You are on page 1of 38

Catalysis Reviews

Science and Engineering

ISSN: 0161-4940 (Print) 1520-5703 (Online) Journal homepage: http://www.tandfonline.com/loi/lctr20

Preparation of monolithic catalysts

T. Alexander Nijhuis , Annemarie E. W. Beers , Theo Vergunst , Ingrid Hoek ,


Freek Kapteijn & Jacob A. Moulijn

To cite this article: T. Alexander Nijhuis , Annemarie E. W. Beers , Theo Vergunst , Ingrid Hoek ,
Freek Kapteijn & Jacob A. Moulijn (2001) Preparation of monolithic catalysts, Catalysis Reviews,
43:4, 345-380, DOI: 10.1081/CR-120001807

To link to this article: http://dx.doi.org/10.1081/CR-120001807

Published online: 03 Feb 2007.

Submit your article to this journal

Article views: 1852

View related articles

Citing articles: 272 View citing articles

Full Terms & Conditions of access and use can be found at


http://www.tandfonline.com/action/journalInformation?journalCode=lctr20

Download by: [Texas A & M University--Kingsville] Date: 30 April 2017, At: 15:02
CATALYSIS REVIEWS, 43(4), 345380 (2001)

PREPARATION OF MONOLITHIC CATALYSTS

T. Alexander Nijhuis,* Annemarie E. W. Beers, Theo Vergunst,


Ingrid Hoek, Freek Kapteijn, and Jacob A. Moulijn

Delft University of Technology, DelftChemTech, Section Industrial


Catalysis, Julianalaan 136, 2628 BL Delft, The Netherlands

I MONOLITHIC REACTORS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346


II REQUIREMENTS FOR USING A MONOLITH AS A CATALYST
SUPPORT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 348
III PREPARATION OF MONOLITHIC CATALYSTS . . . . . . . . . . . . . . . . . . . 350
A. Coating a Monolith with a Catalyst Support Material . . . . . . . . . . . . . . 351
B. Deposition of the Active Phase on a Monolithic Support . . . . . . . . . . . 362
C. Coating Ready-Made Catalysts on a Monolith . . . . . . . . . . . . . . . . . . . . 370
D. Other Types of Monolithic Catalyst . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 374
IV CONCLUSIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377
ACKNOWLEDGMENTS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377
REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 377

ABSTRACT

Monolithic catalysts can be attractive replacements for conventional catalysts


in randomly packed beds or slurry reactors. The conventional procedures for
preparing catalysts, however, cannot simply be applied to monolithic
catalysts. Different procedures are discussed on how to put a coat layer of a
catalyst support material like alumina, silica, or carbon on a monolith body by
either filling the pores in that support or by putting a layer on that support.
Different methods to apply an active phase to the support are discussed as

*Address correspondence to T. Alexander Nijhuis. E-mail: t.a.nijhuis@tnw.tudelft.nl

345

Copyright q 2001 by Marcel Dekker, Inc. www.dekker.com


ORDER REPRINTS

346 NIJHUIS ET AL.

well. Finally, methods to convert ready-made catalysts into monolithic


catalysts are presented.

Key Words: Monolith; Structured catalyst; Preparation; Catalyst

I. MONOLITHIC REACTORS

Monolithic catalyst supports can be an attractive replacement for


conventional carriers in heterogeneous catalysts. Monolithic structures, either
metal or ceramic, consist of single blocks of small (0.5 4 mm) parallel channels
with a catalytic wall. Figure 1 shows an example of a ceramic monolithic structure.
At the moment, the largest application of monoliths is in the automotive industry
for the cleanup of exhaust gases. Other widespread applications are the selective
catalytic reduction of off-gases of power stations and the ozone destruction in
airplanes. A recent article discussing the application of monoliths in catalytic gas-
phase oxidations is by Geus and van Giezen (1).
The use of monolithic supports for other applications in the chemical industry
is at this moment very limited, even though, especially in multiphase reactions,
monolithic reactors have clear advantages over the conventional slurry and fixed-
bed reactors. A comparison between these types of multiphase reactor is given in
Table 1. In this table, it can be seen that a monolith reactor has many operational
advantages (energy input, efficiency, safety, and catalyst separation). The
preparation of monolithic catalysts is discussed in this article to help make this
more facile and thereby minimize the disadvantage of limited experience in this
respect. The lower catalyst loading is compensated by the higher catalyst efficiency
due to the good mass-transfer characteristics (2). The (more difficult) catalyst
replacement makes it important that durable monolithic catalysts be developed.
At present, there is only one large-scale industrial application of monolithic
catalysts in a multiphase process: the production of hydrogen peroxide using the
anthraquinone process (3). In this process, anthraquinones are reduced in one
reactor and oxidized in another, at which time hydrogen peroxide is produced. The
quinone reduction reactor contains a monolithic palladium-based catalyst. The
main reason for choosing a monolithic catalyst in this case is that the transport of
fine catalyst particles with the liquid to the oxidation reactor can cause severe
problems. An incomplete removal of particles prior to the transport of the reaction
mixture to the oxidation reactor would cause the decomposition of hydrogen
peroxide. Applying a structured catalyst reduces this risk significantly. An
extensive discussion on the advantages of monolithic catalyst systems in
multiphase applications is given by Kapteijn et al. (4).
When monolithic supports are compared to other types of structured catalyst
supports, such as the well-known Sulzer packings, it turns out that both support
types have similar advantages over the conventional reactors. An advantage of the
Sulzer-type packings over monolithic supports is that the industrial experience
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 347

Figure 1. Photographs of 200, 400, and 600 cpsi (cells per square inch) cordierite monoliths.

with the Sulzer packings is much greater, even though most of this experience is
obtained in nonreactive applications (e.g., as packings in distillation columns).
Monoliths, however, also have significant advantages over the other types of
structured packing. First, the (most commonly used) porous ceramic material is
easier to use as a catalyst support than the metal of the conventional structured
packings (the bonding of the catalyst to the support material is more facile). When

Table 1. Comparison Between Monolithic and Slurry and Packed-Bed Reactors

Monolith Reactor Slurry Reactor Trickle-Bed Reactor

Energy Input Low Medium (stirring) High (pressure drop)


Catalyst efficiency High, thin active layer High, small particles Low, large particles
required for
pressure drop
Safety High, self-draining Medium, easy cooling, Low, difficult
reactor difficult to separate cooling, catalyst
catalyst from liquid bed retains liquid
Catalyst separation Easy Costly filtering necessary Easy
Preparation Medium, new technology, Easy Easy
methods have now been
developed
Catalyst loading Medium, very open Lowmedium High, dense bed
structure, for
washcoated systems:
low
Catalyst replacement Difficult, shutdown Easy, continuous during Mediumdifficult,
required, monoliths operation shutdown required
have to be carefully
stacked
Experience Gas-phase: extensive; Extensive Extensive
liquid and multiphase:
very limited
ORDER REPRINTS

348 NIJHUIS ET AL.

coating metal supports with a catalyst or catalyst support material, an intermediate


layer of a ceramic material is often used for a better binding (5). Second, the cost
of monolithic supports is relatively low, mainly due to the large-scale production
for the automotive industry. The cost for a basic monolithic structure can be as low
as US$ 3 per liter, mainly due to the relatively simple production method (i.e. via
an extrusion process).
The main focus in this article will be on how to convert a bare (ceramic)
monolithic body into a proper catalyst. The availability of monolith materials will
be discussed and how a coat layer of the desired catalyst support material can be
applied on the monolith if the support itself is not available in a monolithic shape.
The deposition techniques for the active phase on the support will be discussed,
and, finally, the methods of putting a ready-made powder catalyst on a monolithic
structure will be presented. As not much open literature is available, this review
relates heavily to research carried out in our laboratory.

II. REQUIREMENTS FOR USING A MONOLITH AS A


CATALYST SUPPORT

In the application of a monolithic catalyst, one should first determine what


the requirements for the support are. The most common material for monolithic
structures is cordierite (a ceramic material consisting of magnesia, silica, and
alumina in the ratio of 2:5:2), because this material is very well suited for the
requirements of the automotive industry. The main reasons for this are that it has a
high mechanical strength, can stand high temperatures and temperature shocks,
and has a low thermal expansion coefficient (6). For stationary operation in
chemical plants, these demands are much less important. On the other hand, the
lifetime of the catalyst will be much more important and a typical operating
lifetime of 2000 h such as for a three-way catalyst would be far too short for most
industrial applications. Especially, the fact that a monolithic catalyst will be
somewhat more expensive than a fixed-bed catalyst and changing the catalyst will
be at least as laborious will make the catalyst lifetime of great importance. The
very open structure of a monolith, on the other hand, makes the system less
sensitive to fouling, as small particles easily travel through the channels. Only
large particles, able to block a complete channel, can cause problems, but they are
easily caught by an upstream gauze filter.
Other materials whose monolith structures are commercially available are
metals, mullite (mixed oxide of silica and alumina, ratio 2:3) and silicon carbide.
Disadvantages of all these materials are that, similar to cordierite, they have a low
Brunauer, Emmett, and Teller (BET) specific surface area (e.g., for cordierite,
typically 0.7 m2/g), they are rarely used as support materials for conventional
catalysts, and the metal support interaction is usually very low. Monolithic
elements out of carbon, silica, and g-alumina are available as research samples and
can be produced once a significant demand exists. For these materials, surface
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 349

Table 2. Characteristics and Geometrical Properties of Available Materials for Ceramic Monoliths

Cordierite
Material (2MgO:5SiO2:2Al2O3) Silica g-Alumina

Characteristics
Cell density (cpsi) 25 1600 400 400
Pore volume (Hg porosimetry, mL/g) 0.19 0.18 0.42
Pore volume (N2 BET, mL/g) 0.08 0.47
Surface area (N2 BET, m2/g) #4 90 120 190

Cell Density (cpsi)

200 400 600

Geometrical Properties
Porosity (structure) 0.69 0.74 0.80
Channel diameter (mm) 1.50 1.09 0.93
Wall thickness (mm) 305 178 109
Geometrical area (m2/m3) 1850 2710 3450

areas of 200 m2/g are easily available; the mechanical strength, however, is
significantly lower than that of cordierite. The most important characteristics of
ceramic monoliths supplied by Corning to our laboratory are listed in Table 2.
As a compromise to use the best of both materials, it is possible to deposit a
coat layer of, for example, alumina on the cordierite support. In this manner, a
small amount of the catalyst support material covers the cordierite carrier material.
Only if a monolithic catalyst is to be used for a very slow reaction for which a large
amount of catalyst is needed, this is not a practical solution. In these cases, it can
be decided to use a larger wall thickness, increasing the amount of catalyst in the
reactor further and increasing the mechanical strength. This is attractive because
the reactor porosity of the standard monolith (0.75, this does not include the
porosity in the walls) is much higher than that of a conventionally used fixed-bed
reactor for slow reactions. A disadvantage, however, is that diffusion limitations
can occur in a thicker wall.
Ceramic monolithic elements are available with cell densities of 25
1600 cpsi (cells per square inch, equal to a cell size of 5 0.6 mm). The advantage
of using a higher cell density is that the geometrical surface area increases and,
thus, the catalyst can be used more effectively. Disadvantages of higher cell
densities are a somewhat more difficult manufacturing process, more difficult
washcoating, and a higher pressure drop over the reactor. However, the pressure
drop remains very low for high-cell-density monoliths (typically a factor of 10
lower) compared to a packed-bed reactor, because of the straight monolith
channels. Because the monolith porosity changes slightly with cell density (at
increasing cell density the number of walls increases but usually the wall thickness
ORDER REPRINTS

350 NIJHUIS ET AL.

Figure 2. Paths leading to a monolithic catalyst.

decreases), the reaction rate and possible consequential side reactions will have an
influence on the choice of the optimal cell density.

III. PREPARATION OF MONOLITHIC CATALYSTS

A schematic representation of the different paths leading to a monolithic


catalyst is given in Figure 2. All of the preparational steps in this chart will be
discussed in this article, apart from the catalyst activation procedures, which are
not significantly different for a monolithic catalyst compared to conventional
catalysts. Which route should be taken toward the preparation of the monolithic
catalyst system will strongly depend on the application and the catalyst.
If the catalyst has sufficient mechanical strength, the catalyst material can be
extruded together with some binder materials in a monolithic shape. For
applications in which the catalyst efficiency is less important and a larger amount
of catalyst is preferred, this type of preparation is possible. A disadvantage is that
this will only be economical for larger-scale applications.
Ready-made catalysts or other catalysts that do not need a specific support
material, such as zeolites, can be coated or synthesized directly on a monolith
body. Coating ready-made catalysts on a monolith body has the advantage that
well-developed conventional catalyst preparation procedures can be utilized. For
catalysts which do require a specific support material, two possibilities are
available: If the monolith body is available in the required support material, the
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 351

active phase can be directly deposited on the monolith, and if the monolith is not
available in the right material, this support material should be coated on the
monolith body first. The advantage of using the coating technique is that the
catalyst is used more efficiently, because the diffusional distance toward the active
species will be small. This is valid for both the coated ready-made catalysts and
the catalyst of which the support is coated and the active phase is deposited later.
Different techniques will be discussed for putting a catalyst support coat
layer on a monolithic body: colloidal coating, sol gel coating, slurry coating, and
polymer coating. For the deposition of the active phase on a monolithic carrier,
conventional procedures can be utilized, although some adaptations have to be
made in the deposition and drying steps to prevent an active-phase maldistribution
over the monolith body. These modified procedures are also discussed.

A. Coating a Monolith with a Catalyst Support Material

A bare monolithic structure (Fig. 3) can be coated with a catalyst support


layer to increase the BET surface area and to have a support material that has a
better interaction with the active catalyst material. This procedure is most
commonly called washcoating. In this subsection, the most convenient methods
will be discussed for coating a monolith with alumina, silica, and carbon. A
discussion on the basic principles of coating a monolith body is available in Ref. 7.
The macroporous structure of ceramic monoliths facilitates the anchoring of
the washcoat layer. The manner in which washcoating is carried out can be divided
into two methods: the macroporous support can be (partly) filled with the high-
surface-area washcoat material, or a washcoat can be deposited as a layer in the
pores on the ceramic support. This is shown schematically in Figure 4. Pore filling
results in the strongest interaction between monolith and washcoat, as most of the
coat layer is actually fixated inside the pores of the support instead of only being
attached to the external surface of the monolith channels. This type of coating is
carried out by using a solution (or sol) of the material to be deposited or by using a
solution containing very small colloidal particles. The disadvantage of coating by
means of pore filling is that the amount of coating that can be deposited is limited
(the macropore volume is limited, mercury porosimetry shows a porosity for
cordierite monoliths of 0.3), because at one stage, pores will be completely filled
and the washcoat will become inaccessible.
Coating a layer on the monolith wall has the advantages that higher loadings
are possible and that diffusion into the thicker walls does not influence the
reaction. This type of coating is carried out by coating with a suspension of
particles of similar size as the macropores in the cordierite (typically 5 mm). The
principle by which the slurry-coating procedures operate is the following. The
monolith is placed in the liquid containing suspended particles. The pores in the
wall take up liquid, depositing the particles on the monolith walls, because the
particles cannot enter the pores, leaving a layer of deposited particles. This
ORDER REPRINTS

352 NIJHUIS ET AL.

Figure 3. SEM micrographs of bare cordierite (400 cpsi monolith). (a) Cross-sectional view of
channel wall (center: cordierite, left side: steep edge of wall); (b) top view of channel. The layered
cordierite structure is clearly visible.

principle is sometimes described as slip-casting or filter-cake formation. The


disadvantage of the slurry-coating procedure is that the bonding of the coat layer to
the monolith support is more difficult because the coating is not confined to the
pores of the monolith. The advantage of the shorter diffusion distance results in a
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 353

Figure 4. Schematic representation of difference between slurry-coating and pore-filling-coating


methods.

more efficient catalyst for fast reactions and the possibilities for a higher
selectivity in case of a reaction in which an unwanted consecutive reaction over
the catalyst is also possible. For example, in the selective hydrogenation of
benzaldehyde to benzyl alcohol, it has been demonstrated that monolithic catalysts
are much more selective than extrudates (2). The general methodology of
washcoating is discussed in the next paragraphs.
A washcoating solution or slurry is prepared in which a dried monolith is
immersed for a short period (dipped). The monolith is removed from the liquid and
most liquid is shaken out, the remainder being gently blown out by pressurized air.
The easiest way to do this is using an air-knife, a thin slit blowing pressurized
air, because, in this manner, a complete row of channels is cleared simultaneously.
The monolith is then dried in the horizontal position, being rotated continuously
around its axis, to prevent gravity from causing an uneven washcoat distribution.
Finally, the coating is fixated to the monolith by a high-temperature calcination
step. The washcoat loading obtained is typically 5 10 wt% for most methods. If
for a slow reaction a higher loading is required, the coating procedure should be
repeated. This can be done after calcination or the monolith can be dipped again
after drying. The disadvantage of repeated coating without intermediate
calcination is, however, that part of the coat layer deposited in a first dip
dissolves in the second dip and, therefore, the coated amount does not increase as
much as when the monolith has been calcined first. The advantage of skipping the
intermediate calcination is that the preparation time is shorter.
The length of monolithic blocks that can be washcoated is mainly controlled
by the viscosity of the washcoating solution. If the viscosity of this solution is kept
below 30 mPa:s, washcoating monolith bodies up to 25 cm long poses no
problems. Higher viscosities or longer monolith lengths require higher air pressure
to blow out the excess liquid. This will, however, cause high air velocities through
ORDER REPRINTS

354 NIJHUIS ET AL.

the channels once the liquid has left the channel. As a result, the amount of
washcoat will be lower, because this high air velocity will also blow out part of the
remaining wet coating film. Furthermore, it is also possible that liquid is blown out
of the porous walls of the monoliths, as the air will use the porous walls as a way to
relieve the pressure.
When coating monolithic bodies, the adherence of the coating is very
important. Especially, wrongly prepared slurry coatings can be lost easily. For
example, when a hydrophobic perovskite catalyst was coated on a monolithic
structure using the method presented in Section 3.3, the coating adherence was
extremely poor. Mild adherence testing by blowing a gas stream through the
monolith, or by tapping the monolith on a surface, resulted in significant weight
losses, making this type of coated catalyst unusable. For properly adhering
systems, ultrasonic testing in water was used as a severe method of adhesion
testing. Coatings prepared in our group and discussed in this article that did not
lose mass during ultrasonic treatment did not lose mass during catalytic testing
either. Heat-shock testing is sometimes applied as a more extreme manner of
adhesion testing for monolithic catalysts used in automotive applications (8), but
this type of testing is less appropriate for monolithic systems to be applied in
reactors operated at stationary conditions.

1. Coating Using Colloidal Solutions

The easiest way to put a washcoat layer on a monolithic substrate is by using


a colloidal solution of the washcoat material. For both silica and alumina, these are
readily available, but other solutions are also available on demand. This coating
method has been applied for different systems by Beauseigneur et al. (9). The
colloidal coating works via the pore-filling method, having the advantage that the
open frontal area of a monolith (determining the pressure drop) is hardly reduced
by the coating.
A typical procedure for coating a monolith in this manner is to dry the
monolith (a few hours at 383 K) and then to submerge it in the colloidal coating
solution for a few seconds after cooling the monolith. The excess liquid is shaken
out and the liquid still remaining in the channels is gently blown out using

Table 3. Results of Coating a 600-cpsi Cordierite Monolith with Colloidal Solutions

Alumina Coating Silica Coating

Solution Alfa colloidal alumina Ludox AS-40


Solids (wt%) 20 40
Density (kg/m3) 1190 1300
Particle size (nm) 50 20
Washcoat (wt%) 6.9 14.7
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 355

Figure 5. SEM micrographs of Ludox AS-40 coated cordierite monolith (600 cpsi): (a) cross-
sectional view of channel wall (right side: steep edge of wall); (b) top view of coating on walls.

pressurized air. The monoliths are dried horizontally at room temperature while
continuously being rotated around their axes. Finally, the monoliths are calcined in
air (for alumina or silica typically at 723 K).
Some typical coating results are given in Table 3. Investigation of this type of
coating using SEM (scanning electron microscopy) micrographs (Fig. 5) does not
ORDER REPRINTS

356 NIJHUIS ET AL.

show a visible layer on the cordierite. However, it can be seen that the texture of the
support is different compared to that of bare cordierite (Fig. 3a). The typical layered
structure of bare cordierite (Fig. 3b) is no longer visible, neither in the channels nor in
the pores inside the channel walls, indicating a complete coverage of the monolithic
structure. If the thickness of the coating is calculated from the specific surface area of
the cordierite (0.7 m2/g) and a washcoat loading of 10 wt% with a density of
1600 kg/m3, a theoretical coating thickness of 90 nm is obtained. Considering that the
coating procedure has been carried out with very small particles, such a thickness is
very well possible and, thus, explains why no layer is visible in the SEM micrographs.
The specific surface area of the coating will be somewhat lower than that of the
starting material because of the sintering during the drying and calcination step
necessary for the fixation of the coating to the cordierite.

2. Sol Gel Coating

In sol gel coating, a sol is produced containing the material to be used as


catalyst support. In this procedure, the support is actually dissolved in the liquid
and not present as suspended particles. The advantage is that the coating can easily
reach all pores in the cordierite support.
A typical procedure for alumina coating using the sol method consists of the
following steps (10). A sol is prepared from pseudo-boehmite (AlOOH, Pural SB1
from Condea), urea, and 0.3 M nitric acid in a weight ratio of 2:1:5. These are
vigorously mixed using a high-shear mixer. The acid makes the alumina form
positively charged agglomerates, which repel each other, and prevents the
formation of large three-dimensional alumina networks. The addition of the polar
urea helps keep a low-viscosity sol by preventing the gelating of the AlOOH to
proceed too far. The given ratio for pseudo-boehmite, acid and urea results in the
formation of a stable sol with a typical viscosity of 20 mPa:s. A dried monolith is
dipped in this sol. The dipping time does not influence the coating process. The
monolith is then emptied by shaking and by pressurized air and then dried
horizontally while continuously being rotated around its axis. Finally, the
monolith is calcined at 723 K to produce a g-alumina coating. In this calcination
step, the carbon dioxide and nitric oxide produced from the oxidation of the urea
will help in the formation of micropores in the alumina coat layer.
Similar to the coating using colloidal solutions, no separate coat layer is
observable by SEM of a coated monolith using this sol gel method. On the other
hand, the typical cordierite structure is no longer visible, indicating a uniform
coating. Such a coating procedure results in a washcoat loading of typically
10 wt%, having a BET surface area of approximately 250 m2/gwashcoat.
A sol gel coating procedure for silica on metallic monoliths has been
developed by Zwinkels et al. (11) using colloidal silica sols together with
potassium silicate. Different procedures resulted in washcoats 20 50 mm thick,
with a surface area of 60 140 m2/g. Coating with the colloidal silica only resulted
in a thickness of only a few micrometers.
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 357

3. Slurry Coating

In slurry coating of monolithic substrates, the coat layer is primarily


deposited on the channel walls of the support. The main advantages of this method
are a shorter diffusion distance to the active catalyst species for the reactants
flowing through the channels and that the maximum loading of the coating is not
limited by the monolith macropore volume. To achieve this, slurry coating is
usually carried out with a slurry of particles of a comparable size as the larger
macropores of the support (typically a few micrometers). The slurry coating of
different materials is discussed by Addiego et al. (12).
The basic principle of slurry-coating procedures is demonstrated in Figure 6.
Relatively large particles to be coated are present in the slurry together with binder
particles, typically two orders of magnitude smaller in size. This slurry is put on
the monolith channel walls (Fig. 6a) and starts drying slowly (Fig. 6b). Initially,
only the large particles start touching each other as the liquid evaporates (Fig. 6c).
As the remainder of the liquid evaporates, the small binder particles are drawn by
capillary forces to the points were the larger particles touch each other (Fig. 6d).
Consequently, the binder particles will be present at the location where they are
most effective. Because the coated particles are relatively large, the interaction
between these particles in the absence of the small binder particles will be
insufficient to have a properly attached coat layer.
In slurry coating, the first step is to wet-mill the material to be coated to a
particle size of typically 5 mm. A detailed discussion on wet-milling of alumina for
slurry coating of monoliths is given by Blachou et al. (13). For alumina coating,
acid should be added to obtain a pH of 3 4. At both higher and lower pH values,
the viscosity increases, resulting in less effective milling and more difficult
coating. At a pH of about 3.5, the alumina particles are slightly charged and repel
each other and not too much alumina is being dissolved. Because the coating is
carried out using relatively large particles, the contact surface between these
particles and the support is small. Therefore, a binding agent should be added to
increase the contact surface. For alumina coating, this can be a colloidal alumina
solution or, for example, pseudo-boehmite. For silica coating, colloidal silica or
waterglass can be used. Colloidal solutions can be added before the milling of the
material, but for pseudo-boehmite or waterglass, it is recommended that these be
added after the milling process because these binders increase the viscosity of the
slurry. The amount of binder should be about 10 wt% of the total silica or alumina
content; the total amount of solids in the slurry should be between 40 and 50 wt%.
A detailed procedure for alumina coating is given in the next paragraph.
In a 3-L alumina ball-mill with agate balls (Gerhards ball-mill D-652354)
150 g of alumina (Condea Puralox, SBA-200, average particle size: 40 mm) is
milled together with 130 mL of demineralized water and 83 g of 20 wt% colloidal
alumina (Alfa, details in Table 3). Nitric acid is added to lower the pH to 3.5. The
slurry is milled for approximately 7.5 h. Periodically, samples of the slurry are
taken for particle size analysis using a Malvern Master Particle Sizer. The milling
ORDER REPRINTS

358 NIJHUIS ET AL.

Figure 6. Schematic representation of the drying steps in the slurry-coating process: (a) wet slurry
on surface [large white circles: material to be coated (catalyst support or catalyst, typically 5 mm);
black dots: binder (typically 20 50 nm)]; (b) first stages of drying, particles still suspended in liquid;
(c) large support/catalyst particles touch each other, binder still free-floating; (d) binder deposited at
interfaces between particles by capillary forces during final stages in drying.

is stopped once the (mass) average particle size is 5 mm. The slurry is then put into
a beaker in which a (dried) monolith is dipped for approximately 1 min. As
mentioned earlier, the dipping time does not significantly influence the result. The
excess liquid is shaken out of the monolith and the blocked channels are cleared
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 359

Figure 7. SEM micrographs of g-alumina-slurry-coated cordierite monolith (400 cpsi): (a) cross-
sectional view of channel wall (center: cordierite support: right: washcoat layer: right side: steep
edge of washcoat on wall); (b) top view of coating on walls.
ORDER REPRINTS

360 NIJHUIS ET AL.

using pressurized air. The monolith is dried horizontally in static air while
continuously being rotated around its axis. Finally, the monoliths are calcined in
air at 723 K for 4 h (heating and cooling rate of 10 K/min). Monoliths coated in this
manner typically contain 10 12 wt% alumina washcoat. Figure 7 shows a SEM
micrograph of the alumina coat layer deposited in this manner. A coat layer of
approximately 15 mm is present on the walls and hardly any alumina particles are
visible in the channel walls. In the corners of the channels, cracks can be observed
in the coat layer, most likely the results of a difference in thermal expansion
between the cordierite and the alumina. Ultrasonic testing of the coat layer
adherence in water, however, did not result in loss of coating.
For the coated 400-cpsi cordierite monolith, the theoretical coating thickness
can be calculated using the geometrical surface area of the monolith
(2740 m2/m3total ; the density of the coated monolith (485 kg/m3total ; and the
density of the coat layer (1360 kg/m3coat layer ; using alumina density of
3700 kg/m3Al2 O3 and porosity of particles of 0.4 and of layer of 0.3). This
theoretical thickness of 14 mm corresponds very well to the observed thickness of
15 mm. The BET surface area of the coat layer is 200 250 m2/g, which is the same
as the BET surface area of the original Puralox SBA-200.
In case a higher loading of washcoat is required, this can be obtained in two
ways. The monolith can be coated again after it has been dried, or it can be coated
again after calcination. Recoating after calcination yields a similar weight increase
as for the first coat layer. Coating again after drying typically has an increase in
loading of about 70% of the first coating amount, because part of the first coating
will detach during the second pass. Therefore, it is important to minimize the
dipping time to reduce this effect as much as possible. Coating smaller amounts
can be easily done by diluting the coating slurry.

4. Polymerization Coating of Carbon

After silica and alumina, carbon is mentioned as the third most used catalyst
support material (14); however, the number of practical applications is limited.
Carbon does have advantages over other types of support. For example, it is very
stable in acidic and basic media (15). The support is generally more inert than, for
example, alumina, thus reducing the side reactions catalyzed by the support itself.
Carbon has less tendency to react with the active phase (metal), such as nickel does
with alumina forming hard-to-reduce aluminates (16). Furthermore, it can have a
large surface area (600 m2/g or greater), it can easily be functionalized, and the
functionalized carbon by itself can have the desired catalytic activity for some
reactions. The greatest disadvantage of carbon as a catalyst support is the low
thermostability in air. A carbon (coated) monolith is not only attractive as a
catalyst support, but its application as a very low-pressure-drop adsorber is also
possible. At the moment, the largest application of (activated) carbons is in
adsorption processes. An extensive review on the preparation, properties, and
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 361

applications of carbon and carbon-coated monoliths is given by Vergunst et


al. (17).
When coating carbon on a support, the inertness of the carbon material
limits the possible coating options. Compared to silica and alumina, the
physical interaction between support and coating is small. A satisfactory way to
link the carbon coating to the support is, therefore, to deposit the carbon in the
pores and cracks of the support, using a pore-filling procedure. Such a
procedure for carbon coating has been developed by Vergunst et al. (18,19).
Alternatively, the coating can be carried out by dissolving a carbonaceous
polymer and using a dip-coating procedure with this solution. The basics of the
most thoroughly investigated carbon coating by furfuryl alcohol polymerization
will be discussed here briefly.
Furfuryl alcohol is placed in a well-mixed vessel, which has to be cooled to
remove the reaction heat of the polymerization reaction. Insufficient heat removal
easily results in a runaway with the mixture boiling out of the vessel. Over a period
of 45 min, the polymerization catalyst (65% nitric acid) is added (a total of 2.5 mL is
added to 100 mL of furfuryl alcohol). The mixture is to be kept at room temperature.
After the addition of the acid, the polymerization is allowed to proceed for another
hour. The monolith is dipped in the partially polymerized furfuryl alcohol for a few
minutes. After removing the monolith from the liquid, excess liquid is blown out by
pressurized air. After 15 min, the channels, which are again blocked by liquid from
the porous walls, are cleared again by pressurized air. The oligomerized coating on
the monolith is allowed to complete the polymerization process, first for a few hours
at room temperature and then overnight at 353 K. The polymer coating is
carbonized for 2 h in an inert gas stream (argon) at 823 K (heating rate 10 K/min).
Using this method, a carbon coating of around 16 wt% on a cordierite monolith
could be obtained, having a surface area (from CO2 adsorption) of typically
600 m2/gcoating. This method of carbon coating in literature is applied also for the
preparation of carbon membranes (20); surface areas (CO2 adsorption) reported for
this type of polymeric carbon are up to 900 m2/gcoating (21).
A different method for carbon coating ceramic monoliths is by impregnation
with a phenolic resole (22,23). This is a relatively easy procedure, which can be
carried out using low viscous (100 cP) commercially available solutions (22).
After impregnation, the monoliths are emptied by pressurized air, dried to remove
the water from the aqueous resole, and cured to cross-link the resin (423 K).
Finally, the coating is carbonized and activated (if desirable). Typical carbon
loadings obtained in a single coating step are about 6 7 wt%.
A problem that occurs in polymerization coating of carbon on ceramic
monoliths is that during the carbonization step, the coating shrinks. As a result, the
ceramic support will be partially exposed. From SEM micrographs, it is estimated
that about 40% of the support is exposed or only covered by a very thin carbon
layer (Fig. 8). The main advantage of the resistance of the carbon to corrosive
media will consequently be lost, because the ceramic support is still exposed to the
reactant mixture. The two other advantages of a carbon coating (inertness to side
ORDER REPRINTS

362 NIJHUIS ET AL.

Figure 8. Composition backscattered SEM micrograph of carbon-coated monolith. The light parts
were determined by energy-dispersive analysis by x-rays (EDX) to be cordierite, the dark gray parts
are the carbon coating.

reactions by support and the absence of reactions with the active phase) will hardly
be affected by the cracks, because even if all ceramic supports were exposed, the
surface area would be negligible (typically 0.25 m2/gcatalyst) compared to that of
the carbon layer (t100 m2/gcatalyst). Possibilities of reducing support exposure are
to repeatedly coat the monolith with intermediate carbonization, each time
reducing the amount of exposed support but also, unfortunately, the macropore
volume. Alternatively, small previously carbonized (not shrinking) carbon
particles can be added to the polymerization mixture. However, for the most
extreme conditions, the use of all-carbon monoliths will be the best solution,
although the preparation of those is more complex. Methods to prepare all-carbon
monoliths are discussed by Tennison (24) and DeLiso et al. (25). A method to
produce mixed carbon cordierite extruded monoliths is described by Gadkaree
and Jaroniec (26).

B. Deposition of the Active Phase on a Monolithic Support

The deposition of an active phase (metal) on a (washcoated) monolith body


is carried out in a similar manner as with regular catalyst support materials.
However, the large structure of a monolith can easily cause serious problems if the
metal is not deposited homogeneously. For example, if one step in the preparation
were to cause the active phase to be deposited at the external surface of the
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 363

support, in case of a simple extrudate catalyst, this would result in an egg-shell


type of active-phase distribution, which, for many processes, can be advantageous.
If, on the other hand, this occurs for a monolithic support, the result is an
inefficient monolith with only the outer part of the structure having a significant
catalytic activity. The steps in which such a process could happen are deposition
and drying; therefore, these will be discussed separately. For other steps
(calcination, reduction, and passivation of the active phase), a monolithic support
is not different from regular supports; therefore, these steps will not be discussed.
A detailed discussion on the deposition of metals on monolithic substrates can be
found in Ref. 19 and will be discussed here briefly.
Metal deposition procedures available are impregnation, ion exchange,
and deposition precipitation. Which of these procedures is best to use will
primarily depend on the metal precursor to be deposited and the desired
loading.

1. Deposition of the Active Phase by Impregnation

The simplest way to deposit a metal on a monolith is by impregnation.


For conventional catalyst supports, wet and dry (also called pore volume
or incipient wetness) impregnations are possible. For a large structure,
however, dry impregnation is difficult, as it is hard to supply the monolith
with exactly the amount of liquid corresponding to its pore volume, as the
liquid will have to travel a long distance to reach all the pores. The center of
the monolith can then easily remain dry while the external part of the
monoliths contains excess liquid, causing an uneven distribution. A variant on
the dry impregnation for platinum on alumina has been successfully applied
by Klinghoffer et al. (27). In this procedure, the monolith is allowed to suck
up the liquid by capillary forces, and after 5 min, the excess liquid is allowed
to drip out of the channels.
In the case of wet impregnation, first the amount of liquid a monolith will
adsorb must be determined. The metal precursor is dissolved so that the
concentration of metal in the liquid to be taken up by the monolith will produce
the desired metal loading. A dry monolith is immersed in this solution,
removed, and excess liquid blown out. To prevent an uneven distribution of
metal, especially if the metal precursor shows a significant interaction with the
support, the dipping procedure should be carried out in as short a time as is
practically feasible to prevent an excess of metal adsorbing on the support.
Because drying can also result in maldistribution, it is important to continue
immediately with this step, because many solvents will start evaporating
immediately after the monolith is removed from the liquid. If, out of practical
considerations, this is not possible, the wet monoliths should be kept in a
horizontal position while continuously being rotated to prevent gravity from
causing the liquid to flow to one side of the monolith.
ORDER REPRINTS

364 NIJHUIS ET AL.

Figure 9. Effect of drying after impregnation of cordierite monoliths with a nickel nitrate solution
(adapted from Ref. 19), the nickel oxide on the monolith has a dark color: (A) conventional drying in
static air: nickel oxide accumulation is visible at the outer rim of the monolith; (B) forced-airflow
drying: metal accumulation is visible at the point the airstream entered the monolith (top); (C)
microwave drying: a fairly even distribution is obtained, although the center contains the most nickel
(darkest); (D) freeze-drying: a homogeneous distribution is obtained. At the right bottom of each
photograph, the nickel on the monolith pieces is schematically depicted (darker color = more nickel).
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 365

Drying a Wet-Impregnated Monolith

When the metal has no interaction with the support but is only present as
solute in the liquid in the pores of the (coated) monolith, care should be taken that
the metal does not move through the support. If one part of the monolith is dry,
liquid will move from wetter parts of the monolith to the drier parts as a result of
capillary forces. Figure 9A shows the nickel distribution (the dark color) of a
monolith impregnated with a nickel nitrate solution and dried in static air at 363 K.
Almost all nickel has accumulated at the outer rim of the monolith where the water
evaporates fastest. Capillary forces redistribute the water over the monoliths,
taking the dissolved metal with it. Because most water evaporates at the outside of
the monolith, most metal will be deposited at this location. A metal accumulation
at the outer surface of the monolith during drying, as is visible in Figure 9A, was
also reported by Wahlberg et al. (28) for an incipient-wetness impregnation of
copper nitrate on both alumina and titania washcoated monoliths.
Ways to solve maldistribution of metal are to (1) dry the monolith evenly, (2)
dry faster than the liquid is able to redistribute, and (3) prevent movement of liquid.
The first way would be the ideal solution for the metal migration problem. Drying
in a microwave (Fig. 9C) is one way to heat a monolith more homogeneously and
thus dry more evenly. The metal distribution in this case is much more even.
Drying faster than the liquid redistribution (method 2) can be accomplished in two
ways: by slowing down the liquid movement or by increasing the drying rate.
Increasing the liquid viscosity (e.g., by adding glycerol or cellulose to water or by
changing the solvent) can slow down the liquid movement. This can improve the
metal distribution; however, the presence of these compounds will also slow down
the drying rate. It is easier to increase the drying rate. This can be done, for
example, by drying in a forced gas flow (Fig. 9B). Drying the same monolith as
was shown in Fig. 9A using a forced gas flow results in a slightly more even nickel
distribution, but a clear accumulation of metal at the top, where the gas stream
entered the monoliths and liquid evaporated fastest, is visible. Drying at higher
temperatures can also partly solve the metal movement problem; however, the
stability of the metal complex will restrict the allowable drying temperature.
One problem for both microwave drying and drying in a forced airflow is
that, for larger-sized monoliths, the distance for the evaporated solvent to travel
before it leaves the monolith becomes larger. If at one point in the monolith the gas
becomes more saturated with solvent, the drying at that point will be slower and
thus influence the distribution. The ideal drying procedure is therefore one in
which any liquid movement is completely ruled out (method 3), which can be done
by freeze-drying. In Figure 9D, it can be seen that the metal is evenly distributed in
a freeze-dried monolith. The major disadvantage of this procedure is that it is more
expensive than the other drying methods.
A general remark that should be made for all drying methods is that if the
drying (or for freeze-drying the solidifying of the solvent) takes more than a few
ORDER REPRINTS

366 NIJHUIS ET AL.

minutes, it is highly recommended to dry (freeze) horizontally while continuously


rotating the monolith around its axis to prevent movement of liquid by gravity.

2. Deposition of the Active Phase by Ion Exchange

A commonly applied method for depositing metals on a support is via ion


exchange (14). With this method, the support is put into a solution containing the
metal salts, in either positively or negatively charged complexes. These complexes
can adsorb on (or react with) surface groups of the support. An article by Hepburn
et al. (29) discusses the impregnation of different noble metals on g-alumina
monoliths and the resulting concentration profiles of the metals into the walls of
the monolith channels.
Platinum can be put on an alumina support by using, for example,
chloroplatinic acid (H2PtCl6). Each PtCl22 6 complex reacts strongly with two
surface Al OH groups. Typically g-alumina has 3-mmol/m2-surface OH groups
(14). With a typical surface area of 200 m2/g, this makes a platinum loading of 12
wt% possible if one were to use all surface OH groups. A platinum loading of
1 wt% on the support is more often desired. Simply using less acid for obtaining a
lower loading can cause an inhomogeneous distribution, because platinum
adsorption occurs very fast. For conventional catalyst supports, this will result in
egg-shell-type metal distributions. For a monolith, this can result in the majority of
the platinum being deposited at the channel entrance of the monolith. Adding a
competitively adsorbing acid helps in obtaining an even distribution. The effect of
the competitively adsorbing acid is clearly visible when the platinum distribution
over the monolith length is observed if a monolith is placed in ion-exchange
solution containing chloroplatinic acid (Fig. 10a). Without additional acid, most
platinum is deposited at the bottom of the monolith, where the liquid enters first.
Many other metal complexes used in ion-exchange metal depositions, such
as ammonium hexachlororuthenate(IV) on alumina, adsorb much slower on the
support and require ion-exchange times of a few hours to obtain 1 wt% loading.
For these cases, the risk of an extreme maldistribution as occurs for platinum will
not be present.
The effects of salts addition to obtain more even metal distributions in ion-
exchange preparations of monolithic catalysts is also discussed by Hepburn et al.
(30). All ion-exchange procedures for conventional catalysts that yield
homogeneous metal distributions on the support, in general, can also be used
for preparing monolithic catalysts. For these preparations, the circulation of liquid
through the monolith, necessary if the amount of liquid exceeds the monolith
volume, will not yield inhomogeneous distributions.
The advantage of an ion-exchange metal deposition procedure is that the
metal interacts with the support and metal maldistribution problems during drying
will not occur easily. However, if this interaction between metal and support is not
sufficiently strong, movement of metal can still occur and the drying rate then
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 367

Figure 10. Metal concentration as a function of the monolith length. (a) Platinum distribution on
g-alumina monolith for ion exchange of chloroplatinic acid solution for 30 min, with and without
hydrochloric acid present; (b) nickel distribution for a deposition precipitation using urea
decomposition with nickel nitrate on a cordierite monolith. (a and b) Metal deposited on a 25-cm
long, 1-cm diameter, 400-cpsi monolith. Determined using XRF (x-ray fluorescence spectroscopy)
on finely ground 2.5-cm-long sections of the monolith.
ORDER REPRINTS

368 NIJHUIS ET AL.

influences the metal distribution, as has been observed by Hepburn et al. (31) for
the deposition of rhodium on g-alumina monoliths.

3. Deposition of the Active Phase by Deposition Precipitation

A relatively easy method of adding an active phase on a monolithic support is


by deposition precipitation (32). The advantage of this method is that an
insoluble metal salt is deposited on the support, which can no longer move during
a drying step. Deposition precipitation can be carried out in two ways.
The first method for conventional deposition precipitation is to dissolve a
metal salt with a good solubility in a well-mixed vessel, together with the catalyst
support, and then to slowly add a second salt, which causes the first salt to
precipitate. This can be done, for example, by dissolving nickel nitrate and then
adding sodium hydroxide. If the metal is not available as a salt with a sufficient
solubility for the required metal loading, as an alternative both salt (dissolved) and
precipitating agent can be added simultaneously to the mixed vessel containing the
catalyst support. However, a problem arises if this method is applied for the
deposition of a metal on a monolithic support: In the liquid bulk, the metal salt
precursor complex is formed and, at one stage, the concentration will become
higher than the solubility because of the addition of the precipitating agent.
Addition of the complexing agent throughout all liquid is not possible and,
therefore, the liquid has to be well mixed. For deposition to occur on a monolith,
the liquid should also be circulating through the monolith, because otherwise only
the external surface will contain significant amounts of metal. The liquid reaches
supersaturation outside the monolith and the metal wants to deposit on the support.
However, an inhomogeneous deposition easily occurs. The monolithic support has
small channels, and the supersaturated liquid will deposit its metal at the first
possibility; therefore, the entrance of the monolith will contain more metal than
the exit. Only rapid circulation rates can help minimize this problem.
A better procedure for deposition precipitation when using a monolithic
support is one in which the supersaturation of the liquid occurs at the same rate in
all of the liquid, thus also in the liquid inside the monolithic channels. This can be
achieved by a depositionprecipitation step in which both the soluble metal salt
and precipitating agent are present in the liquid from the very beginning. For
example, the decomposition of urea to deposit nickel hydroxide from a nickel
nitrite solution can be used (33). Urea dissolved in water will decompose at
temperatures above 333 K according to Eq. (1), with a consecutive nickel
hydroxide deposition according to Eq. (2):
T.333K
CONH2 2 aq 1 3H2 Ol ! CO2 g 1 2NH1 2
4 aq 1 2OH aq 1

Ni21 aq 1 2OH2 aq !NiOH2 s 2


ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 369

The advantage of this method is that by slowly and uniformly heating a


monolith submerged in a metal salt solution including a decomposing
precipitating agent will cause a simultaneous supersaturation throughout the
liquid. Circulation is not necessary in this case and should be avoided. A
requirement for this procedure is that the solubility of the metal salt be
sufficiently high. The metal concentration in the starting solution should also be
sufficiently high for the required metal loading. The monolith should be placed
in a vessel as large as the monolith itself to prevent liquid outside the monolith
from causing an excess of metal loading at the external surface of the monolith.
Figure 10b shows the nickel distribution as a function of the monolith length
for a nickel-on-alumina washcoated monolith prepared in this manner. If the
solubility is insufficient for the required loading, the best way is to repeat this
procedure a number of times. Placing a monolith in a larger liquid volume and
then recirculating the liquid through the monolith will again cause an uneven
metal distribution, because the bulk of the liquid will also become
supersaturated and the metal will be deposited mostly at the monolith entrance.

4. Coating a Resin Catalyst on a Monolithic Support

Apart from metallic catalysts, nonmetallic catalysts are often used as well.
An example of these types of catalysts is Nafion (DuPont), a strongly acidic
polymeric catalyst used to catalyze many solid acid-catalyzed reactions. In this
subsection, the preparation of a Nafion-coated monolithic catalyst will be
discussed. The Nafion resin is a perfluorinated ion-exchange polymer, with a
backbone structure similar to Teflon, with pendant sulfonic acid groups (34).
Nafion by itself has a very low surface area (0.02 m2/g or less), and dispersing it on
a support or in a matrix enormously increases the activity (35).
The coating of a monolith is performed by immersing the monolithic support
in a solution of 5 wt% Nafion containing a mixture of lower-aliphatic alcohols and
10% water that is commercially available from DuPont. After dipping, the
monolith is removed from the dip solution, excess liquid is shaken out, and the
channels are cleared with air. The Nafion-coated monolith is then dried overnight
at room temperature while rotating in a horizontal position to establish a uniform
Nafion distribution over the monolith. Next, the structures are dried at 383 K for
16 h (overnight). To convert the active sites of the Nafion-coated monoliths into
the active form, the monolithic catalyst has to be exchanged in 25 wt% nitric acid
(35,36) and washed with demineralized water. Prior to the catalytic activity test,
the monolith is dried at 383 K for 15 h (overnight).
The catalyst loading of the Nafion coatings can be determined by
thermogravimetric analysis (TGA). Nafion loadings of 2 3 wt% are achieved
after dipping the monoliths once into the Nafion dip mixture. From texture
analysis, it can be concluded that the (macro)pores of the cordierite carrier are not
blocked by the catalyst particles. Only a small decrease in pore volume and pore-
ORDER REPRINTS

370 NIJHUIS ET AL.

size diameter is observed after dipping. Pore blocking is, in fact, not likely to occur
because the size of the Nafion particles in the dip mixture is approximately 50 nm
and the pore size from the carrier 1 10 mm (as determined by Hg porosimetry).
The activity of Nafion coated on a silica monolith in the acylation of anisole
with octanoic acid (0.21 0.25 L/gnafion/h, first-order rate constant) is comparable
to that of the commercially available Nafion silica composite SAC25 (a Nafion
silica composite with 25 wt% of Nafion: 0.19 L/gcat/h) (37). It can be concluded
that a very active monolithic catalyst system can be prepared in this way for acid-
catalyzed reactions.

C. Coating Ready-Made Catalysts on a Monolith

For the preparation of many types of catalysts, well-optimized procedures


already exist (e.g., in the preparation of zeolites or high-dispersion supported
metal catalysts). If these catalysts can simply be coated on a monolithic support,
these catalyst production methods can still be used and an optimized monolithic
catalyst is formed. A slurry-coating procedure (Fig. 6) like that discussed earlier
with particles of approximately 5 mm will produce an easily accessible coat layer.
Coating with much smaller particles is not recommended, because for small
particles, a dense layer can be formed with small pores in which diffusion
hindrance can occur. Coating with larger particles causes an irregular coat layer.
The result of this manner of coating is usually a system with a bimodal pore
distribution. Small pores are found, originating of the pores present in the coated
particles, with larger pores (typically 2 mm originating from the support and
interparticle voids between the coated particles).
A problem that will arise in the coating of an existing catalyst on the
monolithic support is the binding to the monolithic carrier. For example, if catalyst
particles are deposited, firing to a high temperature to bind the particles to the
support can result in the loss of a crystalline structure or in sintering for supported
metal catalysts. Therefore, it is advisible to coat utilizing a binding material.
The use of a binding material to coat a catalyst on a monolith has an
important disadvantage: For supported metallic catalysts, the binder can cover the
active (metallic) phase, and for zeolites, it can block the pores. To overcome this
drawback as much as possible, the amount of binding material should be
minimized and the covering potential should be reduced as much as possible. If a
colloidal material (silica or alumina) is used as a binder, the binder consists of
particles and is therefore less likely to fully block active regions of the catalyst.
The use of more or less completely dissolved binders (or binders consisting of
nanometer-sized particles) like pseudo-boehmite or sodium silicate (waterglass) is
not recommended, because they are able to cover active regions. When coating a
monolith with support material, this binder problem is, of course, not relevant,
because then the binder is the same material as the support, and the active phase is
applied at a later stage.
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 371

For coating of a ready-made catalyst on a monolith, first a general procedure


will be discussed and then a detailed recipe for coating zeolite BEA (Beta) will be
given. Other recipes for zeolite coating can be found, for example, in Refs. 38 and 39.
To prepare a coating slurry, a number of components are needed: a solvent,
properly sized catalyst particles, the binder, and, optionally, a surfactant and a
temporary binder. The solvent to be used depends on the catalyst to be coated. For
most catalysts, water will do as a solvent; for hydrophobic catalysts, this will not
work properly. For example, if a coating slurry in water is made for the coating of
a cobalt lanthanum perovskite, this will demix into two separate phases if mixing
is stopped and the production of a homogeneous slurry will be hindered by foam
formation during mixing. If this highly unstable slurry will be used for coating,
this results in an inhomogeneous and irreproducible coating. For such type of
catalyst, it is better to use an organic solvent like diethylether, ethylacetate, or
butylacetate. These solvents have the additional advantage that they evaporate
faster and have a lower surface tension. Another consideration for selecting a
solvent is the viscosity of the dipping slurry. For example, if zeolite BEA
(Si=Al 37:5; Zeolyst CP 811E-75) is dispersed in butylacetate, a zeolite content
above 10 wt% causes the solution to be too viscous to use for coating, whereas in
water, a 20-wt% slurry is still usable. The interaction between catalyst and solvent
determines the amount that can be put into the slurry. For example, if a coating is
made with zeolite BEA, a 20-wt% slurry has a very high viscosity of around
1000 mPa:s,* usable only for coating small monoliths or large-channel monoliths,
whereas for a 30-wt% slurry of MFI (Si=Al 15; Zeolyst CBV 3024E) the
viscosity is typically as low as 6 mPa:s.
The binder is preferably colloidal silica or alumina, depending on the
application. Alumina has the advantage that it is more thermostable than silica,
because in high-temperature applications, silica binder can be lost by steaming. On
the other hand, alumina can introduce acidity into the system, making the use of a
silica binder more favorable for some applications. The amount of binder to be
used is preferably kept minimal. Patil and Lachman (40) demonstrated the
activity-decreasing effect of the binder in coating monoliths with silicalite for the
decomposition of methanol; it should be noted that they used large amounts of
binder (.10 wt% of total solids). In our study, we demonstrated that for coating a
cordierite monolith with zeolite BEA, even 1 wt% (of total solids) colloidal silica
binder is sufficient to properly attach the zeolite to the support. This has been
checked by placing a monolith coated with 7 wt% of zeolite BEA in an ultrasonic
waterbath for 15 min, after which the weight loss was only 0.14 wt% (equal to
losing 2 wt% of the coating). About the same weight loss occurred for an uncoated
cordierite monolith, so the real loss of coating is much smaller.

*This slurry is a pseudoplastic (non-Newtonian) liquid while mixing, and during the emptying of the
channels by the pressurized area, the apparent viscosity is much lower.
ORDER REPRINTS

372 NIJHUIS ET AL.

The fact that this small amount of binder is able to properly adhere the zeolite
to the support can be easily explained: When the still-wet monolith with coating is
drying, the large slurry particles start contacting each other and the support as the
solvent evaporates. The colloidal binder particles stay in the remaining liquid.
During drying, the liquid will be drawn by the capillary forces toward the capillary
spaces, which include the contact points between the slurry particles and support.
As a result, the colloidal particles will accumulate at the point where they are
needed. For this principle to work, two demands have to be met. First, the solvent
must wet both support and slurry particles well, because otherwise the liquid will
be drawn to one of them during drying and bonding will not occur (e.g., when
coating relatively hydrophobic particles in water as a solvent). Second, the binder
particles must be much smaller than the slurry particles, as otherwise the
movement of binder particles with the drying liquid principle will not work
properly. If micrometer-sized particles are coated with a colloidal silica or alumina
as a binder (typically a few tens of nanometers in size), this demand is met. It
should naturally be noted that if the size of the binder particles is similar or smaller
than the micropores of the catalyst to be coated, the catalyst pores might be
blocked by the binder particles. The catalyst then would lose efficiency, thus
limiting the minimum usable size for the binder particles.
Optionally, a surfactant can be added to the coating slurry to facilitate the
dispersion of the particles in the slurry. Typically, one would add 2% of the total
solids weight. A temporary binder should be added to bind the particles to the
support prior to calcination. Because only a small amount of permanent binder is
used, the interaction between particles and support is relatively small, as long as
the binder particles are not melted in between the larger particles during the
calcination step. To prevent particles from falling off during the handling of the
monolith before the calcination, a temporary binder, like a cellulose compound,
can help to keep the particles attached to the monolith. During calcination, this
compound is burned off. If a surfactant is used, this compound is also able to
increase the physical interaction between the particles and the support and a
separate temporary binder is usually not needed. The amount of temporary binder
should be sufficient to cover all particles, including the support, and is usually
equal to a few weight percent of the total solids.
All components of the coating slurry should be well mixed using a high-shear
mixer until the slurry is homogeneous. A dried monolith should be dipped into this
slurry for a short period. Very short (a few seconds) dipping times result in a
slightly higher loading (5 10% relative increase) than longer dipping times (a few
minutes), because a dry monolith will rapidly absorb liquid and draw extra
particles against the wall (slip-casting). A longer period of dipping will reduce this
amount. After dipping, excess liquid is shaken out of the monolith and the
channels are cleared using pressurized air. The monolith is dried horizontally
while continuously being rotated around its axis. Finally, the monolith is calcined
in air, typically at 673 1173 K depending on the catalyst coated. The minimum
calcination temperature of 673 K to obtain a good physical interaction among
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 373

Figure 11. SEM micrographs of BEA Si=Al 37:5-coated cordierite monolith (400 cpsi):
(a) cross-sectional view of channel wall; (b) top view of coating on walls.
ORDER REPRINTS

374 NIJHUIS ET AL.

catalyst, binder, and support is, of course, limiting the applicability of this method
to catalysts thermostable to at least this temperature.

1. Slurry-Coating a Cordierite Monolith with Zeolite BEA

The optimally developed procedure for coating a 400-cpsi cordierite


monolith with zeolite BEA consists of the following steps. One hundred fifty
grams of demineralized water are mixed with 0.96 g of colloidal silica (40% in
water, Ludox AS-40) and 0.64 g of surfactant [Teepol (Shell), an alkylarylsulfo-
nate-type surfactant]. While mixing with a high-shear mixer (13,000 rpm), 36 g of
BEA (Si=Al 37:5; Zeolyst CP 811E-75, average particle size: 7 mm, 90% larger
than 3.5 mm and smaller than 14 mm) is added, after which the slurry is mixed for
another minute. A dried monolith is submerged in the slurry for 3 min; the
remaining liquid is shaken out and the monolith is dried overnight. Finally, the
monolith is calcined in air at 723 K for 4 h (heating and cooling rate 10 K/min). For
a 400-cpsi monolith, this results in loadings of typically 5 7 wt%. A SEM
micrograph of the BEA-coated monolith is shown in Figure 11. It can be
calculated that this loading should have a layer thickness of approximately 7 mm,
which is comparable to the particle size of the coated particles, and also explains
why a top view of the coat layer shows some exposed cordierite. Mercury
porosimetry for a 7-wt% coated monolith showed an increased macropore volume
of 0.3 mL/g (bare sample of same cordierite 0.2 mL/g), indicating a macroporous
coating structure. Activity measurements in the esterification of 1-octanol with
hexanoic acid (41) showed that the activity of the coated system was only
approximately 20% lower than the activity of the same zeolite in a slurry system,
indicating that the zeolite hardly loses activity as a result of coverage by binder or
other particles. Ultrasonic testing of coating adhesion showed no significant loss of
coating.

D. Other Types of Monolithic Catalyst


1. Growing Zeolites on a Monolithic Support

Zeolites can also be applied on a support like a monolith by performing the


zeolite synthesis in the presence of the support. This one-step, binderless method
in which the zeolites are grown directly on the surface of the carrier is already
reviewed in literature (42,43).
Also, numerous patents have come out discussing the growing of a zeolite on
a substrate. The preparation of MFI-type zeolite coatings on cordierite monolithic
supports was discussed by Grasselli et al. (44). Methods for growing zeolites X, Y,
MFI, and mordenite on cordierite supports were presented by Lachman and Patil
(45). The synthesis of MFI-type zeolite coatings on stainless-steel supports was
discussed by Geus et al. (46). The preparation of an ultrathin layer of zeolite on a
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 375

substrate by dipping this substrate in the synthesis mixture prior to synthesis was
described by Suzuki (47).
The in situ synthesized zeolite coating can be prepared in several ways,
such as immersing the support in a zeolite synthesis mixture or in a liquid that
only contains template. This latter method can only be applied to a (silica
containing) support which acts as a source for the zeolite synthesis. When
applying this in situ synthesis method (at least for the MFI synthesis), the
composition of the synthesis mixture, the temperature, and the duration are of
extreme importance for the resulting thickness of the zeolite coating and the
orientation of the zeolite crystals. Also important is whether the solution is
saturated. When applying this technique for preparing a zeolite coating, possible
corrosion of the support should be taken into account. To avoid this, mild
synthesis conditions can be employed or a protective layer can be applied onto
the support (42). Also, the type of support should be take into account. To control
the zeolite coating by means of synthesizing the zeolite in the presence of a
support, an understanding of the general theory of heterogeneous nucleation and
crystallization is needed (48).
An example is the application of BEA onto a distillation column packing for
the synthesis of ETBE (ethyl-tert-butyl ether) from ethanol and isobutene (49).
Synthesis of BEA resulted in zeolite coverages of up to 14.5 g/m2 of support. In a
monolith with a channel diameter of 2 mm, this corresponds to a zeolite loading of
,30 kg/m3 of packing. When comparing the performance of this synthesized BEA
coating with commercial BEA powder in the synthesis of ETBE, it was found that
the turnover numbers of the BEA powder was four times higher than those of the
coating, but the selectivity was comparable.
It has been demonstrated that it is possible to grow zeolites on monolithic
structures directly. The main disadvantage of this method, however, is that it is
considerably more complex than slurry-coating a zeolite on a monolith. The main
advantage, on the other hand, is that growing methods result in a supported zeolite
of which a binder cannot hinder the accessibility of reactants to the catalytically
active inside of the zeolite particles.

2. Extruded Monolithic Catalysts

As an alternative to coating a catalyst on a monolithic support, it is possible


to extrude the catalyst support material or the catalyst itself with some additional
binder into a monolithic shape. Lachman et al. (50 52) and Lee et al. (53)
described different methods for preparing such systems. The main advantage of
such a system is that as the monolith wall consists of catalyst, the total amount of
catalyst available can be much larger than for a coated system. For applications
like the pressure-swing separation of air described by Li et al. (54,55), this is a
great advantage. Disadvantages are, however, that the fabrication of such a system
is much more specialistic and requires dedicated extruders. Furthermore, the
ORDER REPRINTS

376 NIJHUIS ET AL.

catalyst itself determines aspects like the mechanical strength. Most importantly,
if the catalyst is buried in a wall, the effectiveness of the catalyst is much lower
than the same catalyst in a washcoated form. Patil and Lachman (40) compared
activities for washcoated and extruded systems of a silicalite catalyst for methanol
decomposition and found an activity of only 30% for the extruded system of the
value they obtained for washcoated systems. This difference in activity was
attributed to catalyst blockage.

3. Other Types of Monolith Catalyst Preparations

Although usually the active phase and the catalyst support material are
deposited separately on a monolithic body, these two steps can also be combined.
The catalyst precursor is then dissolved in the solvent (usually water) in which the
coat slurry of the support material is made. An advantage of this approach is that
the number of catalyst synthesis steps is reduced. A problem that might arise,
however, is that the active phase is also subjected to the higher calcination
temperature that is usually necessary for fixating the coated particles to the
monolithic support. Furthermore, freeze-drying, which works best for immobiliz-
ing a less strongly bound active phase during drying, is not recommended for
depositing the catalyst support, because with this drying method, the support
particles are not drawn to each other while the solvent evaporates. For active
phases of which the precursor has a stronger interaction with the support, this
combination of synthesis steps is possible. An example of the successful combined
deposition of the support and the active phase on a monolithic carrier is given by
Gadkaree et al. (56), who provided a method to carbon-coat a ceramic monolith
while simultaneously depositing different metals, resulting in highly active
catalyst systems.
A combined way of depositing the active phase and support was described by
Wahlberg et al. (28) for the deposition of copper on an alumina-coated monolithic
body. In this study, copper nitrate is deposited on an alumina or titania slurry by
means of urea decomposition, after which the slurry is coated on the monolith.
Because the metal was deposited on the support in this manner, a homogeneous
metal distribution was obtained. A disadvantage of this manner of preparation
turned out to be that the washcoat obtained was brittle and did not adhere as well as
the bare alumina or titania that was coated.
Alternative ways of coating a catalyst, rarely mentioned in the literature in
reference to monolithic supports, but very well possible are, for example, chemical
vapor deposition and sputtering. Both of these techniques have the advantage that
very thin and homogeneous catalytic layers can be deposited on a support.
Because deposition is carried out in the gas phase, maldistribution as a result of
liquid movement (either by gravity or by capillary forces) during drying is
dismissed. During the deposition, it is, however, important that the monolith
temperature is uniform and that the gas is properly circulated through the structure,
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 377

because otherwise homogeneous distributions will not be obtained. The major


disadvantages of both chemical vapor deposition and sputtering are the
complexity and high cost. One recent article discussed the application of
chemical vapor deposition to the preparation palladium catalytic monoliths (57).
The palladium distributions obtained, however, are very inhomogeneous (all
palladium is deposited within a few millimeters of the point where the palladium
containing gas stream enters the monolith), indicating that this technique still
needs further development before it can be applied in practice.

IV. CONCLUSIONS

The preparation of monolithic catalysts requires other steps than commonly


applied for traditional systems. Coating procedures are available for many
supports and ready-made catalysts, each having their own advantages and
disadvantages.
Applying the active phase on a monolith requires extra care to avoid
maldistribution over the support. The large size of the structure makes this more
important than for conventional (small) carriers. Proper deposition procedures
for metals on monoliths are ion exchange, homogeneous deposition
precipitation (e.g., by urea decomposition), and impregnation followed by
freeze-drying.
The availability of well-developed procedures for preparing monolithic
catalysts discussed in this article, the low costs of these carriers, combined with the
operational advantages (no stirring, low pressure drop, no catalyst separation, etc.)
make the large-scale application of monolithic catalysts feasible.

ACKNOWLEDGMENTS

Corning Inc. is acknowledged for the supply of monolithic supports. Nynke Aalders
is thanked for her work on the development of solid-acid-coated monolithic catalysts.

REFERENCES

1. Geus, J.W.; Giezen, J.C. van. Monoliths in Catalytic Oxidation. Catal. Today 1999,
47, 169 180.
2. Nijhuis, T.A.; Kreutzer, M.T.; Romijn, A.C.J.; Kapteijn, F.; Moulijn, J.A.
Monolithic Catalysts as Efficient Three-Phase Reactors. Chem. Eng. Sci. 2001,
56, 823 829.
3. Berglin, C.T.; Herrmann, W. Method in the Production of Hydrogen Peroxide. US
Patent 4,552,748, 1985 (assigned to EKA AB).
4. Kapteijn, F.; Heiszwolf, J.J.; Nijhuis, T.A.; Moulijn, J.A. Monoliths in Multiphase
Catalytic ProcessesAspects and Prospects. Cattech 1999, 3, 24 41.
ORDER REPRINTS

378 NIJHUIS ET AL.

5. Zwinkels, M.F.M.; Jaras, S.G.; Menon, P.G.; Asen, K.I. Preparation of Anchored
Ceramic Coatings on Metal Substrates: A Modified Sol Gel Technique Using
Colloidal Silica Sol. J. Mater. Sci. 1996, 31, 6345 6349.
6. Elmer, T.H. Ultra-low Expansion Ceramic Articles. US Patent 3,958,058, 1976
(assigned to Corning Glass Works).
7. Kolb, W.B.; Papadimitriou, A.A.; Cerro, R.L.; Leavitt, D.D.; Summers, J.C. The Ins
and Outs of Coating Monolithic Structures. Chem. Eng. Prog. 1993, February,
63 67.
8. Heck, R.M.; Farrauto, R.J. Catalyst Characterization. In Catalytic Air Pollution
ControlCommercial Technology; John Wiley & Sons: New York, 1995, 29 47.
9. Beauseigneur, P.A.; Lachman, I.M.; Patil, M.D.; Swaroop, S.H.; Wusirika, R.R. Pore
Impregnated Catalyst Device. US Patent 5,334,570, 1994 (assigned to Corning Inc.).
10. Duisterwinkel, A.E. Clean Coal Combustion with In-Situ Impregnated Sol Gel
Sorbent. Ph.D. Thesis, Delft University of Technology, Delft, 1991.
11. Zwinkels, M.F.M.; Jaras, S.G.; Menon, P.G. Preparation of Combustion Catalysts by
Washcoating Alumina Whiskers-Covered Metal Monoliths Using a Sol Gel
Method. Studies Surface Sci. Catal. 1995, 91, 85 94.
12. Addiego, W.P.; Lachman, I.M.; Patil, M.D.; Williams, J.L.; Zaun, K.E. High Surface
Area Washcoated Substrate and Method for Producing Same. US Patent 5,212,130,
1993 (assigned to Corning Inc.).
13. Blachou, V.; Goula, D.; Phillippopoulos, C. Wet Milling and Preparation of Slurries
for Monolithic Structures Impregnation. Ind. Eng. Chem. Res. 1992, 31, 364 369.
14. Geus, J.W.; van Veen, J.A.R. Preparation of Supported Catalysts. Studies Surface
Sci. Catal. 1993, 79, 335 360.
15. Radovic, L.R.; Rodrguez-Reinoso, F. Carbon Materials in Catalysis. Chem. Phys.
Carbon 1997, 25, 243 358.
16. Kapteijn, F.; Moulijn, J.A.; Tarfaoui, A. Temperature Programmed Reduction and
Sulphiding. Studies Surface Sci. Catal. 1993, 79, 401 417.
17. Vergunst, Th.; Linders, M.J.G.; Kapteijn, F.; Moulijn, J.A. Carbon Based Monolithic
Structures. Catal. Rev. 2001, 3, 291 314.
18. Vergunst, Th.; Kapteijn, F.; Moulijn, J.A. Carbon Coating of Ceramic Monolithic
Substrates. Studies Surface Sci. Catal. 1998, 118, 175 183.
19. Vergunst, Th. Carbon Coated Monolitic CatalystsPreparation Aspects and Testing
in the Three-Phase Hydrogenation of Cinnamaldehyde. Ph.D. Thesis, Delft
University of Technology, Delft, 1999.
20. Chen, Y.D.; Yang, R.T. Preparation of Carbon Molecular Sieve Membrane and
Diffusion of Binary Mixtures in the Membrane. Ind. Eng. Chem. Res. 1994, 33,
3146 3153.
21. Moreno-Castilla, C.; Mahajan, O.P.; Walker, P.L., Jr.; Jung, H.-J.; Vannice, M.A.
Carbon as a Support for CatalystsIII. Glassy Carbon as a Support for Iron. Carbon
1980, 18, 271 276.
22. Gadkaree, K.P. Carbon Honeycomb Structures for Adsorption Applications. Carbon
1998, 36, 981 989.
23. DeLiso, E.M.; Gadkaree, K.P.; Mach, J.F.; Streicher, K.P. Carbon-Coated Inorganic
Substrates. US Patent 5,451,444, 1995 (assigned to Corning Inc.).
24. Tennison, S.R. Phenolic-Resin-Derived Activated Carbons. Appl. Catal. A 1998,
173, 289 311.
ORDER REPRINTS

PREPARATION OF MONOLITHIC CATALYSTS 379

25. DeLiso, E.M.; Lachman, I.M.; Patil, M.D.; Zaun, K.E. Activated Carbon Structures.
US Patent 5,356,852, 1994 (assigned to Corning Inc.).
26. Gadkaree, K.P.; Jaroniec, M. Pore Structure Development in Activated Carbon
Honeycombs. Carbon 2000, 38, 983 993.
27. Klinghoffer, A.A.; Cerro, R.L.; Abraham, M.A. Catalytic Wet Oxidation of Acetic
Acid Using Platinum on Alumina Monolith Catalyst. Catal. Today 1998, 40, 59 71.
28. Wahlberg, A.; Petterson, L.J.; Bruce, K.; Andersson, M.; Jansson, K. Preparation
Evaluation and Characterization of Copper Catalysts for Ethanol Fuelled Diesel
Engines. Appl. Catal. B: Environ. 1999, 23, 271 281.
29. Hepburn, J.S.; Strenger, H.G.; Lyman, C.E. Distributions of HF Co-impregnated
Rhodium, Platinum and Palladium in Alumina Honeycomb Supports. Appl. Catal.
1989, 55, 271 285.
30. Hepburn, J.S.; Strenger, H.G.; Lyman, C.E. Co-impregnation of Rhodium into
Alumina Honeycombs with Acids and Salts. Appl. Catal. 1998, 56, 107 118.
31. Hepburn, J.S.; Strenger, H.G.; Lyman, C.E. Effects of Drying on the Preparation of
HF Co-impregnated Rhodium Al2O3 Catalysts. Appl. Catal. 1989, 55, 287 299.
32. de Jong, K.P. Deposition Precipitation onto Pre-shaped Carrier Bodies. Possibilities
and Limitations. Studies Surface Sci. Catal. 1991, 63, 19 36.
33. Knijff, L.M.; Bolt, R.H.; van Yperen, R.; van Dillen, A.J.; Geus, J.W. Production of
Nickel-on-Alumina Catalysts from Preshaped Support Bodies. Studies Surface Sci.
Catal. 1991, 63, 165 174.
34. Waller, F.J.; Warren van Scoyoc, R. Catalysis with Nafion. Chemtech 1987, July,
438 441.
35. Harmer, M.A.; Farneth, W.E.; Sun, Q. High Surface Area Nafion Resin/Silica
Nanocomposites: A New Class of Solid Acid Catalyst. J. Am. Chem. Soc. 1996, 118,
7708 7715.
36. Olah, G.A.; Iyer, P.S.; Surya Prakash, G.K. Perfluorinated Resinsulfonic Acid
(Nafion-H) Catalysis in Synthesis. Synthesis 1986, July, 513 531.
37. Beers, A.E.W.; Hoek, I.; Nijhuis, T.A.; Downing, R.S.; Kapteijn, F.; Moulijn, J.A.
Structured Catalysts for the Acylation of Aromatics. Topics Catal. 2000, 13,
275 280.
38. Patil, M.D.; Socha, L.S.; Lachman, I.M. Dual Converter Engine Exhaust System for
Reducing Hydrocarbon Emissions. US Patent 5,125,231, 1992 (assigned to Corning
Inc.).
39. Lachman, I.M.; Patil, M.D.; Socha, L.S.; Swaroop, S.H.; Wusirika, R.R. Molecular
Sieve-Palladium-Platinum Catalyst on a Substrate. US Patent 5,244,852, 1993
(assigned to Corning Inc.).
40. Patil, M.D.; Lachman, I.M. Methanol Conversion on Ceramic Honeycombs Coated
with Silicalite. In Perspectives in Molecular Sieve Science; Flank, W.H., Ed.;
American Chemical Society: Washington, DC, 1988; 492 499.
41. Beers, A.E.W.; Spruijt, R.A.; Nijhuis, T.A.; Kapteijn, F.; Moulijn, J.A. Esterification
in a Structured Catalytic Reactor with Counter-Current Water Removal. Catal.
Today 2001, 56, 175 181.
42. Jansen, J.C.; Koegler, J.H.; van Bekkum, H.; Calis, H.P.A.; van den Bleek, C.M.;
Kapteijn, F.; Moulijn, J.A.; Geus, E.R.; van der Puil, N. Zeolitic Coatings and Their
Potential Use in Catalysis. Micropor. Mesopor. Mater. 1998, 21, 213 226.
43. Coronas, J.; Santamara, J. Catalytic Reactors Based on Porous Ceramic Membranes.
Catal. Today 1999, 51, 377 389.
ORDER REPRINTS

380 NIJHUIS ET AL.

44. Grasselli, R.K.; Lago, R.M.; Socha, R.F.; Tsikoyiannis, J.G. Synthesis of Zeolite
Films Bonded to Substrates, Structures and Uses Thereof. US Patent 5,310,714, 1994
(assigned to Mobil Oil Corporation).
45. Lachman, I.M.; Patil, M.D. Method of Crystallizing a Zeolite on the Surface of a
Monolithic Ceramic Substrate. US Patent 4,800,187, 1989 (assigned to Corning
Inc.).
46. Geus, E.R.; Bakker, W.J.W.; Moulijn, J.A.; van Bekkum, H.; Jansen, J.C. Module
Containing Zeolite-Based Membrane and Preparation Thereof. US Patent 5,744,035,
1998 (assigned to Exxon Chemical).
47. Suzuki, H. Composite Membrane Having a Surface Layer of an Ultrathin Film of
Cage-Shaped Zeolite and Processes for Production Thereof. US Patent 4,699,892,
1987.
48. Jansen, J.C.; Kashchiev, D.; Erdem-Senatalar, A. Preparation of Coatings of
Molecular Sieve Crystals for Catalysis and Separation. Studies Surface Sci. Catal.
1994, 85, 215 250.
49. Oudshoorn, O.L.; Janissen, M.; van Kooten, W.E.J.; Jansen, J.C.; van Bekkum, H.;
van den Bleek, C.M.; Calis, H.P.A. A Novel Structured Catalyst Packing for
Catalytic Distillation of ETBE. Chem. Eng. Sci. 1999, 54, 1413 1418.
50. Lachman, I.M.; Bardhan, P.; Nordlie, L.A. Preparation of Monolithic Catalyst
Support Structures Having an Integrated High Surface Area Phase. US Patent
4,631,268, 1986 (assigned to Corning Glass Works).
51. Lachman, I.M.; Golino, C. Preparation of High Surface Area Agglomerates for
Catalyst Support and Preparation of Monolithic Support Structures Containing
Them. US Patent 4,657,880, 1987 (assigned to Corning Glass Works).
52. Lachman, I.M.; Patil, M.D.; Williams, J.L.; Wusiraka, R.R. Catalytically Active
Materials and Method for Their Preparation. US Patent 4,912,077, 1990 (assigned to
Corning Inc.).
53. Lee, L.Y.; Perera, S.P.; Crittenden, B.D.; Kolaczkowski, S.T. Manufacture and
Characterisation of Silicalite Monoliths. Adv. Sci. Technol. 2000, 18, 147 170.
54. Li, Y.Y.; Perera, S.P.; Crittenden, B.D. Zeolite Monoliths for Air Separation, Part 1:
Manufacture and Characterization. Chem. Eng. Res. Des. 1998, 76, 921 930.
55. Li, Y.Y.; Perera, S.P.; Crittenden, B.D. Zeolite Monoliths for Air Separation, Part 2:
Oxygen Enrichment, Pressure Drop and Pressurization. Chem. Eng. Res. Des. 1998,
76, 931 941.
56. Gadkaree, K.P.; Patil, M.D.; Dawes, S.B. Method of Making Activated Carbon
Having Dispersed Catalyst. US Patent 5,488,023, 1996 (assigned to Corning Inc.).
57. Cominos, V.; Gavriilidis, A. Preparation of Axially Non-uniform Monoliths by
Chemical Vapour Deposition. Appl. Catal. A: Gen. 2001, 210, 381 390.
Request Permission or Order Reprints Instantly!

Interested in copying and sharing this article? In most cases, U.S. Copyright
Law requires that you get permission from the articles rightsholder before
using copyrighted content.

All information and materials found in this article, including but not limited
to text, trademarks, patents, logos, graphics and images (the "Materials"), are
the copyrighted works and other forms of intellectual property of Marcel
Dekker, Inc., or its licensors. All rights not expressly granted are reserved.

Get permission to lawfully reproduce and distribute the Materials or order


reprints quickly and painlessly. Simply click on the "Request
Permission/Reprints Here" link below and follow the instructions. Visit the
U.S. Copyright Office for information on Fair Use limitations of U.S.
copyright law. Please refer to The Association of American Publishers
(AAP) website for guidelines on Fair Use in the Classroom.

The Materials are for your personal use only and cannot be reformatted,
reposted, resold or distributed by electronic means or otherwise without
permission from Marcel Dekker, Inc. Marcel Dekker, Inc. grants you the
limited right to display the Materials only on your personal computer or
personal wireless device, and to copy and download single copies of such
Materials provided that any copyright, trademark or other notice appearing
on such Materials is also retained by, displayed, copied or downloaded as
part of the Materials and is not removed or obscured, and provided you do
not edit, modify, alter or enhance the Materials. Please refer to our Website
User Agreement for more details.

Order now!

Reprints of this article can also be ordered at


http://www.dekker.com/servlet/product/DOI/101081CR120001807

You might also like