You are on page 1of 56

This article was downloaded by: [McGill University Library]

On: 20 November 2012, At: 08:56


Publisher: Taylor & Francis
Informa Ltd Registered in England and Wales Registered Number: 1072954 Registered office: Mortimer
House, 37-41 Mortimer Street, London W1T 3JH, UK

Catalysis Reviews: Science and Engineering


Publication details, including instructions for authors and subscription information:
http://www.tandfonline.com/loi/lctr20

ONBOARD FUEL CONVERSION FOR HYDROGEN-FUEL-


CELL-DRIVEN VEHICLES
a b
David L. Trimm & Z. Ilsen Önsan
a
University of New South Wales, Sydney, 2052, Australia
b
Department of Chemical Engineering, Bogaziçi University, 80815, Bebek, Istanbul,
Turkey
Version of record first published: 02 Nov 2011.

To cite this article: David L. Trimm & Z. Ilsen Önsan (2001): ONBOARD FUEL CONVERSION FOR HYDROGEN-FUEL-CELL-
DRIVEN VEHICLES, Catalysis Reviews: Science and Engineering, 43:1-2, 31-84

To link to this article: http://dx.doi.org/10.1081/CR-100104386

PLEASE SCROLL DOWN FOR ARTICLE

Full terms and conditions of use: http://www.tandfonline.com/page/terms-and-conditions

This article may be used for research, teaching, and private study purposes. Any substantial or systematic
reproduction, redistribution, reselling, loan, sub-licensing, systematic supply, or distribution in any form to
anyone is expressly forbidden.

The publisher does not give any warranty express or implied or make any representation that the contents
will be complete or accurate or up to date. The accuracy of any instructions, formulae, and drug doses
should be independently verified with primary sources. The publisher shall not be liable for any loss, actions,
claims, proceedings, demand, or costs or damages whatsoever or howsoever caused arising directly or
indirectly in connection with or arising out of the use of this material.
CATALYSIS REVIEWS, 43(1&2), 31–84 (2001)

ONBOARD FUEL CONVERSION FOR


Downloaded by [McGill University Library] at 08:56 20 November 2012

HYDROGEN-FUEL-CELL-DRIVEN VEHICLES

David L. Trimm1,* and Z. Ilsen Önsan2

1
School of Chemical Engineering and Industrial Chemistry,
University of New South Wales, Sydney 2052, Australia
2
Department of Chemical Engineering, Bogaziçi University, 80815
Bebek, Istanbul, Turkey

I. INTRODUCTION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32
II. BASIC REACTIONS . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
III. INDIRECT PARTIAL OXIDATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
A. Total Oxidation to Supply Heat . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
B. Steam Reforming . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39
C. Combined Reactions: Indirect Partial Oxidation . . . . . . . . . . . . . . . . . . . . . 46
IV. DIRECT PARTIAL OXIDATION . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
V. LOW-TEMPERATURE REMOVAL OF CARBON MONOXIDE . . . . . . . 56
A. Low-Temperature CO Oxidation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
B. Low-Temperature CO Oxidation in H 2-Rich Gas Streams . . . . . . . . . . . 65
VI. WHICH FUEL? WHICH SYSTEM? . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71
ACKNOWLEDGMENT . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
REFERENCES . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78

*Corresponding author.

31

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


ORDER REPRINTS

32 TRIMM AND ÖNSAN

ABSTRACT

Increasingly stringent legislation controls emissions from internal combustion


engines to the point where alternative power sources for vehicles are necessary.
The hydrogen fuel cell is one promising option, but the nature of the gas is such
that the conversion of other fuels to hydrogen on board the vehicle is necessary.
The conversion of methanol, methane, propane, and octane to hydrogen
is reviewed. A combination of oxidation and steam reforming (indirect partial
oxidation) or direct partial oxidation are the most promising processes. Indirect
partial oxidation involves combustion of part of the fuel to produce sufficient
Downloaded by [McGill University Library] at 08:56 20 November 2012

heat to drive the endothermic steam reforming reaction. Direct partial oxida-
tion is favored only at high temperatures and short residence times but is highly
selective. However, indirect partial oxidation is shown to be the preferred
process for all fuels.
The product gases can be taken through a water–gas shift reactor, but
still retain ⬃2% carbon monoxide, which poisons fuel-cell catalysts. Selective
oxidation is the preferred route to removal of residual carbon monoxide. Low-
temperature oxidation in the absence and presence of an excess of hydrogen
is reviewed. Au-based catalysts show much promise, but precious metal cata-
lysts such as Pt/zeolite have some advantages.

Key Words: Hydrogen; Fuel cell; Onboard hydrogen generation; Hydrocarbon


partial oxidation.

I. INTRODUCTION

The development of fuel-efficient engines that produce less pollutants has


been a major objective for many years. Regulatory requirements have been and
will be of major importance in defining standards (1). Increasingly stringent legis-
lation directs attention at factors such as cold-start emissions (2) and sulfur levels
in fuel (3).
The response of engine manufacturers, oil refiners, and catalyst industries
to this legislative pressure has been remarkable. A steady increase in fuel economy
and a decrease in unwanted emissions has been achieved. However, it is now
believed that a new approach to transport motive power may be required.
Most promise has been identified in four key areas:
• Direct-injection compression ignition engines
• Hybrid electric vehicles
• Fuel-cell-powered electric vehicles
• The use of lightweight materials
Some of these concepts are already commercially available. Hybrid electric vehi-
cles are available on the market, although at a somewhat subsidized price. How-
ever, it is the development of fuel-cell-powered vehicles and of compression igni-
tion engines that excites most interest.
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 33

The most promising fuel-cell technology for transport applications appears


to be the Polymer Electrolyte Membrane Fuel Cell (PEMFC) fueled by hydrogen
(4–6). Hydrogen is ionized to protons and electrons at the anode of the cell. Pro-
tons migrate through a membrane to the cathode where reaction to produce water
occurs (7,8).
Overall, fuel is oxidized electrochemically and at low temperatures (⬃80°C),
which increases efficiency and avoids high-temperature generation of nitrogen
oxides. Each cell produces only ⬃0.6–0.7 eV, so an assembly of cells is required
(8). Cells may be small and assemblies can be produced with high-power densities
Downloaded by [McGill University Library] at 08:56 20 November 2012

(⬃1 kW/L). Improvements in cell design continue, with an acceptable final cost
of the order of US$ 50–75/kW—a value which appears to be within the bounds
of possibility.
Both hydrogen and methanol have been considered as fuels, but the kinetics
of methanol conversion are slow, and catalyst deactivation occurs. As a result,
hydrogen fuel cells are preferred (4,6,8,9), but these, in turn, present obvious prob-
lems with generation, storage, and distribution. Pressurized vessels or metal hy-
drides in a vehicle occupy space, increase weight, and decrease fuel efficiency.
They may be practicable in a bus or lorry (4,5,9) but are unlikely to be used in a
car. Perhaps one possibility—the storage of hydrogen in carbon nanofibers (10)—
offers a feasible alternative, but the production and distribution of the gas remains
a difficulty.
As a result, attention has been focused on the design and operation of com-
pact and efficient devices designed to generate hydrogen on board a vehicle. Jamal
and Wyszynski (11) have produced an excellent review of onboard generation of
hydrogen for use as an alternative fuel for spark ignition engines—an application
that appears to have much promise. For fuel-cell applications, however, a supply
of pure hydrogen is desired and more selective processing is required. This usually
involves a mixture of oxidation and steam reforming or partial oxidation, and it
is this conversion of fuels to a level of purity of hydrogen acceptable to fuel cells
that is the subject of the present review.
By far, the most work reported has been focused on methanol conversion
and successful onboard operation has been demonstrated (4,9,12,13). This work
is reviewed in the next section. Methanol, however, is not readily available at
service stations and the fuel would require a large investment in production, stor-
age, and distribution. As a result, the production of hydrogen from natural gas,
LPG (liquefied petroleum gases, mainly containing ethane and propane), gasoline,
and diesel is of interest and is also considered.

II. BASIC REACTIONS

Onboard processing of liquid fuels is regarded as the most promising method


of supplying hydrogen for fuel cells mounted in a vehicle. Devices have been
developed that offer optimal energy density and low volumes.
ORDER REPRINTS

34 TRIMM AND ÖNSAN

Reactions such as thermal reforming, steam reforming, carbon dioxide re-


forming, partial oxidation, and autothermal reforming have been suggested (11),
but nearly all processes considered for commercial exploitation are based on
autothermal reforming or partial oxidation. To some extent, this is dictated by the
preferred fuel.
The production of hydrogen from fuels can result from direct decomposition,
steam reforming, or partial oxidation (14):
CH 4 ⫽ C ⫹ 2H 2 , ∆H°298 ⫽ 74.8 kJ/mol (1)
Downloaded by [McGill University Library] at 08:56 20 November 2012

CH 4 ⫹ H 2 O ⫽ CO ⫹ 3H 2 , ∆H°298 ⫽ 206.2 kJ/mol (2)


CH 4 ⫹ 1/2O 2 ⫽ CO ⫹ 2H 2 , ∆H°298 ⫽ ⫺35.7 kJ/mol (3)
Carbon monoxide may be used to produce more hydrogen via the water–gas shift
reaction:
CO ⫹ H 2 O ⫽ CO 2 ⫹ H 2 , ∆H°298 ⫽ ⫺41.2 kJ/mol (4)
Direct decomposition is energy demanding and produces coke. As a result, the
reaction is not a serious contender for hydrogen production.
Steam reforming may be applied both to alcohols and to hydrocarbons (15):
CH 3 OH ⫹ H 2 O ⫽ CO 2 ⫹ 3H 2 , ∆H°298 ⫽ 49.6 kJ/mol (5)
CH 4 ⫹ H 2 O ⫽ CO ⫹ 3H 2 (2)
The reaction is endothermic and heat must be supplied to the system. For mobile
applications, this is usually done by combusting part of the fuel (12,14):
CH 4 ⫹ 2O 2 ⫽ CO 2 ⫹ 2H 2 O, ∆H°298 ⫽ ⫺802 kJ/mol (6)
The combination of total oxidation, steam reforming, and the water–gas shift
reaction will be referred to as indirect partial oxidation.
The oxidation of part of the fuel provides sufficient heat to drive the steam
reforming reaction and produces some water [reaction (6)]. This is usually insuf-
ficient for steam reforming/water–gas shift requirements and further water must
be added to the system (15).
The steam reforming of hydrocarbons to produce hydrogen is favored at
higher temperatures, where coke formation can be a problem (14,15). Sufficient
water must also be added to minimize coking, although a choice of catalyst may
reduce the amount of coke to be gasified (15). With very heavy fuels, it may be
preferable to move to noncatalytic partial combustion (14,16).
Direct partial oxidation also produces hydrogen, but the carbon monoxide :
hydrogen ratio is lower:
CH 4 ⫹ 1/2 O 2 ⫽ CO ⫹ 2H 2 (3)
Again, the water–gas shift reaction can be used to produce more hydrogen:
CO ⫹ H 2 O ⫽ CO 2 ⫹ H 2 (4)
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 35

Direct partial oxidation is thermodynamically favored only at high temperatures


(⬃750°C⫹) and short residence times (14). Longer residence times or postreactor
reactions may result in indirect partial oxidation or in carbon dioxide reforming:

CH 4 ⫹ CO 2 ⫽ 2CO ⫹ 2H 2 , ∆H°298 ⫽ ⫺247 kJ/mol (7)

Direct partial oxidation is much faster than any of the indirect reactions, and re-
sponse times to changes in fuel supply are shorter.
There is some question of whether direct partial oxidation is totally a cata-
lytic reaction. It has now been established that a catalyst and the pore structure
Downloaded by [McGill University Library] at 08:56 20 November 2012

of a support does influence product spectra (14), indicating the importance of


catalysis to the process.
Both indirect and direct partial oxidation are considered in detail below. It
should be noted that the distinction between indirect and direct partial oxidation
is not always clarified in the literature.
Catalysts used in both processes are subject to catalyst deactivation, particu-
larly by poisoning or thermal reorganization (17). An additional problem arises
with fuel cells, in that levels of carbon monoxide above about 10–100 ppm will
poison the catalyst in the cell (4,7). As a result, any fuel-conversion system must
be designed to reduce carbon monoxide to below these levels.
In order to meet the requirements of high energy density and small size,
oxidation and steam reforming functions may be combined or novel reactor geom-
etries developed. For the purposes of this review, however, it is convenient first
to consider the individual reactions and then to consider possibilities of combina-
tion. Fuels to be considered include methanol, methane, LPG, gasoline, and diesel.
The safety of the operation is also important, in that the hazards to be consid-
ered include those associated with the fuels and with hydrogen, as well as those
possibly arising from pyrophoric catalysts. Such issues are not considered in this
review.

III. INDIRECT PARTIAL OXIDATION


A. Total Oxidation to Supply Heat

Hydrogen may be produced by steam reforming or by partial oxidation, but


neither reaction is significant at room temperature. As a result, it is necessary to
heat the reactor to an appropriate operating temperature—which will depend on
the activity of a catalyst and the reactivity of a fuel. Both methanol and hydrogen
will oxidize at room temperature over a precious metal catalyst (13,18), but hydro-
carbon oxidation requires higher temperatures (19,20).
Electric heating of a system is possible, but significant battery capacity is
required and fuel must be consumed to generate electric charge. As a result, cata-
lytic combustion of part of the fuel is widely accepted as the best means of heating
the system.
ORDER REPRINTS

36 TRIMM AND ÖNSAN

Although total oxidation produces most heat, partial oxidation may be impor-
tant as the concentration of oxygen decreases:
CH 3 OH ⫹ 3/2 O 2 ⫽ CO 2 ⫹ 2H 2 O, ∆H°298 ⫽ ⫺677 kJ/mol (8)
CH 3 OH ⫹ 1/2 O 2 ⫽ CO 2 ⫹ 2H 2 , ∆H°298 ⫽ ⫺152 kJ/mol (9)
This may not produce problems, provided the partial oxidation products can be
dealt with in subsequent processing; for example,
CO ⫹ H 2 O ⫽ CO 2 ⫹ H 2 (4)
Downloaded by [McGill University Library] at 08:56 20 November 2012

However, the higher heat generation of total oxidation makes it the preferred reac-
tion.
Catalytic combustion of methanol and of products originating from metha-
nol-fueled vehicles has been studied over Pt wire (19,21), over supported Rh, Pd,
Pt, and Ag, and over Cu–Cr-based catalysts (22,23). Perhaps the most complete
study was carried out by Jiang et al. (13,24) in the context of fuel-cell operation.
Catalytic combustion was studied over supported Pt and Cu catalysts, in the ab-
sence and presence of water. The onset of significant oxidation was assessed in
terms of the temperature at which 10% of methanol was oxidized (light-off tem-
perature) and the maximum bed temperature was also recorded.
Considerable differences were observed between oxidized and reduced cata-
lyst. Both Pt- and Cu-based catalysts only became active at ⬃200°C if in the
oxidized state. Both catalysts oxidized methanol at room temperature if they were
prereduced. Prereduced copper is pyrophoric and could not be used. Prereduced
Pt was found to be difficult to reoxidize and, once reduced, gave reproducible
room-temperature light off. Maximal bed temperatures were sufficient for steam
reforming.
Jiang (12,24) observed only total oxidation, but McCabe and Mitchell (22)
have reported partial oxidation at lower temperatures.
The catalytic oxidation of methanol will begin at low temperatures over a
supported Pt catalyst and the Pt loading was found not to be significant. What
was also obvious was that the final bed temperature could rise above 330°C, a
value at which Cu-based catalysts start to sinter. Copper-based catalysts are often
used to facilitate steam reforming of methanol (13).
The catalytic oxidation of hydrogen is well documented (25–27). Ma (18)
has shown that light off on Pt/δ-Al 2 O 3 occurs at room temperature (H 2 :O 2 ⫽
3.7:1), the bed temperature rising to values dependent on the concentration of
the reagents. Again, the maximal bed temperature could be greater than 350°C.
Carbon monoxide is also a product of steam reforming, and oxidation can
be initiated at room temperature. The process is discussed in some detail in Sec-
tion VI.
In contrast, the catalytic oxidation of saturated hydrocarbons is considerably
more difficult. Precious metals are known to be highly active, with Hiam et al.
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 37

(19) establishing activity patterns many years ago. Metal oxide (28–30) and, par-
ticularly, manganese-oxide-based catalysts (31) are also active for catalytic com-
bustion.
Despite the fact that considerable attention has been focused on low light-
off temperature catalysts for vehicle emission control, significant oxidation of hy-
drocarbons is observed only above about 170–200°C (19,32). Supported precious
metals (20,33,34) and metal oxides (19,28,30,33) have been examined, with best
performance depending on the support. In general, precious metals are preferred,
with Pd favored for methane oxidation (20,27) and Pt for other fuels (19,30).
Downloaded by [McGill University Library] at 08:56 20 November 2012

The overall reactivity depends also on the support. Pd-exchanged zeolites


are more active than supported Pd, as a result of the higher metal surface area.
Yazawa et al. (35) find that Pd supported on a moderate acid support such as
silica–alumina is best for propane combustion. Ma (18) reported that the light-
off temperature for Pt/δ-Al 2 O 3 was ⬃310°C, whereas Baldwin and Burch (36)
report a value of 267°C for Pd/Al 2 O 3 .
By far the greatest body of work has been focused on vehicle emission con-
trol catalysts (32). Using ceria and ceria–zirconia supports (37), light-off values
of ⬃170°C (fresh) and 200°C (used) have been reported. Generally, the hydrocar-
bons combusted at this temperature do not include methane, but Epling and
Hoflund (38) have recently reported that a Pd/ZrO 2 catalyst will light off methane
at 200°C. This low value is surprising, but Carstens et al. (39) report a similar
value. Fujimoto et al. (40), however, report PdO/ZrO 2 to be active for methane
oxidation only at 280°C. Light off at any of these temperatures is very interesting,
although Epling and Hoflund (38) report that catalytic activity decreases somewhat
after high-temperature exposure.
Ma (18) has also explored the reactivity of methane, ethane, and propane
over Pt/δ-Al 2 O 3 and Ni/MgO in the context of fuel-cell operations. The Pt catalyst
was found to be much more active than Ni, although once the Ni was reduced,
comparable activities could be achieved. The light-off temperatures are summa-
rized in Table 1, from which it is seen that reactivity increases with carbon chain
length. However, the light-off temperature of octane was ⬃200°C over the same
catalyst (42), so further improvement is unlikely.
The light-off temperatures decrease somewhat as the fuel : oxygen ratio in-
creases. However, as seen from the ratio of carbon monoxide to carbon dioxide
(Table 1), the possibility of partial oxidation or water–gas shift increases as meth-
ane increases. This is also reflected to some extent in the bed temperatures reached.
Thus, light off, an overall measure of reactivity, also reflects the importance of
the different reactions.
Ma (18) demonstrated that it was possible to reach hydrocarbon light-off
temperatures by oxidizing methanol. An obvious consequence is to investigate
the light off of a mixture of methanol and hydrocarbon. Studies of the light off
of methanol, iso-octane, and mixtures of the two fuels showed that the light-off
temperature of methanol was not affected by the presence of octane (42). Thus,
Downloaded by [McGill University Library] at 08:56 20 November 2012

38

Table 1. Light-Off Temperatures (TL ), Bed Temperatures (Tmax ), Maximum Conversion (X max ), and Carbon Oxides Ratio Achieved at Different Hydrocarbon-
ORDER

to-Oxygen Ratios over Pt/Al 2 O 3

Hydrocarbon Methane Ethane Propane


HC/O 2 0.27 0.9 2.53 5.04 0.89 1.46 2.78 3.79 0.48 0.97 1.33 1.86
TL (°C) 451 368 370 316 242 229 215 207 185 178 161 152
REPRINTS

Tmax (°C) 618 669 636 548 653 533 463 438 664 626 577 461
X max (%) 100 74.5 26.1 12.5 60.4 28.3 17.2 10.8 87.3 62.3 30.2 13.6
CO/CO 2 0 1 0.25 0.13
TRIMM AND ÖNSAN
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 39

the use of mixtures of fuel allows total oxidation to start at low temperatures, the
heat produced increasing the temperature to the point where less reactive fuels
light off.
It should be noted that light-off temperatures are not always an exact measure
of reactivity (43) and that values may be affected by the presence of more than one
gas (44). Nonetheless, they provide a reasonable estimate of the energy required to
initiate a given oxidation reaction.
From the perspective of onboard production of hydrogen, it is clear that the
total oxidation of hydrogen or methanol over a Pt-based catalyst will start at room
Downloaded by [McGill University Library] at 08:56 20 November 2012

temperature and will supply the heat needed for steam reforming. The oxidation
of hydrocarbons, on the other hand, will not start until higher temperatures and
some means must be used to increase the catalyst bed from ambient to light-off
temperatures. The oxidation of hydrogen stored from a previous run is an obvious
source of heat.

B. Steam Reforming

The production of hydrogen from other fuels involves steam reforming, an


endothermic reaction. The necessary heat is generated by total oxidation (see Sec-
tion III.A), but it is then necessary to consider hydrogen production from alterna-
tive fuels. The fuel most often considered is methanol.

1. Methanol

The steam reforming of methanol to produce hydrogen or synthesis gas has


been of interest for some time, although detailed studies of the mechanism and
kinetics of the reaction have been completed only recently.
The endothermic reaction is catalyzed by a range of metal- and metal-oxide-
based catalysts. By analogy with methanol synthesis, many workers have focused
on copper-based catalysts, with Cu–Zn (45,46), Cu–Al (47–50), Cu–Zn–Al (51–
53), Cu–Mn (47), Cu–Si (47,54), Cu–Zn–Al–V, Cu–Zn–Al–Zr, and Cu–Zn–
Al–La (55) being studied. However, Fe–Cr, Zr–Cr, Ni–Al, Pd–Al, Ni–Ti, Ti–
K–Al (56), and supported Pd catalysts (57) have also been used. Base metals such
as Ni, Co, Fe, Mn, Mo, and Cr, together with precious-metal-based catalysts favor
the production of syngas from methanol (58,59), whereas Cu-based catalysts are
preferred when hydrogen is the desired product (54). However, the operating con-
ditions over some base–precious metal catalysts may be such that hydrogen and
carbon dioxide are the major products.
Kobayashi et al. (60) compared the efficiency of physical mixtures of the
oxides of copper with those of Cr, Zn, Al, Sn, and Si, but such mixtures negate
the beneficial effects of additives that produce highly dispersed copper crystallites
on the surface of the catalyst and that are reasonably thermally stable (61,62).
ORDER REPRINTS

40 TRIMM AND ÖNSAN

Cu–ZnO-based catalysts, prepared either by the Raney method (63) or by copre-


cipitation (50,51) have been most widely used, although sintering of such catalysts
at temperatures greater than ⬃330°C (24) is still a major problem. Shimomura
et al. (64) prepared a range of Cu–Zn–Al-based monolithic catalysts promoted
by the oxides of Ca, Zr, Cr, and V, but they did not report on any improvements
in thermal stability. Inui et al. (65) reported that some deactivation could also
occur during methanol steam reforming as a result of the formation of copper
oxides, but this is less important than thermal sintering as a result of probable
reduction of oxides by product hydrogen.
Downloaded by [McGill University Library] at 08:56 20 November 2012

Steam reforming of methanol over copper-based catalysts was originally


thought to involve the decomposition of methanol to carbon monoxide:
CH 3 OH ⫽ CO ⫹ 2H 2 (10)
CO ⫹ H 2 O ⫽ CO 2 ⫹ H 2 (4)
However, it soon became clear that this explanation was too simple. Takahashi
et al. (66) were the first to suggest the formation of methyl formate as a major
intermediate:
2CH 3 OH ⫽ CH 3 OCHO ⫹ 2H 2 (11)
CH 3 OCHO ⫹ H 2 O ⫽ CH 3 OH ⫹ HCOOH (12)
HCOOH ⫽ CO 2 ⫹ H 2 (13)
This suggestion was supported by Jiang et al. (51,52), who noted that the
rate of methanol steam reforming was the same as the rate of methanol dehydroge-
nation reported by Tonner et al. (67).
However, Peppley et al. (53), using a similar catalyst to study the mechanism
and kinetics of methanol steam reforming, found no evidence of adsorbed methyl
formate. DRIFTS studies identified methoxy and formate groups on the catalyst
surface, but no methyl formate or carbon dioxide.
This apparent discrepancy was explained by Fisher and Bell (68), who studied
methanol decomposition on Cu/SiO 2 , ZrO 2 /SiO 2, and Cu/SiO 2 /ZrO 2 . Adsorbed
methoxy groups were observed which reacted to form formate groups in the ab-
sence of methanol in the gas phase. If gas-phase methanol was present, methoxy
reacted to produce methyl formate. Thus, it is clear that both methyl formate and
adsorbed formate species may be present under normal operating conditions.
Both Jiang et al. (51,52) and Peppley et al. (53) identified the dehydrogena-
tion of methoxy groups as the rate-determining step in methanol steam reforming.
However, the importance of carbon monoxide is open to more question. Jiang et
al. found that no carbon monoxide was formed in their system and that the addition
of carbon monoxide to a methanol: steam feed had no effect on the product spectra.
They suggested that the water–gas shift reaction was unimportant, as competitive
adsorption favored methanol rather than carbon monoxide (69).
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 41

In contrast, Peppley et al. (53) found significant amounts of carbon monoxide


even when methanol conversion was incomplete. They suggested that direct de-
composition of methanol could be significant and that the water–gas shift reaction
was important.
The absence or presence of carbon monoxide is hard to reconcile, but the
product is important because fuel cells are poisoned by traces of CO (4,5). Even
in the ‘‘Jiang’’ system, carbon oxides will be involved eventually, once most of
the methanol has been reacted. Thus, inclusion of the water–gas shift equilibria
and kinetics is necessary in any consideration of the system.
Downloaded by [McGill University Library] at 08:56 20 November 2012

Water–gas shift equilibria are well reported (70) and Peppley et al. (53) have
reported the kinetics of the reaction over a Cu/ZnO catalyst (Table 2). Such cata-
lysts are widely used to promote the reaction on an industrial scale (62,70) but
may be pyrophoric if exposed to oxygen. Direct methanol decomposition is much
slower (Table 3) and may or may not be responsible for the production of carbon
monoxide.
The DRIFTS evidence that CO is not adsorbed on the catalyst is hard to
reconcile with the suggestion that catalyzed CO conversion is significant (53). It
may well be that CO adsorption only becomes important once most methanol
has been converted. Any carbon monoxide that is formed can be converted to an
equilibrium value by the water–gas shift reaction, but further reduction of carbon
monoxide is still required for fuel cells. The means of doing this are discussed
later.
It is obvious that the kinetics of the various reactions and the thermodynamic
equilibria are of considerable interest in the context of hydrogen production for
fuel-cell use. The most relevant kinetic expressions are summarized in Table 2.
Some ambiguity exists in establishing values of C i in the equations developed by
Peppley et al. (53).
One intriguing possibility is the steam reforming of methanol over other
catalysts (57,71,72). Supported palladium is perhaps the most active system (71),
with Pd/ZnO offering high activity and selectivity (57,72).
The main difficulty with Cu-based catalysts results from sintering of the
metal at temperatures above about 330°C. This may be a major disadvantage if
part of the methanol is to be combusted in order to raise the catalyst to tempera-
tures suitable for steam reforming (⬃200°C ⫹). Temporary overheating may re-
sult in sintering and irreversible deactivation.
No such difficulties arise with supported Pd, and the catalyst is active at
⬃200°C. Both activity and selectivity to hydrogen improve as the temperature is
increased (72). A selectivity to hydrogen of 96% has been reported over 1% Pd/
ZnO at ⬃270°C.
Comparisons of the activity and selectivity of various supported Pd catalysts
were reported by Iwasa et al. (57). As with transition metal catalysts, carbon mon-
oxide and hydrogen were major products, possibly resulting from direct decompo-
sition of methanol. However, Pd/ZnO catalysts gave excellent selectivity to hydro-
Downloaded by [McGill University Library] at 08:56 20 November 2012

Table 2. Kinetics of Methanol Conversion


42

Reaction Catalyst Kinetic Expression Ref.


⫺1/2
CH 3 OH ⫹ H 2 O ⫽ CO 2 ⫹ 3H 2 Cu–ZnO–Al 2 O 3 (A) K 1 PM K 1/2
2 P H 51,52
r ⫽ k1 1/2 ⫺1/2 2
(1 ⫹ K 1 K 2 P M P H ⫹ K ⫺1/22 P 1/2
H )

Cu–ZnO–Al 2 O 3 (B) PM P 3H PC 53
r ⫽ k2 K4 1⫺ C S1 C S2
P 1/2
冢 冣冢 H K′2 PM PW 冣
⫺1
PM PW 1/2
⫻ 1 ⫹ K4 ⫹ K 5 PC P 1/2
H ⫹ K6 7 PH
1 ⫹ K 1/2
冤冢 冢 冣 P 1/2
H P 1/2
H
冢 冣冣冢 冣冥
ORDER

CO ⫹ H 2 O ⫽ CO 2 ⫹ H 2 Cu–ZnO–Al 2 O 3 (B) PCO PW PH PC 53


r ⫽ k3 K6 1⫺ C 2S1
冢 P 1/2
H
冣冢 K′3 PCO PW 冣
⫺2
PM PW
⫻ 1 ⫹ K 4 1/2 ⫹ K 5 PC P 1/2
H ⫹ K6
冤 PH 冢 冣 P 1/2
H
冢 冣冥
REPRINTS

CH 3 OH ⫽ CO ⫹ 2H 2 Cu–ZnO–Al 2 O 3 (B) PM P 2H PCO 53


r ⫽ k4 K4 1⫺ C S3 C S4
冢 冣冢 冣P 1/2
H K′4 PM
⫺1
PM PW
⫻ 1 ⫹ K4 ⫹ K 6 1/2 1 ⫹ K 71/2 P H1/2
冤冢 冢 冣 冢 冣冣冢 P 1/2
H PH 冣冥
Note: A: BASF S3-85 (32 wt% CuO)
B: BASF K3-110 (40 wt% Cu)
C: carbon dioxide
C S j : concentration of sites
k i : rate constant
K i : adsorption coefficients
M: methanol
r: rate of reaction
TRIMM AND ÖNSAN

W: water
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 43

gen, the reaction proceeding either via methyl formate (57) or via reactive HCHO
species (71). At 350°C and a residence time of 0.47 s, conversion exceeded 99%
and selectivity to hydrogen was 92%.
Amounts of carbon monoxide observed were small, as would be expected
from a methyl-formate-based reaction. However, the concentration was significant
in the context of fuel-cell operation.
Pd/ZnO must certainly be considered as a candidate catalyst for methanol
steam reforming.
Downloaded by [McGill University Library] at 08:56 20 November 2012

2. Hydrocarbons

The steam reforming of light hydrocarbons to produce hydrogen is a well-


established industrial process (70,73) that has been described in considerable de-
tail (70,73,74). As a result, only an outline of the process is presented and the
points considered are mainly relevant to the operation of fuel cells.
Steam reforming is an endothermic reaction which is favored by high tem-
peratures (70,73):
CH 4 ⫹ H 2 O ⫽ CO ⫹ 3H 2 (2)
C 2 H 6 ⫹ 2H 2 O ⫽ 2CO ⫹ 5H 2 , ∆H°298 ⫽ 347.3 kJ/mol (14)
C 3 H 8 ⫹ 3H 2 O ⫽ 3CO ⫹ 7H 2 , ∆H°298 ⫽ 497.7 kJ/mol (15)
At lower temperatures, the production of methane may also occur:
n⫺1 3n ⫹ 1 n⫺1
C n H 2n⫹2 ⫹ H2O ⫽ CH 4 ⫹ CO (16)
2 4 4
The reactions are essentially at equilibrium during industrial processing, and
the water–gas shift reaction is important to the final product spectra:
CO ⫹ H 2 O ⫽ CO 2 ⫹ H 2 (4)
The equilibrium constant of this reaction, written as (70)
B
ln K p ⫽ A ⫹
T
where A ⫽ ⫺3.79762 and B ⫽ 4159.54, favors the production of carbon dioxide
at low temperatures.
Although nickel-based catalysts are not the most efficient promoters (70,73),
they are most cost-effective and are used industrially. Rhodium-based systems
are more efficient, but more costly.
The high temperatures required for steam reforming promote carbon forma-
tion (75–77), and steam: carbon ratios have to be kept higher than stoichiometric
in order to gasify coke. Rostrup-Nielsen (15) has presented an excellent summary
of the tendency to form coke during methane steam reforming and of the thermo-
ORDER REPRINTS

44 TRIMM AND ÖNSAN

dynamic carbon production limits in the presence of steam and carbon dioxide.
The possibility of carbon dioxide reforming
CO 2 ⫹ CH 4 ⫽ 2CO ⫹ 2H 2 (7)
is considered, although kinetic measurements (15) show the reaction to be slower
than steam reforming. As an approximation, a steam :carbon ratio of ⬃2.5 is suf-
ficient to avoid coking. Exact values for individual hydrocarbons may vary.
The tendency to form coke limits steam reforming to light hydrocarbons for
the industrial production of hydrogen, although the reforming of some feedstocks
Downloaded by [McGill University Library] at 08:56 20 November 2012

containing significant amounts of aromatics has been reported (78).


The specific activities of metals supported on alumina or magnesia have been
found to be (79)
Rh, Ru ⬎ Ni, Pd, Pt ⬎ Re ⬎ Co
Nickel is the most cost-effective catalyst (70,73) and is used universally in indus-
trial catalysts. Coke formation is much less over Rh and Ru (74).
The composition of the support used for nickel catalysts is adjusted to mini-
mize coke formation (70,73). Alkaline components such as magnesia or potassia
favor the gasification of coke (80). Recent studies have shown that ceria supports
may also improve activity and reduce coke formation (81).
The kinetics of the steam reforming of methane and the related water–gas
shift reaction have been the subject of much study. Considerable discrepancies
have been observed, with the reaction showing positive, zero, or negative orders
(82) depending on the range of steam partial pressures used. Elnashaie et al. (82)
recognized the kinetic equations generated by Xu and Froment (83) to be most
reliable, but they showed that many of the discrepancies in the reported values
could be explained in terms of nonmonotonic behavior of the system. The Lang-
muir–Hinshelwood equations reported by Xu and Froment were for three reac-
tions:
CH 4 ⫹ H 2 O ⫽ CO ⫹ 3H 2 (2)

r1 ⫽ k1
冢 PCH4 PH 2O
P 2.5
H2

P 0.5
H 2 PCO

K1 冣 DEN⫺2

CO ⫹ H 2 O ⫽ CO 2 ⫹ H 2 (4)

冢 冣
PCO PH 2O PCO2
r2 ⫽ k2 ⫺ DEN⫺2
PH 2 K2

CH 4 ⫹ 2H 2 O ⫽ CO 2 ⫹ 4H 2 (17)

r3 ⫽ k3
冢 PCH4 P 2H 2O
P 3.5
H2

P 0.5
H 2 PCO 2

K1 K2 冣 DEN⫺2
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 45

where
PH 2 O K H 2 O
DEN ⫽ 1 ⫹ K CO PCO ⫹ K H2 PH2 ⫹ K CH4 PCH4 ⫹
PH 2
Some idea of the variation that can arise can be seen from a comparison
with the rate of steam reforming of methane reported over Ni/MgO (74):
⫺n


r ⫽ kPCH4 1 ⫹ K H 2 P 1/2
H 2 ⫹ K H 2O
PH 2 O
PH 2 冣
Downloaded by [McGill University Library] at 08:56 20 November 2012

Power rate law kinetics are often used to describe steam reforming (86). A
selection of values reported is summarized in Table 3.
The composition of the products of fuel conversion depends both on the
kinetics of the reactions and on thermodynamics (70,73). It is well known that
steam reforming of hydrocarbons at relatively low temperatures produces methane
and, even at ⬃900°C, ⬃4% methane and 20% carbon monoxide emerge from
methane steam reforming (89). The latter may be reduced by applications of the
water–gas shift reaction at lower temperatures or by selective oxidation (see later).
Methane is harder to remove and is an unwanted emission gas. Even more methane
may be produced during the steam reforming of other hydrocarbons if catalytic
methanation can occur (90).
Many of these expected trends were confirmed by Ma (18). The conversion
of propane was effectively complete at 650°C over a nickel-based catalyst, but
methane conversion was still incomplete at 750°C. Little methane was produced
over a Pt-based reforming catalyst, but nickel-catalyzed methanation and consider-
able amounts were produced at about 450°C.
This creates an interesting activity/selectivity problem. Nickel-based cata-
lysts are active for steam reforming at about 100°C⫹ degrees lower than Pt-based
systems, but the selectivity to hydrogen is higher over Pt, particularly at the higher

Table 3. Power Rate Law Kinetics for Hydrocarbon Steam Reforming


Temp.
Range Pressure Orders EA
Hydrocarbon Catalyst (K) (MPa) RH H2O H2 (kJ/mol) Ref.

CH 4 Ni/MgO 723–823 0.1 1 — — 110 74


CH 4 Ni/MgO 623–673 0.1 0.96 ⫺0.17 0.25 60 18
C2H6 Ni/MgO 723–823 0.1 0.6 ⫺0.4 0.2 76 84
C2H6 Ni/MgO 583–623 0.1 0.95 ⫺0.46 0.38 80.6 18
C3H6 Ni/αAl 2 O 3 773–913 0.1 0.75 0.6 — 67 85
C3H8 Ni/MgO 583–623 0.1 0.93 ⫺0.53 0.86 45 18
C 4 H 10 Ni/γAl 2 O 3 698–748 3 0 1 — 54 75
C 4 H 10 Ni/SiO 2 673–723 0.1 0 1 — 87
C 7 H 16 Ni/MgO 723–773 0.5–3 0.1–0.3 ⫺0.2 0.8 67 88
ORDER REPRINTS

46 TRIMM AND ÖNSAN

bed temperatures. Given that the light-off temperature is achieved by the total
combustion of fuel, higher bed temperatures may mean less fuel availability for
steam reforming, but greater selectivity to hydrogen, depending on whether the
catalyst favors methanation.
Precious metals are more active, in general, for steam reforming than nickel.
Turnover numbers have been reported, relative to Ni, for methane and ethane
steam reforming. For silica-supported catalysts used for methane steam reforming,
the comparisons are (91)
Rh (1.6) ⬎ Ru (1.4) ⬎ Ni (1) ⬎ Pd (0.6) ⬎ Pt (0.5)
Downloaded by [McGill University Library] at 08:56 20 November 2012

whereas for alumina-supported systems with ethane (74,78)


Rh (13) ⬎ Ru (9.5) ⬎ Pd (1.0) ⬃ Ni (1.0) ⬎ Pt (0.9)
The economics of the choice are thus the basis the industrial use of Ni-based
catalysts.

3. Summary

It is clear that methanol and lower-molecular-weight hydrocarbons can be


steam reformed to produce hydrogen. Copper–zinc oxide or palladium–zinc oxide
are the preferred catalysts for methanol, but careful control of temperature is nec-
essary with copper–based catalysts.
Nickel-based systems are more cost-effective for light hydrocarbons, al-
though rhodium is an excellent catalyst. Coking is a potential problem which
increases as the molecular weight and aromaticity/olefinicity of the fuel in-
creases. Suitable steam : carbon ratios must be maintained in order to minimize
coking.

C. Combined Reactions: Indirect Partial Oxidation

Steam reforming reactions are endothermic, and heat must be supplied either
externally or by combustion of reactants and products. The latter concept is well
established with the temperature of a secondary reformer in the ammonia synthesis
train being raised by combustion with air of the exit gases from the primary re-
former (74). However, it was Jenkins (92,93) who first thought through and devel-
oped a system that could operate independently to convert fuels to hydrogen. It
is regrettable that this seminal work in the area has not been recognized more
widely, although Johnson Matthey Hot Spottm reactors, developed from his work,
are a continuing commercial success (12).
Autothermal operations are stand-alone systems in which fuel is partly oxi-
dized and partly steam reformed to produce hydrogen. The system must start to
operate at low temperatures, with the balance between oxidation and steam re-
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 47

forming depending on the heat demands of the particular fuel. Reactors based on
methanol are widely available, but other fuels may be used if reactor operation
is adjusted.

1. Methanol

The production of hydrogen by the autothermal reforming of methanol in-


volves the injection of methanol, water, and air into a catalytic reactor at ambient
temperature (94). Part of the methanol is oxidized to raise the temperature to the
Downloaded by [McGill University Library] at 08:56 20 November 2012

point where steam reforming can occur.


Although the Hot Spot reactor is probably the most convenient system, it
is useful first to consider a two-bed system, as this illustrates many of the problems
that are overcome in more advanced reactors.
The two-bed system designed and tested by Jiang et al. (24,95) consisted of
a bed of Pt/Al 2 O 3 catalyst followed by a bed of Cu/ZnO on alumina. Because
the injection of methanol and air was known to initiate total oxidation but to
produce temperatures in excess of those needed to sinter copper, water was added
to the feed. This gave two beneficial effects: the heat of vaporization of water
limited the temperature rise and the heat capacity of the steam allowed more effec-
tive heat to be carried through the reactor.
It is clear that the temperature of both the oxidation bed and the steam re-
forming bed is very dependent on the air-to-methanol ratio and the water-to-meth-
anol value. The presence of water allowed the control of temperature at values
less than those relevant to sintering (Table 4).

Table 4. Hydrogen Production from Methanol in a Two-Bed System


Tfinal Hydrogen
Liquid Tmax Steam Capacity
Feed Oxidation Reformer Methanol SCM
Rate Water/ Air/ Bed Exit Bed Exit Conversion Hydrogen/ Yield
(mL/h) Methanol Methanol (°C) (°C) (%) (lcat/h) (%)

10 1.38 2.16 320 300 87 4.69


10 1.38 1.64 280 200 81.5 5.72
10 1.38 1.20 230 180 44 1.92
52 3.5 3.3 335 235 99.6 15.3 80
32 3.5 3.5 305 235 99.8 7.1 60
45 4.0 3.3 310 240 96.2 8.3 29
10 3.12 2.56 310 235 96.1 1.4 47.5
45 4.01 3.44 325 250 100 8.2 55
32 3.5 3.04 357 245 91 7.1 60
27 2.98 3.03 340 240 92.4 7.0 62
23 2.53 2.60 340 245 91.1 6.0 55

Note: Bed 1 ⫽ 0.5 g Pt/Al 2 O 3 ; bed 2 ⫽ 2 g Cu/ZnO/Al 2 O 3 ; initial T ⫽ 25°C.


Source: Ref. 18.
ORDER REPRINTS

48 TRIMM AND ÖNSAN

Jiang et al. (51) reported that only small amounts of carbon monoxide
(⬃0.5%) were produced and that some exothermicity could be observed in the
second bed when copper oxide was being reduced at high conversions.
Although the two-bed system allowed the understanding of factors important
to the process, such a reactor is not the most efficient system. As discussed later,
Ma (18) has shown that two catalytic functions on the same support offers best
performance.
A multicomponent noble metal–base metal catalyst forms the basis of the
Hot Spot reactor (12). A feed of methanol/air/water is fed to the system and
Downloaded by [McGill University Library] at 08:56 20 November 2012

methanol is said to be converted by a mixture of partial oxidation and steam re-


forming to hydrogen. One suspects that total oxidation may be more probable,
but either reaction produces the heat needed for steam reforming from an initial
temperature of 0°C or less. The maximum temperature inside the reactor is re-
ported to be 400°C and as much as 2.4 mol of hydrogen are produced for every
mole of methanol consumed.
One very real advantage of the Hot Spot system is the modularity of the
reactors. The present reactor design is reported to produce 750 L/h of hydrogen
from a 245-cm 3 reactor (12). This is equivalent to a power density of 3 kW/L
for each reactor. A combination of several such reactors can provide higher kilo-
watt power outputs, and suitable geometries are described by Golunski (12). The
start-up of each individual reactor is facilitated by the addition of extra air, with
steady-state oxidation/steam reforming of an eight-unit reactor being achieved
within 100 s.
Ma et al. (95) have considered the optimization of the process for use at
atmospheric and higher pressures. Somewhat surprisingly, pressure was predicted
not to have too large an effect on the reaction, with temperature being more criti-
cal. Conversion and selectivity were higher at higher temperatures, but copper
sintering again limited the allowable range.
Velu et al. (96) have considered the use of CuZnAl ternary oxides for the
production of hydrogen from methanol. Extra thermal stability might be expected,
but the catalysts were tested only up to 250°C. Hydrogen selectivity approached
100% at 200°C, but carbon monoxide production increased at higher temperatures.
They recommended operation at 200°C to give 40–60% conversion of methanol
and ⬎95% carbon dioxide. They did not investigate the initiation of the reaction
from ambient conditions.
It is necessary to address the terminology used to describe these studies.
Velu et al. (96) described their results in terms of partial oxidation:
CH 3 OH ⫹ 1/2 O 2 ⫽ CO 2 ⫹ 2H 2 (9)
but it seems more probable that the reaction involves a combination of total oxida-
tion and steam reforming:
CH 3 OH ⫹ 3/2 O 2 ⫽ CO 2 ⫹ 2H 2 O (8)
CH 3 OH ⫹ H 2 O ⫽ CO 2 ⫹ 3H 2 (5)
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 49

Thus, it is possible to describe the reaction as direct partial oxidation [reac-


tion (4)] or indirect partial oxidation [reactions (1) and (2)], a difference which
will become more important when hydrocarbons are considered.
Cubeiro and Fierro (72) have studied the production of hydrogen from meth-
anol over Pd/ZnO and suggest that indirect partial oxidation involving methyl
formate is important. Selectivities of ⬃80% hydrogen from 80% conversion of
methanol in the absence of added water were reported, and higher values could
be expected if water had been introduced to the reactor.
However, the main interest in this article is the use of Pd/ZnO, a catalyst
Downloaded by [McGill University Library] at 08:56 20 November 2012

which could be expected to be stable at temperatures in excess of those at which


Cu-based catalyst sinter (⬃330°C). Pd/ZnO may not be the best catalyst [some
alloy formation was observed at higher temperatures (72)], but Pd catalysts would
seem to offer a real alternative. Regrettably, Cubeiro and Fierro did not study
ambient light-off possibilities, but Pd-based catalysts should be active at low tem-
peratures and stable to higher temperatures than Cu-based systems.

2. Hydrocarbons

The production of hydrogen from hydrocarbons is a well-established process,


with a combination of oxidation and steam reforming of methane forming the
basis of an important industrial process (14,70,73). However, the constraints on
the overall process in the context of supplying hydrogen for fuel cells introduce
the possibility of new reactions and different catalysts.
The industrial process involves mainly the steam reforming of methane in
two reforming reactors, with air being added before the secondary reformer to
increase the temperature. Equilibrium is finally reached via the water–gas shift
reaction:

CH 4 ⫹ H 2 O ⫽ CO ⫹ 3H 2 (2)

CH 4 ⫹ 2O 2 ⫽ CO 2 ⫹ 2H 2 O (6)

CO ⫹ 1/2 O 2 ⫽ CO 2 (18)

2H 2 ⫹ O 2 ⫽ 2H 2 O (19)

CO ⫹ H 2 O ⫽ CO 2 ⫹ H 2 (4)

The high temperatures necessary to produce hydrogen (70,73) favor the formation
of coke, and catalyst composition is adjusted to promote gasification (75,76).
Steam reforming [reaction (2)] produces a H 2 :CO ratio of 3 :1, which is
unsuitable for many industrial applications such as methanol synthesis. As a result,
so-called dry reforming has been studied (14,99):

CO 2 ⫹ CH 4 ⫽ 2CO ⫹ 2H 2 (7)
ORDER REPRINTS

50 TRIMM AND ÖNSAN

Table 5. Initiation of Hydrocarbon Oxidation

Fuel Needed to Achieve


Light-off Light Off
Temp. Methanol Hydrogen
Hydrocarbon (°C) (g/kg cal) (L/kg cal)
Methane 415 63 178
Ethane 265 29.7 98
Propane 195 19 76
Downloaded by [McGill University Library] at 08:56 20 November 2012

Source: Ref. 18.

This reaction has the disadvantages that the tendency to form coke is rela-
tively high and that the reaction is slower than steam reforming (70,73). As a
result, steam reforming will predominate if water is present or is produced.
One other problem is that the conversion of hydrocarbons cannot be initiated
at room temperature (Section III). However, it is possible to combine two systems.
Thus, for example, Ma (18) has measured the light-off temperatures for higher
hydrocarbons and used the catalytic oxidation of methanol or hydrogen (which
can be initiated at room temperatures) to reach these values (Table 5). Once the
hydrocarbons starts to oxidize, no further supply of the initiating fuel is required.

Hydrogen from Methane

The conversion of methane to hydrogen is a well-established reaction. Steam


and methane, at a ratio of ⬃2.5 or greater, is passed over Ni-based catalysts at
high temperature to produce hydrogen. Details of the system have been considered
earlier.
The catalytic partial oxidation of methane to synthesis gas has recently re-
ceived much attention. Verykios et al. (100) report that this is due not only to the
formation of more favorable H 2 : CO ratios but also to the fact that the system
requires no external heating, there is no release of undesirable emissions apart
from carbon dioxide, and fact that the reaction is highly selective.
On nickel-based catalysts, the reaction appears to proceed via total oxidation
coupled to steam reforming (100), except at very low oxygen partial pressures.
The deposition of carbon was investigated by Claridge et al. (101), who reported
whisker carbon and encapsulant carbon (75). Van Looij and Geus (102) report
that the oxidation of methane occurred on nickel oxide, whereas reforming took
place further down the bed on metallic nickel sites. Goula et al. (103) noted that
the degree of reducibility of nickel was, in part, controlled by the nature of the
support. Lu et al. (104) report that the use of a calcium aluminate spinel support
gave a conversion of 83% with a selectivity of ⬎90% and carbon formation of
⬍1 wt% for operation at 600°C. Tsipouriari et al. (100) used Ni/La 2 O 3 to produce
selectivities and conversions close to thermodynamic predictions with minimal
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 51

carbon. The lanthana was suggested to distribute nickel and to facilitate gasifica-
tion.
The conversion of methane to hydrogen using Pt to promote oxidation and
Ni to promote steam reforming has been studied in depth (18,51). The use of two
beds of catalyst, of mixed oxidation and steam reforming catalyst, and of a Ni
catalyst on which Pt was deposited have been considered. The result reflected
more the effect of mass and heat transfer in the system, with a bifunctional catalyst
giving optimal performance (Fig. 1). Methane conversions of up to 92% were
obtained with hydrogen product efficiencies of ⬃80% at an operating temperature
Downloaded by [McGill University Library] at 08:56 20 November 2012

of 630°C. The carbon:oxygen ratio at the inlet was 1.55 and the steam-to-carbon
ratio was 2.34.
The process studied was almost certainly total oxidation plus steam re-
forming, but the picture changes significantly when considering precious metals.
Rostrup-Nielsen and Bak Hansen (99) had found that Rh and Ru were more active
than Ni for reforming methane and produced less carbon. Although availability
and price favor Ni-based catalysts, direct partial oxidation and carbon dioxide
reforming could be favored over precious metals because less steam would be
required.
Nakamura et al. (105) were one of the first to investigate partial oxidation
over Rh/SiO 2 and found that the reaction involved total oxidation, followed by
reforming with both carbon dioxide and steam. Nakagawa et al. (106) extended
these studies to find that Ir/TiO 2 and Rh/SiO 2 involved total oxidation plus re-
forming, but that different reactions were favored for Rh/TiO 2 and Rh/Al 2 O 3 .
They suggested that decomposition of methane to carbon and hydrogen occurred

Figure 1. Comparison of relative hydrogen production efficiencies of various bed configurations.


ORDER REPRINTS

52 TRIMM AND ÖNSAN

over these catalysts, the deposited carbon (or CH x species) being subsequently
oxidized by carbon dioxide or residual oxygen.
Bitter et al. (107) suggested a somewhat similar mechanism to explain car-
bon dioxide reforming over Pt- and Rh-based catalysts. They noted that Mark and
Maier (108) had found that methane decomposed on Rh to give carbon, which
was subsequently oxidized by carbon dioxide. For Rh on zirconia or alumina,
Bitter preferred coadsorption of methane and carbon dioxide to produce CH x and
adsorbed O, which reacted together to form CO and hydrogen. The support was
considered to have little effect with Rh catalysts but to be more important
Downloaded by [McGill University Library] at 08:56 20 November 2012

with Pt.
The effect of support on the reactions is a matter of some uncertainty, as
different patterns have been reported (Table 6). Part of the explanation of these
differences comes from the work of Ruckenstein and Wang (109), who have found
that reducible oxides could migrate onto the surface of Rh to block sites.
It also seems possible that all authors have not recognized the relative impor-
tance of individual reactions in the overall partial oxidation process. Thus, for
example, direct partial oxidation has been suggested—at least in the case of Pt
on ceria—to involve oxygen from the lattice (110). Subsequent reoxidation may
or may not involve the precious metal.
Carbon dioxide reforming, on the other hand, has been suggested to proceed
via carbonaceous intermediates (107), which do not necessarily demand lattice
oxygen. Thus, the relative importance of the two reactions may influence the effect
of the support.
What is clear is that partial oxidation—either direct or indirect—is favored
by precious metals. Carbon formation is less than on Ni and the overall activity
may be higher (99,103). Thus, for example, Mark and Maier (108) reported ⬃70%
conversion of methane at 750°C over 1% Rh/ZrO 2 for carbon dioxide reforming
with little coke formation.

Hydrogen from Higher Hydrocarbons

The wide availability of LPG, gasoline, and diesel would make them ideal
as fuels for hydrogen production. Regrettably, not enough attention has been fo-
cused on the catalytic conversion, particularly of LPG.

Table 6. Activity Patterns of Supported Precious Metals Used to Process Methane


Reaction Catalyst Support Activity Ref.

CO 2 /CH 4 Rh Al 2 O 3 ⬎ TiO 2 ⬎ SiO 2 112


CO 2 /CH 4 Ir TiO 2 ⱖ ZrO 2 ⱖ Y2 O 3 ⬎ La 2 O 3 ⬎ MgO ⱖ Al 2 O 3 ⱖ SiO 2 113
Partial oxidation Ir As above 114
CO 2 /CH 4 Ru TiO 2 ⬎ Al 2 O 3 ⬎ C
Partial oxidation Rh MgO ⱖ γ-Al 2 O 3 ⬎ La 2 O 3 110
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 53

The essential reason would seem to be coke formation. Uncatalyzed partial


oxidation reactions are used to produce hydrogen, but, even in the gas phase, coke
formation is a problem. Such processes have been well reviewed by Jamal and
Wyszynski (11) and a gasoline-reforming fuel cell developed by Arthur D. Little
and Chrysler has been described (97). The fuel conversion is based on partial
oxidation initiated by spark ignition, and initial temperatures are raised by electric
heating. Any sulfur in the fuel is removed as hydrogen sulfide, and steam is added
before passing the exit gases through a low-temperature-shift catalyst. The final
removal of CO (to less than 10 ppm) involves selective oxidation over a Pt-based
Downloaded by [McGill University Library] at 08:56 20 November 2012

catalyst. The device is reported to operate satisfactorily with a range of fuels (97).
The operation of such devices in the context of gasoline–hydrogen-fueled
spark ignition engines has been well reviewed (11). Houseman and Cerini (16,98)
reported that there is always a tendency to form soot from gasoline unless a cata-
lyst was present. An air: gasoline ratio of 5.2 or greater was recommended, but
this would be expected to increase with heavier fuels. The presence of a catalyst
was reported to reduce the ratio to 3.5–4.0 for gasoline (16,98).
Other potential problems with the system could include unwanted emissions
or too large quantities of S or N in the fuel (32), which will poison metallic cata-
lysts.
There is reported to be considerable interest in such partial oxidation sys-
tems, but the use of a catalyst would seem to offer advantages, particularly with
respect to soot formation. With light hydrocarbons, the possibility of indirect par-
tial oxidation would appear to be feasible. The light-off temperatures are low
(⬃200°C) and not much initial heating is required. Conversion of ethane or pro-
pane is complete at 500°C (18,73,111) with increased amounts of hydrogen being
produced as the steam : carbon ratio increased. Some methane was produced over
Ni-based catalysts (18). Preliminary results showed that 35% conversion of ethane
to hydrogen could be obtained at a steam :carbon ratio of 2.5—the approximate
level at which coking does not occur. A yield of 50% was observed at a steam :
carbon ratio of 4 over a Ni/MgO catalyst held at 400°C.
It is regrettable that further investigation of the indirect and direct partial
oxidation of light hydrocarbons has not been conducted. Coking would be ex-
pected to be controllable or, possibly, avoidable over precious metal catalysts
(99,113), and hydrogen yields should be high. Some studies have been reported
for propane (111).
The most extensive study of hydrocarbons to hydrogen was made by Jenkins
and Shutt (93) using the Hot Spot reactor. Autoignition was not possible, but hydro-
gen generation from several fuels was reported using a 1% Pt:3% chromia:silica
catalyst. Coke formation would be expected to be less over precious metal catalysts
(99). Yields and selectivities were much more dependent on carbon:oxygen ratios
as compared to methanol, but excellent selectivities were reported (Table 7).
The gasoline used was lead-free and contained 45% aromatics, up to 20%
olefins, and balance saturates. Runs of up to 6 h were reported with little deactiva-
tion, even at an oxygen :carbon ratio of 0.27.
ORDER REPRINTS

54 TRIMM AND ÖNSAN

Table 7. Conversion of Fuels to


Hydrogen over Pt/Cr 2 O 3 /SiO 2

H 2 Selectivity
Fuel O 2 :C (%)

Gasoline 1 25
1.5 50
2.0 80
2.5 50
3 8
Downloaded by [McGill University Library] at 08:56 20 November 2012

Hexane 1 20
1.5 60
2 50
2.5 15
Diesel 1.5 15
2 45
2.5 30
Source: Ref. 93.

The authors report that results with diesel were erratic. These results are
extremely interesting, but the present authors have not been able to reproduce
them, possibly as a result of significant expertise in catalyst preparation by Jenkins
and Shutt (93).
More detailed studies are in progress for hydrogen production from iso-
octane (42). The hydrocarbon lights off at ⬃200°C over Pt/chromia/silica with
indirect partial oxidation was apparently favored at less than 600°C and carbon
formation/carbon dioxide reforming at higher temperatures.

3. Summary

It is clear that the indirect oxidation of methanol and of light hydrocarbons


can be used to produce hydrogen. It may also be possible to convert higher hydro-
carbons (93) provided coking is minimized.
The indirect oxidation of hydrogen and methanol can be initiated at room
temperature. Initial heating must be used to initiate reactions involving hydrocar-
bons.
The chemical reactions show that more hydrogen is produced from hydrocar-
bons than from methanol, but the overall analysis is more complex. The choice
of best fuel and best catalyst is discussed in Section V.

IV. DIRECT PARTIAL OXIDATION

The direct partial oxidation reaction


CH 4 ⫹ 1/2 O 2 ⫽ CO ⫹ 2H 2 (3)
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 55

is known to produce advantageous CO :H 2 ratios for further processing of synthe-


sis gas. However, in the context of fuel cells, maximal hydrogen is desired, and
the CO: H 2 ratio of 2 is less than optimal. Nonetheless, the reactions offer a possi-
ble route to hydrogen.
The direct partial oxidation reaction has been recognized for some time
(115), but it was the work of Ashcroft and his group (116,117) that has led
to recent interest in the reaction. Working with reactors at high temperatures
(⬃ 750°C) and short residence times (0.1 s), they reported methane conversions
above 90% and hydrogen selectivities of 95–99% over precious metal catalysts.
Downloaded by [McGill University Library] at 08:56 20 November 2012

Direct partial oxidation of various hydrocarbons has been extensively ex-


plored by Schmidt et al. (118–122). Limited work has also been carried out with
the direct partial oxidation of methanol (123).
Working at very low contact times (⬃5 ms), the direct partial oxidation of
methane was found to proceed to high conversions and high (⬃90%) selectivity
to hydrogen over Rh washcoated monoliths maintained at ⬃1000°C (118–120).
Coke formation was avoided under these conditions, but the role of the catalyst
and the importance of pore structure in the washcoat was established (118). Al-
though homogeneous reactions may be important, heterogeneous catalysis was
established beyond doubt.
Choudhary et al. (124–127) have adopted a similar approach using a nickel-
based catalyst. Alkaline and rare earth oxide supports were used in pellet form,
and a low contact time (4.8 ms) was maintained. Using a pure methane :pure
oxygen ratio of 1.8, 91% methane conversion was obtained at 800°C with 95%
selectivity to hydrogen. Direct partial oxidation was confirmed with a H 2 :CO ratio
of 2.0 in the products.
An economical method of maximizing H 2 production by combined partial
oxidation of CH 4 and water–gas shift reaction was described by Maiya et al. (128).
The partial oxidation of CH 4 at ⬃900°C to produce syngas (containing ⬃2 mol
of H 2 per mole of CH 4 ) was achieved over an Rh-based catalyst by oxygen trans-
ported through a ceramic membrane tube. The syngas produced was subjected to
the WGS reaction in a second reactor containing a Cu/Zn catalyst. Experiments
and thermodynamic calculations showed a maximization of H 2 production by
WGS at ⬃400°C and a H 2 O/CO of ⬃2, which resulted in the production of 2.9
mol of H 2 per mole of CH 4 reacted.
Schmidt and co-workers have extended the studies to methanol (123) and
higher hydrocarbons (121–123). An interesting balance was found between the
production of olefins (e.g., ethylene from ethane) and hydrogen (e.g., butane oxi-
dation). For butane, the hydrogen selectivity was over 90%; for pentane, the hy-
drogen selectivity was ⬃0.3 and the olefin selectivity was ⬃0.45. Each fuel was
considered separately and the reasons for the change in selectivity between hydro-
carbons were not extensively discussed.
The decomposition of methanol and partial oxidation was also explored
(123). Over Rh gauze, the reactions were significant at about 330°C and extensive
at ⬃730°C: Pt gauze reactions occurred at 370°C (significant) and 950°C (exten-
sive).
ORDER REPRINTS

56 TRIMM AND ÖNSAN

The advantages of the direct partial oxidation rest on the small size of the
reactor, the rapid response to changes, and in the possible absence of coking prob-
lems. The disadvantages lie with the CO : H 2 ratio of 0.5 and the fact that fuel
and air must be premixed. The proportions are such that such a mixture may be
flammable or even explosive, particularly if small variations (e.g., as a result of
pumping and vaporizing liquid hydrocarbons) are possible.
However, the advantages of the small reactor and of rapid response make
direct partial oxidation a process worth consideration.
Downloaded by [McGill University Library] at 08:56 20 November 2012

V. LOW-TEMPERATURE REMOVAL OF CARBON MONOXIDE

The most promising fuel-cell technology for transport applications has been
shown to be the polymer electrolyte membrane fuel cell operating with hydrogen.
The cell operates at relatively low temperatures (⬃80–120°C) and requires an
onboard unit to convert liquid or gaseous fuels to hydrogen.
The most efficient fuel-cell anodes are based on precious metals (4–6,129),
the amount required having been reduced to the point where the anode is no longer
a cost determinant. However, precious metals are sensitive to deactivation by car-
bon monoxide adsorption at low temperatures (130,131) and it is essential to re-
duce the amount of CO entering the cell. Tungsten carbide has been found to be
an alternative material that is not sensitive to CO adsorption (132), but the activity
is very much lower than that of precious metal catalysts.
The reformate coming from the fuel processing section contains ⬃75 vol%
H 2 , ⬃24 vol% CO 2 , and ⬃1–3 vol% CO. At the low temperatures concerned,
the Pt or Pt–Ru anodes are poisoned by CO, which strongly chemisorbs on the
active sites, thus blocking the sites where the dissociation/oxidation of H 2 can take
place and deteriorating cell performance. Latest technology anodes are typically
tolerant to 20–100 ppm CO in the feed gas (133,134). It is therefore necessary
to minimize the amounts of CO in the reformate stream.
Reduction to ppm levels can be achieved in several ways. One obvious route
is to diffuse hydrogen through a Pd/Ag membrane (23), but this requires a high-
pressure differential and fairly high temperatures. It is also expensive. By analogy
with the ammonia synthesis train, carbon monoxide could be methanated, but this
consumes 3 molecules of hydrogen per molecule of CO and results in the emis-
sions of methane—an unwanted greenhouse gas. Methanation of carbon dioxide
could also occur. As a result, attention has been focused on selective oxidation
of carbon monoxide, which has to be achieved in the presence of CO 2 and H 2 O,
with minimum H 2 oxidation.
The fuel conversion system involves a partial oxidation process coupled to
or followed by a water–gas shift system. Exit gas temperatures required to achieve
reasonably low concentrations of CO from the water–gas shift are of the order
of 150–250°C. The fuel cell operates at ⬃80–100°C. Thus, the preferred tempera-
tures for removal of CO are of the order of 80–100°C.
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 57

Selective oxidation is the preferred process. Golunski (12) described the oxi-
dation of CO in several stages—presumably to reduce the possibility of overheat-
ing. CO levels were reduced from ⬎2% to 10 ppm at the expense of 6% of the
hydrogen output.
No details of the catalyst were given, but recent work has thrown light on
possible systems. It is useful to discuss these first in terms of the low-temperature
oxidation of carbon monoxide (a problem in itself) and then in terms of the selec-
tive oxidation.
Downloaded by [McGill University Library] at 08:56 20 November 2012

A. Low-Temperature CO Oxidation

The oxidation of CO is a most important reaction in emission control, partic-


ularly in the cleanup of vehicle emissions (136). Supported noble metal catalysts
are typically used to promote the reaction. The precious metals used are efficient,
but only start to act at about 170°C (137,138). This is true for most conventional
catalysts, which are not active at low temperatures and low O 2 /CO ratios because
O 2 and CO compete for the same sites. O 2 adsorption has more severe site require-
ments in that it must dissociate. Under these conditions, CO adsorption dominates
the metal surface and prevents O 2 adsorption and surface reaction. Over base–
metal oxide catalysts, on the other hand, O 2 is held too strongly to be displaced by
CO at low temperatures. Thus, an efficient low-temperature CO oxidation catalyst
should accommodate both CO chemisorption and the simultaneous dissociative
adsorption of O 2 . This has suggested the use of composite materials with different
components, each having activity for one or other of these functions (137). Noble
metal reducible oxide combinations, base metal oxides (139,140), and perovskite-
type catalysts (141,142) have been tested for their low-temperature CO oxidation
performance.
One common feature can be expected and is found. If the low-temperature
oxidation involves reaction between CO adsorbed on one component and oxygen
species adsorbed on or originating from a second component, then the interface
between the components will be critical. As a result, preparation techniques are
expected to be—and are—vital to the preparation of active catalysts.

1. Noble Metal Reducible Oxide Catalysts

It was demonstrated by Bond et al. (143) in the mid-1970s that Pd supported


on tin(IV) oxide gives substantially higher catalytic activity for CO oxidation by
providing suitable sites for both CO and O 2 chemisorption. Significant reaction
rates were reported over Pt/SnO 2 and Pd/SnO 2 at temperatures below 0°C (144–
146). The phrase ‘‘noble metal reducible oxide’’ (NMRO) was used (147) for this
class of catalysts, which include Pt/SnO 2 , Pt/MnO x , Pt/CoO x , Pt/CeO 2 , Pd/SnO 2 ,
Pd/MnO x , Pd/CeO 2 , Pd/ZrO 2 , Au/MnO x , Au/CeO 2 , Au/Fe 2 O 3 , Au/Co 3 O 4 ,
Au/TiO 2 , Ag/MnO x , Rh/TiO 2 , and the like.
ORDER REPRINTS

58 TRIMM AND ÖNSAN

The catalysts consist of a zero-valent noble metal such as Pt, Pd, Au, or Rh
dispersed over or intermixed with a metal oxide that is reduced under reaction
and/or pretreatment conditions. The noble metal and the oxide may further be
supported on SiO 2 or Al 2 O 3 . Factors that determine oxygen availability and the
gaseous impurities play a major role in catalyst design and performance.
Low-temperature noble metal reducible oxide catalysts must exhibit strong
metal-support interaction (SMSI), because neither the noble metal nor the reduc-
ible oxide alone can catalyze CO oxidation at temperatures ⬍100°C. One or more
of three types of synergetic interaction between the two catalyst components are
Downloaded by [McGill University Library] at 08:56 20 November 2012

said to be responsible for the high efficiency observed at low temperatures


(137,147):

1. The two components may each have independent functions in the cata-
lytic CO oxidation mechanism.
2. The properties of one component may be modified by the presence of
the other.
3. The two components may associate at the atomic level in such a way
as to form unique active sites.

The first type of interaction can be illustrated by CO oxidation on TiO 2-


supported Au, where CO has been suggested to adsorb on the metal and oxygen
is provided from the support (148). The second is illustrated by the observation
that the electronic structure of small metal particles (and hence their adsorption
properties) can be altered by electron transfer to or from the support. Thus, Ru/
MgO and Ru/TiO 2 leads to electron-rich Ru particles, whereas Ru/SiO 2 –Al 2 O 3
and Ru/TiO 2 –Al 2 O 3 leads to electron deficiency (149).
Association at an atomic level leads to possible dissolution or alloy forma-
tion, such as occurs on extensive reduction of Pt–Sn (150), Pd–Ag (23), and Pt–
Rh (151). The new material may be a better or worse catalyst.
The relative importance of these effects is best considered in the context of
catalysts known to promote low-temperature CO oxidation.

Gold Catalysts

The first such catalyst involves supported gold. Although long regarded as
a less active catalyst, studies conducted by Haruta and co-workers (152–154) and
by several other groups (148,155,156) have shown that low-temperature CO oxi-
dation is successfully catalyzed by gold nanoparticles highly dispersed on reduc-
ible oxides such as TiO 2 (157–160), Fe 2 O 3 (158,161,162), Co 3 O 4 (158,163), ZrO 2
(164), MnO x (155,156,165), and NiO (163) or on hydroxides of alkaline earth
metals such as Mg(OH) 2 and Be(OH) 2 (154,158). Research conducted by Haruta
and co-workers has recently been reviewed (152,166). A more comprehensive
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 59

review by Bond and Thompson (167) on catalysis by gold, covering all aspects
of chemisorption and oxidation of CO on supported gold catalysts, has also been
recently published.
Metal–support interactions have a significant effect on the performance of
supported gold catalysts. For example, although pure gold or pure titania do not
catalyze CO oxidation below 230°C, Au/TiO 2 catalysts are active at ambient tem-
peratures (159,168,169). TPD data obtained on Au/TiO 2 and TiO 2 indicate that
the O 2 and CO uptakes of TiO 2 alone are small, but appreciable increases are
observed in the desorption peaks over TiO 2 deposited with Au (158).
Downloaded by [McGill University Library] at 08:56 20 November 2012

In fact, the activity of the catalysts depends markedly on Au particle size.


The turnover frequencies (TOFs) of supported gold catalysts in CO oxidation in-
crease as the gold particle size decreases (166,167), with recent investigations
indicating an optimum gold particle size of 2–3 nm (170,171). Haruta et al. (158)
have demonstrated that although the CO oxidation activities of metal oxide sup-
ports may vary considerably (Co 3 O 4 ⬎ α-Fe 2 O 3 ⬎ TiO 2 ), the TOFs based on
surface Au atoms are almost independent of the kind of metal oxide support but
significantly dependent on Au particle diameter. Comparative studies of Au/TiO 2
and Au/ZrO 2 catalysts prepared by the same technique (164) show that the gold
particle size is similar on both catalysts (⬇2 nm), whereas the exposed gold sites
are different, depending on the oxide support and the pretreatment.
Because this is the case, it is not surprising that significant differences in
catalytic behavior result from different preparation techniques. Higher catalytic
activities are obtained if deposition–precipitation or coprecipitation is used instead
of impregnation. For example, rates of CO oxidation at 27°C over impregnated
1% Au/TiO 2 are found to be 10⫺4 times smaller than that on 1% Au/TiO 2 prepared
by deposition–precipitation and 10⫺3 times smaller than the rate on impregnated
1% Pt/TiO 2 (172). This may be due either to the metal particle size effect or to
the pretreatment used. Bollinger and Vannice (148) report that impregnated Au/
TiO 2 becomes active after a sequential pretreatment consisting of high-tempera-
ture reduction at 500°C, calcination at 400°C, followed by low-temperature reduc-
tion at 200°C. In fact, after such a pretreatment, an impregnated catalyst with an
average Au crystallite size of 25 nm had activity comparable to a coprecipitated
catalyst with an average crystallite size of 4.5 nm.
A conclusive quantitative analysis of the kinetics and mechanism of CO
oxidation over supported gold catalysts has yet to be presented, because of the
scarcity and irregularity of results. CO oxidation over supported gold catalysts
may be occurring (1) on the support between the oxygen species and the adsorbed
CO spilling over from gold particles (158), or (2) on the gold particles between
adsorbed CO and O coming from the support by reverse spillover (152,166), or
(3) mainly near the perimeter interface between the gold particles and the oxide
support (159), as also suggested by Haruta (166). The third mechanism seems to
be indicated by the strong dependence of the CO oxidation activity on the Au
cluster size with a maximum around 3.2 nm (173).
ORDER REPRINTS

60 TRIMM AND ÖNSAN

Active-site configurations suggested on the basis of the foregoing mecha-


nisms have been summarized by Dekkers et al. (157) in four groups: (1) both Au
and metal oxide participate in the catalytic process with the Au–metal oxide inter-
face as the reaction site, (2) active species are very small gold particles stabilized
by the metal oxide, (3) active species is ionic gold stabilized by the metal oxide,
or (4) active sites are specific ensembles of gold atoms.
Haruta et al. (158) proposed a mechanism in which O 2 adsorbs at the Au–
oxide interface, possibly as O 2⫺ , whereas the CO reversibly adsorbed on Au parti-
cles migrates toward the Au–oxide interface, where it reacts with adsorbed O 2⫺
Downloaded by [McGill University Library] at 08:56 20 November 2012

to form bidentate carbonate species. The carbonate intermediate was suggested


either to decompose or to react further with adsorbed CO to form gaseous CO 2 .
Fourier transform infrared (FTIR) measurements indicated the presence of carbon-
ate species. Kinetic studies showed an oxygen dependence of 0.2–0.3 for Au/
Co 3 O 4 and Au/TiO 2 and zero for Au/α-Fe 2 O 3 , as well as a zero-order dependence
on CO for all three catalysts. Activation energies for CO oxidation were 3.9 kcal/
mol (Au/Co 3 O 4 ), 8.2 kcal/mol (Au/TiO 2 ), and 8.4 kcal/mol (Au/α-Fe 2 O 3 ) in the
⫺10°C to 65°C temperature range.
Au/MnO 2 has been suggested to be a promising catalyst with high long-
term CO oxidation activity. Gardner et al. (155,165) have shown that coprecipi-
tated Au/MnO x catalysts are about five times more active than Pt/SnO x catalysts
prepared by deposition–precipitation. Activities of coprecipitated Au/MnO 2 cata-
lysts tested both in powder form and as monolith-coated samples decreased con-
siderably at O 2 ratios less than stoichiometric and increased remarkably in an O 2-
rich environment (155,174). The reaction was found to be first order in O 2 , with
no simple reaction order for CO. Carbon monoxide inhibited the reaction at stoi-
chiometric CO/O 2 ratios, but the apparent order in CO was zero in excess oxygen.
Although reaction data indicate that Au and MnO x interact synergetically, the Au–
MnO x interaction and the CO oxidation mechanism are not clearly understood.
This interaction appears to be a unique property associated with the support oxide
and the Au crystallite size.
There is also some question as to the level of CO reduction that can be
achieved over the catalysts. No direct reports have been found, but residual CO
levels have been found to be greater than over precious-metal-based catalysts.
Gold oxide catalysts tend to deactivate when CO 2 is also present in the reac-
tion mixture (155). Recent studies on catalyst durability indicate that activity can
be maintained over extended periods, particularly under strongly oxidizing condi-
tions (167).

Platinum Catalysts

Among the most promising catalysts investigated for the efficient oxidation
of CO below 100°C and at ambient temperatures is Pt supported on tin(IV) oxide,
with or without additional promoters (137,175). These catalysts require a reductive
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 61

pretreatment and traces of H 2 or H 2 O in the reaction mixture, in order to exhibit


their maximum activity. Long-term activity decay, due partly to CO 2 retention,
is observed at high CO 2 partial pressures.
Because Pt/SnO 2 has much higher catalytic activity for CO oxidation than
either Pt or SnO 2 alone (144,145), a synergetic interaction involving separate but
complementary sites on the Pt and SnO 2 phases or at the Pt–SnO 2 interface clearly
exists (176). The activity for CO oxidation on Pt/SnO 2 increases with Pt loading
up to 19.5 wt% Pt/SnO 2 or 14.5 wt% Pt/SnO 2 /SiO 2 , where maximum activity is
attained (155). The addition of a small amount of Pd enhances CO oxidation activ-
Downloaded by [McGill University Library] at 08:56 20 November 2012

ity (175), and it has been reported that the addition of an iron promoter improves
both the activity and the decay profiles (177).
The nature, temperature, and duration of pretreatment affect the subsequent
catalytic activity of Pt/SnO 2 (146,178). A relatively mild reductive pretreatment
using CO or H 2 at temperatures between 125°C and 225°C for about 1 h increased
catalytic activity relative to no pretreatment or pretreatment with air or inert gas.
Reduction at 125°C led to conversion of Pt oxides to Pt(OH) 2 , whereas reduction
at 225°C led to reduction to Pt metal.
Pretreatment of Pt/SnO 2 catalysts at elevated temperatures or for extended
periods was found to result in an initial fall in the CO 2 yield that persisted for
hours or days before the higher steady-state yield was reached (178). It was shown
that this dip in the CO 2 yield is caused by surface dehydration and can easily be
remedied by humidifying either the catalyst or the reaction mixture. The addition
of moisture increases the activity even for unpretreated Pt/SnO 2 . The rate en-
hancement in humidified reaction gases is found to be an advantage over the water-
sensitive Hopcalite catalyst.
A complete analysis of the low-temperature CO oxidation mechanism over
supported or unsupported Pt/SnO 2 catalysts has not thus far been presented. On the
basis of reaction rate measurements, Boulahouache et al. (179) have explained the
synergetic effect between Pt and SnO 2 in terms of the dissociative adsorption of
oxygen on reducible SnO 2 sites followed by oxygen reverse-spillover onto the Pt
metal and reaction with CO chemisorbed on the metal. The results of CO and O 2
titration experiments conducted by the same group on Pt/SnO 2 catalysts support
the hypothesis that the oxygen is moving from the oxide and show the influence
of the Pt dispersion on the CO oxidation rate. The presence of additional sites at the
phase boundary or of some remote control mechanism could not be distinguished.
Bond et al. (180) have proposed that CO oxidation on Pd/SnO x involves a
synergy between the metal and the support, with CO migrating from the metal
to the oxide where the oxidation takes place:
CO ⫹ O 2⫺ ⫺ Sn 4⫹ ⫺ O 2⫺ ⫽ CO 2 ⫹ ⫺Sn x⫹ ⫺ O 2⫺
The oxygen vacancies in the SnO x are then removed upon direct O2 adsorption.
The Pt/SnO x results obtained by Sermon et al. (181) are consistent with this pic-
ture. Hydrogen chemisorption on Pt/SnO x revealed the close interaction between
Pt and SnO x , because even a mild reduction caused loss in adsorption capacity
ORDER REPRINTS

62 TRIMM AND ÖNSAN

as a result of the decoration of the Pt surface by SnO x . The Pt–SnO x interaction


suppresses the H 2 adsorption capacity and increases the specific CO oxidation
activity after a low-temperature reductive pretreatment at 200°C.
Grass and Lintz (176) studied the kinetics of CO oxidation by oxygen in the
0–80°C temperature interval on Al 2 O 3-supported Pt/SnO 2 catalysts. The results of
rate measurements and titration experiments indicated that the chemisorption of
CO is restricted to Pt, whereas oxygen is adsorbed both on Pt and SnO 2 . At low
CO concentrations, the reaction was first order in CO, whereas at high CO and
O 2 concentrations, it is zero order in both CO and O 2 . The sharp decrease between
Downloaded by [McGill University Library] at 08:56 20 November 2012

the two domains is typical and was explained by the transition from an oxygen-
covered surface to a CO-covered one.
In addition to the abundant literature on Pt/SnO 2 , Mergler et al. (182) have
studied a number of Pt/SiO 2 catalysts promoted by metal oxides other than SnO2 .
They described the CO oxidation behavior of Pt/SiO 2 , Pt/CoO x /SiO 2 , and Pt/
MnO x /SiO 2 catalysts and compared them with commercially available Pt/Al 2 O 3 ,
Pt/Rh/Al 2 O 3 , and Pt/CeO x /Al 2 O 3 . Because the catalysts differed in metal disper-
sion and metal loading, comparison was made on the basis of turnover frequencies
(TOFs). The order in CO oxidation activity after a reductive pretreatment was
found as Pt/CoO x /SiO 2 ⬎ Pt/MnO x /SiO 2 ⬎ Pt/CeO x /Al 2 O 3 ⬎ Pt/Al 2 O 3 ⬎ Pt/
Rh/Al 2 O 3 ⬎ Pt/SiO 2 . It was shown that the addition of metal oxides such as CoO x
and MnO x to a standard Pt/SiO 2 catalyst does not change Pt particle size. The
results suggest that MnO x and CoO x are active components of Pt-based catalysts
in both CO oxidation and NO reduction reactions. CO oxidation occurs readily
at room temperature over Pt/CoO x /SiO 2 . A number of possible models accounting
for the high activity observed on Pt/CoO x /SiO 2 are discussed. Those in which O
vacancies on CoO x serve as dissociation centers for O 2 in the oxidation of CO
were found to be most acceptable.
The low-temperature oxidation of CO over Pt/Al 2 O 3 , CoO x /Al 2 O 3 , and Pt/
CoO x /Al 2 O 3 monolith catalysts has also been studied (183,184). Preoxidized CoO x-
containing catalysts were found to be highly active for CO oxidation at temperatures
as low as ⫺80°C and showed light-off temperatures of ⫺60°C. The activity of
cobalt-containing catalysts below 130°C was found to be independent of the pres-
ence of Pt, which made spillover processes unlikely, and the surface-coverage-de-
pendent sticking coefficients for CO and O 2 were suggested to be the reason for
the higher activity of cobalt-containing catalysts, compared with the Pt-only catalyst.
It appeared that CO was unable to block the cobalt surface to oxygen adsorption
as it does on Pt. At temperatures above 130°C, the presence of Pt in Pt/CoO x /Al 2 O 3
seemed to increase the reduction and oxidation rates of the cobalt oxide by activating
the H 2 and O 2 molecules, which could then participate in the spillover process.

Palladium Catalysts

Palladium can also form the basis of a low-temperature CO oxidation cata-


lyst. Bond et al. (143,180) showed that better activity is achieved on a mixture
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 63

of Pd over a transition-metal-oxide (stannia) support. Pd/SnO 2 catalysts prepared


by impregnation or by Pd ion exchange in a high-area SnO 2 gel showed that for
CO oxidation between 82°C and 140°C, (1) SnO 2 is a better support than SiO 2
and Al 2 O 3 , (2) the enhanced activity obtained is a result of spillover catalysis,
and (3) the spillover of reactants is from the metal onto the reducible oxide sup-
port. The reverse spillover of oxygen atoms was proposed to be much slower due
to the CO coverage on Pd.
A kinetic analysis was later conducted by Sheintuch et al. (185) who verified
the synergetic effect of Pd/SnO 2 in catalyzing CO oxidation at low temperatures.
Downloaded by [McGill University Library] at 08:56 20 November 2012

A model incorporating spillover of CO from Pd onto the SnO 2 support followed


by reaction with surface oxygen directly adsorbed from the gas phase was proposed.
The synergetic effect was supported by measurements for the 20–160°C interval,
which showed that both the CO inhibition effect and the activation energy on Pd/
SnO 2 (9.5–12 kcal/mol) are much lower than that on Pd/SiO 2 (45–47 kcal/mol).
In addition to Pd/SnO 2 catalysts, (Pd ⫹ Pt)/SnO 2 catalysts have also been
used for better activity decay characteristics. Upchurch et al. (175) have observed
a significant increase in activity for a Pd loading of 5 wt% on silica-gel-supported
Pt/SnO 2 catalysts with high Pt loadings. This was confirmed by Schryer et al.
(174) for unsupported 15–20 wt% Pt/SnO 2 catalysts for which the optimum Pd
loading was 5 wt%. Also, 1 wt% Pd ⫹ 1 wt% Pt over SnO 2 was found to give
higher CO conversions than 2 wt% Pt on SnO 2 . The cumulative evidence sup-
ported a reaction scheme in which two interdependent active sites participate in
CO oxidation. One site (precious metal) is mainly associated with the metal func-
tion of the catalyst, whereas the other is intrinsically dependent on the metal–
support interface.
Pd/MnO x and Ru/MnO x catalysts have been synthesized and tested for CO
oxidation in low stoichiometric concentrations of CO and O 2 at 30–75°C for pe-
riods extending up to 300 h (186). Although both catalysts were less active than
the other MnO x-supported noble metal catalysts (Au ⬎ Ag ⬎ Pd ⬎ Ru), significant
conversions were nevertheless observed, and it was suggested that they may be
improved by optimizing preparation and pretreatment techniques.
Carbon monoxide oxidation was studied over SnO 2 , Mn 2 O 3 , and coprecipi-
tated (SnO 2 ⫹ Mn 2 O 3 ) with or without 2 wt% Pd metal impregnation (187).
Mn 2 O 3 showed a high degree of labile oxygen, which can be effectively used for
low-temperature CO oxidation. The results suggested significant synergetic effects
between the two oxides and provided evidence for lattice oxygen incorporation
during CO oxidation. The CO oxidation activities of the oxides were in the order
SnO 2 ⬍ Mn 2 O 3 ⬍ (SnO 2 ⫹ Mn 2 O 3 ).
The mechanism of CO oxidation on the mixed oxide was proposed to consist
of the reduction of Mn 3⫹ to Mn 2⫹ followed by the donation of oxygen ions from
Sn 4⫹ to Mn 2⫹, where Sn cations acted as the oxygen carrier. The impregnation
of Pd metal drastically increased the CO conversion to about 70% at room temper-
ature.
Pavlova et al. (188) studied the non-steady-state and steady-state kinetics of
low-temperature CO oxidation over Pd on TiO 2 , SiO 2, and γ-Al 2 O 3 using a pulse/
ORDER REPRINTS

64 TRIMM AND ÖNSAN

flow system. In situ FTIR was used to monitor the Pd surface. Room-temperature
CO oxidation was demonstrated to be structure sensitive and to proceed via the
interaction between weakly bound CO (both linear and bridged) and oxygen lo-
cated at defect centers.

2. Base–Metal Oxide Catalysts

Base–metal oxide catalysts, especially copper oxides, have received atten-


Downloaded by [McGill University Library] at 08:56 20 November 2012

tion because they are known to exhibit CO oxidation activity per unit surface
comparable to those of precious metal catalysts (189). Hopcalite, which is a mix-
ture of CuO/MnO 2 plus small quantities of other oxides, has been the oldest cata-
lyst used for respiratory protection applications, and this has led to various studies
on the preparation and characterization of Cu–Mn oxides. However, the Hopcalite
catalysts used so far in life-rescue equipment are not resistant to moisture.
For CO oxidation over Cu/Al 2 O 3 and single-crystal Cu catalysts, it is re-
ported (190,191) that oxygen is strongly adsorbed on Cu, forming CuO x at temper-
atures ⬎300°C, whereas CO is weakly adsorbed and is inhibited by O 2 adsorption.
CO oxidation occurs with an activation energy of 18–22 kcal/mol via a Lang-
muir–Hinshelwood mechanism between CO and O 2 adsorbed on similar sites.
Elevated pressure kinetics, atomic emission spectrometry (AES), and TPD studies
on the CO oxidation activity of a Cu (100) single-crystal catalyst using different
CO/O 2 ratios (190) showed that the presence of a certain level of surface oxygen
is advantageous, but, under stoichiometric conditions, an oxide layer is formed
which reduces the catalytic activity observed on metallic Cu.
Platinum may act as a promoter of Cu catalysts for CO oxidation (189). Pt/
Al 2 O 3 , Cu/Al 2 O 3 , and Pt–Cu/Al 2 O 3 prepared by impregnation and subsequent
reduction at 500°C were used for CO oxidation at 120–140°C in gas mixtures
containing 1.2 vol% O 2 , 1.2 vol% CO, and balance N 2 . The increase observed in
the activity of the Cu-incorporated Pt catalyst (0.3 wt% Pt/8 wt% Cu) over the
monometallic catalysts was attributed not only to an SMSI phenomenon but also
to metal–metal interaction (Cu–Pt alloy formation).
In γ-Al 2 O 3-supported transition-metal-oxide catalysts, metal aluminate for-
mation that occurs during high-temperature pretreatment decreases activity. Small
amounts of ZnO increased the specific surface areas and total pore volumes of
CuO/Al 2 O 3 catalysts by preventing sintering and CuAl 2 O 4 formation (192). A
pretreatment temperature of 600°C was the maximum for obtaining increased CO
oxidation activity on CuO–ZnO/Al 2 O 3 catalysts. ZnO addition did not modify
the mechanism of the catalyzed reaction but decreased the concentration of active-
surface CuO crystallites.
Zhou et al. (193) investigated CuO/γ-Al 2 O 3 , CuO/ZrO 2 –Al 2 O 3 , and CuO/
ZrO 2 catalysts for CO oxidation at temperatures between 50°C and 300°C. The
results indicated that catalytic activity and stability against sintering increased as
the amount of ZrO 2 in the support increased: CuO/ZrO 2 ⬎ CuO/ZrO 2 –Al 2 O 3 ⬎
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 65

CuO/Al 2 O 3 . The synergy between Cu and Zr was caused by the high dispersion
of CuO on ZrO 2 , which increased the reduction ability and desorptability of sur-
face oxygen species.
Luo et al. (194) have prepared several ceria-supported copper oxide (CuO/
CeO 2 ) catalysts with CuO loadings between 0.25 and 15 wt%. The activities of
impregnated catalysts for CO oxidation were found to be higher than coprecipi-
tated catalysts. CeO 2 promoted the H 2 reduction of copper, so that CuO/CeO 2
catalysts behaved differently from pure CuO. Two types of reducible copper spe-
cies were observed in all catalysts: A highly dispersed CuO that can adsorb CO
Downloaded by [McGill University Library] at 08:56 20 November 2012

is reducible at low temperatures and is responsible for CO oxidation at low temper-


atures, and bulk CuO which cannot adsorb CO is reducible at high temperatures
only and contributes little to the oxidation activity of CuO/CeO 2 .

3. Summary

It is clear that an efficient low-temperature CO oxidation catalyst should


have appropriate sites for both CO chemisorption and simultaneous dissociative
O 2 adsorption at temperatures ⬍100°C. Composite catalysts containing a noble
metal and a reducible oxide component provide for the metal–support inter-
action(s) or bifunctional mechanism(s) that can activate CO oxidation. Although
reaction data indicate the presence of synergetic interactions, the nature of the
interactions and the reaction mechanisms are not fully understood. The major
determinants of catalyst performance are oxygen availability, composition of the
feed gas, and impurities.
Among the many gold-based catalysts studied, Au/MnO x seems to be a
promising catalyst with high long-term activity for low-temperature CO oxidation,
especially in O 2-rich and CO 2-deficient environments. It requires an oxidative pre-
treatment and is not much affected by the presence of some degree of moisture.
The most promising catalysts for efficient low-temperature CO oxidation are
Pt-based and transition-metal-oxide-supported catalysts with or without additional
promoters. These catalysts require a mild reductive pretreatment and some H 2 or
H 2 O in the feed gas to exhibit high activity. The addition of a small amount of
Pd enhances the CO oxidation activity. Although their CO 2 tolerance is better
than Au catalysts, long-term activity decay may be observed at high CO 2 partial
pressures. The addition of suitable promoter(s) improves both the activity and the
decay profiles. Their performance in stoichiometric CO/O 2 mixtures is far better
than other possible catalysts.

B. Low-Temperature CO Oxidation in H 2-Rich Gas Streams

Partial oxidation to convert fuels to hydrogen is usually associated with a


water–gas shift catalyst, the exit stream from which contains H 2 , CO 2 , H 2 O, and
ORDER REPRINTS

66 TRIMM AND ÖNSAN

1–3 vol% CO. As a result, it is necessary to consider the selective oxidation of


CO in the presence of H 2 and of the other gases listed earlier.
Three obvious possibilities can be considered. It may be possible to find a
catalyst that adsorbs CO but not H 2 , and hence favors selective oxidation of CO.
It may be possible to operate at a temperature where CO is oxidized and H 2 is
not. Finally, it may be possible to find a catalyst where both CO and H 2 are oxi-
dized, but kinetic parameters lead to preferential CO oxidation at the cost of only
small amounts of H 2 oxidation.
In all cases, it is critical to carefully control oxygen: carbon monoxide ratios
Downloaded by [McGill University Library] at 08:56 20 November 2012

and temperatures. It may be necessary to use more than one reactor to do so.
Catalysts identified as suitable for low-temperature CO oxidation are possi-
ble candidates for the selective oxidation.

1. Noble Metal Reducible Oxide Catalysts


Gold Catalysts

Haruta et al. (163) have prepared coprecipitated catalysts containing small


(⬃10 nm) Au particles and NiO, α-Fe 2 O 3 , and Co 3 O 4 . The catalysts were highly
active for both CO and H 2 oxidation, possibly due to the combined effect of gold
and the oxides. The oxidation of CO on Au/NiO and Au/Co 3 O 4 was found to be
complete at 30°C, even in the presence of 76% relative humidity. In contrast, 50%
of the H 2 was oxidized only at 73°C and 66°C, respectively.
The general trend observed was that catalytic activity increased with decreas-
ing Au crystallite size in the oxidation of both H 2 and CO. It is interesting that
the catalytic activity of coprecipitated Au/γ-Al 2 O 3 with similar Au particle sizes
was comparable to those of Au/NiO, Au/α-Fe 2 O 3 , and Au/Co 3 O 4 for H 2 oxidation
but was remarkably lower for CO oxidation. The activity for H 2 oxidation was
solely dependent on the gold metal surface exposed, whereas the choice of support
oxide also played a role in CO oxidation through either a metal–support interac-
tion or via a bifunctional mechanism in which gold and the support oxide activate
different steps of CO oxidation. Hydrogen is taken not to chemisorb on Au,
whereas CO chemisorbs weakly on the metal. Thus, the catalytic activity for CO
may be a result of the interaction of species adsorbed on the metal and the support.
Alternatively, the interaction of Au with the reducible oxide support may change
the surface properties of gold particles so that they favor CO adsorption. It was
reported (166,195) that at pretreatment temperatures ⱖ300°C, the decomposition
of oxidic gold species to metallic Au and formation of crystalline hematite (α-
Fe 2 O 3 ) occur to yield hemispherical Au particles ⬃4 nm in diameter stabilized
by epitaxial contact over α-Fe 2 O 3 .
Various transition metal oxide supports giving the largest difference in cata-
lytic activity between CO and H 2 oxidation were screened, and coprecipitated Au/
MnO x was chosen by Haruta et al. (158,163,196,197) as one of the best catalysts
for CO removal from H 2-rich gases. Because Au catalysts require an oxidative
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 67

pretreatment, the stability of Au/MnO x in H 2 was tested. CO oxidation reached


100% conversion at temperatures below 0°C, whereas H 2 oxidation occurred only
at temperatures above 0°C. The temperature difference was found to be closely
related to the Au content and the Au/Mn atomic ratio in the catalyst, larger temper-
ature differences being obtained at Au/Mn ratios ⬎ 1/20. In streams containing
1 vol% CO and 1 vol% O 2 in H 2 , an Au/MnO x catalyst with Au/Mn ⫽ 1/50
selectively oxidized CO to give ⬃50% conversion at 16°C. Conversions ⬎95%
were obtained between 50°C and 80°C where catalyst activity was stable for 2
days and improved with time-on-stream. At 120°C, CO conversion started to de-
Downloaded by [McGill University Library] at 08:56 20 November 2012

cline, implying that O 2 is competitively consumed by H 2 oxidation.


The stability of Au/MnO x catalysts in H 2 atmospheres was attributed to the
reversible transformation between three Mn oxides, MnO, Mn 2 O 3 and Mn 3 O 4 ,
depending on the O 2 concentration and temperature. In comparison with Pt/zeo-
lite-A catalysts (198,199) that exhibit catalytic activity at temperatures around
200°C under similar conditions, Au/MnO x catalysts operating below 130°C were
found to have several advantages: (1) they were more active for CO oxidation
than for H 2 oxidation, (2) their catalytic activity was increased by the presence
of moisture, and (3) they were less sensitive to the CO 2 present in the reformate
gases.
The selective oxidation of CO in H 2-rich gas streams was investigated over a
number of Au/γ-Al 2 O 3 catalysts prepared by the deposition–precipitation method
(200). The major finding of the study was that the oxidation of low concentrations
of CO in a H 2-rich gas mixture (H 2 /CO/O 2 /He ⫽ 48/1.0/0.5/50.5) can be
achieved with 50% selectivity at 100°C over Au/γ-Al 2 O 3 . Selectivity was defined
as the O 2 consumption for CO oxidation divided by the total O 2 consumption, or
(CO 2 )/(CO 2 ⫹ H 2 O). Both the activity and the selectivity were found to depend
on the Au crystallite size, with an optimum particle size between 5 and 10 nm,
which is in good agreement with results reported for low-temperature CO oxida-
tion on Au catalysts in the absence and presence of H 2 (170,171,173). The increase
in activity with decreasing Au particle size was explained not by increased oxygen
activation at the Au–Al 2 O 3 interface as proposed by Haruta et al. (158), but rather
by the differences in the rates and strengths of adsorption of H 2 and CO on Au
particles.
It was found (200) that the metal particle size on γ-Al 2 O 3 can be controlled
by the addition of Mg citrate to the precursor solution during the deposition–
precipitation of chlorohydroxy Au clusters, as also reported by Haruta and co-
workers (154). This is explained by the fact that citrate is a strong ligand for the
Au 3⫹ ion and can compete with the hydroxyls for coordination to the Au ion,
displacing some of the hydroxyl ligands and thus breaking up the Au clusters.
The preferential oxidation of CO in H 2-rich gas was investigated over Au/
α-Fe 2 O 3 , using simulated reformate (75 kPa H 2 , 0.025–1.5 kPa CO, balance N 2 )
at atmospheric pressure with p O 2 /p CO ratios between 0.25 and 10 (201). The α-
Fe 2 O 3 support was chosen because of the low T1/2 value (the temperature at which
50% conversion occurs) of Au/α-Fe 2 O 3 for low-temperature CO oxidation
ORDER REPRINTS

68 TRIMM AND ÖNSAN

(158,202). A strong initial deactivation of ⬃30% was observed in the first 2 h


time-on-stream, which slowed down during the next 8 h and finally reached steady
state; a similar deactivation behavior was also reported for Au/TiO 2 catalysts
(148,158).
Kinetic studies conducted at 80°C gave CO oxidation rates in terms of
p O 2 and p CO : the reaction orders in O 2 and CO were calculated to be 0.27 and
0.55, respectively, with an apparent activation energy of 31 kJ/mol in the 40–
100°C interval. Kahlich et al. (201) have found the H 2 oxidation rate to be indepen-
dent of p CO , consistent with a reaction mechanism where oxygen adsorbed at the
Downloaded by [McGill University Library] at 08:56 20 November 2012

metal–oxide interface reacts with hydrogen and CO weakly adsorbed on supported


Au clusters. At constant p CO , constant selectivity was obtained over the entire
p O 2 range, indicating that the reaction order with respect to O 2 is the same for
both CO and H 2 oxidation. At 80°C and high p CO , the selectivity defined as CO 2 /
(CO 2 ⫹ H 2 O) reached 75%, but it was considerably lower at low p CO . Selectivity
for CO oxidation increased with decreasing temperature, reflecting the higher ap-
parent E A for H 2 oxidation (50 kJ/mol) than for CO oxidation (31 kJ/mol). The
water–gas shift reactions were negligible.
When compared with the commonly used Pt/γ-Al 2 O 3 catalyst, which has an
optimum temperature of 200°C (203), Au/α-Fe 2 O 3 achieved comparable activity
at the lower temperature of 80°C. However, there is still the question as to how
much carbon monoxide can be removed.
When considering the choice of Au-based catalysts, lower-temperature cata-
lysts offer higher selectivity. However, the reformate stream exits the reactor at
250–350°C and the fuel cell operates at 80–100°C. It may be preferable to accept
lower selectivities rather than pass through a cooling/oxidation/heating cycle.

Platinum-Group Metal Catalysts

Catalysts proposed for selective CO oxidation in H 2-rich gases are alumina-


supported Pt, Ru, and Rh, operating at temperatures in the 120–160°C range (203–
207). In a H 2-rich mixture, containing ⬃1 vol% CO, the CO is completely oxi-
dized by four times the stoichiometric amount of oxygen, corresponding to a
process selectivity of ⬃25%. The temperature dependence of the CO consumption
rate and the CO oxidation selectivity in H 2-rich gases were investigated for Al 2 O 3-
supported Pt metals (204,205,207) and zeolite-supported Pt (198). Oh and Sinkev-
itch (204) have observed selectivities of ⬃40%, ⬃80%, and ⬃80% for Pt/γ-
Al 2 O 3 , Ru/γ-Al 2 O 3 , and Rh/γ-Al 2 O 3 , respectively, at CO conversions close to
100% in H 2-rich gas. The selectivities observed by Brown et al. (205) under condi-
tions more similar to reformate gases are ⬃25% on Ru/γ-Al 2 O 3 and Rh/γ-Al 2 O 3 ,
and those over Pt/zeolite-A and Pt/γ-Al 2 O 3 range between 40% and 50% for al-
most complete CO conversion (198).
Although the selectivities are inferior to Au-based catalysts, they offer better
performance than methanation
CO ⫹ 3H 2 ⫽ H 2 O ⫹ CH 4 (20)
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 69

and Pt-group catalysts have been used for many different applications. As a result,
confidence is higher for precious metal catalysts than for Au-based systems!
A variety of catalytic materials, including noble metals (Pt, Pd, Rh, Ru, all
with 0.5 wt% metal dispersed on alumina) and base metals (Co/Cu, Ni/Co/Fe,
Ag, Cr, Fe, Mn), were screened by Oh and Sinkevitch (204). Using a feed stream
containing nearly 10-fold excess of H 2 over CO (0.85 vol% H 2 , 900 ppm CO, 0–
2300 ppm O 2 , balance N 2 ), temperatures needed to achieve 50% conversion (T1/2 )
were recorded. Among the base metal catalysts tested, the cobalt-containing cata-
lysts exhibited the highest CO oxidation activity, but because their T1/2 values
Downloaded by [McGill University Library] at 08:56 20 November 2012

were found to be 70–80°C higher than the Pt/γ-Al 2 O 3 catalyst, they were not
studied further. Both Ru/γ-Al 2 O 3 and Rh/γ-Al 2 O 3 were the most active catalysts
for CO oxidation at 100°C, their T1/2 values being ⬃70°C lower than that for Pt/
γ-Al 2 O 3 . Over the temperature range of interest (i.e., between 100°C and 200°C),
CO oxidation activity in H 2-rich gas decreased in the order Ru/γ-Al 2 O 3 ⬎ Rh/
γ-Al 2 O 3 ⬎ Pt/γ-Al 2 O 3 ⬎ Pd/γ-Al 2 O 3 .
The CO conversions on all four noble metal catalysts exhibited a maximum
with increasing temperature, and the decline in CO conversion observed at higher
temperatures was attributed to the water–gas equilibrium leading to increased con-
sumption of the limited O 2 supply by H 2 . These results indicate that high-tempera-
ture operation is to be avoided because, under such conditions, H 2 is preferentially
removed rather than the CO contaminant. For each noble metal catalyst, there was
an optimum O 2 concentration at which complete CO conversion was achieved,
and increasing the O 2 concentration beyond the optimum simply increased H 2
oxidation.
In addition to their higher CO oxidation activity, Ru/γ-Al 2 O 3 and Rh/γ-
Al 2 O 3 catalysts consumed substantially less H 2 than the Pt/γ-Al 2 O 3 catalyst. Over
Pt/γ-Al 2 O 3 , complete removal of CO was accompanied by 20% H 2 conversion,
whereas the H 2 conversion over Ru/γ-Al 2 O 3 and Rh/γ-Al 2 O 3 could be kept
below 5%.
The detailed mechanism of the simultaneous oxidation of CO and H 2 over
noble metals is not yet fully understood. H 2 oxidation is strongly inhibited by the
presence of CO (27,208), because CO chemisorption on noble metal surfaces is
much stronger than H 2 or O 2 chemisorption (209,210). Consequently, CO blankets
the metal surface, displacing the weakly chemisorbed H 2 and O 2 species, and
prevents reaction unless the temperature is high enough to desorb some of the
CO on the surface. This indicates that the light-off behavior of noble metals in
CO–H 2 –O 2 mixtures is dominated by the kinetics of CO oxidation rather than by
the kinetics of H 2 oxidation (204). Alternatively, the H 2 in the feed can interact
with CO chemisorbed on the surface to form a complex such as H–CO; its easier
desorption from the surface may increase the CO oxidation activity significantly
(27,208).
Kahlich et al. (203) have studied the selective oxidation of CO on Pt/γ-
Al 2 O 3 , using simulated reformer gas (75 vol% H 2 with N 2 background) over a
range of CO concentrations (0.02–1.5 vol%) at low stoichiometric O 2 excess. The
optimum temperature for the preferential oxidation process over Pt/γ-Al 2 O 3 was
ORDER REPRINTS

70 TRIMM AND ÖNSAN

around 200°C. Kinetic studies conducted for the 150–250°C interval showed that
the rate can be expressed by a simple power-law expression with reaction orders
of 0.80 for p O2 and ⫺0.40 for p CO and an apparent activation energy of 71 kJ/
mol. These results are consistent with the reaction occurring on a surface predomi-
nantly covered by adsorbed CO. This is responsible for the high selectivity of
40% for CO oxidation. The selectivity loss above 200°C is marked by the onset
of CO desorption accompanied by an increase in H 2 oxidation.
Removal of CO from H 2-rich fuels by selective oxidation has been studied
at 150–200°C over zeolite-supported Pt catalysts containing 5.8–6.0 wt% Pt
Downloaded by [McGill University Library] at 08:56 20 November 2012

(199). Pt particle sizes were less than 1 nm on Pt–zeolite A, Pt–mordenite, and


Pt–zeolite X, whereas ⬃7 nm particles were observed on Pt/Al 2 O 3 . Using a reac-
tion mixture consisting of 1% CO, 2% O 2 , and the balance H 2 , the selectivity was
found to be affected by the type of support used in the following order: Pt–zeolite
A ⬎ Pt–mordenite ⬎ Pt–zeolite X ⬎ Pt–alumina. The oxygen content in the feed
was an important parameter. The zeolite-supported Pt catalysts oxidized CO much
more selectively in a large excess of H 2 at low O 2 partial pressures than the con-
ventional Pt/Al 2 O 3 catalyst, approaching 100% conversion with decreasing oxy-
gen concentrations at ⬃200°C. Conventional Pt/Al 2 O 3 required 3% oxygen to
remove CO, but Pt–mordenite required only 1.5% oxygen addition, which could
be reduced down to 0.7% by using a two-stage reactor. Pt–mordenite was sug-
gested as a promising catalyst for CO removal from H 2-rich streams because it
also had resistance to water in the feed.

2. Copper-Based Catalysts

Supported Cu catalysts were investigated for CO removal in postreforming


gas mixtures (211–213). Both CO oxidation and the water–gas shift reaction were
effective in reducing the CO concentration in the presence of a large amount of
H 2 O and H 2 . The positive effect of O 2 addition was also investigated. The mixed-
oxide catalyst Cu/Al 2 O 3 –ZnO exhibited excellent activity for removal of a small
amount of CO in the reformate from methanol. Catalytic activity was strongly
related to Cu particle size. Spinel-type Cu–Al–Zn oxides gave rise to fine Cu
particles. Catalysts calcined at 500–700°C achieved 90% conversion for CO re-
moval at the reaction temperature of 150°C in the presence of ⬃2% O 2 .
No studies of the removal of CO from hydrocarbon reforming streams have
been found. By analogy with Hopcalite, low-temperature oxidation can be ex-
pected to be inhibited by the higher amounts of water.

3. Summary

Reforming gas streams from the fuel processor contain H 2 O and CO 2 as well
as CO and H 2 . The exit temperatures for achieving relatively low CO concentra-
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 71

tions are around 150–250°C. Fuel cells operate at 80–100°C and can be poisoned/
inhibited by the CO present in the reformate. Excess H 2 O from the reformer can
be removed by condensation, but the gas will still be fully humidified.
The amount of oxygen to be injected for the removal of CO by selective
oxidation is very important, as excess O 2 in the system may also oxidize the H 2
in the feed to the fuel cell. Therefore, a catalyst that can selectively oxidize CO
in the presence of H 2 O and CO 2 and using stoichiometric amounts of O 2 is needed.
High-temperature operation is undesirable because high temperatures promote H 2
oxidation.
Downloaded by [McGill University Library] at 08:56 20 November 2012

Gold-based catalysts having reducible oxide components would appear to


be promising, because they are more active for CO oxidation than for H 2 oxidation
and achieve high activity at temperatures lower than those required by the Pt-group
metals. These catalysts are resistant to moisture and appear to be less sensitive to
CO 2 when excess H 2 is present in the reaction mixture (196,197). Because Au-
based catalysts require an oxidative pretreatment, their stability in H 2 atmospheres
is likely to depend on the transformations between the oxide forms of the support
material and, hence, on the O 2 level and the temperature. The level to which CO
can be reduced over such catalysts is also open to question.
Among the noble metal catalysts proposed for selective CO oxidation in H 2-
rich gases are zeolite- or Al 2 O 3-supported Pt, Ru, and Rh catalysts. Research re-
lated to these catalysts seems to be rather limited. Studies on the selective oxida-
tion of CO over Pt/γ-Al 2 O 3 (203) indicate that the presence of H 2 in a CO–O 2
mixture reduces the ignition temperature by about 30°C, corresponding to substan-
tial rate enhancement at about 150°C. A similar enhancement of the CO oxidation
rate is also produced by adding water vapor (214). The limited information avail-
able on selective CO oxidation over Ru and Rh catalysts also looks promising.
One has to accept that familiarity does influence the choice of catalyst. As
a result, the use of Pt–zeolite or Pt–mordenite or promoted Pt/Al 2 O 3 catalysts
would not be unexpected in preference to Au-based catalysts. Temperature control
by the use of more than one reactor would be desired and is a feature of the
Johnson–Matthey system (12).

VI. WHICH FUEL? WHICH SYSTEM?

The choice of a fuel and of a conversion system is far from simple. Unless
there is a breakthrough in fuel-cell technology, the use of hydrogen is essential.
It is necessary, then, to consider the hydrogen-producing potential of various fuels
and the merits of indirect and direct partial oxidation for application in vehicles.
Although this review is focused on catalytic methods of fuel conversion,
serious consideration has been given to gas-phase noncatalytic conversion (97).
The process has the advantage that it can be used with a wide range of fuels, but
it has the disadvantage that coke and unwanted emissions may be produced (11).
The use of a catalyst has been recommended to minimize coking (11,16).
ORDER REPRINTS

72 TRIMM AND ÖNSAN

The main catalytic reactions involved in hydrogen production are steam re-
forming, water–gas shift, and direct partial oxidation:
CH 4 ⫹ H 2 O ⫽ CO ⫹ 3H 2 (2)
CH 4 ⫹ 1/2 O 2 ⫽ CO ⫹ 2H 2 (3)
CO ⫹ H 2 O ⫽ CO 2 ⫹ H 2 (4)
Steam reforming would appear to produce more hydrogen, but the reaction is
endothermic and part of the fuel has to be burned to provide the necessary heat.
Downloaded by [McGill University Library] at 08:56 20 November 2012

This affects the overall yield of hydrogen from fuel.


Because the reactions are used to produce hydrogen, extra water may be
needed to adjust the water–gas shift equilibrium:
PCO2 PH2
K⫽
PCO PH2O

Any extra water must be gasified, and this will require more heat and more total
oxidation of input fuel.
In the case of steam reforming, extra water may also be needed to avoid
coking (76).
Catalysts used in direct and indirect oxidation and in the water–gas shift
reaction are metallic, and it is essential to consider fuel-containing minimal sulfur.
Avci et al. (215) have carried out preliminary analysis of the merits of indi-
rect and direct partial oxidation, based on thermodynamic equilibria. The conver-
sion of methanol, methane, propane (as a model for LPG), and octane (as a model
for gasoline) to hydrogen has been considered. It is useful to illustrate the approach
using the conversion of propane before summarizing the results for other fuels.
Considering first the indirect partial oxidation, the temperature at which pro-
pane steam reforming becomes very slow (light-out temperature) has been re-
ported to be 350°C (18). The thermodynamic equilibrium data for steam reforming
and water–gas shift at this temperature allow calculation of the concentrations of
exit gases, given an assumed value of steam:
C 3 H 8 ⫹ 3H 2 O ⫽ 3CO ⫹ 7H 2 (21)
CO ⫹ H 2 O ⫽ CO 2 ⫹ H 2 (4)
Accepting these temperatures and concentrations, it is possible to calculate
backward to obtain conversion levels versus temperature as a function of inlet
concentrations. Because steam reforming is endothermic, the temperature in-
creases as propane conversion decreases.
Total oxidation of propane is needed to provide the heat needed to drive the
steam reforming reaction and to gasify/heat the inlet reagents. The greater the
total oxidation, the larger the temperature rise that can be achieved and the lower
the amount of propane available for steam reforming.
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 73

Optimization procedures then allow a match of the amount of propane totally


oxidized with the amount of propane to be steam reformed, using the maximum
temperature as the controlling parameter. This maximum occurs at the end of the
oxidation regime and the start of the steam reforming regime.
Because hydrogen is required, sufficient steam must be present to favor the
water–gas shift reaction equilibrium and to avoid coking during steam reforming
(76). Extra water is needed, and the water must be gasified—an additional load
for heat production.
Propane conversion may be incomplete at the conditions relevant to light
Downloaded by [McGill University Library] at 08:56 20 November 2012

out. At the same time, ⬃25% of hydrogen supplied to the fuel cell will be rejected
(12). Residual propane and rejected hydrogen may be catalytically oxidized in a
separate bed, to provide heat to the incoming gases (12). This will result in less
propane being oxidized in the main reactor and more propane being available for
steam reforming.
The light-out temperature is ⬃350°C, but the water–gas shift reaction favors
hydrogen at lower temperatures. Heat exchange between the inlet and exit gases
to the main reactor helps to maximize efficiency on the inlet and reduce the tem-
perature of the product stream.
Avci et al. (215,216) have carried out a full analysis of the overall process
for obtaining hydrogen from various fuels, which takes account of the factors
outlined earlier. Typical results for thermodynamic calculations using stoichio-
metric ratios of fuel and oxygen are summarized in Table 8.
The yield of hydrogen is seen to be very dependent on the fuel and on the
fuel :oxygen and fuel : steam ratios (Table 8). The theoretical yields vary as the
molecular formula of the fuel:
C n H 2n⫹2 ⫹ nH 2 O ⫽ nCO ⫹ (2nH)H 2
nCO ⫹ nH 2 O ⫽ nCO 2 ⫹ nH 2
and the difference from the calculated yield reflects the amount of fuel oxidized
in the system.
It should be noted that the calculations refer to steady state. A given fuel
will only start to oxidize at the appropriate light-off temperature (18). Heating
the inlet gases from ambient to this temperature may require oxidation of stored
hydrogen or electrical heating. In the case of methanol, the system will light off
at room temperature, whereas for methane, heating to ⬃300°C is needed.
Calculations were then carried out to identify the optimal water: carbon ratios
for different fuels. These are summarized in Table 9 and Figure 2. Conversion
first increased with increasing water :carbon ratios, reflecting the effect of extra
water on the water–gas shift equilibrium. Eventually, yields decrease with increas-
ing ratios, reflecting a decrease in the amount of fuel available for steam reforming.
Optimal water :fuel ratios and hydrogen yields are reported in Table 9.
The maximum for methanol occurs at lower steam :carbon ratios, due to the
presence of oxygen in the molecule. It should be noted that the maximum yield
Downloaded by [McGill University Library] at 08:56 20 November 2012

74

Table 8. Data Used for Thermodynamic Calculations: Hydrogen Yields


Indirect Partial Oxidation Direct Partial Oxidation
ORDER

Light-off Yield a Yield a


c d b
Fuel (°C) H 2 O/C ox (%) sr (%) Theory Calculated Yield H 2 O/C Theory Calculated Yield b
Methane 360 3.8 16.0 70.6 400 237 690 0.8 300 240 1470
Propane 169 2.0 20.0 97.2 1000 770 1306 0.9 700 665 1300
REPRINTS

Iso-octane 240 2.1 25.0 99.9 2500 1864 1220 0.9 1700 1646 1250
Methanol 25 1.1 5.0 99.9 300 283 1070 0.8 200 60 250
a
Yield ⫽ (moles of H 2 produced/moles of fuel fed to the reactor) ⫻ 100.
b
Yield ⫽ volume of H 2 produced/mass of fuel ⫹ water fed (mL/gr).
c
ox ⫽ oxidized fuel.
d
sr ⫽ steam reformed fuel.
TRIMM AND ÖNSAN
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 75

Table 9. Calculated Hydrogen Yields Based on the Carried Weight

Indirect Partial Oxidation Direct Partial Oxidation


a b
Fuel Water: Fuel Yield Water :Fuel a Yield b

Methane 2.8 690 0.8 1470


Propane 3.9 1310 2.7 1300
Iso-octane 10.1 1220 7.2 1250
Methanol 1.0 1070 0.8 250
a
Water: fuel ⫽ moles of injected water/moles of injected fuel to the reaction
Downloaded by [McGill University Library] at 08:56 20 November 2012

system.
b
Yield ⫽ volume of H 2 produced/mass of fuel ⫹ water fed (mL/gr).

is obtained at a steam: carbon ratio of less than 2.5 for methanol, propane, and
octane. Coking is not a problem if copper-based catalysts are used for methanol
(13), but could be a problem if nickel-based catalysts are used for propane and
octane (76). An economic balance has to be drawn between the use of precious
metal catalysts (111) at low steam :carbon ratios (which do not favor coking) and
nickel-based catalysts at higher steam :carbon ratios but smaller yields of hy-
drogen.
Although thermodynamic data for the direct conversion of fuels was avail-
able, product yields had been measured experimentally (14) and were used in
similar calculations.

Figure 2. Change of molar hydrogen yield with steam-to-carbon ratio for steam reforming reaction
in indirect partial oxidation of propane.
ORDER REPRINTS

76 TRIMM AND ÖNSAN

Propane conversion was reported to be 95% with 95% selectivity to hydro-


gen. Because the reaction is carried out at very short residence times and at high
temperatures (⬃1000°C), control of heat transfer from the products to the reactants
is essential. Further hydrogen could be produced by the water–gas shift reaction,
but temperatures must be reduced (70):
C 3 H 8 ⫹ 3/2 O 2 ⫽ 3CO ⫹ 4H 2 (22)
CO ⫹ H 2 O ⫽ CO 2 ⫹ H 2 (4)
At steady state, the exit gases were assumed to be heat exchanged (70% efficiency)
Downloaded by [McGill University Library] at 08:56 20 November 2012

with the inlet gases, after which liquid water was injected into the exit stream to
cool the products. This had the effect of both cooling the gases and applying the
water needed to favor the production of hydrogen from the water–gas shift reac-
tion. Product yields were dictated by the water–gas shift equilibrium calculated
for 100°C (Fig. 3).
Total oxidation of propane was not required, because heat exchange from
the exit to the inlet plus the exothermicity of the partial oxidation plus the catalytic
oxidation of the rejected fuel-cell hydrogen was sufficient to provide the steady-
state heating needs.
Also reported are yields of hydrogen in terms of weight of fuel plus water
carried, because extra weight decreases fuel efficiency. It is clear that propane
(used as a model for LPG) or iso-octane (gasoline) are the best fuels in this respect
(Table 9).

Figure 3. Change of molar hydrogen yield with steam-to-carbon ratio for steam reforming reaction
in direct partial oxidation of propane.
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 77

Coking is not a major problem in the process, and low steam :carbon ratios
are no problem with Rh-based catalysis.
In terms of vehicle fuel efficiency, it is the amount of hydrogen per weight
of fuel plus water carried that is critical. Values calculated for the different fuels
and processes are listed in Table 9. It is seen that the hydrogen produced per unit
weight of reactants is highest with propane and octane in both cases (neglecting
methane). The results also show that extra water above that produced by the
reactions is needed, inferring the necessity to refuel a vehicle both with fuel and
water.
Downloaded by [McGill University Library] at 08:56 20 November 2012

Analysis of the kinetics of the various processes is now in progress, allowing


comparisons of the efficiency of the fuels under more realistic conditions (215).
Preliminary results indicate that LPG conversion by indirect partial oxidation re-
mains a preferred system.
It must be noted that the above-considered processes produce ⬃2% CO after
application of the water–gas shift equilibrium (12) and that selective oxidation
of the gas will be needed. Catalysts that can selectively oxidize CO in the presence
of H 2 O and CO 2 and using stoichiometric amounts of O 2 have to be used. Opera-
tion at fuel-cell temperatures will be preferable. Au-based catalysts having reduc-
ible oxide components appear to be promising, if their durability in CO 2-con-
taining H 2-rich atmospheres can be improved. There is some question as to their
efficiency at removing the last traces of CO. The use of Pt–zeolite or promoted
Pt/Al 2 O 3 catalysts that operate well under reducing conditions would seem to be
preferred, but careful control of temperature is essential.
Calculations for stoichiometric ratios of water :fuel are shown in Table 8 for
the various fuels. The theoretical yields are less than those for indirect partial
oxidation as a result of the nature of the chemical reactions involved. Iso-octane
is seen to be the best fuel.
Calculations were repeated to determine optimal water :fuel ratios (Figs. 2
and 3 and Table 9). Because water was only added after the reactor, the yield did
not decrease once the optimum had been reached (Fig. 3). The optimum values
together with hydrogen yield per weight carried are reported in Table 9. Although
methane provides the best yield, this value does not take into account the weight
of gas cylinders, and propane or iso-octane again emerge as the preferred fuels.
Again, these calculations refer to steady-state conditions and start-up heating was
not considered.
It is clear that the indirect oxidation of propane (and, by analogy, LPG)
and octane (gasoline) or the direct oxidation of the two fuels are preferred. The
differences in yields are small, and the difficulties of maintaining closely con-
trolled fuel :air ratios and high temperatures indicate that the indirect partial oxida-
tion of propane or octane is probably the best process.
The yields of hydrogen are not dramatically reduced by operating at a water :
fuel ratio of 2.5. As a result, nickel-based catalysts operating at this ratio (in order
to minimize coking) are the preferred steam reforming catalysts. Platinum should
ORDER REPRINTS

78 TRIMM AND ÖNSAN

be deposited on these catalysts in order to promote low-temperature light off and


oxidation.

ACKNOWLEDGMENT

This work was conducted in Spring 2000, during Professor D. L. Trimm’s


sabbatical leave at Bogaziçi University, Istanbul. The authors wish to express their
gratitude to the university for the opportunity and the support given.
Downloaded by [McGill University Library] at 08:56 20 November 2012

REFERENCES

1. Cooper, B.J. Plat. Metals Rev. 1994, 38, 2.


2. Burke, N.R.; Trimm, D.L.; Howe, R. to be presented at the 12th International Con-
gress on Catalysts, Granada, 2000.
3. Matsumoto, S.; Ikeda, Y.; Suzuki, H.; Ogai, M.; Miyoshi, N. Appl. Catal. B: Envi-
ron. 2000, 25, 115.
4. Ralph, T.R.; Hards, G.A. Chem. Ind. 1998, 327.
5. Raman, V. Chem. Ind. 1997, 771.
6. Prigeut, M.; Dezael, C.; Breele, Y. Proc. 5th IECEC, 1970, p. 111.
7. Vielstich, W.; Iwasita, T. In Handbook of Heterogeneous Catalysis; Ertl, G., Knoz-
inger, H., Weitkamp, J., Eds.; Wiley–VCH: New York, 1997; Vol. 4, 2090.
8. Cameron, D.S. Plat. Metals Rev. 1999, 43, 149.
9. Hoogers, G.; Thompsett, D. Chem. Ind. 1999, 796.
10. Hill, S. New Sci. 1996, 20.
11. Jamal, Y.; Wyszynski, M.L. Int. J. Hydrogen Energy 1994, 19, 557.
12. Golunski, S. Plat. Metals Rev. 1998, 42, 2.
13. Jiang, C.; Trimm, D.L.; Wainwright, M.S. Chem. Eng. Tech. 1995, 18, 1.
14. Pena, M.A.; Gomez, J.P.; Fierro, J.L.G. Appl. Catal. A: Gen. 1996, 144, 7.
15. Rostrup-Nielsen, J.R. In Methane Conversion; Bibley, D.M., et al., Eds.; Elsevier:
Amsterdam, 1998.
16. Houseman, J.; Cerini, D.J. SAE Paper 740600, 1974.
17. Trimm, D.L. In Handbook of Heterogeneous Catalysis; Ertl, G.; Knozinger, H.;
Weitkamp, J., Eds.; Wiley–VCH: New York, 1997; Vol. 3, 1280.
18. Ma, L. Ph.D. thesis, UNSW, Australia, 1995.
19. Hiam, L.; Wise, H.; Chailein, S. J. Catal. 1968, 9, 279.
20. Lee, J.H.; Trimm, D.L.; Cant, N.W. Catal. Today 1999, 47, 353.
21. McCabe, R.W.; McCready, D.F. J. Phys. Chem. 1986, 90, 1428.
22. McCabe, R.W.; Mitchell, P.J. Appl. Catal. 1986, 27, 83.
23. McCabe, R.W.; Mitchell, P.J. J. Catal. 1987, 103, 419.
24. Jiang, C. Ph.D. thesis, UNSW, Australia, 1992.
25. Acres, G.J.K. Plat. Metals Rev. 1966, 10, 60.
26. Dus, R.; Tompkins, F.C. Proc. Roy. Soc. Series A 1975, 343, 477.
27. Dabill, D.W.; Gentry, S.J.; Holland, H.B.; Jones, A. J. Catal. 1978, 53, 164.
28. Moro-Oka, Y.; Ozaki, A. J. Catal. 1966, 5, 116.
29. Moro-Oka, Y.; Morikawa, Y.; Ozaki, A. J. Catal. 1967, 7, 23.
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 79

30. Trimm, D.L. Appl. Catal. 1983, 7, 249.


31. Lahouse, C.; Bernier, A.; Grange, P.; Delmon, B.; Papaefthimiou, P.; Ioannides,
T.; Verykios, X. J. Catal. 1998, 178, 214.
32. Heck, R.M.; Farrauto, R.J. Catalytic Air Pollution Technology: Commercial Tech-
nology; Van Nostrand Reinhold: New York, 1995.
33. Dissanayake, D.; Rosynek, M.P.; Kharas, K.C.C.; Lunsford, J.H. J. Catal. 1991,
132, 117.
34. Farrauto, R.J.; Hobson, M.C.; Kenelly, T.; Waterman, E.M. Appl. Catal. A: Gen.
1992, 81, 227.
35. Yazawa, Y.; Yoshida, H.; Tagaki, N.; Komai, S.; Satsuma, A.; Hattori, T. J. Catal.
Downloaded by [McGill University Library] at 08:56 20 November 2012

1999, 187, 15.


36. Baldwin, T.R.; Burch, A. Appl. Catal. 1990, 66, 337, 359.
37. Fornasiero, P.; Kaspar, J.; Sergu, V.; Graziani, M. J. Catal. 1999, 182, 56.
38. Epling, W.S.; Hoflund, G.B. J. Catal. 1999, 182, 5.
39. Carstens, J.N.; Su, S.S.; Bell, A.T. J. Catal. 1998, 176, 136.
40. Fujimoto, K.; Ribeiro, F.H.; Avales-Borja, M.; Iglesia, E. J. Catal. 1998, 179, 431.
41. Ma, L.; Trimm, D.L.; Jiang, C. Appl. Catal. A: Gen. 1996, 138, 275.
42. Burke, N.; Trimm, D.L. Unpublished results.
43. Paulis, M.; Gandia, L.M.; Gil, A.; Sambeth, J.; Odriozok, J.A.; Montes, M. Appl.
Catal. B: Environ. 2000, 26, 37.
44. Patterson, M.J.; Angrove, D.E.; Cant, N.W. Appl. Catal. B: Environ. 2000, 26, 47.
45. Santacesaria, E.; Carra, S. Riv. Combust. 1978, 32 (7–8), 228.
46. Amphlett, J.C.; Evans, M.J.; Mann, R.F.; Weir, R.D. Can. J. Chem. Eng., 1985,
63, 605.
47. Kobayashi, H.; Takezawa, N.; Minochio, O. J. Catal. 1981, 69, 487.
48. Pour, V.; Barton, J.; Benda, A. Coll. Czech. Chem. Commun. 1975, 40, 2923.
49. Agaros, H.; Cerella, G.; Laborde, M.A. Appl. Catal. 1988, 45, 53.
50. Matsukata, M.; Vemiya, S.; Kikuchi, E. Chem. Lett. 1988, 761.
51. Jiang, C.J.; Trimm, D.L.; Wainwright, M.S. Appl. Catal. 1993, 93, 245; 1993, 97,
145.
52. Wainwright, M.S.; Trimm, D.L. Catal. Today 1995, 23, 29.
53. Peppley, B.A.; Amphlett, J.C.; Kearns, L.M.; Mann, R.F. Appl. Catal. 1999, 179,
21, 31.
54. Takezawa, N.; Kobayashi, H.; Hirose, A.; Shimokawabe M.; Takahashi, K. Appl.
Catal. 1982, 4, 127.
55. Iida, K.; Imai, T.; Njima, S.; Shirohhana, A. Japan Patents JP61-234939, 234940,
234941, and 234942, 1986.
56. Mizuno, K.; Yoshikawa, K.; Wakejima, N.; Takeuchi, Y.; Watanabe, A. Chem.
Lett. 1986, 1969.
57. Iwasa, N.; Kudo, S.; Takahashi, H.; Masuda, S.; Takezawa, N. Catal. Lett. 1993,
19, 211.
58. Mizuno, K.; Watanabe, A.; Takeuchi, Y.; Wakijima, N. Japan Patent JP62-250948,
1987.
59. Tada, A.; Yoshino, T.; Itch, H. Chem. Lett. 1987, 419.
60. Kobayashi, H.; Takezawa, N.; Minochi, C. Chem. Lett. 1976, 1347.
61. Matsukata, M.; Vemiya, S.; Kikuchi, E. Chem. Lett. 1979, 5.
62. Bridger, G.W.; Spencer, M.S. In Catalyst Handbook; Twigg, M.V., Ed.; Wolf Sci-
entific Text: London, 1989; 458.
ORDER REPRINTS

80 TRIMM AND ÖNSAN

63. Wang, D.; Ma, L.; Jiang, C.J.; Trimm, D.L.; Wainwright, M.S.; Kim, D.H. Proceed-
ings 11th International Congress on Catalysis, 1996; Vol. 101, p. 2162.
64. Shimomura, M.; Nojima, S.; Shigern, N. Japan Patent JP62-61641, 1987.
65. Inui, T.; Suehiro, M.; Takegami, Y. J. Jpn. Petrol. Inst. 1982, 25, 63.
66. Takahashi, K.; Takezawa, N.; Kobayashi, H. Appl. Catal. 1982, 2, 383.
67. Tonner, S.P.; Trimm, D.L.; Wainwright, M.S.; Cant, N.W. Ind. Eng. Chem. Res.
Dev. 1984, 23, 3.
68. Fisher, I.A.; Bell, A.T. J. Catal. 1999, 184, 357.
69. Skrzypek, J.; Sloczynski, J.; Ledakowicz, S. Methanol Synthesis; Polish Scientific
Publishers: Warsaw, 1994.
Downloaded by [McGill University Library] at 08:56 20 November 2012

70. Twigg, M.V., Ed. Catalyst Handbook; Wolf Scientific Text: London, 1989.
71. Takezawa, N.; Iwasa, N. Catal. Today 1997, 36, 45.
72. Cubeiro, M.L.; Fierro, J.L.G. J. Catal. 1998, 179, 150.
73. Rostrup Nielsen, J.R. Catalytic Steam Reforming; Danish Technical Press, 1984.
74. Rostrup Nielsen, J.R. In Catalysis Science and Technology; Andersen, J.R., Boud-
art, M., Eds.; Springer-Verlag, New York, 1984; Vol. 5, 1.
75. Trimm, D.L. Appl. Catal. 1983, 5, 263.
76. Rostrup Nielsen, J.R.; Trimm, D.L. J. Catal. 1977, 48, 155.
77. Trimm, D.L. Catal. Today 1999, 49, 3.
78. Grenoble, D.C. J. Catal. 1978, 51, 203.
79. Rostrup Nielsen, J.R. J. Catal. 1973, 31, 173.
80. Ross, J.R.H. In Surface and Defect Properties of Solids; Roberts, M.W., Thomas,
J.M., Eds.; Chemical Society, London, 1974; Vol. 4, 34.
81. Trimm, D.L. Catal. Today 1998, 44, 67.
82. Elnashaie, S.S.E.H.; Adris, A.M.; Al-Ubaid, A.S.; Soliman, M.A. Chem. Eng. Sci.
1990, 45, 491.
83. Xu, J.; Froment, G.F. AIChE J. 1989, 35, 88.
84. Moayeri, M.; Trimm, D.L. J. Appl. Chem. Biotech. 1976, 26, 419.
85. Figueiredo, J.L.; Trimm, D.L. Rev. Port. Quim. 1977, 19, 363.
86. Bhatta, K.S.M.; Dixon, G.M. Ind. Eng. Chem. Prod. Res. Dev. 1969, 8, 324.
87. Saito, M.; Tokuno, M.; Ichiro, A.; Morita, Y. Kogyo Kagaku Zasshi 1970, 73,
2405.
88. Tottrup, P.B. Appl. Catal. 1982, 4, 377.
89. Gas Making and Natural Gas, BP; Ben Johnson and Co., York, 1972.
90. Bond, G.C. Catalysis by Metals; Academic Press: New York, 1970.
91. Kikuchi, E.; Tanaka, E.; Yamazaki, Y.; Morita, Y. Bull. Japan Petrol. Inst. 1974,
16, 95.
92. Jenkins, J.W. US Patent 4,789,540, 1988.
93. Jenkins, J.W.; Shutt, E. Plat. Metals Rev. 1989, 33 (3), 118.
94. Garcia-Fierro, J.L. Proceedings 12th International Congress on Catalysis; Corura,
A., Melo, F.V., Mendioroz, S., Fierro, J.L.G., Eds.; Elsevier: Amsterdam, 2000;
177.
95. Ma, L.; Jiang, C.J.; Adesina, A.A.; Trimm, D.L.; Wainwright, M.S. Chem. Eng.
J. 1996, 62, 103.
96. Velu, S.; Suzuki, K.; Osaki, T. Catal. Lett. 1999, 62, 159.
97. Anonymous. Automot. Eng. 1997, 151.
98. Houseman, J.; Cerini, D.J. SAE Report 769001, 1976.
99. Rostrup-Nielsen, J.R.; Bak Hansen, J.H.; J. Catal. 1993, 144, 38.
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 81

100. Tsipouriari, V.A.; Zhang, Z.; Verykos, X.E. J. Catal. 1998, 179, 283, 292.
101. Claridge, J.B.; Green, M.L.H.; Tsang, S.C.; York, A.P.E.; Ashcroft, A.T.; Battle,
P.D. Catal. Lett. 1993, 22, 299.
102. van Looij, F.; Geus, J.W. J. Catal. 1997, 168, 154.
103. Goula, M.A.; Lemonidou, A.A.; Grunert, W.; Baerns, M. Catal. Today 1996, 32,
149.
104. Lu, Y.; Liu, Y.; Shen, S. J. Catal. 1998, 177, 386.
105. Nakamura, J.; Umeda, S.; Kubushiro, K.; Kunimovi, K.; Uchijima, T. Sekiyo Gak-
kashi 1993, 36, 97.
106. Nakagawa, K.; Ikenaga, N.; Teng, T.; Kobayashi, T.; Suzuki, T. J. Catal. 1999,
Downloaded by [McGill University Library] at 08:56 20 November 2012

186, 405.
107. Bitter, J.H.; Seshan, K.; Lercher, J.A. J. Catal. 1998, 176, 93.
108. Mark, M.F.; Maier, W.F. J. Catal. 1996, 164, 122.
109. Ruckenstein, E.; Wang, H.Y. J. Catal. 1999, 187, 151.
110. Otsuka, K.; Wang, Y.; Sunada, E.; Yamanaka, I. J. Catal. 1998, 175, 152.
111. Barbier, J., Jr.; Duprez, D. Appl. Catal. 1992, 85, 89.
112. Nakamura, J.; Aikawa, K.; Sato, K.; Uchijima, T. Catal. Lett. 1994, 25, 265.
113. Nakagawa, J.; Anzai, K.; Matsui, N. Catal. Lett. 1998, 51, 163.
114. Bradford, M.C.J.; Vannice, M.A. J. Catal. 1999, 183, 69.
115. Prettre, M.; Eichner, C.; Perrin, M. Trans. Faraday Soc. 1993, 43, 335.
116. Ashcroft, A.T.; Cheetham, A.K.; Green, M.L.H.; Vernon, P.D.F. Nature 1991, 352,
225.
117. Vernon, P.D.F.; Green, M.L.H.; Cheetham, A.K.; Ashcroft, A.T. Catal. Today
1992, 13, 417.
118. Hickman, D.A.; Haupfear, E.A.; Schmidt, L.D. Catal. Lett. 1993, 17, 223.
119. Bharadwaj, S.S.; Schmidt, L.D. Fuel Process. Technol. 1995, 42, 109.
120. Hickman, D.A.; Schmidt, L.D. J. Catal. 1992, 138, 267.
121. Bodke, A.S.; Bharadwaj, S.S.; Schmidt, L.D. J. Catal. 1998, 179, 138.
122. Dietz, A.G., III; Carlsson, A.F.; Schmidt, L.D. J. Catal. 1996, 176, 459.
123. Zum Mallen, M.P.; Schmidt, L.D. J. Catal. 1996, 161, 230.
124. Choudhary, V.R.; Rajput, A.M.; Prabhakar, B. Catal. Lett. 1992, 15, 363.
125. Choudhary, V.R.; Rajput, A.M.; Rane, V.H. J. Phys. Chem. 1992, 96, 8686.
126. Choudhary, V.R.; Uphade, B.S.; Mamman, A.S. J. Catal. 1997, 172, 281.
127. Choudhary, V.R.; Uphade, B.S.; Mamman, A.S. J. Catal. 1998, 178, 576.
128. Maiya, P.S.; Anderson, T.J.; Mieville, R.L.; Dusek, J.T.; Picciolo, J.J.; Balachan-
dran, U. Appl. Catal. A: Gen. 2000, 196, 65.
129. Ralph, T.R. Plat. Metals Rev. 1997, 41 (3), 102.
130. Schmidt, V.M.; Bröcherhoff, P.; Höhlein, B.; Menzer R.; Stimmung, U.; J. Power
Sources 1994, 49, 299.
131. Gasteiger, H.A.; Markovic, N.; Ross, P.N.; Cairns, E. J. Phys. Chem. 1994, 98,
617.
132. Ertl, G.; Knözinger, H.; Weitkamp, J., Eds. Handbook of Heterogeneous Catalysis;
VCH–Wiley; New York, 1997; Sec. 7, 2094.
133. Igarashi, H.; Fujimo, T.; Watanabe, M. J. Electroanal. Chem. 1995, 391, 119.
134. Oetjen, H.F.; Schmidt, V.M.; Stimmung, U.; Trila, F. J. Electrochem. Soc. 1996,
143, 3838.
135. Pearce, B.B.; Twigg, M.V.; Woodward, C. Catalyst Handbook; Twigg, M.V., Ed.;
Wolfe: London, 1989; 340.
ORDER REPRINTS

82 TRIMM AND ÖNSAN

136. Armor, J.N. Appl. Catal. B: Environ. 1992, 1, 221.


137. Herz, R.K.; Badlani, D.R.; Schryer, D.R.; Upchurch, B.T. J. Catal. 1993, 141, 219.
138. Farrauto, R.J.; Bartholomew, C.H. Fundamentals of Industrial Catalytic Processes;
Blackie Academic and Professional: London, 1999; 654.
139. Huang, T.J.; Yu, T.C.; Chang, S.H. Appl. Catal. 1989, 52, 157.
140. Lin, W.; Fyltzani-Stephanopoulos, M. J. Catal. 1995, 153, 304.
141. Simonot, L.; Garin, F.; Maire, G. Appl. Catal. B: Environ. 1997, 11, 167.
142. Falcon, H.; Martinez-Lope, M.J.; Alonso, J.A.; Fierro, J.L.G. Appl. Catal. B: Envi-
ron. 2000, 26, 131.
143. Bond, G.C.; Fuller, M.J.; Molloy, L.R. Proceedings 6th International Congress on
Downloaded by [McGill University Library] at 08:56 20 November 2012

Catalysis; Bond, G.C., Wells, P.B., Tompkins, F.C., Ed.; Chemical Society: Lon-
don, 1997; Vol. 1, 356.
144. Stark, D.S.; Harris, M.R. J. Phys. E: Sci. Instrum. 1983, 16, 492.
145. Stark, D.S.; Crocker, A.; Steward, G.J. J. Phys. E: Sci. Instrum. 1983, 16, 158.
146. Drawdy, J.E.; Hoflund, G.B.; Gardner, S.D.; Yngvadottir, E.; Schryer, D.R. Surface
Interf. Anal. 1990, 16, 369.
147. Herz, R.K. NASA Conference Publications No. 3076, Schryer, D.R.; Hoflund,
G.B., Eds.; U.S. GPO: Washington, DC, 1990; 21.
148. Bollinger, M.A.; Vannice, M.A. Appl. Catal. B: Environ. 1996, 8, 417.
149. Doi, Y.; Miyake, H.; Soga, K. J. Chem. Soc., Chem. Commun. 1987, 5, 347.
150. Gardner, S.D.; Hoflund, G.B.; Davidson, M.R.; Schryer, D.R. J. Catal. 1989, 115,
132.
151. Oh, S.H.; Carpenter, J.E. J. Catal. 1986, 98, 178.
152. Haruta, M. Catal. Today 1996, 36, 153.
153. Haruta, M.; Kobayashi, T.; Tsubota, S.; Nakahara, Y. Chem. Express. 1988, 3,
159.
154. Tsubota, S.; Haruta, M.; Kobayashi, T.; Ueda, A.; Nakahara, Y. Preparation of
Catalysts VI; Poncelet, G.C., Jacobs, P.A., Grange, P., Delmon, B., Eds.; 1991;
695.
155. Gardner, S.D.; Hoflund, G.B.; Upchurch, B.T.; Schryer, D.R.; Kielin, E.J.; Schryer,
J. J. Catal. 1991, 129, 114.
156. Hoflund, G.B.; Gardner, S.D.; Schryer, D.R.; Upchurch, B.T.; Kielin, E.J. Appl.
Catal. B: Environ. 1995, 6, 117.
157. Dekkers, M.A.P.; Lippits, M.J.; Nieuwenhuys, B.E. Catal. Lett. 1998, 56, 195.
158. Haruta, M.; Tsubota, S.; Kobayashi, T.; Kageyama, H.; Genet, M.J.; Delmon, B.
J. Catal. 1993, 144, 175.
159. Lin, S.D.; Bollinger, M.; Vannice, M.A. Catal. Lett. 1993, 17, 245.
160. Cant, N.W.; Ossipoff, N.J. Catal. Today 1997, 36, 125.
161. Kozlova, A.P.; Kozlov, A.I.; Sugiyama, S.; Matsui, V.; Akasura, K.; Iwasawa, Y.
J. Catal. 1999, 181, 37.
162. Mimico, S.; Scire, S.; Cristafulli, C.; Visco, A.M.; Galvagno, S. Catal. Lett. 1997,
47, 273.
163. Haruta, M.; Yamada, N.; Kobayashi, T.; Iijima, S. J. Catal. 1989, 115, 301.
164. Grunwaldt, J.D.; Maciejewski, M.; Becker, O.S.; Fabrizioli, P.; Baiker, A. J. Catal.
1999, 186, 458.
165. Gardner, S.D.; Hoflund, G.B.; Davidson, M.R.; Laitinen, H.A.; Schryer, D.R.;
Upchurch, B.T. Langmuir 1991, 7, 2140.
166. Haruta, M. Catal. Surveys Japan 1997, 1, 61.
ORDER REPRINTS

FUEL CONVERSION FOR FUEL-CELL-DRIVEN VEHICLES 83

167. Bond, G.C.; Thompson, D.T. Catal. Rev.—Sci. Eng. 1999, 41 (3–4), 319.
168. Lin, S.D.; Vannice, M.A. Catal. Lett. 1991, 10, 47.
169. Liu, Z.M.; Vannice, M.A. Catal. Lett. 1997, 43, 51.
170. Valden, M.; Lai, X.; Goodman, D.W. Science 1998, 281, 1647.
171. Okumura, M.; Tanaka, K.; Ueda, A.; Haruta, M. Solid State Ionics 1997, 95, 143.
172. Bamwenda, G.R.; Tsubota, S.; Nakamura, T.; Haruta, M. Catal. Lett. 1997, 44, 83.
173. Valden, M.; Pak, S.; Lai, X.; Goodman, D.W. Catal. Lett. 1998, 56, 7.
174. Schryer, D.R.; Upchurch, B.T.; Sidney, B.D.; Brown, K.G.; Hoflund, G.B.; Herz,
R.K. J. Catal. 1991, 130, 314.
175. Upchurch, B.T.; Kielin, E.J.; Miller, I.M. NASA Conference Publication No. 3076;
Downloaded by [McGill University Library] at 08:56 20 November 2012

Schryer, D.R., Hoflund, G.B., Eds.; U.S. GPO: Washington, DC, 1990; 69.
176. Grass, K.; Lintz, H.G. J. Catal. 1997, 172, 446.
177. Hoflund, G.B.; Upchurch, B.T.; Kielin, E.J.; Schryer, D.R. Catal. Lett. 1995, 31,
133.
178. Schryer, D.R.; Upchurch, B.T.; Van Norman, J.D.; Brown, K.G.; Schryer, J.
J. Catal. 1990, 122, 193.
179. Boulahouache, A.; Kons, G.; Lintz, H.G.; Schulz, P. Appl. Catal. A: Gen. 1992,
91, 115.
180. Bond, G.C.; Molloy, L.R.; Fuller, M.J. J. Chem. Soc., Chem. Commun. 1975, 796.
181. Sermon, P.A.; Self, V.A.; Barrett, E.P.S. J. Mol. Catal. 1991, 65, 377.
182. Mergler, Y.J.; van Aalst, A.; van Delft, J.; Nieuwenhuys, B.E. Appl. Catal. B:
Environ. 1996, 10, 245; J. Catal. 1996, 161, 310.
183. Törncrona, A.; Skoglundh, M.; Thormahlen, P.; Fridell, E.; Jobson, E. Appl. Catal.
B: Environ. 1997, 14, 131.
184. Thormahlen, P.; Skoglundh, M.; Fridell, E.; Anderson, B. J. Catal. 1999, 188, 300.
185. Sheintuch, M.; Schmidt, J.; Lechtman, Y.; Yahav, G. Appl. Catal. 1989, 49, 55.
186. Gardner, S.D.; Hoflund, G.B.; Schryer, D.R.; Schryer, J.; Upchurch, B.T.; Brown,
D.R. NASA Conference Publication No. 3076; Schryer, D.R., Hoflund, Eds.; U.S.
GPO: Washington, DC, 1990; 123.
187. Kulshreshtha, S.K.; Gadgil, M.M. Appl. Catal. B: Environ. 1997, 11, 291.
188. Pavlova, S.N.; Sadykov, V.A.; Bulgakov, N.N.; Bredikhin, M.N. J. Catal. 1996,
161, 517.
189. Praserthdam, P.; Majitnapakul, T. Appl. Catal. A: Gen. 1994, 108, 21.
190. Szanyi, J.; Goodman, D.W. Catal. Lett. 1993, 21, 165.
191. Choi, K.J.; Vannice, M.A. J. Catal. 1991, 131, 22.
192. El-Shobaky, H.G.; Mokhtar, M.; El-Shobaky, G.A. Appl. Catal. A: Gen. 1999, 180,
335.
193. Zhou, R.X.; Jiang, X.Y.; Mao, J.X.; Zheng, X.M. Appl. Catal. A: Gen. 1997, 162,
121.
194. Luo, M.F.; Zhong, Y.J.; Yuan, X.X.; Zheng, X.M. Appl. Catal. A: Gen. 1997, 162,
121.
195. Kageyama, H.; Kajimo, N.; Kobayashi, T.; Haruta, M. Physica B 1989, 158, 183.
196. Torres Sanchez, R.M.; Ueda, A.; Tanaka, K.; Haruta, M. J. Catal. 1997, 168, 125.
197. Haruta, M.; Kobayashi, T.; Sano, H.; Yamada, N. Chem. Lett. 1987, 405.
198. Watanabe, M.; Uchida, H.; Igarashi, H.; Suzuki, M. Chem. Lett. 1995, 21.
199. Igarashi, H.; Uchida, H.; Suzuki, M.; Sasaki, Y.; Watanabe, M. Appl. Catal. A:
Gen. 1997, 159, 159.
200. Bethke, G.K.; Kung, H.H. Appl. Catal. A: Gen. 2000, 194–195, 43.
ORDER REPRINTS

84 TRIMM AND ÖNSAN

201. Kahlich, M.J.; Gasteiger, H.A.; Behm, R.J. J. Catal. 1999, 182, 340.
202. Tsubota, S.; Yamada, N.; Haruta, M.; Kobayashi, T.; Nakahara, Y. Chem. Express.
1990, 5 (6), 349.
203. Kahlich, M.J.; Gasteiger, H.A.; Behm, R.J. J. Catal. 1997, 171, 93.
204. Oh, S.H.; Sinkevitch, R.M. J. Catal. 1993, 142, 254.
205. Brown, M.L.; Green, A.W.; Cohn, G.; Andersen, H.C. Ind. Eng. Chem. 1960, 52,
841.
206. Bonacci, J.C.; Otchy, T.G.; Ackerman, T. US Patent 4,238,468, 1980.
207. Schubert, M.M.; Gasteiger, H.A.; Behm, R.J. J. Catal. 1997, 172, 256.
208. Stetter, J.R.; Blurton, K.F. Ind. Eng. Chem. Prod. Res. Dev. 1980, 19, 214.
Downloaded by [McGill University Library] at 08:56 20 November 2012

209. Oh, S.H.; Fisher, G.B.; Carpenter, J.E.; Goodman, D.W. J. Catal. 1986, 100, 360.
210. Berlowitz, P.J.; Peden, C.H.F.; Goodman, D.W. J. Phys. Chem. 1988, 92, 5213.
211. Utaka, T.; Sekizawa, K.; Eguchi, K. Appl. Catal. A: Gen. 2000, 194–195, 21.
212. Cheng, W.H. React. Kinet. Catal. Lett. 1996, 58 (2), 329.
213. Sekizawa, K.; Yano, S.; Eguchi, K.; Arai, H. Appl. Catal. A: Gen. 1998, 169, 291.
214. Muraki, H.; Matunaga, S.I.; Shinjoh, H.; Wainwright, M.S.; Trimm, D.L. J. Chem.
Tech. Biotechnol. 1991, 52, 415.
215. Avci, A.K.; Önsan, Z.I.; Trimm, D.L. Unpublished data.
216. Avci, A.K.; Trimm, D.L.; Onsan, Z.I. In Proceedings 12th International Congress
on Catalysis; Corina, A., Melo, F.V., Mendioroz, S., Fierro, J.L.G., Eds.; Elsevier:
Amsterdam, 2000; 2753.
Request Permission or Order Reprints Instantly!

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

All information and materials found in this article, including but not limited
to text, trademarks, patents, logos, graphics and images (the "Materials"), are
the copyrighted works and other forms of intellectual property of Marcel
Dekker, Inc., or its licensors. All rights not expressly granted are reserved.
Downloaded by [McGill University Library] at 08:56 20 November 2012

Get permission to lawfully reproduce and distribute the Materials or order


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

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

Order now!

Reprints of this article can also be ordered at


http://www.dekker.com/servlet/product/DOI/101081CR100104386

You might also like