Anda di halaman 1dari 24

This article was published as part of the

In-situ characterization of heterogeneous


catalysts themed issue


Guest editor Bert M. Weckhuysen


Please take a look at the issue 12 2010 table of contents to
access other reviews in this themed issue



D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4928 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
Analysing and understanding the active site by IR spectroscopyw
Alexandre Vimont, Fre de ric Thibault-Starzyk and Marco Daturi*
Received 7th June 2010
DOI: 10.1039/b919543m
IR spectroscopy is a technique particularly adapted for understanding the mechanism of catalytic
reactions, being able to probe the surface mechanisms at the molecular level. In this critical review
the main advances in the eld are presented, both under the aspects of the in situ and operando
approaches. A broad view of the most authoritative literature of the domain is given, based
largely on the experience built up at the LCS laboratory in the last decades. After having presented
the general methodology to observe a potential active site directly or by probe molecule adsorption,
several examples illustrate the qualitative and quantitative analysis of the physicalchemical
properties of the surface entities. The last part of the review is dedicated to the discrimination
of the role of the active site and its links with the catalytic steps; the hot problem of the reaction
intermediates and their visibility via spectroscopic techniques is critically addressed (138 references).
Introduction
Heterogeneous catalysis is nding an increasing importance in
everyday life. The industrial heterogeneous catalysts are
shaped multifunctional devices intended to oer to the reacting
agents a multitude of active sites, distributed on the external
surface or inside the porosity of the materials, in order to
optimise the contact between the reacting molecule and the
transformation centre in the sense of the Sabatiers principle.
On these sites the reacting molecules are adsorbed, the inter-
mediates are generated and the products formed. Physically,
the active centres are the sites where the reaction crosses the
energy barrier, as classically represented in Fig. 1.
Therefore, the active sites are intrinsically linked with the
intermediate species and they can be aected by the poisons
which can be formed during the reaction. An accurate design
of the active sites (in terms of quality, strength, position, . . .) is
obviously the key to obtain the optimum catalyst. On the way
to catalyst optimisation and rational design, we have to
thoroughly characterize the active sites, particularly when in
action, considering that they are not static entities, but they
undergo modications depending on the reaction conditions
and surface restructuration phenomena. For this purpose, IR
spectroscopy is one of the most adapted tools, being extremely
sensitive to the molecular vibrations and able to discriminate
the dierent geometrical distortions of the molecules according
to the adsorption state on a site.
In this contribution, we will present how it is possible to
evidence the adsorption sites on a catalyst, particularly using
Laboratoire Catalyse et Spectrochimie, ENSICAEN,
Universite de Caen, CNRS, 6 Bd Marechal Juin, F-14050 Caen,
France. E-mail: marco.daturi@ensicaen.fr; Fax: +33-231452822;
Tel: +33-231452730
w Part of the themed issue covering recent advances in the in-situ
characterization of heterogeneous catalysts.
Alexandre Vimont
Alexandre Vimont (born 1972,
France) received his PhD in
2000 from the University of
Caen in the eld of in situ
and operando infrared spectro-
scopy applied to catalysis,
under the supervision of Jean-
Claude Lavalley and Frederic
Thibaut-Starzyk. He then
joined the laboratoire Catalyse
et Spectrochimie as a permanent
CNRS research engineer. His
current research interests focus
on the comprehension at the
molecular scale of the adsorp-
tion sites in acid materials and
in particular in metal organic frameworks, using in situ and
operando Infrared spectroscopy.
Fre de ric Thibault-Starzyk
Frederic Thibault-Starzyk was
born in Saint-Lo (France) in
1965. He received his PhD
in synthetic organic chemistry
in 1992 from the University of
Caen. After post-doctoral
work with Pierre Jacobs at
the University of Leuven, he
became Charge de Recherche
(1995) and Director of Research
in the CNRS (2009) at the
Catalysis and Spectrochemistry
Laboratory. In 20034, he was
an overseas fellow, Churchill
College, working with David
King at the Chemistry Depart-
ment, University of Cambridge. His research interests are in
heterogeneous catalysis and infrared spectroscopy, including
operando spectroscopy, time resolved measurements, and new
spectroscopic approaches for the characterisation of solids and
zeolites.
CRITICAL REVIEW www.rsc.org/csr | Chemical Society Reviews
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 49284950 4929
probe molecules. We will describe the most accurate methodo-
logies to obtain information on physicalchemical properties
of the sites, as well as on their concentration and strength. We
will show then how to dierentiate an active site from a bare
adsorption centre by coupling FT-IR with catalytic evidences.
We will nally discuss briey the problem of the detection
limit of intermediate species (intrinsically linked to active sites)
using infrared spectroscopy.
Surface adsorption site identication on a catalyst
The best identication of the potential active sites by IR
spectroscopy is the direct detection of their IR ngerprint.
The best examples are probably those given by the zeolites, for
which the strong Brnsted-acid sites are often identied
through the n(OH) band of hydroxyl groups. If this is not
possible, specic methods must be employed to obtain a
spectroscopic response for the sites. The most common is the
adsorption of probe molecules, which provides IR spectra
specic to the interaction with a single site. From the spectro-
scopic point of view, several reviews have given criteria about
the choice of the right molecule,
35
However, without any
further considerations, it is worthwhile noticing that whatever
the probe molecule used, it generates a perturbation of the
surface of the catalysts: indirect perturbations such as electron
withdrawing eect, or direct perturbations like chemical reactions
(protonation, electron transfer, decomposition, . . .). These
phenomena can be considered as invasive, but according to
the probe used and the chemical reaction considered, this can
also give information on the dynamic response of a surface to
the molecule adsorption. Therefore, it can be concluded that
whatever the probe molecule used, the best one is the reactant
itself, providing the same perturbation as during the chemical
catalytic reaction. In this view, the best conditions for a study
appear to be those during the reaction, i.e. the operando
conditions (real in situ);
6
we will develop this point towards
the end of the review.
The spectroscopic study of the adsorption sites on a catalyst
surface requires a good knowledge of the surface state itself, as
well as an excellent reproducibility of the characterisation, to
x the parameters and the limits of the surface description. In
this view, it is necessary to be aware of the presence and nature
of surface impurities, to establish how to clean the surface and
make the sites accessible.
Impurities and surface activation
IR is a very sensitive technique for detecting surface impurities
such as water, carbonates (formed by catalyst contact with
ambient atmosphere), organics and other residual species after
synthesis, such as nitrates, sulfates, templates, etc., because
these species present characteristic bands in the IR spectra.
Although many of these species do not represent a real
problem for catalytic reactions, being often eliminated at the
temperature at which the process takes place, or even taking a
benecial role in the reaction itself, their presence may inhibit
(at least partially) probe adsorption, or simply disturb the
correct spectral interpretation by band overlapping. Band
position and intensity can vary depending on the sample
nature; an abundant literature exists describing these spectra
(see for example ref. 5). Thermal stability of these species
strongly depends on acidbase properties of the material: on
relatively acidic compounds (alumina, zirconia, . . .) carbonate
and nitrate species can be removed at mild temperatures. On
basic oxides, such as MgO, carbonate impurities can still be
observed after a thermal treatment at 750 1C. In the case of
sulfates, a reducing treatment at higher temperature is even
necessary to clean the surface, but traces of sulfur can remain.
7
Nevertheless, we should consider the fact that these residual
species are essentially bulk moieties, almost inert during the
catalytic process; therefore it is better to leave them in
place rather than to heat the sample at very high temperature
(with the aim to clean all the impurities), so producing a
drastic sintering of its surface.
It is much more dicult to detect the presence of alien
anionic entities such as sulfur, Cl

and F

, or cationic species
Fig. 1 reaction prole and intermediates in the simple case of a
process going through two elemental reaction steps A -I -B: a is
the prole for a non catalytic thermal reaction; b is a catalytic reaction
using a good catalyst; c is a catalytic reaction using a bad catalyst.
If I is an intermediate with a low stability, it is hard to detect; if I
0
is a
stable intermediate it will easily be detected. The transition state A* or
I* is not detectable.
1,2
Marco Daturi
Marco Daturi was born in
Genoa (Italy) in 1964. After
having obtained a Master in
Physics, he studied Chemistry,
receiving a PhD in Chemical
Engineering in 1996 from the
University of Genoa (director:
Prof. G. Busca). After post-
doctoral work with Dr J.-C.
Lavalley at the LCS Catalysis
and Spectrochemistry Labo-
ratory, he became Lecturer
(1998) then Professor (2002)
at the University of Caen,
where he teaches thermo-
dynamics and spectroscopy.
His research topics deal with in situ and operando IR spectro-
scopy, heterogeneous catalysis, material surface investigations
and design. He applies fundamental studies to the domains of
pollutants abatement and environmental protection.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4930 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
such as alkaline cations in the zeolites, due to the absence of
specic bands. An indirect method consists in the observation
of the n(OH) range of the catalyst spectrum: OH/Cl, OH/F
substitutions and OH/ONa
+
alkali exchange strongly modify
the intensity and the position of the n(OH) bands. These
impurities and contaminants strongly inuence the acidbase
properties of the samples. As an example, Cl

, F

and sulfate
species can increase the strength of the acid sites on alumina,
8
whereas sodium (even if present in as small quantities as
300 ppm) poisons the strongest Lewis-acid sites,
9,10
Traces of
chloride also hinder oxygen mobility on cerium-based com-
pounds
11
and inhibit the oxidation properties of the supported
metals.
12
Eect of the temperature of activation
The increase of the temperature of activation for metal oxides
not only removes the impurities and molecularly adsorbed
water but also leads to surface reconstruction, due to the
elimination of hydroxyls, according to the mechanism involving
two MOH groups: 2 MOH - MOM + H
2
O (with the
possible creation of a vacancy). In extreme conditions, if the
activation temperature exceeds the calcination conditions
during sample synthesis, sintering phenomena can also occur.
13
The thermal treatment leads to surface reconstruction, and
the amounts of exposed OH groups, cationic, anionic and
defective sites, which govern their acidbase and redox properties,
are not predictable, but can be estimated by IR absorption
experiments of probe molecules at various temperatures. The
most complete IR studies about the eect of the temperature
of activation on the surface modication spectroscopy probably
deal with MgO and alumina.
1417
Direct observation of the active sites
Case of Brnsted-acid sites. Direct observation of the poten-
tial active sites is possible when Brnsted-acid sites are con-
sidered and this possibility is well illustrated by the zeolites or
silica, for which the hydroxyl groups can be directly observed
through their n(OH) bands. The spectrum of hydroxyl groups
of steamed HY zeolite (Fig. 2) allows one to observe the acidic
OH group located in dierent sites: bridged hydroxy groups in
supercages (3625 cm
1
), or in sodalite cages (3545 cm
1
).
Perturbation of bridged OH groups by extraframework
entities generates highly acidic species with specic n(OH)
band (3602 cm
1
for those in the supercages, 3520 in the
sodalite cages). For this zeolite, the presence of the 3602 cm
1
band is often related to the activity of the catalyst in strongly
demanding reactions like n-hexane cracking,
18,19
The OH
bands in the dealuminated Y zeolites have been studied in
detail and a new model has been proposed recently.
20
Such a
sensitivity to the environment is well illustrated with the IR
spectra of the mordenite, which presents three nOH bands,
depending on the location (Fig. 2b).
21
The spectra of the
sample during the conversion of xylene show that only the
hydroxy groups present inside the main channels act as active
sites towards the isomerisation reaction. Sites located in the
narrow side pockets are not catalytically active, but have a
strong inuence on selectivity. This case illustrates nicely the
fact that the direct observation of the potential sites can give
valuable indications on the reaction protagonists.
22
Thermally activated divided metal oxides present residual
OH groups having IR stretching frequencies related to the
nature of surface cations. The multiplicity of the n(OH) bands
of oxides such as alumina has been mainly explained invoking:
(i) the multifold coordination of the hydroxyls themselves
(linear species giving rise to n(OH) bands at a higher frequency
than those characterising bridged species), (ii) the coordination
number of the cation to which OH groups are bound, (iii) the
presence of morphologic defects on the surface (edges, corners,
etc.). In the case of many oxides such as Al
2
O
3
, Ga
2
O
3
, CeO
2
,
ZrO
2
, the higher n(OH) wavenumber component presents
a basic rather than acidic character, as shown by CO
2
experiments.
2325
Their relation to defect crystallographic
sites has also been invoked considering the sensitivity of this
component to the presence of coordinated species in the
neighborhood.
24
This shows that the co-existence of several
types of sites on the surface complicates the relation between a
n(OH) band and a well dened Brnsted surface sites in the
case of metal oxides.
Case of Lewis-acid sites. Direct IR observation of a
Lewis-acid site itself is not possible since a coordinatively
unsaturated site is not a vibrator. Considering the Lewis site
and its rst coordination sphere, the vibrations (metaloxygen
vibrations for metal oxides) are generally coupled with the
very intense bands from the bulk vibration modes and only
studied by mixing the samples with KBr; therefore only
hydrated solids are generally studied. However in few cases,
defect sites are detected in the IR spectrum of activated
samples by specic IR bands, mainly in the low frequency
range: bands situated at 3782/880 cm
1
on beta zeolite
26
and
around 960 cm
1
on Ti-silicalite
27
can be considered as a
ngerprint of Lewis-acid sites and have been partly related
to the enhanced activity of these Lewis-acid catalysts in the
catalyzed MeerweinPondorfVerley (MPV) reaction.
28
Surface AlO species (at ca. 1050 cm
1
) on alumina
29
and
strained siloxane bridges on highly dehydroxylated silica
(880940 cm
1
) are dissociation sites for water, alcohols, H
2
S.
30
However, the number of cases in which direct characterisation
Fig. 2 IR spectra of the hydroxy groups of a steamed Y zeolites
(a) and of a mordenite zeolite (b) after activation at 450 1C.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 49284950 4931
of Lewis acidity is possible is limited, and complementary IR
adsorption experiments using probe molecules are often
necessary.
Probe molecule use. The direct detection of adsorption sites
by IR spectroscopy is often not possible, therefore specic
methods must be employed to obtain a spectroscopic response
of the sites. The most common is the adsorption of probe
molecules which gives IR spectra specic to the interaction
with the site, as mentioned above. Concerning the choice of
the appropriate probe molecule, Lercher et al.
3
gave special
emphasis to the criteria that have to be met to arrive at a
characterization of materials that are useful for catalytic
application, selecting the right molecule for the right site. The
choice of the adapted probe will depend on many parameters:
the chemical function providing the interaction with the site
under study; the size of the molecule, depending on the site
accessibility; the optimum of the interaction (sucient to
furnish valuable information, not excessive so as to limit
surface modications); the spectral response, producing a signal
intense enough, with band positions sensitive to the interaction;
stability on the surface catalyst to avoid decomposition; and
sucient vapour pressure to be easily introduced in an IR cell.
The investigated sites are very often cations, acting as Lewis
centres, but infrared spectroscopy has also been widely used
in order to characterize the metallic centres in oxides and
deposited metal complex. Even if direct investigations on
metaloxygen vibrations are reported
3133
most of the studies
related with catalysis are dealing with the adsorption of probe
molecules. Among these molecules, N
2
, methanol, NO and CO
are frequently used and the latter two are also by far the most
common in the literature. In 2002 a very complete review
concerning the infrared spectra of chemisorbed carbon
monoxide as a characterization tool for the cationic sites of
oxides
34
was published. A specic property of CO is that the
slightly antibonding HOMO 5s orbital is occupied. This
orbital is very important for the electron-donating properties
of CO, because a decrease of electron occupation on this
orbital leads to stabilization of the entire molecule and thus
to an increase of the n(CO) wavenumber (compared to the
n(CO) for the gas phase at 2143.5 cm
1
). On the contrary, the
addition of electron density from a metal d orbital to one of
the 2p* LUMO orbitals (so called p back donation) leads to a
substantial decrease of the vibrational frequency of CO, i.e. to
a weakening of the CO bond. As far as the red-ox properties
are concerned and in the ideal case, the expected information
with CO as a probe are the following:
oxidation state of the cations on the surface,
coordination state of these cations,
location of the cations on at planes or other surface
structures,
location of the active phase on the support,
surface phase analysis,
existence of strong oxidizing agents on the surface.
The characterization of the various sites on mixed oxides
can be advantageously carried out by CO adsorption at
various equilibrium pressures at low temperature, followed
by evacuation at increasing temperatures to obtain infor-
mation about the stabilities of the various species. Although
the CO stretching frequency is the most informative parameter,
the data determining the stabilities of the various species can
be decisive for the assignment of the bands. Multiple carbonyls
adsorbed on the same metal cation are possible and in order to
identify them, isotopic mixtures should be used. This was the
case, for example, of a PtNamordenite sample, where the
use of a
12
CO
13
CO isotopic mixture combined with analysis
of the second derivatives of the spectra was very useful
for proving the polycarbonyl structures.
35
Again, using iso-
topic labeled
13
CO and
15
NO molecules mixed with their
most abundant analogues, it was possible to describe the
multiplicity and symmetry of CO and NO ligands on a
complex coordinated with Rh
2+
in a Rh-ZSM-5 sample;
geometrical structures and band assignments were supported
by DFT computational results.
36
However, sometimes the
polycarbonyls are very stable and in this case, if
12
CO is
adsorbed rst and then
13
CO introduced, mixed species may
not form at ambient temperature.
Concerning NO, Hadjiivanov
37
reported that the coordi-
nation of the NO molecule to a cationic site via the nitrogen
atom is accompanied by a partial charge transfer from the
5s orbital together with an increase in the bond order, just as
in the case of CO. Formation of a p-back-bond, although not
as easy as with CO, is also possible, and this results in a
decrease in the NO stretching modes. The dierent surface
mononitrosyls absorb in a wide spectral range: 19661710 cm
1
.
When only a s bond is formed, a frequency above that of
gaseous NO (1876 cm
1
) is expected, whereas with low-valent
cations, rich in d-electrons, p-back donation is possible and
the NO stretching modes can fall below 1876 cm
1
. Cations
having no d-electrons produce mononitrosyls only; on the
contrary, dimeric molecules are very often the principal
adspecies on transition metal cations. This is the case of NO
adsorption on V-, Cr-, Mo-, W-, Fe-, Cu- and Co-containing
oxide systems where the metal cations are not in their highest
oxidation state. Thus, it is clear that a p-back donation
stabilizes dinitrosyls. Examples of complementary information
obtained by CO and NO coadsorption are also available.
38
Acid sites. A review by Busca
39
describes the bases of IR
spectroscopic methods for the characterization of the surface
acidity (both of Lewis and Brnsted type), on dierent mixed
oxides. A systematization is proposed associating the surface
acidity with the ionicity/covalency of the elementoxygen
bond, mainly aected by the size and charge of the cation.
In a parallel work, the results obtained for the characterization
of the Lewis-acid strength of more than 30 binary and ternary
mixed oxides are interpreted on the basis of the dierent
polarizing powers of the involved cations.
40
Recent progress
on acidity characterization is reviewed elsewhere and
described to be related to the broadening of the spectral range
(investigation of overtones, combination bands, and low-
frequency modes) and to the adsorption of new non-traditional
probe molecules for identication of acid sites.
41,42
Some examples on oxide and zeolitic compounds. The choice
of the probe molecule is crucial to obtain an overall view of the
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4932 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
acidity. Its size has to be small enough to interact with all
available sites and to avoid connement eects
43
but its basic
strength has to be strong enough to interact even with the
weakest acidic sites. Ammonia seems to be a good candidate
for this but due to the high polarity of the NH bonds,
hydrogen bonding with basic entities governs the coordination
of adsorbed species and direct conclusions about acidic
strength are not straightforward.
44
That is why most often
the adsorption of several probe molecules is required. For
example, the FAU and the MOR structures are made of big
and small cavities in which some acidic hydroxyls are out of
reach of basic molecules such as pyridine. Co-adsorbing the
strongly basic trimethylamine (TMA) and NH
3
(see Fig. 3), we
were recently able to give for the rst time an infrared evidence
of three distinct acidic hydroxyls in defect-free HY,
45
to give
an assignment for the corresponding wavenumbers and to
characterize their respective acidic strength.
46
Moreover,
TMA desorption associated with the recovery of hydroxyls
at 3656 and 3638 cm
1
(and two corresponding n(NH) bands
reveal the presence of at least two distinct acidic strengths for
the hydroxyls located inside the supercages. For the same site
location, the local chemical factor should then play a role: the
aluminium distribution in the framework is not necessarily
homogeneous, and the number of Al next-nearest neighbours
inuences the acidic strength of a given site. Another explana-
tion for the unusual 3656 cm
1
component could be that part
of the O
4
crystallographic sites is a proton holder for this low
Si/Al ratio HY sample; in such a case, all the four theoretically
forecasted sites in the zeolitic FAU structure would have been
observed by IR spectroscopy.
46
The combined use of these
two molecules moreover helped us to better characterize the
various coordinated NH
4
+
and determine the activity ranking
between the ammonium species and coordinated ammonia
over Lewis sites during NO
x
SCR.
47
In the zeolites, some strong Lewis-acid sites can be obtained
during steaming leading to the formation of extra-framework
aluminium species. Mild Lewis sites may be naturally present
(in the alkaline form) or generated upon ionic exchange with
transition metals which are necessary for NO
x
SCR with
hydrocarbons. For over-exchange level, some Lewis species
may remain on the external surface and coadsorption of the
bulky ortho-toluonitrile and CO was recently reported to
identify the dierent Co
n+
species and their location in a Co
H-MFI zeolite.
48
In this respect, the use of the nitrile probe
provided more precise evidences about the distribution of Co
species in active CH
4
-SCR Co-HMFI than those arising
from previous UVvis, EXAFS and XRD data.
4953
The
oTN (ortho-toluonitrile) and NO co-adsorption allowed to
determine that a signicant amount of cobalt species is at the
external surface, mostly in the form of divalent cobalt. On
the other hand, in the internal surface part of Co species are
trivalent, together with predominant divalent Co ions. These
observations coupled with those coming from the operando
study were valuable for reactivity explanation, inferring that
the active sites for CH
4
-SCR in Co-MFI are Co
3+
species
(presence of a nitrosyl n(NO) band at 1930 cm
1
) located in
the cavities, but likely in non-classical cation positions, which
are able to convert NO to an adsorbed bridging nitrate species,
that can be later decomposed to give gas phase NO
2
.
54
Moreover, the cavity may contribute to the stabilization of
aggregates containing trivalent cobalt. At the same time, the
presence of Co-isocyanates involved in the SCR suggests that
a possible route for the reaction implies the reduction of
nitrate-like species by methane, forming H
2
O and isocyanates,
which could later react with NO producing N
2
and CO
2
. On
the contrary, it seemed that substitutional Co
2+
ions did
not play a key role in the reaction, being very likely almost
redox-inactive. Co
2+
-dinitrosyls formed on them being
decomposed well below the reaction temperature, they did
not seem to be involved in the reaction.
54
These considerations
linking the active site with the reactivity of the species
coordinated on it and with the possible intermediates can be
generalized to all the reactions and will be discussed later in
this review.
Metalorganic framework: the ideal case for spectroscopic
identication of adsorption sites. In the case of the metal
oxides, surface relaxation and reconstruction phenomena
might not allow to accurately identify the nature of the
Lewis-acid sites (coordination, oxidation degree, unsaturation
degree) via the study of spectra after the adsorption of probe
molecules. On the contrary, this is not the case for the majority
of hybrid organicinorganic solids such as MOFs (metalorganic
frameworks), which are crystalline nanoporous solids, with a
framework of inorganic units (clusters, chains or planes) and
organic linkers (phosphonates, carboxylates, sulfonates). The
majority of MOFs present framework metal sites that can
exhibit coordinative vacancies upon solvent removal; such
sites represent Lewis-acid centers of well dened symmetry
and oxidation degree. As has been shown in the case of
MIL-100
55
and H-KUST
56,57
it is possible to assign in a clear
and indisputable way the bands due to the adsorbed species,
like those (still debated nowadays) for carbonyls or nitrosyls
coordinated on cations such as chrome, copper and iron. CO
adsorption on MIL-100/101(Cr) is an example of the con-
tribution of IR to the identication of the potential active sites
in MOF.
55,58,59
Three n(CO) bands are observed at 2207,
2200 and 2193 cm
1
(Fig. 4), showing that Cr
3+
sites are
Fig. 3 Infrared spectra of the HY sample upon TMA adsorption and
NH
3
saturation evidencing OH in the supercages at 3637 cm
1
, OH in
the sodalite units at 3548 cm
1
and OH in the hexagonal prism at
3501 cm
1
(from ref. 45). Reproduced by permission of the American
Chemical Society.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 49284950 4933
not equivalent. This heterogeneity is attributed to the possible
presence of uoride ions (synthesis being performed in the
presence of hydrouoric acid) on the metallic trimers, with 2, 1
or no uorine ion, respectively, in the neighborhood of the
coordinatively unsaturated (CUS) Cr
3+
considered site.
55
Quantitatively, the number of free Cr
3+
sites in the activated
compound is exactly that expected, considering that one
corner over the three octahedra is occupied by one anion.
On MIL-101(Cr), this methodology allows identifying these
sites as the grafting centres of catalytically active sites.
59
Relationship between probe molecules and activity: the case of
acetonitrile. Two possible aspects of acidity are generally
considered. The rst one is the hydrogen bond that can be
established between a Brnsted site and the basic probe
molecule (for example when carbon monoxide is interacting
with acidic zeolites at liquid-nitrogen temperature). The second
one is the extent of proton transfer, or more exactly the
amount of protonated probe molecule (e.g. pyridine) on the
surface of the catalyst at room temperature. These two facets
of acidity, however, can not reliably be used for explaining the
catalytic activity in acid-catalyzed reactions. The H-bond
between a basic molecule and the various acid sites in a solid
is strongly inuenced by temperature. The linear relation-
ship between the strength of a H-bond with CO at 100 K
(often used for comparing solid catalysts) and the activation of
proton transfer to a hydrocarbon in a chemical reactor at
700 K is far from being established. Acetonitrile has been
increasingly used as an infrared probe molecule for solid
catalysts, and might well provide an integrated approach of
acidity. Both H-bond and protonation can be observed, and it
can be used to probe the actual proton transfer in reaction
conditions.
In the absence of water, heating acetonitrile on an acidic
zeolite leads to the reversible protonation of the probe
molecule.
60
Protonation of acetonitrile on zeolite Brnsted
sites at high temperature has been used to build an acidity
scale agreeing with catalytic activities in the conversion of
saturated hydrocarbons: acidity is determined under conditions
near to that of catalytic reactions (high temperature), leading
to a more relevant parameter for the prediction of catalytic
activity than H-bonds. Acid catalysis involves protonation
of the reactant, and a scale built on actual protona-
tion, preferably in reaction conditions, is more interesting.
The protonation temperature was measured on a series of
zeolites, and depends very much on the pore size.
61
The nearer
the pore size is to the size of the adsorbed molecule, the lower
the protonation temperature. For example, in mordenite, two
main locations exist for the Brnsted site: in the main channels
and in the side pockets. From the point of view of CO at
low temperature, a stronger H-bond is created in the main
channels, and the stronger site would therefore be in the main
channels. However, protonation has only been observed in the
small side pockets, where maximum connement takes place.
The measurement of the protonation temperature is a way to
know how easy the proton transfer is from the acid catalyst
to the basic adsorbed molecule. It is therefore a new spectro-
scopic measurement for the acidity of the solid, one that
does not only involve the interactions strength between the
adsorbed molecule and the surface, but the actual proton
transfer, the real nature of acidity. The probe is here not
the adsorbed molecule itself, but rather the actual catalytic
protonation reaction. The new scale obtained between various
zeolites was compared to the activity in catalytic conversion of
saturated hydrocarbons at high temperature, a reaction where
activity is linked to the acid strength of very strong sites.
Contrarily to what was observed with H-bonding of CO or
with pyridine protonation at room temperature, the scale of
acetonitrile protonation temperature was perfectly linked with
the catalytic activity.
62
Modication of the basicity of the probe molecule by the acid
site. Acetonitrile has also shown that the basicity of the probe
is inuenced by the adsorption on the solid, especially on
zeolites where connement is important. During molecular
dynamics simulations of acetonitrile on mordenites,
63
the
electric dipole on the molecule was signicantly modied in
the move from the main channel to the inside of the small
lateral cavity. In the purely siliceous mordenite used for the
simulation, the molecular dipole of acetonitrile was enhanced
from 3 D (in the large channels or in the gas phase) to 4 D in
the small cavities. Such enhanced dipole could also imply that
the basicity of acetonitrile would be increased by a third on
entering the small cavities. Moreover, the probe molecule is
locked in the cavity, and does not escape easily. It is thus
kept in close distance to the OH group, thus enhancing the
probability for the proton transfer (Fig. 5).
When the size of the molecule compares with that of the
cavity, electron densities in the basic molecule are modied,
the basicity of the probe can be enhanced, and the proton
transfer can happen more easily. All these parameters show
that the basicity of the probe molecule can not be considered
without the solid, and that it is always a pair at work.
Modication of the acidity of the site by the probe molecule.
In some cases, the probe molecule itself modies the surface
acid sites: during comparative study of the Lewis acidity
between SiO
2
B
2
O
3
and alumina by adsorption of pyridine
(py), CD
3
CN (MeCN) and CO, no coordination of carbon
monoxide has been observed on silicaboria even at low
temperature whereas CO strongly coordinates on alumina.
64
Similarly, coordinated pyridine and acetonitrile show a
much lower thermal stability on SiO
2
B
2
O
3
than on Al
2
O
3
,
Fig. 4 Left: CO adsorption sites of the trimers of chromium
octahedra in MIL-100(Cr). Right: n(CO) bands of CO adsorbed on
MIL-100(Cr).
55
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4934 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
indicating that silicaboria presents much weaker Lewis-acid
sites than g-Al
2
O
3
. On the other hand, coordinated pyridine
and acetonitrile species show that infrared frequency shifts
(n
8a
, n
19b
and n(CN), respectively) are larger on B
2
O
3
SiO
2
than on Al
2
O
3
, suggesting that charge transfer from these
probe molecules is more important on the B
3+
than on the
Al
3+
Lewis acid (Fig. 6). DFT calculations of the interaction
of these probe molecules with models representing Al
3+
and
B
3+
Lewis-acid sites adequately reproduce these experimental
observations (Fig. 7). The weakness of the B
3+
Lewis-acid
sites is ascribed to the p-character of BO bonds, which
disfavours the conversion of boron from a trigonal planar
conformation to a tetrahedral conformation upon adsorption
of probe molecules and decreases the adsorption energy of
pyridine and acetonitrile despite a strong charge transfer.
The absence of interaction noted during the adsorption of
CO on SiO
2
B
2
O
3
has been explained by its basicity, not
strong enough to compensate for the energy required for the
conformational change of the B
3+
Lewis-acid centre.
Transformation LewisBrnsted sites. Brnsted acidity can
be generated by water on the surface of crystalline or amorphous
solids. On metal oxides, it is well known that water can
transform Lewis into Brnsted acidity by dissociative water
adsorption on the M
d+
O
d
acidbase pairs, with the con-
sequent creation of MOH acidic groups. On aluminium
uorides, water addition transforms Lewis-acid sites into
Brnsted sites on activated compounds.
65
Due to the strong
Lewis acidity, water is strongly coordinated on uorides and
generates strong Brnsted sites, as evidenced by CO adsorp-
tion. The absence of suciently strong basic sites on uorides,
as shown by an unpublished study of propyne adsorption,
inhibits dissociative adsorption of water on such materials.
The consequences of the adsorption of protic molecules
(water and alcohols) on the acidity of MOFs, have been well
identied by IR spectroscopy:
55,58
adsorption of water on
activated MIL-100(Cr) leads to the formation of coordinated
species well characterized by two narrow n(OH) bands at
about 3700 and 3580 cm
1
. CO adsorption at 100 K shows
that coordinated water induces the creation of Brnsted-acid
sites with a strength close to that reported in the case of
phosphated silica. The successive addition of CO molecules on
hydrated MIL-100(Cr) shows that each coordinated water
molecules creates two Brnsted-acid sites (see Scheme 1).
This LewisBrnsted acid conversion was used to modulate
the strength of the created Brnsted-acid sites according to the
Fig. 5 Molecular dynamic simulation of acetonitrile in a purely
siliceous mordenite. The Y coordinate corresponds to the distance of
the probe molecule from the centre of the main channel. The Z
coordinate corresponds to the distance along the main channel.
(A) Positions of the centre of gravity of acetonitrile during the
simulation, showing it can go from the main channels to the side
pockets. (B) Molecular dipole and (C) location of the molecule during
the simulation vs. distance from the main channel in the mordenite
structure along the axis of the side pocket (Z B2 A

) (adapted from
ref. 63). Reproduced by permission of the American Chemical Society.
Fig. 6 IR spectra of activated Al
2
O
3
(top) and boriasilica (bottom)
after introduction of pyridine at room temperature (left) and CO at
100 K (right-spectra at various coverage).
64
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 49284950 4935
nature of the adsorbate used: the comparison of the results
deduced from the grafting of CH
3
OH, H
2
O, CF
3
CH
2
OH and
(CF
3
)
2
CHOH shows that the stronger the acidity of the
adsorbate, the higher the acidity of the generated Brnsted
sites. Their strength is directly related to the nature of the
grafted molecules and can reach that of the zeolitic Brnsted-
acid sites when uorinated alcohols are used (see Fig. 8).
58
Basic sites. Heterogeneous catalysis using basic solids has
been much less studied than acidic catalysis. It seems even that
a concern exists about the denition of basicity itself. It may be
useful to recall here that, in spite of the common use, there is
not a physical dierence between the so-called Brnsted and
Lewis basicity, because for both it is due to electrons on
oxygen atoms. In fact, for Brnsted, an acid site is a proton
donor (i.e. an hydroxyl on heterogeneous catalysts), whereas
Fig. 7 Evolution of the total interaction energy (DE
tot
, ), deformation energy of the Lewis-acid center (DE
def
(A), ), deformation energy of the
probe molecule (DE
def
, --), and interaction energies of the probe with the Lewis center at their geometry in the complex (DE
int
, ---) for six acidbase
complexes as a function of the ML distance (M = B or Al; L = C or N).
64
The deformation energy is the dierence between the energies of the
isolated species at their equilibrium geometries and the energy of the isolated species at their geometry in the complex.
Scheme 1 Interaction of CO with coordinated water molecules in
MIL-100(Cr).
58
Reproduced by permission of the American Chemical
Society.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4936 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
for Lewis it is an electron pair acceptor (i.e. a cation); but
looking at basicity denitions, in the sense of Brnsted a basic
site is a proton acceptor, while for Lewis it is an electron pair
donor, therefore an oxygen site in both cases, or an halogen
atom for halogen based catalysts. Some simple concepts
addressing this problem have also been mentioned by
Zecchina et al.
66
Concerning the diculties in studying surface basicity, we
might explain it with the fact that FT-IR is a technique
sensitive to molecular bonds: in the case of acidity, when using
a probe molecule to characterize Lewis sites, the molecular
deformations of the molecule induced by the cationic polarizing
eect are measured. In the case of a basic site, the surface is
donating electrons and the back-donation eect on a molecule
is a more complicated phenomenon to quantify. When studying
proton donation or acceptance, the spectra are always com-
plicated by broad and badly resolved features. From a general
point of view, anyway, protonic molecules seem particularly
adapted for probing basic sites, taking into account the
basicity denition itself (and impacting them only by weak
interactions).
As an acid site is always associated to its conjugated basic
site, reviews dealing with characterization of acidity by infrared
also report interesting data about basicity (see for example
ref. 41) and even describe some typical probe molecule inter-
actions: CH-acids such as chloroform (Cl
3
CH(D)), acetylene
(C
2
H
2
) and methylacetylene (CH
3
C
2
H) are shown to be
potentially suitable probe molecules for basic properties using
the H-bonding method.
67
All three molecules undergo
Oz
2
HC hydrogen bonding and the induced red-shift of
the CH stretching frequency permits a ranking of the base
strength of a given series of materials. Many other probe
molecules were tested for the specic study of the surface
basicity of divided metal oxides, and Lavalley reviewed some
years ago the infrared spectrometric studies of the surface
basicity of metal oxides and zeolites using adsorbed probe
molecules.
4
Results obtained from carbon monoxide (CO),
carbon dioxide (CO
2
), sulfur dioxide (SO
2
), pyrrole (C
4
H
5
N),
chloroform (CHCl
3
), acetonitrile (CH
3
CN), alkanes, thiols,
boric acid tri-Me-ether, ammonia (NH
3
), and pyridine
(C
5
H
5
N) were discussed in that well cited review. As we
already noticed in the case of the acidity study, the author
reminds us that no probe can be used universally. CO
2
for
weakly basic metal oxides and for basic OH groups, CO for
the characterization of highly basic structural defects on metal
oxides activated at high temperature and pyrrole in the case of
alkaline zeolites, appear to be quite suitable probes. Moreover,
both NH
3
and pyridine (generally used as probes for the
measure of the acidity of catalysts) are also described to
adsorb on basic oxides via dissociative chemisorption. In this
respect we should stress that a strongly interacting probe (such
as CO
2
and NO
2
, for example) highly modies the surface
behaviour, often leading to false interpretations on the
strength of the basic site. Moreover, a probe which dissociates
is an unfriendly tool for site concentration characterisation
when using volumetric methods. Again, a non-dissociating,
weakly interacting protonic probe would appear as a most
adapted tool for surface basic site characterization.
A pioneering review on the basic properties of zeolites was
proposed by Barthomeuf.
68
More recently, NO
2
disproportio-
nation on alkaline zeolites was used to generate nitrosonium
(NO
+
) and nitrate ions whose infrared vibrations are shown
to be very sensitive to the cation chemical hardness and to the
basicity of zeolitic oxygen atoms.
69
Recently, Michalska et al.
70
have pointed out that propyne
is an excellent probe for the study of oxygen basicity. Addi-
tionally, they have veried that probe dissociation does not
depend on the site strength, but can be due to the presence
of Lewis-acid sites coupled with the basic ones: the formation
of the hydrogen bond weakens the bond between RC and H.
Fig. 8 Brnsted-acid strength of OH groups from various grafted species on MIL-100(Cr) measured by CO adsorption: correlation between the
n(OH) shifts, the H
0
values and the n(CO) position.
58
The y axis label should be read as |Dn(OH)|. Reproduced by permission of the American
Chemical Society.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 49284950 4937
In the presence of an acid site, which can host the CH
3
CRC
moiety, the surface protonation is therefore easily achieved.
Therefore, propyne dissociation is a good probe for the
presence of acidbase pairs on a surface.
70
Basicity case study. Among the catalysts having basic
properties, ceria is probably the most common owing to its
properties in the domains of oxidation and hydrogen produc-
tion. It is the base material for car exhaust control devices,
notably for Three Way Catalyst (TWC) applications. For
these reasons an extensive characterisation of its properties
has been undertaken for at least two decades. Pyrrole adsorp-
tion on CeO
2
leads to dissociative adsorption characterized by
stretching ring vibrations at 1444 and 1367 cm
1
typical of
pyrrolate ions and n(OH) vibration at 3628 cm
1
typical
of surface hydroxyls formed upon proton transfer.
71
This
complete dissociation of C
4
H
5
N is indicative of the high
basicity of CeO
2
surface O
2
ions but does not allow an
investigation of its variation upon reducing ceria. CO
2
was
further adsorbed as it acts as a Lewis acid toward either O
2
surface ions (with the production of carbonates) or residual
basic OH surface species (with the production of hydrogen
carbonates (HC)). The study extended to CeZr mixed oxides
72
indicates that hydrogen-carbonates (indicative for basic OH
groups) are mainly observed for rich ceria compounds and
that the intensity of carbonate species is directly proportional
to the cerium content, as expected according to the basic
properties of this element. Identication of the spectral
features typical of each species arising from CO
2
adsorption
was claried by studying the splitting of the n
3
band of
carbonates and their thermal stability.
Nowadays, basicity is becoming a more and more important
parameter, and basic materials are involved in numerous
industrial processes, such as ne chemical production via green
chemistry routes (replacing homogeneous by heterogeneous
procedures, for example in esterication reactions in the
absence of any solvent), or environmental catalysis. One of
the most investigated methods for nitrogen oxide removal is
NO
x
-trapping. In this process, NO
x
are stored and concen-
trated in highly basic compounds, before being submitted to a
reduction. The materials used in this respect are typically
barium, alkaline metals, alkaline-earths and rare earth
oxides.
73
In such a case CO
2
and NO
2
are the most adopted
probe molecules, giving rise to carbonates and nitrates,
respectively. The storing materials are obviously submitted
to severe surface and bulk restructuration, but in the same way
they will be under service, exposed to the reacting lean and
reach ows, containing NO
x
and high concentrations of CO
2
.
General indications on nitrate characterization and their
coordination on a number of solids can be found in the
excellent review by Hadjiivanov.
37
A comparative study of
barium and potassium based formulation
74
indicated upon
NO
2
adsorption the formation of both ionic and covalent-like
NO
3

species over PtRh/Ba/Al


2
O
3
, whereas only very stable
ionic potassium nitrates (sharp peak at 1373 cm
1
) were
detected over Pt/K/Mn/CeAl
2
O
3
. This is due to the higher
basicity of the potassium sites which furthermore enlarges
the adsorbing temperature window and delays the nitrate
release during the rich step impeaching NO sudden outlet.
A comparison of the latter composition with a great number
of other NSR catalysts was reported in ref. 75, where the exact
chemistry of nitrite and nitrate formation was investigated,
and their coordination on specic structural sites of the oxide
determined thanks to the parallel use of TEM analyses.
Another interesting example is the synthesis of phyto-
sterol esters from transesterication of a fatty methyl ester
(dodecanoate) with b-sitosterol carried out in the presence of
basic solid catalysts, such as lanthanum oxides.
76
The
acidbase properties of La
2
O
3
were characterized by IR
spectroscopy, which revealed the presence of residual unidentate,
bidentate, polydentate, and mineral carbonate species inside
all of the solids even after activation, suggesting dierent basic
and catalytic characteristics of the samples. The carbonate
strengths were determined by propyne adsorption, using the
shift of the n(CRC) stretching mode for the adsorbed species:
the lower the position of that vibration, the greater the basicity
of the corresponding site. A ranking of the basic strength of
the surface carbonates species of the lanthanum oxycarbonate
samples was thus possible and it was correlated to the catalytic
activity: the lower the basicity of carbonates, the higher the
phytosterol ester yield. Moreover, a thorough spectral
characterization of KBr-diluted samples indicated that the
higher the intensity of unidentate carbonate bands (1499 and
1382 cm
1
), the higher the sterol ester yield, suggesting those
moieties of medium basic strength play a central role in the
catalytic mechanism.
It was shown in our laboratory that COS hydrolysis
77
and
CS
2
hydrolysis
78
could be used as test-reactions for metal
oxide hydroxyl basicity. However, the use of IR spectroscopy
should add value to this form of characterisation: adsorption
of CS
2
on a series of metal oxides (Al
2
O
3
, ZrO
2
, ZnO and
CeO
2
) gave rise to a specic interaction with O
2
sites, leading
to the formation of xanthate (COS
2
)
2
species characterized
by bands in the 12001000 cm
1
range whose intensities
correlate well with the relative basicity of the analyzed oxides.
Co-adsorption experiments of CS
2
with either CO
2
or pyridine
showed that its adsorption sites are mainly those giving rise to
bidentate carbonates, showing that this new probe is more
specic than CO
2
. Moreover, the transformation of xanthate
into carbonate species at low temperature could account also
for the surface oxygen mobility.
79
Metallic and redox sites. Transition metal oxides, rare-earth
oxides and various metal complexes deposited on their surface
are typical catalytic phases leading to redox properties. For
each of these phases, complementary tools exist for an appro-
priate characterization of the metal coordination number,
oxidation state or nuclearity. Among all the techniques, IR
provides information by characterizing the characteristic
vibrations of intrinsic (hydroxyls) or extrinsic (methanol,
CO, . . .) probes.
We have already discussed the principles of the use of probe
molecules for the characterisation of surface species in the
section concerning probe molecule use. CO and NO can
provide highly valuable information on the supported metal
dispersion and coordination, as well as on the oxidation degree
of such moieties. Typical examples can be found in the
works of Binet
8083
and Bazin.
84,85
This methodology appears
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4938 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
particularly adapted to the case of redox supports, where the
alternative hydrogen chemisorption will lead to imprecise
results. Another positive point for the use of such probes
is their small size, allowing site accessibility even in small
cavities. Moreover, the fair interaction energy between the
probe and the site is a guarantee for a minor perturbation
of the surface state upon interaction. A useful volumetric/
spectroscopic CO adsorption combined method allows
metal dispersion calculation in a simple and reliable way, by
integrating the bands relative to CO adsorption on the metal
sites vs. the molar amount of the introduced probe, as shown
in Fig. 9.
85,86
In such a way CO adsorption mode, sites and
quantity are continuously monitored by IR spectra upon
calibrated doses introduction.
Additionally, carbon monoxide and nitrogen monoxide
adsorption can provide very useful information on the coordi-
nation mode of such molecules (of primary importance for
environmental issues) and on the complexes formed with
the metal particles, potential sites for pollutant abatement.
For example, it was found that the adsorption of CO on a non-
reduced Pt/TiO
2
sample reveals the existence of Pt
3+
and Pt
2+
cations as well as some amount of metallic platinum. However,
Pt
4+
species are also present on the sample but, being
coordinatively saturated, cannot adsorb CO, while NO forms
nitrosyl species with bare Pt
2+
and Pt
0
sites, but it is not
coordinated to Pt
3+
ions. Therefore, it appears that NO is a
more sensitive probe than CO for testing the state of Pt
2+
cations. Moreover it must be underlined that probe adsorption
is not totally innocent: adsorbed CO slowly reduces the
platinum cations, whereas NO oxidizes metallic platinum
even at ambient temperature.
87
Similar results were found on
Rh-ZSM-5, where a new kind of rhodium gem-dicarbonyls
was discovered. The shift of the n(CO) vibration allowed under-
standing Rh position in the porous structure, its oxidation
state and the capacity to host dierent chemical species having
dierent stability, especially in the presence of water. These
data are fundamental for understanding the mechanism of
dierent catalytic reactions.
88
According to this methodology,
CO and NO adsorption on zeolithe supported Rh nano-
particles containing dierent promoter elements permitted to
both characterise the eect of the additive and the catalytic
activity of the noble metal.
89
Many studies were carried out on copper oxidation state,
due to its importance when inserted in ZSM-5 zeolites for NO
reduction. Its characteristic carbonyl band at 2158 cm
1
provides quantitative results on integrating its molar extinc-
tion coecient.
90
Hadjiivanov et al.
91
described the water
eect (which is always present during deNOx real conditions)
and reported that bands at 2158 and 2134 cm
1
may charac-
terize CO bound to dry and wet Cu
+
centres, respectively,
the latter being also possibly assigned to a CuO-like phase.
However a comparative study using both Cu-ZSM-5 and
CuO/Al
2
O
3
allowed Praliaud et al. to propose the n(CO)
at 21232133 cm
1
to be due to non-isolated Cu
+
species
(Cu
+
surrounded by Cu
2+
ions) arising from the partial
reduction of bulk CuO. Whereas, the band at 21522157 cm
1
would characterize isolated Cu
+
ions, which are described to
be responsible for the high activity in NO reduction into N
2
.
92
Concerning NO, its adsorption even at room temperature may
lead to Cu
+
oxidation to Cu
2+
and therefore its use for the
determination of copper oxidation state distribution is rather
problematical. However the formation of Cu
+
mono-
and dinitrosyls is observable for high temperature (770 K)
Cu-ZSM-5 activated under vacuum and Datka et al.
93
even
reported the possible existence of two distinct Cu
+
NO
species at 1812 and 1825 cm
1
, associated to two distinct
Cu
+
sites diering in the density of oxygen packing. A
discussion around Cu
+
carbonyls can be also found in
the Strauss
94,95
and Zecchina
96
reports. Further detailing the
characterization of Cu
+
by NO, it has been shown that the
zeolitic structure also inuences the symmetry of the dinitrosyl
species and for a CuMOR sample the Cu
+
NO (1813 cm
1
)
transformation into Cu
+
(NO)
2
leads to dierent spectral
feature depending on the Cu
+
location: a doublet with
n
s
and n
as
at 1828 and 1730 cm
1
in the main channels
and another doublet with n
s
and n
as
at 1870 and 1785 cm
1
in the constrained side-pockets.
97
The use of both probes can also be interesting when each one
is sensitive to an oxidation state of an element, as in the case of
copper, probed by NO in the Cu
2+
state, whereas CO adsorp-
tion would be more specic to the Cu
+
state (since CO com-
plexes with Cu
2+
are only stable at very low temperature).
98
Another example concerns the adsorption of both CO and
NO as informative for the determination of the oxidation state
of vanadium on vanadiatitania catalysts. Although the
carbonyl bands for V
4+
CO, V
3+
CO and Ti
4+
CO species
almost coincide,
99
the fact that NO forms dinitrosyls with V
n+
but not with Ti
4+
allows the eective use of NO as a probe
molecule.
100
Iron-containing ZSM-5 catalysts were also studied for their
potential application in deNOx by hydrocarbons. Interesting
results identifying dierent Fe
2+
species as active sites for
NO
x
SCR with propene can be found in ref. 101 and 102. Fe
3+
characterization in the Y structure was also previously studied
Fig. 9 Integrated intensity of the n(CO) bands bound to platinum as
a function of introduced CO amount (reproduced from ref. 85). The
abscissa of the line intersection corresponds to a CO monolayer on
the exposed metal surface. This allows calculating the number of
accessible Pt atoms, i.e. the metal dispersion. Reproduced by permission
of Elsevier.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 49284950 4939
using CO adsorption,
103
even if IR spectroscopy of probe
molecules is more suitable for characterization of Fe
2+
than
Fe
3+
cations. More recently, we tried to resolve this inter-
pretation conict present in the specialized literature, reporting
NO adsorption followed by infrared spectroscopy to charac-
terize iron cations in Fe-ferrierite.
104
Dierent iron sites
forming mononitrosyl species were identied. The comple-
mentary use of Mo ssbauer spectroscopy enabled us to deter-
mine that the iron oxidation state is essentially +2. For low Fe
loading (by ion exchange), the main fraction of Fe
2+
cations is
suggested to be located in highly accessible positions of the
ferrierite, where ionic exchange takes place in the easiest
way. When the amount of Fe is increased, a second site in a
less accessible position is detected. When an oxidative
pretreatment is applied, only the iron cations in the conned
positions lead to the formation of Fe
3+
OH species. More-
over, NO appears to be able to form polynitrosyl species with
these conned Fe
2+
cations. It thus appears that both the
oxidation and the coordination states of conned Fe
2+
may
change easily, which makes them excellent candidates for
active redox sites.
104
Combining CO and NO as molecular
probes, we were able to go into very ne detail for site
characterisation in Fe-FER. It was ascertained that type I
sites are the most populated and the Fe
2+
ions located in the
so-called G-positions are the most symmetric ones. Their
unique geometry allows two guest molecules approaching
the site from dierent cages, resulting in the formation of
monocarbonyls (2195 cm
1
), converted then, at low tempera-
tures, into dicarbonyls (2188 cm
1
). Type I Fe
2+
cations are
hardly sensitive to oxidizing treatment and show little tendency
to yield Fe
3+
cations. Type II sites are less populated and less
symmetric: Fe
2+
ions in these B-positions form, with CO,
monocarbonyls (2189 cm
1
) and, with NO, mononitrosyls
(1880 cm
1
) practically coinciding in wavenumber with the
nitrosyls formed with type I Fe
2+
ions. These cations are
sensitive to oxidizing treatment and are easily oxidized to
Fe
3+
ions most probably associated to the formation of
a-oxygen species. When the iron concentration in the samples
increases, a third site (F-site) is occupied. Iron ions in this
position change easily and reversibly their oxidation state from
Fe
2+
to Fe
3+
thus forming Fe
3+
OH or Fe
3+
O

species.
When in the Fe
2+
state, iron ions form the most stable
carbonyl species (2196 cm
1
) which can be converted, at
low temperature, into di- (B2188 cm
1
) and tricarbonyls
(B2180 cm
1
). With NO these Fe
2+
ions form nitrosyls
absorbing at 1895 cm
1
. With time, in NO atmosphere, the
Fe
2+
cations are displaced from their original positions in
order to form tetranitrosyl species.
105
The specic structure of MOF compounds gives particular
adsorption, separation and catalytic properties to these
materials. For example, the controlled reduction of a large-pore
iron(III) trimesate with unsaturated iron sites, MIL-100(Fe),
strongly increases the strength of interaction with unsaturated
gas molecules, such as propylene and CO, that have either a
double or triple bond. Therefore, this property leads to a
dramatic improvement of not only preferential gas sorption
but also separation performance for the investigated hybrid
compound: it could be used for the removal of CO impurity to
protect the deactivation of Pt electrodes in low temperature
fuel cells, for example, or the removal of CO from CO
2
-rich
mixtures arising from the production of hydrogen from
biomass. The use of mixed-valent MIL-100(Fe) may be also
considered for further applications involving the selective
separation or purication of olens and acetylenes from
hydrocarbon mixtures. The reducibility of MIL-100(Fe) to
form Fe
II
CUS has been unambiguously shown by in situ IR
spectroscopic analysis using CO as a probe, which also
allowed to quantify the concentration of Fe
2+
/Fe
3+
sites as
a function of the activation treatment (see Scheme 2 and
Fig. 10). Moreover it provided evidences that unsaturated
sites can be created not only through the removal of water
but also to a much lesser extent from the departure of other
Scheme 2 Representations of MIL-100(Fe): (a) one unit cell, (b) two types of mesoporous cages shown as polyhedra, (c) formation of
Fe
III
CUS and Fe
II
CUS in an octahedral iron trimer of MIL-100(Fe) by dehydration and partial reduction from the departure of anionic ligands
(X

= F

or OH

) (from ref. 106). Reproduced by permission of Wiley-VCH.


D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4940 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
molecules, such as trimesic acid or anionic ligands (F and OH)
coordinated on terminal sites of Fe
III
.
106
As already mentioned, ceria and ceriazirconia are catalyst
components of primary importance, notably for their red-ox
properties. This arises from the ability of the Ce cation
oxidation number in ceria and ceria-zirconia to easily change
between 3+ and 4+. The surface state (reduced or oxidized)
and composition (ratio between cerium and zirconium cations)
is available using methanol (CH
3
OH) adsorption.
107
Its dissocia-
tion leads to cation coordinated methoxy and hydroxyl
formation. Both the n(CO) wavenumbers associated to methoxy
species
108,109
and the n(OH) associated to surface hydroxyls
110
depend on the cerium oxidation state. The Ce
4+
/Ce
3+
surface
ratio is thus available from the quantitative study of the
corresponding methoxy intensities. Fig. 11 shows the con-
secutive adsorption of oxygen (O
2
) calibrated doses after
methanol dissociation over pre-reduced ceriumzirconium
mixed oxide which enable the determination of the oxygen
storage capacity of the sample.
Moreover, the methoxy species is very sensitive to the local
coordination site and it allows discriminating between the
dierent cations on a surface. Concerning ceriazirconia solid
solutions, for example, methoxys on Ce
4+
and Zr
4+
surface
sites present specic IR ngerprints. Therefore, the surface
cationic composition can easily been obtained upon methanol
adsorption at room temperature.
25
Rare-earth compounds
also present electronic transitions, arising for ions in internal
structural defects; for ceria, at temperatures above 523 K
a new band appears at 2120 cm
1
and is attributed to the
2
F
5/2
-
2
F
7/2
electronic transition of Ce
3+
, thus indicating the
beginning of bulk oxide reduction.
111
Accessibility of sites. To be active in a catalytic reaction, a
surface site must be accessible to reactants and products.
Isotope labelling is here again a very useful tool. Deutera-
tion of the surface OH groups is only possible if they are
accessible to the deuterated molecule used for the exchange. In
silica, for example, internal silanol groups are distinguished
from external ones based on accessibility to deuterated water
molecules.
112
An accessibility index (ACI) was derived from infrared
spectroscopy of substituted alkylpyridines with dierent sizes
(pyridine: 0.57 nm, 2,6-lutidine: 0.67 nm, collidine: 0.74 nm)
over hierarchical ZSM-5 crystals. The samples were prepared
by selective silicon extraction of a parent commercial sample in
alkaline medium (desilication) and contained dierent degrees
of intracrystalline mesoporosity. The enhanced accessibility of
acid sites in the hierarchical zeolites was shown. A relatively
bulky molecule such as collidine, which probes practically no
acid sites of the parent medium-pore MFI structure, can access
up to 40% of the Brnsted sites in the mesoporous sample.
The ACI is a powerful tool to standardize acid site accessibility
in zeolites and can be used to rank the eectiveness of
synthetic strategies towards hierarchical zeolites (mesoporous
crystals, nanocrystals, and composites).
113
Time resolved studies of probe molecules: 2D-pressure jump
spectroscopy. Time-resolved IR spectroscopy can be used in
probe molecule studies. Microsecond infrared spectroscopy
can be used to monitor adsorbed probe molecules after a
pressure jump on the surface of a porous catalyst. The analysis
of the time behavior of the adsorbed molecule can be obtained
by a Fourier Transform of the IR spectra over time. A 2D map
is obtained, showing frequency response vs. IR spectra for the
adsorbed probe molecule on the catalyst. This technique has
Fig. 10 (a) IR spectra of MIL-100 under a stream of 10% CO at 25 1C after activation under a helium ux at various temperatures over 3 or 12 h.
(b) Amount of Fe
III
CUS and Fe
II
CUS detected by IR analysis upon CO adsorption at 173 1C on MIL-100(Fe) activated under high vacuum at
dierent temperatures (from ref. 106). Reproduced by permission of Wiley-VCH.
Fig. 11 Progressive reoxidation of a mixed Ce
0.80
Zr
0.20
O
2
compound
by introduction of small doses of O
2
after reduction under hydrogen
at 673 K.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 49284950 4941
been used to study platinum catalysts supported on MFI
zeolites. The 2D map can be read as a map showing loca-
tion vs. size for the metal particles. Small particles can be
distinguished from large ones, and their location can be
determined in the pores or on the outer surface of the zeolite.
The role of the pores was demonstrated for the protection
of small particles during ageing of the catalyst, and sintering
was limited by the pore diameter where the particle was
located.
86
Quantitative analysis by coupling IR with gravimetry. Thermo-
gravimetric analysis (TGA) allows monitoring weight changes
in the sample. It has been combined with IR to give new
information on surface sites. Reliable quantitative information
is the key to understanding the catalytic role of surface sites,
and this combination of techniques oers just that. It was used
to determine the molar absorption coecients for IR bands of
OH groups on silica and HY zeolites, as well as for adsorbed
probe molecules.
112
Important information was obtained
about the quantity of the OH groups located in the dierent
cages of HY zeolite and also about the quantity of inner and
surface silanol groups on silica. The number of silanol groups
accessible to water molecules was shown to be constant
whatever the sample of precipitated silica, as well as the ratio
of water/silanol groups under room atmosphere. Combined
thermogravimetry and IR was also applied to operando con-
ditions, and denoted as AGIR (Analysis by Gravimetry and
IR): a modied microbalance was used to follow mass changes
(mg accuracy) inside a catalytic reactor equipped with infrared
windows. The spectroscopic response of water and ammonium
ions coadsorbed together on zeolites was shown to vary depending
on the conditions. The molar absorption coecients for d(H
2
O)
and d(NH
4
+
) at 1640 and 1540 cm
1
for water and ammonia on
a HY zeolite were studied under dry gas ow at variable
temperature. It showed the inuence of coverage on the infrared
response of adsorbed species in zeolites. Adsorption sites change
with coverage, and bands are shifted and their shape and
intensity are modied. Other interesting facts were observed:
water modies strongly the aspect of the d(NH
4
+
) vibration
band in ammoniated zeolites, without changing the absorption
coecient. Measuring the sample mass while at the same time
recording its infrared spectrum showed the key importance of
conditions under which the measurement is done. The presence
of co-adsorbed species (water in particular) strongly modies the
spectrum of surface species. Under reaction condition, this new
technique is especially important for a correct assignment of
infrared features and catalytic behaviour, and above all it makes
IR measurements really quantitative.
114
Other coupled techniques. As we have mentioned in the
introduction, the purpose of this review is to critically discuss
the characterization of the active site in catalytical processes
by using IR analysis compared with catalytic tests. Sometimes,
additional information can be provided by the coupling of
complemental techniques able to yield enhanced insight into
material local properties (structural, textural, physical, . . .)
and reactional events at the nano-scale. For more information
about this point we refer the reader to specic literature,
notably within the present themed issue. We just want to
briey mention a few examples:
Numerous techniques have been coupled with IR, thus
broadening the experimental data set for subsequent mecha-
nistic considerations. Of course, complementary information
can be obtained by parallel experiments; nevertheless, to be
sure that results are comparable, it is often preferred that the
investigation is carried out in a single setup.
115
The most natural technique to associate with IR is certainly
Raman, to cover the totality of the vibrational spectrum. The
rst in situ Raman and FTIR spectroscopic characterization of
a catalytic system under reaction conditions using a single
bench-top instrument with a dedicated cell was published in
2003 by Payen et al.
116
Following both the evolution of the
molecular structure of the active phase by Raman spectro-
scopy and the nature of the dierent surface adsorbed species
by IR and Raman spectroscopy during deNOx reaction they
were able to provide an exhaustive identication of the
adsorbed species and the corresponding coordination sites.
Similar information can also be obtained by coupling
DRIFT with hard X-ray diraction, which can highlight
phenomena taking place in nanoparticulate catalysts, such as
the formation of simultaneous surface and bulk species and
their evolution.
117
Also combining, at high time resolution, a transmission
based structural probes, dispersive EXAFS, with diuse
reectance infrared spectroscopy, can highlight fundamental
steps occurring during gassolid interactions, like that of
oxidation and reduction of alumina-supported Rh at 573 K
using NO and H
2
, and the structuralreactive role of linear
Rh-nitrosyl species within these processes.
118
Chemometric tools. Chemometrics is of course a very ecient
way of decomposing complex bands, when they are progressively
changing during reaction or during progressive adsorption. CO
adsorption on Pt sites leads to complex bands on zeolites where
dierent metal particle sizes can exist.
86
Small particles lead to a
stronger inuence of corners and edges compared to faces, and
large particles produce bands with a stronger inuence of large
faces. Chemometrics allows separating contributions, and the
spectra of CO on small and large particles can be distinguished
quantitatively. This was revealed to be very important for
studying the inuence of ageing of the catalyst on the catalytic
sites.
86
Active sites intermediates and spectators
(discrimination of the aforementioned sites
by using catalytic evidences)
Intermediates, active sites and spectator species are well
dened concepts in catalysis. We will summarize them again
here for a clear discussion of their spectroscopic characteri-
zation. Heterogeneous catalysis involves adsorption of
reactants on the surface, their reaction on the active site, with
the possible formation of intermediates species, and the
formation of nal products (with a possible change of adsorp-
tion site), and the further desorption of the products which will
go to the exit of the reactor.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4942 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
During the adsorption step, several sites can participate.
Some of them will indeed be the place for the reaction
(and these are the active sites) while some others will only be
adsorption sites on which no reaction will take place. These
non-active sites can adsorb strongly reactants or products
and block any movement of surface species, resulting in a
deactivation of the catalyst. Several reactions can take place
on the surface. Some species will have no role in the reaction
under study, and will thus be spectator species. Some other
species can hamper or slow down the reaction, either by
staying adsorbed on the active site or by blocking the circula-
tion of surface species, and will be poison species. Another
type of species will be formed on the surface but will be later
transformed into the nal expected product, and these are
intermediate species. Intermediates are stable species on the
surface (at least for a short duration), and should not be
confused with transition states. Transition states are postulated
between two chemical species on the surface, but do not actually
exist since they have no actual lifetime. They are only pathways
between two states on the surface (on an active site).
Correlation between probe studies and activity
In situ analyses of a surface provide photographs of the
material state, detailing the concentration of the supercial
entities directly by their corresponding IR spectra or via the
spectral response of a molecule adsorbed on the site and their
relative strengths vs. the applied probe, as aforementioned.
The most general way to link catalytic activity to a specic site
is the correlation method. It consists of measuring activities of
a series of materials (starting from the initial reaction rate, or
the conversion at the steady state), and to compare this value
to the number of sites estimated by the intensity of a specic
IR band. In general, this site concentration corresponds to the
entities able to adsorb probe molecules. If a linear correlation
between these two sets of data is found, the activity can
reasonably be attributed to the presence of such adsorption
sites. Nevertheless, all the sites able to adsorb a molecule are
not necessarily active: the presence of the correlation means
that the ratio between the active and the adsorption sites
remains constant when their global amount varies. Additionally,
we should consider the limitations intrinsic to the probe
molecule (probe sensitivity to the strength, nature and number
of sites) and the diculties to obtain for both spectroscopic
and catalytic measurements in similar activation conditions.
The use of appropriate probe molecules permits nevertheless
to formulate reliable hypotheses on the nature of the corres-
ponding sites. Moreover, when the quantitative analysis of the
sites is possible, a calculation of the turn over frequency (TOF)
can also been performed. A demonstrative example concerns
active site study via CO adsorption in hydrotreating reactions
such as hydrodesulfurization (HDS). They are usually related
to anionic vacancies (coordinatively unsaturated sites, CUS)
located on the edges of mixed sulde nanosized particles
(CoMoS or NiMoS) supported on high specic surface
area alumina. Nitrogen monoxide has been the most employed
but a partial oxidation of the sulde phase may occur, even at
very low temperatures. Infrared spectra of CO adsorbed on
the sulded promoted CoMo/Al
2
O
3
catalysts displays a strong
nCO band at 2070 cm
1
which is correlated to the HDS
catalytic activity.
119
It has been assigned to CO interacting
either with a Co atom or with a Mo atom adjacent to a Co
atom. Similar correlations between the intensity of specic nCO
bands and the activity in HDS catalytic activity have been also
found with Mo/Al
2
O
3
catalyst. Due to the drastic procedure of
activation of sulde catalyst, it is worth noting that some
researchers of our laboratory have designed a new IR cell,
called CellEx, in order to characterize in situ sulded catalysts
under a pertinent H
2
S/H
2
ow, with pressure varying from 0.1
up to 4.0 MPa. This cell allows obtaining similar suldation
procedure for both IR characterizations and catalytic tests.
120
Another signicant example can be mentioned in the case of
the NO
x
selective reduction by hydrocarbons on oxide-based
catalysts. Working on Ag/Al
2
O
3
samples for NO
x
reduction
by ethanol, we observed the formation of cyanide and
isocyanate species. The IR band assignment was not straight-
forward: contrarily to what is usually reported in the litera-
ture, n(NCO) of Ag
I
(NCO) species is located at 2204 cm
1
(and not at 2230 cm
1
), whereas that of Ag
0
(NCO) species is
at 2243 cm
1
. Thus, during SCR reaction of NO on silver/
alumina catalyst, the isocyanate species generally observed as
intermediate compounds around 2230 and 2260 cm
1
are not
linked to silver sites but to the support, the main role of silver
being to favour NCO formation and concentration on the
support. The hypothesis that the two observed bands are due
to dierent alumina coordination sites (Al
octa
and Al
tetra
) for
isocyanates was determined as the most probable by isotopic
substitutions.
121
This information was useful for the compre-
hension of the NO SCR pathway. In fact comparing the
selective catalytic reduction of NO
x
in excess of oxygen using
ethanol as reducing agent on silver/alumina and on bare
alumina showed the connection between the presence of silver
and isocyanate species on a catalyst. Then, a detailed investi-
gation concerning these groups was undertaken to understand
their formation, their localization, and their reactivity in order
to propose the pathway of the NO
x
SCR into N
2
. Three
elemental sequences were suggested explaining, rst, the
formation of silver cyanide and its transformation into
Al
3+
NCO, then isocyanate hydrolysis into ammonia, and
nally the reaction of the latter species with NO in the
presence of oxygen giving rise to nitrogen (Fig. 12).
122
However, a working surface can have a totally dierent
behaviour, with sites changing their nature. The real valence of
a site can be ascertained only catching it in full action, using
operando techniques.
Operando studies on Fe-FER indicated that these samples
present interesting NO
x
SCR eciency for temperature as low
as 433 K. Our investigations also indicate that for such
low reaction temperature ammonia and nitrogen monoxide
compete for adsorption onto the Fe
2+
species (whose ne
characterisation has been reported in the section Metallic and
redox sites), which are active for the NO-to-NO
2
oxidation.
Our results are also consistent with this last reaction being
the rate determining step of the global SCR process. We nally
concentrated our study on the eect of SO
2
on this NO-to-NO
2
reaction. We must conclude that for the 2.5 wt% Fe-FER
sample, the majority of iron sites are poisoned by sulfate
formation, although some Fe
2+
sites are thio-resistant.
123
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 49284950 4943
Operando IR spectroscopy shows out all its added values
when very complicated and multifunctional systems such as
NO
x
-trap catalysts are investigated. In this case the spectro-
scopic evidences are the only support for material modelling
and scale-up, since kinetic calculations are totally ineective
in the transient reaction conditions. On the contrary, the
synchronous analysis of gaseous and adsorbed phases all
along the experiments gave highly valuable information in
order to establish consecutive steps of NO
x
adsorption and
reduction on this kind of materials. It has, therefore, been
established that this reaction proceeds initially through NO
oxidation. The NO
2
produced that way is subsequently adsorbed
on storage basic sites (for example Ba sites for a typical Toyota
NO
x
-trap catalyst, or potassium sites in the case of the
formulation developed with Rhodia: Pt/K/Mn/Al
2
O
3
CeO
2
)
in the form of nitrite species. The last step of this mechanism is
thus the oxidation of nitrites to nitrates, probably involving
NO
2
oxidizing molecules. Two dierent periods have also been
determined for the NO
x
storage process, and linked to the
progressive adsorption of NO
x
all along the catalytic reactor,
as well as to the presence of two types of adsorbing sites, i.e.
surface and bulk alkaline or earth-alkaline sites.
75
Concerning
the catalyst working under the alternate lean/rich ow, unique
dynamic response details can be obtained by this technique, as
illustrated by Fig. 13.
Surface and gas species comparison by time resolved analysis
(as reported in Fig. 14 for the system described above) is the
best way to establish reaction pathways. It is possible, for
example, to observe at what time dierent intermediates or
byproducts are emitted and to deduce which site is responsible
for a given catalytic issue.
In such a way it is possible, step by step, to reconstruct all
the reaction mechanism as for a jigsaw puzzle, evidencing the
dierent steps, the chemical species taking part into them and
the sites at play for the catalytic reaction, as depicted in Fig. 15
for the Pt/K/Mn/Al
2
O
3
CeO
2
NO
x
-trap catalyst.
124
Basicity: identication of the active sites in the reaction of COS
hydrolysis
The COS hydrolysis reaction (COS + H
2
O -CO
2
+ H
2
S) is
generally used to eliminate COS from gaseous euent and
improves the Claus process, an industrial process to transform
H
2
S into sulfur. The comparison of the results from catalytic
test and from infrared experiments using CO
2
as acidic probe
molecule on Al
2
O
3
, ZrO
2
,TiO
2
and ZnO, has revealed that the
OH-basic sites (type I according to the nomenclature of
Kno zinger
16
) are the active sites of the reaction, except for
ZnO, inactive for this reaction.
77
Interaction of COS with
these basic sites leads to hydrogen-thiocarbonates according
to the reaction MOH + COS - M(HSCO
2
)

, which play
the role of intermediates in the COS hydrolysis reaction
(Fig. 16).
It was concluded that this reaction can be used as a test
reaction for the basicity of OH groups of metal oxides.
77
Selective site poisoning: application to acid reactions
One of the most powerful methods to determine the nature of
the active sites in a catalytic reaction is the selective poisoning
of the adsorption sites, added before or during the reaction a
molecule, which will remain irreversibly adsorbed. In this way,
some adsorption sites will be specically blocked and the
consequence on the solid activity will be clear. The operando
IR reactor-cell is the ideal tool using a specic probe molecule,
because it allows monitoring the surface poisoning in a
qualitative (nature of sites aected by poisoning) and quanti-
tative way (amount of poisoned sites) in parallel to the changes
of the catalytic activity (conversion and selectivity). As examples
we can cite the study of the activity of the dierent Brnsted-
acid site of HY zeolites in n-hexane cracking
125
and cyclo-
hexene conversion,
126
or the role of the acidbase pairs of
modied alumina in tetrahydrothiophene conversion.
10
If site
poisoning is due to the reaction products themselves, when
those are spectroscopically identied we are in the ideal
case for the operando study, under real conditions, with the
possibility to determine deactivation mechanisms.
What is visible by IR during the reaction?
The question of what can be detected by IR spectroscopy
during the reaction naturally follows. IR is the spectroscopy of
chemical bonding, not of individual atoms. IR can directly
detect sites, on the condition that they are made of several
atoms (as in Brnsted sites with the OH group), and it can
detect species formed by adsorption of reactants. During
the reaction of hydrocarbons on a protic acid catalyst, for
example, the reactant is adsorbed on the Brnsted-acid site
before the reaction takes place. The Brnsted site is perturbed
by this adsorption, and the intensity of the n(OH) vibration
band for the free OH group decreases. After the catalytic cycle
has taken place, the hydrocarbon is desorbed, and the n(OH)
vibration band reappears. At a given moment on the surface,
the ratio of free to occupied acid sites follows a statistic
distribution reecting the coverage level. When the tempera-
ture increases, the adsorption equilibrium shifts in favor of the
gas phase, and the relative amount of occupied sites decreases,
with a corresponding decrease of the visible perturbation of
the infrared band. The reaction speed on the catalytic cycle
nevertheless increases, and the catalytic conversion increases
while the number of active sites in play seems to decrease. This
is the classical paradox of operando spectroscopy: active sites
seem to take no part in a very active catalyst. The reason is
Fig. 12 Proposed formation mechanism of NCO groups in a CO + NO reaction on Ag/Al
2
O
3
catalyst.
122
Reproduced by permission of Elsevier.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4944 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
simply that at a given moment the relative amount of active
sites interacting with surface species is too low to be detected,
it does not reach the detection limit of the spectroscopic
technique. In an ecient and active catalyst, the biggest part
Fig. 13 Example of IR analysis of the surface of a NO
x
-trap catalyst (Pt/K/Mn/Al
2
O
3
CeO
2
) submitted to an alternate rich/lean ow. The
presence of nitrate, carbonate and isocyanate species (mutually replacing during the rich/lean phases) is highlighted. In the upper part of the gure,
the trend of the integrated intensities of the bands assigned to nitrates (blue), carbonates (orange) and isocyanates (green) is plotted vs. time.
124
Fig. 14 Intensity evolution vs. time of the intensities of the main
surface (nitrates and isocyanates) and gas species (NO
x
, ammonia, . . .)
when the sample is alternatively exposed to the lean (light blue
background) or rich (pink background) ows at 400 1C. The
zoomed part corresponds to N
2
production overlapping the CO signal
in MS. The vertical blue dashed line represents the time limit to
obtain a maximum in selectivity, skipping isocyanate and ammonia
formation.
124
Fig. 15 Schematic representation of the reaction mechanisms taking
place on a Pt/K/Mn/Al
2
O
3
CeO
2
NO
x
-trap catalyst during a lean
(upper part) and rich ow (lower part). In particular, the oxidation
and storage sites are highlighted, showing the superuous role of
platinum during the lean step. During the rich period, nitrate reduc-
tion via the exothermic CO and H
2
oxidation is evidenced, as well as
the dissociation of NO
x
species on metal particle sites. The presence of
secondary reactions is also mentioned. All the species highlighted in
yellow and the respective coordination sites can be observed and
quantied by IR spectroscopy.
124
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 49284950 4945
of active sites are not in interaction with products, and the active
site spends most of its time at rest. Limiting steps can only be
observed by decreasing activity (by decreasing temperature for
example) and by slowing down the system. Adsorbed species
need to accumulate enough to get over the detection limit.
Another way to concentrate species on a site over a limited
period of time is to use transient conditions: sending a pulse of
one of the reacting agents suddenly increases the concentration
of the intermediates in the reaction chain, making them visible.
Data treatment methods can be a very valuable help for
spectroscopic studies of active sites. Operando often leads to
large amounts of data, and the eye might not spot meaningful
small band variations, especially when the species detected is
not expected from the known reaction mechanism. 2D correla-
tion spectroscopy (2D-COS) or chemometric techniques
such as principal component analysis or MCR-ALS allows
extracting meaningful information.
2D-COS was rst used in catalysis in 2000, and it allowed
extracting information from fairly noisy spectra at high
temperature. It showed that the reason for improved selectivity
of H-MFI for para-xylene over time-on-stream was due to the
lling of pore defects by coke related species.
127
After this
lling of the defects (detected by the disappearing of their
spectral signature, correlated with the signals for coke species),
the pore structure was restored to its initial shape, and so was
the eciency for selecting para from the mixture of the three
isomers of xylene. More recently, 2D-COS was used to identify
catalytically active sites amongst the various Brnsted sites in
hierarchical zeolites.
128
Kinetics and surface concentrations
The surface elementary reaction rate is proportional to the
surface concentration of the adsorbed species involved in
the reaction process (LangmuirHinshelwood theory). An
increase of this concentration will result in a corresponding
increase of the reaction speed. Inversely, any change aecting
the reaction speed will modify the surface concentration of the
intermediate surface species. The reactant pressure is generally
an eective way for such a perturbation.
The determination of real reaction intermediates in the
synthesis of methanol from CO and H
2
is dicult, and many
surface species appear on the catalyst: carbonyls, formates,
carbonates and methoxy groups. This reaction has been
studied on zinc aluminate supported copper catalysts, and it
is a very good example of the combined use of operando
spectroscopy and on-line analysis for the determination of
mechanism and reaction intermediates.
129
Two separate reac-
tion mechanisms are possible. Insertion of CO under a surface
OH group can lead to surface formate groups, further con-
verted into methoxy surface groups, and leading to methanol
after hydrogenation. In the case of an active copper catalyst,
neither copper formate nor copper methoxy species are detected:
methoxy groups on the support only build up (in high amounts)
on the surface. Desorption is the limiting step, and the
methanol formation rate is linked to the product of surface
concentrations of methoxy groups and hydrogen. Both species,
methoxy groups on the support and hydrogen on copper, are
located on two separate adsorption sites. The second possible
route is hydrogenation of copper carbonyls via formyl species.
The corresponding kinetic equation is the product of surface
concentrations of carbonyl groups and of hydrogen. To check
which is the main mechanism, the approach used was to keep
hydrogen partial pressure constant, while varying that of
carbon monoxide. In the formate mechanism, the equation
becomes very simple. The surface concentration of dissociated
H
2
is constant and the reaction speed only depends on the
surface methoxy group concentration:
V
MeOH
= k
1
[CH
3
O][H] = k
1
0
[CH
3
O]
The formate mechanism was easily excluded, since the surface
methoxy concentration clearly was not linearly linked to the
amount of methanol produced. In the case of the carbonyl
mechanism, the LangmuirHinshelwood theory leads to two
equations, depending on the competitive or non-competitive
adsorption of CO and hydrogen. If the adsorption is non-
competitive, and with constant hydrogen pressure:
V
MeOH
= k
2
00
(K
CO
P
CO
)/(1 + K
CO
P
CO
)
For the competitive adsorption, the result is slightly
dierent:
V
MeOH
= k
2
00
(K
CO
P
CO
)/(1 + K
CO
P
CO
)
2
K
CO
was measured easily because the surface carbonyl
concentration is directly linked to P
CO
. The equations corres-
ponding to the competitive and non-competitive adsorption
were t to the experimental point at the highest P
CO
. A very
nice t was obtained for the competitive adsorption (Fig. 17).
Thus, the reaction mechanism was determined together with
the adsorption site. Combining surface and gas species in the
kinetic modelling of the system is one of the most powerful
approaches for operando spectroscopy.
Transient studies
Kinetic measurements can be performed with two main
approaches: steady state or transient states. In steady state
studies, changing a concentration in the gas feed will change
Fig. 16 Variation of COS conversion with the area I of the band
d(OH) (at 1220 cm
1
) of HOCO
2

species formed from CO


2
adsorption
on dierent metal oxides.
77
Reproduced by permission of Springer.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4946 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
concentrations in surface species and in products according to
the rate equations, thus allowing assessment of the validity of
the rate equations, as we have seen above. Studying transient
states is performed by taking advantage of the time resolution
of spectroscopic and analytic techniques. Infrared spectra can
be recorded fairly quickly, and a time resolution of 0.1 s is
routine on any commercial FT instrument. In order to draw
conclusions on the kinetics at the time scale of 0.1 s, the whole
chain of events in the catalytic setup needs to be controlled and
determined at the same time resolution (at least). On-line
analyses can be performed quickly with IR gas cells. Mass
spectra can be recorded at ca. 1 data point per second, and fast
GC can lead to a chromatogram per minute. These information
can be combined, and the time scale of the observation will
depend on the technique used. One important parameter is the
time resolution of the catalytic reactor itself. With a long
residence time, gases will be mixed in the reactor, with a
consequent loss in the time resolution of the whole experiment.
Short residence time can only be achieved with a low dead
volume in the reactor. IR micro reactor-cells have been
designed which allow very short residence time and space
velocity of 50 000 h
1
can be achieved. In this sense, our
laboratory has performed considerable eorts in the last two
decades.
Freysz et al.
130
have used two complementary techniques
for gas products analysis: FTIR using a gas microcell
(multireection cell with an inner volume of 88 ml) and mass
spectrometry. The dead volumes are minimized in the whole
system in order to allow time-resolved analysis in particular
inside the IR reactor cell.
131
The time resolution of
both surface analysis and gas analysis is around 12 s. This
methodology was applied to study the reduction of NO by CO
on a silica-supported platinum catalyst by the addition of very
small CO pulses (20 ml) in a continuous NO ow.
130
The
simultaneous qualitative and quantitative analysis and the
correlation of both surface species and gas phase evolution
during the transient-pulse catalytic reaction indicated that a
direct reaction between CO and NO (adsorbed or not) must be
rejected. When a pulse of CO is sent, two distinct successive
periods are observed (Fig. 18). The rst period concerns the
reduction of Pt by CO. The diminution of the NO/Pt concen-
tration at the beginning of the pulse does not correspond to a
signicant reduction of NO (no signicant production of
N
2
/N
2
O), but to a desorption of nitrogen monoxide: the
amount of observed gas NO increases during this period.
The adsorbing CO displaces NO from the surface but does
not react with it. The concomitant production of CO
2
can then
only be explained by a reaction between CO and the initially
Fig. 17 Experimental () rates of methanol formation as a function
of P
CO
. Curve 1 computed assuming a non-competitive adsorption
between CO and H
2
, curve 2 computed assuming a competitive
adsorption (from ref. 129). Reproduced by permission of Elsevier.
Fig. 18 Pulse of 20 ml of CO in a 1000-ppm NO in He continuous ow on a Pt/SiO
2
catalyst; T = 498 K. (A) IR spectra of the gas phase
(one spectrum per 2 s). (B) IR spectra of the surface species (one spectrum per 2 s). (C) Correlation between the surface species and the gas phase IR
analysis.
130
Adapted and reproduced by permission of Elsevier.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 49284950 4947
adsorbed oxygen: CO reduces the platinum surface. The
second period corresponds to the reduction of NO. At the
end of the pulse, CO progressively disappears from the
surface, whereas NO re-adsorbs on the Pt sites. The important
NO consumption does not correspond only to this re-adsorption,
but mainly to an important N
2
/N
2
O production (i.e. to NO
x
reduction). Remaining CO
lin
/Pt
red
undergoes a total oxidation
into CO
2
, because no more CO is observed in the gas phase
and no more carbonaceous species are observed on the surface
(if we except the initial CO
lin
/Pt
ox
spectator species). This
second CO
2
production corresponds to the N
2
/N
2
O forma-
tion, and NO is the only possible oxidant: CO
lin
/Pt
red
is the
origin of NO reduction. At the end of the pulse, the NO/Pt
concentration is higher than at the beginning. In fact, adsorbed
nitrosyl species now occupy Pt sites which were unavailable
before the pulse as other species were already adsorbed
(O
ads
or nitrates). The surface then slowly returns to its initial
equilibrium.
Studying very short lived intermediates cannot be done with
standard catalytic setups and spectrometers. Both parameters
have to be controlled, the time resolution of the catalytic
setup, and of the spectroscopic observation. Very high time
resolution (B10 ns) can be achieved by special recording
techniques such as step scan interferometry. Step scan inter-
ferometry only records one point of the interferogram over
time, and the measurement has to be repeated many times so
that all the points in the interferogram can be collected. The
IR spectra themselves are obtained after Fourier transform of
all the recomposed interferograms. The catalytic experiment
therefore needs to be well controlled in time but also highly
reproducible. If the start of the reaction is well controlled in
time, the various states of the surface can be recorded over
time and the observation can be reproduced so that results are
co-added to lower the detection limit. Fast valves allow
millisecond control over the reaction, which can be coupled
with millisecond step scan interferometry.
132
Nanosecond
control of a reaction is achieved with laser triggered reaction,
mostly in photochemistry.
133
Thermal reactions can also be
observed, and that of CO and NO on Ag supported on
alumina was induced by an ultrafast pulsed laser, after which
the state of the adsorbed species was monitored by step scan
FT IR at 33 ns time resolution.
134
The transformation of
cyanides on silver into isocyanates on alumina is the limiting
step of the reaction. This step could be recorded, and a bridged
intermediate was detected between the silver particle and the
alumina surface. This intermediate only lasts 2 microseconds,
and it leads to the formation of an isocyanide on Al
VI
(Al
VI
is
statistically the most accessible site), which in turn is quickly
oxidized into an isocyanate and transferred to the stronger
Lewis acid Al
IV
. The interface between the Ag particle and the
alumina support plays a key role in the catalytic activity,
and this helps understanding the role of the particle size
(see Scheme 3).
Spatial resolution
A fundamental point for a complete and trustful catalyst
characterisation is the identication of the exact position of
the sites along the catalytic bed. Shaped catalysts are in fact far
from being homogeneous objects: extrudates or pebbles are
often quite heterogeneous (even coreshell objects), while
the washcoating process is more an art than a science and
active phase deposition experts proceed using empirical
techniques. We should also mention that for both monoliths
and pebble shaped catalysts, binders are unavoidable com-
ponents, constituting up to 30% of the total material mass.
Furthermore, monolith shaped catalysts are often multi-
layered, so that the space identication of the site is a real
and crucial problem.
As shown out by Weckhuysen in his recent review about
spatial resolution in catalysis,
135
the few tentative studies of
monitoring a catalyst by infrared light use IR microscopy.
This technique is detailed in another contribution of this issue.
It is interesting to remark that using synchrotron light in spite
of a classical IR source, the spatial resolution is signicantly
boosted, allowing to catch sites in action at an impressive
spatial resolution, such as distinguishing between the dierent
morphological elements of a zeolitic single-crystal.
136
The
combination of synchrotron-based IR spectral imaging and
label-free coherent anti-Stokes Raman scattering can further
enhance the details of a working surface at the level to show
the multidimensional distribution of reactants and products in
zeolite crystals with micrometre resolution, highlighting the
importance of diusion barriers and capillary forces imposed
by a pore network for the catalytic activity.
137
We wish to mention here that very recently we have
obtained the rst spectra of a catalyst shaped onto a monolith
using direct transmission IR spectroscopy, allowing a precise
and quantitative analysis of the adsorbed species, in real
operando conditions (keeping the reaction shape of the honey-
comb, as well as the real reaction gas composition). Preliminary
Scheme 3 Detailed mechanism for the limiting step of the conversion
of NO and CO on alumina supported silver catalyst. The cyanide
intermediate is transformed after the laser pulse into a isocyanide via a
bridged species (lasting 2 ms on the surface) between the silver particle
and an aluminium VI atom on the alumina support.
134
Reproduced by
permission of American Association for the Advancement of Science.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4948 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
results for this achievement were presented at the Operando
III conference
138
and a paper is in progress.
Conclusions
IR spectroscopy can provide valuable and pertinent infor-
mation on the adsorption sites on a catalyst, both directly or
via the adsorption of adapted probe molecules (the best probe
is always the reactants themselves). Physicalchemical properties
of the sites can be described and ranked vs. reference com-
pounds: acidity, basicity, redox properties are veried and
quantied in strength and as site concentrations.
Dierent operando techniques can complete such a view
discriminating which sites play a role in the catalytic reaction,
behaving as active sites or just hosting reacting agents or
products. A catalytic site in action cannot be separated from
its respective intermediate, so that the problem of site detection
during the catalytic reaction is strongly connected to that of
intermediates. In this view, we have also tried to answer the
frequently asked question of what can really be seen by IR
spectroscopy during the reaction process.
Notes and references
1 M. Boudard and G. Dje ga-Mariadassou, Cinetique des reactions
en catalyse heteroge`ne, Masson, Paris, 1982.
2 C. Lamberti, E. Groppo, G. Spoto, S. Bordiga and A. Zecchina,
Adv. Catal., 2007, 51, 174.
3 J. A. Lercher, C. Grundling and G. EderMirth, Catal. Today,
1996, 27, 353376.
4 J. C. Lavalley, Catal. Today, 1996, 27, 377401.
5 E. Payen, J. C. Lavalley, M. Daturi and F. Mauge , in Handbook
of Vibrational Spectroscopy, ed. J. M. Chalmers and P. R.
Grith, Wiley, 2001, vol. 4: Applications in Industry, Materials
and the Physical Sciences, pp. 30053041.
6 F. Meunier and M. Daturi, Catal. Today, 2006, 113, 12.
7 P. Bazin, O. Saur, F. C. Meunier, M. Daturi, J. C. Lavalley,
A. M. Le Govic, V. Harle and G. Blanchard, Appl. Catal., B,
2009, 90, 368379.
8 G. Leofanti, A. Marsella, B. Cremaschi, M. Garilli, A. Zecchina,
G. Spoto, S. Bordiga, P. Fisicaro, G. Berlier, C. Prestipino,
G. Casali and C. Lamberti, J. Catal., 2001, 202, 279295.
9 A. B. M. Saad, V. A. Ivanov, J. C. Lavalley, P. Nortier and
F. Luck, Appl. Catal., A, 1993, 94, 7183.
10 F. Can, A. Travert, V. Ruaux, J. P. Gilson, F. Mauge , R. Hu and
R. F. Wormsbecher, J. Catal., 2007, 249, 7992.
11 A. Badri, C. Binet and J. C. Lavalley, J. Chem. Soc., Faraday
Trans., 1997, 93, 21212124.
12 J. M. Gatica, P. Fornasiero, J. Kaspar, T. Lesage, S. Aiello and
M. Daturi, Phys. Chem. Chem. Phys., 2002, 4, 381388.
13 A. Zecchina, D. Scarano, S. Bordiga, G. Spoto and C. Lamberti,
Adv. Catal., 2001, 46, 265397.
14 G. Spoto, E. N. Gribov, G. Ricchiardi, A. Damin, D. Scarano,
S. Bordiga, C. Lamberti and A. Zecchina, Prog. Surf. Sci., 2004,
76, 71146.
15 S. Coluccia, S. Lavagnino and L. Marchese, Mater. Chem. Phys.,
1988, 18, 445464.
16 H. Knozinger and P. Ratnasamy, Catal. Rev. Sci. Eng., 1978, 17,
3170.
17 C. Morterra and G. Magnacca, Catal. Today, 1996, 27,
497532.
18 F. Lonyi and J. H. Lunsford, J. Catal., 1992, 136, 566577.
19 S. Jolly, J. Saussey, J. C. Lavalley, N. Zanier, E. Benazzi and
J. F. Joly, Ber. Bunsen-Ges. Chem., 1993, 97, 313315.
20 N. Malicki, P. Beccat, P. Bourges, C. Fernandez, A.-A. Quoineaud,
L. J. Simon, F. Thibault-Starzyk, Z. G. J. C. Ruren Xu and
Y. Wenfu, Stud. Surf. Sci. Catal., 2007, 170, 762770.
21 O. Marie, P. Massiani and F. Thibault-Starzyk, J. Phys. Chem. B,
2004, 108, 50735081.
22 O. Marie, F. Thibault-Starzyk and P. Massiani, J. Catal., 2005,
230, 2837.
23 M. Bensitel, O. Saur, J. C. Lavalley and G. Mabilon, Mater.
Chem. Phys., 1987, 17, 249258.
24 A. Vimont, J. C. Lavalley, A. Sahibed-Dine, C. O. Arean,
M. R. Delgado and M. Daturi, J. Phys. Chem. B, 2005, 109,
96569664.
25 M. Daturi, C. Binet, J. C. Lavalley, A. Galtayries and
R. Sporken, Phys. Chem. Chem. Phys., 1999, 1, 57175724.
26 A. Vimont, F. Thibault-Starzyk and J. C. Lavalley, J. Phys.
Chem. B, 2000, 104, 286291.
27 G. Ricchiardi, A. Damin, S. Bordiga, C. Lamberti, G. Spana,
F. Rivetti and A. Zecchina, J. Am. Chem. Soc., 2001, 123,
1140911419.
28 P. J. Kunkeler, B. J. Zuurdeeg, J. C. van der Waal, J. A. van
Bokhoven, D. C. Koningsberger and H. van Bekkum, J. Catal.,
1998, 180, 234244.
29 J. C. Lavalley and M. Benaissa, J. Chem. Soc., Chem. Commun.,
1984, 908909.
30 B. A. Morrow and I. A. Cody, J. Phys. Chem., 1976, 80,
19982004.
31 I. Nova, L. Lietti, L. Casagrande, L. DallAcqua, E. Giamello
and P. Forzatti, Appl. Catal., B, 1998, 17, 245258.
32 L. Capek, V. Kreibich, J. Dedecek, T. Grygar, B. Wichterlova,
Z. Sobalik, J. A. Martens, R. Brosius and V. Tokarova,
Microporous Mesoporous Mater., 2005, 80, 279289.
33 E. M. El-Malki, R. A. van Santen and W. M. H. Sachtler,
J. Catal., 2000, 196, 212223.
34 K. I. Hadjiivanov and G. N. Vayssilov, Adv. Catal., 2002, 47,
307511.
35 M. Mihaylov, K. Chakarova, K. Hadjiivanov, O. Marie and
M. Daturi, Langmuir, 2005, 21, 1182111828.
36 E. Ivanova, M. Mihaylov, H. A. Aleksandrov, M. Daturi,
F. Thibault-Starzyk, G. N. Vayssilov, N. Ro sch and
K. I. Hadjiivanov, J. Phys. Chem. C, 2007, 111, 1041210418.
37 K. I. Hadjiivanov, Catal. Rev. Sci. Eng., 2000, 42,
71144.
38 S. Coluccia, L. Marchese and G. Martra, Res. Chem. Intermed.,
2000, 26, 15.
39 G. Busca, Phys. Chem. Chem. Phys., 1999, 1, 723736.
40 G. Busca, Catal. Today, 1998, 41, 191206.
41 L. M. Kustov, Top. Catal., 1997, 4, 131144.
42 G. Spoto, S. Bordiga, A. Zecchina, D. Cocina, E. N. Gribov,
L. Regli, E. Groppo and C. Lamberti, Catal. Today, 2006, 113,
6580.
43 O. Marie, F. Thibault-Starzyk and J. C. Lavalley, Phys. Chem.
Chem. Phys., 2000, 2, 53415349.
44 A. Zecchina, L. Marchese, S. Bordiga, C. Paze and E. Gianotti,
J. Phys. Chem. B, 1997, 101, 1012810135.
45 F. R. Sarria, O. Marie, J. Saussey and M. Daturi, J. Phys. Chem.
B, 2005, 109, 16601662.
46 F. R. Sarria, V. Blasin-Aube, J. Saussey, O. Marie and M. Daturi,
J. Phys. Chem. B, 2006, 110, 1313013137.
47 F. R. Sarria, J. Saussey, J. P. Gallas, O. Marie and M. Daturi,
Stud. Surf. Sci. Catal., 2005, 158, 821828.
48 T. Montanari, O. Marie, M. Daturi and G. Busca, Catal. Today,
2005, 110, 339344.
49 D. Kaucky , J. Dedecek and B. Wichterlova , Microporous
Mesoporous Mater., 1999, 31, 7587.
50 J. Dedecek and B. Wichterlova, J. Phys. Chem. B, 1999, 103,
14621476.
51 J. Dedecek, D. Kaucky and B. Wichterlova , Microporous
Mesoporous Mater., 2000, 3536, 483494.
52 L. Drozdova, R. Prins, J. Dedecek, Z. Sobalik and
B. Wichterlova, J. Phys. Chem. B, 2002, 106, 22402248.
53 V. Schwartz, R. Prins, X. Wang and W. M. H. Sachtler, J. Phys.
Chem. B, 2002, 106, 72107217.
54 T. Montanari, O. Marie, M. Daturi and G. Busca, Appl. Catal.,
B, 2007, 71, 216222.
55 A. Vimont, J. M. Goupil, J. C. Lavalley, M. Daturi, S. Surble,
C. Serre, F. Millange, G. Ferey and N. Audebrand, J. Am. Chem.
Soc., 2006, 128, 32183227.
56 L. Alaerts, E. Seguin, H. Poelman, F. Thibault-Starzyk,
P. A. Jacobs and D. E. De Vos, Chem.Eur. J., 2006, 12,
73537363.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
This journal is c The Royal Society of Chemistry 2010 Chem. Soc. Rev., 2010, 39, 49284950 4949
57 S. Bordiga, L. Regli, F. Bonino, E. Groppo, C. Lamberti,
B. Xiao, P. S. Wheatley, R. E. Morris and A. Zecchina, Phys.
Chem. Chem. Phys., 2007, 9, 26762685.
58 A. Vimont, H. Leclerc, F. Mauge, M. Daturi, J. C. Lavalley,
S. Surble, C. Serre and G. Ferey, J. Phys. Chem. C, 2007, 111,
383388.
59 Y. K. Hwang, D. Y. Hong, J. S. Chang, S. H. Jhung, Y. K. Seo,
J. Kim, A. Vimont, M. Daturi, C. Serre and G. Ferey, Angew.
Chem., Int. Ed., 2008, 47, 41444148.
60 J. Czyzniewska, S. Chenevarin and F. Thibault-Starzyk, Stud.
Surf. Sci. Catal., 2002, 142, 335342.
61 S. Joly, J. Saussey, A. Janin, F. Thibault-Starzyk and
J. C. Lavalley, J. Chim. Phys., 1997, 94, 18381847.
62 F. Thibault-Starzyk, A. Travert, J. Saussey and J. C. Lavalley,
Top. Catal., 1998, 6, 111118.
63 K. S. Smirnov and F. Thibault-Starzyk, J. Phys. Chem. B, 1999,
103, 85958601.
64 A. Travert, A. Vimont, J. C. Lavalley, V. Montouillout,
M. R. Delgado, J. J. C. Pascual and C. O. Arean, J. Phys. Chem.
B, 2004, 108, 1649916507.
65 M. Daturi, A. Vimont and J. M. Wineld, in Functionalized
Inorganic Fluorides, ed. A. Tressaud, Wiley, Chichester, 2010,
pp. 101139.
66 A. Zecchina, C. Lamberti and S. Bordiga, Catal. Today, 1998, 41,
169177.
67 H. Knozinger and S. Huber, J. Chem. Soc., Faraday Trans., 1998,
94, 20472059.
68 D. Barthomeuf, Catal. Rev. Sci. Eng., 1996, 38, 521612.
69 O. Marie, N. Malicki, C. Pommier, P. Massiani, A. Vos,
R. Schoonheydt, P. Geerlings, C. Henriques and F. Thibault-
Starzyk, Chem. Commun., 2005, 10491051.
70 A. Michalska, M. Daturi, J. Saussey, I. Nowak and M. Ziolek,
Microporous Mesoporous Mater., 2006, 90, 362369.
71 C. Binet, M. Daturi and J. C. Lavalley, Catal. Today, 1999, 50,
207225.
72 M. Daturi, C. Binet, J. C. Lavalley and G. Blanchard, Surf.
Interface Anal., 2000, 30, 273277.
73 S. Roy and A. Baiker, Chem. Rev., 2009, 109, 40544091.
74 T. Lesage, C. Verrier, P. Bazin, J. Saussey, S. Malo, C. Hedouin,
G. Blanchard and M. Daturi, Top. Catal., 2004, 3031, 3136.
75 T. Lesage, J. Saussey, S. Malo, M. Hervieu, C. Hedouin,
G. Blanchard and M. Daturi, Appl. Catal., B, 2007, 72, 166177.
76 S. Valange, A. Beauchaud, J. Barrault, Z. Gabelica, M. Daturi
and F. Can, J. Catal., 2007, 251, 113122.
77 A. Aboulayt, F. Mauge , P. E. Hoggan and J. C. Lavalley, Catal.
Lett., 1996, 39, 213218.
78 A. Aboulayt, PhD thesis, El Jadida, 1999.
79 A. Sahibed-Dine, A. Aboulayt, M. Bensitel, A. B. M. Saad,
M. Daturi and J. C. Lavalley, J. Mol. Catal. A: Chem., 2000,
162, 125134.
80 F. Mauge , C. Binet and J.-C. Lavalley, in Catalysis by Metals, ed.
H. Jobic, Springer-Verlag, Berlin, 1997, p. 6L1.
81 C. Binet, A. Jadi and J. C. Lavalley, J. Chim. Phys., 1989, 86,
451470.
82 C. Binet, A. Jadi, J. C. Lavalley and M. Boutonnetkizling,
J. Chem. Soc., Faraday Trans., 1992, 88, 20792084.
83 A. Badri, C. Binet and J. C. J. Lavalley, J. Chim. Phys., 1995, 92,
13331343.
84 P. Bazin, O. Saur, J. C. Lavalley, M. Daturi and G. Blanchard,
Phys. Chem. Chem. Phys., 2005, 7, 187194.
85 V. Perrichon, L. Retailleau, P. Bazin, M. Daturi and
J. C. Lavalley, Appl. Catal., A, 2004, 260, 18.
86 M. Rivallan, E. Seguin, S. Thomas, M. Lepage, N. Takagi,
H. Hirata and F. Thibault-Starzyk, Angew. Chem., Int. Ed.,
2010, 49, 785789.
87 E. Ivanova, M. Mihaylov, F. Thibault-Starzyk, M. Daturi and
K. Hadjiivanov, J. Mol. Catal. A: Chem., 2007, 274, 179184.
88 E. Ivanova, M. Mihaylov, F. Thibault-Starzyk, M. Daturi
and K. Hadjiivanov, J. Catal., 2005, 236, 168171.
89 M. Lepage, T. Visser, F. Soulimani, A. Iglesias-Juez and
B. M. Weckhuysen, J. Phys. Chem. C, 2010, 114, 22822292.
90 T. Pieplu, F. Poignant, A. Vallet, J. Saussey, J. C. Lavalley and
J. Mabilon, in Catalysis and Automotive Pollution Control III, ed.
A. Frennet and J. M. Bastin, Stud. Surf. Sci. Catal., 1995, vol. 96,
pp. 619629.
91 K. Hadjiivanov, D. Klissurski, G. Ramis and G. Busca, Appl.
Catal., B, 1996, 7, 251267.
92 H. Praliaud, S. Mikhailenko, Z. Chajar and M. Primet, Appl.
Catal., B, 1998, 16, 359374.
93 J. Datka, P. Kozyra and E. Kukulska-Zajac, Catal. Today, 2004,
90, 109114.
94 J. J. Rack, J. D. Webb and S. H. Strauss, Inorg. Chem., 1996, 35,
277278.
95 S. H. Strauss, J. Chem. Soc., Dalton Trans., 2000, 16.
96 A. Zecchina, S. Bordiga, G. T. Palomino, D. Scarano,
C. Lamberti and M. Salvalaggio, J. Phys. Chem. B, 1999, 103,
38333844.
97 F. Xamena, P. Fisicaro, G. Berlier, A. Zecchina, G. T. Palomino,
C. Prestipino, S. Bordiga, E. Giamello and C. Lamberti, J. Phys.
Chem. B, 2003, 107, 70367044.
98 A. Le Nestour, M. Gaudon, G. Villeneuve, M. Daturi,
R. Andriessen and A. Demourgues, Inorg. Chem., 2007, 46,
40674078.
99 P. Concepcion, B. M. Reddy and H. Knozinger, Phys. Chem.
Chem. Phys., 1999, 1, 30313037.
100 K. Hadjiivanov, P. Concepcion and H. Knozinger, Top. Catal.,
2000, 1112, 123130.
101 R. Joyner and M. Stockenhuber, J. Phys. Chem. B, 1999, 103,
59635976.
102 H. Y. Chen, E. M. El-Malki, X. Wang, R. A. van Santen and
W. M. H. Sachtler, J. Mol. Catal. A: Chem., 2000, 162, 159174.
103 J. Novakova, L. Kubelkova, B. Wichterlova, T. Juska and
Z. Dolejsek, Zeolites, 1982, 2, 1722.
104 V. Blasin-Aube, O. Marie, J. Saussey, A. Plesniar, M. Daturi,
N. Nguyen, C. Hamon, M. Mihaylov, E. Ivanova and
K. Hadjiivanov, J. Phys. Chem. B, 2009, 113, 83878393.
105 E. Ivanova, M. Mihaylov, K. Hadjiivanov, V. Blasin-Aube,
O. Marie, A. Plesniar and M. Daturi, Appl. Catal., B, 2010, 93,
325338.
106 J. W. Yoon, Y.-K. Seo, Y. K. Hwang, J.-S. Chang, H. Leclerc,
S. Wuttke, P. Bazin, A. Vimont, M. Daturi, E. Bloch,
P. L. Llewellyn, C. Serre, P. Horcajada, J.-M. Grene` che,
A. E. Rodrigues and G. Fe rey, Angew. Chem., Int. Ed., 2010,
49, 59495952.
107 F. Fally, V. Perrichon, H. Vidal, J. Kaspar, G. Blanco,
J. M. Pintado, S. Bernal, G. Colon, M. Daturi and
J. C. Lavalley, Catal. Today, 2000, 59, 373386.
108 C. Binet and M. Daturi, Catal. Today, 2001, 70, 155167.
109 M. Daturi, E. Finocchio, C. Binet, J. C. Lavalley, F. Fally,
V. Perrichon, H. Vidal, N. Hickey and J. Kaspar, J. Phys. Chem.
B, 2000, 104, 91869194.
110 M. Daturi, E. Finocchio, C. Binet, J. C. Lavalley, F. Fally and
V. Perrichon, J. Phys. Chem. B, 1999, 103, 48844891.
111 E. Finocchio, M. Daturi, C. Binet, J. C. Lavalley and
G. Blanchard, Catal. Today, 1999, 52, 5363.
112 J.-P. Gallas, J.-M. Goupil, A. Vimont, J.-C. Lavalley, B. Gil,
J.-P. Gilson and O. Miserque, Langmuir, 2009, 25, 58255834.
113 F. Thibault-Starzyk, I. Stan, S. Abello , A. Bonilla, K. Thomas,
C. Fernandez, J.-P. Gilson and J. Pe rez-Ram rez, J. Catal., 2009,
264, 1114.
114 P. Bazin, A. Alenda and F. Thibault-Starzyk, Dalton Trans.,
2010, 39, 84328436.
115 S. J. Tinnemans, J. G. Mesu, K. Kervinen, T. Visser,
T. A. Nijhuis, A. M. Beale, D. E. Keller, A. M. J. van der Eerden
and B. Weckhuysen, Catal. Today, 2006, 113, 315.
116 G. Le Bourdon, F. Adar, M. Moreau, S. Morel, J. Rener,
A. S. Mamede, C. Dujardin and E. Payen, Phys. Chem. Chem.
Phys., 2003, 5, 44414444.
117 M. A. Newton, M. Di Marco, A. Kubacka and M. Fernandez-
Garcia, J. Am. Chem. Soc., 2010, 132, 45404541.
118 M. A. Newton, A. J. Dent, S. G. Fiddy, B. Jyoti and J. Evans,
Catal. Today, 2007, 126, 6472.
119 F. Mauge, A. Vallet, J. Bachelier, J. C. Duchet and J. C. Lavalley,
J. Catal., 1996, 162, 8895.
120 L. Oliviero, L. Mariey, M. A. Lelias, S. Aiello, J. van Gestel and
F. Mauge, Catal. Lett., 2010, 135, 6267.
121 N. Bion, J. Saussey, C. Hedouin, T. Seguelong and M. Daturi,
Phys. Chem. Chem. Phys., 2001, 3, 48114816.
122 N. Bion, J. Saussey, M. Haneda and M. Daturi, J. Catal., 2003,
217, 4758.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online
4950 Chem. Soc. Rev., 2010, 39, 49284950 This journal is c The Royal Society of Chemistry 2010
123 I. Malpartida, E. Ivanova, M. Mihaylov, K. Hadjiivanov,
V. Blasin-Aube, O. Marie and M. Daturi, Catal. Today, 2010,
149, 295303.
124 P. Bazin, O. Marie and M. Daturi, Journee detude SIA
Le post-traitement Diesel: perspectives et developpements,
Rouen, France, November 29, 2005.
125 S. Jolly, J. Saussey and J. C. Lavalley, J. Mol. Catal., 1994, 86,
401421.
126 J. F. Joly, N. Zanierszydlowski, S. Colin, F. Raatz, J. Saussey and
J. C. Lavalley, Catal. Today, 1991, 9, 3138.
127 F. Thibault-Starzyk, A. Vimont, C. Fernandez and J. P. Gilson,
Chem. Commun., 2000, 10031004.
128 C. Fernandez, I. Stan, J.-P. Gilson, K. Thomas, A. Vicente,
A. Bonilla and J. Pe rez-Ramrez, Chem.Eur. J., 2010, 16, 62246233.
129 F. Lepeltier, P. Chaumette, J. Saussey, M. M. Bettahar and
J. C. Lavalley, J. Mol. Catal. A: Chem., 1997, 122, 131139.
130 J. L. Freysz, J. Saussey, J. C. Lavalley and P. Bourges, J. Catal.,
2001, 197, 131138.
131 T. Lesage, C. Verrier, P. Bazin, J. Saussey and M. Daturi, Phys.
Chem. Chem. Phys., 2003, 5, 44354440.
132 W. Wasylenko and H. Frei, Phys. Chem. Chem. Phys., 2007, 9,
54975502.
133 H. Frei and Y. H. Yeom, in In situ spectroscopy of catalysts, ed.
B. M. Weckhuysen, American Scientic Publishers, Washington
DC, 2004.
134 F. Thibault-Starzyk, E. Seguin, S. Thomas, M. Daturi, H. Arnolds
and D. A. King, Science, 2009, 324, 10481051.
135 B. M. Weckhuysen, Angew. Chem., Int. Ed., 2009, 48,
49104943.
136 E. Stavitski, M. Kox, I. Swart, F. de Groot and B. Weckhuysen,
Angew. Chem., Int. Ed., 2008, 47, 35433547.
137 M. Kox, K. Domke, J. Day, G. Rago, E. Stavitski, M. Bonn and
B. Weckhuysen, Angew. Chem., Int. Ed., 2009, 48, 89908994.
138 V. Blasin-Aube , P. Bazin, F. C. Meunier, O. Marie and
M. Daturi, 3rd International Congress on Operando Spectroscopy,
Rostock-Warnemu nde, Germany, 2009.
D
o
w
n
l
o
a
d
e
d

b
y

U
N
I
V

F
E
D
E
R
A
L

D
O

R
I
O

D
E

J
A
N
E
I
R
O

o
n

1
1

O
c
t
o
b
e
r

2
0
1
1
P
u
b
l
i
s
h
e
d

o
n

2
9

O
c
t
o
b
e
r

2
0
1
0

o
n

h
t
t
p
:
/
/
p
u
b
s
.
r
s
c
.
o
r
g

|

d
o
i
:
1
0
.
1
0
3
9
/
B
9
1
9
5
4
3
M
View Online

Anda mungkin juga menyukai