Anda di halaman 1dari 23

Environment International 31 (2005) 445 467

www.elsevier.com/locate/envint

Review article

Pollution by nitrogen oxides: an approach to NOx abatement by using


sorbing catalytic materials
M.A. Gomez-Garca1, V. Pitchon, A. Kiennemann*
LMSPC, Laboratoire de Materiaux, Surfaces et Procedes pour la Catalyse, UMR 7515 du CNRS-ECPM-ULP, 25 Rue Becquerel,
67087 Strasbourg Cedex 2, France
Received 29 April 2004; accepted 8 September 2004
Available online 11 November 2004

Abstract
This article summarises the abatement of NOx pollution by using sorbing catalytic materials with especial relevance to the challenge
presented in fixed installations sources. A general vision of the origins of the different pollutants, with emphasis on nitrogen oxides
formation, is presented as introduction. The impact of NOx pollution comprises additionally a quick view of its toxicity and environmental
effects. Actual solutions are presented especially the case of the selective catalytic reduction (SCR) process with its advantages and
difficulties. The new concepts for NOx abatement are also analysed. In such a way, updated information on solid sorbents for NOx removal is
provided by including metal oxides, spinelles, perovskites, double-layered cuprates, zeolites, carbonaceous materials, heteropolyacids
(HPAs), and supported heteropolyacids. The possibility of reducing those sorbed NOx is also underlined. Sorption mechanisms are analysed
and clarified by emphasising convergence and disagreement points.
D 2004 Elsevier Ltd. All rights reserved.
Keywords: Nitrogen oxide; Selective catalytic reduction; Double-layered cuprate

Contents
1.

Introduction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .
1.1. Combustion background . . . . . . . . . . . . . . . . . . . .
1.2. The pollution problem . . . . . . . . . . . . . . . . . . . . .
1.2.1. Nitrogen oxides . . . . . . . . . . . . . . . . . . . .
1.3. The impact of NOx . . . . . . . . . . . . . . . . . . . . . . .
1.3.1. Toxicity . . . . . . . . . . . . . . . . . . . . . . . .
1.3.2. Environmental impact . . . . . . . . . . . . . . . . .
1.4. Emission regulations . . . . . . . . . . . . . . . . . . . . . .
1.5. Actual approaches to NOx abatement. . . . . . . . . . . . . .
1.5.1. Selective catalytic reduction (SCR) of NOx . . . . . .
1.5.2. Conclusions on bactual approaches to NOx abatementQ
1.6. New techniques for NOx abatement . . . . . . . . . . . . . .
1.6.1. Selective NOx recirculation (SNR). . . . . . . . . . .
1.6.2. NOx storage and reduction (NSR) . . . . . . . . . . .
1.6.3. Regeneration of NOx sorbents . . . . . . . . . . . . .
1.6.4. Conclusions on bnew techniques for NOx abatementQ .
* Corresponding author. Tel.: +33 390 242766; fax: +33 390 242768.
E-mail address: kiennemann@chimie.u-strasbg.fr (A. Kiennemann).
1
On leave from Universidad Nacional de Colombia, Sede Manizales.

0160-4120/$ - see front matter D 2004 Elsevier Ltd. All rights reserved.
doi:10.1016/j.envint.2004.09.006

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

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

446
447
447
447
449
449
449
449
450
451
452
452
452
452
452
453

446

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

1.7.

NOx -sorbing catalyst materials . . . . . . . . . . . . . . .


1.7.1. Metal oxides . . . . . . . . . . . . . . . . . . . .
1.7.2. Spinels (AB2O4) . . . . . . . . . . . . . . . . . .
1.7.3. Perovskites (ABO3) . . . . . . . . . . . . . . . .
1.7.4. Double-layered cuprates, La2x Bax SrCu2O6 . . . .
1.7.5. Zeolites . . . . . . . . . . . . . . . . . . . . . .
1.7.6. Carbonaceous materials . . . . . . . . . . . . . .
1.7.7. Heteropolyacids (HPAs) . . . . . . . . . . . . . .
1.7.8. Conclusions on bNOx -sorbing catalytic materialsQ .
1.8. Conclusions. . . . . . . . . . . . . . . . . . . . . . . . .
References. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

.
.
.
.
.
.
.
.
.
.
.

1. Introduction

exhaust in front of the catalyst was tried. Although this


provided substantially greater levels of NOx reduction in
laboratory experiments, the consensus is that it is not yet
a technically available approach. An alternative to the
catalytic approach is based on NOx adsorption. In it, NOx
is stored in the catalyst under lean conditions and the
NOx trap is regenerated for a short period under fuel-rich
conditions. This technique has showed good applicability
in Japan and Sweden, but, in the rest of Europe, there is
still an issue with NOx traps because sulphur oxides
poison them and fuels still contain too high level of
sulphur to allow these NOx traps to be used for long
periods. As it will be discussed later in this chapter, in
spite of the intense research, these are only partial
solutions resulting in serious drawbacks.
In the case of emissions coming from stationary
sources, the selective catalytic NOx reduction by NH3
is currently the most used method. Among the particular
advantages to use NH3 as reducing agent is the high
selectivity of NH3 reaction with NO in the presence of
oxygen and the promotional effect of oxygen on the
kinetics of this reaction. The most significant technical
parameters are the positioning of catalyst SCR compared
to the source of gas and the design of ammonia
injection. The ammonia must be perfectly distributed to
ensure an adequate value of the NH3/NOx ratio (NOx
elimination depends on this ratio). However, the use of
ammonia generates many problems due to leak (the
ammonia emissions are also the object of regulations as
NOx ) and to the difficulties of transport and storage. A
future technology will have imperatively to take into
account ammonia not converted in reactions SCR
(Forzatti, 2000).
The problem of NOx pollution is introduced here by
including its types, sources and emission levels; their
influence on the environment and public health, their
regulation, and the approaches for their removal are also
reviewed. Updated information on solid sorbents for NOx
removal is provided by including metal oxides, spinelles,
perovskites, double-layered cuprates, zeolites, carbonaceous
materials, and heteropolyacids (HPAs). Sorption mechanisms are analysed and clarified by emphasising convergence and disagreement points.

The combustion of fossil fuels to meet the society


requirements of energy discharge large quantities of
pollutants to the environment. Among these contaminants
are the nitrogen oxides (NOx ) which are the source of severe
environmental problems such as acid rain, smog formation,
global warming, and ozone layer weakening. For the past 30
years, great efforts have been made in research directed to
find solutions for the NOx problem. Many NOx -controlling
technologies have been developed, but these are still not
enough to meet the more stringent regulation in future. For
example, in France, NOx emissions estimate for 2001 were
about 1378 kt showing a diminution tendency principally
due to the development of new catalytic techniques.
However, NOx emissions are still 37% higher than the
target fixed for 2010 by the Gothenburg protocol (CITEPA,
2002). To face these requests, emissions control technologies have to be improved constantly. Currently, two main
methods for the removal of NOx from emission gases are
employed:
(i)
(ii)

The three-way catalyst (TWC) developed for mobil


sources that use gasoline.
The selective catalytic reduction of NOx with NH3
(SCR), which is applied for stationary sources such as
power plants.

The use of TWC is an established technology for the


catalytic reduction of NOx produced by gasoline engines
operating at stoichiometric conditions (not discussed
here). However, there is still no appropriate catalytic
technology for NOx emission abatement in vehicles with
diesel and lean-burn gasoline engines. The main difficulties with potential catalysts are their narrow temperature
window of operation and their poisoning by sulphur
oxides (Shelef and McCabe, 2000). The development of
fuel-efficient lean-burn engines has resulted in major new
challenges for emission control systems. Direct NOx
reduction under lean-burn conditions was first attempted
by the residual hydrocarbons in the exhaust. When the
use of residual hydrocarbons proved unsuccessful, the
direct addition of a hydrocarbon reductant into the

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

.
.
.
.
.
.
.
.
.
.
.

454
454
457
457
458
459
460
461
463
464
464

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

1.1. Combustion background


The major source of nitrogen oxides are the combustion
processes such as coal-, oil- and gas-fired power plants
including turbines, engines, waste incinerators, refineries
and gasifiers (stationary sources), and the engines of vehicles
and aeroplanes. In both cases, combustion processes are so
fast that the thermodynamic equilibrium to CO2 and H2O is
not reached. The amount of oxidant (oxygen or air) just
sufficient to burn the carbon, hydrogen, and sulphur in a fuel
to carbon dioxide, water vapour and sulphur dioxide is the
theoretical or stoichiometric oxygen or air requirement. The
chemical equation for complete combustion of a fuel is:
Cx Hy Oz Sw 4x y  2z 4w=4O2 ! CO2
y=2H2 O wSO2
where x, y, z, and w are stoichiometric coefficients for
carbon, hydrogen, oxygen and sulphur in the fuel,
respectively. For example, 1 mol of octane requires 12.5
mol of oxygen for complete combustion to 8 mol of carbon
dioxide and 9 mol of water. In practice, more than the
theoretical amount of air is necessary to achieve complete
combustion. This excess air is expressed as a weight
percentage of the theoretical air amount. The equivalence
ratio is defined as the ratio of the actual air/fuel ratio to the
stoichiometric air/fuel ratio. Equivalence ratio values less
than 1.0 corresponds to fuel-rich mixtures (or conditions).
Conversely, values greater than 1.0 correspond to fuel-lean
mixtures. Hence, in the example presented above, for the
combustion of octane the stoichiometric air/fuel ratio is
equal to 15.13 kgair kgoctane1.
1.2. The pollution problem
Key combustion-generated air pollutants include sulphur
oxides (principally SO2), particulate matter, carbon monoxide, unburned hydrocarbons, and nitrogen oxides (NOx ).
Table 1 shows typical concentrations range of emissions in
the flue gases from efficient electric power generation plants
as a function of the fuel utilised (Ramachandran et al., 2000).
Natural gas, for example, is a relatively clean fuel with
respect to NOx formation when compared with heavy fuel or
coal. Natural gas does not contain any fuel nitrogen, i.e.,
nitrogen atoms bound in the hydrocarbon molecules, and,

Table 1
Typical concentrations and ranges of emissions in flue gases from power
generation plants (Ramachandran et al., 2000)
NOx (ppm)
SOx (ppm)
CO2 (%)
O2 (%)
H2O (%)
N2
a

Natural gas

Fuel oila

Coal

25160
b0.520
512
318
819
Balance

100600
2002000
1214
25
912
Balance

1501000
2002000
1015
35
710
Balance

Dependent on fuel oil, e.g., No. 6 fuel oil.

447

therefore, the only NOx production route is via oxidation of


the molecular nitrogen contained in the combustion air.
Sulphur occurs in fuels as inorganic minerals (primarily
pyrite, FeS2), organic structures, sulphate salts, and elemental sulphur. Sulphur contents range from parts per
million (ppm) in pipeline natural gas, to a few tenths of a
percent in diesel and light fuel oils, to 0.5% to 5% in heavy
fuel oils and coals. Sulphur compounds are pyrolized during
the volatilisation phase of oil and coal combustion and react
in the gas phase to form predominantly SO2 and some SO3.
Conversion of fuel sulphur to these oxides is generally high
(85% to 90%) and is relatively independent of the
combustion conditions. From 1% to 4% of the SO2 are
further oxidised to SO3; the latter is highly reactive and
extremely hygroscopic. It combines with water to form
sulphuric acid aerosol. Combustion-related particulate
emissions may consist of one or more of the following
types, depending on the fuel: mineral matter derived from
ash constituents of liquid and solid fuels can vaporise and
condense as submicronsize aerosols, sulphate particles
formed in the gas phase which can condense and unburned
carbon includes unburned char (the remaining solid particle), coke, and soot (occurs as fine particles (0.02 to 0.2
Am), often agglomerated into filaments or chains which can
be several millimetres long). Carbon monoxide is a key
intermediate in the oxidation of all hydrocarbons. In a welladjusted combustion system, essentially all the CO is
oxidised to CO2 and final emission of CO is very low
indeed (a few parts per million). However, in systems which
have low temperature or which are in poor adjustment, CO
emissions can be significant. Additionally, various unburned
hydrocarbon species may be emitted from hydrocarbon
flames. In general, there are two classes of unburned
hydrocarbons: (1) small molecules that give the intermediate
products of combustion (for example, formaldehyde) and
(2) larger molecules that are formed by pyro-synthesis in
hot, fuel-rich zones within flames, e.g., benzene, toluene,
xylene, and various polycyclic aromatic hydrocarbons.
1.2.1. Nitrogen oxides
Several types of nitrogen oxides with diverse and well
known physical and chemical properties exist. However,
their characterisation as adsorbed species on different
surfaces has no unified interpretation (Shriver and Atkins,
1999; Hadjiivanov, 2000). Among these oxides, the
components of the polluting emission gases formed by the
combustion of biomass and fossil fuels are mainly nitric
oxide (NO) and nitrogen dioxide (NO2) which are collectively named NOx . NO is the primary form in combustion
products (typically 95% of total NOx ; Bosch and Janssen,
1988; Fritz and Pitchon, 1997; Janssen, 1999). NOx is
produced not only by burning of fuels and biomass but also
by lightening, oxidation of NH3 (produced by microbial
decomposition of proteins in the soil), and volcanic activity.
Fig. 1 shows the sources of global NOx emissions and their
relative contributions (Schnelle and Brown, 2002). The

448

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

The proposed mechanism involves a chain reaction of O*


and N* activated atoms (Eqs. (1) and (2)). It was found that
the quantity of NO produced in a combustion process is
related to the quantity of N2 and O2 in the combustion
products and to the heat of combustion, but it is not related
to the nature of the fuel. The rate of NO production by the
Zeldovich mechanism is given with good accuracy by:
dNO
2kO*N2 :
dt

Fig. 1. Relative contributions of anthropogenic NOx emission sources


(Schnelle and Brown, 2002).

large amount of NOx generated at coal-fired electric power


plants is evident and the very large contribution from motor
vehicles and other forms of transportation is pronounced.
Anthropogenic activity (such as burning of fuels by power
plants and vehicles) is the main source of nitrogen oxides
emissions while endogenic are only a small fraction.
According to modern estimates 21 million ton of NOx are
emitted per year in the U.S. alone and 95% of it is due to
power sources and vehicles (Armor, 1997).
Another nitrogen oxide that is relevant due to its
important role in atmospheric pollution is N2O. It is
produced mainly in the nature by microbial activity.
However, its level in the atmosphere continues to increase
mainly due to anthropogenic activities (Li and Amor,
1993a,b). The human contribution to the release of N2O
to the atmosphere is around 4.7 to 7 million ton per year
(3040% of the total emissions). This includes activities
such as adipic acid production for Nylon 6,6, nitric acid
manufacture, fossil fuels and biomass combustion, and the
use of fertilizers. N2O is converted into N2 and NO in the
stratosphere and it contributes to the destruction of the
earths protective ozone layer and to the green house effect
(Kapteijn et al., 1996).
Three reaction paths, each having unique characteristics,
are responsible for the formation of NOx during combustion
processes: (1) thermal NOx , which is formed by the
combination of atmospheric nitrogen and oxygen at high
temperatures; (2) fuel NOx , which is formed from the
oxidation of fuel-bound nitrogen (FBN); and (3) prompt
NOx , which is formed by the reaction of fuel-derived
hydrocarbon fragments with atmospheric nitrogen (Perry et
al., 1997; Schnelle and Brown, 2002).

A value of k=1.8 1011exp(38370/T) m3 kmol1 s1 (T


in K) has been suggested by Miller and Bowman (1989). As
indicated, the rate of NO formation increases exponentially
with temperature and, of course, oxygen, and nitrogen must
be available for thermal NOx to form. Thus, thermal NOx
formation is rapid in high-temperature lean zones of flames.
Eq. (3) shows that the formation of NO is essentially
controlled by reaction (1). It also reveals the importance of
both temperature and atomic oxygen concentration. The
Zeldovich radical chain mechanism dominates NO formation under most engine conditions. However, other pathways (involving N2O intermediates and fuel radicals such as
CH) can be significant (Perry et al., 1997).
Lowering the combustion temperature by operating the
engine under fuel-lean conditions (excess air) or by
recirculating exhaust gas through the engine is often used
to lower the NOx emission from the engine, although these
approaches are not so effective.
The conversion of NO to nitrogen dioxide (NO2) occurs
at low temperatures when exhaust gases are vented to the
atmosphere. This reaction is represented by Eq. (4).
2NO O2 Y2NO2

The NO/NO2 ratio for nitrogen oxides passing into


polluted atmospheres is equal to 10/1, values calculated
from rate constants between 400 and 600 K (Schnelle and
Brown, 2002). However, in the presence of air, the ratio
between NO and NO2 could change because of the
thermodynamic equilibrium as represented in Fig. 2.

1.2.1.1. Thermal NOx .


The majority of NO is formed by
the reaction between nitrogen and oxygen following the
mechanism established by Zeldovich (1946):
N2 O4YNO N4

N2 O2 Y2NO

Fig. 2. Thermodynamic equilibrium between NO and NO2 from an initial


gas mixture of 500 ppm NO2, 5% O2, and 10% water.

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

In fact, it is possible to observe from this figure that, from


NO2 and oxygen at 298 K and increasing the temperature,
there is a progressive formation of NO. At around 673 K,
the equimolar mixture is reached and, after 1073 K, only
NO will be present.
1.2.1.2. Fuel NOx .
Fuel-bound nitrogen (FBN) is the
major source of NOx emissions from combustion of
nitrogen-bearing fuels such as heavy oils, coal, and coke;
and its formation reaction can be described as follows:
CNYIN

IN Oor O2 ; OHYNO . . .

where C(N) denotes the nitrogen in char while I(N)


represents nitrogen-containing intermediate species like
CN, HCN, NH and NH2. Under reducing conditions
surrounding the burning droplet or particle, the FBN is
converted to fixed nitrogen compounds such as HCN and
NH3. These, in turn, are readily oxidised to form NO. Among
these reactions, the reduction of NOx over the char surface is
quite complex and not yet fully understood (Tomita, 2001).
1.2.1.3. Prompt NOx . Hydrocarbon fragments (such as C,
CH, and CH2) may react with atmospheric nitrogen under
fuel-rich conditions to yield fixed nitrogen species such as
NH, HCN, H2CN, and CN as proposed by Fenimore (1972).
These, in turn, can be oxidised to NO in the lean zone of the
flame. In most flames, especially those from nitrogencontaining fuels, the prompt mechanism is responsible for
only a small fraction of the total NOx . Its control is important
only when attempting to reach the lowest possible emissions.
1.3. The impact of NOx
1.3.1. Toxicity
NOx has two types of toxicity, one in liquid and
concentrate state and another like a pollutant gas. As a liquid,
it can be found in industries where acid is produced and/or
laboratories where this gas is used and stored in pressured
bottles. Contact with skin or eyes could cause severe burns.
Additionally, because this liquid has its boiling point at 294.3
K, at 1 atm, risks are rather large for inhalation. In France,
exposition limit allowed for bperoxideQ, dimmer of NO2,
N2O4, is 3 ppm (6 mg m3) and for NO is 25 ppm (30 mg
m3). An exposition, for periods longer than 15 min, at
concentrations higher than 5 ppm of N2O4 (9 mg m3), is
known to produce intolerable irritations or irreversible
modifications on pulmonary tissues. However, lower concentrations could be irritable.
1.3.2. Environmental impact
Nowadays, the deforestation of the northern hemisphere
by contamination is considered one of the most important
ecological problems. Although responsibility might be
directed toward a variety of factors, that of acid rain remains

449

a major contribution among them. Rain is generally a slightly


acidic, and its pH value is around 56, although acid rain pH
has lower values, between 4 and 4.5. Acid rains are due to
SO2 and NO2 present in the atmosphere. In this process,
nitrogen oxides play an influential role in the photochemistry
of both troposphere and stratosphere (Armor, 1996; Cohen
and Murphy, 2003).
The family of nitrogen oxide species is partitioned
between active radicals (NOx , NO3), reservoir species (e.g.,
peroxyacetylnitrate (PAN)) which can convert back into NO2
and terminal species (HNO3, organic nitrates), which no
longer contribute to photochemistry and are efficiently
deposited. At low wintertime temperatures, PAN is stable
and can be transported to the upper troposphere and remote
regions. In the summer, however, the lifetime of PAN is short
(few hours), hence concentrations may remain low despite
abundant photochemical radicals that promote PAN formation (Kermikri, 1995). Thus, temperature directly affects the
partitioning of nitrogen oxides, which will in turn affect
deposition. Nitric oxide is rapidly oxidised by ozone while
radicals such as OH* and HO2 are transformed into NO2,
HNO2, and HO2NO2. Acid rains usually form high up in the
clouds, a locality where nitrogen oxides react with water,
oxygen, and other oxidants. These compounds are transformed into HNO3 that then acidifies rain, snow or fog
because of its solubility in water. These acid precipitations
greatly perturb the aquatic ecosystems and can cause the
biological death of lakes and rivers.
1.4. Emission regulations
Starting in the early 1970s, a policy has gradually evolved
in various countries, notably Japan, the United States, and
Germany, to establish a program of reducing emission levels
of pollutants in flue gases from stationary sources. Table 2
gives an overview of different targets (Erisman et al., 2003).
The last protocol, the Gothenburg Protocol, is unique in
the sense that it establishes reductions of four pollutants to
abate three effects (acidification, eutrophication, and the
effects of tropospheric ozone on human health and vegetation). The protocol has been signed by 29 European countries
together with Canada and the U.S. The countries themselves
estimate the critical loads. Apart from the protocols, the
member states have to fulfil several directives. In May 1999,
the European Commission (EC) presented a proposal for a
Directive on national emission ceilings (NECD) for the same
pollutants as CLRTAP and, for the first time, for ammonia. In
the U.S., for example, the Title IV of EPAs b1990 Clean Air
Act AmendmentsQ is directed at SO2 and NOx emissions from
utility power plants to control the precursors of acid
deposition. Annual emissions reported by parties to the
CLRTRAP are summarised by EMEP (a co-operative
program for monitoring and evaluation of the long-range
transmission of air pollutants in Europe). Updated information related to pollutant emissions could be obtained at the
European Environment Agency (EEA), which interprets

450

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

Table 2
Air emission reduction targets for the EU (Erisman et al., 2003)
Policy/Pollutant

Base year

Target year

Reduction (%)

UNECECLRTAP
Sulphur dioxidea
Sulphur dioxideb
Nitrogen oxidesc
Nitrogen oxidesb
Ammonia

1980
1990
1987
1990
1990

2000
2010
1994
2010
2010

62
75
Stabilisation
50
12

5 th Environmental Plan
Sulphur dioxide
1985
Nitrogen dioxide
1990

2000
2000

35
30

ceiling (NECD) c
1990
2010
1990
2010
1990
2010

78
55
21

Dir. on Nat. Emission


Sulphur dioxide
Nitrogen oxides
Ammonia
a

Target from 1994 second sulphur protocol. The different emission


ceiling for each member state corresponds to a 62% emission reduction for
U.S.
b
Targets from multipollutant bGothemburgQ Protocol.
c
Targets from first NOx protocol.

emission data, co-ordinates the development of the spatial


distribution of emissions, and provides information on
policies and scenarios (Vestreng et al., 2000); and in the
European Pollutant Emission Register (EPER) which is the
first Europe-wide register of emissions into air and water
from large- and medium-sized industrial facilities in Europe.
The first EPER report includes data for the year 2001 from
about 10 000 industrial facilities in the European Union and
Norway (EPER, web site).
1.5. Actual approaches to NOx abatement
A first approach consists in modifying the operational
conditions to decrease the NOx formation. Among new
developed technologies, the furnaces with low production of
NOx represent an important contribution. Low NOx burners
that minimize thermal formation of NOx by reducing
combustion temperatures and controlling flame stoichiometry make an important contribution and, under well-tuned
optimized conditions, NOx reductions of 40% or more can be
achieved. The reduction of NOx occurs by reducing the
temperature of combustion and by controlling of the added
air. However, this procedure implies side effects like the
increase of carbon in ashes, a greater CO formation, as well
as problems of corrosion and slags because of a local
reducing environment.
A second approach consists in destroying or allowing
NOx to react with other reagents, making it possible to
eliminate them or to modify them. From a thermodynamic
point of view, NO and NO2 are unstable:
NO X 1=2 N2 1=2 O2 DG8  86kJ
NO2 X 1=2 N2 O2 DG8  51kJ

mol1

mol1

Despite this thermodynamic instability, kinetic studies


have revealed that the activation energy for homogeneous
decomposition of NO is high (~335 kJ mol1) (Garin,
2001). Therefore, a catalyst is necessary to lower this
activation energy in order to facilitate this decomposition. In
general, exhaust gas treatment methods are preferred for
NOx control because they provide for utility system loads
wider range than possible by combustion controls. Among
the methods being investigated to clean exhaust gas streams
of NOx are (Lyon et al., 1990; Ramachandran et al., 2000):
(1)

Injection of methanol into flue gases to convert NO to


NO2 (potential by-products are CO and CH2O), which
is removed in a dliquid-modifiedT wet limestone SO2
scrubber. Various configurations of the latter step
include the use of a spray dryer with aqueous slurried
lime with NaOH as an additive, where the NOx
removal rate is 3550%.
(2) The addition of some metal chelates, such as ferrous
ethylenediaminetetraacetate (Fe(II)*EDTA), increases
the adsorption of NO by reacting to produce ferrous
nitrosyl complex (Fe(II)*EDTA*NO). The resulting
spent scrubber solution is then biologically treated to
remove the nitrosyl adduct (bound nitric oxide) from
the ferrous EDTA and to reduce the oxidized ferric
EDTA (Fe(III)EDTA) to the ferrous form that is
active for NOx scrubbing. In addition, the process
water can be amended with limestone.
(3) Use of adsorbents, e.g., utilising alumina impregnated
with sodium carbonate in a fluidised bed, where it was
found that SO2 must be present for efficient NOx
removal.
(4) Selective nocatalytic reduction (SNCR), using aqueous
urea (Eq. (7)) or ammonia injection into the furnace at
11481423 K,
CO NH2 2 2NO 1=2 O2 X 2N2 CO2 2H2 O

(5)

7
The reduction of NOx is between 30% and 75%.
Selective catalytic reduction (SCR) of NO using
ammonia (or other compounds such as hydrocarbons)
as the reductant to form nitrogen and water. The
principal types of catalysts that have been investigated
for SCR are:
(a) Supported noble metal catalysts, e.g., Pd/Al2O3.
(b) Base metal oxide catalysts, e.g., those containing
vanadium.
(c) Metal ion exchanged zeolites, e.g., Cu-ZSM-5.

The typical characteristics of a SCR process including


effectiveness of NOx reduction can reach 85% and SO2
conversion between 1% and 2% (Orsenigo et al., 1996).
While NO reduction with NH3 can occur as shown in Eq.
(8), both the SCR and SNCR processes are carried out in the
presence of air (O2), as represented by Eq. (9), with the
reaction represented by Eq. (10) occurring as a side reaction

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

that is favoured by oxygen excess. In the presence of


oxygen, NO can be oxidised to NO2 (favoured by lower
reaction temperatures), but NO2 is also reduced by NH3, as
represented by Eq. (11).
6NO 4NH3 X 5N2 6H2 O

4NO 4NH3 O2 X 4N2 6H2 O

4NO 4NH3 3O2 X 4N2 O 6H2 O

10

2NO2 4NH3 O2 X 3N2 6H2 O

11

The catalytic methods are probably the most used and


adequate because they transform NOx into nitrogen, which
represents a true solution to the pollution problem. That is
why, the SCR method will be described in a more detailed.
1.5.1. Selective catalytic reduction (SCR) of NOx
Many challenges exist when applying catalysis/catalysts
to fuel combustion processes (stationary applications) for
NOx emission control. The basic chemistry involves the
following reactions:
Selective or desired reactions:
4NO 4NH3 O2 X 4N2 6H2 O
2NO2 4NH3 O2 X 3N2 6H2 O
Nonselective (other possible) reactions:
2SO2 O2 X 2SO3
4NH3 5O2 X 4NO 6H2 O
4NH3 3O2 X N2 6H2 O
Ammonia or, in some cases, urea reacts selectively to
reduce the NOx . The nonselective reaction consumes the
reagent and reduces the NOx conversion. In situations
where sulphur compounds are present, the conversion to
SO3 must be minimised to prevent salt formation and
deposits on heat transfer surfaces, which reduces the heat
transfer efficiency (Satterfield, 1991; Svachula et al., 1993).
The composition of the gas stream (flue gas) must be
monitored permanently and the quantity of NH3 reductant
added must be tightly controlled to minimize slip into the
exhaust stream. In general, SCR removes between 60% and
85% of NOx using between 0.6 and 0.9 mol NH3 for 1 mol
of NOx and leaves between 1 and 5 ppm of unreacted NH3
(slip). The various characteristics of different SCR catalysts
are given in Table 3 (Heck, 1999).
The catalyst that is used in large-scale applications is a
combination of V2O5 and TiO2 supported on a monolith or
wire screen plate where NH3 is used as the reductant. One
of the most effective catalysts is about 6 wt.% vanadia on
titania (Vogt et al., 1991) which sometimes is promoted
with WO3 or MoO3. Because of the variant exhaust gas
compositions, particulate loading, and contaminants, there

451

Table 3
Operating characteristics of different SCR catalysts (Heck, 1999)
Low temperatureLT catalyst (Pt-based)
423573 K
Narrow temperature window
Temperature window shifts
Not sulphur tolerant
Medium temperatureVNXk catalyst (V2O5/TiO2)
533700 K
Most broadly used
1015 years of experience
Sulphur tolerant
High temperatureZNXk catalysts (zeolite)
618863 K
Very high NOx conversion
Very low NH3 slip
NH3 destruction
Sulphur tolerant 698 K

are different catalyst support structures. The extruded


catalyst and the metallic support are typically used in
high dust conditions and have low cell densities (10100
cells in.2 or cpsi) and the composite catalyst (either on a
metallic or ceramic monolith) is used in low dust
conditions and has a higher cell density (from 64 to 400
cpsi). However, this catalyst has a rather narrow temperature range of work, with WO3 added to stabilise the
catalyst up to 723 K or broaden the temperature window to
lower reaction temperatures (Gutberlet and Schallert,
1993). The optimum temperature is in the region of
573673 K. NO can be completely reduced at lower
temperatures, but SO2 is oxidised to SO3 by the vanadia.
Compounds, such as NH4HSO4 and (NH4)2S2O7, that will
deposit on the catalyst below 523573 K, are also formed.
The oxidation of NH3 becomes appreciable at 673 K. It
has also been reported that WO3 and MoO3 make the
catalyst more poison-resistant (Satterfield, 1991). Catalyst
behaviour depends on level of the V2O5 (Machej et al.,
1990a,b; Yang and Cheng, 1992), and it must be controlled
to minimize the SO3 formation (Machej et al., 1990c). For
V2O5 on titania catalysts, bench-scale results have shown
that at NH3: NOx ratios smaller than 1, the NOx
conversion linearly increases with increasing ratio, and
the reaction rate depends on the concentration of ammonia.
For ratios higher than 1, the reaction rate is dependent on
the concentration of NOx . Chemical and mechanistic
aspects of the selective catalytic reduction of NOx by
ammonia over oxide catalyst were reviewed by Busca et
al. (1998) and Parvulescu et al. (1998).
Ammonia causes not only the problem of slip but also of
transportation and storage. In most applications, the NH3
slip or NH3 in the exhaust is also specified or regulated
along with the NOx emissions. That is why a powerful
stimulus for research is the discovery that SCR can be
carried out using hydrocarbons as reducing agents (HCSCR: Amiridis et al., 1996; Bosch and Janssen, 1988). The

452

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

first catalyst reported to possess HC-SCR activity is the


classic CuZSM-5 zeolite (Iwamoto and Hamada, 1991).
More recently, it has been shown that different cationexchanged zeolites (e.g., Co, Ni, Cr, Fe, Mn, Ga, In) are
active in this reaction (Kikuchi et al., 1996). Different base
oxides/metals (e.g., Al2O3, TiO2, ZrO2, MgO, and these
oxides promoted by, e.g., Co, Ni, Cu, Fe, Sn, Ga, In, Ag
compounds) have also been proposed as supports (Burch et
al., 2002). An important stage in the HC-SCR development
is the discovery by Li and Armor (1992) that, over Co
ZSM-5, NOx reduction can be performed by methane. This
could allow replacing ammonia as a reducer. Lately, the
perimeter of the potential catalysts operating with methane
has significantly increased (Armor, 1996). Therefore, the
selectivity of CH4 toward reacting with NO is much higher
on Ga-ZSM-5 (100% at 723 K, which means that CH4
reacts with NO only without combustion needless) than on
Co-ZSM-5 (46% at 723 K), and this difference increases
with temperature (82% and 22% at 823 K for Ga-H-ZSM-5
and Co-ZSM-5, respectively). It appears that two different
mechanisms may be operating for the NO reduction over
Co-ZSM-5 vs. Ga-H-ZSM-5, and this difference is probably
related to their very different active centres (over Co-ZSM5, NO adsorbed at Co sites while over Ga-H-ZSM-5, CH4 is
activated on Ga sites).
1.5.2. Conclusions on bactual approaches to NO x
abatementQ
Although that the selective catalytic reduction by
ammonia is currently the most widespread method for the
clean up of flue-gas from stationary sources, there are still
many problems to overcome. Among the advantages of NH3
as the reducing agent is high selectivity toward the reaction
with NO in the presence of O2 and the promoting effect of
oxygen on the rate of NONH3 reaction. Typically, during a
commercial SCR process over V2O5WO3TiO2 systems,
stoichiometric control of the ammonia must be maintained
to avoid emissions of unreacted 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. The use
of hydrocarbons in the ammonia place seems promising.
1.6. New techniques for NOx abatement
The NOx abatement has been improving in the last years
by two new techniques via the car industry: Selective NOx
Recirculation (SNR) and NOx Storage and Reduction
(NSR), both of them have the adsorption of NOx on a
sorbent as common characteristic but the decomposition
procedure are different. We will present some of their
characteristics in the following. The crucial point is the
efficiency of the sorbent material. In general, this material
should have a high NOx trapping capacity, a high selectivity
toward NOx in a complex mixture of gases, a temperature of

desorption low enough for the process to be profitable with


an energy input as low as possible and finally, an essential
necessity, and a high resistance to SO2 poisoning.
1.6.1. Selective NOx recirculation (SNR)
SNR technique has been developed by DaimlerChrysler
in 1994, and it is concerned with the treatment of exhaust
gases coming from motor fuel-lean or diesel (Bogner et al.,
1995; Kiennemann et al., 1998). Two adsorbers operating
alternately in the adsorption and desorption modes are
arranged in parallel in the exhaust system. Upstream of the
adsorber system, a control valve is provided to separate the
exhaust stream from one line to the other. Downstream of
the adsorbers, return valves are arranged for the recirculation of desorbed NOx . For the coordinated operation of the
adsorptiondesorption mode, upstream- and downstreamlocated valves are in communication. Therefore, the
principle of this concept consists in the concentration and
recirculation of NOx into the combustion chamber of the
engine where they will be decomposed thermally. Because
thermal decomposition depends highly in NOx concentration, the performance of the adsorbent material is one of the
clues for the success of the SNR concept.
1.6.2. NOx storage and reduction (NSR)
This concept has emerged whereby the problem of
destruction of NOx is undertaken by a two-stage operation,
the combination of two air/fuel ratios instead of a single
fixed ratio. During a fuel-lean stage, NOx is trapped upon a
selective sorbent in the form of nitrate (NO3) as proposed
by Takahasi et al. (1995). Because there is plenty of oxygen
present at this stage, HC, H2, and CO are readily oxidised to
water and carbon dioxide. Then, when the engine is
switched to operation fuel-rich (normal air/fuel ratio), the
resulting exhaust becomes comparatively oxygen-deficient
and HC, H2, and CO remains unoxidised. Hence, the three
components react with the NO3 into harmless nitrogen,
water, and carbon dioxide (Fig. 3; Matsumoto, 1996;
Misono and Inui, 1999).
The first catalyst of this type was developed by Toyota
(Kato et al., 1993) which composition was based upon that
of a three-way catalyst. The preparation was by impregnation of noble metals upon alumina (essentially Pt) of several
alkali and alkali earth metals (Na+, K+, Ba+2) and of rare
earth oxides (mainly La2O3) (Matsumoto, 2000). The
basicity of the storage component determines the amount
of NOx adsorbed. Nevertheless, the performance of platinum, especially the hydrocarbon oxidation, decreases with
support basicity. The particle size of platinum and BaO and
the distance between them are other factors which govern
the NOx storage: small particles in contact with each other
adsorb more than large particles (Takahasi et al., 1996).
1.6.3. Regeneration of NOx sorbents
Continuous gas separation requires at least two independent sorbent beds in parallel, which are alternatively used for

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

453

Fig. 3. Adsorption and reduction of NOx by NSR (NOx Storage and Reduction) concept (Misono and Inui, 1999). NSC: NOx Storage Compound.

sorption and regeneration steps. The resultant mechanical


complexity imits the practical application of those processes,
but recent attempts have been made to overcome this
problem. The solidNOx interactions obey chemical equilibrium which depends on temperature, pressure, and atmosphere (reducing or oxidizing). Thus, the regeneration of NOx
sorbents can employ pressure swing, or thermal (temperature)
swing operations, or cyclic oxidationreduction treatment.
1.6.3.1. Pressure swing process. Fig. 4 shows schematically the principle of pressure and thermal swing adsorption
(PSA and TSA, respectively). In the PSA process, adsorption takes place at high partial pressure and desorption takes
place at low partial pressure. Pressure swing adsorption
(PSA) has widely been applied to various processes, such as
the separation of oxygen or carbon dioxide from air and
purification of hydrogen. The application of this process to
remove NOx requires sorbents with a high capacity for
reversible sorption (Arai and Machida, 1994).
However, for conventional adsorbents, such as activated
carbons, silica gels, and zeolites, the NOx adsorption
capacity is not sufficient for practical uses. Fig. 5 represents
the typical breakthrough and elution curves when the
adsorption and desorption runs were repeated. In Fig. 5,
the solid and broken lines indicate the concentration of NO
with and without adsorbent, respectively. The amount of NO
adsorbed corresponds to the area denoted as a n , whereas
that of NO desorbed is b n . When a n is to be equal to b n ,
reversible sorption is possible.

Fig. 4. Principle of pressure or thermal swing adsorption (Arai and


Machida, 1994).

1.6.3.2. Thermal (temperature) swing process. The TSA


process utilises differences in sorption capacity at different
temperatures (Fig. 4). The sorbent bed operates isothermally
during sorption steps, whereas, in regeneration steps, the bed
is heated to give energy to desorb NOx . The temperatures for
sorptionregeneration steps are cycled by use of external
heater or by alternating feed gas temperatures. This process
is quite useful for thermostable sorbents, the NOx sorption
isotherm of which is strongly temperature-dependent.
Because most NOx solid reactions, with exception of simple
physisorption, are strongly dependent upon the reaction
temperature, a TSA process is expected to give rise to much
larger NOx sorption capacities than PSA processes. On the
other hand, heating and then cooling of a solid sorbent
clearly requires much more time than changing the pressure.
1.6.3.3. Oxidationreduction cycles. A oxidationreduction cycle can be applied not only to regenerate the sorbents
but also to reduce any accumulated NOx to N2. In an
oxidizing atmosphere at relatively low temperatures (b573
K), NO is converted into NO2, which is much more reactive.
NO2 then reacts with solids to produce nitrate or nitrite
precipitates. In a reducing atmosphere, however, nitrate/
nitrite will be decomposed and reduced to N2 over suitable
catalysts such as noble metals.
1.6.4. Conclusions on bnew techniques for NOx abatementQ
Besides such dual-bed operation, NOx sorbing catalysts
having integrated structures of NOx sorbing sites and
catalytically active sites for reducing NOx would be possible.
Provided that these bifunctional surface structures are stable
under reaction conditions, the combination of NOx sorption
and catalysis may give rise to a synergistic effect to the

Fig. 5. Typical breakthrough and elution curves of NO: Solid line, with
adsorbent; broken line, without adsorbent; C 0, initial concentration of NO;
a n , NO adsorbed; b n , NO desorbed (Iwamoto, 1990).

454

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

Table 4
NO adsorption capacities (mg NO m2) for supported metal oxides on
aluminasilica (95:5 wt/wt) at 1 mbar NO (Shelef, 1975; Kiel et al., 1992;
Eguchi et al., 1996)
Metal oxide

Supported
amount
(wt.%)

Temperature
(K)

Surface area
(m2 g1)

Adsorption
capacity

Fe3O4
CuO
NiO
Cr2O3 with
Cr4O11
a-Fe2O3
Co3O4
Pt (oxidised
surface)

8.5
8.9
6.9
10

299
299
299
273

90
104
73
50

0.23
0.12
0.15
0.12

8.5

298
298
273

90
20
8.6

0.06
0.23
0.20

development of NOx -sorbingreducing catalyst. In such a


way, the concept of oxidationreduction cycles seems to be
the most appropriate for its applications to fixed installations.
1.7. NOx -sorbing catalyst materials
Sorptive NOx removal utilises adsorption, absorption
(means bulk-type reactions between the solid sorbent and
NOx ), and/or solidgas reactions. Different material types
are reported for the sorption of nitrogen oxides: some can
adsorb NO or NO2 only, but others can adsorb both of them.
For reviewing, they were classified as follow: metal oxides,
spinels, perovskites, zeolites, carbonaceous materials, and
heteropolyacids. Materials and mechanisms will be discussed in more detail in the following sections.
1.7.1. Metal oxides
1.7.1.1. Transition metal oxides. An extensive data set of
NO-uptake on metal oxides is supplied in the literature
(Winter, 1971; Gandhi and Shelef, 1972, 1973; Yao and
Shelef, 1973; Otto and Shelef, 1969, 1970, 1973, 1975;
Eguchi et al., 1996). The NO sorption is measured for
several metal transition oxides, mostly supported on an
aluminasilica (95:5 wt/wt) carrier. Table 4 gives further
information of the materials.
Note that a partial pressure of 1 mbar NO corresponds to
a concentration of 1000 ppm NO in an inert gas at a total
pressure of 1 bar. Reduced forms of Fe2O3 (as Fe3O4),
chromia (as Cr2O3), MnO2 (as MnO), and Pt-black are
obtained by reduction with CO or H2 at 723 K. They show
distinctly higher sorption capacities than the original oxides.
It is established that the chemisorption of NO on Cr3+, Co2+,
and Mn2+ is especially large because of formation of
covalent bonds between NO and the metal ion. It is
doubtful, however, that these reduced oxides surfaces are
stable in an oxygen-containing hot gas. Fe3O4 and MnO are
even oxidised by NO. CuO is the only metal oxide with a
higher sorption capacity for the oxidised form than for the
reduced form Cu2O. The reasonable NO sorption capacity

of CuO (Table 5), however, cannot be utilised because CuO


is also a good SO2 scavenger via formation of CuSO4 (Kiel
et al., 1992). The NOx capacity for unsupported Pt-particles
(average size of ca. 30 nm) is low per weight of material, but
relatively high per surface area, compared to the supported
oxides (Table 4).
Also on Pd-surfaces, a strong chemisorption is observed,
but no NOx -storage capacities are given (Zuelke et al.,
1968). These results on Pt and Pd indicate that the precious
metals can be valuable for NOx -storage. Pt or Pd on galumina also catalyses the oxidation of NO into NO2
(Karlsson and Rossenberg, 1984). On Pt, however, SO2 is
easily oxidised into SO3, with fast sulphate deposition as a
consequence. This problem could be overcome by the
addition of metal oxides as V2O5, MoO3, or Nb2O3. NO
sorption is also observed on TiO2 (Abad et al., 2004) and
SnO2 (Solymosi and Kiss, 1976) at room temperature (the
capacity of SnO2 is low but increases when part of the Sn4+
ions are reduced into Sn3+).
1.7.1.2. Rare earth oxides.
Several reported studies
indicate that in the presence of ceria, the activation energy
for NO dissociation and/or the activation energy for N2
desorption is lowered (Oh and Eickel, 1988; Loof et al.,
1991; Padeste et al., 1994; Rao et al., 1994; Trovarelli,
1996). When CeO2 is pretreated in oxygen, 3.5 mg NO g1
is sorbed at 4 mbar NO pressure and room temperature.
Nitrogen formation following NO adsorption over metalloaded CeO2-containing catalyst following reductive treatment has also been reported (Niwa et al., 1982). NO
decomposition occurs at a reduced M-CeO2 interface site
with reoxidation of CeO2 by oxygen of NO. Cordatos and
Gorte (1996) studied NO adsorption on ceria-supported Pd
catalysts. They believe that NO adsorbs on reduced ceria.
They observed negligible adsorption on the ceria without Pd,
demonstrating that the oxidised ceria cannot adsorb NO. The
large amount of NO on the reduced sample suggests that NO
reoxidises CeO2x and desorbs as N2. The picture of ceria
supports effects that emerges is one in which oxygen, and
possibly NO, move relatively free between the ceria and the
catalytic metal. Lattice oxygen from CeO2 at the metalceria
boundary provides oxygen to the metal for oxidation
reactions, while reduced sites can be oxidised by NO, giving
up N2 and completing the oxidationreduction cycle. The
relatively free movement of molecules between sites on the
catalytic metal and on the ceria provides a means for
enhancing NO dissociation. This is very important for metal

Table 5
Amount of NO2 (mg gcatalyst1) adsorbed and desorbed on barium aluminate
as a function of the calcination temperature (Hodjati et al., 1998a)
Temperature (K)

Surface (m2 g1)

Adsorbed NO2

Desorbed NO2

1073
1273
1473

161
90
19

24
21
3

24
20
3

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

like Pd, Pt, and Rh which are not active for NO dissociation
(Bhattacharyya and Das, 1999; Garin, 2001).
Introducing foreign cations such as Al+3, Zr+4, or Si+4 into
the CeO2 lattice to form a solid solution, may significantly
improve the stability of the surface and strongly enhance the
redox properties of ceria (Fornasiero et al., 1995; Zhang et
al., 1995; Colon et al., 1999). The increase in oxygen
storage/transport and redox capacities is the result of
enhanced oxygen mobility within the crystal lattice originated from formation of a defective solid solution. This
consideration has been used for example to explain the
higher reducibility of CeO2 and its oxygen storage capacity
in CeO2ZrO2 mixed oxides (Fornasiero et al., 1996; Colon
et al., 1998). Oxygen mobility should be favoured in ceria
zirconia solid solutions because the substitution of Ce4+ with
the smaller Zr4+ cation (radius 0.84 2) causes the shrinking
of the CeO2 fluorite-type lattice and the formation of
structural defects (Cutrufello et al., 2002). The energetics
of oxygen ion transport in the CeO2ZrO2 solid solutions has
been investigated by employing computer simulation techniques. Balducci et al. (1987, 1998) could investigate the
defect properties for the whole composition range and
provide information that is relevant to catalytic behaviour.
Some conclusions define that:
(i)

The activation energy for oxygen migration in the bulk


is found to be low and decreases almost monotonically
with the zirconia content (this indicates facile oxygen
diffusion through the bulk catalyst) and
(ii) The Ce+4/Ce+3 reduction energy is significantly
reduced even by small amounts of zirconia.
To understand the NOx mechanism over ceriazirconia
mixed oxides, Daturi et al. (2000, 2001) developed NO
adsorption studies over prereduced samples (on which
anionic vacancies can be well characterised) analysed by
FT-IR and mass spectroscopy. Following the model
proposed, the reduction capacity of the catalyst will depend
on the amount of surfaces vacancies created by the
reduction treatment and on the overall amount of exchanged
oxygen. From their experimental results, it appears clearly
that the amount of nitrogen arising from NO reduction has
been strikingly increased by a factor ~3.3 for CeO2ZrO2
mixture compared with CeO2. Oxygen surface defects, fast
filled by NO oxidising agent, are recreated by oxygen
migration through the material bulk. In fact, the reduction
process of ceria-based materials may be represented by the
following equations:
2
3
2
2Ce4
s 4Os Y2Ces 3Os 5 OO2g

455

(14)

(15)

Here, the used notations mean: oxygen vacancy with no


electrons: 5; one electron (F+ centre): or two electrons (F
centre):
; (s) indicates surface, and (sb) near surface
species. Eq. (12) will take place on ceria and Eq. (13) on
ceriazirconia. The elimination of one O surface atom
produces one vacancy or one F(s)+centre and one Ce+4/Ce+3
reduction. Eq. (14) is the migration of a near subsurface
oxygen atom to the surface, leading to a subsurface F centre,
the electrons of which are further stabilised as Ce+3 in Eq.
(15). The inverse process gives rise to sample reoxidation.
The mechanism presented Fig. 6 is not only justifies catalyst
activity but also selectivity: N2O seems not to take place,
contrary to what happens when NO adsorbs and decomposes on platinum catalysts.
Even if N2O is produced, it should be rapidly dissociated
by the reduced catalyst, as proved by other studies (Fally et
al., 2000).
1.7.1.3. Alkaline earth oxides. Hodjati (1998) developed a
new type of material from the family of barium aluminates
(BaAl2O4, BaAl12O19). For bulk BaO tested under fuel-lean
condition, the adsorption capacity was 15 mg NOx g1 of
catalyst, but BaO lead only to a small desorption which
began at 823 K, the maximum temperature used in the
standard test. Nevertheless, an increase of temperature to
923 K lead to complete desorption of the NOx previously
adsorbed. For this catalyst, it was not possible to obtain
reproducible quantities of adsorbed NOx during successive

12

13

Fig. 6. Proposed reaction mechanism for NOx reduction on CeO2 and


CeO2ZrO2 (Daturi et al., 2001).

456

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

runs but instead a decrease of adsorption capacity was


observed. For Pt/Al2O3, the adsorption capacity was 10 mg
NOx g1 of catalyst, with a good reproducibility. The
crystalline form of barium aluminate strongly depends on
calcination temperature: below 1473 K, the solid has a
nanocrystalline form, while above 1473 K, a well-defined
crystalline structure is observed. For barium aluminates with
specific surface of 90 m2 g1, the NO2 adsorption capacity
is 21 mg NOx g1 of catalyst; and it is clearly a function of
the calcination temperature, i.e., of the surface area, as
shown in Table 5.
Hodjati (1998) clearly indicate that NO2 does not
need to be formed on the surface to be trapped by the
barium aluminates as observed in the catalyst designed
by Toyota (Kato et al., 1993) where it was postulated
that the role of noble metal was to oxidise NO into NO2
before a subsequent adsorption on barium oxide. Additionally, Hodjati et al. (1998b) made some experiments
in order to check the existence of the different species
formed upon barium compounds (carbonates, nitrates,
and sulphates). Barium oxides and barium aluminates
catalysts were exposed to a gas mixture containing 500
ppm NO, 500 ppm NO2, 5% CO2, 10% O2, 5% H2O
and, in other experiments, 25 ppm SO2. The difference
in behaviour between bulk BaO and barium aluminates
is striking. Principally, when submitted to a mixture of
gas representative of fuel-lean conditions, barium aluminate undergoes selective adsorption of NO2 (20 mg g1
of catalyst).

While this process remains reproducible, it should be


noted that no deactivation was observed at all. The
difference in the behaviour between barium oxide and
barium aluminate can be explained by the fact that strongly
held carbonates are irreversibly formed on the former which
was proved not to be in the case of barium aluminates,
where the competition between nitrate and carbonate
formation always favours the nitrates. Due to the strong
basicity, barium oxides tend to form sulphates in the
presence of SO2. Sulphates are formed at the expense of
nitrates which, in the case of barium aluminates, will result
in the breaking of the bariumaluminium bond leading
therefore to the same type of deactivation as BaO with the
favoured formation of BaSO4 (as verified by infrared (IR)
and XRD). Those sulphate compounds are very stable and
do not decompose even when subjected to calcination at a
temperature as high as 1273 K. However, when the extent of
poisoning is low (less than 30% of BaO involved in sulphate
formation), it is possible to regenerate the catalyst under
reducing conditions and the initial storage capacity can be
recovered, even in the presence of some remaining sulphates
(Courson et al., 2002). Infrared and thermodynamic
calculations were carried out by Breen et al. (2002) in
order to ascertain the temperatures and gas composition
under which barium sulphate can be converted to barium
carbonate or barium oxide. Results are represented in Fig. 7
(for temperaturesNT DGrxn=0, the products of reaction will
spontaneously be formed, and at temperaturesbT DGrxn=0, no
reaction will occur).

Fig. 7. Evolution of Gibbs free energy (DG rxn) as a function of temperature for various reactions of barium sulphate and carbonate (Breen et al., 2002).

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

Fig. 7 shows the inherent stability of BaSO4, and


spontaneous decomposition only occurs at temperatures
N1666 K. BaCO3 is also stable but not as stable as BaSO4.
Indeed, Fig. 7 indicates that BaSO4 will replace BaCO3
((a)(b)) at temperatures up to 1744 K even under
conditions where the partial pressure of CO2 far exceeds
that of SO2. The behaviour of BaSO4 under reducing
conditions is particularly interesting. The reaction of
barium sulphate with hydrogen to give barium oxide,
SO2 and H2O (c) occurs, as expected, at much lower
temperature (1236 K) than the decomposition of barium
sulphate to barium oxide in the absence of hydrogen (1666
K). However, the temperature for the substitution of
barium sulphate can be further reduced to 896 K by
including CO2 in the gas mixture and allowing for the
formation of barium carbonate rather than barium oxide
(d). This is important because it shows that the addition of
CO2 to the gas mixture can considerably lower the
temperature at which the stable sulphate can be removed.
By using only carbon monoxide as the reducing agent, the
temperature of carbonate substitution of sulphates can be
further reduced to 820 K (e). If hydrogen sulphide is
formed instead of sulphur dioxide under rich conditions,
then the DG rxn changes considerably. The decomposition
of BaSO4 to BaO (f) can occur at much lower temperature
(826 K) than the temperature of 1236 K required to
decompose the BaSO4 to BaO with SO2 rather than H2S as
a product (Eqs. (c) and (f)).
1.7.2. Spinels (AB2O4)
The spinel oxides belong to a class of complex oxides
presenting the chemical formulas AB2O4 in which A ions
are generally divalent cations occupying tetrahedral sites
and B ions are trivalent cations in octahedral sites. The
spinel crystal structure is build up by a cubic closed packing
(ccp) array of oxygen anions: 1/8 of the tetrahedral holes are
occupied by cations A and 1/2 of the octahedral holes by
cations B. It has been known that the catalytic activity of
spinels depends essentially on two factors: the degree of A
substitution and the degree of the inversion of the spinel.
Variants are possible, such as inverse the spinel structure,
B[AB]O4, with half of the B cations in interstices and the
other half with the A ions in octahedral ones (MullerBushbaum, 2003). Spinels, such as ZnCo2O4, have good
NO sorption abilities (Panayotov et al., 1986). Octahedral
configuration of Co3+ ion seems important for the sorption
ability. Co3O4 forms an inactive CoAl2O4 phase on an
alumina support, with Co2+ in tetrahedral co-ordination.
However, spinels with partial replacement of Co2+ by Ni2+
or Mg2+ could completely decompose N2O into nitrogen
and oxygen at 473 and 573 K in the absence and presence of
excess of oxygen and water (Yan et al., 2003). Shelef (1975)
derived the following sequence for the uptake of NO per
surface area at 298 K and a pressure of 1080 Torr. NO:
CuAl2 O4 JCuCr2 O4 JNiAl2 O4 JCoAl2 O4

457

1.7.3. Perovskites (ABO3)


Ideal perovskite crystals are characterised by a ccp array
of oxygen anions and large cations A in dodecahedral
position. The smaller cation B (mostly transition metals)
occupies those octahedral holes formed exclusively by
oxide ions. Advantages of perovskites are:
(i)

Their ability of stabilising transitions metal ions in


different valences states.
(ii) Their variable oxygen stoichiometry at elevated
temperatures.
(iii) The strong ligand binding of the transition metal ions
for NO (and CO).
(iv) The variability of the structural framework.
Generally, La, Sr, Ba, Cs, K, and/or Na occupy the
position A in perovskites (Viswanathan, 1992). The reports
available of NO-sorption most deal with Ba and La.
Systematic studies of perovskites as NO-sorbents are based
primarily on a variation of the cation B. The type of cation
A is regarded as a secondary factor, with influence on the
stabilisation and valence of the transition metal B. Tascon et
al. (1985) published a systematic quantitative study of NO
sorption on LaMO3 perovskites (M=Cr, Mn, Fe, Co, Ni).
NO adsorption at room temperature showed maxima for
LaMnO3 and LaCoO3. The number of NO molecules
adsorbed at monolayer coverage on LaMO3 oxides did not
equal estimated numbers of transition metals ions exposed
in the oxide surfaces. This fact implies a special adsorption
mechanism for this type of compounds.
Hodjati et al. (2000a,b) prepared perovskites with A=Ba,
Sr, and Ca, and B=Sn, Zr, and Ti were prepared by solgel
like method. For these materials, the amount of NOx trapped
is always equal to the amount of NOx desorbed upon
heating. Table 6 gives a summary of the adsorption
capacities of all of the samples tested.
A surface process only cannot explain the large quantity
of nitrates formed at 298 K. The presence of bulk nitrates
was revealed by XRD (the presence of BaSnO3, Ba(NO3)2
equally confirming the absence of barium carbonates). The
fact that Ba(NO3)2 appears implies that a fraction of the
original structure of BaSnO3 is transformed into this
compound. This requires at least a partial opening of the
structure during the absorption phase while the perovskite

Table 6
Amount of NO2 trapped (mg g1) on various perovskites (Hodjati et al.,
2000a)
Structure

Amount of NO2

BaSnO3
SrSnO3
CaSnO3
BaZrO3
SrZrO3
CaZrO3
BaTiO3

17
13
9
12
8
3
3

458

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

structure is reformed upon desorption. This leads them to


propose the following mechanism of absorption/desorption:
Absorption phase:
BaSnO3 3NO2YBaNO3 2 SnO2 NO
Desorption phase:
Ba NO3 2 SnO2 YBaSnO3 2NO2 1=2 O2
The variation of NO, NO2, and NOx on BaSnO3 as a
function of temperature is given in Fig. 8 (Kiennemann et
al., 2002). In a first step, the formation of NO during the
absorption phase and the formation of O2 during desorption
were detected. This explains the process of opening of the
perovskite structure, its thermal reconstitution, and the
continuity of the absorptiondesorption process over several
cycles. The thermodynamic effect is evident from the
increase of NO concentration with the temperature (N723
K) at expenses of NO2 concentration (see also Fig. 2).
1.7.4. Double-layered cuprates, La2x Bax SrCu2O6
Machida et al. (1997, 1998) have reported the NOsolid
reaction based on intercalation into substituted doublelayered cuprates, La2x Bax SrCu2O6. This material shows
rapid uptake of gaseous NO up to ca. 30 mg molBa1 via
intercalation at 523 K. The crystal structure of the layered
cuprate is represented in Fig. 9.
In La2x Bax SrCu2O6, Ba and La appear to be situated in
the interlayer between basal planes of CuO5 sheets where
intercalation takes place. The NO intercalation is driven by an
equimolecular interaction with Ba in the interlayer. Part of the
incorporated NO is considered to react with interlayer
oxygens near Ba and produce NO2 species, which lead to
a reversible NO evolution between 573 and 923 K. Other part
of the incorporated NO occupies anion vacancies in CuO5
sheets, Fig. 10(ii), leading to N2 evolution at TN1023 K. The
layered cuprate therefore contains two different types of NO
sites in the interlayer, i.e., the reversible site (interlayer Ba)

Fig. 9. Schema of ideal crystal structure of double layered cuprate,


La2x Bax SrCu2O6 (Machida et al., 1998). Tetrahedrons and circles
represent CuO5 and La/Ba layers, respectively.

and the irreversible site (anion vacancy in CuO5 sheets). In


the irreversible site, Cu species liganded by NO are highly
oxidised so that lattice oxygens are eliminated when heated
above 823 K, Fig. 10(iii). This means that CuO5 oxygen is
partially replaced by NO at the end of this oxygen evolution.
A further increase in temperature results in the thermal
cleavage of the NO bond above 1123 K (Fig. 10(iv)) and
subsequent elimination of N2 while leaving the NO oxygen in
the lattice.
According to this proposed mechanism, it is expected
that the NO sorption/desorption property is influenced by
coexisting gaseous oxygen. Authors have confirmed that,
when the NO sorption is conducted in the presence of
gaseous oxygen, NO is readily oxidised to NO2. The strong
ability for NO2 to coordinate with Ba allows rapid sorption
into a solid as already reported for the BaCuO system.
The solid exposed to NO/O2 mixtures at 523 K showed a
strong IR absorption at 1380 cm1 indicative of nitrate ion

Fig. 8. Absorption and desorption profile on NO and NO2 over BaSnO3 under fuel-lean conditions (Kiennemann et al., 2002): (a) NO formation; (b) NO2
absorption then desorption; and (c) thermodynamic equilibrium NO X NO2.

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

459

Fig. 10. Model for dissociative desorption of NO intercalated by the layered cuprate (Machida et al., 1998). !: Cu.

(NO3). This is in contrast with the formation of nitrite ions


(NO2) for NO sorption in the absence of O2. On the other
hand, irreversible NO species occupying anion vacancies
are expected to decrease in the presence of gaseous oxygen
because the oxygen-defect equilibrium is a function of
oxygen partial pressure. Additionally, a serious problem was
brought about by coexisting CO2 in the gas feed, which
significantly suppressed the NO absorption ability.
1.7.5. Zeolites
Since the pioneering work of Iwamoto (1990), zeolite
studies have been extensively reported in the literature.
Zhang et al. (1993) have examined the uptake of NO on
ZSM-5 zeolites exchanged with various metal ions, at 273 K
and with no other gases present but He. Zeolites exchanged
with alkaline earth, Ce, La, or Na ions, take up NO only
weakly. The highest sorption capacity was measured with
the cations Co, Cu, Ni, and Mn, in this order (Table 7).
Zeolites exchanged with transition metals ions can sorb
NO in both reversible (as NO+) and irreversible (by forming
NO2+, NO2 and NO3+) way (Iwamoto et al., 1981). Most of
NO adsorption is reversible. However, the amount of NO is
also influenced by zeolite structure. The sorption of NO per
ion on different Cu-exchanged zeolites decreases in the
order (Arai and Machida, 1994):
ZSM  5Noffretite=erioniteNmordeniteNLNferrieriteNY
with a maximum of 0.8 mol NO per mol Cu on ZSM-5. This
order is consistent with that of the increment of the Al
content in zeolites. The uptake of NO per ion is not varied
with the exchange level in Cu-ZSM-5. Therefore, maximal
uptake of NO is yielded with the maximal exchange level of
Cu-ZSM-5. The maximal amount of metal ions loaded on
the zeolite is determined by the Al-content of the framework. According to experimental results, an optimal Si/Al
ratio exists for the uptake of NO and it is in the range of 10

20. At lower values, the uptake of NO per Cu-ion is low,


and, at higher values, the maximal metal loading is low.
The uptake of NO on several exchanged zeolites is
screened at 373473 K and with 10% NO in He. The
maximal NO-uptake is ca. 13 mg g1 on Cu-ZSM-5 and
Cu-mordenite and 9.5 mg g1 on Co-Beta at 373 K.
However, tests under more realistic conditions (NOx
concentration around 1000 ppm, 5% O2, and adequate
space velocity) showed only low NO conversion (Truex et
al., 1992). Preadsorbed NO2 and SO2 reduce significantly
the irreversible sorption of NO. CO reduces the capacity and
water has a very detrimental effect: a maximal uptake of 2.6
mg NO g1 was found on Cu-mordenite and 3.2 mg NO
was sorbed per gram of Ba-mordenite at 373 K. NO and

Table 7
Adsorption capacities on cation exchanged ZSM-5
Cation

Cation content
(wt.%)

Degree of
exchange (%)

NO-uptake
(mol g1)

Transitions metals
Co
3.1
Cu
5.9
Ni
2.4
Mn
4.2
Ag
10.9
Fe
2.1
Cr
0.9
Zn
3.8

90
157
68
127
90
62
41
96

28.4
25.7
10.3
8.4
5.3
4.8
2.1
2.0

Alkaline earth metals, others


Ca
1.3
Sr
5.4
Ba
6.4
Mg
0.7
Ce
0.43
La
0.40
Na
2.8

54
105
80
46
8
7
100

4.5
3.9
3.9
1.2
1.0
0.6
0.2

NO concentration: 0.1% in He; temperature 273 K (Zhang et al., 1993).

460

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

water are sorbed, at least partially, on the same sites (Li and
Armor, 1993a,b; Li et al., 1993). Nevertheless, the presence
of oxygen enhances significantly the adsorption of NO via a
N2O3 compound on NaY. Very interesting results were
published by Sultana et al. (2000) when a mixture of
oxidized exhaust gas composed of helium with 500 ppm
NO, 500 ppm NO2, 5% water, 10% O2, and 5% CO2 was
conducted over a bed of NaY zeolite at 423 K. At these
conditions, NO and NO2 are simultaneously adsorbed as
N2O3 by displacing three water molecules. This behaviour is
temperature-dependent and has a maximum at 523 K.
Reducing performance of zeolites using hydrocarbons or
ammonia is also interesting. The Cu-ZSM-5 zeolites showed
increased activity in the reduction of NOx under fuel-lean
conditions when hydrocarbons were added to the feed
(Adelmann et al., 1994). Martens et al. (1998) studied the
reduction of NO2 to N2 at different temperatures over
ferrierite (FER) and mordenite (MOR) zeolites loaded with
5 wt.% of transition metal ions: Ag, Co, Cu, and 1 wt.% Pt.
The silver zeolites are highly actives at all temperatures, in
particular, the Ag/H-FER catalyst transforms more than
60% of the NO2 into N2 between 523 and 723 K. No
deactivation below 873 K was detected even in the presence
of water and, sometimes, 150 ppm of sulfur dioxide.
Because silver cations in zeolites have a low tendency to
undergo hydrolysis and silver oxide particles are thermally
stable, this remarkable catalytic stability is a unique feature
of Ag zeolites. The limiting factor is the hydrothermal
stability of the Brbnsted acid sites of the zeolite. Mechanical
mixtures of platinum ferrierite and silver mordenite, and of
platinum chabazite and silver mordenite, exhibit a synergetic effect on N2 formation in the catalytic reduction of NO
with isooctane and octane in presence of typical oxygen and
water concentrations of fuel-lean exhaust gas. The synergism is explained by the selective NO into NO2 oxidation in
the small pore platinum zeolite and the NO2 reduction in the
large pore silver zeolite (Martens et al., 2001).
The high selectivity of NH3 for NO reduction, combined
with the enhanced reaction rate in presence of O2, makes
NH3 the more preferred reducing agent over zeolite or
molecular sieve catalysts. In zeolite catalysts, the NOx
conversion using NH3 is independent on O2 concentration
over a wide range. The large internal surface area of the
zeolites has the advantage of having large NH3 storage
capacity. As an alternative to noble metals, transition metals
like vanadium (V) and copper (Cu) were the doping elements
for mordenite and Y-zeolites. Among these catalysts, Cudoped Y-zeolite (CuY) showed the best activity with more
than 90% conversion of NO within the temperature range
573653 K. A further increase in temperature, however,
resulted in a strong decrease of NO conversion because of
rapid increase in NH3 oxidation. Presence of Fe3+ cations
were reported to make FeY catalyst active for SCR of NO
by NH3 both in presence and in absence of O2 (Amiridis et
al., 1993). The results of application of an oxidation catalyst
and SCR technology to a stationary diesel and dual-fuel

(natural gasdiesel) engine are presented by Bittner et al.


(1992). The reduced nonmethane hydrocarbons conversion
performance over oxidation catalyst is attributed to the
presence of higher concentrations of light hydrocarbons,
identified as ethane and propane, in the exhaust, whereas for
the diesel mode, the hydrocarbons species in the exhaust are
of higher molecular weight than propane, which are easier to
be converted over the catalyst. The SCR catalyst reduces
NOx emissions up to 90% for both diesel and dual-fuel mode
of operation at a space velocity of 15 000 h1 and at NH3/
NOx mole ratios of 1.0 or higher.
Comprehensive reviews of the literature for the
selective catalytic reduction of NO by hydrocarbons have
been published by Amiridis et al. (1996) and Traa et al.
(1999). They are in agreement with the fact that although
Cu-ZSM-5 is an active catalyst for the selective reduction
of NO by hydrocarbons in the presence of excess O2,
there are significant limitations to its utilisation for
commercial applications, primarily due to the inhibiting
effect of water, its low hydrothermal stability, and reported
vulnerability to poisoning by SO2, under realistic exhaust
compositions. A recent report of the production of
hydrogen cyanide at levels of 2030 ppm over this type
of catalysts under NO-hydrogen SCR conditions further
undermines their practical applicability. The utilisation of
Co-ZSM-5 appears to be more appealing, especially since
it can activate methane as the reducing agent. Unfortunately, Co-ZSM-5-based catalysts also suffer from deactivation in the presence of water and SO2, although, in
general, they are more stable than Cu-ZSM-5. Noble
metal-based catalysts appear to overcome most of these
problems, but face additional ones related to their narrow
window of operability and high selectivities for N2O
formation and the oxidation of SO2 to SO3.
1.7.6. Carbonaceous materials
NO is sorbed on some specific carbonaceous materials
with the highest known capacities per weight of adsorbent.
These materials include carbon fibers and carbon nanotubes. Whilst carbon fibbers consist of graphite sheets
arranged in various orientations along the fibber axis with
many edges (disposed like a fishbone), both single-walled(SWNT) and multiwalled-(MWNT) carbon nanotubes
show a concentric wall structure consisting of ordered
graphite platelets. The morphology and the size of the
above materials, especially because they do present huge
length vs. diameters, can play a significant role in catalytic
applications.
For active carbon fibbers, modified by iron or copper
oxides on their surface, the NO-sorption ranges up to 320
mg g1 at 303 K. Even higher values are reported for NO2.
Maximal NOx desorption occurs between 393 and 443 K.
NO is fully recovered by heating above 473 K. Sorption is
thought to proceed via an enhanced physisorption mechanism comprising dimerisation of NO and a kind of
liquefaction in the micropores (supercritical micropore

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

461

also be adsorbed on the carbon nanotubes, but the adsorption


rate and amount are less than those for NO (Long and Yang,
2001).
Although this carbon material has attractive performances, still two possible drawbacks are not yet cleared out:
the oxidative stability (and hence, regenerability) and the
capture of SO2.

Fig. 11. Schematic model of the NO dimmer in the slit-shaped ACF (Active
Carbon Fibber) micropores (Kaneko and Kobayashi, 1988).

filling). On unmodified active carbon fibber, NO is sorbed


well, but in far lower amounts than with the metal oxide
coating, and it is released at lower temperatures (Kaneko,
1987; Kaneko and Kobayashi, 1988). The mechanism of
micropore filling of supercritical NO is not clear at the
present stage. However, the presence of highly dispersed Fe
species is obviously essential for this phenomenon to take
place. From the temperature dependence of the isotherm,
this type of NO adsorption seems to possess both
chemisorption and physical adsorption characteristics.
Actually, gaseous NO molecules produce strong chemisorption on Fe2O3 highly dispersed around the entrance of the
slit-shaped micropores of activated carbon fibbers. After
adsorption step, almost all the absorbed NO molecules
migrate to fill the micropores as (NO)2 as presented in Fig.
11. This intermolecular interaction plays a key role in the
micropore filling.
Surprisingly, the modified active carbon fibber sorbents
have a high selectivity for NOx : in a mixture with 300 ppm
NO, the presence of O2, SO2, H2O, and CO2 as co-gases did
not prevent an almost complete removal of NO. However,
these materials present several problems like their low
oxidative stability: first traces of CO2, by oxidation of the
sorbent, are released at 383433 K (Kaneko, 1987). An
uptake amount of 78 mg of NO g1 was observed on carbon
nanotubes at room temperature in the presence of oxygen
(5%). This adsorption is also reversible and MWNT could be
regenerated by heating to 573 K. By comparison, SO2 can

1.7.7. Heteropolyacids (HPAs)


The catalytic function of heteropoly-compounds has
attracted much attention particularly in the last two decades.
One of the remarkable characteristics is that some solid
HPAs (group A, hydrogen forms included) sorb easily a
large quantity of polar or basic molecules such as alcohols
and nitrogen bases in the solid bulk (Okuhara et al., 1996).
Its structure, as described by Brown et al. (1977), comprises
primary units called Keggin anions (XM12O403), which
consists of a central XO4 (X=heteroatom or central atom)
surrounded by 12 MO6 octahedra (M=addenda atom). The
Keggins polyanions are linked together by H+(H2O)2
bridges to form the so-called secondary structure (Shikata
et al., 1997; Bardin and Davis, 2000).
Owing to the flexible nature of the secondary structure of
the acid forms of group-A salts, polar molecules are readily
absorbed into the bulk, by substituting water molecules and/
or by changing the interdistance between polyanions: this is
not a diffusion process in micropores. Heteropolycompounds which have absorbed a significant amount of polar
molecules behave in a sense like a concentrated solution
(this state is called bpseudo-liquid phaseQ). Therefore,
reactions on HPAs could proceed as surface- or bulk-type
phase as presented in Fig. 12.
Surface-type reactions proceed like in the case of ordinary
solid catalyst. However, for bulk-type (I) reactions, whole
protons in the bulk take part in catalysis and the environment
of the reactants are quite different. The mechanism bulk-type
(II) is that of a catalyst in which redox cycles of the whole
catalyst bulk are involved in catalysis by virtue of rapid
migration of protons and electrons (Misono et al., 1988).
Two different teams had originally reported sorption of
nitrogen oxides on HPAs: Yang and Chen (1994), and

Fig. 12. Schematic of heteropolyacidreactants possible interactions (Misono et al., 1988).

462

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

Belanger and Moffat (1995a). Belanger and Moffat made


some studies to analyse the sorption of NO2 on those kinds
of metaloxygen clusters compounds analysing the effect of
changes in the elemental compositions of the anions: HPW,
HPMo, HSiW. The results showed that the quantities of NO2
lost from the gas phase follow the order HPWNHSiWN
HPMo, and the quantity of NO2 retained on HPW is strongly
dependent on temperature: increasing from 298 K reaches a
maximum in the range from 423 to 573 K, and decreases to
small values from 773 to 873 K. Supplementary experiments
showed that the maximum quantity of NO taken up by the
solid is approximately equal to those of NO2 (Belanger and
Moffat, 1996, 1998).
Chen and Yang (1995) made some experiments to
analyse the linkage of NO on HPW. Based on TPD,
XRD, and IR results, they could propose several possible
linkage structures for the interaction between those compounds, with 3 NO molecules that substituted 6 H2O
structural molecules in HPW (it was proposed that the NO
was protonated as NOH+ in the pseudo-liquid phase). If the
reaction goes to completion, this represents 3 wt.% NO
absorbed by HPW. The bcc lattice constant decreased from
12.2 to 11.7 2 because of this reaction. They also showed
that the presence of CO2 and SO2 do not affect the NO
sorption rate into the HPW. In that way, an absorption
mechanism was proposed:

Because NO sorption is a bulk reaction, those results


indicate that CO2 and SO2 do not substitute the H2O
linkages in the secondary structure as does NO. Those
results also indicate that CO2 and SO2 do not influence the
diffusion rates of NO in the bulk substitution. In contrast,
they found a strong effect over sorption rate of NH3
(decreasing) and oxygen (increasing).
Hodjati et al. (2000a,b,c, 2001a,b) used a different
approach in order to elucidate the sorbing mechanism and
to understand the interaction between NOx (NO and NO2)
and HPW. Their first results, based on XRD, TGA, and
FTIR, put in evidence the thermal fragility of the HPW and
the degradation of structural properties because of the loss of
H+(H2O)2 bridges which was accompanied by a loss of the
capacity of sorption. From those results, it was possible to
define that the operation temperature limit for HPW is 673 K,
which is within the range of temperatures for practical
applications. One of the key factors for the NOx absorption/
desorption process is the presence of water. The absence of
water during the sorption phase leads to a reduction in the
desorption temperature of ca. 293 K. This is an indicative of
a competition between NO/NO2 and water for the absorption

sites. In the absence of water from the phase of desorption,


no desorption is observed at all. However, as soon as water is
added to the mixture, desorption of NOx becomes immediately possible (at temperatures below 393 K).
From all those results presented above, it was possible to
propose a mechanism for absorption and desorption of NOx
on HPW as described in Fig. 13.
The sorption proceeds via a reversible substitution of
water molecules by NOx to form an [H+(NO2,NO+)]
complex. Substitution results from co-absorption of NO
and NO2 in an equimolar ratio. Once substitution is
achieved, this complex maintains the cohesion of the
secondary unit. The presence of H+ in the structure is
essential to the mechanism because the complete substitution by Na+ led to a solid with no NOx sorption capacity.
For practical applications, it is important to improve the
physical properties of HPW (e.g., mechanic and thermal
resistance). This could be reached by depositing HPW on a
suitable support while preserving its chemical properties
(absorption capacity). Dispersing HPA on solid supports is
important for catalytic application because the specific
surface of unsupported HPA are usually low (although
interstitial voids are created by the terminal oxygen atoms
linking the hydrated protons because these are not interconnected the resulting solid acid have low BET (N2)
surface areas 110 m2 g1). In general, HPW strongly
interact with supports at low loading levels, while the bulk
properties of HPW prevail at high loading levels.
It is possible to find in the literature a number of studies
focused on improving the thermal stability of HPW via
deposition of the species on different supports. Several
authors found that basic solids like Al2O3 and MgO tend to
decompose heteropolyacids (Cheng and Luthra, 1988; Rao
et al., 1989). There is a big influence of peripheral metal
elements of the anion and of central heteroatom in catalytic
properties of heteropolyacids when using acid supports as
SiO2 (Belanger and Moffat, 1995b; Kozhevnikov et al.,
1996; McCormick et al., 1998; Dias et al., 2003). Marme et
al. (1998) presents additional results for either lamellar and
mesoporous silicates compounds (clays and kenyaite).
Lopez-Salinas et al. (2000) supported HPW on zirconia
and observed that Keggin structures could be retained on the
support from 20 wt.% HPW on ZrO2. Similar values where
obtained when using mesoporous pure-silica molecular
sieve MCM-41 and pillared layered (Kozhevnikov et al.,
1995; Hu et al., 1996). Edwards et al. (1998) analysed the
surface structure of HPW supported on extruded, pressed,
and powdered TiO2. H3PW12O40 supported on certain active
carbons can firmly entrap HPW (Schwegler et al., 1991,
1992). HPW supported on multiwalls carbon nanotubes
leads to an improvement of both absorption capacity and
efficiency at high temperature (Gomez-Garca et al., 2004a).
Enhanced catalytic activity of HPW was found when they
were supported on a strongly acidic ion-exchange resin,
Amberlyst-15 (Baba and Ono, 1986). In general, as
evidenced in those above studies, the catalytic activity of

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

463

Fig. 13. Mechanism of NOx absorptiondesorption on HPW (Hodjati et al., 2000a).

supported HPW is related to the type of carrier, the loading


and conditions of pre-treatment. In relation to supporting
procedure, it is well known that a particle of metal oxide in
suspension tends to polarise and to be electrically charged
(e.g., in an acid medium the particle is positively charged
and in a basic medium the reverse is true, and, from this
behaviour, it is possible to define the isoelectric point for a
solid given, IEPS (Brunelle, 1978)). Various oxides were
tried out as supports with the goal of wash coating HPW on
a monolith. The aim is to provide a material strongly
anchored to the support that retains high NOx sorption
capacity. The supports were classified according to their
isoelectric point. The results for the case of 50%:50% wt/wt
ratio between HPWsupport are reported in Table 8.
Referring to the isoelectric point of a solid (Brunelle,
1978), if the support is moderate to strongly basic (e.g.,
Al2O3 and MgO), the interaction with HPW is too strong
resulting in a loss of crystallinity (XRD) of the heteropolyacid with a complete degradation of the absorption
properties. If the support is strongly acidic (e.g., SiO2), the
absorption remains possible and the XRD structure of the
heteropolyacid exists, but the anchorage is not secured and
the submission of a monolith to a dry air flow will lead to
the loss of HPW in the stream. In the case of medium acidity
(e.g., TiO2, SnO2), the structural properties are retained
(XRD) and the absorption capacity is high. Equally, the
anchorage is strong and the monolith resists perfectly to the
blow test. In the case of SnO2, the value of NOx trapped is
equal to 36 mg NO2 gHPW1, but in the case of TiO2 is
Table 8
Influence of support isoelectric point on HPW structure and absorption
capacity (all supports were impregnated with 50% wt/wt HPW; Hodjati et
al., 2001b)
Support

Isoelectric point
(IEPS)

XRD
structure

Capacity
(mg NO2 gHPW1)

MgO
Al2O3
ZrO2
TiO2
SnO2
SiO2

12.112.7
7.09.0
6.7
6.0
5.5
1.02.0

No
No
No
Yes
Yes
Yes

0
0
34
39
36
17

particular because the value of 39 mg NO2 gHPW1 matches


perfectly the one of bulk HPW, which proves that, with this
support, the properties of the heteropolyacid are entirely
preserved (Thomas et al., 2004). However, as evidenced
experimentally, IEPS condition is necessary but not
sufficient in itself. The specific surface and HPW loading
must also be the parameters to take into consideration.
The possibility of reducing the NOx absorbed in HPW
was recently studied by using a system containing HPW-Pt
(Vaezzadeh et al., 2002). Reduction capacities were
measured under representative exhaust lean gas conditions
and by periodic switching of lean and rich (CO, H2, and/or
C3H6) mixtures. The presence of Pt leads to an increase of
the efficiency of absorption from 50% to 85% with a partial
reduction of NO2 into NO and N2. The reduction behaviour
of this system was improved by supporting HPW-Pt on Ce
Zr-based materials. We have demonstrated that H2, CO, and
CH4 separately or a mixture of two of them could reduce
NOx into nitrogen (30% to 70%) at moderate temperature
(523 K) after periodic switching between lean-rich mixtures.
These kind of catalytic systems are highly selective because,
in any case, neither N2O nor NH3 formation were detected
(Gomez-Garca et al., in press(b)).
1.7.8. Conclusions on bNOx -sorbing catalytic materialsQ
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. Therefore, it is possible to say that NOx
sorption mechanisms could be classified into five general
types:
(i)

(ii)

Adsorbed surface species (NO, NO+, NO2, N2O3)


partially desorb as NOx with increased temperature.
For the remaining part, they decompose into nitrogen
and oxygen at further increase of temperature (e.g.,
with Cu-ZSM-5 or some perovskites). The adsorption
capacity is proportional to the surface area of the
adsorbent.
Formation of nitrate or nitrite salts on basic oxides.
This reaction occurs with sufficient rate only at

464

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

elevated temperature and it is not limited to the


surface area of the sorbents. At higher temperatures,
the salts are decomposed to their oxides with
increased formation of NOx and mostly O2.
(iii) A third sorption mechanism called intercalation has
been reported for layered oxides (e.g., cuprates,
titanates, hydrotalcites). NOx molecules are intercalated between the layers without further modification
of crystal structure.
(iv) A fourth mechanism comprises carbon materials. This
sorption type, however, is less relevant for the lean
exhaust gas application because it is only operative in
a small and low temperature range.
(v) A fifth mechanism involves the bulk absorption of
molecules by forming solvates as in the case of
heteropolyacids. In that way, polar molecules undergo
catalytic reactions not only on the surface but also in
the bulk of the crystalline HPA. Thus, solid HPAs
behave like highly concentrated solutions, i.e., nearly
all the HPA protons, not only the surface proton sites
participate in the catalytic reaction.
1.8. Conclusions
Several challenges have to be faced when trying to solve
the problem of NOx pollution with a catalytic system:
selectivity, operational temperature, and poisoning. The
fundamental points are the sorption and desorption capacity
of sorbing material in a complex mixture of gases during
short periods. The development of practical regeneration
processes, such as oxidationregeneration cycles have
opened novel applications for catalytic materials which
can integrate dual functions. Thus, the presence of noble
metals (synonymous for platinum group metals) is essential
during the reduction process due to their ability to activate
reductant agents and their relative resistance to poisoning.
Finally, overcoming the problems of preserving sorbents
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.

References
Abad J, Bohme O, Roman E. Dissociative adsorption of NO on TiO2 (1 1 0)
argon ion bombarded surfaces. Surf Sci 2004;549:134 42.
Adelmann BJ, Lei GD, Schatler WMH. Coadsorption of nitrogen
monoxide and nitrogen dioxide in zeolithe de-NOx . Catal Letters
1994;28:119 30.
Amiridis MD, Puglisi F, Dumesic JA, Millmann WS, Topsbe NY. Kinetic
and infrared spectroscopic studies of FeY zeolites for the selective
catalytic reduction of nitric oxide by ammonia. J Catal 1993;142:
572 84.
Amiridis MD, Zhang T, Farrauto RJ. Selective catalytic reduction of nitric
oxide by hydrocarbons. Appl Catal, B Environ 1996;10:203 27.
Arai H, Machida M. Removal of NOx through sorptiondesorption cycles
over metal oxides and zeolites. Catal Today 1994;22:97 109.

Armor JN. NOx /hydrocarbon reactions over gallium loaded zeolites: a


review. Catal Today 1996;31:191 8.
Armor JN. Catalytic solutions to reduce pollutants. Catal Today
1997;38:163 7.
Baba T, Ono Y. Heteropolyacids and their salts supported on acidic ionexchange resin as highly active solid-acid catalysts. Appl Catal
1986;22:321 4.
Balducci G, Kaspar J, Fornasiero P, Graziani M, Saiful Islam M, Gale
JD. Computer simulation studies of bulk reduction and oxygen
migration in CeO2ZrO2 solid solutions. J Phys Chem, B 1987;
101:1750 3.
Balducci G, Kaspar J, Fornasiero P, Graziani M, Saiful Islam M. Surface
and reduction energetics of the CeO2ZrO2 catalysts. J Phys Chem, B
1998;102:557 61.
Bardin BB, Davis RJ. Effect of water on silica-supported phosphotungstic
acid catalysts for 1-butene double bond shift and alkane skeletal
isomerization. Appl Catal, A Gen 2000;200:219 31.
Belanger R, Moffat JB. A comparative study of the adsorption and reaction
of nitrogen oxides on 12-tungstophosphoric, 12-tungstosilicic, and 12molybdophosphoric acids. J Catal 1995a;152:179 88.
Belanger R, Moffat JB. Interaction of NO and NO2 on 12-tungtophosphoric
acid. Catal Letters 1995b;32:371 8.
Belanger R, Moffat JB. The interaction of nitrogen oxides with metaloxygen cluster compounds (heteropolyoxometalates). J Mol Catal, A
Chem 1996;114:319 29.
Belanger R, Moffat JB. The sorption and reduction of nitrogen oxides by
12-tungstophosphoric acid and its ammonium salt. Catal Today
1998;40:297 306.
Bhattacharyya S, Das RK. Catalytic control of automotive NOx : a review.
Int J Energy Res 1999;23:351 69.
Bittner RW, Aboujaoude FW. Catalytic control of NOx , CO and NMHC
emissions from stationary diesel and dual-fuel engines. Trans ASME, J
Engng Gas Turbines Power 1992;114:597 601.
Bosch H, Janssen F. Formation and control of nitrogen oxides. Catal Today
1988;2:369 79.
Bogner W, Kr7mer M, Krutzsch B, Piscinger S, Voigtl7nder D, Wenniger
G, et al. Removal of nitrogen oxides from the exhaust of a lean-tune
gasoline engine. Appl Catal, B Environ 1995;7:153 71.
Breen JP, Marella M, Pistarino C, Ross JRH. Sulfur-tolerant NOx storage
traps: an infrared and thermodynamic study of the reactions of alkali
and alkalineearth metal sulfates. Catal Letters 2002;80:123 8.
Brown GM, Noe-Spirlet MR, Busing MR, Levy HA. Dodecatungstophos3
phoric acid hexahydrate, (H5O+2 )3(PW12O40
) The true structure of
Keggins bpentahydrateQ from single-crysral X-ray and neutron diffraction data. Acta Crystallogr, B 1977;33:1038 46.
Brunelle JP. Preparation of catalysts by metallic complex adsorption on
mineral oxides. Pure Appl Chem 1978;50:1211 29.
Burch R, Breen JP, Meunier FC. A review of the selective reduction of NOx
with hydrocarbons under lean-burn conditions with non-zeolitic oxide
and platinum group metal catalysts. Appl Catal, B Environ
2002;39:283 303.
Busca G, Lietti L, Ramis G, Berti F. Chemical and mechanistic aspects of
the selective catalytic reduction of NOx by ammonia over oxide
catalysts: a review. Appl Catal, B Environ 1998;18:1 36.
Chen N, Yang RT. Activation of nitric oxide by heteropoly compounds:
structure of nitric-oxide linkages in tungstophosphoric acid with keggin
units. J Catal 1995;157:76 86.
Cheng WC, Luthra NP. NMR study of the adsorption of phosphomolybdates on alumina. J Catal 1988;109:163 9.
CITEPA. Emissions dans lair en France Metropolitaine. Avril; 2002.
Cohen RC, Murphy JG. Photochemistry of NO2 in Earths stratosphere:
constraints from observations. Chem Rev ASAP Article 2003;103:
4985 98.
Colon G, Pijolat M, Valdivieso F, Vidal H, Kaspar J, Finocchio E, et al.
Surface and structural characterization of Cex Zr1x O2 CEZIRENCAT
mixed oxides as potential three-way catalyst promoters. J Chem Soc,
Faraday Trans 1998;94:3717 26.

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467


Colon G, Valdivieso F, Pijolat M, Baker RT, Calvino JJ, Bernal S. Textural
and phase stability of Cex Zr1x O2 mixed oxides under high temperature
oxidising conditions. Catal Today 1999;50:271 84.
Cordatos H, Gorte RJ. CO, NO and H2 adsorption on ceria-supported Pd. J
Catal 1996;159:112 8.
Courson C, Khalfi A, Mahzoul H, Hodjati S, Moral N, Kiennemann A, et al.
Experimental study of the SO2 removal over a NOx trap catalyst. Catal
Commun 2002;3:471 7.
Cutrufello MG, Ferino I, Monaci R, Rombi E, Solinas V. Acidbase
properties of zirconium, cerium and lanthanum oxides by calorimetric
and catalytic investigation. Top Catal 2002;19:225 40.
Daturi M, Finocchio E, Binet C, Lavalley JC, Fally F, Perrichon V, et al.
Reduction of high surface area CeO2ZrO2 mixed oxides. J Phys Chem,
B 2000;104:9186 94.
Daturi M, Bion N, Saussey J, Lavalley JC, Hedouin C, Seguelong T, et al.
Evidence of a lacunar mechanism for deNOx activity in ceria-based
catalysts. Phys Chem Chem Phys 2001;3:252 5.
Dias JA, Caliman E, Dias SCL, Paulo M, Thyrso A, de Souza CP. Preparation
and characterization of supported H3PW12O40 on silica gel: a potential
catalyst for green chemistry processes. Catal Today 2003;85:39 48.
Edwards JC, Thiel C, Benac B, Knifton JF. Solid-state NMR and FT-IR
investigation of 12-tungstophosphoric acid on TiO2. Catal Letters
1998;51:77 83.
Eguchi K, Watabe M, Machida M, Arai H. Selective removal of NO by
absorption in mixed oxide catalysts. Catal Today 1996;27:297 305.
Erisman JW, Grennfelt P, Sutton M. The European perspective on nitrogen
emission and deposition. Environ Int 2003;29:311 25.
Fally F, Perrichon V, Vidal H, Kaspar J, Blanco G, Pintado JM, et al.
Modification of the oxygen storage capacity of CeO2ZrO2 mixed
oxides after redox cycling aging. Catal Today 2000;59:373 86.
Fenimore CP. Formation of nitric oxide in premixed hydrocarbon flames.
13th Symp Combustion, p. 373 80.
Fornasiero P, Di Monte R, Rao GR, Kaspar J, Meriani S, Trovarelli A,
et al. Rh-loaded CeO2ZrO2 solid-solutions as highly efficient
oxygen exchangers: dependence of the reduction behaviour and the
oxygen storage capacity on the structural-properties. J Catal 1995;
151:168 77.
Fornasiero P, Balducci G, Di Monte R, Kaspar J, Sergo V, Gubitosa G, et al.
Modification of the redox behaviour of CeO2 induced by structural
doping with ZrO2. J Catal 1996;164:173 83.
Forzatti P. Environmental catalysis for stationary applications. Catal Today
2000;62:51 65.
Fritz A, Pitchon V. The current state of research on automotive lean NOx
catalysis. Appl Catal, B Environ 1997;13:1 25.
Gandhi HS, Shelef M. The adsorption of nitric oxide and carbon monoxide
on nickel oxide. J Catal 1972;24:241 9.
Gandhi HS, Shelef M. The adsorption of nitric oxide on copper oxides. J
Catal 1973;28:1 7.
Garin F. Mechanism of NOx decomposition. Appl Catal, A Gen
2001;222:183 219.
Gomez-Garca MA, Pitchon V, Kiennemann A, Corrias M, Kalck Ph,
Serp Ph. Sorption-desorption of NOx from a lean gas mixture on
H3PW12O40d 6H2O supported on carbon nanotubes. Top Catal
2004a;30/31:229 34.
Gomez-Garca MA, Pitchon V, Kiennemann A. The removal of NOx
from lean exhaust gas using storage/reduction on H3PW12O40d 6H2O
supported on Cex Zr4x O8. Environ Sci Technol 2004b [in press].
Gutberlet H, Schallert B. Selective catalytic reduction of NOx from coal
fired power plants. Catal Today 1993;16:207 35.
Hadjiivanov KI. Identification of neutral and charged Nx Oy surface species
by IR spectroscopy. Catal Rev, Sci Eng 2000;42:71 144.
Heck RM. Catalytic abatement of nitrogen oxidesstationary applications.
Catal Today 1999;53:519 23.
Hodjati, S. The`se de doctorat. Universite Louis Pasteur. Strasbourg; 1998.
Hodjati S, Bernhardt P, Petit C, Pitchon V, Kiennemann A. Removal of
NOx : Part I Sorption/desorption processes on barium aluminate. Appl
Catal, B Environ 1998a;19:209 19.

465

Hodjati S, Bernhardt P, Petit C, Pitchon V, Kiennemann A. Removal of


NOx : Part II Species formed during the sorption/desorption processes
on barium aluminates. Appl Catal, B Environ 1998b;19:221 32.
Hodjati S, Petit C, Pitchon V, Kiennemann A. Absorption/desorption of
NOx process on perovskites Nature and stability of the species formed
on BaSnO3. Appl Catal, B Environ 2000a;27:117 26.
Hodjati S, Petit C, Pitchon V, Kiennemann A. The mechanism of the
selective NOx adsorption on 12-tungstophosphoric acid hexa-hydrate.
Stud Surf Sci Catal 2000b;130:1265 70.
Hodjati S, Vaezzadeh K, Petit C, Pitchon V, Kiennemann A.
Absorption/desorption of NOx process on perovskites: performances
to remove NOx from a lean exhaust gas. Appl Catal, B Environ
2000c;26:5 16.
Hodjati S, Petit C, Pitchon V, Kiennemann A. Removal of NOx from a lean
exhaust gas by absorption on heteropolyacids: reversible sorption of
nitrogen oxides in H3PW12O40d 6H2O. J Catal 2001a;197:324 34.
Hodjati S, Vaezzadeh K, Petit C, Pitchon V, Kiennemann A. The
mechanism of the selective NOx sorption on H3PW12O40d 6H2O
(HPW). Top Catal 2001b;1617:151 5.
Hu C, He Q, Zhang Y, Liu Y, Zhang Y, Tang T, et al. Synthesis of new types
of polyoxometallate pillared anionic clays. J Chem Soc, Chem
Commun 1996;2:121 3.
Iwamoto M. Catalytic decomposition of nitrogen monoxide. Stud Surf Sci
Catal 1990;54:121 43.
Iwamoto M, Hamada H. Removal of nitrogen monoxide from exhaust gases
through novel catalytic processes. Catal Today 1991;10:57 71.
Iwamoto M, Yokoom S, Sakai K, Kagawa S. Catalytic decomposition of
nitric oxide over copper(II)-exchanged, Y-type zeolites. J Chem Soc,
Faraday Trans 1981;77:1629 39.
Janssen F. Emission control from stationary sources. In: Jansen FJJ, van
Santen RA, editors. Environmental catalysis. Imperial College Press;
p. 234 93.
Kaneko K. Anomalous micropore filling of nitric oxide on alpha-iron
hydroxide oxide-dispersed activated carbon fibers. Langmuir
1987;3:357 63.
Kaneko K, Kobayashi A. The dimer state of NO in micropores of Cu(OH)2dispersed activated carbon fibres. J Chem Soc, Faraday Trans I
1988;84:1795 804.
Kapteijn F, Mirasol JR, Moulijn JA. Heterogeneous catalytic decomposition
of nitrous oxide. Appl Catal, B Environ 1996;9:25 64.
Karlsson HT, Rossenberg HS. Flue gas denitrification Selective catalytic
oxidation of nitric oxide to nitrous oxide. Ind Eng Chem Process Des
Dev 1984;23:808 14.
Kato K, Nohira H, Nakanishi K, Igushi S, Kihara T, Muraki H. 1993.
Europ. Patent. 0573672 A1.
Kermikri I. Lair des villes rend bien malade. La Recherche 1995;279:884 8.
Kiel JHA, Prins W, van Swaaij WPM. Performance of silica-supported
copper oxide sorbents for SOx /NOx -removal from flue gas: I Sulphur
dioxide absorption and regeneration kinetics. Appl Catal, B Environ
1992;1:13 39.
Kiennemann A, Martens JA, Kasemo B, Chaize E, Webster D, Krutzsch B, et
al. Reduction of NOx in lean exhaust by selective NOx -recirculation
(SNR-Tecnique). Part I. System and decomposition process. SAE. 1998;
no. 982592.
Kiennemann A, Petit C, Roger AC, Pitchon V. Perovskites: a versatile
material in heterogeneous catalysis. Curr Top Catal 2002;3:147 60.
Kikuchi E, Ogura M, Terasaki I, Goto Y. Selective reduction of nitric oxide
with methane on gallium and indium containing H-ZSM-5 catalysts:
formation of active sites by solid-state ion exchange. J Catal
1996;161:465 70.
Kozhevnikov IV, Sinnema A, Jansen RJJ, van Bekkum H. New acid catalyst
comprising heteropoly acid on a mesoporous molecular sieve MCM-41.
Catal Letters 1995;30:241 52.
Kozhevnikov IV, Kloetstra KR, Sinnema A, Zandbergen HW, van Bekkum
H. Study of catalysts comprising heteropoly acid H3PW12O40 supported
on MCM-41 molecular sieve and amorphous silica. J Mol Catal, A Chem
1996;114:287 98.

466

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467

Long RQ, Yang RT. Carbon nanotubes as a superior sorbent for nitrogen
oxides. Ind Eng Chem Res 2001;40:4288 91.
Lopez-Salinas E, Hernandez-Cortez JG, Schifter I, Torres-Garcia E,
Navarrete J, Gutierrez-Carrillo A, et al. Thermal stability of 12tungstophosphoric acid supported on zirconia. Appl Catal, A Gen
2000;193:215 25.
Li Y, Armor JN. Catalytic reduction of nitrogen oxides with methane in the
presence of excess oxygen. Appl Catal, B Environ 1992;1:L31 40.
Li Y, Armor JN. Simultaneous, catalytic removal of nitric oxide and nitrous
oxide. Appl Catal, B Environ 1993a;3:55 60.
Li Y, Armor JN. Selective catalytic reduction of NOx with methane over
metal exchange zeolites. Appl Catal, B Environ 1993b;2:239 56.
Li Y, Battari PJ, Armor JN. Effect of water vapor on the selective reduction
of NO by methane over cobalt-exchanged ZSM-5. J Catal 1993;
142:561 71.
Loof P, Kasemo B, Anderson S, Frestad A. Influence of ceria on the
interaction of CO and NO with highly dispersed Pt and Rh. J Catal\
1991;130:181 91.
Lyon RK, Cole JA, Kramlich JC, Chen SL. The selective reduction of SO3
to SO2 and the oxidation of NO to NO2 by methanol. Combust Flame
1990;81:30 9.
Machej T, Remy M, Ruiz P, Delmon P. Studies on the V2O5TiO2 system:
Part 1 TiO2(rutile)V2O5. J Chem Soc, Faraday Trans 1990a;86:715 23.
Machej T, Remy M, Ruiz P, Delmon P. Studies on the V2O5TiO2
system: Part 2 TiO2(anatase)V2O5. J Chem Soc, Faraday Trans
1990b;86:723 30.
Machej T, Ruiz P, Delmon P. Studies on the V2O5TiO2 system: Part 3
Monolayers of V2O5. J Chem Soc, Faraday Trans 1990c;86:731 8.
Machida M, Murakami H, Kitsubayashi T, Kijima T. NO-intercalation
properties of a double-layered cuprate, La2x Bax SrCu2O6 Dissociative
desorption of intercalated NO. Chem Mater 1997;9:135 40.
Machida M, Murakami H, Kijima T. Temperature-swing sorption/desorption cycles of nitric oxide through intercalation by double layered
cuprate. Appl Catal, B Environ 1998;17:195 203.
Marme F, Coudurier G, Vedrine JC. Acid-type catalytic properties of
heteropolyacid H3PW12O40 supported on various porous silica-based
materials. Microporous Mesoporous Mater 1998;22:151 63.
Martens JA, Cauvel A, Francis A, Hermans C, Jayat F, Remy M, et al. NOx
abatement in exhaust from lean-burn combustion engines by reduction
of NO2 over silver-containing zeolite catalysts. Angew Chem, Int Ed
1998;37:1901 3.
Martens JA, Cauvel A, Jayat F, Vergne S, Jobson E. Molecule sieving
catalysts for NO reduction with hydrocarbons in exhaust of lean
burn gasoline and diesel engines. Appl Catal, B Environ 2001;
29:299 306.
Matsumoto S. DeNOx catalyst for automotive lean-burn engine. Catal
Today 1996;29:43 5.
Matsumoto S. Catalytic reduction of nitrogen oxides in automotive exhaust
containing excess oxygen by NOx storage-reduction catalyst. Cat
Technol 2000;4:102 9.
McCormick RL, Boonrueng SK, Herring AM. In situ IR and temperature
programmed desorption-mass spectrometry study of NO absorption and
decomposition by silica supported 12-tungstophosphoric acid. Catal
Today 1998;42:145 57.
Miller JD, Bowman CT. Mechanism and modelling of nitrogen chemistry in
combustion. Pror Energy Combust Sci 1989;15:287 338.
Misono M, Inui T. New catalytic technologies in Japan. Catal Today
1999;51:369 75.
Misono M, Okuhara T, Mizuno N. Catalysis by heteropoly compounds.
Stud Surf Sci Catal 1988;44:267 78.
Mqller-Bushbaum H. The crystal chemistry of AM2O4 oxometallates. J
Alloys Compd 2003;349:49 104.
Niwa M, Furukawa Y, Murakami Y. Adsorption of nitric oxide on cerium
oxide. J Colloid Interface Sci 1982;86:260 5.
Oh SH, Eickel CC. Effects of cerium addition on CO oxidation
kinetics over alumina-supported rhodium catalysts. J Catal 1988;
112:543 55.

Okuhara T, Mizuno N, Misono M. Catalytic chemistry of heterocompounds. Adv Catal 1996;41:113 252.
Orsenigo RC, Beretta A, Forzatti P, Svachula J, Tronconi E, Bregani F,
et al. Theoretical and experimental study of the interaction between
NOx reduction and SO2 oxidation over DeNOx -SCR catalysts. Catal
Today 1996;27:15 21.
Otto K, Shelef M. The adsorption of nitric oxide on chromia supported on
alumina. J Catal 1969;14:226 37.
Otto K, Shelef M. The adsorption of nitric oxide on iron oxides. J Catal
1970;18:184 92.
Otto K, Shelef M. The adsorption of nitric oxide on platinum black. J Catal
1973;29:138 43.
Padeste C, Cant NW, Trimm DL. Reactions of ceria supported rhodium
with hydrogen and nitric oxide studied with TPR/TPO and XPS
techniques. Catal Letters 1994;28:301 12.
Panayotov D, Matyshak V, Skyarov V, Vlasenko V, Mehandjev D.
Interaction between NO and CO on the surface of CuCo2O4 spinel.
Appl Catal 1986;24:37 51.
Parvulescu VI, Grange P, Delmon B. Catalytic removal of NO. Catal Today
1998;46:233 316.
Perry RH, Green DW, Maloney JO. Perrys chemical engineering handbook. seventh edition. McGraw Hill; 1997.
Ramachandran B, Hermann RG, Choi S, Stenger HG, Lyman CE, Sale JW.
Testing zeolite SCR catalyst under protocol conditions for NOx
abatement from stationary emission sources. Catal Today 2000;
55:281 90.
Rao KM, Gobetto R, Iannibello A, Zecchina A. Solid state NMR and IR
studies of phosphomolybdenum and phosphotungsten heteropoly
acids supported on SiO2, g-Al2O3, and SiO2Al2O3. J Catal
1989;119:512 6.
Rao GR, Kaspar J, Di Monte R, Meriani R, Graziani M. NO decomposition
over partially reduced metallized CeO2ZrO2 solid solutions. Catal
Letters 1994;24:107 12.
Satterfield CH. Heterogeneous catalysis in industrial practice. second
edition. Hill7 McGraw; 1991.
Schnelle KB, Brown CA. Air pollution control technology handbook. CRC
Press; 2002.
Schwegler MA, van Bekkum H, de Munok NA. Heteropolyacids as
catalysts for the production of phthalate diesters. Appl Catal
1991;74:191 204.
Schwegler MA, van Bekkum H, de Munok NA. Activated carbon as a
support for heteropolyanion catalysts. Appl Catal 1992;80:41 58.
Shelef M. Nitric oxide: surface reactions and removal from auto exhaust.
Catal Rev, Sci Eng 1975;11:1 40.
Shelef M, McCabe RW. Twenty-five years after introduction of automotive
catalysts: what next? Catal Today 2000;62:35 50.
Shikata S, Nakata S, Okuhara T, Misono M. Catalysis by heteropoly
compounds: 32 Synthesis of methyltert-butyl ether catalyzed by
heteropolyacids supported on silica. J Catal 1997;166:263 71.
Shriver DF, Atkins PW. Inorganic chemistry. third edition. Oxford Press;
1999.
Solymosi F, Kiss J. Adsorption and reduction of NO on tin(IV) oxide
catalysts. J Catal 1976;41:202 11.
Sultana A, Loenders R, Monticelli O, Kirschhock C, Jackobs PA, Martens
JA. DeNOx of exhaust gas from lean-burn engines through reversible
adsorption of N2O3 in alkali metal cation exchanged faujasite-type
zeolites. Angew Chem, Int Ed 2000;39:2934 7.
Svachula J, Ferlazzo N, Forzatti P, Tronconi E. Selective reduction of
nitrogen oxides (NOx ) by ammonia over honeycomb selective catalytic
reduction catalysts. Ind Eng Chem Res 1993;32:1053 60.
Takahasi N, Shinjoh H, Ijima T, Suzuki T, Yamakazi K, Yokota K, et al.
Proc 1st Int Cong Env Catal, Pisa; 1995. p. 45.
Takahasi N, Shinjoh H, Ijima T, Suzuki T, Yamakazi T, Yokota K, et al. The
new concept 3-way catalyst for automotive lean-burn engine NOx
storage and reduction catalyst. Catal Today 1996;27:63 9.
Tascon JMD, Tejuca LG, Roschester CH. Surface interactions of NO and
CO with LaMO3 oxides. J Catal 1985;95:558 66.

M.A. Gomez-Garca et al. / Environment International 31 (2005) 445467


Thomas S, Vaezzadeh K, Pitchon V. Supported heteropolyacids for NOx
storage and reduction. Top Catal 2004;30/31:207 13.
Tomita A. Suppression of nitrogen oxides emission by carbonaceous
reductants. Fuel Process Technol 2001;71:53 70.
Traa Y, Burger B, Weitkamp J. Zeolite-based materials for the selective
catalytic reduction of NOx with hydrocarbons. Microp Macrop Mat
1999;30:3 41.
Trovarelli A. Catalytic properties of ceria and CeO2 containing
materials. Catal Rev, Sci Eng 1996;38:439 520.
Truex TJ, Searles RA, Sun DC. Catalysis for nitrogen oxides control
under lean burn conditions. Platinum Metals Rev 1992;36:2 10.
Vaezzadeh K, Petit C, Pitchon V. The removal of NOx from a lean exhaust gas
using storage and reduction on H3PW12O14d 6H2O. Catal Today
2002;73:297 305.
Vestreng V, Stbren E. Analysis of the UNECE/EMEP Emission Data.
MSC-W Status Report 2000. Norwegian Meteorological Institute:
Blindern, Oslo; 2000.
Viswanathan B. CO oxidation and NO reduction on Perovskite oxides.
Catal Rev, Sci Eng 1992;34:337 54.
Vogt ETC, Boot M, Dillen VAJ, Geus JW, Jansen FJJG, Kerkhof F.M.G.V.D..
The catalytic reduction of nitric oxide by ammonia over clean and
vanadium oxide-coated platinum foil. J Catal 1991;129:186 94.
Winter ERS. The catalytic decomposition of nitric oxide by metallic oxides.
J Catal 1971;22:158 70.

467

Yan L, Ren T, Wang X, Ji D, Suo J. Catalytic decomposition of N2O over


Mx Co1x Co2O4 (M=Ni, Mg) spinel oxides. Appl Catal, B Environ
2003;45:85 90.
Yang JP, Cheng RT. Role of WO3 in mixed V2O5WO3/TiO2 catalysts for
selective catalytic reduction of nitric oxide with ammonia. Appl Catal,
A Gen 1992;80:135 48.
Yang RT, Chen N. A new approach to decomposition of nitric oxide using
sorbent/catalyst without reducing gas: use of heteropoly compounds.
Ind Eng Chem Res 1994;33:825 31.
Yao HC, Shelef M. The surface interaction of O2 and NO with manganous
oxide. J Catal 1973;31:377 83.
Zeldovich Y. The oxidation of nitrogen in combustion and explosions. Acta
Physicochimica USSR 1946;21:577 628.
Zhang W, Yahiro H, Mizuno N, Izumi J, Iwamoto M. Removal of nitrogen
monoxide on copper ion-exchanged zeolites by pressure swing
adsorption. Langmuir 1993;9:2337 43.
Zhang Y, Anderson A, Muhammed M. Nanophase catalytic oxides: I
Synthesis of doped cerium oxides as oxygen storage promoters. Appl
Catal, B Environ 1995;6:325 37.
Zuelke RW, Skibba M, Gottlieb C. Sorption and magnetic susceptibility
studies on metal-free-radical systems: nitric oxide on palladium. J Phys
Chem 1968;72:1425 31.

Anda mungkin juga menyukai