expressions for
e
and R
W
:
and these were used in subsequent 2-D modeling of foam reactor
applications.
It should be emphasized that eqs 7-9 refer to 30-PPI cases
only and do not apply to other pore densities or the presence of
washcoat. They should be used with caution until more
information is available or more exact expressions are derived
from theoretical analysis. The present form is appropriate for a
first pass analysis; however, at some point, the parameters
must be calibrated by comparisons to measured temperature
profiles.
Examples of Exothermic and Endothermic Processes
The properties previously outlined suggest replacing packed
beds of traditional catalyst particles with cartridges of ceramic
foam could be beneficial to certain industrial processes.
Important advantages are (1) the ability to match the shape and
size of the reactor, giving easier loading of many long and
narrow tubes; (2) greatly reduced pressure drop, saving process
energy costs and providing for larger space velocities; (3) larger
external surfaces, providing higher effectiveness factors; and
(4) enhanced heat transfer, decreasing axial and radial temper-
ature profiles and avoiding hot spots. These qualities must be
balanced with a lower solid loading, compared to packed beds
(i.e., the foam structure only fills 10%-20% of the reactor
volume with solid, compared to 45%-55% for particles). If
the decreased activity per unit volume of reactor is not recovered
by a higher effectiveness factor (i.e., the particles already have
a high effectiveness factor and the process is now mass-transfer
limited), activity must be improved by adding a high-surface-
area washcoat or increasing the loading of the active component.
Practical benefits from using foam supports include easier
operation, smaller reactors, higher yields and selectivities, and
improved temperature control and heat management. Because
of complex interactions between many factors, technical and
economic assessment for replacing packed particles with
structured foam supports must be considered on a case-by-case
basis, as demonstrated in the examples discussed below.
Heat-Transfer-Limited Processes. Many industrial reactor
operations are heat-transfer-limited, both for endothermic and
exothermic reactions. Examples of important industrial processes
of this type are given in Table 3. The necessary surface for
heat transfer into or from the process is achieved with multiple
long reactor tubes of small diameter, causing a significant
pressure drop. To overcome this, the catalyst pellets are made
larger, which increases radial heat transfer but the effectiveness
factor decreases and more catalyst volume is needed. Ceramic
foam cartridges fabricated to fit the tube precisely can relieve
these problems by delivering higher heat transfer, a lower
pressure drop, and larger effectiveness factors.
Brown
19
demonstrated possible benefits for these processes
with laboratory experiments on ethylene epoxidation, carbon
dioxide methanation, and Fischer-Tropsch synthesis. These
reactions are examples of exothermic processes with different
constraints on the foam. Table 4 summarizes industrial process
conditions and the rationale for selecting these processes. The
important differences are observed relative to surface area
(Fischer-Tropsch > methanation > epoxidation) and effective-
ness factors (epoxidation > methanation > Fischer-Tropsch).
Because the purpose was to demonstrate the feasibility and
advantages for the foam-supported catalysts, no attempt was
made to incorporate promoters to improve activity/selectivity
or to operate under optimum process conditions.
(1) Ethylene Epoxidation. Ethylene epoxidation involves two
main reactions:
where reaction 11 is undesirable. The commercial catalyst is
10 wt % Ag/R-Al
2
O
3
in the form of pellets 3-12 mm in size
packed into tubular reactors 2-5 cm in diameter, operating at
220-235 C. The conversion per pass is limited to 10%,
because temperature increases in excess of 30-40 C are
detrimental to both catalyst stability and C
2
H
4
O selectivity.
Spheres of R-Al
2
O
3
without added washcoat were impregnated
with a silver nitrate solution, dried at 120 C, and calcined at
450 C. The foams were treated in the same manner, except
they were rotated at a rate of 12 rpm during drying in a

e
) 6.84 10
-5
(
1
S
V
)
T
3
+ 42.2Re
S

g
(8)
R
W
) 0.0692S
V
Re
S
0.48

g
(9)
Table 3. Classification of Some Industrial Catalytic Processes for
Which Ceramic Foam Catalyst Supports Could be Beneficial
a
type example
Exothermic Nature
partial oxidation ethylene to ethylene oxide
partial oxidation o-xylene to phthalic anhydride
partial oxidation propene to acrylic acid
partial oxidation butane to maelic anhydride
partial oxidation methanol to formaldehyde
alkylation benzene to ethyl benzene
alkylation diethylbenzene or cumene
oxidative rearrangement water gas shift
oxychlorination acetic acid to vinyl acetate
oxychlorination ethylene to ethylene dichloride
hydrogenation methanol synthesis
hydrogenation methanation of CO/CO
2
hydrogenation Fischer-Tropsch synthesis
Endothermic Nature
dehydrogenation ethylbenzene to styrene
dehydrogenation cyclohexane to benzene
dehydrogenation cyclohexanol to cyclohexanone
dehydrogenation butanol to methyl ethyl ketone
steam reforming natural gas to synthesis gas
steam reforming naphtha to synthesis gas
a
Data taken from ref 11.
Table 4. Conditions for Three Exothermic Examples
a
ethylene
epoxidation
carbon dioxide
methanation
Fischer-Tropsch
synthesis
catalyst Ag metal Ru metal Co metal
support R-Al2O3 -Al2O3 -Al2O3
BET low, <2 m
2
/g moderate, 100 m
2
/g high, >200 m
2
/g
washcoat no yes yes
BET 1-2 m
2
/g 10 m
2
/g 30-40 m
2
/g
reactions 2 1 many
a
Data taken from ref 19.
C
2
H
4
+ 0.5O
2
) C
2
H
4
O (H
r
) -105 kJ/mol) (10)
C
2
H
4
+ 3O
2
) 2CO
2
+ 2H
2
O (H
r
) -1334 kJ/mol)
(11)
4170 Ind. Eng. Chem. Res., Vol. 46, No. 12, 2007
microwave oven for 1 min to ensure uniform deposition
throughout the foam. Composition and structure of the samples
were confirmed by X-ray diffraction. Rate measurements were
made under differential flow conditions (<10% conversion)
using a 13-mm inner diameter (ID) quartz reactor containing a
length of 1.27 cm of 100 mesh powder, spheres, or ceramic
foam. Kinetic measurements revealed that the reaction is first
order in ethylene and zero order in oxygen. Further details are
given by Brown.
19
Table 5 shows that the activity per gram of
silver is preserved on the foam using this preparation technique,
confirming that this is a viable substitution. However, the
effectiveness factor for the 1.8-mm pellet is 1, so the activity
per unit volume of the reactor with foam catalyst is lower.
Therefore, it is necessary to increase the metal loading for the
same volume activity.
A better understanding of the heat management impact of
the foam replacement comes from a 2-D model simulation
(Figure 8) of the commercial process for ethylene epoxidation
under typical industrial conditions.
58
Strong axial and radial
temperature gradients develop at the front of the bed, reaching
a center peak at 269 C, with a radial gradient of 20 C and
temperature runaway predicted at 274 C, so there is very little
flexibility. To prevent the occurrence of runaway, conversion
is limited to 9.3%, which requires expensive recycle of the
unconverted ethylene. The high-temperature peak leads to
catalyst deactivation (and selectivity loss), for which compensa-
tion usually is made industrially by increasing the reactor inlet
temperature; however, the model indicates that the increase
cannot exceed 247 C, so the life of the process operations is
thereby restricted.
Table 6 compares these results with those of the foam catalyst,
using the same kinetics but inserting heat-transport correlations
(eqs 8 and 9) for the foam. In case 1, the silver loading of the
foam is increased to 12.5 wt % to give the same final conversion
and productivity as the commercial pellets. The most dramatic
change is the decrease in the difference between the wall and
center temperature from 20 to 2 C (i.e., the radial profile is
negligible) and the increase of the inlet temperature for runaway
to 264 C. Case 2 examines the effect of increasing the silver
loading to 22.5 wt %. Although the radial temperature profile
is slightly higher and the maximum inlet temperature decreases,
productivity improves by 60%. The benefits that are observed
using foam with 12.5 wt % silver are ease of loading, lower
pressure drop, and improved stability and lifetime, but for 22.5
wt % silver, there is also less recycle and increased productivity.
These are a few examples of the types of benefits possible. Other
benefits, such as an increased reactor diameter, resulting in
decreased capital costs, can also be considered and economic
comparisons can be used to guide process improvements.
(2) Carbon Dioxide Methanation. Methanation, which is
an example of a highly exothermic hydrogenation with a
moderate catalyst surface area, has been widely used com-
mercially in synthetic natural gas (SNG) and ammonia plants.
The exothermic reaction is
which is reversible and reaches almost complete conversion of
CO
2
at 200 C. The reaction is easily catalyzed by ruthenium
on -Al
2
O
3
. Brown
19
calcined 1.8-mm -Al
2
O
3
spheres at 900
C to achieve a BET surface area of 100 m
2
/g and impregnated
the spheres with an ethanolic ruthenium chloride solution that
contained a small amount of HCl. After soaking for several
hours, the spheres were removed, dried, and calcined at 550
C. The 30-PPI R-Al
2
O
3
foam equivalent was washcoated with
10 wt % of crushed -Al
2
O
3
spheres and impregnated as done
previously. Rate measurements were conducted in the same
manner, except the samples were reduced at 350 C in hydrogen.
Rate comparisons (Table 7) show that, unlike ethylene epoxi-
dation, the pellets have a low effectiveness factor that is almost
compensated by the higher value for the foam (with a slight
decrease due to the larger size of the ruthenium crystallite), and
the foam volume activity is almost twice that of the pellets,
Table 5. Ethylene Epoxidation Rate Comparisons
a
powder
1.8 mm
spheres foam
BET surface area (m
2
/g) 1.37 1.37 1.66
bed density (g/cm
3
, bed)) 1.06 0.60
Ag loading (wt %) 9.90 9.90 12.1
Ag crystallite size (nm) 8.9 8.9 7.2
rate of reaction (mol g
-1
s
-1
)
0.5 atm C
2
)
, 225 C
1.4 10
-6
1.3 10
-6
1.5 10
-6
effectiveness factor 1 1 1
rate at 1 atm
(mol s
-1
cm
-3
, bed))
1.38 10
-6
0.90 10
-6
a
Data taken from ref 19.
Figure 8. Simulation of an epoxidation commercial reactor with a foam
catalyst. Reactor: 4 cm in diameter, 12 m in length; catalyst: 10% Ag/R-
Al2O3, 5-mm rings. Inlet conditions: 46% C2H4, 4% CO2, 11% O2, and
39% CH4; gas hourly space velocity, GHSV ) 2890; 242 C; 20 atm (from
ref 58).
Table 6. Comparison of Simulation Results of a Commercial
Reactor Tube (Same Conditions as Figure 6) Loaded with
Commercial Catalysts and Ceramic Foam Structures
a
property pellet bed
foam bed
case 1
foam bed
case 2
Ag loading (%) 10 12.5 22.5
C2H4 inlet (mol/h) 489 489 489
C2H4 conversion (%) 9.3 9.3 14.8
e (W/m) 3.26 28.4 28.4
Rw (W/m
2
) 732 336 336
maximum center temp (C) 269 251 263
maximum radial gradient (C) 20 2 5
inlet temperature for runaway 247 264 247
productivity ((kg of C2H4)/h) 1.38 1.38 2.21
a
Data taken from ref 58.
Table 7. Carbon Dioxide Methanation Rate Comparisons
a
powder
18-mm
pellets foam
wt % Ru 0.90 0.90 0.85
BET surface area (m
2
/g) 139 139 13.3
Ru crystallite size (nm) 16.1 16.1 19.4
rate of reaction (mol s
-1
g
-1
)
0.2 atm CO2, 225 C
9.59 10
-3
2.22 10
-3
6.98 10
-3
effectiveness factor 1 0.23 1
rate (mol s
-1
cm
-3
, reactor) 2.44 10
-3
4.47 10
-3
a
Data taken from ref 19.
CO
2
+ 4H
2
) CH
4
+ 2H
2
O (H
r
) -175 kJ/mol)
(12)
Ind. Eng. Chem. Res., Vol. 46, No. 12, 2007 4171
indicating potentially larger productivity for reactors. Model
studies on this system are in progress and will be reported later.
(3) Fischer-Tropsch Synthesis. This is the most demanding
example, because of the complexity of both the catalyst and
the reactions involved. Brown
19
simulated commercial Co/-
Al
2
O
3
Fischer-Tropsch catalysts by impregnating 1.8-mm
-Al
2
O
3
spheres with a cobalt nitrate solution to give 20 wt %
cobalt after drying at 105 C and calcining at 350 C. To match
this, the surface area for the foam should be 40 m
2
/g, which
requires a washcoat with a high surface area. After experiment-
ing with several alternative approaches, a coprecipitation
procedure was adopted that gave the highest surface area at 20
wt % cobalt. XRD showed the presence of Co
3
O
4
and -Al
2
O
3
.
The powder was slurried with water and foam samples were
washcoated with repeated cycles to give a washcoat loading of
17 wt %, corresponding to a surface area of 33 m
2
/g.
Rate comparison test results (Table 8) showed the 1.8-mm
pellets with an effectiveness factor of 0.63 and the foam
successfully reproducing the volume activity of the pellets with
an increase in selectivity to higher hydrocarbons. This example
demonstrates that complex reaction chemistry may be success-
fully incorporated into the foam structure. Future studies will
examine the effect of improved heat transfer.
These examples demonstrate the advantages of ceramic foam
supports for different types of exothermic reactions. For
endothermic processes, an important example is the production
of synthesis gas with steam or carbon dioxide reforming of
methane.
(4) Methane Reforming. Richardson et al.
59
demonstrated
the kinetic superiority of ceramic foams for dry reforming of
methane,
using 3.2-mm, 0.5-wt % Rh/-Al
2
O
3
spheres and 0.5 wt % Rh/
30-PPI R-Al
2
O
3
foam with -Al
2
O
3
washcoat. Rates were
measured in a differential reactor and converted to turnover
frequencies using rhodium surface areas measured by H
2
chemisorption. The results in Figure 9 clearly indicate that the
turnover frequencies of rhodium on the foam are 10 times
larger than those on the pellets, as anticipated from a high
effectiveness factor that was possible with the foam.
Methane steam reforming is a major industrial process that
proceeds via reaction 14:
The results of a 2-D simulation comparing commercial pellets
and equivalent foam supports is shown in Figure 10 for typical
industrial conditions.
60
The foam catalyst reaches maximum
equilibrium conversion within approximately one-half the tube
length of conventional catalysts, and temperature gradients in
the front of the bed are considerably reduced. This is a
consequence of improved radial heat transfer and, to a lesser
extent, a higher effectiveness factor, which provides a faster
approach to equilibrium. These results support earlier predictions
reported by Twigg and Richardson,
10
using a 1-D model (Table
9). Simulations have also been reported by Thurgood et al.
61
and Reitzmann et al.
62
for methanol-steam reforming and
o-xylene oxidation, respectively, with similar results.
These experiments and simulations confirm the inherent
advantages of ceramic catalyst supports. Other examples from
the literature agree with this observation and are summarized
in the following section.
Other Process Applications
Most reported applications of ceramic catalyst supports are
concerned primarily with the chemistry that occurs on the strut
surfaces rather than the transport properties previously outlined.
They fall into the following classifications: short-contact-time
reactions, catalytic combustion, fuel processing, solar methane
reforming, and miscellaneous processes.
Table 8. Fischer-Tropsch Synthesis Rate Comparisons
a
powder
1.8-mm
pellets foam
wt % Co 20 20 20
BET surface area (m
2
/g) 183 183 33.0
Co crystallite size (nm) 8.0 8.0 6.4
rate of reaction (mol s
-1
g
-1
)
0.2 atm CO2, 225 C
6.21 10
-3
3.93 10
-3
9.33 10
-3
effectiveness factor 1 0.63 1
rate (mol s
-1
cm
-3
, reactor) 6.13 10
-3
5.97 10
-3
% selectivity (23% conversion)
C1 54 52 24
C2-C7 41 47 71
>C8 5 1 5
a
Data taken from ref 19.
CH
4
+ CO
2
/2CO + 2H
2
(13)
CH
4
+ H
2
O ) CO + 3H
2
(14)
Figure 9. Comparison of turnover frequencies for 30 PPI R-Al2O3 foam
and pellet catalysts (from ref 59).
Figure 10. Simulation of a methane steam reforming reactor tube under
industrial conditions (from ref 60).
Table 9. Simulation Comparison between a Conventional
Packed-Bed Reformer and a Ceramic Foam Support
a
property conventional foam
tube length for reaction 916 cm 439 cm
effectiveness factor at outlet 0.05 1
average heat flux 31.2 kW m
2
67.1 kW m
2
pressure drop 1.21 atm 0.14 atm
a
Data taken from ref 10. Conditions were as follows: tube diameter,
10 cm; CH4 flow, 1500 mol/h; H2O/CH4 ratio ) 3; wall temperature, 800
C; inlet temperature, 550 C; pressure, 20 atm; convention catalyst,
multihole cylinders; ceramic foam, 30 PPI R-Al2O3.
4172 Ind. Eng. Chem. Res., Vol. 46, No. 12, 2007
Short-Contact-Time Reactions. An extensive series of
papers have appeared on a special application of ceramic
foams.
63-99
Unlike the previous applications that enhance the
performance of already existing processes, these papers, which
are inspired by Schmidt and co-workers, explore the chemistry
at the strut surface and point the way to novel processes. As
part of a general study on structured supports (foams, honey-
comb monoliths, and other structures), a wealth of information
has developed on the chemistry of important reactions at short
contact times (on the scale of milliseconds), facilitated by low-
pressure drop at high flow rates. This is desirable for reactions
in which desired intermediates are generated during consecutive
reactions. The studies generally involve partial oxidation
63-79
and oxidative dehydrogenation
80-99
(i.e., reactions with oxygen
under rich conditions that are exothermic and produce fuels and
olefins, respectively). In the first type of reaction, CH
4
reacts
with limited O
2
, as in reaction 15:
Normally, CO and H
2
are produced first and are then converted
to CO
2
and H
2
O, consuming all the O
2
. The unconverted CH
4
is subsequently reformed via reactions 13 and 14 back to CO
and H
2
. At contact times on the order of milliseconds, Schmidt
et al. measured extremely high selectivities for CO and H
2
,
which they believe are the primary products. Although there is
some controversy about this observation, they clearly demon-
strated the advantage of the ceramic foam catalysts, not only in
providing the short contact time but also in reducing the reactor
to a fraction of its normal size.
This is especially true when the concept is applied to higher
hydrocarbons,
73-76
alcohols,
77
and biofuels,
79,99
where the
incentive is to develop small-scale reformers for liquid hydro-
carbons and renewable feeds for fuel-cell hydrogen production.
The objective of oxidative dehydrogenation
80-99
is the
production of valuable olefin feedstock from alkanes through
extremely low-contact reactions, such as that in reaction 16,
while not permitting oxidation of the alkene product. Alkenes
from C
2
-C
4
have been formed over ceramic foams at mil-
lisecond contact times during autothermal operation with high
alkene selectivity (70%) and alkane conversion (70+%), without
carbon formation using catalyst beds 10-100 times smaller than
those observed in current technologies. This work, in conjunc-
tion with reaction modeling, resulted in greater insight into
chemical mechanisms for these reactions, including the role of
surface-assisted homogeneous reactions.
Catalytic Combustion. An early use for ceramic foams was
gas-burner enhancement, where the turbulent mixing of air and
fuel and improved radiation within the structure gives very
uniform combustion characteristics for flames.
6,7,9
By adding
catalytic components to the foam, combustion occurs at lower
temperature, thereby decreasing NO
x
emissions. Catalytic
combustion using a wide range of ceramic foams and catalytic
agents have reinforced these concepts.
100-107
Two applications
are apparent: (1) for process heating up to 1000 C, high
radial heat-transfer rates with foams decrease the size of the
catalyst beds, and (2) for gas-turbine applications, the adiabatic
bed has better mixing than its honeycomb counterpart and there
is less possibility of the development of hot spots.
107
Thermal
shock and size issues should be the same, because identical
materials are used for both foams and monoliths.
Fuel Processing. Many of the features previously discussed
suggest that ceramic foam catalysts could fill the niche required
for producing hydrogen for fuel-cell applications, either for on-
board vehicle processing or at hydrogen refueling stations. Their
efficiency in reforming a wide range of hydrocarbons has been
demonstrated, with enhanced heat transfer allowing a reduction
in reactor size, thereby permitting compact units. This has been
demonstrated in several studies, using both metal and ceramic
foams for alcohol and hydrocarbon reforming, water gas shift
conversion, and selective CO oxidation.
108-113
Solar Methane Reforming. An interesting application of
ceramic foams has been solar methane reforming with CO
2
.
114-118
The concept is to use the endothermic reaction (eq 13) to absorb
heat from a concentrated solar flux, producing CO and H
2
that
can be stored and later used to generate heat via reaction 17:
with methane that is formed being recycled in a closed system.
Most applications envisage a parabolic dish collector with a
ceramic foam absorber receiver (reactor) at the focal point. The
receiver comprises a thin, but wide, dish-shaped ceramic foam
loaded with reforming catalyst (rhodium) located in a metal
reactor with a quartz window. This, indeed, is a true test of the
pre-shaping feature of ceramic foam supports, as well as
operation under cyclic and transient conditions. The ceramic
foam catalysts performed well throughout many trials.
Miscellaneous Processes. Other processes involving ceramic
foam catalysts that have received recent attention include
selective methanol oxidation,
119
methanol to olefins,
120
diesel
exhaust gas,
121
the destruction of chlorinated hydrocarbons by
steam refoming
20
and photocatalysis,
122
and countercurrent two-
phase flow reactors.
123
Commercial Aspects
Ceramic foam-based catalysts have many advantageous
properties that can be exploited in industrial applications. In
addition to the research work referenced in this paper, there
have been plant trials involving relatively small and quite large
ceramic foam-based catalysts. In several instances, the catalytic
performance was encouragingly good; however, as yet, there
seems to be no full-scale use of these catalysts. Here, we explore
why foam-based catalysts have not yet replaced traditional
catalysts in two important areas: monolithic autocatalyst
applications and processes using conventional pellets, granules,
or small extruded catalysts.
Monolithic Autocatalysts. The development of catalysts for
controlling exhaust gas emissions from cars began in earnest
in the early 1970s, following the U.S. Clean Air Act of 1970.
At first, an oxidation catalyst was used alone, then two-stage
reduction/oxidation systems were fitted that, in turn, were
replaced by single three-way catalysts (TWCs).
124-127
Much
later, diesel-powered cars were fitted with oxidation catalysts.
Initially, some companies used traditional pellet catalysts in flat,
relatively thin reactors, in which conversion was limited by gas
bypass. Loose pellets in these reactors suffered attrition
problems, because of gas pulsations and vehicle vibrations.
Other companies experimented with ceramic and metal foil
monolithic honeycombs that were washcoated with a thin layer
of high-surface-area catalyst. Finally, extruded cordierite mono-
liths gained prominence, after extrusion die technology had been
mastered.
128,129
At this time, ceramic foam had been introduced
CH
4
98
O
2
CO + 2H
2
98
O
2
CO
2
+ 2H
2
O (15)
i-C
4
H
10
+
1
2
O
2
) i-C
4
H
8
+ H
2
O (16)
CO + 3H
2
) CH
4
+ H
2
O (17)
Ind. Eng. Chem. Res., Vol. 46, No. 12, 2007 4173
for filtering liquid metal in casting aluminum and its alloys,
and other potential applications were sought. By the early 1980s,
there was a flurry of activity seeking auto catalyst applications
for ceramic foams. Washcoating was used, as with extruded
ceramic monoliths, to overcome the low surface area of the
ceramic foam. The resulting catalyst had high activity; however,
by then, extruded monolithic honeycomb technology was well-
established, so the introduction of foam-based autocatalysts was
hindered by several factors, including the following:
(1) Extruded cordierite monoliths were well-established and
were difficult to displace.
(2) Ceramic foam was brittle and did not seem to have the
strength and durability of extruded ceramic monoliths.
(3) The importance of providing foam catalyst with an outer
solid skin to facilitate incorporation into a metal can was not
properly appreciated at an early stage.
(4) Foam structures of similar dimensions as an extruded
monolith had higher backpressure.
(5) The commercial introduction of ceramic foam catalyst
was thought to be relatively costly; because the manufacturing
process was complex and expensive, new investment was
required to manufacture it.
Recently, there has been much research on the use of ceramic
foam catalyzed-soot filters for diesels.
130-133
Again, it is
necessary to displace existing well-established products that
generally have a lower backpressure, and, more importantly,
the filtration efficiency of ceramic foam is less than that of a
ceramic wall-flow filter.
For ceramic foam catalysts to enter the autocatalyst area,
development work on very large cartridge-like units (e.g., 14
cm 15 cm) that do not crack during thermal cycling is
necessary. They should also have a ceramic skin to enable them
to be canned in the same way that cordierite monoliths are today.
Moreover, it would be necessary to select the most appropriate
pore density and catalyst length carefully to achieve satisfactory
catalyst internal surface area while not having too high a
backpressure.
Process Catalysts. Ceramic foam catalysts can have advan-
tages over their traditional counterparts in the process chemical
industry, especially in situations that involve strongly exothermic
or endothermic reactions, or where an open pore structure helps
to control selectivity. To enter mass production, the cost of
manufacturing foam catalyst pieces must be comparable to
catalyst costs for traditional pellets, larger Raschig rings, and
multihole pellets. It seems likely that an automated robotic
process will be needed to have an efficient production process.
Ceramic foam catalysts can be quite brittle, and breakage
during manufacture and loading into a reactor could cause
potential problems; however, this could be overcome by having
a continuous outer protective ceramic skin (for example, on a
cylinder-shaped foam catalyst that has a skin either on the
outside perimeter or on the top and bottom faces).
Finally, appropriate reactors must be used to take full
advantage of the special properties of the ceramic foam, and
new reactor designs may well be required. For example, with
highly exothermic or endothermic reactions, heat and mass
transport into and out of a foam catalyst is better than its
conventional counterpart; however, in a practical situation, using
a foam catalyst may just result in different limitations restricting
the rate of the overall process, so the full benefit of the foam
catalyst is not realized. An example might be a strongly
endothermic process using thick-walled multitubular reactors.
Replacing traditional catalysts with foam catalysts shifts the
limitation to heat transfer through the reactor wall, rather than
within the tubes to the catalyst surface; therefore, only a
proportion of the ceramic foam catalyst benefit is realized.
Complications are possible when long lengths of preformed
ceramic foam are used. There could be a practical problem with
charging and discharging reactor tubes that are never made
perfectly straight, and, in use, this can become extenuated. For
instance, bowed and distorted tubes are common in reactors such
as steam reformers. This may not be as serious a problem in
exothermic processes, where temperatures are generally much
lower than those in a steam reformer; however, it would be
better to have new reactor designs without inherent difficulties.
Conclusions
Recently, there have been a considerable number of studies
on the fundamentals of fluid transport in ceramic foam catalyst
supports. This has led to a greater understanding of pressure
drop and heat transfer in ceramic foam beds and has resulted
in correlations that, although still empirical, are useful in
designing reactors to take advantage of the unique properties
of the foam. A large amount of research on applying ceramic
foam supports to important reactions has also been reported.
The most noteworthy reports fall into two categories: those that
benefit from a low-pressure drop at high flow rates, so that
consecutive reactions can be run at short contact times to
optimize yields of valuable intermediates, and those that are
highly endothermic or exothermic and are normally limited by
heat transfer. Techniques for successfully loading active com-
ponents onto ceramic struts have been demonstrated, and
simulation studies have verified the considerable advantages that
could be gained over conventional processes.
One unique property of the foam is transparency, i.e., the
ability for high-intensity light to penetrate deep into the structure
through radiation from one strut to another. This has been
exploited in two applications: the absorption of concentrated
solar radiation to drive endothermic reactions that essentially
store the solar energy, and use of photocatalytic reactions at
lower temperatures to destroy contaminants such as chlorinated
hydrocarbons in water.
Apart from small applications, such as on-board fuel proces-
sors for fuel-cell vehicles, these research achievements have not
been applied to replace traditional catalysts in large-scale
commercial operations, partly because of resistance to untried
innovations and partly because foams may not overcome the
advantages of monoliths (e.g., lowest pressure drop) for certain
applications. Clearly, additional development work is needed.
Ceramic foam catalysts must be manufactured economically,
using new production techniques, and must be robust and
durable to withstand the rigors of real world use. In addition,
special features must be devised to make the foams compatible
with process reactors. However, it is also possible to exploit
foam advantages by designing new types of reactors or use
reactors that operate on different principles.
Acknowledgment
We are grateful to S. Brown, H. Gadali, and W. Quon for
providing unpublished results.
Literature Cited
(1) Moulijn, J. A.; Makkee, Ir. M.; van Dieper, A. E. Chemical Process
Technology; Wiley: New York, 2001.
(2) Cybulski, A., Moulijn, J. A., Eds. Structured Catalysts and Reactions;
Marcel Dekker: New York, 1998.
(3) Cybulski, A., Moulijn, J. A., Eds. Structured Catalysts and Reactions,
Second Edition; CRC/Taylor and Francis: Boca Raton, FL, 2006.
4174 Ind. Eng. Chem. Res., Vol. 46, No. 12, 2007
(4) Gibson, L. J.; Ashby, M. F. Cellular Solids, Structure and Properties;
Pergamon Press: Oxford, U.K., 1988.
(5) Gibson, L. J.; Ashby, M. F. Cellular Solids, Structure and Properties,
Second Edition; University of Cambridge Press: Cambridge, U.K., 1999.
(6) Bhaduri, S. B. Science and technology of ceramic foams. AdV.
Perform. Mater. 1994, 1, 205-220.
(7) Sweeting, T. B.; Norris, D. A.; Strom, L. A.; Morris, J. R. Reticulated
ceramics for catalyst support applications. Mater Res. Soc. Symp. Proc.
1995, 368, 309-314.
(8) Carty, W. M.; Lednor, P. W. Monolithic ceramics and heterogeneous
catalysts: Honeycombs and foams. Curr. Opin. Solid State Mater. Sci. 1996,
1, 88-95.
(9) Colombo, P. Ceramic foams: Fabrication, properties and applications.
Key Eng. Mater. 2002, 206-213, 1913-1918.
(10) Twigg, M. V.; Richardson, J. T. Preparation and properties of
ceramic foam catalyst supports. In Preparation of Catalysts VI; Poncelet,
G., Martens, J., Delmon, B., Jacobs, P. A., Grange, P., Eds.; Studies in
Surface Science and Catalysis 91; Elsevier: Amsterdam, New York, 1994;
pp 345-359.
(11) Twigg, M. V.; Richardson, J. T. Theory and applications of ceramic
foam catalysts. Chem. Eng. Res. Des. 2002, 80, 183-189.
(12) Schwartzwalder, A. E.; Somers, A. V. Method of making porous
ceramic articles. U.S. Patent 3,090,094, 1963.
(13) Lange, F. F.; Miller, K. T. Open-cell, low-density ceramics
fabricated from reticulated polymer substrates. AdV. Ceram. Mater. 1987,
2, 827-831.
(14) Saggio-Woyansky, J.; Scott, C. E.; Minnear, W. P. Processing of
porous ceramics. Am. Ceram. Soc. Bull. 1992, 71, 1676-1682.
(15) Luyten, J.; Thijs, I.; Vandermeulen, W.; Mullens, S.; Wallaeys,
B.; Mortelmans, R. Strong ceramic foams from polyurethane templates.
AdV. Appl. Ceram. 2005, 104, 4-8.
(16) Twigg, M. V; Sengelow, W. M. U.S. Patent 4,863,712, 1989.
(17) Twigg, M. V.; Sengelow, W. M. U.S. Patent 4,810,685, 1989.
(18) Richardson, J. T. Principles of Catalyst DeVelopment; Plenum
Press: New York, 1989.
(19) Brown, S. A. The Kinetics of Exothermic Reactions on Ceramic
Foam Catalysts. Ph.D. Dissertation, Department of Chemical Engineering,
University of Houston, Houston, TX, 2001.
(20) Moates, F. C.; McMinn, T. E.; Richardson, J. T. Radial reactor for
trichloroethylene steam reforming. AIChE J. 1999, 45, 2411-2418.
(21) Hall, M. J.; Bracchini, M. Characterization of ceramic foams
through Fourier analysis of digital images. J. Am. Ceram. Soc. 1997, 80,
1298-1302.
(22) Richardson, J. T.; Peng, Y.; Remue, D. Properties of ceramic foam
catalyst supports: Pressure drop. Appl. Catal., A 2000, 204, 19-32.
(23) Boretto, G.; Merlo, A. M.; Amato, I. Characterization of ceramic
foams by computed tomography with high spatial resolution. AdV. Sci.
Technol. 1999, 16, 529-536.
(24) Peng, Y. Transport Properties of Ceramic Foam for Catalyst Support
Applications. Ph.D. Dissertation, Department of Chemical Engineering,
University of Houston, Houston, TX, 2001.
(25) Appoloni, C. R.; Fernandes, C. P.; Innocentini, M. D. de M.;
Macedo, A. Ceramic foam porous microstructure characterization by X-ray
microtomography. Mater. Res. 2004, 7, 557-564.
(26) Buciuman, F. C.; Kraushaar-Czarnetzki, B. Ceramic foam monoliths
as catalyst carriers. 1. Adjustment and description of the morphology. Ind.
Eng. Chem. Res. 2003, 42, 1863-1869.
(27) Giani, L.; Groppi, G.; Tronconi, E. Mass transfer characterization
of metal foams as supports for structured catalysts. Ind. Eng. Chem. Res.
2005, 44, 4993-5002.
(28) Kwon, Y. W.; Cooke, R. E.; Park, C. Representative unit cell models
for open cell metal foams with or without elastic filler. Mater. Sci. Eng.
2003, 343, 63-70.
(29) Brothers, A. H.; Dunand, D. C. Amorphous metal foams. Scr. Mater.
2005, 54, 513-520.
(30) Ergun, S.; Orning, A. A. Fluid flow through randomly packed
columns and fluidized beds. Ind. Eng. Chem. 1949, 41, 1179-1184.
(31) Philipse, A. P.; Schram, H. L. Non-Darcian airflow through ceramic
foams. J. Am. Ceram. Soc. 1991, 74, 728-732.
(32) Montillet, A.; Comiti, J.; Legrand, J. Axial dispersion in liquid flow
through packed retriculated metallic foams and fixed beds in different
structures. Chem. Eng. J. 1993, 52, 63-71.
(33) Du Plessis, P.; Montillet, A.; Commiti, J.; Legrand, J. Pressure drop
prediction for flow through high porosity metallic foams. Chem. Eng. Sci.
1994, 49, 3545-3553.
(34) Lopes, R. A.; Segadaes, A. M. Microstructure, permeability and
mechanical behavior of ceramic foams. Mater. Sci. Eng., A 1996, A209,
149-155.
(35) Hall, M. J.; Hiatt, J. P. Measurements of pore scale flows within
and exiting ceramic foams. Exp. Fluids 1996, 20, 433-440.
(36) Hackert, C. L.; Ellzey, J. L.; Ezekoye, O. A.; Hall, M. J. Transverse
dispersion at high Peclet numbers in short porous media. Exp. Fluids 1996,
21, 286-290.
(37) Seguin, D.; Montilletm, A.; Comiti, J.; Huet, F. Experimental
characterization of flow regimes in various porous media. II. Transition to
turbulent regime. Chem. Eng. Sci. 1998, 53, 3897-3909.
(38) Innocentini, M. D. M.; Salvini, V. R.; Macedo, A.; Pandolfelli, V.
C. Prediction of ceramic foams permeability using Erguns equation. Mater.
Res. 1999, 4, 283-289.
(39) Innocentini, M. D. M.; Salvini, V. R.; Pandolfelli, V. C.; Coury, J.
R. The permeability of ceramic foams. Am. Ceram. Soc. Bull. 1999, 78,
78-84.
(40) Innocentini, M. D. M.; Salvini, V. R.; Coury, J. R.; Pandolfelli, V.
C. Assessment of Forchheimers equation to predict the permeability of
ceramic foams. J. Am. Ceram. Soc. 1999, 82, 1945-1948.
(41) Smit, G. J. F.; Du Plessis, J. P. Modeling of non-Newtonian purely
viscous flow through isotropic high porosity synthetic foams. Chem. Eng.
Sci. 2001, 54, 645-654.
(42) Moreira, E. A.; Innocentini, M. D. M.; Salvini, V. R.; Pandolfelli,
V. C.; Coury, J. R. Permeability of cellular ceramics. ReV. UniV. Rural,
Ser. Cienc. Exatas Terra 2002, 21, 177-185.
(43) Fourie, J. G.; Du Plessis, J. P. Pressure drop modeling in cellular
metallic foams. Chem. Eng. Sci. 2002, 57, 2781-2789.
(44) Phanikumar, M. S.; Mahajan, R. L. Non-Darcy natural convection
in high porosity metal foams. Int. J. Heat Mass Transfer 2002, 45, 3781-
3793.
(45) Moreira, E. A.; Coury, J. R. The influence of structural parameters
on the permeability of ceramic foams. Braz. J. Chem. Eng. 2004, 21, 23-
33.
(46) Moreira, E. A.; Innocentini, M. D. M.; Coury, J. R. Permeability
of ceramic foams to compressible and incompressible flow. J. Eur. Ceram.
Soc. 2004, 24, 3209-3218.
(47) Stemmet, C. P.; Jongmans, J. N.; van der Schaaf, J.; Kuster, B. F.
M., Schouten, J. C. Hydrodynamics of gas-liquid counter-current flow in
solid foam packings. Chem. Eng. Sci. 2005, 60, 6422-6429.
(48) Macdonald, F. I.; El-Sayed, M. F. S.; Mow, K.; Dullien, F. A. L.
Flow through porous mediasThe Ergun equation revisited. Ind. Eng. Chem.
Fundam. 1979, 18, 199-208.
(49) Zumbrunnen, R.; Viskanta, R.; Incropera, F. P. Heat transfer through
porous solids with complex internal geometries. Int. J. Heat Mass Transfer
1986, 29, 275-284.
(50) Lee, K. B.; Howell, J. R. Theoretical and experimental heat and
mass transfer in highly porous media. Int. J. Heat Mass Transfer 1991, 34,
2123-2132.
(51) Younis, L. B.; Viskanta, R. Experimental determination of the
volumetric heat transfer coefficient between stream of air and ceramic foam.
Int. J. Heat Mass Transfer 1993, 36, 1425-1434.
(52) Schlegel, A.; Benz, P.; Buser, S. Wameubertragung und Drukabfall
in keramischen Schaumstrukturen bei erzwungener Stromung. Warme-
Stoffubertrag 1993, 28, 259-266.
(53) Jones, S.; Catton, I. Non-intrusive heat transfer coefficient deter-
mination in highly porous metal/ceramic foams. Proc. ASME Heat Transfer
DiV. 1998, 5, 579-584.
(54) Lu, T. J.; Stone, H. A.; Ashby, M. F. Heat transfer in open-cell
metal foams. Acta Metall. 1998, 46, 3619-3635.
(55) Boomsma, K.; Poulikakos, D. On the effective thermal conductivity
of a three-dimensionally structured fluid-saturated metal foam. Int. Heat
Mass Transfer 2001, 44, 827-836.
(56) Richardson, J. T.; Remue, D.; Hung, J. K. Properties of ceramic
foam catalyst supports: Mass and heat transfer. Appl. Catal., A 2003, 250,
319-329.
(57) Fourie, J. G.; Du Plessis, J. P. Effective and coupled thermal
conductivities of isotropic open-cellular foams. AIChE J. 2004, 50, 547-
556.
(58) Pereira, J. C. F.; Malico, I.; Hayashi, T. C.; Raposo, J. Experimental
and numerical characterization of the transverse dispersion at the exit of a
short ceramic foam inside a pipe. Int. J. Heat Mass Transfer 2004, 48,
1-14.
(59) Peng, Y.; Richardson, J. T. Properties of ceramic foam catalyst
supports: one-dimensional and two-dimensional heat transfer correlations.
Appl. Catal., A 2004, 266, 235-244.
(60) Gadalli, H.; Richardson, J. T. Private communication, 2006.
(61) Richardson, J. T.; Garrait, M.; Hung, J. K. Carbon dioxide reforming
with Rh and Pt-Re catalysts dispersed on ceramic foam supports. Appl.
Catal., A 2003, 255, 69-82.
(62) Quon, W.; Richardson, J. T. Private communication, 2006.
Ind. Eng. Chem. Res., Vol. 46, No. 12, 2007 4175
(63) Thurgood, C. P.; Amphlett, J. C.; Mann, R. F.; Peppley, B. A. A
comparison between ceramic foam catalyst and packed-beds for methanol-
steam reforming. Presented at AIChE Annual Meeting Conference Proceed-
ings, Cincinnati, OH, October 30-November 4, 2005.
(64) Pias Peleteiro, P. M.; Reizmann, A.; Bareiss, A.; Kraushaar-
Czarnetzki, B. Simulation of a reactor for the partial oxidation of o-xylene
to phthalic anhydride packed with ceramic foam monoliths. DGMK
Tagungsbericht, In Proceedings of the DGMK/SCI-Conference, Oxidation
and Functionalization: Classical and AlternatiVe Routes and Sources,
Milan, Italy, 2005.
(65) Hickman, D. A.; Schmidt, L. D. Synthesis gas formation by direct
oxidation of methane over Pt monoliths. J. Catal. 1992, 138, 267-282.
(66) Hickman, D. A.; Schmidt, L. D. Production of synthesis gas by
direct catalytic oxidation of methane. Science 1993, 259, 343-346.
(67) Hickman, D. A.; Haupfear, E. A.; Schmidt, L. D. Synthesis gas
formation by direct oxidation of methane over rhodium monoliths. Catal.
Lett. 1993, 17, 223-237.
(68) Hickman, D. A.; Schmidt, L. D. Synthesis gas formation by direct
oxidation of methane over monoliths. ACS Symp. Ser. 1993, 523, 416-
426.
(69) Hickman, D. A.; Schmidt, L. D. Steps in methane oxidation on
platinum and rhodium surfaces. High-temperature reactor simulations.
AIChE J. 1993, 39, 1164-1177.
(70) Huff, M.; Torniainen, P. M.; Hickman, D. A.; Schmidt, L. D. Partial
oxidation of CH
4, C2H6, and C3H8 on monoliths at short contact times. In
Studies in Surface Science and Catalysis 81; Curry-Hyde, H. E., Howe, R.
F., Eds.; Elsevier: Amsterdam, 1994; pp 315-320.
(71) Schmidt, L. D.; Huff, M. Partial oxidation of CH4 and C2H6 over
noble metal-coated monoliths. Catal. Today 1994, 21, 443-454.
(72) Schmidt, L. D.; Dietz, A., III. Monoliths for partial oxidation
catalysts. Mater. Res. Soc. Symp. Proc. 1995, 368, 299-307.
(73) Huff, M.; Schmidt, L. D. Ethene formation by oxidative dehydro-
genation of ethane over monoliths at very short contact times. J. Phys. Chem.
1993, 97, 11815-11822.
(74) Huff, M.; Schmidt, L. D. Partial oxidation and cracking of alkanes
over noble metal coated monoliths. In Methane and Alkane ConVersion
Chemistry; Bhasin, M. M., Slocum, D. W., Eds.; Plenum Press: New York,
1995.
(75) Dietz, A. G., III; Carlsson, A. F.; Schmidt, L. D. Partial oxidation
of C
5 and C6 alkanes over monolith catalysts at short contact times. J. Catal.
1998, 176, 459-473.
(76) Leclerc, C. A.; Schmidt, L. D. Rapid light off of octane reforming.
Am. Chem. Soc., DiV. Fuel Chem. 2002 47, 718-719.
(77) Leclerc, C. A.; Schmidt, L. D. Rapid light off of octane reforming.
Presented at 224th ACS National Meeting, Boston, MA, August 18-22,
2002.
(78) Krummenacher, J. J.; West, K. N.; Schmidt, L. D. Catalytic partial
oxidation of higher hydrocarbons at millisecond contact times: Decane,
hexadecane, and diesel fuel. J. Catal. 2003, 215, 332-343.
(79) Wanat, E. C.; Suman, B.; Schmidt, L. D. Partial oxidation of
alcohols to produce hydrogen and chemicals in millisecond-contact time
reactors. J. Catal. 2005, 235, 18-27.
(80) Williams, K. A.; Schmidt, L. D. Catalytic autoignition of higher
alkane partial oxidation on Rh-coated foams. Appl. Catal., A 2006, 299,
30-45.
(81) Dauenhauer, P.; Schmidt, L. D. Autothermal reforming of renewable
feedstocks. Presented at 38th Central Regional Meeting of the American
Chemical Society, Frankenmuth, MI, May 16-20, 2006.
(82) Huff, M.; Schmidt, L. D. Production of olefins by oxidative
dehydrogenation of propane and butane over monoliths at short contact
times. J. Catal. 1994, 149, 127-141.
(83) Huff, M.; Schmidt, L. D. Oxidative dehydrogenation of isobutane
over monoliths at short contact times. J. Catal. 1995, 155, 82-94.
(84) Yokoyama, C.; Bharadwaj, S. S.; Schmidt, L. D. Platinum-tin-
platinum-copper catalysts for autothermal oxidative dehydrogenation of
ethane to ethylene. Catal. Lett. 1996, 38, 181-188.
(85) Witt, P. M.; Schmidt, L. D. Effect of flow rate on partial oxidation
of methane and ethane. J. Catal. 1996, 163, 465-475.
(86) Flick, D. W.; Huff, M. C. Acetylene formation during the catalytic
oxidative dehydrogenation of ethane over Pt-coated monoliths at short
contact times. Catal. Lett. 1997, 47, 91-97.
(87) Flick, D. W.; Huff, M. C. Oxidative dehydrogenation of ethane
over a Pt-coated monolith versus Pt-loaded pellets: Surface area and thermal
effects. J. Catal. 1998, 178, 315-327.
(88) Schmidt, L. D.; Hohn, K. L.; Davis, M. B. Catalytic partial oxidation
of methane at extremely short contact times: Production of acetylene. In
Studies in Surface Science and Catalysis 119; Parmaliana, A., Sanfilippo,
D., Frusteri, F., Vaccari, A., Arena, F., Eds.; Elsevier: Amsterdam, 1998;
pp 397-402.
(89) Schmidt, L. D. Fast light off and transients in millisecond chemical
reactors. Presented at 223rd ACS National Meeting, Orlando, FL, April
7-11, 2002.
(90) Schmidt, L. D. Keynote Address: Hydrogen and olefins from higher
alkanes in millisecond reactors. Presented at 226th ACS National Meeting,
New York, September 7-11, 2003.
(91) Huff, M.; Schmidt, L. D. Partial oxidation and cracking of alkanes
over noble metal coated monoliths. In Proceedings of American Chemical
Society Symposium on Methane and Alkane ConVersion Chemistry, San
Diego, CA, March 13-17, 1995; American Chemical Society: Washington,
DC, pp 227-239.
(92) Williams, K. A.; Leclerc, C. A.; Schmidt, L. D. Rapid lightoff of
syngas production from methane: A transient product analysis. AIChE J.
2005, 51, 247-260.
(93) Flick, W.; Huff, M. Oxidative dehydrogenation of ethane over
supported chromium oxide and Pt modified chromium oxide. Appl. Catal.
1999, 187, 13-24.
(94) Liebmann, L. S.; Schmidt, L. D. Oxidative dehydrogenation of
isobutane at short contact times. Appl. Catal., A 1999, 179, 93-106.
(95) Flick, D.; Huff, M. Oxidative dehydrogenation over promoted
chromia catalysts at short contact times. In 12th International Congress on
Catalysis; Corma, A., Melo, F. V., Mendioroz, S., Fierro, J. L. G., Eds.;
Studies in Surface Science and Catalysis 130; Elsevier: Amsterdam, 2000;
pp 779-784.
(96) Pfefferle, W. C.; Castaldi, M.; Etemad, S.; Karim, H.; Lyubovsky,
M.; Roychoudhury, S.; Smith, L. Catalysts for improved process efficiency.
Presented at the 223rd ACS National Meeting, Orlando, FL, April 7-11,
2002.
(97) Schmidt, L. D. Chemicals and energy from alkanes in millisecond
reactors. Presented at the 223rd ACS National Meeting, Orlando, FL, April
7-11, 2002.
(98) Krummenacher, J. J.; Schmidt, L. D. High yields of olefins and
hydrogen from decane in short contact time reactors: Rhodium versus
platinum. J. Catal. 2004, 222, 429-438.
(99) Subramanian, R.; Schmidt, L. D. Renewable olefins from biodiesel
by autothermal reforming. Angew. Chem., Int. Ed. 2005, 44, 302-305.
(100) Donsi, F.; Williams, K. A.; Schmidt, L. D. A. multistep surface
mechanism for ethane oxidative dehydrogenation on Pt- and Pt/Sn-coated
monoliths. Ind. Eng. Chem. Res. 2005, 44, 3453-3470.
(101) Schmidt, L. D.; Subramanian, R.; Salge, J. R.; Deluga, G. A.
Hydrogen and chemicals from renewable fuels by autothermal reforming.
Indian Chem. Eng. 2005, 47, 100-105.
(102) Inui, T.; Kuroda, T.; Otowa, T. Catalytic combustion of methane
over composite catalysts supported on a ceramic foam. Nenryo Kyokaishi
1985, 64, 270-277.
(103) Inui, T.; Adachi, T.; Nanya, M.; Miyamoto, A. Catalytic combus-
tion of C
1-C14 paraffins on Pt-CeO2 composite catalyst supported on
ceramic foam and ceramic fiber. Chem. Exp. 1986, 4, 255-258.
(104) Jiratova, K.; Moravkoca, L.; Malecha, J.; Koutsky, B. Ceramic
foam in catalytic combustion of methane. Collect. Czech. Chem. Commun.
1995, 60, 473-481.
(105) Jiratova, K.; Moravkova, L.; Malecha, J.; Koutsky, B. l. Ceramic
foam-supported perovskites as catalysts for combustion of methane. Collect.
Czech. Chem. Commun. 1997, 62, 875-883.
(106) Aleksandrov, Y. A.; Tsganova, E. I.; Ivanovskaya, K. E.;
Vorozheikin, I. A. Kinetic regularities of the heterogeneous catalytic
oxidation of carbon monoxide on ceramic foam-supported catalysts. Russ.
J. Gen. Chem. 2002, 72, 13-16.
(107) Tsganova, E. I.; Didenkulova, I. I.; Shekunova, V. M.; Aleksan-
drov, Y. A. Oxidation of CO on Zr-, Fe-, Cr-, Ni-, and Cu-containing
catalysts prepared by pyrolysis of metal -diketonates on a synthetic ceramic
foam. Russ. J. Gen. Chem. 2005, 75, 1354-1358.
(108) Tsganova, E. I.; Didenkulova, I. I.; Shekunova, V. M.; Aleksan-
drov, Y. A. Oxidation of CO on Mn-containing catalysts prepared by
pyrolysis of metal -diketonates on synthetic ceramic foam. Russ. J. Gen.
Chem. 2005, 75, 1599-1602.
(109) Richardson, J. T.; Shafiei, M.; Cantu, T.; Machiraju, D.; Telleen,
S. Low NO
x emission combustion catalyst using ceramic foam supports.
In Natural Gas Conversion VII; Bao, X., Xu, Y., Eds.; Studies in Surface
Science and Catalysis 147; Elsevier: Amsterdam, 2004; pp 457-462.
(110) Thompson, L. T. Advances in fuel processing for PEM fuel cells.
Prepr. Symp.sAm. Chem. Soc., DiV. Fuel Chem. 2003, 48, 320-321.
(111) Worner, A.; Friedrich, C.; Tamme, R. Development of a novel
Ru-based catalyst system for the selective oxidation of CO in hydrogen
rich gas mixtures. Appl. Catal., A 2003, 245, 1-14.
(112) Liguras, D. K.; Goundani, K.; Verykios, X. E. Production of
hydrogen for fuel cells by catalytic partial oxidation of ethanol over
structured Ni catalysts. J. Power Sources 2004, 130, 30-37.
4176 Ind. Eng. Chem. Res., Vol. 46, No. 12, 2007
(113) Liguras, D. K.; Goundani, K.; Verykios, X. E. Production of
hydrogen for fuel cells by catalytic partial oxidation of ethanol over
structured Ru catalysts. Int. J. Hydrogen Energy 2004, 29, 419-427.
(114) Woerner, A.; Tamme, R.; Mueller-Steinhagen, H. Ceramic foam
based catalysts: preparation and testing for fuel processing. Presented at
the 7th World Congress of Chemical Engineering, Glasgow, U.K., July 10-
14, 2005.
(115) Sirijaruphan, A.; Goodwin, J. G., Jr.; Rice, R. W.; Wei, D.;
Butcher, K. R.; Roberts, G. W.; Spivey, J. J. Metal foam supported Pt
catalysts for the selective oxidation of CO in hydrogen. Appl. Catal., A
2005, 281, 1-9.
(116) Buck, R.; Muir, J. F.; Hogan, R. H.; Skocypec, R. D. Carbon
dioxide reforming of methane in a solar volumetric receiver/reactor: The
CAESAR project. Solar Energy Mater. 1991, 24, 449-463.
(117) Takano, A.; Tagawa, T.; Goto, S. Carbon dioxide reforming of
methane on supported nickel catalysts. J. Chem. Eng. Jpn. 1994, 27, 727-
731.
(118) Takano, A.; Tagawa, T.; Goto, S. Carbon dioxide reforming over
nickel catalyst supported on ceramic foam. J. Ceram. Soc. Jpn. 1996, 104,
444-446.
(119) Worner, A.; Tamme, R. CO
2 reforming of methane in a solar
driven volumetric receiver-reactor. Catal. Today 1998, 46, 165-174.
(120) Kodama, T.; Kiyama, A.; Moriyama, T.; Yokoyama, T.; Shimizu,
K. I.; Andou, H.; Satou, N. Ni-Mg-O catalyst driven by direct light
irradiation for catalytically-activated foam absorber in a solar reforming
receiver-reactor. Energy Fuels 2003, 17, 914-921.
(121) Pestryakov, A. N.; Tetranovskii, V. P.; Pfander, N.; Knop-Gericke,
A. Supported foam-copper catalysts for methanol selective oxidation. Catal.
Commun. 2004, 5, 777-781.
(122) Patcas, F. C. The methanol-to-olefins conversion over zeolite-
coated ceramic foams. J. Catal. 2005, 231, 194-200.
(123) Labhsetwar, N. K.; Biniwale, R. B.; Kumar, R.; Bawase, M. A.;
Rayalu, S. S.; Mitsuhashi, T.; Haneda, H. Application of catalytic materials
for diesel exhaust emission control. Curr. Sci. 2004, 87, 1700-1704.
(124) Kim, H.; Lee, S.; Han, Y.; Park, J. Preparation of dip-coated TiO
2
photocatalyst on ceramic foam pellets. J. Mater. Sci. 2005, 40, 5295-5298.
(125) Stemmet, C. P.; van der Schaaf, J.; Kuster, B. F. M.; Schouten, J.
C. Structured foam packings for multiphase reactors: Mass transfer
characteristics for counter-current gas-liquid flow. Presented at the 7th
World Congress of Chemical Engineering, Glasgow, U.K., July 10-14,
2005.
(126) Twigg, M. V. Progress and future challenges in controlling
automotive exhaust gas emission. Appl. Catal., B 2006, 70 (1-4), 2-15.
(127) Twigg, M. V. Controlling automotive exhaust emissions: successes
and underlining science. Philos. Trans. R. Soc. London, Ser. A 2005, 363,
1013-1033.
(128) Twigg, M. V. Twenty-five years of auto catalysts. Platinum Met.
ReV. 1999, 44, 168-171.
(129) Acres, G. J. K.; Harrison, B. The development of catalysts for
emission control from motor vehicles: early research at Johnson Matthey.
Top. Catal. 2004, 28, 3-11.
(130) Benbow, J.; Bridgewater, J. Paste Flow and Extrusion; Oxford
University Press: Oxford, U.K., 1993.
(131) Benbow, J. J.; Lord, L. W.; Heath, D. J. Support and Catalyst.
GB Patent 138590M, 1992.
(132) Pryor, M. J.; Gray, T. J. Molten Metal Filter. U.S. Patent 3,893,-
917, 1975.
(133) Pryor, M. J.; Gray, T. J. Ceramic Foam Filter. U.S. Patent 3,-
947,363, 1976.
(134) Pryor, M. J.; Pryor, M. J.; Gray, T. J. Method of Preparing Molten
Metal Filter. U.S. Patent 4,056,586, 1977.
(135) van Setten, B. A. A. L.; Bremmer, J. S.; Jelles, S. J.; Makkee,
M.; Moulijn, J. A. Ceramic foam as a potential molten salt oxidation catalyst
support in the removal of soot from diesel exhaust gas. Catal. Today 1999,
55, 613-621.
ReceiVed for reView August 24, 2006
ReVised manuscript receiVed January 8, 2007
Accepted January 9, 2007
IE061122O
Ind. Eng. Chem. Res., Vol. 46, No. 12, 2007 4177

Vous aimerez peut-être aussi