Vous êtes sur la page 1sur 15

Applied Energy 86 (2009) 22832297

Contents lists available at ScienceDirect

Applied Energy
journal homepage: www.elsevier.com/locate/apenergy

Catalysis for NOx abatement


Sounak Roy a, M.S. Hegde a, Giridhar Madras a,b,*
a
b

Solid State and Structural Chemistry Unit, Indian Institute of Science, Bangalore 560 012, India
Department of Chemical Engineering, Indian Institute of Science, Bangalore 560 012, India

a r t i c l e

i n f o

Article history:
Received 4 December 2008
Received in revised form 17 March 2009
Accepted 18 March 2009
Available online 25 April 2009
Keywords:
NOx reduction
Metal ion substitution
Environmental catalysis

a b s t r a c t
Research in the eld of NOx abatement has grown signicantly in the past two decades. The general trend
has been to develop new catalysts with complex materials in order to meet the stringent environmental
regulations. This review discusses briey about the different sources of NOx and its adverse effect on the
ecosystem. The main portion of the review discusses the progress and development of various catalysts
for NOx removal from exhaust by NO decomposition, NO reduction by CO or H2 or NH3 or hydrocarbons.
The importance of understanding the mechanism of NO decomposition and reduction in presence of
metal ion substituted catalysts is emphasized. Some conclusions are made on the various catalytic
approaches to NOx abatement.
2009 Elsevier Ltd. All rights reserved.

Contents
1.

2.

3.

Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.1.
NOx sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.2.
Noxious effect of NOx . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.3.
Legislations . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Catalytic deNOx . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.
NOx decomposition . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.1.1.
Photocatalytic decomposition and reduction of NOx . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.2.
NOx reduction by CO . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.3.
deNOx by H2 and NH3 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.4.
Selective catalytic reduction (SCR) of NO by H2 or NH3 or HC . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.
Other methods. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.1.
NSR (NOx storage and reduction) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.2.
Selective NOx recirculation (SNR) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.3.
Selective non-catalytic reduction (SNCR) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
2.5.4.
Ozone injection . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
Conclusions and future perspectives. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

1. Introduction
Environmental catalysis can be dened as technologies using
catalysts to reduce the emission of environmentally unacceptable

* Corresponding author. Address: Department of Chemical Engineering, Indian


Institute of Science, Bangalore 560 012, India. Tel.: +91 80 2293 2321; fax: +91 80
2360 1310.
E-mail address: giridhar@chemeng.iisc.ernet.in (G. Madras).
0306-2619/$ - see front matter 2009 Elsevier Ltd. All rights reserved.
doi:10.1016/j.apenergy.2009.03.022

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.
.

2283
2284
2284
2284
2285
2285
2286
2287
2289
2289
2292
2292
2293
2293
2293
2293
2294

compounds [1,2]. The exhausts from automobiles and stationary


sources such as power plants contain CO, NOx and hydrocarbons.
The conversion of these pollutants to CO2, N2 and H2O using catalysts is a challenge. In the last two decades, signicant developments have occurred in this eld leading to a better
understanding of the catalytic NOx abatement. A number of reviews regarding various aspects of NOx abatement have been published, as discussed in the following sections. This review, however,
is a compendium of all aspects of NOx abatement that include NO

2284

S. Roy et al. / Applied Energy 86 (2009) 22832297

decomposition, NO reduction by CO, H2, NO storage and selective


catalytic reduction of NO.
1.1. NOx sources
The major source of nitrogen oxides is the combustion of fossil
fuels such as petroleum in the engines of vehicles or coke in the
electrical power plants. The origin of NOx is generally categorized
into mobile and stationary sources. Fig. 1 describes the different
sources of NOx in US and in European countries [3,4].
NOx is a generic term for mono-nitrogen oxides namely NO and
NO2, which are produced during combustion at high temperatures.
At ambient temperatures, oxygen and nitrogen do not react with
each other. However, in an internal combustion engine, high temperatures lead to reactions between nitrogen and oxygen to yield
nitrogen oxides. In the presence of excess oxygen, nitric oxide will
be converted to nitrogen dioxide.
Bosch and Janssen [5] categorize three types of NOx formed during the combustion process. NOx from engine exhaust typically
consists of a mixture of 95% NO and 5% NO2. The rst category,
thermal NOx, is formed by the oxidation of N2 at high temperatures.


N2 O2 () 2NO; DH298 180:6 kJ=mol

This reaction takes place above 1300 K and follows the Zeldovich mechanism of chain reactions involving N* and O* activated
atoms:

N2 O ! NO N

N O2 ! NO O

The rate of NO formation is essentially controlled by reaction


(2) and increases exponentially with temperature. The Zeldovich
mechanism dominates NO formation under most engine conditions [6]. The NOx emission from the engine can be controlled by
lowering the combustion temperature by operating the engine under excess air (fuel-lean) conditions but most these approaches are
not very effective [6], though recent approaches based on high
temperature air combustion (HiTAC) are effective.
The second category of NOx is called fuel NOx and is formed from
the oxidation of nitrogen present in fuels such as coal and heavy
oils. In contrast to thermal NOx, fuel, NOx formation is relatively
independent of temperature at normal combustion temperatures
[6].

% of NOx from different sources

60

USA
Europe

50
40

The third category of NOx is called prompt NO0x (also termed as


Fenimore NO) which is formed by the reaction of hydrocarbon
fragments with atmospheric nitrogen to yield products such as
HCN and H2CN. These can be subsequently oxidized to NO in the
lean zone of the ame. NO can further react with oxygen to NO2
or N2O.


NO 1=2O2 () NO2 ;

DH298 113 kJ=mol

2NO () N2 O 1=2O2 ;

DH298 99 kJ=mol

4
5

Prompt NOx formation is proportional to the number of carbon


atoms present per unit volume and is independent of the identity
of the parent hydrocarbon. The quantity of HCN formed increases
with the concentration of hydrocarbon radicals. Prompt NOx can
be formed in a signicant quantity at low-temperature, fuel-rich
conditions and where residence times are short.
Another route of formation of NO is via nitrous oxide. In this
mechanism, O-atom attacks molecular nitrogen in presence of a
third molecule that results in the formation of N2O. This subsequently reacts with O atom to form NO, N2O + O ? 2 NO with
an activation energy of 97 kJ/mol. This reaction route is overlooked because the total NO formed by this reaction is not significant. However, lean conditions suppress Fenimore NO and low
temperatures suppress Zeldovich NO. As high pressures promote
this reaction, the formation of NO by this route occurs primarily
in lean premixed combustion in high pressure gas turbine
engines.
1.2. Noxious effect of NOx
The oxides of nitrogen play a major role in the photochemistry
of the troposphere and stratosphere. NOx catalyzed ozone destruction occurs via the following reactions:

NO O3 ! NO2 O2

NO2 O ! NO O2

These reactions are largely responsible for the ozone decline in


middle to high latitudes from spring to fall [7,8]. Another adverse
effect of NOx is acid rain, which can perturb the ecosystems and
can cause biological death of lakes and rivers. Peroxyacetylene nitrates (PAN) can also be formed from nitric oxide and contribute
signicantly to global photo-oxidation pollution [9].
Some biological studies have shown NO as an essential messenger, which transmits the necessary information to the white blood
cells within the bloodstream to destroy tumor cells and to the
neurotransmitters to dilate the blood vessels [9,10]. However, the
biologically active NO is a poisonous product of the in vivo
enzyme-catalyzed transformation of the amino acid, arginine. NO
diffuses through the alveolar-cells and capillary vessels of the
lungs and damages the alveolar structures and their functions
throughout the lungs provoking both lung infections and respiratory allergies like bronchitis, pneumonia, etc. [11,12].

30

1.3. Legislations
20
10
0
Power

Industry Res. Fuel Comb. Agriculture Transportation

Fig. 1. Illustration of emission of NOx by source category in USA and European


countries.

Because of the ecological and health hazards due to the presence of NOx in the environment, regulations have been proposed
to control NOx emissions. There is a wide variation among countries with respect to both the type and the level of regulation employed. The Gothenburg protocol establishes reductions of four
main pollutants to reduce acidication, eutrophication and the effect of ozone. Other than Canada and USA, 29 European countries
have signed this protocol and these countries have estimated the

2285

S. Roy et al. / Applied Energy 86 (2009) 22832297


Table 1
Emission standards for India [14].
Year

NO ! 1=2N2 1=2O2 ;

Emission standards for 3-wheel


gasoline vehicles (g/km)

Emission standards for 2-wheel


gasoline vehicles (g/km)

CO

HC

HC + NOx

CO

HC

HC + NOx

1230
6.75
1.00

812

5.40
2.00

1230
4.50
2.00

812

3.60
2.00

Engine Power (P)

Date

1991
1996
2000

P 6 19 kW
19 kW < P 6 50 kW
50 kW < P 6 176 kW
176 kW < P 6 800 kW
Date

2004.01
2005.07
2004.01
2004.07
2004.01
2004.11

Emission standards for diesel engines 6800 kW


for generator sets

NO reduction by CO is an important reaction in three-way catalysis (TWC).

NO CO ! 1=2N2 CO2 ;

DH298 328 kJ=mol

HC
(g/kWh)

NOx
(g/kWh)

PM
(g/kWh)

Smoke
(1/m)

iii. NO reduction by H2 and NH3

5.0
3.5
5.0
3.5
3.5
3.5

1.3
1.3
1.3
1.3
1.3
1.3

9.2
9.2
9.2
9.2
9.2
9.2

0.6
0.3
0.5
0.3
0.3
0.3

0.7
0.7
0.7
0.7
0.7
0.7

NO can be reduced in presence of hydrogen or ammonia.

CO
(mg/Nm3)

NMHC
(mg/Nm3)

NOx
ppm(v)

PM
(mg/Nm3)

150
150
150

150
100
100

1100
970
710

75
75
75

ii. NO reduction by CO

CO
(g/kWh)

Emission limits for diesel engines >800 kW for generator


sets

Until 2003.06
2003.072005.06
2005.07

DH298 86:6 kJ=mol

NO H2 ! 1=2N2 H2 O; DH298 287 kJ=mol

10

6NO 4NH3 ! 5N2 6H2 O

11

iv. SCR (selective catalytic reduction) of NO


While NO can be reduced with ammonia based on Eq. (11), SCR
processes are usually carried out in presence of oxygen. Thus the
equations for the reduction of ammonia and hydrocarbons in presence of oxygen are given below,


4NO 4NH3 O2 ! 4N2 6H2 O; DH298 1627 kJ=mol

12

NO CH4 3=2O2 ! 1=2N2 CO2 2H2 O;


critical loads themselves. An overview of the different targets [13]
is available.
The ongoing emission standards in Europe in Euro IV while Euro
V will be effective from September 2009. The latter regulates that
the emission will be less than 0.18 g/km for diesel and 0.06 for petrol driven engines, respectively. Since the year 2000, India is adopting Euro I for four-wheeled light-duty and for heavy-duty vehicles.
For 2- and 3-wheelers, Bharat Stage II (Euro II) was applicable from
April, 2005 and Stage III (Euro III) standards have come in force
from April, 2008 [14]. Table 1 shows the emission standards for
different Indian vehicles and from new diesel engines used in generator sets.

DH298 847 kJ=mol

13

The above four sections can be reclassied as two major paths,


one in which the catalytic decomposition of NOx occurs in absence
of a reducing agent (i) and the second in which the catalytic reduction of NOx occurs in the presence of a reducing agent (iiiv). The
second pathway can be sub-classied to cases where the compositions of the exhaust gas are near stoichiometric conditions (ii, iii) or
under fuel-lean conditions (iv). We discuss the above four methods
of deNOx in the following sections (Sections 2.12.4) and also discuss other methods of deNOx in a separate section (Section 2.5).
2.1. NOx decomposition

2. Catalytic deNOx
NO molecule has the electron conguration r2g r2u rg ; pu p1g .
Due to the unpaired p antibonding electron the molecule is paramagnetic and partly cancels the effect of p bonding electrons. The
bond order is 2.5, consistent with an inter-atomic distance of
1.15 that is intermediate between triple bond distance in NO+
of 1.06 and double bond 1.20 . NO has an unpaired electron
in its 2p* orbital and this has led to the notion that amphoteric
bonding for NO on a surface is a useful consideration. NO can either
donate its electron to the surface (like CO) or it can accept electron
density from the surface into the half-lled 2p* orbital, and therefore show a very wide variety of chemistry on surface.
NOx is thermodynamically unstable. However, it does not
decompose because its high activation energy (364 kJ/mol). Therefore, a catalyst is needed to lower the activation energy in order to
facilitate the decomposition. Research in this domain is extensive,
and can be divided into four major paths of deNOx activities based
on automobile and stationary sources.
i. NO decomposition
The direct decomposition of nitric oxide to nitrogen and oxygen
is one of the most attractive methods because this reaction is thermodynamically favorable and does not need any reductants.

NO is a molecule that can adsorb either dissociatively or molecularly depending on the metal. NO normally dissociates on the
base metals and show molecular adsorption at room temperature
on noble metals [15]. However, NO dissociation often depends on
surface temperature, surface coverage, crystal plane and surface
defects. The structure of adsorbed NO molecule on metal surfaces
has been discussed employing experimental techniques like electron energy loss spectroscopy (EELS), low energy electron diffraction (LEED), photo electron diffracion (PED), and theoretical
simulation using density-functional theory (DFT) calculations [16].
Brown and King compiled NO vibrational frequencies and assigned sites for NO adsorption on various single crystal metals
[17]. NO can be adsorbed on metal surfaces in different geometries
like linear (atop), bent, bridge etc, as shown in Fig. 2. A general
observation from those vibrational frequencies is that the NO
stretching frequency in the atop NO is more than that of the
bent NO. Thus, the NO dissociation will be facile if the NO is adsorbed on the surface in bent geometry than atop geometry. NO
has an unpaired electron in its 2p* orbital. In a metalNO bonding,
NO makes a 5rd bond with metal atoms as in MCO, and backbonding takes place from metal d orbital to 2p* orbital of NO. Thus,
the metalN bond will be stronger and NO bond will be weaker. If
NO obtains an electron in its antibonding orbital it becomes NO,
which is isoelectronic to O2. O2 adsorption is always on a sideon geometry, which dissociates to form M@O. Similarly, when

2286

S. Roy et al. / Applied Energy 86 (2009) 22832297

the bent or linearly (atop) adsorbed NO molecule becomes


NO, it changes its geometry to a side-on NO intermediate and
dissociates. Rh{1 0 0} or Ni {1 0 0} shows this kind of behavior,
which is absent for Pt surface [18,19]. Fig. 3 describes the geometry. Lambert et al. showed that Rh surfaces are more efcient than
Pt surfaces to dissociate NO [20].
If the metal d orbital is promoted by an extra electron from alkali metal or alkaline-earth metal, then the charge transfer from
metal d to 2p* orbital of NO should increase and dissociate NO.
Lambert and co-workers have shown that sodium or potassium
promoted Rh surface dissociates NO more than the clean surface
[21]. However, the problem in using a single crystal metal to dissociate NO is that the dissociated oxygen either oxidizes the metal
surface to metal oxides or makes an oxygen adlayer over the metal
surface that hinders the dissociation process.
In one study [22], the catalytic activities of 40 metallic (Mg, Zn,
Ni, etc.) oxides were examined in the temperature range of 500
800 C for the decomposition of NO to nitrogen and oxygen. Under
conditions where the O2 formed was continuously removed as
NO2, the reaction was rst order with respect to NO pressure. In
a closed reaction system, O2 retards the decomposition of NO by
competing for the same surface centers. The oxygen self-poisoning effect with respect to NO decomposition is slower on ceriasupported catalysts, which may be attributed to oxygen spillover
from the noble metal to the reduced ceria [23]. Cobalt oxide
(Co3O4) is one of the most active single-component metal oxides
for NO decomposition [24]. The high activity of Co3O4 seems to
be due to the relatively weak CoO bond, which leads to an easy
desorption of lattice oxygen in the lower temperature range.
Co3O4 has a spinel structure with Co(II) and Co(III). During NO
decomposition, Co(II) may oxidize to Co(III) and form Co2O3 like
phase on the surface. But Co2O3 is not stable and decomposes to
Co3O4 thus facilitating NO dissociation. With the addition of Na
[25] or Ag [26], the activity enhances primarily due to excess electron density on the surface. Fang and White [27] studied the
chemisorption of NO on platinised TiO2 by Fourier transform infrared spectroscopy. Two adsorption bands at about 1760 and
1700 cm1, which were assigned as NO adsorbed on Pt open sites
and on closed packed sites, respectively, were found to vary in
intensity with reduction and O2 pretreatment. NO uptake also
diminished on reduced Pt/TiO2 [27]. A similar observation has been
made with Rh/TiO2 [28,29]. The room-temperature chemisorption
of NO was observed to decrease with increasing catalyst reduction

:O:

O:

:N

M:

Linear(~sp)

:O:

:O:

N
M

Bent
(~sp2)

Bridging
(~sp2)

Fig. 2. Schematic representation of the bonding in NO complexes [17].

O:
N

:O:
Or :N

M
M

M:

O
M

M
Fig. 3. Schematic representation of NO dissociation [20].

O
M

temperature. Thus, reduced Pt and Rh on TiO2 are less active for NO


dissociation.
The adhesive and catalytic properties of rhodium, palladium,
and platinum on a-Al2O3 have been examined and it has been observed that the variation of the Fermi level as a function of metalceramic interface indicate that the ceramic is not inert. The oxygenmetal interface is unstable for rhodium and was predicted
to be particularly favorable both for dissociative adsorption of
NO as well as for a coupling reaction [30].
The decomposition of NO over Cu-exchanged zeolites has been
studied extensively since the pioneering work of Iwamoto et al.
[3134]. NOx decomposition over CuMFI zeolites (ZSM-5 has been
named as MFI with silica/alumina = 23.3) proceeds via a redox type
mechanism. NO could be adsorbed as NO+, NO and NO
2 species on
the Cuzeolite. Cu+ ions generated through pretreatment, at elevated temperature, are oxidized to Cu2+ ions by oxygen and a part
of the resulting Cu2+ ions acts as adsorption sites for NO+. The NO
molecule has an unpaired electron in a degenerate antibonding
orbital and the electron transfer of this unpaired electron from
the antibonding orbital of NO to empty or partially lled 3d orbitals of transition metal ions can then occur easily. This is followed
by lone pair donation from NO and p back-bonding to the NO orbitals, generating a nitrosyl complex. Although the molecular orbital
structure of the nitrosyl is rather complex, the process can be conveniently schematized as: Cu2 NO () Cu NO . A fraction of
the Cu2+ ion becomes reduced to Cu+ through spontaneous desorption of oxygen. Reoxidation with NO regenerates the oxidized Cu2+
sites and N2 is formed. From the IR studies, Giamello et al. supported the above data and showed the formation of the nitrosylic
adduct [35]. Cheung et al. concluded that the catalytic performance
of Cu in a zeolite matrix for the decomposition of NO was linked to
the Cu () Cu2 redox cycle, which is essential for the formation
of the ONCu2 NO
2 complex, and thus dependent on the relative
stability of the Cu+ and Cu2+ species [36]. However, the low hydrothermal stability of these materials for gasoline engines is a serious
drawback.
Many of the theoretical calculations provide an insight on the
mechanism of NOx decomposition on zeolites. Schneider and coworkers have proposed a mechanism pathway involving two successive O-atom transfers to an isolated, zeolite-bound Cu+ center,
initiated by formation of a short-lived and difcult to detect isonitrosyl intermediate, and yielding sequentially N2O and Cu-bound O followed by N2 and Cu-bound O2 [37]. A recent density functional
theory (DFT) calculated electronic structure and excitation energy
spectra for the model system (HO)3AlOCuOCu, show that an anionic NO dimer (ONNO) is formed on (OCuOCu) chain [38].
However, NO decomposition over metal surfaces, oxides or zeolites has some serious practical difculties in practice. NO decomposition process is a high temperature phenomenon, which is
undesirable. NO dissociation over single crystal metal surface oxidizes the metal surface, which in turn hinders the NO dissociation
process. So, a reductant is required to scavenge the dissociated
oxygen. From an exhaust catalysis view, it is not sufcient to dissociate NO, other pollutants like CO and hydrocarbons (HC) should
also be oxidized. Therefore, NO reduction by CO, HC, H2, and NH3 is
more attractive for NOx abatement.
2.1.1. Photocatalytic decomposition and reduction of NOx
An alternative method of dissociating and reducing NOx is by
the use of photocatalysis at room temperature. Direct photocatalytic decomposition of NO would yield N2 and O2. Anpo and his
co-workers have studied photocatalytic decomposition of NO on
TiO2 and found anatase TiO2 exhibit a high efciency for the
decomposition of NO in a ow system [39]. Lim et al. have investigated the photocatalytic decomposition of NO on Degussa P-25
TiO2 in an annular ow type reactor and showed the conversion

S. Roy et al. / Applied Energy 86 (2009) 22832297

of NO to NO2, N2O and N2 by photocatalysis increases with light


intensity, residence time and decreasing initial NO concentration
[40]. Bowering et al. have showed that the photocatalytic activity
of Degussa P-25 TiO2 decreases with the increasing pretreatment
temperature [41]. We have shown the reduction of NO over the
catalyst Ti1xPdxO2d was two orders of magnitude higher than
unsubstituted TiO2 [42]. Direct NO decomposition into N2 and
N2O was also studied and it was observed that NO dissociates in
presence of UV light at room temperature. The products found
were N2, N2O and O2. The O2 evolved reacts with NO to give NO2.
NO2 formed is adsorbed by catalyst. Prolonged NO2 adsorption
thus makes the surface inactive for NO dissociation [42]. However,
NO dissociation starts again when a reductant like CO is passed to
scavenge the evolved dissociated O2.
2.2. NOx reduction by CO
NO reduction by CO is a primary reaction in three-way catalysis
(TWC). The introduction of catalytic treatment of automotive exhausts began with the removal of the incomplete combustion
products, CO and residual HC. These reactions are

2CO O2 ! 2CO2

14

2Cx Hy 2x y=2O2 ! 2xCO2 yH2 O

15

Simple oxidation catalysts, referred to as two-way Catalysis,


accomplished this. The regulations necessitating the catalytic removal of nitrogen oxides, formed in the combustion chamber, began an essential part of exhaust treatment and thus simultaneous
oxidation as well as reduction was required. A combustion exhaust
of a slightly rich mixture of fuel and air was fed to the upstream
catalyst bed, and reduction of NOx was fast and nearly complete.
Secondary air was then injected ahead of a downstream oxidation
catalyst to remove the CO and HCs. However, due to the presence
of hydrogen in the exhaust, the reduction of NOx in the upstream
catalyst resulted mainly in ammonia. When re-oxidized on the
downstream catalyst, the ammonia converted back to NOx, violating the whole approach [43,44]. It was subsequently proposed that
if one could equilibrate the combustion exhaust of an exactly stoichiometric combustion mixture, it was thermodynamically possible to remove all the three pollutants, namely NOx, CO and HC
leaving only water, CO2 and N2 [45]. This is known three-way
catalysis and three-way catalytic converters have been at the heart
of vehicle emission control systems since the 1980s. The airfuel
ratio is the mass ratio of air to fuel present during combustion
and when all the fuel is combined with all the free oxygen, this
is known as the stoichiometric ratio and designated as k = 1. For
gasoline, the stoichiometric air/fuel mixture is approximately
14.7 times the mass of air to fuel. If A/F ratio is below 14.7
(k < 1), then it is called fuel-rich condition and if A/F ratio exceeds
this value (k > 1), then it is called fuel-lean condition. This requires
tight stoichiometry constraints and in a randomly oscillating dynamic combustion, this is not easily attained.
Before 1975, research on automotive catalysts was devoted to
non-noble metals. However, a catalyst in the rigorous automotive
exhaust condition should have (a) intrinsic reactivity, (b) poison
resistance ability and (c) durability and it soon became apparent
that the base metal oxides of Ni, Co, Mn, Cr do not have these properties [46,47]. The noble metals in typical three-way catalysts
(Pt, Pd, Rh) are dispersed as individual atoms or small clusters of
atoms over an oxide support. In the early TWCs, Rh became the catalytic element of choice for NOx control [48]. In the case of Rh/
Al2O3 catalysts, presence of the metal-support interaction has been
shown to be present between zero valent Rh atoms at the interface
and the oxygen ions of the support [49]. The oxidation and reduction of Rh affects dynamic behavior because oxidation of Rh se-

2287

verely decreases its NO reduction activity [50]. Hecker and Bell


subsequently studied NO + CO reaction over Rh/SiO2 and showed
that conversion of NO in presence of oxidized Rh is higher compared to that over reduced Rh [51]. The product selectivity is one
of the very important factors in NO + CO reaction. With the formation of N2, the unselective product N2O is also formed. Hecker and
Bell [51] also showed that the pretreatment of the catalyst (i.e. oxidation) increases N2 selectivity in NO + CO reaction. Further insight
on the mechanism of NO + CO reaction was given by Oh et al. [52],
who studied this reaction over alumina-supported Rh catalyst and
over Rh(1 1 1) single crystal. It was shown that the NO dissociation
is the key step in the reaction. In late 1980s and in early 1990s,
more emphasis was placed on the reaction mechanism, order of
reaction and determining the intermediate products. Dictor observed isocyanate species when NO + CO was reacted over Rh/
Al2O3 [53]. Cho derived a detailed mechanism of NO reduction by
CO and showed that the reaction N2O + CO is an important intermediate reaction in overall NO + CO reaction and the intermediate
N2O + CO reaction in NO + CO reaction is faster by an order of magnitude than the isolated N2O + CO reaction over Rh/Al2O3 [54,55].
This theoretical comparative mechanistic investigation provided by Cho [54,55] was the starting point with successive reviews made by Zhdanov [56] and Cho [57]. In a subsequent work
[58], it was studied extensively and theoretical calculations performed over monometallic and bimetallic Pt and Rh clusters [59],
showed that N2O adsorbs via the terminal N atom, whereas the
conguration with N2O bound to the metal via the oxygen is thermodynamically unstable over noble metals. Hence, both theoretical and kinetic results suggest molecular N2O adsorption and
subsequent dissociation of N2O involving active sites.
The major problem with Rh/Al2O3 catalyst is ageing and loss of
Rh reducibility at 600 C and due to the formation of a solid solution of the type Al2xRhxO3 [60]. Since oxygen is balanced to both
Al3+ and Rh3+ ion in solid solution, Rh3+ ion is difcult to reduce
and loses its catalytic activity.
As discussed earlier, different congurations of adsorbed NO
molecules exhibit a different reactivity towards dissociation [17].
Parallel comparisons can be made on polycrystalline catalysts that
contain Rh on the nature of active NO species towards the dissociation with nitrosyl and/or dinitrosyl species that are involved in
the formation of the reaction products. In this respect, in situ
infrared spectroscopy was used to examine the intermediates
and determine the inuence of the oxidation state of rhodium in
three-way catalysts [61]. Different spectroscopic features have
been observed on Rh and Pt after NO exposure related to the build
up of positively, neutral and negatively charged NO species on Rh,
with neutral NO species predominating on Pt. The formation of
RhNO+ and gem-dicarbonyl species RhI(CO)2 during the adsorption of NO and the CO + NO reaction has been associated mainly
with surface Rh oxidation at low-temperature [62]. In this context,
an excellent overview on the mechanistic aspects of the NO reduction with CO in presence of Rh based catalysts has been provided
[63].
Several studies [5156] on deNOx processes in TWC make use of
nano metal clusters supported on Al2O3 or SiO2. Associating noble
metals with a reducible support has been investigated in detail. For
example, the inuence of Ce additive on the catalytic performances
of three-way bimetallic PtRh/Al2O3 in the CO + NO reaction
showed a benecial effect of ceria on the conversion of NO mainly
at low temperatures and had no effect at temperatures above
280 C [64]. A bi-functional mechanism involving reaction of oxygen from ceria with adsorbed CO molecules on the metals followed
by dissociation of adsorbed NO molecules on anionic vacancies
successfully predicted the experimental data. At high temperatures, ceria is extensively reduced into Ce3+, leading to the formation of anionic vacancies in the vicinity of metal particles. The

S. Roy et al. / Applied Energy 86 (2009) 22832297

that NO reduction by CO over Pd catalyst is not only a structure


sensitive reaction, but also it depends on the surface coverage of
the gaseous molecules, surface temperature and on crystal orientation [75] and concluded that the formation of the stable inactive
atomic nitrogen species plays an important role. NO reduction by
CO over Pd clusters supported by MoO3 (Pd8Mo) showed NO
adsorption on different oxidation state of Pd and Mo forms stable
isocyanate species [77]. Recently, NO + CO reaction over Pd(1 1 1)
was investigated by molecular beam method and molecular as well
as dissociative chemisorption of NO was observed over Pd(1 1 1)
[78]. In many of the above studies, the metal is in the zero valent
state or move into higher valent state during reaction.
Signicant observations have been provided by in situ and operando spectroscopic studies that justify this conclusion. Particularly,
the dependency of the activity and selectivity of palladium particles in the production of N2O during the overall NO/CO reaction
has been discussed extensively [79,80]. The nitrogen species
formed during the reaction was analyzed by determining the surface changes on the catalyst and this was dependent on the oxidation state of Pd.
Ionically substituted Pd in CeO2 or TiO2 showed an order of
magnitude higher rate of NO reduction by CO and also higher
selectivity towards N2 [81,82]. The catalysts, Ce1xPd1xO2d or
Ce1xTixPdyO2d, were synthesized by solution combustion method. Pd was in 2+ and more ionic than PdO. Due to lower valent ionic substitution in CeO2, oxide ion vacancies are created. These
oxides showed high CO oxidation activity primarily due to CO
adsorption on Pd2+ site and O2 dissociation in oxide ion vacancy
site. With a similar analogy, the mechanism of NO reduction by
CO has been found to be very much site specic. A bi-functional
mechanism based on Pd2+ being the adsorbent of NO and CO and
the oxide ion vacancies for NO dissociative chemisorption was proposed and this model ts the experimental data well. The ionic catalysts not only have the higher TOF at lower temperature but also
higher selectivity of N2 compared to the other catalysts reported in
the literature [8385].
The oxide support in the catalysts plays a crucial role in the catalytic activity. Non-reducible oxide supports like Al2O3 and SiO2
disperse the precious metals and enhance the surface area. However, reducible rare-earth oxides like CeO2 can release oxygen creating oxide ion vacancies. This is commonly known as oxygen
storage capacity (OSC) and is a necessary step in the incorporation
of oxygen ions from diatomic gaseous oxygen into a solid. The main
100

Ce0.98Pt0.02O2-
Ce0.98Rh0.02O2-
Ce0.98Pd0.02O2-

80

100

60

80

N2 selectivity (%)

strong interaction generated with noble metals alters their electronic properties and adsorption properties [64].
NO molecule needs to be dissociated for its conversion to nitrogen. NO, when dissociatively chemisorbed, produces N2 while
molecularly adsorbed NO leads to N2O. Therefore, the NO reduction
by CO should be a site-specic reaction wherein CO should be
molecularly adsorbed and NO should be dissociatively chemisorbed. Thus a catalyst should be designed in such a fashion that
there should be sites for CO adsorption and NO dissociation. Towards this objective, an entirely alternative approach towards
TWC was performed in our laboratory where noble metals in their
ionic form were substituted in the reducible support like CeO2 or
TiO2 by a novel solution combustion synthesis method leading to
Ce1xMxO2d or Ti1xMxO2d, where M = Rh, Ru, Pd, Pt, Cu, Ag, Au.
The idea here was to have higher metal dispersion in the form of
ions and to create oxide ion vacancies due to lower valent ions
substituted in CeO2 or TiO2. The solid solution Ce1xRhxO2d, where
Rh in 3+ ionic state substituted for Ce4+, showed higher rates and
more product selectivity in NO + CO reaction compared to Rh metal-supported CeO2. The noble metal ion acts an adsorbent for CO
while the oxide ion vacancy dissociates NO to nitrogen. In this catalyst, the Rh3+OCe4+ interaction plays a key role towards the
higher activity of Ce1xRhxO2d [65]. This catalyst also showed high
rates of oxidation of CO and propane.
Several studies investigated the use of Rh-free catalysts and it
was found that Pt showed good three-way catalysis properties
[66]. Several metal oxides were added to Pt/SiO2 and their effects
on CO oxidation and NO reduction reactions were studied. It was
observed that Pt/CoOx/SiO2 showed remarkably good low-temperature CO oxidation activity, while Pt/MnOx/SiO2 showed good performance in the NO reduction by CO [66]. Granger et al. have shown
that at low-temperature the overall conversion of NO over Pt is better than the other noble metals [67]. Pt supported on CeO2 synthesized by solution combustion method forms a solid solution where
Pt is in ionic state and creates an oxide ion vacancy. The NO reduction by CO over Ce1xPtxO2d showed orders of magnitude higher
rate than Pt/Al2O3 catalyst, and that enhanced rate has been attributed to the dissociation of NO in the oxide ion vacancy [68]. However, the selectivity of N2 and N2O was not measured. Further it
was proved that the rate of the reaction increases with the increase
in oxide ion vacancies by substituting La or Y in CeO2 with Pt [69].
Supported bimetallic catalysts often exhibit certain desirable
properties (e.g. improved activity, selectivity, thermal stability,
and poison resistance), which are absent in each of the individual
metals. PtRh bimetallic catalysts exhibit synergism [70] and show
substantially higher catalytic activity than the physical mixture of
the Pt and Rh catalysts. Similarly, the catalytic synergy effects between cobalt phases and noble metals have been observed in Co
Pt(Pd, Rh)/CeAlO catalysts [71].
Recently we have shown that the enhanced catalytic activity of
Ce1xPtx/2Rhx/2O2d (x = 0.01) bimetal ionic catalysts compared to
mono-metal ionic catalysts, Ce1xPtxO2d and Ce1xRhxO2d. The enhanced catalytic activity of these bimetal ionic catalysts is attributed to synergism due to the easy reduction of Rh3+ ion by Pt2+ ion
[72]. Other than this, palladium can also be substituted in the catalyst. Palladium, among all other noble metals, is not only more plentiful, but also it has been found to be more durable at higher reaction
temperature [73]. Thus a catalytic converter utilizing a Pd-only catalyst can be positioned nearer the engine than a PtRh catalyst that
could potentially decrease the cold-start emission [44]. Pd-based
catalyst exhibits a high rate of deNOx activity at low-temperature
and also shows higher product selectivity than Pt or Rh based catalysts [74]. Fig. 4 shows the conversion of NO and N2 selectivity over
the three catalysts, Ce1xMxO2d (M = Rh, Pt, Pd).
Goodman and co-workers have contributed signicantly to the
understanding of NO reduction over Pd metal [75,76]. They found

% NO conversion

2288

40

20

60

40

20

0
150

200

250

300

350

400

450

500

Temperature ( C)

0
100

200

300

400

500

600

700

800

900

1000

Temperature ( C)
Fig. 4. TPR prole of NO in NO + CO reaction over of Ce0.98Pt0.02O2d,
Ce0.98Rh0.02O2d and Ce0.98Pd0.02O2 [74]. The N2 selectivity is shown in the inset.

S. Roy et al. / Applied Energy 86 (2009) 22832297

role of this is to extend the three-way window on the lean side of


stoichiometry by acting as a sink for gas-phase oxygen during
rich-to-lean transients. Further, the oxygen storage component
can also promote oxidation of reductants, like CO and hydrocarbons
during lean-to-rich transients. CeO2 can be reduced to Ce2O3 and
oxidized back to CeO2 [86]. Attempts to stabilize the interactions
between the noble metals and ceria were successful since mid
1990s. Ceriazirconia mixed oxide systems have been considered
as a substitute for ceria on the basis of their greater oxygen storage/release. This could potentially decrease the cold-start emission,
mainly by allowing the catalyst to be located in position closer to
engine manifold [87,88]. It was shown that in NO reduction by
CO, the turnover frequency of NO over Pd/Ce0.6Zr0.4O2 is about
two orders of magnitude higher than that of Pd/CeO2/Al2O3
[89,90]. Substitution of Zr in CeO2 to the extent of 30%, i.e. Ce0.7Zr0.3O2, helps Ce4+ to reduce to 3+ state more easily than CeO2 itself
enhancing OSC. Enhancement of OSC due to Zr substitution in CeO2
is shown to be due to destabilization of oxide ion sublattices in CeO2
leading to 4 + 4 coordination around both Ce and Zr ions [91]. This
leads to creation of long CeO and short CeO bonds and oxygen
from the longer (weaker) CeO bond can be easily extracted by
CO forming CO2. Thus, the creation of oxide ion vacancy is facilitated by the substitution of Zr in CeO2. It is not yet clear if La3+
ion substitution leads to such enhancing catalytic property, though
some work showed that La3+ ions can be directly incorporated into
the ceria promoter to increase its oxygen storage capacity through
extrinsic defects [92]. Pd supported CeTi solid solution has also
showed a promising TWC activity due to its enhanced reducibility.
The labile lattice oxygen was attributed to play an important role in
low-temperature catalytic activity of NO reduction by CO [93].
2.3. deNOx by H2 and NH3
As H2 is present in the exhaust, it can act as a TWC reductant to
NO. In fact, NO + H2 is a comparatively low-temperature reaction
compared to that of NO + CO. The main products of this reaction
are N2 and H2O. However, unselective products like N2O and NH3
are also formed. The NO reduction by H2 over oxidized and reduced
Rh supported catalysts has been investigated but the selectivity of
N2 to N2O was low [94]. A small enhancement of selectivity was
observed when Rh was used with Sn. No signicant change was observed in the reaction rate and the product selectivity with other
Group VIII metals.
NO reduction by H2 and the complete reaction mechanism over
the PtRh(1 0 0) alloy single crystal surface was also studied [95]
and found that Rh exhibits an excellent activity for the selective
reduction of NO towards N2, because of more NO dissociation
and more probability of pairing-up of N-adatoms on Rh surface.
Pt, on the other hand, is more selective in the reduction of NO towards NH3 and N2O under net-reducing conditions. Nieuwenhuys
and co-workers used Pt as a catalyst for the reaction, but the formation of ammonia was signicant at high concentrations of H2
[96]. Unlike the NO + CO reaction where only NO dissociates for
more N2 formation, in this case, both NO and H2 dissociate. We
have shown that over Ru, Rh, Pd, Pt ion substituted TiO2, rates of
NO reduction by H2 depends the reducibility of the catalysts. By
H2-uptake and electrochemical analysis, it was found that the catalyst, which has better reducibility at lower temperature or at lower potential, shows more NO + H2 reaction activity. Among the four
catalysts, Ti0.99Pd0.01O2d showed better H2 adsorption/dissociation
property and consequently better catalytic activity [97]. The reduction of NO by hydrogen on Pt/Al2O3 as a function of temperature
has been examined. N2O is formed from the NO chemisorbed on
Pt metal at low temperatures. However, at higher temperatures,
the formation of N2 is predominant because the reaction rate for
N2 formation is higher than that for N2O formation indicating that

2289

reductive conditions and appropriate reaction temperatures are


important factors in determining the selective formation of nitrogen from the reduction of NO.
A detailed overview on the reduction of NO with hydrogen in
presence of supported Pt-, Rh- and Pd-based catalysts has been reported and the activity follows the order: Pt/Al2O3 > Pd/Al2O3 > Rh/
Al2O3 [98]. The rate enhancement observed for the NO + H2 reaction has been mainly related to the involvement of a dissociation
step of chemisorbed NO molecules assisted by adjacent chemisorbed H atoms. The kinetics mechanisms conrm that Pd and
Rh are predominantly covered by chemisorbed NO molecules and
the lack of ammonia formation on Rh/Al2O3 during the reaction. Indeed the generation of chemisorbed H atoms may assist the dissociation of NO that explains signicant shifts of the light-off curves
towards lower temperatures in comparison with those obtained
when CO is used as a reducing agent.
NO reduction by NH3 in absence of oxygen has also been investigated [99] and the reaction was over Ce0.95Cu0.05O2d, thus
achieving a N2 to N2O ratio of 7.

10NH3 10NO ! 7N2 N2 O 9H2 O 4NH3

16

2.4. Selective catalytic reduction (SCR) of NO by H2 or NH3 or HC


DeNOxing from stationary sources and mobile sources can be
efciently achieved by using the SCR process in which NO is reduced by hydrogen or ammonia or hydrocarbons in presence of excess oxygen. Perovskite-type oxides of general formula ABO3
(where A is usually a rare-earth metal coordinated by 12 oxygen
atoms and B is usually a transition metal surrounded by six oxygen
atoms in octahedral coordination) have been investigated for catalytic converter applications since the early 1970s [100]. The oxidation state of B cations and the structural defects can be changed by
partially substituting A and/or B with metals (A0 , B0 correspondingly) of different oxidation states. Perovskites like LaCoO3,
LaCo1xCuxO3d, have been suggested as potential NOx abatement
catalyst for automobile exhaust control [101,102]. The partial substitution of lanthanum and manganese to form mixed oxides
La1xAxMn1yByO3 and its effect on the catalytic activity was investigated. La3+ has been partially replaced by A+, A2+, or A4+ ions in order to obtain Mn ions in various oxidation states or to create O2
vacancies in the lattice, which can dissociate NO. La0.8K0.2MnO3
showed more selective NO reduction than LaMnO3 [103]. Mn3+
was also partially substituted by other catalytically active transition metals, such as Cu2+, leading to much higher activity for the
CO + NO reaction [104,105].
Incorporating noble metals into a perovskite structure can stabilize the noble metal against sintering, reaction with the support,
or volatilization. It was found that the perovskite La0.5CexSr0.5xMnO3 exhibited a relatively low catalytic activity but the addition of
0.1 wt% Pt resulted in a catalyst showing high activity and N2 selectivity for the SCR [106]. This was attributed to the oxygen vacancy
sites of the support located next to the small Pt clusters that could
provide the means for the formation of adsorbed NO with the N
atom located on the Pt metal and the O atom on the oxygen vacancy. In addition, adsorbed oxygen formed on the oxygen vacancies of the support located at the metalsupport interface is more
active in removing adsorbed hydrogen from the Pt surface [107].
Rh containing LaMn1xRhxO3 shows better TWC activity than LaMnO3 [108]. However, the rate of the reactions and the selectivity of
products are not better than ceria-supported noble metal catalysts.
Recently, another perovskite-type catalyst containing Pd, LaFe0.57Co0.38Pd0.05O3, was reported, which had high catalytic activity
and high metal dispersion due to the structural responses to redox
atmospheres at 800 C [109]. Structural analysis showed that cationic Pd occupied the B-site of the ABO3 type perovskite crystal

2290

S. Roy et al. / Applied Energy 86 (2009) 22832297

structure and that PdO segregated out to form a metallic alloy of


PdCo in the reductive atmosphere. This movement of Pd suppressed the agglomeration and growth of the metal particles and
was termed as the self-regeneration of Pd. The mobility of Pd
in and out of the perovskite lattice requires a high energy of activation and indeed this catalyst has been operated at very high temperature (900 C).
However, additional fundamental approaches [110,111] show
that these processes may occur at relative low-temperature
corresponding to the usual temperature conditions of three-way
catalysts. Because Co is a carcinogenic, another perovskite,
LaFe0.95Pd0.05O3 was synthesized [112,113], which also showed
self-regeneration and is a commercial automotive catalyst in Japan.
The stabilization of well-dispersed PdOx entities in strong interaction with the perovskite structure could be attractive for promoting the decomposition of N2O and may represent an
interesting practical issue for the replacement of Rh based catalysts. An alternative strategy than that previously developed by
Uenishi et al. [114] was proposed for the promotion of well-dispersed Pd/LaCoO3 with lower Pd content and higher surface Pd
concentration at the surface. This consists of a two-step procedure
with a conventional wet impregnation of Pd precursor and appropriate successive reductive and oxidative thermal treatments
[115]. Those thermal treatments are accompanied with surface
structural modications that favor the re-dispersion of oxidic palladium species. According to this procedure [115], a signicant rate
enhancement is observed which cannot be related to weak interactions between PdO and LaCoO3 but mainly to the occurrence of redispersion processes which could be governed by surface reconstructions of the perovskite structure during those successive thermal ageing. In that case, oxidic Pd species strongly interacting with
LaCoO3 would be stabilized in unusual oxidation states. The deposition of palladium on reducible supports such as LaCoO3 leads to
higher activity from temperature-programed experiments in comparison with conventional supports such as alumina. Interestingly,
successive reductive and oxidative thermal treatments in the reactant mixture at high temperature enhance the conversion of N2O
particularly on perovskite support and show a re-dispersion and
the stabilization of palladium species in unusual oxidation states
which would originate a rate enhancement in the decomposition
of N2O [116].
The industrial catalysts for SCR from stationary sources are
based on TiO2 supported V2O5WO3 and/or V2O5MoO3 oxides
[117120] while mobile systems use zeolite based catalysts. Anatase form of TiO2 is the support of choice mainly because SO2 poi-

O
O

soning does not take place on TiO2 [121,122]. Generally the active
sites on the V2O5WO3/TiO2 and V2O5MoO3/TiO2 industrial SCR
catalysts are vanadium oxide species [121123]. It is interesting
to note that V2O5/TiO2 (anatase) is unstable where TiO2 (anatase)
is a metastable polymorph that converts into thermodynamically
stable form rutile at higher temperature and pressure. V2O5 favors
this transformation and the anatase sintering and loss of surface
area. However, with WO3 and MoO3, both the reduction of surface
area and transformation of anatase to rutile are lower [124,125].
Further these catalysts act as inhibitors for SO2 oxidation
[125,126]. SCR reaction is a redox process that occurs with a redox
or Marsvan Krevelen-type mechanism on vanadium-based catalysts. NH3 adsorbs on pure V2O5, on V2O5TiO2, on V2O5/SiO2
TiO2, on V2O5WO3/TiO2 and on V2O5MoO3/TiO2 in two different
strongly held species: (i) molecularly adsorbed ammonia, through
a Lewis-type interaction and (ii) ammonia observed as ammonium
ions, over Bronsted acidic OH surface hydroxyl groups [127129].
15
N NMR experiments showed that the adsorption on Lewis site is
predominant in V2O5TiO2 [130]. The TiO2-anatase supports, only
shows Lewis acidity [131], whereas ammonium ions are formed
on VOH sites. The adsorption of NO (the other SCR reactant) has
also been extensively investigated in the literature. It has been
shown that the interaction of NO is very weak over many V2O5based catalysts. By adsorption of NO over V2O5TiO2, Ramis et al.
observed the formation of a surface nitrosyl species, coordinated
to Ti4+ sites [132]. However, NO does not adsorb on an ammoniacovered surface because NH3 has greater basicity and blocks the
Ti4+ adsorption sites. This data is in good agreement with the kinetic observations on V2O5-based catalysts where zero and rst order kinetics with respect to NH3 and NO, respectively [133136] is
observed. In vanadia-based catalysts, NH3 adsorbs mainly on the
Bronsted acid sites while insignicant NO adsorption is observed.
Thus an EleyRideal mechanism has been suggested for SCR. Takagi et al. proposed one of the rst reaction schemes for SCR over
V2O5-based catalysts [137]. Subsequently, Inomata and his group
proposed a popular reaction mechanism [138140] in which the
reaction of NH
4 species occurs with gaseous NO through an activated complex (Fig. 5). According to Janssen et al., the polyvanadate species, O@VAOAV@O adsorbs ammonia over and leads to
the intermediate species VONH2 [141]. In 1990s, Ramis et al. proposed the reaction pathway [132] in which ammonia is adsorbed
over a Lewis acid site that activates ammonia to an amide NH2 species resulting in catalyst reduction. This amide species then reacts
with gas-phase NO to nitrosamide as an intermediate, which
decomposes to nitrogen and water. The reduced catalyst sites are

OH

Vs

O
O

H2O
Reaction (2)

NH3
fast

OH

Reaction (1)

O2

Vs

NO
O
N H

OH
O

Vs

O
O

H
OH

OH
O

Vs

N2 + H2O
Activated Complex
Fig. 5. Mechanism of SCR over vanadium oxide catalysts [138].

S. Roy et al. / Applied Energy 86 (2009) 22832297

80

Ti0.9Cu0.1O2-

40
0
80

Ti0.9Co0.1O2-

% NO conversion

40
0
80

Ti0.9Fe0.1O2-
40

Ti0.9Mn0.1O2-

0
80
40
0
80

Ti0.9Cr0.1O2-

40
0
100

200

300

400

500

Temperature ( C)
Fig. 6. NH3-SCR over Ti0.9M0.1O2d (M = Cr, Mn, Fe, Co, Cu) [158].

then regenerated by gas-phase oxygen. This is the rst mechanism


proposed for vanadia-based catalyst that suggested activation of
ammonia on Lewis acid cites.
This concept was popularized and proven by the theoretical
density-functional theory (DFT) calculations [142144]. A recent
density-functional theoretical calculation shows that on V2O9H8
cluster surface, SCR reaction is initiated more favorably by the
ammonia activation on Brnsted acidic VOH site, where NH4+
ion is formed. Then with interaction with gaseous NO, NH2NO adduct is formed, which decomposes into N2 and H2O [145].
Despite its problems like accounting for band-gaps and longrange forces [146], it is a versatile tool that can be used for verifying the kinetic mechanisms. We discuss a couple of examples
where DFT has been successfully used for determining the mechanisms in SCR. Among the various catalysts for SCR, Ag/Al2O3 is one
of the superior catalysts. In the SCR of NOx, the adsorbed NO
3 species on the Al2O3 and the Ag/Al2O3 play an important role in the
kinetics. The adsorbed NO
3 could be bonded to metal cations on
the surface via the following adsorption structure: bridging, bidentate and monodentate [147]. The interpretations of the infrared
spectra proposed by researchers can be different. For example,
the band observed around 1300 cm1 has been attributed to monodentate nitrate [148] bidentate nitrate [147] or solvated nitrate
[149]. Thus the assignment and structures of the nitrates species
on Ag/Al2O3 and Al2O3 were still not clear. The vibrational modes
of nitrate species were simulated by DFT and analyzed by Gaussian
98 and Hyperchem 7.0 program. Based on the density-functional
theory calculations, the band at 1304 cm1 was assigned to an isolated bidentate nitrates species, whose formation could involve lattice oxygen of Al2O3 [150]. A possible mechanism for NOx reduction
by C2H5OH over Ag/Al2O3 was proposed to be similar to that of
C3H6 [151] However, this mechanism cannot explain why
C2H5OH has a higher efciency for the SCR of NOx over Ag/Al2O3

2291

than hydrocarbons such as C3H6. Thus a different mechanism


based on surface enolic species on Ag/Al2O3 was proposed. DFT calculations were used to conrm the structure of adsorbed enolic
species on Ag/Al2O3. Thirty-two models of adsorbed species were
calculated and different kinds of adsorbed species and the interaction of the surface with the adsorbed species were considered.
However, DFT calculations were in good agreement with the
experimental value for only the enolic species conrming the
mechanism of NOx reduction with ethanol in presence of Ag/
Al2O3 [152] Thus, DFT can be successfully used to test and conrm
mechanisms of NOx reduction.
The three major aspects of SCR reaction are to reduce the reaction temperature, enhance the SCR window and increase the product selectivity. The V2O5-based catalysts do operate at
comparatively higher temperature. Extensive studies on SCR of
NO have been carried out over other rst row transition metal
oxide than vanadium. Mn oxides based catalysts are becoming
popular for SCR reactions because of its low-temperature activity.
Kapteijn and his group did extensive work on Mn-based catalysts
for SCR process [153155]. In unsupported Mn-oxide studies it
was found that SCR activity in the following order: MnO2 > Mn5O8 > Mn2O3 > Mn3O4. In alumina-supported manganese oxide,
they found two Lewis-type coordinatively unsaturated Mn3+ ions
as the active species. A comparatively recent work based on Mnoxide-supported hombicat TiO2 has shown NH3 adsorption onto
the Lewis acid sites of Mn4+, and has given a conclusive mechanism
for unselective N2O formation [156,157]. Though the Mn-based
catalysts operate at lower temperature, the product selectivity is
poor. In our work on rst row transition metal substituted anatase
TiO2, SCR catalytic activity of the materials was tested.
Ti0.9Mn0.1O2d showed low-temperature activity whereas
Ti0.9Fe0.1O2d showed the widest SCR window and better product
selectivity (Fig. 6). The active phase for the catalytic activity is
the substituted anatase phase, not the illmenite phase. NH3-TPD
studies showed that Ti0.9Mn0.1O2d has the highest Bronsted acidity and Ti0.9Fe0.1O2d has the highest Lewis acidity among all other
catalysts [158]. Therefore, a solid solution of Ti0.9Mn0.05Fe0.05O2d
was synthesized and found to be the best catalyst in terms of
low-temperature reactivity and product selectivity. Several papers
[159,160] have investigated the kinetics over conventional V2O5
WO3/TiO2 catalysts and shown that NH3 is stored in the Ti or W
surface and when the feed NH3 concentration is lowered, the
stored ammonia reacts with the NO to form N2.
Cu2+-exchanged ZSM-5 zeolites are active catalysts for the
reduction of nitric oxide with ammonia in the presence of oxygen.
It was shown that NO reduction by NH3 over Cu(II) ion-exchanged
Y-type zeolites [Cu(II)NaY] followed LangmuirHinshelwood
kinetics [161]. Many studies suggest that the ammonia SCR reaction takes place by the strong adsorption of ammonia on the catalyst, and that the molecularly adsorbed ammonia species interacts
with NO from the gas phase or from a weakly adsorbed state
through an EleyRideal type mechanism [162]. It has been reported that Fe2O3 [163], Fe containing mixed oxides [164] and
Fe-exchanged materials [165], and in particular FeZSM-5 [166
169] show signicant SCR activity.
A comparison of different metal-exchanged zeolites showed iron
beta as having good activity, nitrogen selectivity and ageing characteristics [170], however studies of low-temperature SCR of NO with
ammonia over zeolite based catalysts are few [171]. The reaction
shows considerable sensitivity to the nature of the oxide support
and a comparative study of iron oxide on different oxide supports
including MgAl2O4, SiO2, TiO2 and ZrO2 has been performed [162].
Besides using zeolites ZSM-5 and zeolite beta (BEA) in these reactions, different synthetic and natural zeolitic materials such as
mordenite (MOR), heulanditeclinoptilolite (HEU), ferrierite (FER)
and chabazite (CHA) have also been investigated [172]. Among

2292

S. Roy et al. / Applied Energy 86 (2009) 22832297

them, Femordenite and Feclinoptilolite have shown catalytic


properties for the SCR of NO with ammonia, similar to that of Fe
ZSM-5 [172]. Natural zeolites [173] have also been used because
they have a good combination of their crystalline structures and exhibit interesting physicochemical properties with various metal
oxy-hydroxide phases that are naturally embedded inside their
pores [174]. However, potential commercialization of natural zeolites as catalysts is not feasible due to the cost of homogenisation
and purication of the zeolite-bearing tuffs [175].
Unreacted ammonia can escape in the exhaust (called slip) and
thus the emission controls also restrict ammonia in the exhaust.
Further, the transportation and storage of ammonia is not very
cost-effective. This led to research on the NOx reduction by hydrocarbons as reducing agents.
The rst catalyst that was found to have a good hydrocarbon NOx
reduction capability in oxygen rich conditions was Cu/ZSM-5
[26,34]. Subsequently, this reaction has been extensively studied
[176,177] over different cation exchanged zeolites. It has also been
extensively studied over different base oxides and these oxides promoted by Co, Ni, etc. as supports [178]. However, many of these catalysts are deactivated in presence of water vapor [179].
Subsequently, another system Ag/c-Al2O3 catalyst was found to
be active for HC-SCR [180]. The optimal silver content was reported
to be 2 wt% [181,182]. Thus a catalytic system based on silver supported alumina was developed by Klingstedt et al. [183] and
showed the optimum silver content, the correct choice of the alumina support, the ratio of hydrocarbon and NOx on the high performance of the HC-SCR deNOx [184]. While a maximum conversion of
90% was observed at 450 C, improved conversion could be obtained
if the catalyst was divided into a few layers [185]. However, one of
the drawbacks of this catalyst is its poor activity below 300 C and
large formation of CO. But the presence of a small amount of hydrogen in the feed improved the catalytic activity signicantly
[186,187]. This has given added impetus to this area of research
[188,189]. Hydrogen promotes the NOx reduction over Ag/c-Al2O3
catalysts when using a range of lower alkanes and alkenes and higher alkanes [190] as reductants. Silver has also been used in zeolites
for HC-SCR of NOx with various reducing agents [191193].
SOx is detrimental to activity at low temperatures but that at
higher temperatures the effect of SOx is minimal or may even enhance NOx conversion. Meunier and Ross [194] found that for a
1.2 wt% Ag/c-Al2O3 catalyst the activity decreased rapidly in the
presence of 100 ppm SO2 in the reaction feed and found that the
sulfation of Ag rather than c-Al2O3 was primarily responsible for
the deactivation. Abe et al. [195] noted that silver sulfate decomposes at a lower temperature (427 C) than aluminium sulfate
(727 C) and explained why catalysts tested in the literature maintain their activity in the presence of SOx at higher temperatures
(>427 C) but tend to be inactive at temperatures lower than this.
The effect of SOx is also inuenced by the nature of the reductant. For example, SOx enhanced the activity of the 5 wt% Ag/cAl2O3 catalyst [196] and a pre-sulfated Pt/Al2O3 catalyst [197]
when propene was used as a reductant but caused severe deactivation when propane was used. This is because the reaction occurs
mainly on the support for propane and is hindered by strongly
bound sulfates on the support. However, the reaction takes place
predominantly on the Pt sites which do not adsorb sulfur strongly
when using propene as a reductant. Reactive RSOx species seem to
be key intermediates in the SCR reaction with propene over Ag/
Al2O3 catalyst [198]. The effect of SOx on Ag/c-Al2O3 catalyst for
the NOx-SCR with hydrogen and hydrocarbon was investigated
[199] and that the effect of ageing was a function of the gas mix
and temperature of ageing. The effect of sulfur on the deactivation
was dependent on the temperature and was explained by the
activity of the catalyst for the oxidation of SO2 to SO3 and the relative stability of silver and aluminium sulfates.

For mobile SCR application urea can be used instead of NH3


[200]. Urea is usually applied as an aqueous solution in urea-SCR.
When this solution is atomized into the hot exhaust gas stream,
water evaporates from the droplets, thus leading to solid or molten
urea [201]. Urea then subsequently decomposes to ammonia and
isocyanic acid. The rate of HNCO hydrolysis is much higher than
that of the SCR reaction and thus ammonia is the active reducing
agent even when urea is used [201]. Due to compactness of the deNOx systems, only short residence times are realized and this leads
to considerable mixing problems for NOx with the reducing agent.
2.5. Other methods
In addition to the above methods, a few other techniques are
used for NOx reduction. The simplest approach, of course, is to
modify the operational conditions such that NOx formation is decreased. Thus furnaces with low production of NOx are devised
such that the NOx burners minimize thermal formation of NOx by
reducing combustion temperatures and controlling ame stoichiometry. However, this increases formation of CO and also has problems of corrosion and slags because of a local reducing
environment [202]. Another method is to inject methanol into ue
gases so that NO is converted to NO2. This is subsequently removed
using limestone scrubber. In the 1990s, two techniques namely the
NOx storage and reduction (NSR) and selective NOx recirculation
(SNR) were developed by the car industry. In both techniques,
NOx is adsorbed on a sorbent but the subsequent decomposition
procedure is different. In general, this material should have a high
NOx trapping capacity, a high selectivity and a high resistance to
SO2 poisoning. Finally, we discuss another method, namely the
non-catalytic selective catalytic reduction.
2.5.1. NSR (NOx storage and reduction)
The reduction of NOx takes place by a two-stage operation. During the fuel-lean stage, the NOx is trapped on an adsorbent in the
form of a nitrate. Then, the engine is switched to a fuel-rich condition where the hydrocarbons, hydrogen and CO react with the nitrate to yield nitrogen, water and carbon dioxide.
During the lean-burn stage in presence of excess oxygen, NOx is
oxidized by oxygen to NO2 over the platinum site and stored on the
barium oxide as barium nitrate. In this stage, hydrocarbons, H2,
and CO are readily oxidized into water and carbon dioxide. When
the engine is switched to operation with the normal airfuel mixture, HC, H2, and CO do not oxidize. So these reductants react with
the NO
3 stored in the catalyst and forms nitrogen, water, and carbon dioxide. Adopting this mode of operation, NSR catalysts were
developed and commercialized by Matsumoto from Toyota in
1994 [203205]. NOx storage materials consist of alkaline-earth
metals or alkaline metals and noble metals such as platinum and
rhodium dispersed on the support, e.g. PtBa/Al2O3. The support
of NSR catalysts has been changed from Al2O3 to CeO2 to CeO2
ZrO2. However, formation of BaAl2O4 and BaCeO3 has been observed around 850 C [206,207].
A study on Pt/Ba/Al2O3 catalyst for a NSR reaction with H2,
showed that at relatively low-temperature, the NOx storage capacity is sufcient, but that the rate of NOx release and reduction (NOx
regeneration) is slow and insufcient. Further, with a large amount
of H2 injection, a large amount of NH3 was detected as product
[208,209]. One other comparative study between Pt/Ba/Al2O3 and
Pt/K/Al2O3 catalysts for NSR reaction, showed that the under ow
conditions during a complete NSR cycle, barium containing catalyst shows better NSR performance. However, no ammonia was
produced [210] for the reaction over Pt/K/Al2O3. A main problem
of the NSR catalysts is sulfur poisoning. Sulfur oxides (SOx) in exhaust gas react on the catalyst in the same way as NOx. There
can be two types of sulfur poisoning [204] wherein SO2 reacts with

S. Roy et al. / Applied Energy 86 (2009) 22832297

alumina to form aluminium sulfate that hinders reactivity and


where SO2 competes with the NOx for BaO to form barium sulfate
(BaSO4). The presence of TiO2 suppresses the sulfate formation
without reducing the NOx storage capacity signicantly [211].
2.5.2. Selective NOx recirculation (SNR)
SNR technique deals with the treatment of exhaust gases from a
fuel-lean or diesel and was developed by DaimlerChrysler in
1994. In this method [212], two adsorbers are arranged in parallel
and operate in the adsorption and desorption modes alternately.
The principle of the concept consists in the concentration and
recirculation of NOx into the combustion chamber of the engine
where they are decomposed thermally [213]. Because thermal
decomposition depends on NOx concentration, a high performance
of the adsorbent material is the key for the success of the SNR concept [202,214].
It is implied from the information presented above that the
amount of sorbed NOx strongly depends on the sorption mechanism as well as on the physical properties of sorbents and sorbates.
Thus, for future practical applications, the sorbents have to retain
their adsorption capability for prolonged periods in presence of
water, CO2 and SO2.
2.5.3. Selective non-catalytic reduction (SNCR)
NOx reduction by NH3 or urea in presence of excess oxygen
without catalysts is also practiced and is called selective non-catalytic reduction (SNCR). Processes using ammonia, urea, or cyanuric
acid for the SNCR of nitric oxides from engine exhaust gases were
compared. Ammonia, urea, and cyanuric acid were found to be
most effective at low, intermediate, and high oxygen concentrations, respectively [215]. The reaction of NO and ammonia only occurs after a minimum temperature, while ammonia decomposes to
NO at higher temperatures. Thus the reaction needs a specic temperature window to be efcient. Further, the reaction also needs
sufcient reaction time in that temperature window and need to
be sufciently mixed to avoid ammonia slip [216]. It has been reported that injection of some additives together with the reducing
agents in SNCR processes can lower and widen the optimum reaction temperature window for NOx reduction [217]. Though selective non-catalytic reduction can achieve the same efciency of
about 90 % as SCR without a catalyst, practical constraints of temperature, time, and mixing often lead to worse results in practice.
One possible way of using SNCR, at lower molar ratios and minimum NH3 slip, may be the combination of SNCR processes with a
back-end NO oxidation process [216].
2.5.4. Ozone injection
Ozone injection is an attractive technology that is more energy
efcient than plasma and electron beam processes directly applied
to the exhaust gas for the oxidization of NO [218,219]. In this technique, small amounts of oxygen or air can be discharged to produce
ozone, which can then be injected into the ue gas. The key step in
this method is the oxidization property of NO in ue gas. Both kinetics and mixing play a key role in the application of this technology
and a good understanding the ozoneNOx reaction jet is fundamental to the design and optimization of this technology [220]. The effectiveness of ozone injection on the selective catalytic reduction of
nitrogen oxides has been investigated [221]. Nitric oxide in the exhaust gas is initially oxidized to nitrogen dioxide by ozone, and then
the exhaust gas containing the mixture of NO and NO2 is reduced to
nitrogen in a catalytic reactor. The ozone injection method was
determined to be very efcient for the oxidation of NO to NO2 in a
wide range of temperatures, and the increase in the content of NO2
by the ozone injection remarkably improved the performance of
the catalytic reactor for the reduction to nitrogen [222].

2293

3. Conclusions and future perspectives


There has been a signicant research in the area of NOx abatement recently. This review initially briey discussed the different
NOx sources, the different legislations existing in various countries.
The synthesis of various catalysts and the reaction mechanisms for
NOx removal by NO decomposition, NO reduction by CO or H2 or
NH3 or hydrocarbons was discussed.
Currently, two main methods for the removal of NOx from emission gases are employed namely the TWC developed for mobile
sources that use gasoline and the SCR, which is applied for stationary sources such as power plants and for lean-burn engines. The
use of TWC is a much more established technology for the catalytic
reduction of NOx produced by gasoline engines than compared to
the catalysts for vehicles with diesel and lean-burn gasoline engines. The industrial catalysts for SCR from stationary sources are
based on TiO2 supported V2O5WO3 and/or V2O5MoO3 oxides
while mobile systems for SCR use zeolite based catalysts.
The TWC has three functions, namely the oxidation of CO, the
oxidation of unburnt hydrocarbons (HCs) and the reduction of
NOx to nitrogen. They work well during operation at or close to
stoichiometric conditions. However, when the engine runs at
fuel-rich conditions, the overall exhaust gas composition is reducing in nature, and it is difcult to carry out oxidation reactions on
the catalyst surface. TWCs have been developed to incorporate
ceria that stores oxygen during leaner periods of the operating cycle, and releases oxygen during richer periods of the operating cycle. However, ceria, when doped with metal such as Pd loses
surface area at high temperatures of 800 C. Thus mixed oxides
of ceriazirconia are often used instead of ceria as the oxygen storage component. Ceria itself is a rare-earth metal with restricted
suppliers and ceriazirconia is a relatively expensive material
when available commercially, and it would be desirable to nd a
material having at least as good oxygen storage performance as
ceriazirconia, but utilizing less expensive materials.
Although selective catalytic reduction by ammonia is currently
the most widespread method for the clean up of ue gas from stationary sources, many problems still exist. The advantages of NH3
as the reducing agent are high selectivities toward the reaction
with NO in the presence of O2 and the promoting effect of oxygen
on the reaction kinetics of NO with NH3. The most signicant technical parameters are the positioning of catalyst from the source of
gas and the design of ammonia injection. Typically, during a commercial SCR process over V2O5WO3TiO2 systems, stoichiometric
control of the ammonia must be maintained to avoid emissions of
unreacted ammonia. In addition, there are difculties of transport
and storage of ammonia. It would be necessary to develop new catalysts operating with low or high temperature; and it would be
desirable to substitute ammonia by another reductant because of
the dangers of storage, leakage, and transport of liquid ammonia.
For mobile applications, the problems are more complicated.
The rst generation of commercial SCR catalysts for mobile applications were monoliths made of anatase TiO2 supporting V2O5 or
WO3, similar to the vanadium-based catalysts used for stationary
SCR applications. However, the stringent legislation on NOx emissions, the necessity of catalysts to be active up to higher temperatures and the toxicity of vanadium have driven the research focus
towards hydrogen- and metal-exchanged zeolites. Since the discovery of HC-SCR technology, various types of catalysts have been reported, such as ion-exchanged zeolites, supported precious
metals, and metal oxide-based catalysts. Ion-exchanged zeolites
showed very high performance on HC-SCR. However, they are
unstable in hydrothermal conditions that cause deactivation of catalysts. Supported precious metal catalysts, especially supported Pt
catalysts, show high stability and high tolerance to sulfur oxides

2294

S. Roy et al. / Applied Energy 86 (2009) 22832297

(SOx) and water vapor. They show very high activity at lower temperatures, but the non-selectivity to nitrogen and a narrow temperature range for NOx reduction are problems. Metal oxide-based
catalysts show high stability and moderate tolerance to SOx and
water vapor and in this context, Ag/c-Al2O3 has been developed.
However, one of the drawbacks of silver supported alumina catalyst
is its poor activity below 300 C and large formation of CO. Thus it
has been combined with a suitable catalyst to oxidize the formed
CO and unburned hydrocarbons and thus an dual-catalytic system
has been developed wherein the platinum catalyst is placed after
the Ag/alumina catalyst. However, the selectivity to nitrogen of this
dual system is signicantly lower and thus further investigations
are needed to obtain a good dual system without loss of activity
and selectivity. In this context, the self-regenerating Pd perovskite
oxides seem promising. However, the activation of oxygen in the
perovskite support is not signicant and needs to be improved.
An alternative to the catalytic approach is based on NOx adsorption. NOx is stored in the catalyst under lean conditions and is regenerated for a short period under fuel-rich conditions. However, this
technique fails when fuels contain high levels of sulfur. Overcoming
the problems of preserving adsorbents catalytic activities in presence of water, CO2, and SO2 and the selectivity of noble metals to
N2O and NH3 formation are some of the keys for the future practical
application of sorbing catalytic material in NOx depollution.
New approaches to develop novel catalysts may also be investigated. One can modify the surface of an oxide by substituting a
fraction of metal atoms with another metal. The presence of another metal disrupts the bonding in the oxide. This leads to two
phenomena: the oxygen atoms whose proximity is close to the metal become chemically reactive and aid oxidation while the metal
also becomes reactive and is able to adsorb/activate oxygen. Several methods have been developed for synthesizing these types
of compounds. However, one can further the OSC if one is able to
substitute lower valent metal ions in the reducible matrix. Thus,
there would be two distinct sites for reducing and oxidizing molecules in the ionic catalysts unlike metal surfaces. Since the adsorption sites are next to each other, the electron transfer from
reducing molecules to oxygen would be facilitated by the lattice
leading to very high catalytic activity. In such a compound, the metal is fully dispersed as ions and they cannot sinter due to ionic
repulsion. Thus one could have a compound in which higher metal
dispersions in the form of ions are observed and oxide ion vacancies are created due to lower valent ions substituted in CeO2 or
TiO2. These catalysts may yield high rates of reaction for SCR.
Several challenges have to be faced when trying to solve the
problem of NOx pollution with a catalytic system: selectivity, operational temperature, and poisoning. From the studies highlighted
in this article, it is clear that more work needs to be carried out
to understand the mechanism of NO decomposition and reduction
under various operating conditions.
References
[1] Armor JN. Environmental catalysis. Appl Catal B: Environ 1992;1:22156.
[2] Centi G, Ciambelli P, Perathoner S, Russo P. Environmental catalysis: trends
and outlook. Catal Today 2002;75:315.
[3] Klingstedt F, Arve K, Eranen K, Murzin DU. Toward improved catalytic low
temperature NOx removal in diesel powered vehicles. Acc Chem Res
2006;39:27382.
[4] Helms GT, Vitas JB, Nikbakht PN. Regulatory options under the US clean air
act: the federal view. Water Air Soil Pollut 1993;67:20716.
[5] Bosch H, Janssen F. Formation and control of nitrogen oxides. Catal Today
1988;2:36979.
[6] Garin F. Mechanism of NOx decomposition. Appl Catal A: Gen
2001;222:183219.
[7] Crutzen PJ, Bruh C. Catalysis by NOx as the main cause of the spring to fall
stratosphere ozone decline in the northern hemisphere. J Phys Chem A
2001;105:157982.

[8] Ravishankara AR. Introduction: atmospheric chemistry long-term issues.


Chem Rev 2003;103:45058.
[9] Fritz A, Pitchon V. The current state of research on automotive lean NOx
catalysis. Appl Catal B: Environ 1997;13:125.
[10] Snyder S, Bredt D. Les fonctions biologiques du monoxyde dazote. Pour La Sci
1992;177:707.
[11] Chiron M. Air pollution by automobile exhaust and public health. Stud Surf
Sci Catal 1987;30:110.
[12] Mauzerall DL, Sultan B, Kim N, Bradford DF. NOx emissions from large point
sources: variability in ozone production, resulting health damages and
economic costs. Atmos Environ 2005;39:285166.
[13] Erisman JW, Greenfelt P, Sutton M. The European perspective on nitrogen
emission and deposition. Environ Int 2003;29:31125.
[14] <http://www.dieselnet.com/standards/in>.
[15] Broden G, Rhodin TN, Bruker C, Benbow R, Hurych Z. Synchrotron radiation
study of chemisorptive bonding of CO on transition metals polarization
effect on Ir(1 0 0). Surf Sci 1976;59:593611.
[16] Egawa C, Onawa K, Iwai H, Oki S. Interaction of CO and NO with Fe thin lms
grown on Rh(1 0 0) surface. Surf Sci 2004;557:3140.
[17] Brown WA, King DA. NO chemisorption and reactions on metal surfaces: a
new perspective. J Phys Chem B 2000;104:257895.
[18] Whitman LJ, Ho W. The kinetics and mechanisms of alkali metal-promoted
dissociation: a time resolved study of NO adsorption and reaction on
potassium-precovered Rh(1 0 0). J Chem Phys 1988;89:762146.
[19] Ho W. Time resolved electron energy loss spectroscopy of surface kinetics.
Electron Spectrosc Relat Phenom 1987;45:118.
[20] Palermo A, Lambert RM, Harkness IR, Yentekaksi IV, Marina O, Vayenas CG.
Electrochemical promotion by Na of the platinum-catalyzed reaction
between CO and NO. J Catal 1996;161:4719.
[21] Goddard PJ, West J, Lambert RM. Adsorption, coadsorption, and reactivity of
sodium and nitric oxide on Ag(1 1 1). Surf Sci 1978;71:44761.
[22] Winter ERS. The catalytic decomposition of nitric oxide by metallic oxides. J
Catal 1971;22:15870.
[23] Oh SH, Eichel CC. Effects of cerium addition on CO oxidation kinetics over
alumina-supported rhodium catalysts. J Catal 1988;112:54355.
[24] Boreskov GK. Forms of oxygen bonds on the surface of oxidation catalysts.
Discuss Faraday Soc 1968;41:2639.
[25] Park PW, Kil JK, Kung HH, Kung MC. NO decomposition over sodiumpromoted cobalt oxide. Catal Today 1998;42:5160.
[26] Iwamoto M, Hamada H. Removal of nitrogen monoxide from exhaust gases
through novel catalytic processes. Catal Today 1991;10:5761.
[27] Fang SM, White JM. Chemisorption of nitric oxide on platinized titania. J Catal
1983;83:18.
[28] Pande NK, Bell AT. A study of the interactions of NO with Rh/TiO2 and TiO2promoted Rh/TiO2. J Catal 1986;97:13749.
[29] Chin AA, Bell AT. Kinetics of nitric oxide decomposition on silica-supported
rhodium. J Phys Chem 1983;87:37006.
[30] Ward TR, Alemany P, Hoffman R. Adhesion of rhodium, palladium, and
platinum to alumina and the reduction of nitric oxide on the resulting
surfaces: a theoretical analysis. J Phys Chem 1993;97:76919.
[31] Iwamoto M, Yokoo S, Sakai K, Kagawa S. The carbonylation of alcohols
catalyzed by Cu (I) carbonyl. J Chem Soc Faraday Trans 1981;77:162936.
[32] Iwamoto M, Furukawa H, Mine Y, Uemura F, Mikuriya S, Kagawa S. Copper(II)
ion-exchanged ZSM-5 zeolites as highly active catalysts for direct and
continuous decomposition of nitrogen monoxide. J Chem Soc, Chem Commun
1986;11:12723.
[33] Iwamoto M, Yahiro H, Mine Y, Kagawa S. Excessively copper ion-exchanged
ZSM-5 zeolites as highly active catalysts for direct decomposition of nitrogen
monoxide. Chem Lett 1989;19:2137.
[34] Iwamoto M, Yahiro H, Tanda K, Mizuno N, Mine Y, Kagawa S. Removal of
nitrogen monoxide through a novel catalytic process. 1. Decomposition on
excessively copperion-exchanged ZSM-5 zeolites. J Phys Chem
1991;95:372730.
[35] Giamello E, Murphy D, Magnacca G, Morterra C, Shioya Y, Nomura T, et al. The
interaction of NO with copper ions in ZSM5: an EPR and IR investigation. J
Catal 1992;136:51020.
[36] Cheung T, Bhargava SK, Hobday M, Foger K. Adsorption of NO on Cu
exchanged zeolites, an FTIR study: effects of Cu levels, NO pressure, and
catalyst pretreatment. J Catal 1996;158:30110.
[37] Schneider WF, Hass KC, Ramprasad R, Adams JB. First-Principles analysis of
elementary steps in the catalytic decomposition of NO by Cu-exchanged
zeolites. J Phys Chem B 1997;101:43537.
[38] Zakharov II, Ismagilov ZR, Ruzankin SP, Anufrienko VF, Yashnik SA, Zakharova
OI. Density functional theory molecular cluster study of copper interaction
with nitric oxide dimer in CuZSM-5 catalysts. J Phys Chem C
2007;111:30809.
[39] Zhang J, Ayusawa T, Minagawa M, Kinugawa K, Yamashita H, Matsuoka M,
et al. Investigations of TiO2 photocatalysts for the decomposition of NO in the
ow system: the role of pretreatment and reaction conditions in the
photocatalytic efciency. J Catal 2001;198:18.
[40] Lim TH, Jeong SM, Kim SD, Gyenis J. Photocatalytic decomposition of NO by
TiO2 particles. J Photochem Photobio A: Chem 2000;134:20917.
[41] Bowering N, Walker GS, Harrison PG. Photocatalytic decomposition and
reduction reactions of nitric oxide over Degussa P25. Appl Catal B: Environ
2006;62:20816.

S. Roy et al. / Applied Energy 86 (2009) 22832297


[42] Roy S, Hegde MS, Ravishankar N, Madras G. Creation of redox adsorption sites
by Pd2+ ion substitution in nanoTiO2 for high photocatalytic activity of CO
oxidation, NO reduction, and NO decomposition. J Phys Chem C
2007;111:815360.
[43] Shelef M. Nitric oxide: surface reactions and removal from auto exhaust. Catal
Rev Sci Eng 1975;11:140.
[44] Trovarelli A. Catalysis by ceria and related materials. Imperial College Press;
2002.
[45] Gross H. US patent 3,370,914.
[46] Kummer JT. Catalysts for automobile emission control. Prog Energy Combus
1980;6:17799.
[47] Gandhi HS, Graham GW, McCabe RW. Automotive exhaust catalysis. J Catal
2003;216:43342.
[48] Taylor KC, Schlatter JC. Selective reduction of nitric oxide over noble metals. J
Catal 1980;63:5371.
[49] Van Zon JBAD, Koningsberger DC, Vant Blick HFJ, Sayers DE. An EXAFS study
of the structure of the metalsupport interface in highly dispersed Rh/Al2O3
catalysts. J Chem Phys 1985;82:574255.
[50] Schlatter JC, Taylor KC, Sinkevltch RM. General Motors Research Pub. GMR3112; 1979.
[51] Hecker WC, Bell AT. Reduction of NO by CO over silica-supported rhodium:
infrared and kinetic studies. J Catal 1983;84:20015.
[52] Oh SH, Fisher GB, Carpenter JE, Goodman DW. Comparative kinetic studies of
COO2 and CONO reactions over single crystal and supported rhodium
catalysts. J Catal 1986;100:36076.
[53] Dictor R. An infrared study of the behavior of CO, NO, and CO + NO over Rh/
Al2O3 catalysts. J Catal 1988;109:8999.
[54] Cho BK. Mechanistic importance of intermediate N2O + CO reaction in
overall NO + CO reaction system: I. Kinetic analysis. J Catal 1992;138:
25566.
[55] Cho BK. Mechanistic importance of intermediate N2O + CO reaction in overall
NO + CO reaction system. II. Further analysis and experimental observations. J
Catal 1994;148:697708.
[56] Zhdanov VP. Does the N2OCO subreaction play an important role in the NO
CO reaction on Rh? J Catal 1996;162:1479.
[57] Cho BK. Emphasizing the Mechanistic Importance of Intermediate N2O + CO
reaction in overall NO + CO reaction system. J Catal 1996;162:14951.
[58] Granger P, Malfoy P, Leclercq L, Leclercq G. Kinetics of the CO + N2O reaction
over noble metals: II Rh/Al2O3 and PtRh/Al2O3. J Catal 2004;223:14251.
[59] Granger P, Malfoy P, Esteves P, Paul J-F, Leclercq L, Leclercq G. Europacat IV,
Book of abstracts, Rimini; 1999.
[60] Duprez D, Delahay G, Abderrahim H, Grimblot J. Characterization of Rh/Al2O3
catalysts by gas adsorption and X-ray photoelectron spectroscopy. J Chim
Phys 1986;83:46577.
[61] Dujardin C, Mamede A-S, Payen E, Sombret B, Huvenne J-P, Granger P.
Inuence of the oxidation state of rhodium in three-way catalysts on their
catalytic performances: an in situ FTIR and MS study. Top Catal 2004;30
31:34752.
[62] Granger P, Praliaud H, Billy J, Leclercq L, Leclercq G. An infrared investigation
of the transformation of NO over supported Pt and Rh based three-way
catalysts. Surf Interf Anal 2002;34:926.
[63] Granger P, Dujardin C, Paul JF, Leclercq G. An overview of kinetic and
spectroscopic investigations of three-way catalysts: Mechanistic aspects
of the CO+NO and the CO + N2O reactions. J Mol Catal 2005;228:
24153.
[64] Granger P, Delannoy L, Lecomte JJ, Dathy C, Praliaud H, Leclercq L, et al.
Kinetics of the CO + NO reaction over bimetallic platinumrhodium on
alumina: effect of ceria incorporation into noble metals. J Catal
2002;207:20212.
[65] Gayen A, Priolkar KR, Sarode PR, Jayaram V, Hegde MS, Subbana GN, et al.
Ce1xRhxO2d solid solution formation in combustion-synthesized Rh/CeO2
catalyst studied by XRD, TEM, XPS, and EXAFS. Chem Mater
2004;16:231728.
[66] Mergler YZ, Aalst AV, Delft JV, Nieuwenhuys BE. Promoted Pt catalysts for
automotive pollution control: characterization of Pt/SiO2, Pt/CoOx/SiO2, and
Pt/MnOx/SiO2 catalysts. J Catal 1996;161:3108.
[67] Granger P, Dathy C, Lecomte JJ, Leclercq L, Prignet M, Mabilon G, et al. Kinetics
of the NO and CO reaction over platinum catalysts: I. Inuence of the support.
J Catal 1996;173:30414.
[68] Bera P, Patil KC, Jayaram V, Subbana GN, Hegde MS. Ionic dispersion of Pt and
Pd on CeO2 by combustion method: effect of metalceria interaction on
catalytic activities for NO reduction and CO and hydrocarbon oxidation. J
Catal 2000;196:293301.
[69] Gayen A, Baidya T, Ramesh GS, Srihari R, Hegde MS. Design and fabrication of
an automated temperature programmed reaction system to evaluate 3-way
catalysts Ce1xy(La/Y)x PtyO2d. J Chem Sci 2006;118:4755.
[70] Oh SH, Carpenter JE. Platinumrhodium synergism in three-way automotive
catalysts. J Catal 1986;98:17890.
[71] Meng M, Lin P-Y, Fu Y-L, Yu S-M. The structures, catalytic properties and
spillover effects of the catalysts Co-Pt(Pd, Rh)/Ce-Al-O. In: Guerrero-Ruiz A,
Rodriguez-Ramos I, editors. Studies in surface science and catalysis, vol.
138. Elsevier; 2001. p. 38794.
[72] Gayen A, Baidya T, Biswas K, Roy S, Hegde MS. Synthesis, structure and three
way catalytic activity of Ce1xPtx/2Rhx/2O2d (x = 0. 01 and 0.02) nanocrystallites: synergistic effect in bimetal ionic catalysts. Appl Catal A: Gen
2006;315:13546.

2295

[73] Gaspar AB, Dieguez LC. Dispersion stability and methylcyclopentane


hydrogenolysis in Pd/Al2O3 catalysts. Appl Catal A: Gen 2000;201:24151.
[74] Roy S, Hegde MS. Pd ion substituted CeO2: a superior de-NOx catalyst to Pt or
Rh metal ion doped ceria. Catal Commun 2008;9:8115.
[75] Xu X, Chen P, Goodman DW. A comparative study of the coadsorption of
carbon monoxide and nitric oxide on Pd(1 0 0), Pd(1 1 1), and silica-supported
palladium particles with infrared reectionabsorption spectroscopy. J Phys
Chem 1994;98:92426.
[76] Rainer DR, Koranne M, Vesecky SM, Goodman DW. CO + O2 and CO + NO
reactions over Pd/Al2O3 catalysts. J Phys Chem 1997;101:1076974.
[77] Schmal M, Baldanza MAS, Vannice MA. Pdx Mo/Al2O3 catalysts for NO
reduction by CO. J Catal 1999;185:13851.
[78] Thirunavukkarasu K, Thirumoorthy K, Libuda L, Gopinath CS. A molecular
beam study of the NO + CO reaction on Pd(1 1 1) surfaces. J Phys Chem B
2005;109:1327282.
[79] Mamede A-S, Leclercq G, Payen E, Grimblot G, Granger P. Surface resonance
Raman spectroscopy study of NO transformation over Pd-based catalysts.
Phys Chem Chem Phys 2003;5:44026.
[80] Mamede A-S, Gengembre L, Payen E, Granger P, Leclercq G, Grimblot J. XPS
characterization of adsorbed reaction intermediates on automotive exhaust
gas catalysts. The NO and NO + CO interaction with Pd. Surf Interf Anal
2002;34:10511.
[81] Roy S, Marimuthu A, Hegde MS, Madras G. High rates of NO and N2O
reduction by CO, CO and hydrocarbon oxidation by O2 over nano crystalline
Ce0.98Pd0.02O2d: catalytic and kinetic studies. Appl Catal B: Environ
2007;71:2331.
[82] Roy S, Marimuthu A, Hegde MS, Madras G. High rates of CO and hydrocarbon
oxidation and NO reduction by CO over Ti0.99Pd0.01O1.99. Appl Catal B: Environ
2007;73:30010.
[83] Granger P, Lecomte JJ, Dathy C, Leclercq L, Leclercq G. Kinetics of the CO + NO
reaction over rhodium and platinumrhodium on alumina: II. Effect of Rh
incorporation to Pt. J Catal 2007;175:194203.
[84] Holles JH, Switzer MA, Davis RJ. Inuence of ceria and lanthana promoters on
the kinetics of NO and N2O reduction by CO over alumina-supported
palladium and rhodium. J Catal 2000;190:24760.
[85] Schmal M, Baldanza MAS, Vannice MA. PdxMo/Al2O3 catalysts for NO
reduction by CO. J Catal 1999;185:13851.
[86] Herz RK. Dynamic behavior of automotive catalysts. 1. Catalyst oxidation and
reduction. Ind Eng Chem Prod Res Dev 1981;20:4517.
[87] Balducci G, Kasper J, Fornasiero P, Graziani M, Islam MS. Surface and
reduction energetics of the CeO2ZrO2 catalysts. J Phys Chem B
1998;102:55761.
[88] Fernandez-Garcia M, Martinez-Arias A, Iglesis-Juez A, Hungria AB, Anderson
JA, Conesa JC, et al. Behavior of bimetallic PdCr/Al2O3 and PdCr/(Ce,Zr)Ox/
Al2O3 catalysts for CO and NO elimination. J Catal 2003;196:214. and 22033.
[89] Di Monte R, Kaspar J, Fornasiero P, Graziani M, Paze C, Gubitosa G. NO
reduction by CO over Pd/Ce0.6Zr0.4O2Al2O3 catalysts: in situ FT-IR studies of
NO and CO adsorption. Inorg Chim Acta 2002;334:31826.
[90] Martinez-Arias A, Fernandez-Garcia M, Iglesis-Juez A, Hungria AB, Anderson
JA, Conesa JC, et al. New Pd/CexZr1xO2/Al2O3 three-way catalysts prepared by
microemulsion: Part 2. In situ analysis of CO oxidation and NO reduction
under stoichiometric CO + NO + O2. Appl Catal B: Environ 2001;31:5160.
[91] Dutta G, Waghmare UV, Baidya T, Hegde MS, Priolkar KR, Sarode PR.
Reducibility of Ce1xZrxO2: origin of enhanced oxygen storage capacity. Catal
Lett 2006;108:16572.
[92] Trovarelli A, de Leitenburg C, Dolcettis G. Design better cerium-based
oxidation catalysts. Chemtech 2006;32:114.
[93] Baidya T, Marimuthu A, Hegde MS, Ravishankar N, Madras G. Higher catalytic
activity of nano-Ce1xyTixPdyO2d compared to nano-Ce1xPdxO2d for CO
oxidation and N2O and NO reduction by CO: role of oxide ion vacancy. J Phys
Chem C 2007;111:8309.
[94] Hecker WC, Bell AT. Reduction of NO by H2 over silica-supported rhodium:
infrared and kinetic studies. J Catal 1985;92:24759.
[95] Hirano H, Yamada T, Tanaka KI, Siera J, Cobden P, Nieuwenhuys BE.
Mechanisms of the various nitric oxide reduction reactions on a platinum
rhodium (1 0 0) alloy single crystal surface. Surf Sci 1992;262:97112.
[96] Mergler YZ, Nieuwenhuys BE. NO reduction by H2 over promoted Pt catalysts.
Appl Catal B: Environ 1997;12:95110.
[97] Roy S, Hegde MS, Sharma S, Lalla NP, Marimuthu A, Madras G. Appl Catal B:
Environ 2008;84:34150.
[98] Granger P, Dhainaut F, Pietrzik S, Malfoy P, Mamede A-S, Leclercq L, et al. An
overview: comparative kinetic behaviour of Pt, Rh and Pd in the NO + CO and
NO + H2 reactions. Top Catal 2006;39:6576.
[99] Bera P, Aruna ST, Patil KC, Hegde MS. Studies on Cu/CeO2: a new NO reduction
catalyst. J Catal 1999;186:3644.
[100] Libby WF. Promising catalyst for auto exhaust. Science 1971;171:499500.
[101] Tanaka H, Tan I, Uenishi M, Kimura M, Dohmae K. In: Kruse N, Frennet A,
Bastin JM, editors. Top catal, vols. 16/17. New York: Kluwer Academic
Publishers; 2001. p. 6370.
[102] Sorenson SC, Wronkiewicz JA, Sis LB, Wirtz GP. Properties of LaCoO3
as a catalyst in engine exhaust gases. Bull Am Ceram Soc 1974;53:
44652.
[103] Voorhoeve RJH, Remeika JP, Trimble LE. Perovskites containing ruthenium as
catalysts for nitric oxide reduction mater. Res Bull 1974;9:13938.
[104] Nitadori T, Kurihara S, Misono M. Catalytic properties of La1x A0x MnO3
(A0 = Sr, Ce, Hf). J Catal 1986;98:2218.

2296

S. Roy et al. / Applied Energy 86 (2009) 22832297

[105] Mizuno N, Fujiwara Y, Misono M. Pronounced synergetic effect in the


catalytic properties of LaMn1xCuxO3. J Chem Soc, Chem Commun
1989:3168.
[106] Costa CN, Stathopoulos VN, Belessi VC, Efstathiou AM. An Investigation of the
NO/H2/O2 (lean-deNOx) reaction on a highly active and selective Pt/
La0.5Ce0.5MnO3 catalyst. J Catal 2001;197:3504.
[107] Costa CN, Savva PG, Andronikou C, Lambrou PS, Polychronopoulou K, Belessi
VC, et al. An investigation of the NO/H2/O2 (lean de-NOx) reaction on a highly
active and selective Pt/La0.7Sr0.2Ce0.1FeO3 catalyst at low temperatures. J Catal
2002;209:45671.
[108] Guilhaume N, Primet M. Three-way catalytic activity and oxygen storage
capacity of perovskite LaMn0.976Rh0.024O3+d. J Catal 1997;165:197204.
[109] Nishihata Y, Mizuki J, Akao T, Tanaka H, Uenishi M, Kimura M, et al. Selfregeneration of a Pd-perovskite catalyst for automotive emissions control.
Nature 2002;418:1647.
[110] Twagirashema I, Frere SM, Gengembre L, Dujardin C, Granger P. Structural
regeneration of LaCoO3 perovskite-based catalyst during the NO + H2 + O2
reactions. Top Catal 2002;4243:1716.
[111] Dhainaut F, Pietrzyk S, Granger O. Kinetics of the NO/H2 reaction on Pt/
LaCoO3: a combined theoretical and experimental study. J Catal
2008;258:296305.
[112] Nishihata Y, Mizuki J, Tanaka H, Uenishi M, Kimura M. Self-regeneration of
palladium-perovskite catalysts in modern automobiles. J Phys Chem Solids
2005;66:27482.
[113] Tanaka H, Taniguchi M, Kajita N, Uenishi M, Tan I, Sato N, et al. Top catal, vols.
30/31. New York: Kluwer Academic Publishers; 2004. p. 38996.
[114] Uenishi M, Tanigushi M, Tanaka H, Kimura M, Nishihata Y, Mizuki J, et al. The
reducing capability of palladium segregated from perovskite-type LaFePdOx
automotive catalysts. Appl Catal B: Environ 2005;57:26773.
[115] Twagisrashema I, Engelmann-Pirez M, Frre M, Gengembre L, Burylo L,
Dujardin C, et al. An in situ study of the NO + H2 + (O2) reaction on Pd/LaCoO3
based catalysts. Catal Today 2007;119:1005.
[116] Dacquin JP, Dujardin C, Granger P. Surface reconstruction of supported Pd on
LaCoO3: consequences on the catalytic properties in the decomposition of
N2O. J Catal 2008;253:3749.
[117] Wood SC. Select the right Nox control technology consider the degree of
emission reduction needed, the type of fuel, combustion device design, and
operational factors. Chem Eng Prog 1994;90:328.
[118] Bosch H, Janssen FJ. Preface. Catal Today 1998;2:58.
[119] Cho SM. Properly apply selective catalytic reduction for NOx removal. Chem
Eng Prog 1994;90:3944.
[120] Forzatti P, Lietti L. Recent advances in de-NOxing catalysis for stationary
applications. Heterogen Chem Rev 1996;3:3351.
[121] Alemany LJ, Berti F, Busca G, Ramis G, Robba D, Toledo GP, et al.
Characterization and composition of commercial V2O5WO3TiO2 SCR
catalysts. Appl Catal B: Environ 1996;248:299311.
[122] Went GT, Leu LJ, Bell AT. Quantitative structural analysis of dispersed vanadia
species in TiO2 (anatase)-supported V2O5. J Catal 1992;134:47991.
[123] Orlik SN, Ostapyuk VA, Martsenyuk-Kurkharuk MG. Selective reduction of
nitrogen-oxides with ammonia on V2O5/TiO2 catalysts. Kinet Katal
1995;36:2849.
[124] Oliveri G, Busca G, Lorenzelli V. Structure and surface area evolution of
vanadia-on-titania powders upon heat treatment. Mater Chem Phys
1989;22:51121.
[125] Cristiani C, Bellotto M, Forzatti P, Bregani F. On the morphological properties
of tungstentitania de-NOxing catalysts. J Mater Res 1993;8:201926.
[126] Hums E, Burzlaff H, Rothammel W. Evidence of the interaction between
arsenious oxide and molybdenum oxide in deNOx catalysts by structure
analysis of MoAs2O7. Appl Catal 1991;73:L1924.
[127] Inomata M, Mori K, Miyamoto A, Ui T, Murakami Y. Structures of supported
vanadium oxide catalysts. 1. Vanadium(V) oxide/titanium dioxide (anatase),
vanadium(V) oxide/titanium dioxide (rutile), and vanadium(V) oxide/
titanium dioxide (mixture of anatase with rutile). J Phys Chem
1983;87:75461.
[128] Topsoe NY. Characterization of the nature of surface sites on vanadiatitania
catalysts by FTIR. J Catal 1991;128:499511.
[129] Dines TJ, Rochester CH, Ward AM. Infrared and Raman study of the surface
acidity of titania-supported vanadia catalysts. J Chem Soc Faraday Trans
1991;87:16117.
[130] Hu S, Harris SP. 15N NMR study of the adsorption of NO and NH3 on titaniasupported vanadia catalysts. J Catal 1996;158:199204.
[131] Busca G, Saussey H, Saur O, Lavalley JC, Lorenzelli V. FT-IR characterization of
the surface acidity of different titanium dioxide anatase preparations. Appl
Catal 1985;14:24560.
[132] Ramis G, Busca G, Bregani F, Forzatti P. Fourier transform-infrared study of
the adsorption and coadsorption of nitric oxide, nitrogen dioxide and
ammonia on vanadiatitania and mechanism of selective catalytic
reduction. Appl Catal 1990;64:25978.
[133] Wong WC, Nobe K. Kinetics of nitric oxide reduction with ammonia on
chemical mixed and impregnated vanadium(V) oxide-titanium(IV) oxide
catalysts. Ind Eng Chem Prod Res Dev 1984;23:5648.
[134] Nam I, Eldridge JW, Kittrell IR. Model of temperature dependence of a
vanadiaalumina catalyst for nitric oxide reduction by ammonia: fresh
catalyst. Ind Eng Chem Prod Res Dev 1986;25:18692.
[135] Beekman JW, Hegedus LL. Design of monolith catalysts for power plant
nitrogen oxide (NO) emission control. Ind Eng Chem Res 1991;30:96978.

[136] Pinoy LJ, Hosten LH. Experimental and kinetic modelling study of DeNOx on
an industrial V2O5WO3/TiO2 catalyst. Catal Today 1993;17:1518.
[137] Takagi M, Kowai T, Soma M, Onishi I, Tamaru K. The mechanism of the
reaction between NOx and NH3 on V2O5 in the presence of oxygen. J Catal
1977;50:4416.
[138] Inomata M, Miyamoto A, Murakami Y. Mechanism of the reaction of NO and
NH3 on vanadium oxide catalyst in the presence of oxygen under the dilute
gas condition. J Catal 1980;62:1408.
[139] Miyamoto A, Kobayashi K, Inomata M, Murakami Y. Nitrogen-15 tracer
investigation of the mechanism of the reaction of nitric oxide with ammonia
on vanadium oxide catalysts. J Phys Chem 1982;86:294550.
[140] Inomata M, Miyamoto A, Ui T, Kobayashi K, Murakami Y. Activities of
vanadium pentoxide/titanium dioxide and vanadium pentoxide/aluminum
oxide catalysts for the reaction of nitric oxide and ammonia in the presence
of oxygen. Ind Eng Chem Prod Res Dev 1982;21:4248.
[141] Janssen F, Van den Kerkhof F, Bosch H, Ross JJ. Infrared and x-ray
photoelectron spectroscopy study of carbon monoxide and carbon dioxide
on platinum/ceria. Phys Chem 1987;91:59317.
[142] Wolf M, Yang DL, Durant JL. A Comprehensive Study of the Reaction
NH2 + NO ? products: reaction rate coefcients, product branching
fractions, and ab initio calculations. J Phys Chem A 1997;101:624351.
[143] Marcy TP, Heard DE, Leone SR. Product Studies of Inelastic and Reactive
Collisions of NH2 + NO: effects of vibrationally and electronically excited NH2.
J Phys Chem A 2002;106:824955.
[144] Duan X, Page M. Theoretical characterization of structures and vibrational
frequencies for intermediates and transition states in the reaction of NH2
with NO. J Mol Struct (Theochem) 1995;333:23342.
[145] Soyer S, Uzun A, Senkan S, Onal I. A quantum chemical study of nitric oxide
reduction by ammonia (SCR reaction) on V2O5 catalyst surface. Catal Today
2006;118:26878.
[146] von Barth U. Basic density-functional theory an overview. Phys Scripta
2004;T109:93942.
[147] Kameoka S, Ukisu Y, Miyadera T. Selective catalytic reduction of NOx
with CH3OH, C2H5OH and C3H6 in the presence of O2 over Ag/Al2O3
catalyst: role of surface nitrate species. Phys Chem Chem Phys
2000;2:36774.
[148] Shimizu K, Shibata J, Satsuma A, Hattori T. Mechanistic causes of the
hydrocarbon effect on the activity of AgAl2O3 catalyst for the selective
reduction of NO. Phys Chem Chem Phys 2001;3:8805.
[149] Underwood GM, Miller TM, Grassian VH. Transmission FT-IR and Knudsen
cell study of the heterogeneous reactivity of gaseous nitrogen dioxide on
mineral oxide particles. J Phys Chem A 1982;103:61849.
[150] Zhang X, He H, Gao H, Yu Y. Experimental and theoretical studies of surface
nitrate species on Ag/Al2O3 using DRIFTS and DFT. Spectrochim Acta Part A:
Mol Biomol Spectrosc 2008;71:144651.
[151] Meunier FC, Zuzaniuk V, Breen JP, Olsson M, Ross JRH. Mechanistic
differences in the selective reduction of NO by propene over cobalt- and
silver-promoted alumina catalysts: kinetic and in situ DRIFTS study. Catal
Today 2000;59:28791.
[152] He H, Yu Y. Selective catalytic reduction of NOx over Ag/Al2O3 catalyst: from
reaction mechanism to diesel engine test. Catal Today 2005;100:3747.
[153] Kapteijn F, Singoredjo L, Dekke NJJ, Moulijn JA. Kinetics of the selective
catalytic reduction of nitrogen oxide (NO) with ammonia over manganese
oxide (Mn2O3)tungsten oxide (WO3)/c-alumina. Ind Eng Chem Res
1993;32:44552.
[154] Kapteijn F, Singoredjo L, van Driel M, Andreini A, Moulijn JA, Ramis G, et al.
Alumina-supported manganese oxide catalysts: II. Surface characterization
and adsorption of ammonia and nitric oxide. J Catal 1994;150:10516.
[155] Singoredjo L, Korver R, Kapteijn F, Moulijn J. Alumina supported manganese
oxides for the low-temperature selective catalytic reduction of nitric oxide
with ammonia. Appl Catal B: Environ 1992;1:297316.
[156] Kapteijn F, Singoredjo L, Andreini A, Moulijn JA. Activity and selectivity of
pure manganese oxides in the selective catalytic reduction of nitric oxide
with ammonia. Appl Catal B: Environ 1994;3:17389.
[157] Pena DA, Uphade BS, Reddy EP, Smirniotis PG. Identication of Surface
species on titania-supported manganese, chromium, and copper oxide lowtemperature SCR catalysts. J Phys Chem B 2004;108:992736.
[158] Roy S, Viswanath B, Hegde MS, Madras G. Low-temperature selective
catalytic reduction of NO with NH3 over Ti0.9M0.1O2d (M = Cr, Mn, Fe, Co,
Cu). J Phys Chem C 2008;112:600212.
[159] Nova I, Lietti L, Tronconi E, Forzatti P. Dynamics of SCR reaction over a TiO2supported vanadiatungsta commercial catalyst. Catal Today 2000;60:7382.
[160] Nova I, Lietti L, Tronconi E, Forzatti P. Transient response method applied to
the kinetic analysis of the deNOx-SCR reaction. Chem Eng Sci
2001;56:122937.
[161] Adelmana BJ, Sachtler WMH. The effect of zeolitic protons on NOx reduction
over Pd/ZSM-5 catalysts. Appl Catal B: Environ 1997;14:111.
[162] Apostolescu N, Geiger B, Hizbullah K, Jan MT, Kureti S, Reichert D, et al.
Selective catalytic reduction of nitrogen oxides by ammonia on iron oxide
catalysts. Appl Catal B: Environ 2006;62:10414.
[163] Chmielarz L, Kustrowski P, Zbroja M, asocha W, Dziembaj R. Selective
reduction of NO with NH3 over pillared clays modied with transition metals.
Catal Today 2004;90:439.
[164] Willey RJ, Lai H, Peri JB. Investigation of iron oxide-chromia-alumina aerogels
for the selective catalytic reduction of nitric oxide by ammonia. J Catal
1991;130:31931.

S. Roy et al. / Applied Energy 86 (2009) 22832297


[165] Long RQ, Yang RT. Selective catalytic reduction of NO with ammonia over
mordenite
(Fe-MOR):
catalytic
performance,
Fe3+-exchanged
characterization, and mechanistic study. J Catal 2002;207:27485.
[166] Salker AV, Weisweiler W. Catalytic behaviour of metal based ZSM-5 catalysts
for NOx reduction with NH3 in dry and humid conditions. Appl Catal A: Gen
2000;2:2219.
[167] Krishna K, Seijger GBF, Van den Bleek CM, Makkee M, Mul G, Calis HPA.
Selective catalytic reduction of NO with NH3 over FeZSM-5 catalysts
prepared by sublimation of FeCl3 at different temperatures. Catal Lett
2003;86:12132.
[168] Chen H, Sun Q, Wen B, Yeom Y, Weitz E, Sachtler WMH. Reduction over
zeolite-based catalysts of nitrogen oxides in emissions containing excess
oxygen: unraveling the reaction mechanism. Catal Today 2004;96:110.
[169] Krocher O, Devadas M, Elsener M, Wokaun A, Soger N, Pfeifer M, et al.
Investigation of the selective catalytic reduction of NO by NH3 on FeZSM5
monolith catalysts. Appl Catal B: Environ 2006;66:20816.
[170] Rahkamaa-Tolonen K, Maunula T, Lomma M, Huuhtanen M, Keiski RL. The
effect of NO2 on the activity of fresh and aged zeolite catalysts in the NH3-SCR
reaction. Catal Today 2005;100:21722.
[171] Qi G, Yang RT. Low-temperature SCR of NO with NH3 over noble metal
promoted FeZSM-5 catalysts. Catal Lett 2005;100:2436.
[172] Long RQ, Yang RT. Selective catalytic oxidation (SCO) of ammonia to nitrogen
over Fe-exchanged zeolites. J Catal 2001;201:14552.
[173] Kim H, Hwang U-C, Nam I-S, Kim YG. The characteristics of a copperexchanged natural zeolite for NO reduction by NH3 and C3H6. Catal Today
1998;44:5765.
[174] Katranas TK, Vlessidis AG, Tsiatouras VA, Triantafyllidis KS, Evmiridis NP.
Dehydrogenation of propane over natural clinoptilolite zeolites. Micropor
Mesopor Mater 2003;61:18998.
[175] Ates A. Characteristics of Fe-exchanged natural zeolites for the
decomposition of N2O and its selective catalytic reduction with NH3. Appl
Catal B: Environ 2007;76:28290.
[176] Lombardo EA, Sill GA, dItri JL, Hall WK. The Possible Role of Nitromethane in
the SCR of NOx with CH4 over MZSM5 (M = Co, H, Fe, Cu). J Catal
1998;173:4409.
[177] Desai AJ, Kovalchuk VI, Lombardo EA, dItri JL. CoZSM-5: why this catalyst
selectively reduces NOx with methane. J Catal 1999;184:396405.
[178] Burch R, Breen JP, Meunier FC. A review of the selective reduction of NOx with
hydrocarbons under lean-burn conditions with non-zeolite oxide and
platinum group metal catalysts. Appl Catal B: Environ 2002;39:283303.
[179] Parvulescu VI, Granger P, Delmon B. Catalytic removal of NO. Catal Today
1998;46:233316.
[180] Miyadera T. Alumina-supported silver catalysts for the selective reduction of
nitric oxide with propene and oxygen-containing organic compounds
alumina-supported catalysts for the selective reduction of nitric oxide by
propene. Appl Catal B: Environ 1993;2:199205.
[181] Hoost TE, Kudla RJ, Collins KM, Chattha MS. Characterization of Ag/c-Al2O3
catalysts and their lean-NOx properties. Appl Catal B: Environ
1997;13:5967.
[182] Martinez-Arias A, Fernandez-Garcia M, Iglesias-Juez A, Anderson JA, Conesa
JC, Soria J. Study of the lean NO(x) reduction with C3H6 in the presence of
water over silver/alumina catalysts prepared from inverse microemulsions.
Appl Catal B: Environ 2000;28:2941.
[183] Klingstedt F, Eranen K, Lindfors L, Andersson S, Cider L, Landberg C, et al. A
highly active Ag/alumina catalytic converter for continuous HC-SCR during
lean-burn conditions: From laboratory to full-scale vehicle tests. Top Catal
2004;30/31:2730.
[184] Arve K, Capek L, Klingstedt F, Eranen K, Lindfors L, Murzin DY, et al.
Preparation and characterisation of Ag/alumina catalysts for the removal of
NOx emissions under oxygen rich conditions. Top Catal 2004;30/31:916.
[185] Eranen K, Lindfors L, Klingstedt F, Murzin DY. Continuous reduction of NO
with octane over a silver/alumina catalyst in oxygen-rich exhaust gases:
Combined heterogeneous and surface-mediated homogeneous reactions. J
Catal 2003;219:2540.
[186] Shibata J, Takada Y, Shichi A, Satokawa S, Satsuma A, Hattori T. Ag cluster as
active species for SCR of NO by propane in the presence of hydrogen over Ag
MFI. J Catal 2004;222:36876.
[187] Satokawa S. Enhancing the NO/C3H8/O2 reaction by using H2 over Ag/Al2O3
catalysts under lean-exhaust conditions. Chem Lett 2000;3:2945.
[188] Shibata J, Takada Y, Shichi A, Satokawa S, Satsuma A, Hattori T. Inuence of
zeolite support on activity enhancement by addition of hydrogen for SCR of
NO by propane over Agzeolites. Appl Catal B: Environ 2004;54:13744.
[189] Richter M, Fricke R, Eckelt R. Unusual activity enhancement of NO conversion
over Ag/Al2O3 by using a mixed NH3/H2 reductant under lean conditions.
Catal Lett 2004;94:1158.
[190] Burch R, Breen JP, Hill CJ, Krutzsch B, Konrad B, Jobson E, et al. Exceptional
activity for NOx reduction at low temperatures using combinations of
hydrogen and higher hydrocarbons on Ag/Al2O3 catalysts. Top Catal
2004;3031:1925.
[191] Shibata J, Shimizu K, Takada Y, Shichi A, Yoshida H, Satokawa S, et al.
Structure of active Ag clusters in Agzeolites for SCR of NO by propane in the
presence of hydrogen. J Catal 2004;227:36774.

2297

[192] Shi C, Cheng M, Qu Z, Yang X, Bao X. On the selectively catalytic reduction of


NOx with methane over AgZSM-5 catalysts. Appl Catal B: Environ
2002;36:17382.
[193] Konova P, Arve K, Klingstedt F, Nikolov P, Naydenov A, Kumar N, et al. A
combination of Ag/alumina and Ag modied ZSM-5 to remove NOx and CO
during lean conditions. Appl Catal B: Environ 2007;70:13845.
[194] Meunier FC, Ross JRH. Effect of ex situ treatments with SO2 on the activity of a
low loading silveralumina catalyst for the selective reduction of NO and NO2
by propene. Appl Catal B: Environ 2000;24:2332.
[195] Abe A, Aoyama N, Sumiya S, Kakuta N, Yoshida K. Effect of SO2 on NOx
reduction by ethanol over Ag/Al2O3 catalyst. Catal Lett 1998;51:59.
[196] Angelidis TN, Kruse N. Promotional effect of SO2 on the selective catalytic
reduction of NOx with propane/propene over Ag/c-Al2O3. Appl Catal B:
Environ 2001;34:20112.
[197] Burch R, Sullivan JA, Watling TC. Mechanistic considerations for the reduction
of NOx over Pt/Al2O3 and Al2O3 catalysts under lean-burn conditions. Catal
Today 1998;42:1323.
[198] Angelidis TN, Christoforou S, Bongiovanni A, Kruse N. On the promotion by
SO2 of the SCR process over Ag/Al2O3: Inuence of SO2 concentration with
C3H6 versus C3H8 as reductant. Appl Catal B: Environ 2002;39:197204.
[199] Breen JP, Burch R, Hardacre C, Hill CJ, Krutzsch B, Bandl-Konrad B, et al. An
investigation of the thermal stability and sulphur tolerance of Ag/c-Al2O3
catalysts for the SCR of NOx with hydrocarbons and hydrogen. Appl Catal B:
Environ 2007;70:3644.
[200] Gabrielsson PLT. Urea-SCR in automotive applications. Top Catal
2004;28:17784.
[201] Koebel M, Elsener M, Kleeman M. Urea-SCR: a promising technique to reduce
NOx emissions from automotive diesel engines. Catal Today 2000;59:33545.
[202] Gomez-Garcia MA, Pitchon V, Kiennemann A. Pollution by nitrogen oxides:
an approach to NOx abatement by using sorbing catalytic materials. Environ
Int 2005;31:44567.
[203] Takahashi N, Shinjoh H, Iijima T, Suzuki T, Yamazaki K, Yokota K, et al. The
new concept 3-way catalyst for automotive lean-burn engine: NOx storage
and reduction catalyst. Catal Today 1996;27:639.
[204] Matsumoto S. DeNOx catalyst for automotive lean-burn engine. Catal Today
1996;29:435.
[205] Takeuchi M, Matsumoto S. NOx storage-reduction catalysts for gasoline
engines. Top Catal 2004;28:1518.
[206] Casapu M, Grunwaldt J-D, Maciejewski M, Wittrock M, Gbel U, Baiker A.
Formation and stability of barium aluminate and cerate in NOx storagereduction catalysts. Appl Catal B: Environ 2006;63:23242.
[207] Casapu M, Grunwaldt J-D, Maciejewski M, Krumeich F, Baiker A, Wittrock M,
et al. Comparative study of structural properties and NOx storage-reduction
behavior of Pt/Ba/CeO2 and Pt/Ba/Al2O3. Appl Catal B: Environ
2008;78:288300.
[208] Sjoblom J, Papadakis K, Creaser D, Odenbrand CUI. Use of experimental
design in development of a catalyst system. Catal Today 2005;100:2438.
[209] Sakamoto Y, Motohiro T, Matsunaga S, Okumura K, Kayama T, Yamazaki K,
et al. Transient analysis of the release and reduction of NOx using a Pt/Ba/
Al2O3 catalyst. Catal Today 2007;121:21725.
[210] Malpartida I, Guerrero-Perez MO, Herrera MC, Larrubia MA, Alemany LJ. MS
FTIR reduction stage study of NSR catalysts. Catal Today 2007;126:1628.
[211] Takahashi N, Suda A, Hachisuka I, Sugiura M, Sobukawa H, Shinjoh H. Sulfur
durability of NOx storage and reduction catalyst with supports of TiO2, ZrO2
and ZrO2TiO2 mixed oxides. Appl Catal B: Environ 2007;72:18795.
[212] Bgner W, Krmer M, Krutzsch B, Piscinger S, Voigtlnder D, Wenniger G.
Removal of nitrogen oxides from the exhaust of a lean-tune gasoline engine.
Appl Catal B: Environ 1995;7:15371.
[213] Hodjati S, Vaezzadeh K, Petit C, Pitchon V, Kiennemann A. NOx sorption
desorption study: application to diesel and lean-burn exhaust gas (selective
NOx recirculation technique). Catal Today 2000;59:323334;.
[214] Trichard JM. Current tasks and challenges for exhaust after-treatment
research: an industrial viewpoint. Stud Surf Sci Catal 2007;171:21133.
[215] Caton JA, Zhiyong X. The selective non-catalytic removal (SNCR) of nitric
oxides from engine exhaust streams: Comparison of three processes. J Eng
Gas Turb Power 2004;126:23440.
[216] Javed MT, Irfan N, Gibbs BM. Control of combustion-generated nitrogen oxides
by selective non-catalytic reduction. J Environ Manage 2007;83:25189.
[217] Bae SW, Roh SA, Kim SD. NO removal by reducing agents and additives in the
selective
non-catalytic
reduction
(SNCR)
process.
Chemosphere
2006;65:1705.
[218] Fua Y, Diwekar UM. Cost effective environmental control technology for
utilities. Adv Environ Res 2004;8:17396.
[219] Mok YS, Lee Heon-Ju. Removal of sulfur dioxide and nitrogen oxides by using
ozone injection and absorptionreduction technique. Fuel Process Technol
2006;87:5917.
[220] Wang Z, Zhou J, Fan J, Cen K. Direct numerical simulation of ozone injection
technology for NOx control in ue gas. Energy Fuels 2006;20:24328.
[221] Mok YS. Oxidation of NO to NO2 using the ozonization method for the
improvement of selective catalytic reduction. Chem Eng Jpn 2004;37:133744.
[222] Mok YS, Shin DN, Koh DJ, Kim KT. Effect of ozone injection on selective
catalytic reduction of nitrogen oxides. Appl Chem 2005;9:21720.

Vous aimerez peut-être aussi