Sunteți pe pagina 1din 37

I NTERNATIONAL J OURNAL OF C HEMICAL

R EACTOR E NGINEERING
Volume 1 2003 Review R6

Recent Research in Catalytic Inorganic


Membrane Reactors
Anthony G. Dixon∗


Worcester Polytechnic Institute, agdixon@wpi.edu

ISSN 1542-6580
Copyright 2003
c by the authors.
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system,
or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording, or
otherwise, without the prior written permission of the publisher, bepress, which has been given
certain exclusive rights by the author.
Recent Research in Catalytic Inorganic Membrane
Reactors
Anthony G. Dixon

Abstract
The two most important, and often the most expensive, steps in a chemical
process are usually the chemical reactor and the separation of the product stream.
Both the process economics and the efficient use of natural resources could be
improved by the combination of these two operations into a single unit opera-
tion, leading to potential savings in energy and reactant consumption and reduced
by-product formation. One promising way to accomplish this combination is the
use of membrane separation and catalytic reaction together in a multifunctional
reactor. Until relatively recently, the use of membranes was restricted to low tem-
perature processes with mild chemical environments, which could be tolerated by
polymeric materials. Recent advances in inorganic materials have expanded the
range of membrane use, to include high temperature and chemically harsh envi-
ronments. This has allowed inorganic membranes to be integrated into catalytic
reactors. This area was reviewed previously by the present author (Dixon, 1999).
The present contribution seeks to review literature and new developments in the
succeeding four and a half year period, since the end of 1998. Research directions
that were previously considered promising are re-evaluated here, and new ideas
since then are presented.

KEYWORDS: Inorganic membrane reactor, catalysis, catalytic membrane reac-


tor
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 1

1. INTRODUCTION
This review begins with a brief survey of well-established uses of membrane reactors. The first is the use of a membrane
to remove a product from a reaction limited by chemical equilibrium, thus allowing higher conversions than would be
possible in a conventional fixed bed catalytic reactor. The second area is the distributed feed of reactant throughout the
reactor volume, and a third is the use of the membrane to control how the reactants contact. The extent to which these
areas are of interest today is illustrated.

The second part of the review presents a look at the current state of membrane reactor evaluation. Despite the
advances of recent years, many challenges and difficulties still face catalytic membrane reactors. Some are technical:
in particular, better materials need to be developed to address problems of stability, and of obtaining high permeation
rate and high permselectivity in the same material. Some recent work addresses commercial or economic evaluation.
Other contributions seek to develop criteria to determine when membrane reactors can outperform conventional reactors.

The main focus of the review is in the third part. Several trends in recent research are examined, including the
application of hydrogen removal membranes in methane utilization, the decline in methane oxidative coupling and the
rise of oxidative dehydrogenation, the use of distributed feed to control temperature in partial oxidations, the increase
in detailed modeling of membrane reactors, the increased appreciation of the effects on the catalyst of the changed
reaction environment in a membrane reactor, and the strong increase in research into zeolite membrane reactors. The
purpose of this contribution is to provide an interpretative review of these and other recent advances, to place them in
context, and to identify trends for the future. In particular, reference will be made to those areas which were identified
as innovations in a previous review (Dixon, 1999), and to how they have progressed.

The scope of the review is limited to inorganic membrane reactors under chemically harsh environments, or
at high temperatures (> 100°C). We will not discuss bioreactors, biological membranes, and enzyme catalysis or
polymeric membranes. Very good recent papers and reviews have discussed biological membrane reactors (Sirkar et al.,
1999) and polymeric membrane reactors (Bengtson et al., 2002) applied to such areas as pervaporation (Lim et al., 2002),
fuel cells (Thampan et al., 2001) and photocatalysis (Molinari et al., 2000). In addition, we will discuss only
configurations involving catalysts. The membrane itself may be catalytically active. It may contain a catalyst, or be in
proximity to a catalyst. This last statement means that the membrane and the catalytic surface directly influence each
other. That is, the reactor and membrane are integrated into one unit. This excludes staged reactor/membrane
configurations where these are two separate operations. An additional limitation of this review is that it will focus on
recent developments in reactor type or configuration. A review of the many new developments in the field of membrane
materials and their properties is not within the scope of the present contribution. Membrane reactor materials will be
mentioned only as necessary for the understanding of new catalytic reactor types, or as the reasons for new
developments.

A comprehensive review of this ever-widening field is all but impossible within a single journal article, even
with the above restrictions. Catalytic membrane reactors has been an often-reviewed field in the last fifteen years,
culminating with the monograph by Sanchez and Tsotsis (2002), which is, at present, the definitive reference for catalytic
membrane reactors. In this review the focus is on those innovative uses of catalytic membrane reactors that either are
very new or have not previously been covered. The list of references in the present review has therefore been restricted
to the years 1999 to present. It is intended that it will represent a fairly complete bibliography for catalytic inorganic
membrane reactors for that period, under the above restrictions of scope.

2. CATALYTIC INORGANIC MEMBRANE REACTORS - BACKGROUND


The present section provides a brief overview of well-established uses of catalytic inorganic membrane reactors. For the
sake of brevity, historical developments and original papers will not be cited and for these the reader is referred to the
sources mentioned in the previous section. The principles to be discussed in this section will be illustrated with recent
references only.

Many useful recent reviews have been published, which focus on particular aspects of catalytic membrane
reactors. The review of Coronas and Santamaria (1999) presents a clear discussion of porous membrane reactors, with

Produced by The Berkeley Electronic Press, 2003


2 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

an emphasis on those that employ microporous membranes, such as zeolites. Julbe et al. (2001b) provide an interesting
view of porous ceramic membrane reactors, while Stoukides (2000) gives an in-depth treatment of solid-oxide membrane
reactors. Bouwmeester (2003) focuses on the use of mixed oxygen-ionic and electronic conducting dense membranes
for syngas production and other methane conversion reactions. Gryaznov (1999) gives a history of Russian work in
hydrogenation and dehydrogenation using palladium membrane reactors. For a discussion of mathematical modeling
principles for catalytic membrane reactors, the reader may consult Hoff et al. (2003).

2.1 Types of Membrane and Reactor Configurations


The materials of membrane construction can be classified as either dense or porous. Dense materials include palladium
membranes that are semipermeable to hydrogen, and solid oxide electrolyte dense membranes such as modified zirconias
and perovskites, which have reasonably high oxygen permeation rates at high temperatures. Porous inorganic membranes
can be divided into macroporous (dp > 50 nm), mesoporous (50 > dp > 2 nm) and microporous (dp < 2 nm). Macroporous
materials, such as "-alumina membranes, provide no separative function, but may be used to support layers of smaller
pore size to form composite membranes, or in applications where a well-controlled reactive interface is required.
Mesoporous materials for membranes have generally had pore sizes in the 4 - 5 nm range, so that permeation is governed
by Knudsen diffusion. Typical materials are Vycor glass, and composite membranes of (-alumina supported on
successively larger-pore layers of "-alumina support. Microporous membranes offer the potential for molecular sieving
effects, with very high separation factors, and materials such as carbon molecular sieves, porous silicas and zeolites have
been studied. The most active areas of development for membrane materials are currently synthesis of supported thin
films such as supported Pd films on porous aluminas or on porous stainless steel, and supported zeolite films.

Classification of membrane reactors with catalysts can be based on the location of the catalyst. Table 1 gives
an updated list of the acronyms which will be used in this paper.

Table 1. Explanation of acronyms used throughout the review.

ACRONYM EXPLANATION

CMR Catalytic Membrane Reactor

PBMR (IMR, OFIMR) Packed Bed Membrane Reactor (Inert


Membrane Reactor, Outward Flow Inert
Membrane Reactor)

PBCMR Packed Bed Catalytic Membrane Reactor

PVMR Pervaporation Membrane Reactor

FBMR Fluidized Bed Membrane Reactor

FBCMR Fluidized Bed Catalytic Membrane Reactor

CNMR Catalytic Non-permselective Membrane Reactor

ZMR Zeolite Membrane Reactor

PFR (PBR, FBR) Plug Flow Reactor (Packed Bed Reactor, Fixed
Bed Reactor)

The catalyst may be deposited within the membrane, including the case where the membrane itself is
intrinsically catalytically active (CMR), catalyst particles may be packed inside or outside a membrane tube (PBMR)
or catalyst may be both in the membrane pores and packed as particles in or around the membrane (PBCMR). Similar
terminology is used if the membrane is used with a fluidized bed, in which case FBMR and FBCMR are used for the
two possible configurations.

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 3

The PBMR acts to change the fluid phase composition in contact with the catalyst particles, either by removing
intermediate or product species, or by gradually supplying reactant species. The CMR provides an alternative way to
contact the reactants and the catalyst. Microporous membranes especially, have a high surface area per gram, and if a
catalyst can be well-dispersed within the pores they can be highly effective. The membrane geometry allows products
to leave the catalytic site without having to diffuse against reactants, and possibly further react. In addition, the
membrane provides effective contact since all molecules can be forced to pass through the membrane, as opposed to less
efficient contacting in a packed bed. Drawbacks include operating in a diffusion-controlled regime, and difficulties in
obtaining sufficient catalyst loading in the membrane.

The following three sections present further details on the three most well-established areas of application of
catalytic membrane reactors: product removal, reactant feed and control of reactant contact. Indeed, Dalmon has
categorized membrane reactors as extractors, distributors and contactors, and this terminology is also becoming more
widely used (Julbe et al., 2001b).

2.2 Membrane Reactors for Preferential Removal of a Species


The most-studied type of membrane reactor is illustrated schematically in Figure 1 (Abashar et al. 2002), the product
removal or “extractor” type.

Figure 1. Schematic representation of a co-current product removal PBMR. (Reprinted


from Appl. Catal. A, 236, 35, Copyright (2002), with permission from Elsevier)

The membrane may be porous or dense, and is most frequently in a tube-and-shell configuration. All three
types, PBMR, CMR and PBCMR have been used. The reactant gases may be fed either to the tube or to the shell side
(the reaction side). The other side (the permeate side) either is at a lower pressure or has an inert, sweep gas flowing
through.

2.2.1 Principles: the function of the membrane is to selectively remove a product from the reaction mixture. If the
reaction is equilibrium-limited, the decreased activity of the species being removed will permit further conversion to
occur, beyond that which would be possible if no species were removed. This is referred to as "equilibrium shifting" or
sometimes "exceeding equilibrium." The equilibrium being referenced in these expressions is the hypothetical
equilibrium of the closed system formed by a fixed mass of the reaction side feed only. It is not the equilibrium of the
actual membrane reactor system, which would include both the feeds to the reaction and permeate sides.

In a conventional isothermal plug-flow reactor (PFR) two important rates govern its performance - the rate of
reaction and the rate of reactant feed per catalyst volume to the reactor. The ratio of these gives the Damköhler number,
Da = (reactor volume)(maximum reaction rate per volume)/(inlet flow rate), which also involves reactor tube dimensions.
The membrane reactor brings in at least one additional rate, the permeation rate of the fastest gas, which can be
represented by the Damköhler-Peclet product, DaPe = (maximum reaction rate per volume)/(maximum permeation rate
per volume). For proper performance of a membrane reactor, it is important that these three rates are correctly balanced.

Produced by The Berkeley Electronic Press, 2003


4 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

If the permeation rate is too low, the membrane has little effect and the reactor behaves like a PFR with the reactant gas
feed. If the permeation rate is too high, the shell and tube sides will equilibrate rapidly. The activity of the catalyst is also
important, as reflected in the value of Da. If the reaction rate is slow, equilibrium will not be approached, and the
removal of a product by a membrane will not affect the ultimate yield. To obtain a driving force for the removal of the
product from the reaction zone, a lower partial pressure must exist on the permeate side than on the reaction side. This
can be achieved by either a difference in total pressure, or by diluting the permeate side with enough inert to lower the
mole fraction of the permeating species.

2.2.2 Applications: although there are still a few publications on reactions that were favorites of early studies, such as
decomposition of H2S (Chan et al., 2000; Fan et al., 2000), cyclohexane dehydrogenation (Itoh et al., 2003) and ethane
dehydrogenation (L. Wang et al., 2003), interest in these has generally declined. An extended discussion of both
hydrogenation and dehydrogenation reactions has been contributed by Dittmeyer et al. (2001). The industrial importance
of the dehydrogenation of ethylbenzene to styrene has ensured continued interest in this reaction (Dittmeyer et al., 1999;
Wu, 1999; Elnashaie et al., 2001; She et al, 2001).

More recent work on product removal has addressed dehydrogenation of higher alkanes, such as propane
dehydrogenation to propylene (Weyten et al., 2000; Schäfer et al., 2001; Hou and Hughes, 2002; Schäfer et al., 2003).
The dehydrogenation of alcohols is also an area that is becoming more popular (Liu et al., 1999; Schramm and Seidel-
Morgenstern, 1999; D. Xue et al., 2001; Keuler and Lorenzen, 2002a; 2002b; Liu et al, 2003).

An area of much current interest is the production of hydrogen and utilization of methane via methane steam
reforming (Lin and Rei, 2001; Lin et al., 2003) and the water-gas shift reaction (Criscuoli et al., 2000; Tosti et al., 2003;
Basile et al., 2003b). Fluidized beds are also being used for methane-based reactions, to solve problems of thermal
control (Jarosch and de Lasa, 1999; Grace et al., 2001; Prasad and Elnashaie, 2002; Chen et al., 2003). The conversion
of methane to C2 and higher hydrocarbons by hydrogen removal through a proton-conducting membrane has been
presented by Li et al. (2002), and by hydrogen removal through a palladium membrane by Ishihara et al. (2002). Non-
oxidative methane aromatization has been studied by Rival et al. (2001), Iliuta et al. (2002, 2003) and Basile et al.
(2003a). The dry reforming of methane to give a 1:1 mixture of H2 and CO has been studied using several different types
of membrane to remove the hydrogen.
CH4 + CO2 6 2 CO + 2 H2
Prabhu and Oyama (2000) used a modified Vycor glass membrane which was stable and provided high permselectivity
via an activated H2 transport mechanism; Ferreira-Aparicio et al. (2002a) used a modified mesoporous mullite membrane,
while Raybold and Huff (2002) used the more usual Pd membrane. They found that at 550 °C their conversion and
selectivity rose from 26% and 73% respectively, to 58% and 99%, due to the hydrogen removal. Paturzo et al. (2003)
used a CMR with Ru deposited in the pores of an alumina membrane tube.

Other hydrogen product removal reactions include the conversion of methylcyclohexane to toluene by Ferreira-
Aparicio et al. (2002b). Li and Zhong (2003) carried out dimethyl carbonate synthesis from CO2 and methanol with water
removal through Si-based membranes. Most product removal reactions extract H2 or H2O from the reaction, but in the
hydrodechlorination of dichloroethane (Chang et al., 1999) the species being preferentially removed was HCl.

2.3 Membrane Reactors for Distributed Addition of a Reactant


In the second main type of membrane reactor, a membrane is used to feed one reactant along the length of the reactor.
This “distributor” configuration is illustrated in Figure 2 in two orientations, inward flow (IMR) and outward flow
(OFIMR), for the partial oxidation of butane to maleic anhydride (Pedernera et al., 2000).

2.3.1 Principles: the idea of a distributed reactant feed is applied to systems with two competing reactions. A typical
example is the partial oxidation of a hydrocarbon. The general reactions have the form
A + L1 B ÷ P (1)
A + L2 B ÷ S (2)
where P is the desired main product, S is the undesired side-product, and A and B are reactants. If the kinetics are written
in power-law form, and B is the distributed reactant, then improved selectivity to P will result if the reaction order in B
of the desired reaction is less than the reaction order in B of the undesired reaction. This principle is complicated by the
change in residence time behavior of the reactants as one of them is fed gradually (Klose et al., 2003).

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 5

Figure 2. Two types of distributed reactant feed membrane reactor. (AIChE J., 2000, 46,
2489. Reproduced with permission. Copyright © 2000 AIChE. All rights reserved)

Porous membranes used for reactant feed are typically mesoporous or macroporous, leading to problems of
control of the rate of addition and distribution of reactant B, and backdiffusion of reactant A. Microporous membranes
often have too low a permeability. Dense membranes are sometimes used to feed hydrogen for hydrogenation reactions.
Examples are Gryazenov et al. (2001) for the hydrogenations of cyclopentadiene and 1,3-pentadiene, and Vincent and
Gonzalez (2002) for acetylene hydrogenation.

Most dense membrane use has been to feed oxygen, using solid oxide electrolytes as membranes (Stoukides,
2000). Solid oxide electrolytes with both ionic and electronic conductivities are very attractive as membrane materials,
as no external circuit is needed. Most work has focused on perovskite materials with aliovalent substitutions. A typical
perovskite has formula ABO3 (e.g. LaAlO3); substitution on the A and B sites of lower valence cations induces anion
disorder and increases both conductivities.

2.3.2 Applications: the dense solid oxide membrane materials require high temperatures (> 700 °C) for good O2 fluxes.
This has led to their being restricted to methane activation, and in particular they have been much-studied for the
oxidative coupling of methane (OCM), and for methane partial oxidation to syngas. Porous membranes do not allow the
separation of O2 from air, but can operate over a wider temperature range. Consequently, a wider range of reactions have
been studied using porous reactant feed membranes than dense ones.

The OCM reaction has been the subject of many hundreds of papers on catalysis and reactor design, owing to
the economic implications that a successful process for turning natural gas into higher hydrocarbons would have, if high
yield (above 35%) could be achieved. Despite all the research, this has not yet been possible. Early work in membrane
reactors was optimistic that higher yields could be obtained by lowering the oxygen partial pressures and suppressing
the total oxidation reactions to COx. Some simulation studies showed that if a dense oxide material could be found that
was permeable to oxygen, producing a surface species that was highly selective to C2 species rather than COx species,
then high yields would be possible. Kao et al. (2003) simulated a porous ceramic membrane reactor under optimal
operation, and obtained a maximal yield of 30%. Recent experimental work using membrane reactors (Assabumrungrat
et al., 2000; Shao et al., 2001; Akin et al., 2001) has not resulted in significant improvements. Lu et al. (2000b, 2000c)
obtained yields of only 16% with dense oxide membranes, whereas with porous membranes the best was 27% (Lu et al.,
2000a). Akin and Lin (2002a) claim high yields using dense ceramic membrane reactors, however they obtained 35%
yield only for reactant feeds of 1% methane, i.e. a highly dilute feed. In fact, they presented a figure giving their results
as yield versus methane partial pressure, which shows that yield decreases with increasing PCH4, and the results of Lu
et al. (2000a), at about 4% methane feed, fall right on top of those presented by Akin and Lin (2002a). These results seem

Produced by The Berkeley Electronic Press, 2003


6 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

to indicate that little progress has been made towards a commercial membrane-based OCM process, and the amount of
new research into this area will probably decline.

The distributed-reactant membrane reactor has also been studied for several oxidative dehydrogenation of
alkanes reactions such as ethane to ethylene (Wang et al., 2002), propane to propylene in an IMR (Ramos et al., 2000;
Hou et al., 2001; Bottino et al., 2002) and in a CMR (Alfonso et al., 1999; 2000) and butane to butene (Tellez et al.,
1999; Raybold and Huff, 2000; Ge et al., 2001; Alfonso et al., 2002; Pedernera et al., 2002), with the two most recent
studies using a CMR. A relatively newer reaction class is oxidative dehydrogenation of alcohols, in particular the
conversion of methanol to formaldehyde (Brinkmann et al., 2001; Diakov et al., 2001; 2002; Diakov and Varma, 2003).
The results for these reactions continue to show promise, with higher yields for the membrane reactor when compared
with a fixed bed, over certain ranges of the operating parameters.

Partial oxidations are being more frequently studied. The selective oxidation of ethane to ethylene was studied
by Akin and Lin (2002b), and the epoxidation of ethylene to ethylene oxide was the subject of extensive study by the
group of Varma (Lafarga and Varma, 2000; Al Juaied et al., 2001). An interesting aspect of their work, was that the
kinetics of ethylene epoxidation did not clearly indicate which reactant species should be distributed. Ethylene
epoxidation over silver is a set of complicated reactions, with inhibition by product CO2 and water, the frequent presence
of additives to improve selectivity in the industrial case, and even the competitive adsorption of the product ethylene
oxide itself, all leading to Langmuir-Hinshelwood-Hougen-Watson kinetics. Lafarga and Varma (2000) showed that
better conversion and selectivity were obtained if ethylene was distributed, rather than oxygen, since high oxygen to
ethylene ratios favored ethylene oxide formation. When oxygen was distributed, worse results were obtained than with
a PFR. It is also possible that the changed residence times may have had an effect, by increasing the contact time of both
ethylene and ethylene oxide.

The partial oxidation of butane to maleic anhydride has been studied extensively by two research groups (Xue
and Ross, 2000; E. Xue et al., 2001; Mallada et al., 2000a; 2000b; Pedernera et al., 2000). Several variations in how best
to conduct the reaction have been studied, including the use of split flow (Mallada et al., 2000a), different reactor flow
configurations (Pedernera et al., 2000; Mallada et al., 2000b), and the addition of CO2 to the reactant feed (E. Xue et al.,
2001; Mallada et al., 2002). The use of an externally-fluidized bed can also improve the thermal control of the reactor
(Alonso et al., 2001).

Aside from these main areas of application of the distributed feed membrane reactor concept, Langhendries et
al. (1999) modeled a liquid-phase hydrocarbon oxidation, in which the oxidising agent was aqueous t-butyl-
hydroperoxide, distributed through a microporous membrane, thus keeping the organic and aqueous phases separate, and
eliminating the need for a solvent. Ren et al. (2003) have reported promising results on the use of solid oxide membranes
for in situ O2 enrichment of air in oxygen-enhanced combustion.

In an attempt to extend the use of dense oxide membranes to a wider range of reactions, Cai et al. (2002) used
a doped BiVOx (BIMEVOX) membrane reactor for the lower temperature (300 - 600 °C) conversion of alkanes. These
materials show promise of good oxygen permeation rates, at moderate temperatures.

2.4 Membrane Reactors for Control of Reactant Contact


The way in which reactants contact each other may be controlled by using the geometry of the membrane in a CMR.
When a conventional packed bed with catalyst particles is employed, the reactants must necessarily contact each other
as they diffuse into the pellet. Similarly, the reactants and products must contact each other as the products diffuse out
of the pellet, and the products will remain in contact with active catalyst until they diffuse out of the pellet. Then there
is always the possibility that the products may contact active catalyst in another pellet downstream. The two-sided
geometry of a membrane presents an opportunity to contact reactants differently. Two different reactants can be fed from
opposite sides, creating an interface within the membrane. Products may be made to diffuse out of the membrane
preferentially to one side, where they do not further contact the reactant fed on the other side. Two major application
areas of this idea have emerged. In the first, a gas and a liquid phase are on opposite sides of the membrane; in the
second, two reactants in the same phase may be kept separate before reaching the catalytic site inside the membrane. This
type of membrane reactor has sometimes been termed an “interfacial contactor” (Miachon et al., 2003).

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 7

2.4.1 Multiphase Membrane Reactor: the principle here is to improve the contact between volatile and nonvolatile
reactants. This is well illustrated by the work of Espro et al. (2001) who studied the selective oxidation of light alkanes
on a Nafion-based catalytic membrane, mediated by a Fe2+ - H2O2 Fenton system. They observed that the liquid phase
reaction had problems of slow kinetics, separation of the catalyst from the reaction media, and separation and recovery
of the reaction products. With a three-phase membrane system the catalyst was immobilized on the membrane, gas and
liquid were separated by the membrane, and the products were easily condensed out of the gas phase.

Similar ideas lie behind the use of such membranes for hydrogenation by Daub et al. (2001) and Centi et al.
(2003), who observed that hydrogen has poor solubility in organic liquids. The membrane forces the gas-liquid interface
to be in contact with the catalyst, at lower pressure, as the hydrogen does not have to disolve into the liquid film then
diffuse through it. Both groups studied the reduction of nitrates in water, and similar research lines were pursued by
Raeder et al. (2003) and Miachon et al. (2003) for wet air oxidation of toxic compounds in liquids. Slightly different
concerns motivated Lapkin et al. (2002) in their study of the homogeneous catalytic hydration of propane by phosphoric
acid. Here, the conventional supported liquid phase (SLP) catalysis allows leaching of the acid, which must be prevented.
Use of the membrane allowed the acid to contact the gas phase through the pores, while keeping it in a closed container.
Finally, as demonstrated by Assabumrungrat et al. (2003), pervaporation is inherently a three-phase membrane operation
as product water is removed as vapor from one side, while the liquid phase remains on the other.

A somewhat different approach was used by Tan and Li (2000) in their experimental study of dissolved oxygen
(DO) removal to provide ultrapure water for the semiconductor industry. They used a highly hydrophobic hollow fiber
membrane, that permeated gas easily, but not water. They fed the water with DO over a catalyst on one side of the
membrane. The DO could diffuse out of the water through the membrane to the permeate side, where it was flushed away
by hydrogen sweep gas (physical stripping). In addition, the H2 could pass through the membrane and reduce the DO
over the catalyst to further increase conversion. This approach combined aspects of product removal, distributed reactant
feed and control of reactant contact.

2.4.2 Non-permselective CMR: in this type of reactor (sometimes called a CNMR), the membrane plays no
permselective role, but simply provides a location for the establishment of a reaction zone. One reactant (B) is fed on
the tube side of the membrane, and the other reactant (A) is fed on the shell side. The partial pressure gradients set up
cause them to permeate toward each other inside the membrane, where they react. Provided the reaction rate is higher
than the transport rate within the membrane, a reaction plane will be set up within the membrane. It is then possible to
avoid reactant slip, and direct products to one side of the membrane.

Neomagus et al. (2000) studied methane combustion in a Pt-activated CNMR. They noted the absence of slip
of the methane, the ease of control of the reaction as they could vary flow rate, composition and pressure of each side
of the membrane independently. The transport-controlled regime also hampered thermal runaway, and pressure drop
across the membrane could be manipulated to increase conversion. Similar results were found for propane combustion
by Saracco and Specchia (2000), who also warned that steady-state multiplicity was posssible in the transition regime
from kinetic to transport control.

3. ASSESSMENT OF CATALYTIC INORGANIC MEMBRANE REACTORS


A continuing paradox, in the field of catalytic inorganic membrane reactors, is the growing number of research
publications, while there are very few reported industrial installations. The growth in publications was highlighted by
Saracco et al. (1999), who presented a graph showing the results of a Chemical Abstracts database search from 1965 to
1997, using the keywords CATALYTIC and MEMBRANE and REACTOR. They found less than ten publications per
year prior to 1985. This was followed, with some fluctuations, by an almost linear increase to 120 publications per year
in 1997. A search for recent years by the present author using the CAPLUS + MEDLINE database with the same
keywords yielded
1997: 90 1998: 74 1999: 82
2000: 110 2001: 144 2002: 127
Despite the different results from the databases in 1997, the overall trend is clear. The strong upwards trend in the
number of publications continued in recent years, with some ups and downs. The fluctuations in the number of
publications in this relatively small field may be related to the timing of proceedings volumes of major conferences, but

Produced by The Berkeley Electronic Press, 2003


8 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

do not detract from the overall trend of increasing interest in membrane reactors. It should be noted that these search
results include polymeric, biological membrane reactors, etc.

The resolution to the paradox may be that the concepts behind catalytic membrane reactors are intellectually
appealing, yet technological difficulties and unfavorable economics have so far prevented their realization. Several
research groups have attempted to summarize the problems (both practical and commercial) faced by catalytic membrane
reactor technology. Others have conducted detailed economic analyses of specific systems. Still others attempt to provide
reactor design guidelines based on engineering principles and illustrative analyses.

3.1 Practical and Commercial Problems


Tennison (2000) considered the question of why we do not see any commercial implementations of membrane reactors.
His article considered porous ceramic membranes only, and was focused on hydrogen removal. Big problems for
membrane reactors include complexity, scale-up and economics. A “killer app” is needed for the successful development
of new technology. It should have low costs, low risks, an available market and immediate benefits. If the membrane
reactor is mission-critical to the process, it will be a problem. We will want many similar applications, to justify the cost
of membrane development and manufacture. For example, the use of palladium membranes for hydrogen removal is not
justified economically, as several studies have shown, except maybe to provide clean hydrogen for fuel cells. Oxygen-
permeable membranes are considered high-cost and high-risk; of the current research areas, syngas may be commercially
viable. Using CMRs or PBMRs as add-ons is attractive, but this can only be done for product removal, not reactant
distribution. Tennison suggests that too much research is focused on low value-added products, and that high value-
added processes such as pharmaceutical oxidation processes may be more promising. Also, there are too many problems
with big plants, and some attention should be paid to microdistributed processes for membrane reactor applications.

An updated review paper focusing on the hurdles to inorganic membrane use was presented by Saracco et al.
(1999). They emphasized the need for defect-free, high-selectivity, high-permeation membranes. They also noted the
difficulty in reproducing laboratory-scale results on a larger scale suitable for commercialization. This may be especially
true for microporous membranes. Ceramic and metallic membrane tubes have to be housed within a reactor assembly
with feed/product lines and other reactor peripherals. Progress is still needed in developing cheap, high-temperature
sealing systems for membrane reactors. Related to this is the need to develop technologies for heat supply/removal and
temperature-control for large-scale modules. Saracco et al. (1999) also point out the need to develop new membrane
catalysts that are less sensitive to coking and poisoning. An increasing awareness of the interaction of the catalyst with
the changed environment caused by the use of a membrane is an area of recent research discussed further below, in
section 4.

Saracco et al. (1999) also point out the need for continued work on the scientific and engineering underpinnings
of the research area. They suggest the development of complex models for large-scale reactor modules. They further call
for design criteria for membrane reactor design: optimal size, flow patterns, stages, recycle, and intermediate feeds.

3.2 Reactor Design Guidelines


A different analysis of effectiveness for catalytic membrane reactors is brought up by the question of whether they
provide any real advantage over more conventional reactors. Ultimately, economics must be the deciding factor,
although very few studies have included an economic evaluation. Two exceptions are the studies of Criscuoli et al.
(2001), who investigated the economics of the water-gas shift reaction, and Onstot et al. (2001), who investigated the
dry reforming of methane for hydrogen production. Onstot et al. (2001) evaluated various plant designs with two
different types of membrane, but they considered only operating energy costs as the costs of commercial membranes
were prohibitively expensive at the current state of development. Their work pointed towards the desirability of
developing microporous ceramic membranes, which would lead to more energy-efficient plants.

Criscuoli et al. (2001) made an economic analysis of palladium membrane reactors for the water gas shift
reaction, with the aim of a fixed pure hydrogen recovery. They evaluated both capital and operating costs for different
inlet compositions and for a membrane reactor alone and for a membrane reactor preceded by hydrogen recovery, using
an industrial synthesis gas feed. Their results are shown in Figure 3.

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 9

Figure 3. Effect of palladium thickness on capital costs and operating costs for membrane reactors.
(Reprinted from J. Membrane Sci., 181, 21, Copyright (2001), with permission from Elsevier)

Criscuoli et al. (2001) concluded that current membrane technology for hydrogen removal was not yet
economically feasible, mainly due to the high cost of the membrane. This could only be justified if higher productivity
per unit area of membrane could be realised. From the graphs above, they suggested that a selective layer of 20 :m
thickness should be the target of future research.

Most of the comparisons are based on conversion, yield or a similar feature of reactor performance, rather than
a complete economic analysis. These have usually been made separately for product removal and distributed reactant
feed membrane reactors.

3.2.1 Product removal: a simple idea for the screening of membrane reactors was put forward by Boudart (Cattech Vol.
2, p. 94, 1997), based on an observation by Weisz (CHEMTECH p. 424, July 1992) that most commercial reactors
operate in a “window of reality” with a space-time yield (STY) in the range 1 to 10 mol m-3 s-1. This ensures a balance
between reactor size and heat and mass transfer limitations. Boudart suggested that the corresponding quantity for
membrane reactors is the areal time yield (ATY) in mol m-2 s-1, which is just the permeation flux of the membrane.
Dividing the ATY by the STY gives the area to volume ratio (A/V) that would be needed for a feasible catalytic
membrane reactor. This concept was well
illustrated for several common membrane
types by van de Graaf et al. (1999a).

The range of ATY available from


the membrane is mapped to the feasible STY
range, giving a range of A/V which any
feasible membrane reactor module needs to
provide, as shown in Figure 4. van de Graaf
et al. (1999a) concluded from this analysis
that for some systems application of
membrane reactors was already feasible,
while for others such as zeolite membrane
reactors, a moderate increase in permeability
was required that might be achieved by
reducing membrane thickness.

More detailed analyses of specific Figure 4. Comparison of STY of catalytic reactors to ATY of
reactions or reaction classes have been some inorganic membranes. (Reprinted from Appl. Catal. A,
considered. Chmielewski et al. (1999) set 178, 225, Copyright (1999), with permission from Elsevier)

Produced by The Berkeley Electronic Press, 2003


10 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

conversion targets in plug flow membrane reactors for isothermal reversible reactions, and illustrated their analysis by
showing conversion improvements in the n-butane to isobutane isomerization. Moon and Park (2000) carried out a
parametric analysis of product removal membrane reactors in terms of Da and Pe, using cyclohexane dehydrogenation
as a model reaction. They plotted conversion contours in the Da-Pe diagram, and identified regions where conversion
was controlled by different mechanisms, such as selective permeation. This analysis allowed them to investigate the
effects of changing membrane permeability, selectivity, and sweep gas flow rate, and to see under what conditions these
changes would have most effect.

Sang et al. (1999) conducted a preliminary assessment of membrane reactors to improve selectivity of
methylamine synthesis from ammonia and methanol. Their intent was to use porous ceramic membranes to selectively
remove monomethylamine and dimethylamine from the product mixture, at the expense of the thermodynamically
favored trimethylamine. They found that loss of the smaller reactant molecules through the membrane presented a severe
problem, as the currently available membranes were not permselective enough. Additionally, they discovered that the
reaction proceeds in two stages. The first stage lasted until all methanol and intermediate dimethylether was consumed.
The product mixture at that point had often erroneously been taken as the equilibrium composition. There then followed
a slow equilibration step to yield the true equilibrium mixture. Their analysis of membrane reactors showed that if more
active catalysts were available, especially for the second kinetic stage, then the requirements on membrane performance
would be easier, and even existing membrane materials might lead to attractive membrane reactor configurations.

Along the same lines of needing improved


catalyst performance, the same research group (Ma
and Lund, 2003) performed simulations of high-
temperature water-gas shift membrane reactors, with
a two-step microkinetic model, and compared the
yields to be obtained from existing palladium
membrane reactors, similar reactors with perfect
membranes and similar reactors with perfect catalysts.
Their results are summarized in Figure 5. The
simulations suggested that the limiting factor to better
performance was the rate of reaction and not the rate
of permeation through the membrane. The best
improvement was judged to be the development of a
catalyst that would not be inhibited by CO2, rather
than the development of better membrane materials.
Both of these studies emphasize the importance of
developing kinetic models that properly reproduce the
behavior of the catalyst and the reaction over the range
of conditions expected in the membrane reactor.

3.2.2 Reactant feed: theoretical work leads us to


expect that when comparing reactant feed PBMR with Figure 5. Comparison of WGS simulated reactor yield to
the co-fed PFR for partial oxidation or similar idealized membrane reactors. (Reprinted with
reactions, the PBMR can give higher yields, but may permission from Ind. Eng. Chem. Res. 2003, 42, 711.
require a much larger reactor. When equal size Copyright 2003 American Chemical Society)
reactors are compared, the decrease in conversion due
to the lower reaction rate in the PBMR may offset the
increase in selectivity due to the favorable kinetics (Dixon, 2001), however the increase in residence time in the PBMR
due to the lower average molar flow may act to offset the conversion decrease (Diakov et al., 2002).

Dixon (2001) made a comparative study of yield to intermediate products obtained from the distributed-feed
membrane reactor and the conventional cooled-tube plug flow reactor. The non-isothermal, non-adiabatic one-
dimensional pseudohomogeneous reactor model was used. The analysis was presented in terms of the usual
dimensionless groups (Damköhler number Da, heat of reaction parameter B, Peclet number Pe, etc.).

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 11

The procedure for determining


when the membrane reactor performance
was better is illustrated in Figure 6. This is a
plot of the ratio of PBMR yield to PFR yield,
against the ratio of PBMR maximum
temperature rise to PFR maximum
temperature rise. The top left quadrant is the
region in which the membrane reactor
satisfies the two performance criteria, that
the membrane reactor should give improved
yield with no higher maximum temperature.
For fixed values of B and Da the membrane
reactor model was solved for a series of
decreasing values of DaPe, or increasing
values of the parameter 1/Pe, which is
proportional to the permeability of the
distributed reactant. This resulted in a
trajectory in the membrane performance
plane, that started at low yield and
temperature rise ratios, and in general both
increased as the amount of reactant
increased. Three types of trajectory are
illustrated: for B=14, Da=1 the PBMR
temperature rise exceeded that of the PFR
while the yield was still less, so no range of Figure 6. Comparison of PBMR to PFR for yield and hot spot as
membrane permeability gave rise to an permeation rate is increased. (Reprinted from Catal. Today, 67,
advantage; for B=14, Da=3 the PBMR to 189, Copyright (2001), with permission from Elsevier)
PFR yield ratio exceeded unity for a range of
permeabilities given by the interval [DaPeH, DaPeL] as shown, before falling below unity again; and for B=20, Da=6 the
PBMR to PFR yield ratio again exceeded unity for a range of permeabilities given by a different interval [DaPeH, DaPeL],
after which the PBMR to PFR temperature rise ratio exceeded unity.

It was concluded that significantly better yields could be obtained by the membrane reactor under some
conditions. These were generally at higher reaction rates, and for more thermally challenging conditions, such as higher
heat of reaction or lower heat removal rate. The membrane reactor, however, used significantly larger amounts of the
distributed reactant. It was found that co-feeding some reactant as well as distributing it could improve yield as well as
decreasing the overall amount of reactant required, compared to a purely distributed feed.

3.3 A Place for Catalytic Membrane Reactors?


From Sections 3.1 and 3.2 above, the road to identifying economically and technically sound applications for catalytic
membrane reactors is clearly a somewhat rocky one. Early optimism has given way to a spirit of healthy scepticism.
Several authors have cited the same technical challenges consistently over the years: deposition of the catalyst in the
membrane, maintaining high permeability and permselectivity, thermal effects, structural stability of the membrane,
sealing etc. Materials development still drives new advances in membrane reactors and their suitability for various
applications.

In response to these difficulties, researchers are making efforts to find niche applications in product removal
and distributed reactant feed, where the use of a membrane brings a clear economic advantage. Heat effects, catalyst
behavior and transportation/reaction interactions are all being studied in greater detail, as will be further illustrated
below. Some areas of research are still strong, such as methane utilization for syngas production, and oxidative
dehydrogenation. There are reports of commercial application of pervaporation reactors and zeolite membrane reactors
for dehydration. It is clear that the ingenuity of researchers will continue to be employed to find new ways to use
membrane reactors. The following section summarizes most, if not all, of those from the last five years.

Produced by The Berkeley Electronic Press, 2003


12 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

4. RECENT RESEARCH IN CATALYTIC INORGANIC MEMBRANE REACTORS


The research areas that have been of recent interest can be categorized as being either continuations of themes begun
a few years ago, or as completely new trends. As a framework for organizing the large amount of recent work, and to
help identify new areas, we shall re-visit the areas of innovation identified previously (Dixon, 1999) and compare them
to present research.

4.1 Thermal Effects in Membrane Reactors


The challenge of managing heat effects in conventional reactors provides an opportunity for membrane reactors. Three
areas for new research that were identified before, are 1) to use distributed reactant feed to control the temperature rise
in exothermic reactions; 2) to manipulate intermediate species in complex reaction networks by control of temperature
and partial pressure by distributing the fluxes of energy and material along the reactor; 3) to use exothermic and
endothermic reactions on the permeate and retentate sides of the membrane to satisfy each others’ heat needs (thermal
coupling).

4.1.1 Progress on Earlier Areas of Innovation: the underlying idea of temperature control by reactant feed is that if
the reactant is distributed along the tube in a PBMR configuration, then for exothermic reactions such as partial
oxidations, the release of energy will be more gradual. This could lead to a lower hot-spot, and to a wider feasible range
of operation. Often, the motivation for the use of a distributed oxygen feed comes less from the prospect of a gain in
selectivity, but more from the opportunity for improvements in stable reactor operation (Varma and Diakov, 2002). A
gradual oxygen feed can allow a higher hydrocarbon inlet concentration, or even operation in the flammable region for
the nominal combined feed. This has often been the case for the partial oxidation of butane to maleic anhydride, where
conventional technology limits the inlet hydrocarbon mole fraction to approximately 1%.

Using a membrane reactor to control the fluxes of reactants, as well as conventional manipulation of heating
and cooling of the reactor tube was suggested to allow control of intermediates and reaction pathways by manipulating
temperature and species concentrations. However, there does not seem to have been any published work involving
membrane reactors yet from this initiative.

A different idea for approaching the problem of heat management in a product removal membrane reactor is
that of thermal coupling. The idea is to run an exothermic reaction on one side of the membrane, and an endothermic
reaction on the other. The reactant for one reaction is a permeate species of the other. If the rates of heat release can be
balanced, with the permeation rate and reaction rate, two problems could be solved. The energy needs of one reaction
could be largely met, at the same time the difficulties of efficiently using the often dilute permeate could be overcome.
Practical difficulties of matching temperatures and heat release/uptake seem to have limited the applications of this
concept, and most integration of two reactions on opposite sides of the membrane may be thought of as examples of
reaction coupling, discussed below. Thermal considerations appear to be secondary, except for one or two instances
where combustion of the permeated hydrogen can supply heat for high-temperature reactions such as methane activation
(Li et al., 2000).

4.1.2 Detailed Modeling of Membrane Reactors: a noticeable trend which may be associated with the thermal behavior
of membrane reactors is the growing number of publications which include nonisothermal models. Early membrane
reactor models were used mainly to interpret data from small laboratory experiments in furnaces, and were usually
isothermal. Nonisothermal treatments began as simple one-dimensional models of tube and shell membrane reactors.
Recent examples are the work of Dixon (2001) for exothermic reactions, Chan et al. (2000) for endothermic reactions,
and Madia et al. (1999) who focused on the methane steam reforming (MSR) reaction. These models allowed an
approach where the effects of the main process and reaction parameters could be investigated. One drawback is that they
bring in lumped parameters such as overall heat transfer coefficients, whose estimation by existing fixed bed transport
correlations is uncertain.

Follow-on work for the MSR was presented by Barbieri et al. (2001) and Marigliano et al. (2001) who did an
in-depth analysis of thermal effects in both PFR and PBFR reactors for this endothermic reaction. They used both a
thermodynamic equilibrium conversion model and a reactor tube model to obtain regions in the conversion-reaction
temperature (X-T) diagram that are reachable by PFR and PBMR. They established limiting conditions for adiabatic and

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 13

isothermal operation, and calculated operating regions for various conditions, such as inert flow rate. They further
analyzed the relative importance of the terms in the energy balance. The reaction energy was found to be most important
near the tube inlet, the terms representing heat exchange with the external oven, and heat exchange across the membrane,
became more important further down the tube, as reaction heat declined in magnitude.

The importance of two-dimensional modeling was emphasized by Brinkmann et al. (2001) and Pedernera et
al. (2002). Koukou et al. (2001) looked at non-ideal flow (dispersion) effects in a two-dimensional model of an adiabatic
full-scale membrane reactor for the water-gas shift (WGS) reaction. They solved a momentum balance equation for both
axial and radial components of velocity, with pressure drop given by the pointwise Ergun equation, and assuming a
pseudohomogeneous approach. They found that there were strong radial profiles of hydrogen, which reduced the
apparent hydrogen transport rates through the membrane, compared to those that would be calculated assuming plug-
flow, or one-dimensional flow. This is evident from the radial hydrogen partial pressure profiles in Figure 7, at different
axial positions. Similar gradients and trends were found for radial temperature profiles.

Figure 7. Radial profiles of hydrogen partial pressure on feed and


separation sides of membrane. (Reprinted from Chem. Eng. Journal,
83, 95, Copyright (2001), with permission from Elsevier)

Hou et al. (2001) used a two-dimensional model in concentration and temperature, and showed that the radial
concentration profile affected selectivity and yield in a reactor for the oxidative dehydrogenation of propane. Hara et
al. (1999) demonstrated that in a H2-removal Pd-membrane reactor, the presence of CO could prevent the hydrogen
permeation through the membrane, at lower temperatures. They attributed this to concentration polarization effects of
the H2, and also to CO hindrance by adsorption of CO on the membrane surface. It may be noted that they needed a two-
dimensional membrane tube model to explain these effects. Sousa et al. (2002) considered the appearance of sharp axial
gradients, and introduced numerical methods for the solution of their one-dimensional models based on wavelet analysis.
Nekhamkina et al. (2000) investigated the appearance of spatio-temporal patterns for reactant-distributed reactors,
showing complex behavior under certain conditions.

4.1.3 Example of Detailed Modeling: as an example of the necessity for detailed modeling, consider the two reactor
configurations shown in Figure 2, the IMR and the OFIMR reactors for the partial oxidation of butane to maleic
anhydride. An experimental study by Mallada et al. (2000b) noted that the yield obtained from the OFIMR was better,
as shown in Figure 8.

Produced by The Berkeley Electronic Press, 2003


14 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

Figure 8. Selectivity-conversion plot with yield contours, comparing


IMR (open symbols) and OFIMR (closed symbols) configurations.
(Reprinted with permission from Ind. Eng. Chem. Res. 2000, 39, 620.
Copyright 2000 American Chemical Society)

Mallada et al. (2000b) postulated that the radial temperature and concentration profiles were different in the
two reactors. A complete explanation was provided by Pedernera et al. (2000) who combined detailed kinetics with a
two-dimensional pseudohomogeneous nonisothermal reactor model. A separate kinetics study led to a dual-site Mars-van
Krevelen mechanism. The VPO catalyst was known to be sensitive to the reaction atmosphere, optimal selectivity and
activity was obtained only in a moderately high oxidation state. For the IMR, the more reduced catalyst (nonselective)
was at the reactor entrance and near the tube axis for the rest of the reactor. For the OFIMR, it was also at the reactor
entrance, but near the tube wall further downstream. Pedernera et al. (2000) calculated from their model the ratio of
active/selective to nonselective sites. For the IMR, the lowest value of this ratio was at the axis, for the OFIMR it was
at the wall. Then using the two-dimensional reactor model, they showed that the temperature profiles for both
configurations were at their maximum values at the tube axis, and were lowest at the tube wall. So, for the IMR, the
unfavorable conditions occurred at the reactor axis, where reaction rates were highest. For the OFIMR, the axis was the
favorable region, and was also where the highest reaction rates took place. This explained the observed better
performance of the OFIMR. This example serves to show that proper understanding of membrane reactor behavior
requires both a realistic kinetic model and detailed calculations of the species concentration and temperature profiles.

4.2 Synergy Between Reaction and Permeation


The problem of increasing membrane permeation rates while maintaining the permselectivity was mentioned above. This
section deals with research in which the reaction occurring on the permeate side of the membrane was used to increase
the permeation rate through the membrane, either by improving the driving force for permeation or by modifying the
membrane characteristics. There has been new work in both areas, but there appear to be no approaches radically
different from these two ideas.

4.2.1 Use of the Reaction to Improve Driving Force: One way to address the problem of low permeability through
a membrane is to conduct a reaction on the permeate side to decrease the partial pressure of the permeate.A typical
example is the work of Moustafa and Elnashaie (2000) who modeled the dehydrogenation of ethylbenzene to styrene
reaction on one side of a thin palladium membrane, and the hydrogenation of benzene to cyclohexane on the other side.
Their calculations suggested that higher styrene yields would be obtained with the reaction on the permeate side, rather
than a sweep gas. There was little or no opportunity for thermal coupling in this case since the endothermic

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 15

dehydrogenation reaction ran at higher temperature for most of the reactor length than the exothermic benzene
hydrogenation.

Related ideas were presented by Alfonso et al. (2001) who used a flow-through two-layer membrane (see
sections 4.3 and 4.5 below) to do both oxidative and non-oxidative dehydrogenation of butane, to achieve thermal and
chemical coupling of both processes. They found that it was crucial to completely consume the O2 in the oxidative
dehydrogenation layer. The coupled arrangement increased both conversion and selectivity compared to a single-layer
membrane. As mentioned briefly above, Li et al. (2002) worked on the non-oxidative conversion of methane, where the
energy for CH4 pyrolysis was supplied by combusting the permeated H2 with oxygen fed on the permeate side, which
also improved the driving force for hydrogen permeation.

4.2.2 Use of the Reaction to Modify the Membrane: the use of dense oxide membranes to feed O2 to partial oxidation
reactions is limited by the need for high temperatures to obtain reasonable permeation rates. The potential exists for
selectivity gains and control of heat effects. One application where dense membranes have a clear role to play is in
methane activation. In particular, the partial oxidation of methane to syngas
CH4 + ½ O2 6 CO + 2 H2
is carried out at high temperatures, usually above 750 °C. Air is not used directly in the reactor, as it is not economical
to separate the nitrogen from synthesis gas downstream. The oxygen separation plant for the feed is the dominant cost
of the process, and its elimination would be desirable. The slightly exothermic partial oxidation reaction is attractive
compared with the strongly endothermic steam reforming of methane.

Early research in this area focused on the use of perovskites and related materials and was summarized in the
previous review (Dixon, 1999). It was established by several groups, that these semi-permeable dense mixed-conducting
oxide materials could have oxygen fluxes of 1 scc/cm2.min under standard permeation testing. Under reacting conditions,
with a strongly reducing atmosphere on the permeate side, the material structure could change and the flux could increase
up to approximately 4 scc/cm2.min. The stability of the material is a crucial issue, and long run times under reacting
conditions are required to demonstrate that the membrane is useful. Recent work by two of the groups involved in the
original exploration of this area have improved performance. Maiya et al. (2000) obtained fluxes of 10 scc/cm2.min in
their non-perovskite material, while obtaining stable 70% conversion of a 80% methane feed. Sammels et al. (2000)
claim 10-12 scc/cm2.min for their Brown-Millerite membranes, with 90% conversion of an 80% methane feed and stable
operation for over one year.

A great deal of work from China has focused on trying to increase oxygen permeation rate and maintain stability

Figure 9. Long term performance of BSCFO membrane for partial oxidation of


methane to syngas. (Reprinted from Catal. Today, 67, 3, Copyright (2001), with
permission from Elsevier)

Produced by The Berkeley Electronic Press, 2003


16 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

of the membrane under highly reducing atmospheres. Jin et al. (2000a; 2000b) worked on a lanthanum cobaltite
perovskite tubular membrane of composition La0.6Sr0.4Co0.2Fe0.8O3-* (LSCFO-type). They investigated the effects of flow
rate, tube length and reaction temperature with a NiO/(-Al2O3 catalyst. Their oxygen fluxes were at the level of 4
scc/cm2.min under reacting conditions, their membrane lasted only 3-7 hours and they found problems with carbon
deposition for methane feeds higher than 6%. Similarly Gu et al. (2002), using an yttria-stabilized zirconia SCF oxide,
were also able to obtain 4.5 scc/cm2.min and 65-70% conversion of a 15% methane feed. Their membrane ran for more
than 220 hours. D. C. Zhu et al. (2003) used a La2NiO4 membrane to investigate the reaction mechanism. The best results
to date have been obtained by the group at Dalian (Shao et al., 2001; Dong et al., 2001; H. Wang et al., 2003), using a
Ba0.5Sr0.5Co0.8Fe0.2O3-* (BSCFO-type) disk membrane in a laboratory configuration, together with a LiLaNiO/(-Al2O3
catalyst. As shown in Figure 9, the permeation rate was approximately 11.5 scc/cm2.min, and they achieved 98.5%
conversion of a 50% methane feed. The reactor ran with no stability problems for upwards of 500 hours. The authors
determined that their material changed drastically under the reducing conditions of the reaction, but that it exhibited
almost complete phase reversibility, which accounted for its stable performance.

Recognizing the need in industry for membranes in tubular form, Ritchie et al. (2001) developed technology
for spray deposition of a thin dense film of LaSrFeGaO on a porous tube. Their reactor ran for 700 hours before problems
developed with the seals, and they claimed 92% conversion of a pure methane feed over an Rh catalyst. Their feed rate
was limited, however, due to the relatively low oxygen permeation rates in this material, only 0.5 scc/cm2.min,
considerably below the results reported above.

Finally, the group at the University of Calabria have taken a different approach to this reaction. The high
conversions cited above are all for reactors run at atmospheric conditions, however natural gas is available from the well
head at much higher pressures, at which the equilibrium conversion is too low to be economical. Rather than waste the
available pressure, they suggest the use of a Pd membrane to remove hydrogen and improve conversion (Basile et al.,
2001; Basile and Paturzo, 2001; Paturzo and Basile, 2002). Future directions may include the use of both types of
membrane to separate oxygen in the feed, and selectively remove product hydrogen.

Returning to the theme of modification of the membrane material by the reactive gas mixture, Julbe et al.
(2001a) have introduced a new concept, the chemical valve membrane. For a distributed feed reactor using an inert
membrane, they contend that the distributed reactant is fed in larger amounts downstream in the reactor, where it is less
needed. This idea is illustrated in Figure 10a, for the case of butane oxidation. The butane concentration decreases with
tube length, as the butane reacts. Additionally, the total pressure decreases along the tube as the fluid passes through the
catalyst packing. This pressure drop is higher than on the distributed feed side, which is often at constant pressure, or
at least is an open tube. Thus, the transmembrane pressure driving force increases down the tube, providing more oxygen
downstream than upstream. There is a mismatch between oxygen supply and need. Their novel idea is illustrated in
Figure 10d, and rests upon the identification of a membrane material that lets more oxygen through in a reducing
environment, and less in an oxidizing environment. Thus the permeability of the membrane is controlled by the red/ox

Figure 10. Schematic of O2 concentration profile in reactor for a) IMR; d) chemical valve membrane.
(Reprinted from Catal. Today, 67, 139, Copyright (2001), with permission from Elsevier)

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 17

properties of the gas, and is higher near the tube inlet, where the butane concentration is also higher, and is lower further
down the tube, where there is less butane.

Julbe et al. (2001a) found that vanadium oxide had such properties; it transformed reversibly from V2O5 porous
crystallites with a closed porous structure when oxidized, to V2O3 crystallites with an open porous structure when
reduced. Their experimental work demonstrated a reproducible synthesis method, and a strong increase in helium and
oxygen permeance with increasing butane/oxygen ratio in the reactor. Although this concept is attractive and the initial
test encouraging, they warn that a more realistic evaluation is needed for large scale reactors and under real reaction
conditions.

4.3 Contact Between Catalyst and Reactants


The ways in which the reactants and catalyst can be made to contact each other using a membrane geometry gave rise
to three areas that seemed promising (Dixon, 1999). For these areas, the membrane was porous and in the CMR
configuration, and either was intrinsically active or had a catalyst deposited within the pores. The three areas were
control of contact time, control of catalyst location, and control of reactant concentration at the catalyst surface. Some
aspects of these research directions continue to be developed, while others have lapsed and new directions have replaced
them.

4.3.1 Progress on Earlier Areas of Innovation: The first area was membranes operated in the cross-flow or flow-
through mode, in which all the reactant was forced to flow through the membrane, by feeding it to one side with a closed
exit. This configuration gave a uniform contact time, which could be tailored to a particular reaction by choice of the
membrane thickness and/or reactant flow rate. The pore size of the membrane controlled the diffusion regime. The idea
was to get efficient contact by forcing the gas to contact the catalyst in the Knudsen diffusion regime, where wall-
molecule collisions dominated over molecule-molecule collisions. This forced all the molecules into good contact with
the catalyst. This configuration is sometimes referred to as a “flow-through contactor” (Miachon et al., 2003).

A recent example is the work of Ilinitch et al. (2000) on the reduction of aqueous nitrates over Pd-Cu/Al2O3
catalyst. Conventional reactors operate with pronounced diffusion limitations in the catalyst particles, but the use of
finely dispersed catalysts is often not wanted due to downstream separation problems. The diffusion limitations were
minimized by the flow-through mode of the membrane reactor, and gave a multi-fold increase in conversion, as shown
in Figure 11. A disk membrane was used in two configurations, first with the outlet closed so that the reactor and
catalytic membrane operated in a “stirred-tank” mode (curve 1). The second configuration had the outlet open so that
all the product stream was forced to flow through the
membrane (curves 2 and 3, for different flow rates).
Conversion was clearly improved when the reactants
were forced through the membrane, rather than passing
into it by diffusion only. Similar results were obtained
by Reif and Dittmeyer (2003) for nitrate reduction and
the dechlorination of chloroform. They compared the
flow-through mode to the catalytic diffuser mode
described in section 2.4, and found higher catalyst
activity for flow-through. They warned, however, that
pore blocking could become a problem for the flow-
through mode. B. Zhu et al. (2003) used a flow-through
CMR for selective oxidation of propane to acrolein, to
get a six- to seven-fold increase in selectivity by
controlling contact time to prevent further oxidation of
products.

Similar results were obtained for propene


epoxidation by Kobayashi et al. (2003), who found a Figure 11. Performance of the catalytic membrane
strong effect on selectivity to the intermediate species in in the stirred tank and flow-through modes.
a sequential reaction. They attributed this to control of (Reprinted from Catal. Today, 56, 137, Copyright
the propylene oxide contact time by flow through the (2000), with permission from Elsevier)

Produced by The Berkeley Electronic Press, 2003


18 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

membrane. Maira et al. (2003) compared different configurations for the photocatalytic oxidation of VOCs. They
obtained generally better conversion with the flow-through mode, but the improvement was not so clear-cut. They
mentioned that they had problems with low activity of their catalyst, which may have clouded the picture. Tsuru et al.
(2003) found that photocatalysis of methanol decomposition increased in rate when reactants were forced to permeate
through a titanium oxide membrane. Miachon et al. (2003) also saw increased catalyst activity for wet air oxidation in
flow-through mode. Alfonso et al., (2001) used flow-through of a two-layer membrane as mentioned above, building
on prior work on contact mode by that group.

In the second area, controlled deposition in the membrane geometry could be used to place the catalyst in the
membrane. The optimal location or distribution of a catalyst in a membrane could strongly influence results. Some
theoretical work showed interesting early results, but there does not seem to have been any recent work on this idea. This
is perhaps due to the need for experimental work, and the difficulty of synthesizing membranes with the required catalyst
distributions. The object in the third area of innovation was to control the partial pressure of the reactants in the phase
in contact with the catalyst. An asymmetric CMR was used with one reactant fed from the support side, so that due to
diffusional resistance the concentration of that reactant was lowered at the catalyst site in the toplayer.There has been
little work directly addressing this idea, but it has developed into the control of the reactant species at the catalytic site,
as discussed below.

4.3.2 Control of Species in Contact with Catalyst: one of the ideas that has been around for some time is the notion
that by using membranes to feed reactant species, different activated forms of the species could be brought into contact
with the catalyst. This was especially attractive for the OCM reaction, given the problems with suppressing the complete
oxidation reactions. Tagawa et al. (1999) made the most recent attempt at this approach, using a YSZ membrane to
control the O2- lattice species contacting the methane, to preferentially react to C2 species. As with many other attempts,
the yield to C2 remained in single figures, well below the commercial need.

An interesting new development along these lines was published by Niwa et al. (2002) for the direct
hydroxylation of benzene to phenol. The traditional route is the three-step cumene process, giving an overall one-pass
yield of less than 5%, and which is very energy-consuming. Niwa et al. permeated hydrogen through a palladium thin
layer membrane supported on a porous alumina tube, into a hydrocarbon liquid though which an oxygen/helium mixture
was bubbled. Hydrogen was dissociated in permeating through the palladium membrane and the dissociated hydrogen
(H*) appearing on the opposite side of the membrane immediately reacted with the oxygen to give HOO@ and H2O2, the
latter then decomposed to HO@, atomic oxygen and water. An active oxygen species then attacked the benzene ring and
was responsible for the hydroxylation to phenol. The authors suggested that oxene might be the active species, but were
unable to definitely identify it. They obtained selectivities to phenol in the range 80-97% at benzene conversions of 2-
16%, and noted that the new process appears to be much simpler and safer than current practice.

4.3.3 Effects of Reaction Mixture Composition on Catalyst: there have been several publications from different fields
of membrane reactors that all appear to illustrate a common theme. All have identified changes in catalyst behavior and
deactivation patterns, that they attribute to the different reaction mixture that the catalyst is exposed to, as a result of a
membrane either adding a reactant or removing a product.

In a distributed-feed membrane reactor, with oxygen as the distributed species, the resulting decrease in the
partial pressure of oxygen in the reaction mixture can cause changes in the oxidation state of oxide catalysts. This was
hinted at before, in the discussion of the oxidation state of the catalyst when comparing IMR and OFIMR reactor
configurations (Mallada et al., 2000b; Pedernera et al., 2000). More recent work from the same group (Mallada et al.,
2002) found that it was beneficial to co-feed CO2 to a distributed-feed IMR for oxidation of butane to maleic anhydride.
They noted that a higher hydrocarbon/oxygen ratio in the gas mixture lowered selectivity to maleic anhydride due to a
less oxidized catalyst surface. Adding the CO2 along with the butane feed compensated for the lower oxygen partial
pressures near the reactor entrance, while using a greenhouse gas that was available as a by-product of the main reaction.

Tellez et al. (1999) explained the differences they observed between a PFR (FBR) and a IMR for the oxidative
dehydrogenation of butane over V/MgO by developing a model that included the effect of the O2 distribution through
the membrane on the state of the catalyst. Their kinetic model assumed two types of surface site, selective (2) and non-
selective (8). They showed through the model that for the IMR the ratio of the selective sites to the non-selective sites
was higher near the reactor entrance, which was where most of the reaction took place. This resulted in higher selectivity

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 19

to the desired product, butadiene.


Their computed profiles of the
ratio 2/8 are shown in Figure 12,
for various flow rates. By the
time the ratio is higher in the FBR
than in the IMR, the butane
concentration has decreased and
the reaction rate is low.

The studies discussed


above all show that the
performance of the catalyst
depends on the mixture
composition to which it is
exposed. This implies that to
understand the behavior of a
membrane reactor, it is essential
to understand how the catalyst
changes when the reactants are
fed differently.
Figure 12. Variation of the 2/8 ratio along the reactor. (Reprinted from Chem.
Eng. Science, 54, 2917, Copyright (1999), with permission from Elsevier) There can also be
changes in the dehydrogenation
catalysts used in product removal membrane reactors. The use of the membrane can lead to hydrogen-lean reaction
mixtures. Hou et al. (1999; 2000) have shown that hydrogen removal increases the tendency to catalyst deactivation due
to poisoning by H2S and carbon deposition in methane steam reforming. Similarly, Rival et al. (2001) warn of more
pronounced catalyst deactivation for CH4 non-oxidative aromatization, if hydrogen is withdrawn. Schäfer et al. (2003)
found that as a result of H2 removal coking was favored in
the propane dehydrogenation reaction. In a series of
publications on using zeolite membrane reactors (see
below) for the dehydrogenation of isobutane, Casanave et
al. (1999) and later Ciavarella et al. (2001) found that
under some conditions they obtained lower experimental
yields than expected from their modeling work, which had
been well-validated. They attributed this discrepancy to
catalyst deactivation or inadequacy of the kinetics when
hydrogen was removed from the reaction mixture. They
suggested that new catalysts may need to be designed to
work with membrane reactors, rather than just assuming
that catalysts developed for use in traditional fixed bed
reactors would also work in the very different environment
of a membrane reactor.

To address this problem for the case of propane


dehydrogenation, Wolfrath et al, (2001) and Kiwi-Minsker
et al. (2002) developed a novel membrane reactor as shown
in Figure 13. The catalyst was fabricated in the form of
filaments, which were placed in the membrane reactor in
two zones. One zone was inside the palladium membrane
tube, the other on the shell side. The catalyst filaments ran
parallel to the membrane tube, and the gas flowed axially
over them. The operation of the reactor was periodic; one
zone would be fed the propane, which was dehydrogenated Figure 13. Schematic of the two-zone
over the filaments to produce propylene and hydrogen, microstructured membrane reactor. (Reprinted from
accompanied by the deposition of coke. Air was fed to the Chem. Eng. Science, 57, 4947, Copyright (2002),
with permission from Elsevier)

Produced by The Berkeley Electronic Press, 2003


20 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

other zone, into which the hydrogen permeated. The oxygen then reacted with the hydrogen and the coke deposited on
the catalyst surface to produce CO2 and H2O, and regenerate the coked catalyst. The heat produced by the combustion
reactions helped to provide the energy needed for the dehydrogenation, and the lowered hydrogen partial pressure on
the permeate side due to the reaction increased the driving force for hydrogen permeation. Thus, this reactor also shows
some aspects of thermal and reaction coupling. In fact, it may be regarded as a variation on the reaction coupling idea,
extended to address the catalyst deactivation problem by periodic operation and using the membrane to partition active
and deactivated catalyst.

4.4 Molecular Sieve Membrane Reactors


The need for highly permselective membranes, which are stable to high temperatures and resist chemical attack, has led
to great interest in microporous materials, especially molecular sieve zeolites. Only a few years ago, review papers in
this area focused on synthesis of thin zeolite films and their use for separation, with only a brief mention of their uses
in reactors (Tavolaro and Drioli, 1999; Caro et al., 2000). The few studies of zeolite membrane reactors were summarized
by Dixon (1999) and Coronas and Santamaria (1999). The great strides forward that have been made in synthesizing
defect-free zeolite membranes or thin films have allowed research to progress on their use in reactors. Some early
indications of this were apparent in the conference reports that appeared at the end of 1998, and at present the literature
on this subject is building fast. In this review, any reactors that use molecular sieve membranes are included here, so that
silica and molecular sieve carbon membranes are included, although zeolitic materials form the largest group.

The use of molecular sieve materials in reactors has put a premium on understanding and using the adsorption
and diffusion interactions in the restricted diffusion regime. A parallel body of work is also rapidly developing on
modeling molecular sieve membranes, using approaches such as the Stefan-Maxwell flux equations, supplemented by
configurational Monte Carlo simulations of adsorption equilibria. Non-equilibrium molecular dynamics simulations are
also being used to understand phenomena at the molecular level. A review of these developments is beyond the scope
of this paper, the interested reader can consult the book by Sanchez and Tsotsis (2002) as a good starting point.

To organize the work in this relatively new area it is convenient to divide the reactor studies into three areas.
In the first, the membranes are inert (IMR), in the second they are either intrinsically catalytically active or have had
active components deposited in them (CMR), and in the third they function as pervaporation membranes, in which a
vapor (usually water) selectively permeates through the membrane from a liquid-phase mixture.

4.4.1 Inert Zeolite Membrane Reactors (IMR): as with the IMR based on a palladium or porous ceramic membrane,
the most common application is hydrogen product removal. However, the wide variety of zeolite materials, cage and
channel sizes and adsorption properties allows these materials to demonstrate good selectivity to other species.
Permselectivities that are much better than Knudsen selectivity can be achieved, and can often be tailored during
synthesis by cation exchange of the zeolite extra-framework ions.

Following work in the research group of Dalmon on the synthesis and testing of a zeolite-type MFI membrane
supported on "-alumina, Casanave et al. (1999) reported the initial reactor studies on the dehydrogenation of isobutane
over a Pt-In catalyst. Further work by Ciavarella et al. (2001) investigated more systematically the effects of changing
the operating conditions - feed flow rate, sweep gas flow rate and co-current versus counter-current sweep gas flow. The
first set of results (Casanave et al., 1999) obtained yields to isobutene that were in the range of 50 - 100% higher than
those obtained in a traditional PFR. By manipulating the operating conditions, Ciavarella et al. (2001) were able to obtain
yields that were four times the PFR results. Some of their results are shown in Figure 14. The left hand graph shows the
effect of the feed flow rate for different configurations at a fixed sweep gas rate (250 ml min-1). The right hand graph
shows the effect of the sweep flow rate for different configurations at a fixed feed flow rate (50 ml min-1).

Clearly the membrane reactor conversions exceed the thermodynamic equilibrium expectation. The effect of
changing sweep mode from co- to counter-current was small, however based on permeation experiments it was found
that the countercurrent mode was more effective in removing hydrogen. Both groups carried out comparisons of
cocurrent sweep mode and countercurrent sweep mode, and found that the reactor had membrane transport limitations
in cocurrent mode, and kinetics limitations in countercurrent mode. They concluded that the membrane reactor could
be further improved, since by simply changing sweep method the limiting factor in the reactor could be membrane or
catalyst. Later work (van Dyk et al., 2003) compared microporous MFI and dense Pd membranes for hydrogen removal.

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 21

Figure 14. Performance of the zeolite membrane reactor for isobutane conversion. (Reprinted from Catal.
Today, 67, 177, Copyright (2001), with permission from Elsevier)

The fact that similar results were obtained from the two very different membranes led the authors to conclude that the
lack of an active enough catalyst was the limiting factor in improving conversion.

Other molecular sieve materials have been put forward for hydrogen removal in reactors. Itoh and Haraya
(2000) developed carbon membranes with molecular sieving properties, in the form of carbon hollow fibers. They studied
the dehydrogenation of cyclohexane to benzene as a test reaction. When the conversion fell short of the model
predictions they suspected concentration polarization of the membrane and possible loss of catalyst activity in the lower
hydrogen atmosphere (see section 4.1.2). Giessler et al. (2003) reported on the performance of a molecular sieve silica
for hydrogen removal from the low-temperature water-gas shift reaction. They compared both hydrophilic and
hydrophobic membranes, the latter were better and gave 99% conversion of CO, well above the equilibrium conversion
achievable with a PFR. Similarly, Lee et al. (2003) prepared silica/porous stainless steel membranes that preferentially
permeated hydrogen from methanol steam reforming product mixtures.

The affinity of some zeolite materials for water has led to their use in membrane reactors for dehydration
reactions. Guiru et al. (1999) used a zeolite 4A membrane for the dehydration of diethylene glycol. They obtained
complete conversion at temperatures of only 160 °C, where for the PFR temperatures of more than 250 °C were required.
Better selectivity was associated with lower temperatures, due to the suppression of side reactions. Salomón et al. (2000)
compared mordenite and NaA zeolite membranes for the selective removal of water from the gas-phase synthesis of
MTBE from tert-butanol and methanol. Higher yields were obtained with the zeolite membrane reactor compared to the
PFR, due to water removal by preferential adsorption on the hydrophilic zeolite.

Other, less usual species may also be preferentially removed using zeolite membranes. van de Graaf et al.
(1999a; 1999b) used a silicalite-1 membrane to separate trans-2-butene from propene and cis-2-butene in the metathesis
of propene reaction. The permselectivity of the trans-2-butene with respect to propene was about 5, due to the fact that
trans-2-butene was more strongly adsorbed into the zeolite pores, and blocked the propene. When they fed pure propene
to the reactor the conversion was worse than for the PFR, and the permselectivity was 2.3. This was attributed to reactant
loss at the reactor entrance, as there was no trans-2-butene to block the propene, which permeated readily and was lost.
When they added a pre-reactor before the membrane reactor to get some trans-2-butene in the system they obtained a
13% increase in propene conversion over the equilibrium value. Piera et al. (2001) used silicalite-1 membranes grown
on stainless steel porous tubes for selective removal of i-octene from the reaction mixture in the liquid-phase
oligomerization of i-butene. They got the same i-butene conversion as for a PFR over the amberlyst catalyst, but much
better C8 selectivity and yield by removing the intermediate product to prevent further reaction to C12 or C16 species. The
reaction was not equilibrium-limited, so it was expected that there would be no effect on conversion, except at higher

Produced by The Berkeley Electronic Press, 2003


22 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

space velocities, due to the increase in residence time in the zeolite membrane reactor.

Finally, the use of a zeolite IMR for a distributed reactant feed was studied by Mota et al. (2001) for butane
oxidation to maleic anhydride over vanadium phosphoros mixed oxides. They showed that with a pure butane feed, the
catalyst near the reactor inlet performed badly due to reduction of the vanadium to V4+ species, and they identified the
need to co-feed some oxygen with the butane. Then the maleic anhydride yield increased to that of the PFR. The
advantage of the membrane reactor was that it was possible to increase the butane feed due to the limitation of the
flammability problems. They also found that if the butane flow was reversed, then contact of the gases with the more
oxidized reactor zone increased transient yield. Further work was promised on the reverse-flow concept for both heat
management and control of catalyst properties. Pârvulescu et al. (2002) used Ni-MCM-41 mesoporous molecular sieve
membranes to control the addition of H2O2 for conversion of benzene and toluene to benzaldehyde and phenol. Gora et
al. (2002) used a zeolite membrane to feed n-hexane along the tube length of a zeolite membrane reactor for the
hydroisomerization of C6-hydrocarbons.

4.4.2 Catalytic Zeolite Membrane Reactors (CMR): in these the object is to contact the gas with the catalyst, so CMR
designs are often operated in a flow-through mode (see section 4.3), or used to control contact time. A recent example
is the study of Masuda et al. (2003) for the conversion of methanol to olefins. They used a ZSM-5 catalytic membrane
on the outer surface of an alumina ceramic filter, as shown in Figure 15.

Figure 15. Concept for recovery of intermediate species at high yields using a
ZSM-5 zeolite membrane reactor. (Reprinted from Chem. Eng. Science, 58, 649,
Copyright (2003), with permission from Elsevier)

The methanol to olefin reaction is a series reaction, in which methanol reacts to dimethylether, then to olefins
(desired) and finally to paraffins and aromatics (not desired). In a fixed bed, dispersion and pellet diffusion result in a
broad range of contact times, and it is hard to design the reactor to give high yield of olefins. The use of the CMR, with
pressure drop across the membrane to control the permeation rate across it, can give a much more uniform residence
time. In Figure 15, the horizontal arrow denotes the rate of reaction in the zeolite layer, while the vertical arrow denotes
the rate of permeation. The idea is to match reaction rates and diffusion rates so that the methanol has reacted to olefins
just as it exits the zeolite layer, then the olefins permeate fast out of the filter layer. Masuda et al. (2003) reported olefin
selectivities of 80-90%, for methanol conversions of 60-98%.

Other uses of the CMR for contact time control have been reported. Hasegawa et al. (2001) studied selective
oxidationof CO by permeation through a Pt-loaded Y-type zeolite membrane. Their concern was with the hydrogen feed

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 23

for fuel cells, which usually has low amounts of CO present, which cannot be tolerated by the nafion membrane
downstream. In their CMR, hydrogen permeates through the membrane much faster than any CO or oxygen, which
therefore have much longer residence times over the catalyst in the PtY membrane, and the CO is converted leaving a
hydrogen permeate stream with greatly reduced CO content. Maira et al. (2003) were discussed earlier for their flow-
through membrane (section 4.3). Their membranes were actually hybrid silicalites with TiO2 catalyst for the VOC
photocatalytic oxidation. Julbe et al. (2000) studied V-MFI membranes for the oxidative dehydrogenation of propane.
They looked at oxygen distribution versus co-feed flow-through modes. They found no improvement with oxygen
distribution, and concluded that the membrane acted to improve the contact between the reactants and the catalyst.
Bottino et al. (2002) used a CMR for the separated feed of reactants in the oxidehydrogenation of propane, and found
that V/Al2O3 catalyst was more effective than V/ZSM-5.

4.4.3 Zeolite Membrane Pervaporation Reactors (PVMR): Tanaka et al. (2001) studied pervaporation-aided
esterification of acetic acid with ethanol at 343 K over an amberlyst catalyst using zeolite T inert membranes. They
obtained almost complete conversion when the water vapor was removed through the membrane, as opposed to
approximately 75% without pervaporation. The main problem in this system was to find a pervaporation membrane that
could withstand the acid reaction mixture.

Ethanol esterification was also studied by Bernal et al. (2002) who used H-ZSM-5 membranes in their CMR.
Water and ethyl acetate were generated on the membrane, which they suggested would favor transport to the permeate
side. They compared the PFR, an IMR with H-ZSM-5 catalyst particles packed inside a Na-ZSM-5 hydrophilic
membrane for water removal, and the H-ZSM-5 CMR. They found that the active membrane (CMR) performed
somewhat better on conversion.

4.5 Integration of Functionality Through Multiple Membranes


There has been ongoing research interest in combining more functions into one unit operation, using either multiple
membranes or having one membrane function in two or more ways. For example, having both a metered feed and a
product removal in one operation, or another combination of any of the functions discussed previously. Prior areas in
which there was research interest included using two membranes to carry out supported liquid phase catalysis, the use
of membranes for simultaneous reactant distribution and product removal, and membranes with catalytic and separation
layers adjacent to each other. There has been no further research appearing in the first area, one or two examples of the
second, while the third has seen an increase in activity
and new developments.

4.5.1 Multiple Membranes: although there were


previously several theoretical studies on the use of two
different membranes in the same reactor, there have
been no experimental studies to date, and interest
appears to be declining. This possibly reflects the
difficulties of finding conditions suitable for two
different membranes to both work well, and the
technical problems of operating such a complex unit.
Prasad and Elnashaie (2002) included membrane tubes
for both oxygen feed distribution and hydrogen
product removal in their fluidized bed model. Omorjan
et al. (1999) continued a series of theoretical studies,
discussing the maximal extent of an isothermal
reversible reaction A ] B + C with two membranes,
each permeable to only one of the products. They
showed that complete conversion was possible with a
double membrane reactor, and illustrated their
Figure 16. Membrane reactor and co-feed reactor with
calculations for the water splitting reaction.
the grained catalyst (Ag0.01Bi0.85V0.54Mo0.45O4).
(Reprinted from J. Membrane Sci., 198, 119, Copyright
4.5.2 Multiple Function Membranes: there has been
(2002), with permission from Elsevier)
some growth in interest in this area, as evidenced by

Produced by The Berkeley Electronic Press, 2003


24 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

the appearance of several new ideas. Kölsch et al. (2002) developed a membrane that could both feed oxygen and
separate acrolein and water, for the direct oxidation of propane to acrolein. Their idea was to dose oxygen to the reactor
via a porous membrane, to keep low oxygen partial pressure and thus high selectivity to acrolein on the reactor side.
also, they proposed to remove acrolein as a reactive intermediate by an oxygenate-selective membrane. The membrane
was prepared by in-situ hydrolysis of TEOS on part of an asymmetric porous alumina membrane. Their permeation tests
showed a permselectivity for acrolein over other reaction products of 2-3. They compared two configurations, as shown
in Figure 16. Both involved the oxygenate-selective membrane, but one co-fed the oxygen along with the propane, and
the other distributed it along the length of the reactor. They were able to get acrolein yields five times larger with the
distributed feed than with the co-feed, and nine times better selectivity. However, the distributed feed conversion was
only 50% of the co-feed conversion, and the authors concluded that yields were still too low for industrial application,
and that a better catalyst was needed to take advantage of the new configuration.

Nomura et al. (2003) combined a separative layer, a porous catalytic layer and a porous metal support layer,
termed a heat exchange (HEX) membrane reactor due to the higher thermal conductivity of the stainless steel support.
Liu et al. (2003) reported on the use of a NiP-Cu multilayer membrane for ethanol dehydrogenation. Their original NiP
membrane gave improved separation of hydrogen, but some reactant was lost through it. They therefore added a catalytic
Cu layer to convert the permeating ethanol that would otherwise be wasted. They obtained higher ethanol conversion
and acetaldehyde yield compared to using the NiP membrane with only separation functionality. Abashar et al. (2002)
looked at dual functionality via spatial catalytic patterns. They incorporated two types of catalyst in their product removal
reactor, mentioned previously (Figure 1). The main catalyst decomposed ammonia over Ni/Al2O3, the auxiliary catalyst
carried out methanation reactions over Ni/Mg Al2O4 spinal, to consume hydrogen. Hydrogen was also removed
conventionally via a palladium membrane. They
modeled various catalyst configurations, and found
that they could get complete conversion at low
temperature.

Finally, the two-layer membrane of Alfonso


et al. (2001), which has been mentioned previously, is
shown in Figure 17. A butane/oxygen/inert mixture
was fed to the membrane and goes through two
different consecutive reactions in the two different
catalyst layers. First is oxidative dehydrogenation over
a V/MgO catalyst, which should be thick enough to
consume all the oxygen in the feed. Then non-
oxidative dehydrogenation takes place over a Pt-Sn/(-
Al2O3 catalytic layer. The two-layer membrane could
increase both conversion and selectivity compared to
the results obtained when a VMgO single-layer
membrane was used.

Figure 17. Two-layer, flow-through catalytic membrane


reactor concept. (Reprinted with permission from Ind. Eng.
Chem. Res. 2001, 40, 1058. Copyright 2001 American
Chemical Society)

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 25

6. CONCLUDING REMARKS
Progress towards commercialization of catalytic membrane reactors continues to be slow, and may not pick up until there
is a definite success reported. From the industrial point of view there are still many issues to be resolved before a case
can be made for membrane reactors, including the cost of the membranes, their production in quantity, and their
reliability. Commercial zeolite membranes for dehydration and pervaporation reaction applications may already be a
reality, while the two large consortia in the USA, one led by Air Products and one by Praxair, that are working towards
dense oxide membrane reactors for syngas production, may be getting close to a commercial unit. However, it will still
probably be difficult to commit to an entire process based on a membrane reactor, especially if the process depends on
the new technology. It may be easier for membrane reactors to gain a foothold in smaller-scale applications, as suggested
by Tennison (2000), such as fuel cells and/or microreactors (Goto et al., 2003). The increased emphasis on membrane
reactors related to fuel cells has been noticeable in the proceedings of recent conferences.

On the research side, interest is still strong, with the trend in the number of publications on catalytic membrane
reactors increasing steadily despite fluctuations from year to year. Progress in the field is strongly tied to membrane
materials development, but reaction engineers are showing ingenuity in finding new configurations and niche
applications for the use of existing catalytic membranes.

Despite being the oldest area of application of catalytic membrane reactors, the removal of product hydrogen
still gives rise to a great deal of work, and new reactions and areas are being found. Distributed feed of reactant oxygen
continues to be a strong area of research, especially for partial oxidation of propane or butane, and for a wide range of
oxidative dehydrogenation reactions. The availability of more versatile zeolitic and other microporous materials. which
allow the selective removal of intermediate species and product species other than hydrogen, shows promise of allowing
the development of new areas of research.

As the subject of catalytic membrane reactors matures, the level of sophistication of experimental and modeling
studies increases. More detailed models are being developed, including the use of two-dimensional reactor tube models
and the inclusion of both concentration and temperature gradients. Researchers are becoming aware of the need for a
better understanding of the reaction kinetics and the subtleties of membrane reactor design. In particular, there is a
growing body of opinion that it is necessary to design the reactor and develop the catalyst together, taking into account
the interactions between them. It is not enough to just put membrane tubes into the usual fixed bed reactor design and
operation.

There is increased use of membranes to control the contact time of reactants and catalyst, particularly in the
flow-through configuration. This is becoming a strong third area of catalytic membrane reactor research. Another area
of strong growth is the use of zeolitic or molecular sieving membrane reactors. In the last five years, the synthesis of
zeolitic thin film materials has progressed to the point where there is now a substantial body of literature on zeolite
membrane reactors. However, in most cases, long-term performance and large scale unit behavior have yet to be tested.
The area of multifunctionality of membrane reactors, in which the membrane does more than one of extract, distribute
or contact, gives rise to a small but diverse range of investigations. Current work in this area promises to lead to some
truly novel ideas.

Despite some gloomy prognoses in the past, the area of catalytic membrane reactors continues to attract a lot
of research interest. The progress that has been made to date suggests that we may be confident that the challenges to
the commercial use of membrane reactors will be met in the future.

ACKNOWLEDGMENTS
This review is based upon a keynote lecture given by the author at the Engineering Conferences International meeting
on Chemical Reactor Engineering - Meeting the Challenges for New Technology, June 29-July 4, 2003, Quebec City,
Quebec. I would like to thank the Conference Chairs Hugo de Lasa and Franco Berutti for the invitation to speak.
Thanks are also due to many colleagues who supplied originals of figures from their own publications for this review.

Produced by The Berkeley Electronic Press, 2003


26 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

REFERENCES
Abashar, M.E.E., Al-Sughar, Y.S., Al-Mutaz, I.S., "Investigation of Low Temperature Decomposition of Ammonia
Using Spatially Patterned Catalytic Membrane Reactors", Appl. Catal. A, Vol. 236, 35-53 (2002).

Akin, F.T., Lin, Y.S., "Oxidative Coupling of Methane in Dense Ceramic Membrane Reactor with High Yields", AIChE
J., Vol. 48, 2298-2306 (2002a).

Akin, F.T., Lin, Y.S., "Selective Oxidation of Ethane to Ethylene in a Dense Tubular Membrane Reactor", J. Memb. Sci.,
Vol. 209, 457-467 (2002b).

Akin, F.T., Lin, Y.S., Zeng, Y., "Comparative Study on Oxygen Permeation and Oxidative Coupling of Methane on
Disk-Shaped and Tubular Dense Ceramic Membrane Reactors", Ind. Eng. Chem. Res., Vol. 40, 5908-5916 (2001).

Al-Juaied, M.A., Lafarga, D., Varma, A., "Ethylene Epoxidation in a Catalytic Packed-Bed Membrane Reactor:
Experiments and Model", Chem. Eng. Sci., Vol. 56, 395-402 (2001).

Alfonso, M.J., Julbe, A., Farrusseng, D., Menendez, M., Santamaria, J., "Oxidative Dehydrogenation of Propane on
V/Al2O3 Catalytic Membranes. Effect of the Type of Membrane and Reactant Feed Configuration", Chem. Eng. Sci.,
Vol. 54, 1265-1272 (1999).

Alfonso, M.J., Menendez, M., Santamaria, J., "Vanadium-Based Catalytic Membrane Reactors for the Oxidative
Dehydrogenation of Propane", Catal. Today, Vol. 56, 247-252 (2000).

Alfonso, M.J., Menendez, M., Santamaria, J., "Coupling of Consecutive Reactions in a Two-Layer, Flow-Through
Catalytic Membrane", Ind. Eng. Chem. Res., Vol. 40, 1058-1064 (2001).

Alfonso, M.J., Menendez, M., Santamaria, J., "Oxidative Dehydrogenation of Butane on V/MgO Catalytic Membranes",
Chem. Eng. J., Vol. 90, 131-138 (2002).

Alonso, M., Lorences, M.J., Pina, M.P., Patience, G.S., "Butane Partial Oxidation in an Externally Fluidized
Bed-Membrane Reactor", Catal. Today, Vol. 67, 151-157 (2001).

Assabumrungrat, S., Paserthdam, P., Goto, S., "Oxidative Coupling of Methane in a Ceramic Membrane Reactor:
Uniform Oxygen Permeation Pattern", J. Chin. Inst. Chem. Engrs., Vol. 31, 19-25 (2000).

Assabumrungrat, S., Kiatkittipong, W., Praserthdam, P., Goto, S., "Simulation of Pervaporation Membrane Reactors for
Liquid Phase Synthesis of Ethyl Tert-Butyl Ether from Tert-Butyl Alcohol and Ethanol", Catal. Today, Vol. 79-80,
249-257 (2003).

Barbieri, G., Marigliano, G., Perri, G., Drioli, E., "Conversion-Temperature Diagram for a Palladium Membrane Reactor.
Analysis of an Endothermic Reaction: Methane Steam Reforming", Ind. Eng. Chem. Res., Vol. 40, 2017-2026 (2001).

Basile, A., Paturzo, L., "An Experimental Study of Multilayered Composite Palladium Membrane Reactors for Partial
Oxidation", Catal. Today, Vol. 67, 55-64 (2001).

Basile, A., Paturzo, L., Laganà, F., "The Partial Oxidation of Methane to Syngas in a Palladium Membrane Reactor:
Simulation and Experimental Studies", Catal. Today, Vol. 67, 65-75 (2001).

Basile, A., Paturzo, L., Vazzana, A., "Membrane Reactor for the Production of Hydrogen and Higher Hydrocarbons from
Methane Over Ru/Al2O3 Catalyst", Chem. Eng. J., Vol. 93, 31-39 (2003a).

Basile, A., Paturzo, L., Gallucci, F., "Co-Current and Counter-Current Modes for Water Gas Shift Membrane Reactor",
Catal. Today, Vol. 82, 275-281 (2003b).

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 27

Bengtson, G., Scheel, H., Theis, J., Fritsch, D., "Catalytic Membrane Reactor to Simultaneously Concentrate and React
Organics", Chem. Eng. J., Vol. 85, 303-311 (2002).

Bernal, M.P., Coronas, J., Menendez, M., Santamaria, J., "Coupling of Reaction and Separation at the Microscopic
Level: Esterification Processes in a H-ZSM-5 Membrane Reactor", Chem. Eng. Sci., Vol. 57, 1557-1562 (2002).

Bottino, A., Capannelli, G., Comite, A., "Catalytic Membrane Reactors for the Oxidehydrogenation of Propane;
Experimental and Modelling Study", J. Memb. Sci., Vol. 197, 75-88 (2002).

Bouwmeester, H.J.M., "Dense Ceramic Membranes for Methane Conversion", Catal. Today, Vol. 82, 141-150 (2003).

Brinkmann, T., Perera, S.P., Thomas, W.J., "An Experimental and Theoretical Investigation of a Catalytic Membrane
Reactor for the Oxidative Dehydrogenation of Methanol", Chem. Eng. Sci., Vol. 56, 2047-2061 (2001).

Cai, R., Tong, J.H., Zhu, H.J., Zhu, B.C., Sheng, S.S., Yang, W.S., "Electrode-electrolyte BIMEVOX Membrane Reactor
for the Oxidation of Alkanes at Moderate Temperatures", paper O-23, 5th Int. Conf. Catalysis in Membrane Reactors,
Dalian, China, June 26-28 (2002).

Caro, J., Noack, M., Kölsch, P., Schäfer, R., "Zeolite Membranes - State of Their Development and Perspective",
Microp. Mesop. Mat., Vol. 38, 3-24 (2000).

Casanave, D., Ciavarella, P., Fiaty, K., Dalmon, J.-A., "Zeolite Membrane Reactor for Isobutane Dehydrogenation:
Experimental Results and Theoretical Modelling", Chem. Eng. Sci., Vol. 54, 2807-2815 (1999).

Centi, G., Dittmeyer, R., Perathoner, S., Reif, M., "Tubular Inorganic Catalytic Membrane Reactors: Advantages and
Performance in Multiphase Hydrogenation Reactions", Catal. Today, Vol. 79-80, 139-149 (2003).

Chan, P.P.Y., Vanidjee, K., Adesina, A.A., Rogers, P.L., "Modeling and Simulation of Non-Isothermal Catalytic Packed
Bed Membrane Reactor for H2S Decomposition", Catal. Today, Vol. 63, 379-385 (2000).

Chang, C.-C., Reo, C.M., Lund, C.R.F., "The Effect of a Membrane Reactor Upon Catalyst Deactivation During
Hydrodechlorination of Dichloroethane", Appl. Catal. B, Vol. 20, 309-317 (1999).

Chen, Z., Yan, Y., Elnashaie, S.S.E.H., "Modeling and Optimization of a Novel Membrane Reformer for Higher
Hydrocarbons", AIChE J., Vol. 49, 1250-1265 (2003).

Chmielewski, D., Ziaka, Z.D., Manousiouthakis, V., "Conversion Targets for Plug Flow Membrane Reactors", Chem.
Eng. Sci., Vol. 54, 2979-2984 (1999).

Ciavarella, P., Casanave, D., Moueddeb, H., Miachon, S., Fiaty, K., Dalmon, J.-A., "Isobutane Dehydrogenation in a
Membrane Reactor. Influence of the Operating Conditions on the Performance", Catal. Today, Vol. 67, 177-184 (2001).

Coronas, J., Santamaria, J., "Catalytic Reactors Based on Porous Ceramic Membranes", Catal. Today, Vol. 51, 377-389
(1999).

Criscuoli, A., Basile, A., Drioli, E., "An Analysis of the Performance of Membrane Reactors for the Water-Gas Shift
Reaction Using Gas Feed Mixtures", Catal. Today, Vol. 56, 53-64 (2000).

Criscuoli, A., Basile, A., Drioli, E., Loiacono, O., "An Economic Feasibility Study for Water Gas Shift Membrane
Reactor", J. Memb. Sci., Vol. 181, 21-27 (2001).

Daub, K., Wunder, V.K., Dittmeyer, R., "CVD Preparation of Catalytic Membranes for Reduction of Nitrates in Water",
Catal. Today, Vol. 67, 257-272 (2001).

Diakov, V., Lafarga, D., Varma, A., "Methanol Oxidative Dehydrogenation in a Catalytic Packed-Bed Membrane

Produced by The Berkeley Electronic Press, 2003


28 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

Reactor", Catal. Today, Vol. 67, 159-167 (2001).

Diakov, V., Varma, A., "Reactant Distribution by Inert Membrane Enhances Packed-Bed Reactor Stability", Chem. Eng.
Sci., Vol. 57, 1099-1105 (2002).

Diakov, V., Blackwell, B., Varma, A., "Methanol Oxidative Dehydrogenation in a Catalytic Packed-Bed Membrane
Reactor: Experiments and Model", Chem. Eng. Sci., Vol. 57, 1563-1569 (2002).

Diakov, V., Varma, A., "Methanol Oxidative Dehydrogenation in a Packed-Bed Membrane Reactor: Yield Optimization
Experiments and Model", Chem. Eng. Sci., Vol. 58, 801-807 (2003).

Dittmeyer, R., Hollein, V., Quicker, P., Emig, G., Hausinger, G., Schmidt, F., "Factors Controlling the Performance of
Catalytic Dehydrogenation of Ethylbenzene in Palladium Composite Membrane Reactors", Chem. Eng. Sci., Vol. 54,
1431-1439 (1999).

Dittmeyer, R., Hollein, V., Daub, K., "Membrane Reactors for Hydrogenation and Dehydrogenation Processes Based
on Supported Palladium", J. Molec. Catal. A: Chemical, Vol. 173, 135-184 (2001).

Dixon, A.G., "Innovations in Catalytic Inorganic Membrane Reactors", in Spivey, J.J. (ed.), Specialist Periodical
Reports: Catalysis, Vol. 14, Royal Society of Chemistry, London, 40-92 (1999).

Dixon, A.G., "Analysis of Intermediate Product Yield in Distributed-Feed Nonisothermal Tubular Membrane Reactors",
Catal. Today, Vol. 67, 189-203 (2001).

Dong, H., Shao, Z., Xiong, G., Tong, J., Sheng, S., Yang, W., "Investigation on POM Reaction in a New Perovskite
Membrane Reactor", Catal. Today, Vol. 67, 3-13 (2001).

Elnashaie, S.S.E.H., Abdallah, B.K., Elshishini, S.S., Alkhowaiter, S., Noureldeen, M.B., Alsoudani, T., "On the Link
Between Intrinsic Catalytic Reactions Kinetics and the Development of Catalytic Processes. Catalytic Dehydrogenation
of Ethylbenzene to Styrene", Catal. Today, Vol. 64, 151-162 (2001).

Espro, C., Arena, F., Frusteri, F., Parmaliana, A., "On the Potential of the Multifunctional Three Phase Catalytic
Membrane Reactor in the Selective Oxidation of Light Alkanes by Fe2+-H2O2 Fenton System", Catal. Today, Vol. 67,
247-256 (2001).

Fan, J., Ohashi, H., Ohya, H., Aihara, M., Takeuchi, T., Negishi, Y., Semenova, S.I., "Analysis of a Two-Stage
Membrane Reactor Integrated with Porous Membrane Having Knudsen Diffusion Characteristics for the Thermal
Decomposition of Hydrogen Sulfide", J. Memb. Sci., Vol. 166, 239-247 (2000).

Ferreira-Aparicio, P., Rodriguez-Ramos, I., Guerrero-Ruiz, A., "On the Applicability of Membrane Technology to the
Catalysed Dry Reforming of Methane", Appl. Catal. A, Vol. 237, 239-252 (2002a).

Ferreira-Aparicio, P., Rodriguez-Ramos, I., Guerrero-Ruiz, A., "On the Performance of Porous Vycor Membranes for
Conversion Enhancement in the Dehydrogenation of Methylcyclohexane to Toluene", J. Catal., Vol. 212, 182-192
(2002b).

Ge, S.H., Liu, C.H., Wang, L.J., "Oxidative Dehydrogenation of Butane Using Inert Membrane Reactor with a
Non-Uniform Permeation Pattern", Chem. Eng. J., Vol. 84, 497-502 (2001).

Giessler, S., Jordan, L., Diniz da Costa, J.C., Lu, G.Q., "Performance of Hydrophobic and Hydrophilic Silica Membrane
Reactors for the Water Gas Shift Reaction", Sep. Purif. Technol., Vol. 32, 255-264 (2003).

Gora, L., Buijsse, E.J.W., Jansen, J.C., Maschmeyer, Th., "Hydroisomerization of C6 in a Zeolite Membrane Reactor",
paper O-2, 5th Int. Conf. Catalysis in Membrane Reactors, Dalian, China, June 26-28 (2002).

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 29

Goto, S., Tagawa, T., Assabumrungrat, S., Praserthdam, P., "Simulation of Membrane Microreactor for Fuel Cell with
Methane Feed", Catal. Today, Vol. 82, 223-232 (2003).

Grace, J.R., Li, X., Lim, C.J., "Equilibrium Modelling of Catalytic Steam Reforming of Methane in Membrane Reactors
with Oxygen Addition", Catal. Today, Vol. 64, 141-149 (2001).

Gryaznov, V., "Membrane Catalysis", Catal. Today, Vol. 51, 391-395 (1999).

Gryaznov, V., Ermilova, M.M., Orekhova, N.V., , "Membrane-Catalyst Systems for Selectivity Improvement in
Dehydrogenation and Hydrogenation Reactions", Catal. Today, Vol. 67, 185-188 (2001).

Gu, X., Jin, W., Chen, C., Xu, N., Shi, J., Ma, Y.H., "YSZ-SrCo0.4Fe0.6O3-* Membranes for the Partial Oxidation of
Methane to Syngas", AIChE J., Vol. 48, 2051-2060 (2002).

Guiru, W., Hongchen, G., Yushan, L., "The preparation of zeolite 4A membrane reactor and its use in dehydration of
diethylene glycol", Proceedings 12th International Zeolite Conference, Materials Research Society, 1795-1802, (1999).

Hara, S., Sakaki, K., Itoh, N., "Decline in Hydrogen Permeation Due to Concentration Polarization and CO Hindrance
in a Palladium Membrane Reactor", Ind. Eng. Chem. Res., Vol. 38, 4913-4918 (1999).

Hasegawa, Y., Kusakabe, K., Morooka, S., "Selective Oxidation of Carbon Monoxide in Hydrogen-Rich Mixtures by
Permeation Through a Platinum-Loaded Y-Type Zeolite Membrane", J. Memb. Sci., Vol. 190, 1-8 (2001).

Hoff, K.A., Poplsteinova, J., Jakobsen, H.A., Falk-Pedersen, O., Juliussen, O., Svendsen, H.F., "Modeling of Membrane
Reactor", Int. J. Chem. Reactor Eng., Vol. 1, Article A9 (2003).
http//www.bepress.com/ijcre/vol1/A9

Hou, K., Fowles, M., Hughes, R., "Potential Catalyst Deactivation Due to Hydrogen Removal in a Membrane Reactor
Used for Methane Steam Reforming", Chem. Eng. Sci., Vol. 54, 3783-3791 (1999).

Hou, K., Fowles, M., Hughes, R., "The Effect of Hydrogen Removal During Methane Steam Reforming in Membrane
Reactors in the Presence of Hydrogen Sulphide", Catal. Today, Vol. 56, 13-20 (2000).

Hou, K., Hughes, R., Ramos, R., Menendez, M., Santamaria, J., "Simulation of a Membrane Reactor for Oxidative
Dehydrogenation of Propane, Incorporating Radial Concentration and Temperature Profiles", Chem. Eng. Sci., Vol. 56,
57-67 (2001).

Hou, K., Hughes, R., "A Comparative Simulation Analysis of Propane Dehydrogenation in Composite and Microporous
Membrane Reactors", J. Chem. Technol. Biotechnol., Vol. 78, 35-41 (2002).

Ilinitch, O.M., Cuperus, F.P., Nosova, L.V., Gribov, E.N., "Catalytic Membrane in Reduction of Aqueous Nitrates:
Operational Principles and Catalytic Performance", Catal. Today, Vol. 56, 137-145 (2000).

Iliuta, M.C., Larachi, F., Grandjean, B.P.A., Iliuta, I., Sayari, A., "Methane Nonoxidative Aromatization Over
Ru-Mo/HZSM-5 in a Membrane Catalytic Reactor", Ind. Eng. Chem. Res., Vol. 41, 2371-2378 (2002).

Iliuta, M.C., Grandjean, B.P.A., Larachi, F., "Methane Nonoxidative Aromatization Over Ru-Mo/HZSM-5 at
Temperatures up to 973 K in a Palladium-Silver/Stainless Steel Membrane Reactor", Ind. Eng. Chem. Res., Vol. 42,
323-330 (2003).

Ishihara, T., Kawahara, A., Fukunaga, A., Nishiguchi, H., Shinkai, H., Miyaki, M., Takita, Y., "CH4 Decomposition with
a Pd-Ag Hydrogen-Permeating Membrane Reactor for Hydrogen Production at Decreased Temperature", Ind. Eng.
Chem. Res., Vol. 41, 3365-3369 (2002).

Itoh, N., Haraya, K., "A Carbon Membrane Reactor", Catal. Today, Vol. 56, 103-111 (2000).

Produced by The Berkeley Electronic Press, 2003


30 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

Itoh, N., Tamura, T., Hara, S., Takahashi, T., Shono, A., Satoh, K., Namba, T., "Hydrogen Recovery from Cyclohexane
as a Chemical Hydrogen Carrier Using a Palladium Membrane Reactor", Catal. Today, Vol. 82, 119-125 (2003).

Jarosch, K., de Lasa, H.I., "Novel Riser Simulator for Methane Reforming Using High Temperature Membranes", Chem.
Eng. Sci., Vol. 54, 1455-1460 (1999).

Jin, W., Gu, X., Li, S., Huang, P., Xu, N., Shi, J., "Experimental and Simulation Study on a Catalyst Packed Tubular
Dense Membrane Reactor for Partial Oxidation of Methane to Syngas", Chem. Eng. Sci., Vol. 55, 2617-2625 (2000a).

Jin, W., Li, S., Huang, P., Xu, N., Shi, J., Lin, Y.S., "Tubular Lanthanum Cobaltite Perovskite-Type Membrane Reactors
for Partial Oxidation of Methane to Syngas", J. Memb. Sci., Vol. 166, 13-22 (2000b).

Julbe, A., Farrusseng, D., Jalibert, J.C., Mirodatos, C., Guizard, C., "Characteristics and Performance in the Oxidative
Dehydrogenation of Propane of MFI and V-MFI Zeolite Membranes", Catal. Today, Vol. 56, 199-209 (2000).

Julbe, A., Farrusseng, D., Cot, D., Guizard, C., "The Chemical Valve Membrane: A New Concept for an
Auto-Regulation of O2 Distribution in Membrane Reactors", Catal. Today, Vol. 67, 139-149 (2001a).

Julbe, A., Farrusseng, D., Guizard, C., "Porous Ceramic Membranes for Catalytic Reactors - Overview and New Ideas",
J. Memb. Sci., Vol. 181, 3-20 (2001b).

Kao, Y.K., Lei, L., Lin, Y.S., "Optimum Operation of Oxidative Coupling of Methane in Porous Ceramic Membrane
Reactors", Catal. Today, Vol. 82, 255-273 (2003).

Keuler, J.N., Lorenzen, L., "Comparing and Modeling the Dehydrogenation of Ethanol in a Plug-Flow Reactor and a
Pd-Ag Membrane Reactor", Ind. Eng. Chem. Res., Vol. 41, 1960-1966 (2002a).

Keuler, J.N., Lorenzen, L., "The Dehydrogenation of 2-Butanol in a Pd-Ag Membrane Reactor", J. Memb. Sci., Vol.
202, 17-26 (2002b).

Kiwi-Minsker, L., Wolfrath, O., Renken, A., "Membrane Reactor Microstructured by Filamentous Catalyst", Chem. Eng.
Sci., Vol. 57, 4947-4953 (2002).

Klose, F., Wolff, T., Thomas, S., Seidel-Morgenstern, A., "Concentration and Residence Time Effects in Packed Bed
Membrane Reactors", Catal. Today, Vol. 82, 25-40 (2003).

Kobayashi, M., Togawa, J., Kanno, T., Horiuchi, J., Tada, K., "Dramatic Innovation of Propene Epoxidation Efficiency
Derived from a Forced Flow Membrane Reactor", J. Chem. Technol. Biotechnol., Vol. 78, 303-307 (2003).

Kölsch, P., Noack, M., Schäfer, R., Georgi, G., Omorjan, R., Caro, J., "Development of a Membrane Reactor for the
Partial Oxidation of Hydrocarbons: Direct Oxidation of Propane to Acrolein", J. Memb. Sci., Vol. 198, 119-128 (2002).

Koukou, M.K., Papayannakos, N., Markatos, N.C., "On the Importance of Non-Ideal Flow Effects in the Operation of
Industrial-Scale Adiabatic Membrane Reactors", Chem. Eng. J., Vol. 83, 95-105 (2001).

Lafarga, D., Varma, A., "Ethylene Epoxidation in a Catalytic Packed-Bed Membrane Reactor: Effects of Reactor
Configuration and 1,2-Dichloroethane Addition", Chem. Eng. Sci., Vol. 55, 749-758 (2000).

Langhendries, G., Baron, G.V., Jacobs, P.A., "Selective and Efficient Hydrocarbon Oxidation in a Packed Bed
Membrane Reactor", Chem. Eng. Sci., Vol. 54, 1467-1472 (1999).

Lapkin, A.A., Tennison, S.R., Thomas, W.J., "A Porous Carbon Membrane Reactor for the Homogeneous Catalytic
Hydration of Propene", Chem. Eng. Sci., Vol. 57, 2357-2369 (2002).

Lee, D.-W., Lee, Y.-G., Nam, S.-E., Sea, B., Lee, K.-H., "Preparation and Characterization of SiO2 Composite

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 31

Membrane for Purification of Hydrogen from Methanol Steam Reforming as an Energy Carrier System for PEMFC",
Sep. Purif. Technol., Vol. 32, 45-50 (2003).

Li, C.-F., Zhong, S.-H., "Study on Application of Membrane Reactor in Direct Synthesis DMC from CO2 and CH3OH
Over Cu-KF/MgSiO Catalyst", Catal. Today, Vol. 82, 83-90 (2003).

Li, L., Borry, R.W., Iglesia, E., "Design and Optimization of Catalysts and Membrane Reactors for the Non-Oxidative
Conversion of Methane", Chem. Eng. Sci., Vol. 57, 4595-4604 (2002).

Lim, S.Y., Park, B., Hung, F., Sahimi, M., Tsotsis, T.T., "Design Issues of Pervaporation Membrane Reactors for
Esterification", Chem. Eng. Sci., Vol. 57, 4933-4946 (2002).

Lin, Y.-M., Rei, M.-H., "Study on the Hydrogen Production from Methanol Steam Reforming in Supported Palladium
Membrane Reactor", Catal. Today, Vol. 67, 77-84 (2001).

Lin, Y.-M., Liu, S.-L., Chuang, C.-H., Chu, Y.-T., "Effect of Incipient Removal of Hydrogen Through Palladium
Membrane on the Conversion of Methane Steam Reforming. Experimental and Modeling", Catal. Today, Vol. 82,
127-139 (2003).

Liu, B., Wu, G., Niu, G., Deng, J.-F., "Rh-Modified Alumina Membranes: Preparation, Characterization and Reaction
Studies", Appl. Catal. A, Vol. 185, 1-10 (1999).

Liu, B., Lian, P.Y., Zhao, X.H., "A Novel NiP-Cu Composite Membrane Reactor for Catalytic Dehydrogenation of
Ethanol", Sep. Purif. Technol., Vol. 32, 281-287 (2003).

Liu, S., Tan, X., Li, K., Hughes, R., "Methane Coupling Using Catalytic Membrane Reactors", Catal. Rev. - Sci. Eng.,
Vol. 43, 147-198 (2001).

Lu, Y., Dixon, A.G., Moser, W.R., Ma, Y.H., "Oxidative Coupling of Methane in a Modified (-Alumina Membrane
Reactor", Chem. Eng. Sci., Vol. 55, 4901-4912 (2000a).

Lu, Y., Dixon, A.G., Moser, W.R., Ma, Y.H., Balachandran, U., "Oxidative Coupling of Methane Using
Oxygen-Permeable Dense Membrane Reactors", Catal. Today, Vol. 56, 297-305 (2000b).

Lu, Y., Dixon, A.G., Moser, W.R., Ma, Y.H., Balachandran, U., "Oxygen-Permeable Dense Membrane Reactor for the
Oxidative Coupling of Methane", J. Memb. Sci., Vol. 170, 27-34 (2000c).

Ma, D., Lund, C.R.F., "Assessing High-Temperature Water-Gas Shift Membrane Reactors", Ind. Eng. Chem. Res., Vol.
42, 711-717 (2003).

Madia, G.S., Barbieri, G., Drioli, E., "Theoretical and Experimental Analysis of Methane Steam Reforming in a
Membrane Reactor", Can. J. Chem. Eng., Vol. 77, 698-706 (1999).

Maira, A.J., Lau, W., N, Lee, C.Y., Yue, P.L., Chan, C.K., Yeung, K.L., "Performance of a Membrane-Catalyst for
Photocatalytic Oxidation of Volatile Organic Compounds", Chem. Eng. Sci., Vol. 58, 959-962 (2003).

Maiya, P.S., Anderson, T.J., Mieville, R.L., Dusek, J.T., Picciolo, J.J., Balachandran, U., "Maximizing H2 Production
by Combined Partial Oxidation of CH4 and Water Gas Shift Reaction", Appl. Catal. A, Vol. 196, 65-72 (2000).

Mallada, R., Menendez, M., Santamaria, J., "Use of Membrane Reactors for the Oxidation of Butane to Maleic
Anhydride Under High Butane Concentrations", Catal. Today, Vol. 56, 191-197 (2000a).

Mallada, R., Pedernera, M., Menendez, M., Santamaria, J., "Synthesis of Maleic Anhydride in an Inert Membrane
Reactor. Effect of Reactor Configuration.", Ind. Eng. Chem. Res., Vol. 39, 620-625 (2000b).

Produced by The Berkeley Electronic Press, 2003


32 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

Mallada, R., Menendez, M., Santamaria, J., "On the Favorable Effect of CO2 Addition in the Oxidation of Butane to
Maleic Anhydride Using Membrane Reactors", Appl. Catal. A, Vol. 231, 109-116 (2002).

Marigliano, G., Barbieri, G., Drioli, E., "Effect of Energy Transport on a Palladium-Based Membrane Reactor for
Methane Steam Reforming Process", Catal. Today, Vol. 67, 85-99 (2001).

Masuda, T., Asanuma, T., Shouji, M., Mukai, S.R., Kawase, M., Hashimoto, K., "Methanol to Olefins Using ZSM-5
Zeolite Catalyst Membrane Reactor", Chem. Eng. Sci., Vol. 58, 649-656 (2003).

Miachon, S., Perez, V., Crehan, G., Torp, E.G., Raeder, H., Bredesen, R., Dalmon, J.-A., "Comparison of a Contactor
Catalytic Membrane Reactor with a Conventional Reactor: Example of Wet Air Oxidation", Catal. Today, Vol. 82, 75-81
(2003).

Molinari, R., Mungari, M., Drioli, E., Di Paola, A., Loddo, V., Palmisano, L., Schiavello, M., "Study on a Photocatalytic
Membrane Reactor for Water Purification", Catal. Today, Vol. 55, 71-78 (2000).

Moon, W.S., Park, S.B., "Design Guide of a Membrane for a Membrane Reactor in Terms of Permeability and
Selectivity", J. Memb. Sci., Vol. 170, 43-51 (2000).

Mota, S., Miachon, S., Volta, J.-C., Dalmon, J.-A., "Membrane Reactor for Selective Oxidation of Butane to Maleic
Anhydride", Catal. Today, Vol. 67, 169-176 (2001).

Moustafa, T.M., Elnashaie, S.S.E.H., "Simultaneous Production of Styrene and Cyclohexane in an Integrated Membrane
Reactor", J. Memb. Sci., Vol. 178, 171-184 (2000).

Nekhamkina, O., Rubinstein, B.Y., Sheintuch, M., "Spatiotemporal Patterns in Thermokinetic Models of Cross-Flow
Reactors", AIChE J., Vol. 46, 1632-1640 (2000).

Neomagus, H.W.J.P., Saracco, G., Wessel, H.F.W., Versteeg, G.F., "The Catalytic Combustion of Natural Gas in a
Membrane Reactor with Separate Feed of Reactants", Chem. Eng. J., Vol. 77, 165-177 (2000).

Niwa, S., Eswaramoorthy, M., Nair, J., Raj, A., Itoh, N., Shoji, H., Namba, T., Mizukami, F., "A One-Step Conversion
of Benzene to Phenol with a Palladium Membrane", Science, Vol. 295, 105-107 (2002).

Nomura, M., Meester, B., Schoonman, J., Kapteijn, F., Moulijn, J.A., "Preparation of Thin Porous Titania Films on
Stainless Steel Substrates for Heat Exchange (HEX) Reactors", Sep. Purif. Technol., Vol. 32, 387-395 (2003).

Omorjan, R.P., Paunovic, R.N., Tekic, M.N., "A Discussion of Maximal Extent of an Isothermal Reversible Gas Phase
Reaction in Double-Membrane Reactors", Chem. Eng. Process., Vol. 38, 355-363 (1999).

Onstot, W.J., Minet, R.G., Tsotsis, T.T., "Design Aspects of Membrane Reactors for Dry Reforming of Methane for the
Production of Hydrogen", Ind. Eng. Chem. Res., Vol. 40, 242-251 (2001).

Pârvulescu, V., Constantin, C., Popescu, G., Su, B.L., "Oxidation of the Aromatic Hydrocarbons in the Nickelsilicate
Membrane Reactor", paper P18 in Proc. 7th Int. Conf. on Inorganic Membranes, Dalian, China, June 23-26 (2002).

Paturzo, L., Basile, A., "Methane Conversion to Syngas in a Composite Palladium Membrane Reactor with Increasing
Number of Pd Layers", Ind. Eng. Chem. Res., Vol. 41, 1703-1710 (2002).

Paturzo, L., Gallucci, F., Basile, A., Vitulli, G., Pertici, P., "An Ru-Based Catalytic Membrane Reactor for Dry
Reforming of Methane - Its Catalytic Performance Compared with Tubular Packed Bed Reactors", Catal. Today, Vol.
82, 57-65 (2003).

Pedernera, M., Mallada, R., Menendez, M., Santamaria, J., "Simulation of an Inert Membrane Reactor for the Synthesis
of Maleic Anhydride", AIChE J., Vol. 46, 2489-2498 (2000).

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 33

Pedernera, M., Alfonso, M.J., Menendez, M., Santamaria, J., "Simulation of a Catalytic Membrane Reactor for the
Oxidative Dehydrogenation of Butane", Chem. Eng. Sci., Vol. 57, 2531-2544 (2002).

Piera, E., Tellez, C., Coronas, J., Menendez, M., Santamaria, J., "Use of Zeolite Membrane Reactors for Selectivity
Enhancement: Application to the Liquid-Phase Oligomerization of i-Butene", Catal. Today, Vol. 67, 127-138 (2001).

Prabhu, A.K., Oyama, S.T., "Highly Hydrogen Selective Ceramic Membranes: Application to the Transformation of
Greenhouse Gases", J. Memb. Sci., Vol. 176, 233-248 (2000).

Prasad, P., Elnashaie, S.S.E.H., "Novel Circulating Fluidized-Bed Membrane Reformer for the Efficient Production of
Ultraclean Fuels from Hydrocarbons", Ind. Eng. Chem. Res., Vol. 41, 6518-6527 (2002).

Raeder, H., Bredesen, R., Crehan, G., Miachon, S., Dalmon, J.-A., Pintar, A., Levec, J., Torp, E.G., "A Wet Air
Oxidation Process Using a Catalytic Membrane Contactor", Sep. Purif. Technol., Vol. 32, 349-355 (2003).

Ramos, R., Menendez, M., Santamaria, J., "Oxidative Dehydrogenation of Propane in an Inert Membrane Reactor",
Catal. Today, Vol. 56, 239-245 (2000).

Raybold, T.M., Huff, M.C., "Oxidation of Isobutane Over Supported Noble Metal Catalysts in a Palladium Membrane
Reactor", Catal. Today, Vol. 56, 35-44 (2000).

Raybold, T.M., Huff, M.C., "Analyzing Enhancement of CO2 Reforming of CH4 in Pd Membrane Reactors", AIChE
J., Vol. 48, 1051-1061 (2002).

Reif, M., Dittmeyer, R., "Porous, Catalytically Active Ceramic Membranes for Gas-Liquid Reactions: A Comparison
Between Catalytic Diffuser and Forced Through Flow Concept", Catal. Today, Vol. 82, 3-14 (2003).

Ren, J.-Y., Fan, Y., Egolfopoulos, F.N., Tsotsis, T.T., "Membrane-Based Reactive Separations for Power Generation
Applications: Oxygen Lancing", Chem. Eng. Sci., Vol. 58, 1043-1052 (2003).

Ritchie, J.T., Richardson, J.T., Luss, D., "Ceramic Membrane Reactor for Synthesis Gas Production", AIChE J., Vol.
47, 2092-2101 (2001).

Rival, O., Grandjean, B.P.A., Guy, C., Sayari, A., Larachi, F., "Oxygen-Free Methane Aromatization in a Catalytic
Membrane Reactor", Ind. Eng. Chem. Res., Vol. 40, 2212-2219 (2001).

Salomon, M.A., Coronas, J., Menendez, M., Santamaria, J., "Synthesis of MTBE in Zeolite Membrane Reactors", Appl.
Catal. A, Vol. 200, 201-210 (2000).

Sammells, A.F., Schwartz, M., Mackay, R.A., Barton, T.F., Peterson, D.R., "Catalytic Membrane Reactors for
Spontaneous Synthesis Gas Production", Catal. Today, Vol. 56, 325-328 (2000).

Sanchez Marcano, J.G., Tsotsis, T.T., Catalytic Membranes and Membrane Reactors, Wiley VCH, Weinheim, 2002.

Sang, C., Chang, C.-C., Lund, C.R.F., "Preliminary Assessment of Membrane Reactors as a Means to Improve the
Selectivity of Methylamine Synthesis", Ind. Eng. Chem. Res., Vol. 38, 4552-4562 (1999).

Saracco, G., Neomagus, H.W.J.P., Versteeg, G.F., van Swaaij, W.P.M., "High-Temperature Membrane Reactors:
Potential and Problems", Chem. Eng. Sci., Vol. 54, 1997-2017 (1999).

Saracco, G., Specchia, V., "Catalytic Combustion of Propane in a Membrane Reactor with Separate Feed of Reactants.
IV Transition from the Kinetics- to the Transport-Controlled Regime", Chem. Eng. Sci., Vol. 55, 3979-3989 (2000).

Schäfer, R., Noack, M., Kölsch, P., Thomas, S., Seidel-Morgenstern, A., Caro, J., "Development of a H2-Selective
SiO2-Membrane for the Catalytic Dehydrogenation of Propane", Sep. Purif. Technol., Vol. 25, 3-9 (2001).

Produced by The Berkeley Electronic Press, 2003


34 International Journal of Chemical Reactor Engineering Vol. 1 [2003], Review R6

Schäfer, R., Noack, M., Kölsch, P., Stöhr, M., Caro, J., "Comparison of Different Catalysts in the Membrane-Supported
Dehydrogenation of Propane", Catal. Today, Vol. 82, 15-23 (2003).

Schramm, O., Seidel-Morgenstern, A., "Comparing Porous and Dense Membranes for the Application in Membrane
Reactors", Chem. Eng. Sci., Vol. 54, 1447-1453 (1999).

Shao, Z., Dong, H., Xiong, G., Cong, Y., Yang, W., "Performance of a Mixed-Conducting Ceramic Membrane Reactor
with High Oxygen Permeability for Methane Conversion", J. Memb. Sci., Vol. 183, 181-192 (2001).

She, Y., Han, J., Ma, Y.H., "Palladium Membrane Reactor for the Dehydrogenation of Ethylbenzene to Styrene", Catal.
Today, Vol. 67, 43-53 (2001).

Sirkar, K.K., Shanbhag, P.V., Kovvali, A.S., "Membrane in a Reactor: A Functional Perspective", Ind. Eng. Chem. Res.,
Vol. 38, 3715-3737 (1999).

Sousa, J.M., Cruz, P., Magalhães, F.D., Mendes, A., "Modeling Catalytic Membrane Reactors Using an Adaptive
Wavelet-Based Collocation Method", J. Memb. Sci., Vol. 208, 57-68 (2002).

Stoukides, M., "Solid-Electrolyte Membrane Reactors: Current Experience and Future Outlook", Catal. Rev. - Sci. Eng.,
Vol. 42, 1-70 (2000).

Tagawa, T., Moe, K.K., Ito, M., Goto, S., "Fuel Cell Type Reactor for Chemicals-Energy Co-Generation", Chem. Eng.
Sci., Vol. 54, 1553-1557 (1999).

Tan, X., Li, K., "Investigation of Novel Membrane Reactors for Removal of Dissolved Oxygen from Water", Chem. Eng.
Sci., Vol. 55, 1213-1224 (2000).

Tanaka, K., Yoshikawa, R., Ying, C., Kita, H., Okamoto, K.-I., "Application of Zeolite Membranes to Esterification
Reactions", Catal. Today, Vol. 67, 121-125 (2001).

Tavolaro, A., Drioli, E., "Zeolite Membranes", Adv. Mater., Vol. 11, 975-996 (1999).

Tellez, C., Menendez, M., Santamaria, J., "Simulation of an Inert Membrane Reactor for the Oxidative Dehydrogenation
of Butane", Chem. Eng. Sci., Vol. 54, 2917-2925 (1999).

Tennison, S., "Current Hurdles in the Commercial Development of Inorganic Membrane Reactors", Membrane
Technology, Vol. 2000, No. 128, 4-9 (2000).

Thampan, T., Malhotra, S., Zhang, J., Datta, R., "PEM Fuel Cell as a Membrane Reactor", Catal. Today, Vol. 67, 15-32
(2001).

Tosti, S., Basile, A., Chiappetta, G., Rizzello, C., Violante, V., "Pd-Ag Membrane Reactors for Water Gas Shift
Reaction", Chem. Eng. J., Vol. 93, 23-30 (2003).

Tsuru, T., Kan-no, T., Yoshioka, T., Asaeda, M., "A Photocatalytic Membrane Reactor for Gas-Phase Reactions Using
Porous Titanium Oxide Membranes", Catal. Today, Vol. 82, 41-48 (2003).

van Dyk, L., Miachon, S., Lorenzen, L., Torres, M., Fiaty, K., Dalmon, J.-A., "Comparison of Microporous MFI and
Dense Pd Membrane Performances in an Extractor-Type CMR", Catal. Today, Vol. 82, 167-177 (2003).

van de Graaf, J.M., Zwiep, M., Kapteijn, F., Moulijn, J.A., "Application of a Silicalite-1Membrane Reactor in Metathesis
Reactions", Appl. Catal. A, Vol. 178, 225-241 (1999a).

van de Graaf, J.M., Zwiep, M., Kapteijn, F., Moulijn, J.A., "Application of a Zeolite Membrane Reactor in the
Metathesis of Propene", Chem. Eng. Sci., Vol. 54, 1441-1445 (1999b).

http://www.bepress.com/ijcre/vol1/R6
Dixon: Recent Research in Catalytic Inorganic Membrane Reactors 35

Vincent, M.J., Gonzalez, R.D., "Selective Hydrogenation of Acetylene Through a Short Contact Time Reactor", AIChE
J., Vol. 48, 1257-1267 (2002).

Wang, H., Cong, Y., Yang, W., "High Selectivity of Oxidative Dehydrogenation of Ethane to Ethylene in an Oxygen
Permeable Membrane Reactor", Chem. Commun., 1468-1469 (2002).

Wang, H., Cong, Y., Yang, W., "Investigation on the Partial Oxidation of Methane to Syngas in a Tubular
Ba0.5Sr0.5Co0.8Fe0.2O3-* membrane reactor", Catal. Today, Vol. 82, 157-166 (2003).

Wang, L., Murata, K., Inaba, M., "Production of Pure Hydrogen and More Valuable Hydrocarbons from Ethane on a
Novel Highly Active Catalyst System with a Pd-Based Membrane Reactor", Catal. Today, Vol. 82, 99-104 (2003).

Weyten, H., Luyten, J., Keizer, K., Willems, L., Leysen, R., "Membrane Performance: The Key Issues for
Dehydrogenation Reactions in a Catalytic Membrane Reactor", Catal. Today, Vol. 56, 3-11 (2000).

Wolfrath, O., Kiwi-Minsker, L., Renken, A., "Novel Membrane Reactor with Filamentous Catalytic Bed for Propane
Dehydrogenation", Ind. Eng. Chem. Res., Vol. 40, 5234-5239 (2001).

Wu, J.C.-S., "Feasibility of Manufacturing Hydrogen and Styrene Through the Use of Porous Ceramic Membranes", Ind.
Eng. Chem. Res., Vol. 38, 4491-4495 (1999).

Xue, D., Chen, H., Wu, G.-H., Deng, J.-F., "Amorphous Ni-B Alloy Membrane: Preparation and Application in Ethanol
Dehydrogenation", Appl. Catal. A, Vol. 214, 87-94 (2001).

Xue, E., Ross, J., "The Use of Membrane Reactors for Catalytic n-Butane Oxidation to Maleic Anhydride with a
Butane-Rich Feed", Catal. Today, Vol. 61, 3-8 (2000).

Xue, E., Ross, J.R.H., Mallada, R., Menendez, M., Santamaria, J., Perregard, J., Nielsen, P.E.H., "Catalytic Oxidation
of Butane to Maleic Anhydride Enhanced Yields in the Presence of CO2 in the Reactor Feed", Appl. Catal. A, Vol. 210,
271-274 (2001).

Zhu, B., Li, H., Yang, W., "AgBiVMo Oxide Catalytic Membrane for Selective Oxidation of Propane to Acrolein",
Catal. Today, Vol. 82, 91-98 (2003).

Zhu, D.C., Xu, X.Y., Feng, S.J., Liu, W., Chen, C.S., "La2NiO4 Tubular Membrane Reactor for Conversion of Methane
to Syngas", Catal. Today, Vol. 82, 151-156 (2003).

Produced by The Berkeley Electronic Press, 2003

S-ar putea să vă placă și